cb9910en

Download as pdf or txt
Download as pdf or txt
You are on page 1of 393

MAIN REPORT

THE STATE
OF THE WORLD’S
LAND AND WATER
RESOURCES FOR FOOD
AND AGRICULTURE 2021
Systems at breaking point
THE STATE
OF THE WORLD’S
LAND AND WATER
RESOURCES FOR FOOD
AND AGRICULTURE 2021
Systems at breaking point

Main report

Food and Agriculture Organization of the United Nations


Rome, 2022
Required citation:
FAO. 2022. The State of the World’s Land and Water Resources for Food and Agriculture – Systems at breaking point. Main report. Rome.
https://doi.org/10.4060/cb9910en

The designations employed and the presentation of material in this information product do not imply the expression of any opinion
whatsoever on the part of the Food and Agriculture Organization of the United Nations (FAO) concerning the legal or development status of
any country, territory, city or area or of its authorities, or concerning the delimitation of its frontiers or boundaries. Dashed lines on maps
represent approximate border lines for which there may not yet be full agreement. The mention of specific companies or products of
manufacturers, whether or not these have been patented, does not imply that these have been endorsed or recommended by FAO in
preference to others of a similar nature that are not mentioned.

The views expressed in this information product are those of the author(s) and do not necessarily reflect the views or policies of FAO.

ISBN 978-92-5-136127-6
© FAO, 2022

Some rights reserved. This work is made available under the Creative Commons Attribution-NonCommercial-ShareAlike 3.0 IGO licence
(CC BY-NC-SA 3.0 IGO; https://creativecommons.org/licenses/by-nc-sa/3.0/igo/legalcode).

Under the terms of this licence, this work may be copied, redistributed and adapted for non-commercial purposes, provided that the work is
appropriately cited. In any use of this work, there should be no suggestion that FAO endorses any specific organization, products or
services. The use of the FAO logo is not permitted. If the work is adapted, then it must be licensed under the same or equivalent Creative
Commons licence. If a translation of this work is created, it must include the following disclaimer along with the required citation: “This
translation was not created by the Food and Agriculture Organization of the United Nations (FAO). FAO is not responsible for the content or
accuracy of this translation. The original [Language] edition shall be the authoritative edition.”

Disputes arising under the licence that cannot be settled amicably will be resolved by mediation and arbitration as described in Article 8 of
the licence except as otherwise provided herein. The applicable mediation rules will be the mediation rules of the World Intellectual Property
Organization http://www.wipo.int/amc/en/mediation/rules and any arbitration will be conducted in accordance with the Arbitration Rules of
the United Nations Commission on International Trade Law (UNCITRAL).

Third-party materials. Users wishing to reuse material from this work that is attributed to a third party, such as tables, figures or images,
are responsible for determining whether permission is needed for that reuse and for obtaining permission from the copyright holder. The risk
of claims resulting from infringement of any third-party-owned component in the work rests solely with the user.

Sales, rights and licensing. FAO information products are available on the FAO website (www.fao.org/publications) and can be purchased
through [email protected]. Requests for commercial use should be submitted via: www.fao.org/contact-us/licence-request.
Queries regarding rights and licensing should be submitted to: [email protected].

Cover photographs (from the top to bottom): ©FAO/Giulio Napolitano, ©FAO/Giuseppe Bizzarri, ©FAO/Alessia Pierdomenico, ©FAO/Sheam Kaheel
CONTENTS
Foreword xiii
PREFACE xv
Methodology: Global Datasets xix
AcknowledgEments xxi

1
Abbreviations and acronyms xxiv
Key messages of SOLAW 2021 xxvii

Status of and trends in land, 1


soil and water resources
1.1 Introduction 2
1.2 Emissions from land and the changing climate 2
1.3 Land-cover status and trends 7
1.4 Land-use trends 13
1.5 Soils under pressure 38
1.6 Land degradation – human pressures on land resources 50
1.7 Water scarcity 63
1.8 Conclusions 82

2
In focus: Fragile mountain agriculture 83
References 86

Socioeconomic Settings 98
2.1 Introduction 100
2.2 Socioeconomic transitions – 100
implications for land and water management
2.3 Diminishing per capita water resources availability 109
2.4 Patterns of landholding 113
2.5 Access to land and water 119
2.6 Competition for land and water – an issue of governance 125
2.7 Conclusions 126
Case study: Gender empowerment resolves 128
water-related conflict in Yemen
References 130

iii
3
RISKS TO LAND AND WATER 138
RESOURCES RUN DEEP
3.1 Introduction 140
3.2 Looking into the future 141
3.3 Land degradation risk 150
3.4 Water scarcity risk to land productivity 156
3.5 Conclusions 161
Case study: Understanding how groundwater responds 163
to climate and anthropogenic abstraction
Case study: Farmers and water utilities voluntarily 164
cooperating to reduce nitrate concentrations in Germany
Case study: Land and water systems at risk in the 166

4
Arab region due to climate change
References 170

Sustainable resources 174


planning and management
4.1 Introduction 176
4.2 Sustainable land resources planning 176
4.3 Reversing human-induced degradation 202
4.4 Planning for drought 214
4.5 Conclusions 223
Case study: Participatory land resources planning to 225
promote sustainable land management in Morocco
Case study: Information-based climate-proof land 227
management in the Lao People’s Democratic Republic
Case study: Restoring degraded land in Rohingya 229
refugee camp in Cox’s Bazar, Bangladesh
In focus: Dryland systems 232
References 246

iv
5
Responses and action areas 254
5.1 Introduction 256
5.2 Response platform: from global to individual efforts 257
5.3 Action area I: Adopting inclusive and adaptive 261
land and water governance
5.4 Action area II: Implementing integrated 269
solutions at scale
5.5 Action area III: Embracing innovative 285
technologies and management
5.6 Action area IV: Investing in long- term sustainability 312
5.7 Conclusions 319
Case study: Water accounting and auditing in the West Bank 320
Case study: Technology impacts on traditional water rights 321
systems in the Near East and North Africa region
Case study: Restoring rangeland productivity, biodiversity 323
and ecosystems in Ethiopia and Jordan
Case study: Unconventional farming in marginal areas 326

6
References 328

Conclusions and recommendations 340


6.1 The state of play offers no room for complacency 342
6.2 Socioeconomic development pathways are diverging 343
6.3 The risks to a food-secure future are proliferating 344
6.4 Responses should be better planned – 346
the tools are in place
6.5 Channel actions in four response areas 347

Annex: Country groupings 353


Glossary 357

v
Maps
1.1 Greenhouse gas emission intensity per unit of land (tonnes CO2-eq per km2) 3
1.2 Reference length of the growing period, average for period 5
1981–2010 (days)
1.3 Mean temperature change, 1961–2020 (°C) 6
1.4 Dominant land-cover classes, Global Land Cover Share data, 2010 9
1.5 Major agricultural systems in 2010 16
1.6 Economic water productivity for rice, average 1996–2005 (USD/m3) 25
1.7 Economic water productivity for wheat, average 1996–2005 (USD/m3) 26
1.8 Estimated inland fishery catch as a percentage of the global inland 29
2007–2016 catch
1.9 Global distribution of forests by climate domain, 2020 31
1.10 Global soil organic carbon, 2019 (tonnes/ha) 39
1.11 Soil organic carbon stocks in croplands, 2019 (tonnes/ha) 40
1.12 Soil organic carbon in the circumpolar region, 2019 (tonnes/ha) 41
1.13 Global map of salt-affected soils (global salt-affected areas) v1.0: 45
(top) topsoil (0–30 cm) and (bottom) subsoil (30–100 cm)
1.14 Biophysical status of land, 2015 53
1.15 Trend in biophysical status, 2015 54
1.16 Land degradation pressures, 2015 55
1.17 Dominant drivers of land degradation, 2015 56
1.18 Land degradation classes based on the severity 58
of human-induced pressures and deteriorating trends, 2015
1.19 Regions at risk based on status and trends of land resources, 2015 62
1.20 Level of water stress of all sectors by major basin, 2018 74
1.21 Level of water stress due to the agriculture sector by basin, 2018 75
1.22 Global water quality risk for three Sustainable Development Goal 6.3.2 78
indicators (nitrogen, electrical conductivity and biological oxygen demand),
modelling of the Global Freshwater Quality Database data 2000–2010 at 50 km
resolution
1.23 Annual anthropogenic phosphorus inputs into freshwater systems from 79
agriculture, industrial and domestic sectors, 2002–2010 (kg P/ha)
1.24 Global regions of concern (global areas susceptible to 80
pesticide pollution), 2010
3.1 Drought frequency on rainfed farming systems, 1984–2018 148

vi
3.2 Drought risk on pastoral farming systems, 1984–2018 148
3.3 Levels of water stress on irrigated areas, 2015 149
3.4 Regions at risk based on status and trends of land degradation, 2021 151
3.5 Irrigated cropland subject to human-induced land degradation in 2014: 158
(a) Africa and Western Asia, (b) South Asia, (c) East Asia, (d) Southeast Asia
and (e) parts of America
A Temperature for the period 1986–2005 and projected change 167
in temperature for the period 2046–2065
B Predicted effects of climate change on water availability for crops, 168
2046–2065
C Predicted effects of climate change on water availability for 169
livestock, 2046–2065
4.1 Potential production for rainfed wheat at high inputs (tonnes/ha), 185
based on average climate from 1981 to 2010 and global soil and
terrain information
4.2 Marginality hotspots – overlapping dimensions of marginality 186
4.3 Distribution and intensity of cropland, 2010 (% of 30 arcsecond grid cell) 187
4.4 Downscaled yield of rainfed maize, 2010 (tonnes/ha) 188
4.5 Yield achievement ratio (100 × actual/potential) for maize 189
under rainfed water supply conditions, 2010
4.6 Yield achievement ratio (100 × actual/potential) for 26 crops in 189
current rainfed cropland, 2009–2011
4.7 Difference in land suitability for rainfed wheat with high inputs on actual 192
cultivated land between baseline climate (1981–2010) and the projected
climate for the period 2070–2099, ensemble mean for Representative
Concentration Pathway 4.5 scenario
4.8 Difference in land suitability for rainfed coffee between baseline 193
climate (1981–2010) and the climate in the period 2070–2099, ensemble
mean for Representative Concentration Pathway 4.5 scenario
4.9 Difference in land suitability for rainfed wheat between baseline 195
climate (1981–2010) and the climate in the period 2070–2099, ensemble
mean for Representative Concentration Pathway 8.5 scenario
4.10 Multiple cropping zone classes for rainfed conditions, climate of 196
1981–2010
4.11 Multiple cropping zone classes for rainfed conditions, and the 196
climate in the period 2070–2099, ensemble mean for Representative
Concentration Pathway 4.5 scenario
4.12 Most suitable crops (suitability index > 40) for 1981–2010 climate 198
conditions based on an analysis of ten crops
4.13 Number of different crops possible to be grown (suitability index > 40) for 199
the period 2070–2099 climate conditions (ensemble mean for Representative
Concentration Pathway 4.5) based on analysis of ten crops

vii
4.14 Agricultural stress index for 1 May 2021 (%)' 218
A The world’s dry lands 233
B Aridity projections – drier types 236

Figures
1.1 Global agrifood system greenhouse gas' emissions by life-cycle stage 3
and per capita emissions, 1990–2019
1.2 Regional distribution of dominant land-cover classes (%)' 10
1.3 Land-cover trends, 1992–2019 (million ha) 11
1.4 Forest-cover trends, 1990–2020 (million ha per year) 14
1.5 Agricultural land-use trends, 1961–2019 (million ha) 14
1.6 Land-use trends (million ha) 17
1.7 Global distribution of irrigated surfaces by geographical region, 2018 22
1.8 Area equipped for irrigation by geographical region, 2000, 2012 22
and 2018 (million ha)
1.9 Irrigated area as a percentage of cultivated area, 2000, 2012 and 2018 (%) 23
1.10 Vegetable yields by region, 2012 (tonnes/ha) 25
1.11 Global land use under permanent meadows and pastures, 1961–2019 27
(million ha)
1.12 Top five inland waters fish capture producing countries, 1999–2018 29
(thousand tonnes)
1.13 Global forest areas in 2020 and net changes by decade, 1990–2020 32
1.14 Annual forest area net change by decade and region, 1990–2020 32
(million ha/year)
1.15 Global fertilizer inputs, 2010–2019 (tonnes) 35
1.16 Global crop protection inputs, 2010–2019 (tonnes/year) 36
1.17 Total factor productivity growth in world agriculture, 1961–2019 37
1.18 Global and continental nitrogen (top) and phosphorus (bottom) 43
fertilizer use, 1961–2019 (tonnes/year)
1.19 Annual global water balance, 2000 63
1.20 Global distribution of large dams and reservoirs, 2016 67
1.21 Total annual internal renewable water resources per capita by 69
geographical region, 2000, 2012 and 2018 (m3/capita)

viii
1.22 Total annual water withdrawals per capita by geographical region, 69
2000, 2012 and 2018 (m3/capita)
1.23 Evolution of global total water withdrawals, 1910–2018 (km3/year) 71
1.24 Level of water stress by geographical region, 2006, 2009, 2012, 73
2015 and 2018
2.1 Population by Sustainable Development Goal' region: estimates, 102
1950–2020 and medium-variant
projection with 95 percent prediction intervals, 2020–2100
2.2 Action tracks in a food system 104
2.3 Population distribution according to country threshold water 111
stress, 2000 (left) and 2018 (right)
2.4 Population density mean (people/km2) by water stress 111
class at major basin level, 2018 (%)
2.5 Worldwide distribution of farms and farmland, by land size class, 114
2010 census data for 129 countries and territories
2.6 Share of value of food production from smallholders (<2 ha), 114
by region and income grouping, 2010 census data for 129 countries
and territories
2.7 Analysing smallholdings and family farms through the sustainable rural 118
livelihoods framework
4.1 Search criteria and options for the Land Resources Planning Toolbox 180
4.2 The Global Agro-Ecological Zones soil suitability downscaling framework 183
4.3 Maize soil suitability at 6 065 Land Potential Knowledge System 184
sampling sites in Ghana based on low-input rainfed farming systems
4.4 Land degradation neutrality conceptual framework 208
4.5 Farmer field school generic sequence of activities 215
4.6 Cycle of disaster management 216
4.7 (Top) Ten steps and (bottom) three pillars of drought policy 217
and preparedness
A Distribution of land uses in dry lands (thousand ha) 233
B Distribution of dry lands among aridity zones (thousand ha) 234
C Total dryland population per ecosystem type 235
5.1 Ecosystem services from land, soil and water 274
5.2 Sources of incentives 274
5.3 The nexus approach 276
5.4 Logical framework for adapting to climate change through 280
land and water management in Eastern Africa
5.5 Sources of incentives 292

ix
5.6 Water accounting brings together all water flows and uses 293
5.7 Gross biomass water productivity, 2020 294
5.8 Farmer-led irrigation development 297
5.9 The water storage continuum 298
5.10 Number of large built-storage dams, 1900–2010 299
5.11 Relationship between rainfall variability and gross domestic 300
product, 1990–2016

Tables
1.1 Land-use class change, 1992 and 2000–2019 (million ha) 8
1.2 Global land-use class extent, 1992, 2000, 2010–2019 (million ha) 15
1.3 Areas equipped for irrigation, 1961, 2012 and 2018 21
1.4 Agricultural lands (cropland, permanent meadows and pastures) 33
per capita, 2000, 2010 and 2017
1.5 Input layers for overall biophysical status, overall trend and 52
cumulative pressure by drivers
1.6 Extent of human-induced land degradation, 2015 (million ha) 57
1.7 Extent of land degradation, 2015 (million ha) 57
1.8 Human-induced land degradation, 2015 59
1.9 Extent of land degradation classes for global land cover, 2015 61
1.10 Areas of agricultural land and forest at risk, 2015 62
1.11 Permanent and seasonal water cover on irrigated land, 64
2019 and 2000–2004 changes (ha)
1.12 Permanent and seasonal water cover on rainfed cropland, 65
2019 and 2000–2004 changes (ha)
1.13 Total water and total freshwater withdrawals for human use, 68
and percentage of total water withdrawals, 2018
3.1 Anticipated increases in crop water requirements induced by 146
demand and climate forcing to 2050
3.2 Productive land at risk from land degradation, 2021 152
4.1 Change in the extent (km2) of land suitability classes for rainfed coffee 194
between baseline climate (1981–2010) and the climate in the period 2070–
2099 (2080s), for ensemble mean for Representative Concentration Pathway
4.5 scenario
4.2 Changes of rainfed multiple cropping potentials between baseline 197
climate (1981–2010) and the 2080 climate (ensemble mean for Representative
Concentration Pathway 4.5)

x
4.3 Changes in production for baseline climate and climate scenario 200
ensembles (ENS) for the 2050s and 2080s with (+) and without
carbon dioxide fertilization on very suitable (VS), suitable (S) and
moderately suitable (MS) land
5.1 Selected nature-based solution investment case studies 315

Boxes
1.1 Global land degradation assessment using the adapted Global Land 52
Degradation Information System method
Five mountain production systems 84
2.1 Globally Important Agricultural Heritage Systems 116
2.2 The sustainable rural livelihoods framework 117
3.1 FAO future of food and agriculture scenarios from a land- and 143
water-use perspective
4.1 Innovative tools and approaches for land-use planning 178
4.2 Localizing/increasing accuracy of Global Agro-Ecological Zones 183
predictions with site-specific soil data collected using a mobile application
4.3 Marginal lands for crop production 186
4.4 Forecast crop yields informing the European Union’s 190
Common Agricultural Policy and drought management
4.5 High-end climate change impact on rainfed crops in Ghana 199
4.6 Summary of anticipated shifts in land suitability 201
4.7 Sustainable land management objectives 205
4.8 Lessons learned from the TerrAfrica Strategic Investment Programme 211
4.9 Restoration interventions in the Great Green Wall for 212
the Sahara and Sahel Initiative
4.10 Farmers field schools develop sustainable land management capacity 215
4.11 Monitoring and early warning systems 218
4.12 Responding to crises in the Horn of Africa 220
4.13 An integrated drought management plan in Central and Eastern Europe 222
A The Aral Sea: impacts of land degradation on ecosystems and human health 237
B Rome Promise on Monitoring and Assessment of Drylands 239
for Sustainable Management and Restoration
C Sustainable land management practices in rangelands in 243
sub-Saharan Africa

xi
5.1 International frameworks: convergence around integrated, 259
sustainable and equitable land and water governance
5.2 Facilitating policy coherence and integrated land and water 262
governance through climate responses
5.3 Strengthening water governance and water tenure rights 265
5.4 Green Negotiated Territorial Development and its contribution 267
to improving livelihoods resilience
5.5 Role of women in water resources management in agriculture 268
5.6 Supporting water governance analysis 270
5.7 Watershed management for resilience and sustainable land 272
management scaling out
5.8 Payment for ecosystem services: investing in nature, investing in people 273
5.9 The nexus approach 277
5.10 Koronivia Joint Work on Agriculture 279
5.11 Nature-based solutions help mitigate droughts and floods 282
5.12 Agroforestry can enhance soil fertility 283
5.13 Reducing risks, addressing vulnerability and enhancing 284
pastoralist resilience in Africa
5.14 Soil biological diversity at the heart of sustainable soil management 287
5.15 Rainwater harvesting serves protected cropping in Lebanon 289
5.16 Montana versus Wyoming: sprinklers, irrigation efficiency 291
and recapturing return flows
5.17 “Following the water” to assess “real” water savings 292
5.18 Water accounting and auditing 293
5.19 The Fertilizer Code 302
5.20 Urban farming: a solution to enhance food security in cities 305
5.21 Intensifying production using rice fallows to grow pulses and vegetables 307
5.22 Reducing food loss and waste in Senegal and the United Kingdom 308
5.23 European Commission’s circular economy action plan and 310
food loss and waste
5.24 Wastewater: a potential water resource in the Central 311
America and Caribbean region
5.25 Valuing wastewater as a source of nutrients for agriculture 312
5.26 Financing nature-based solutions 314

xii
Foreword

The state of the world’s land and water resources for food and agriculture 2021 (SOLAW 2021)
provides new information on the status of land, soil and water resources, and evidence of the
changing and alarming trends in resource use. Together, they reveal a situation that has much
deteriorated in the last decade, when the first SOLAW 2011 report highlighted that many of our
productive land and water ecosystems were at risk. The pressures on land and water ecosystems
are now intense, and many are stressed to a critical point.

Against this background, it is clear our future food security will depend on safeguarding our
land, soil and water resources. The growing demand for agrifood products requires us to look
for innovative ways to achieve the Sustainable Development Goals, under a changing climate and
loss of biodiversity. We must not underestimate the scale and complexity of this challenge. The
report argues that this will depend on how well we manage the risks to the quality of our land and
water ecosystems, how we blend innovative technical and institutional solutions to meet local
circumstances, and, above all, how we can focus on better systems of land and water governance.

The interlinked actions and coalitions resulting from the 2021 United Nations Food Systems
Summit provide an important entry to renew national and global priorities, and as a basis to
advance the transformation of our agrifood systems to be more efficient, inclusive, resilient and
sustainable.

A meaningful engagement with the key stakeholders – farmers, pastoralists, foresters and
smallholders – directly involved in managing soils and conserving water in agricultural land-
scapes is central. These are nature’s stewards and the best agents of change to adopt, adapt and
embrace the innovation we need to secure a sustainable future.

I invite you to read the SOLAW 2021 report with a view to the fundamentals of all terrestrial agri-
food production. Land degradation and water scarcity will not disappear. However, while the scale
of the challenge is daunting, whether as cultivators of land or consumers of food, even small shifts
in behaviours will see the much-needed transformation at the core of our global agrifood systems.

The new FAO Strategic Framework 2022-31 firmly commits the Organization to promote the
sustainable management of our vital land and water ecosystems for better production, better
nutrition, a better environment and a better life for all, leaving no one behind.

Dr Qu Dongyu
Director-General
Food and Agriculture Organization of the United Nations

xiii
preface

Setting the scene


Human use of land and water for agriculture has not yet peaked, but all evidence points to slow-
ing growth in agricultural productivity, rapid exhaustion of productive capacity and generation
of environmental harm. Taking production that is more environmentally responsible and
climate smart to scale can reverse trends in the deterioration of land and water resources and
promote inclusive growth. This aligns with the aspirations of the FAO strategic framework:
“better production, better nutrition, a better environment and a better life”.

The past decade has seen the advent of several important global
policy frameworks including the 2030 Agenda for Sustainable Taking production that
Development, the Paris Agreement on climate change, the Sendai is more environmentally
Framework for Disaster Risk Reduction 2015–2030, the Small responsible and climate
Island Developing States Accelerated Modalities of Action, the New smart to scale can reverse
Urban Agenda and the Addis Ababa Action Agenda on Financing for trends in the deterioration
Development. The frameworks have introduced the Sustainable of land and water
Development Goals (SDGs), nationally determined contributions
resources and promote
inclusive growth.
and land degradation neutrality. In particular, there are dedi-
cated SDGs for water, and targets for land and soil health. The
frameworks are accompanied by global assessments of natural
resources, including soils, forestry, biodiversity, desertification
and climate. The state of the world’s land and water resources for
food and agriculture 2021 (SOLAW 2021) report aims to take stock
of the implications for agriculture and recommend solutions
for transforming the combined role of land and water in global ©Oliver Nguyen
food systems.

The uncertainty of climate change and the complex feedback loops


between climate and land present agriculture with amplified levels
of risk that need to be managed. A global view points to a convergence of factors putting unprec-
edented pressure on land and water resources, leading to a set of human impacts and shocks in
the supply of agricultural products, notably food. The SOLAW 2021 report argues that a sense of
urgency needs to prevail over a hitherto neglected area of public policy and human welfare, that
of caring for the long-term future of land, soil and water.

xv
Shocks, including severe floods, droughts and the COVID-19
pandemic tend to divert attention away from development
priorities. International finance institutions warn of the
widening fault lines between developed and developing coun-
tries in meeting global goals while facing resurgent infections
and rising death tolls from COVID-19. Recovery programmes
offer opportunities to address urgencies and kick-start the
process of change, including in land and water management.

Land, soil and water form the basis of the FAO commit-
©FAO/Giulio Napolitano

ment to the changes advocated in the 2021 United Nations


Food Systems Summit. However, recognition and actions are
needed to redirect the focus onto the land, on which 98 percent
of the world’s food is produced. Taking care of land, water
and particularly the long-term health of soils is fundamental
Taking care of land, to accessing food in an ever-demanding food chain, guar-
water and particularly the anteeing nature-positive production, advancing equitable
long-term health of soils is livelihoods, and building resilience to shocks and stresses
fundamental to accessing arising from natural disasters and pandemics. They all start
food in an ever-demanding from land and water access and governance. Sustainable
food chain. land, soil and water management also underpin nutritious,
diverse diets and resource-efficient value chains in the shift
to sustainable consumption patterns.

What SOLAW 2021 says


The SOLAW 2021 report comes at a time when human pressures on the systems of land, soils
and freshwater are intensifying, just when they are being pushed to their productive limits. The
impacts of climate change are already constraining rainfed and irrigated production over and
above the environmental consequences resulting from decades of unsustainable use.

The SOLAW 2021 report builds on the concepts and conclusions given in the previous SOLAW
2011 report. Much has happened in the intervening years. Recent assessments, projections and
scenarios from the international community paint an alarming picture of the planet’s natural
resources – highlighting overuse, misuse, degradation, pollution and increasing scarcity. Rising
demands for food and energy, competing industrial, municipal and agricultural uses, and the
need to conserve and enhance the integrity of the Earth’s ecosystems and their services make the
picture extremely complex and full of interlinkages and interdependencies.

The SOLAW 2021 report adopts the driver–pressure–state–impact–response approach. This is


a well-established framework for analysing and reporting important and interlinked relation-
ships among sustainable agricultural production, society and the environment. The approach
provides a structure to report on cause–effect relationships to arrive at key policy recommenda-

xvi
tions and enable policymakers to assess the direction and nature of changes needed to advance
sustainable management of land and water resources.

The drivers of demand for land and water resources are complex. By 2050, FAO estimates
agriculture will need to produce almost 50 percent more food, fibre and biofuel than in 2012 to
satisfy global demand and keep on track to achieve “zero hunger” by 2030. Progress made in
reducing the number of undernourished people in the early part of the twenty-first century has
been reversed. The number has risen from 604 million in 2014 to 768 million in 2020. While
prospects for meeting the nutritional requirements of 9.7 billion people by 2050 at the global
level exist, problems with local patterns of production and consumption are expected to worsen,
with increasing levels of undernourishment and obesity among the steadily growing and mobile
population.

Options to expand cultivated land areas are limited. Prime agricultural land is being lost
to urbanization. Irrigation already accounts for 70 percent of all freshwater withdrawals.
Human-induced land degradation, water scarcity and climate change are increasing the levels
of risk for agricultural production and ecosystem services at times and in places where economic
growth is needed most.

Most pressures on the world’s land, soil and water resources derive from agriculture itself. The
increase in use of chemical (non-organic) inputs, uptake of farm mechanization, and overall
impact of higher monocropping and grazing intensities are concentrated on a diminishing stock
of agricultural land. They produce a set of externalities that spill over into other sectors, degrad-
ing land and polluting surface water and groundwater resources.

The impacts from accumulating pressures on land and water are Human-induced land
felt widely in rural communities, particularly where the resource degradation, water scarcity
base is limited and dependency is high, and to a certain extent and climate change are
in poor urban populations where alternative sources of food are increasing the levels of risk
limited. Human-induced deterioration of land, soil and water for agricultural production
resources reduces production potential, access to nutritious food and ecosystem services
and, more broadly, the biodiversity and environmental services at times and in places
that underpin healthy and resilient livelihoods.
where economic growth is
needed most.
A central challenge for agriculture is to reduce land degradation
and emissions and to prevent further pollution and loss of envi-
ronmental services while sustaining production levels. Responses
need to include climate-smart land management attuned to
variations in soil and water processes. Management options are
available to increase productivity and production levels if inno-
©FAO/Giulio Napolitano

vation in management and technology can be taken to scale to


transition to sustainable agrifood systems. However, none of this
can go far without planning and managing land, soil and water
resources through effective land and water governance.

xvii
Increasing land and water productivity is crucial for achieving food security, sustainable
production and SDG targets. However, there is no “one size fits all” solution. A “full package” of
workable solutions is now available to enhance food production and tackle the main threats from
land degradation, increasing water scarcity and declining water quality.

The SOLAW 2021 report indicates how institutional and technical


responses can be packaged to address the challenges of increas-
ing water and food security within land, soil and water domains,
and, more widely, across agriculture and food systems. It stresses
the importance of integrated approaches in managing land and
water resources. Sustainable land management, sustainable soil
management and integrated water resources management are all
©FAO/IFAD/WFP/Michael Tewe

examples of such approaches, which can be blended with tech-


nology innovation, data and policies to accelerate improvement
in resource-use efficiency, raise productivity and align progress
with SDGs.

An important point to recognize is that many agents of change


in the landscape remain excluded from the benefits of technical

Injecting a sense of advances. This applies to disproportionately poorer and socially

urgency into making the disadvantaged groups, with most living in rural areas. While
necessary transformations technical solutions to specific land and water challenges may be
in the core of the global within grasp, much will depend on how land and water resources
food system is essential. are allocated. Inclusive forms of land and water governance
will be adopted at scale only when there is political will, adap-
tive policymaking and follow-through investment. A primary
focus on land and water governance is essential in creating the
transformative changes needed to achieve patterns of sustainable agriculture that can enhance
income and sustain livelihoods while protecting and restoring the natural resource base.

Significant complementary efforts will also be needed in food systems beyond the farm to maxi-
mize synergies and manage trade-offs in related sectors, particularly energy production. For
this to happen, changes in policy, institutional and technical domains that disrupt “business as
usual” models may prove necessary.

Time is of the essence. Current trends in natural resource depletion indicate production from
rainfed and irrigated agriculture is operating at or over the limit of sustainability. Injecting
a sense of urgency into making the necessary transformations in the core of the global food
system is essential.

Chapter 1 of this report provides a base from which to examine the socioeconomic trends in
Chapter 2 and the demand projections for land and water resources and attendant risks in Chap-
ter 3. These assessments provide a rationale for resource planning and management in Chapter
4 and for implementing institutional adaptation and technological innovation to increase crop
production and productivity while conserving natural resources in Chapter 5. Finally, Chapter 6
presents the conclusions drawn from the report and offers overall recommendations and action
in four key areas.

xviii
Methodology –
Global datasets

The global datasets used to assess environmental change have advanced since the first edition
of The state of the world’s land and water resources for food and agriculture (SOLAW) report in
2011. Annual “snapshots” of land-cover classifications are now derived from higher-resolution
imagery under the European Space Agency’s Climate Change Initiative using the FAO land-cover
classification scheme. The Global Forest Resource Assessment provides an up-to-date account
of net global forest loss. Continental coverage of monthly water consumption by growing
vegetation is available in the FAO Water Productivity Open-access Portal. In addition, the devel-
opment of the Global Agro-Ecological Zones (GAEZ) version 4 (v4) data portal now consolidates
the global distribution of land and agroclimatic resources at high resolution (~1 km) to analyse
the distribution of crop production for reference years and the potential for crop production
under climate change.

Translating these changes in land cover and associated energy balances into land and water use
for agricultural production is possible. Trends in agricultural production derived from national
statistics are attributed to the land where agroclimatic conditions and available soil moisture
are adequate for crop growth. Accordingly, the spatial frame of reference for this edition of
the SOLAW 2021 report is the set of agroclimatic and land data compiled for GAEZ v4 and is an
update of the GAEZ v2/3 used in the compilation of SOLAW 2011.

There are two baseline or reference years for GAEZ: 2000 and 2010. Reported agricultural
production in these reference years is distributed across 12 main land-use/land-cover shares in
each 5 arcminute cell. These shares are for: artificial surfaces, cropland, grassland, tree-covered
areas, shrub-covered areas, herbaceous vegetation (aquatic or regularly flooded), mangroves,
sparse vegetation, bare soil, snow and glaciers, water bodies and cropland equipped with full
control irrigation. These are the major land-class layers in GAEZ to which the FAO Statistical
Database (FAOSTAT) national crop production data are distributed (downscaled) through refer-
ence to land cover (FAO Global Land Cover Share) and land equipped for irrigation (FAO Global
Map of Irrigated Areas v5).

The GAEZ v4 unit of analysis is the 30 arcsecond pixel used to compile its reference grid. This
represents approximately 900 m at the equator and 600 m at the poles. The compilation of
climate, soil, land-cover and water source data at this resolution allows GAEZ to depict a nomi-
nal “state” of land and related water resources in a set of land-use types that conform with
FAO’s Global Information System on Water and Agriculture (AQUASTAT) and FAOSTAT produc-
tion data (i.e. they can accommodate reported harvested areas, yield and cropping intensity).

xix
The AQUASTAT database has been regularly updated since 2011, providing up-to-date informa-
tion on water resources for agriculture at the global level. It plays a crucial role in collecting data
and monitoring achievement of Sustainable Development Goal (SDG) 6: “ensure availability
and sustainable management of water and sanitation for all”, and in particular indicators of
SDG target 6.4 on water stress and water-use efficiency. The AQUASTAT method for collecting
data has evolved since 2018, relying on a network of AQUASTAT national correspondents who
ensure data collection and quality. This allows AQUASTAT to align with the country-led and
country-owned processes promoted through the SDGs for gathering data.

xx
Acknowledgements

The preparation of the SOLAW 2021 report has benefited from the support and input of a number
of individuals and institutions.

Overall supervision and review: L. Li and S. Koo-Oshima.

Early stage conceptualization: E. Mansur and O. Unver.

Coordination: F. Ziadat.

Chapter authors: V. Boerger, D. Bojic, P. Bosc, M. Clark, D. Dale, M. England, J. Hoogeveen, S.


Koo-Oshima, P. Mejias Moreno, D. Muchoney, F. Nachtergaele, C. Olivera Sanchez, M. Salman,
S. Schlingloff, O. Unver, R. Vargas, L. Verchot, Y. Yigini and F. Ziadat.

Editorial team: M. Kay (chief editor), S. Bunning and J. Burke.

Independent Advisory Committee: U. Apel, M. Astralaga, A. Bahri, F. Denton, J. Herrick, B.


Hubert, B. Orr, G. de Santi, J. Sara, A. P. Schlosser, A. Szöllösi-Nagy and F. Tubiello.

Further contributions to chapters: W. Ahmad, A. Bhaduri, R. Biancalani, C. Biradar, A. Bres, L.


Caon, R. Cuevas, D. Dale, F. El-Awar, S. Farolfi, V. Graw, N. Harari, J. Maynard, R. Mekdaschi
Studer, B. Orr, E. Pek, N.E. Rodriguez, Y. Tong, J.J. Troell, S. Uhlenbrook, L. Verchot and P.
Waalewijn.

External and internal technical reviewers: E. Aksoy, S. Alexander, J. Barron, M. Bernoux,


T. Brewer, S. Burchi, A. Cattaneo, M. Chaya, T. Darwish, B. Davis, I. Elouafi, C. Giupponi, N.
Harari, S. Hodgson, J. Jorgensen, P. Lidder, J. Lundqvist, R. Mekdaschi Studer, J. Molina Cruz, L.
Montanarella, V. Nangia, T. Oweis, A. Pandya, E. de Pauw, R. Poch, S. Ramasamy, C. Ringler, M.
Torero, S. Uhlenbrook, H. Van Velthuyzen, L. Verchot, P. Waalewijn, Y. Wada and P. Zdruli.

Regional consultation process: M. Alagcan, J. Ariyama, I. Beernaerts, A. Bhaduri, T. Estifanos,


J. Faures, M. Hamdi, T. Hofer, R. Jehle, T. Lieuw, Y. Niino, V. Nzeyimana, J. Quilty, E. Rurangwa
and T. Santivanez.

Process facilitation: R. DeLaRosa, M. Kay, K. Khazal, O. Unver and F. Ziadat.

Copy-editor: C. Brown.

Preparation and review of thematic background reports and case studies: M. Abdel Monem,
D. Agathine, H. Ahmadzai, M. Al Hamdi, F. Arafat, J. Ariyama, C. Batchelor, L. Battistella, O.
Berkat, A. Bhaduri, R. Biancalani, E. Borgomeo, R. Brathwaite, A. Bres, M. Bruentrup, A. Catta-
neo, F. Chiozza, C. Chouchani Cherfane, R. Coppus, D. Dale, B. Davis, P. Dias, M. De Gaetano, E.
Donegan, I. Elouafi, M. England, S. Farolfi, J. Faures, L. de Felice, T. Fetsi, M. Flores Maldonado,

xxi
A. Malhotra, G. Franceschini, E. Ghosh, I. Gil, V. Gillet, G. Grossman, G. Gruere, F. Haddad,
M. Henry, J. Herrick, T. Hoang, A. Huber-Lee, S. Iftekhar, R. Jalal, T. Kahil, P. Kanyabujinja
Nshuti, K. Khazal, B. Kiersch, D. Kulis, F. Kumar Mondal, P. Lidderr, J. Lindsay Azie, C. Lucrezia,
D.A. Lyra, R. Mahamud, Z. Makhamreh, Y. Makino, M. Merlet, F. Nachtergaele, V. Nzeyimana,
V. Onyango, R. Ouchna, P. Panagos, L. Peiser, M. Petri, J. Preissing, C. Richerzhagen, S. Ritu, O.
Rochdi, W. Saleh, N. Santos, W. Scheumann, R. Selvaraju, M. de Souza, T. Tang, A. Tanjim, M.A.
Tomaszkiewicz, H. Tropp, S. Tutundjian, G. Veleasco, L. Verchot and Y. Wada.

Preparation of statistics and maps: G. Ben Hamouda, J. Burke, F. Chiozza, R. Coppus, V. Graw, M.
Hernández, T. Hoang, K. Khazal, M. Marinelli, L. Peiser and A. Sander.

Publishing arrangements, communications and graphic design: A. Asselin-Nguyen, K. Khazal,


P. Mander, J. Morgan and M. Piraux.

Secretarial assistance: A. Grandi.

Institutions involved in preparation: SOLAW 2021 is a collaborative effort led by the Land
and Water Division of FAO in collaboration with several divisions/units at FAO headquarters,
regional and country offices, senior advisers and key partners. Appreciation is given to the
partner institutions that provided data and written contributions:

Agricultural Research for Development

Asian Soil Partnership

Australian Centre for International Agricultural Research

Centre for Rural Development, Humboldt University of Berlin

Deutsches Institut für Entwicklungspolitik – German Development Institute

Environmental Law Institute

Federal Ministry for Food and Agriculture (Germany)

French Agricultural Research Centre for International Development

Future Earth/Water Futures Group

Griffith University

International Center for Agricultural Research in the Dry Areas

International Center for Biosaline Agriculture

International Center for Tropical Agriculture

International Commission on Irrigation and Drainage

International Groundwater Resources Assessment Centre

International Institute for Applied Systems Analysis

International Water Management Institute

xxii
Joint Research Centre, European Commission

Organisation for Economic Co-operation and Development

Stockholm Environment Institute

Stockholm International Water Institute

Thünen Federal Research Institute for Rural Areas, Forestry and Fisheries

World Overview of Conservation Approaches and Technologies

xxiii
Abbreviations
and acronyms

AEZ agroecological zoning

2030 Agenda 2030 Agenda for Sustainable Development

AQUASTAT FAO’s Global Information System on Water and Agriculture

BAU business as usual

CBD Convention on Biological Diversity (United Nations)

CFS Committee on World Food Security

CO2-eq carbon dioxide equivalent

CRC conflict resolution committee

ENS ensemble

FAO Food and Agriculture Organization of the United Nations

FAOSTAT FAO Statistical Database

FFS farmer field school

FLW food loss and waste

FOFA future of food and agriculture

GAEZ Global Agro-Ecological Zones

GDP gross domestic product

GEF Global Environment Facility

GHG greenhouse gas

GIAHS Globally Important Agricultural Heritage Systems

GIS Geographic Information System

GLADIS Global Land Degradation Information System

GLC-SHARE Global Land Cover Share

xxiv
GLOSIS Global Soil Information System

GMIA Global Map of Irrigated Areas

GSP Global Soil Partnership

HWSD Harmonized World Soil Database

ICT information and communications technology

IFI international financing institution

IIASA International Institute for Applied Systems Analysis

IPBES Intergovernmental Science-Policy Platform on Biodiversity and


Ecosystem Services

IPCC Intergovernmental Panel on Climate Change

IRWR internal renewable water resources

IUCN International Union for Conservation of Nature

IWRM integrated water resources management

JRC Joint Research Centre (European Commission)

KJWA Koronivia Joint Work on Agriculture

LADA Land Degradation Assessment in Drylands

LandPKS Land Potential Knowledge System

LDN land degradation neutrality

LRP land resources planning

MARS Monitoring Agricultural Resources

MEWS monitoring and early warning system

NAEZ national agroecological zoning

NbS nature-based solution

NDC nationally determined contribution

NDVI normalized difference vegetation index

NGO non-governmental organization

OECD Organisation for Economic Co-operation and Development

PES payments for ecosystem services

RCP Representative Concentration Pathway

xxv
REDD+ Reducing Emissions from Deforestation and Forest Degradation in
Developing Countries

REWAS Real Water Savings

RUSLE Revised Universal Soil Loss Equation

SAMIS Strengthening Agro-climatic Monitoring and Information Systems

SDG Sustainable Development Goal

SDS sandstorm and dust storm

SI suitability index

SIP Strategic Investment Programme

SLM sustainable land management

SOC soil organic carbon

SOLAW The state of the world’s land and water resources for food and
agriculture

SOM soil organic matter

SPI Science-policy Interface (United Nations Convention to


Combat Desertification)

SRL sustainable rural livelihoods

SSP Shared Socioeconomic Pathway

SSS stratified societies

TSS towards sustainability

UNCCD United Nations Convention to Combat Desertification

UNEP United Nations Environment Programme

UNFCCC United Nations Framework Convention on Climate Change

VGGT Voluntary guidelines on the responsible governance of tenure of land,


fisheries and forests in the context of national food security

WaPOR Water Productivity Open-access Portal

WOCAT World Overview of Conservation Approaches and Technologies

WPI water productivity index

WUA water user association

xxvi
Key messages of SOLAW 2021
The state
The interconnected systems of land, soil and water are stretched to the limit.
Convergence of evidence points to agricultural systems breaking down, with impacts
felt across the global food system.

Current patterns of agricultural intensification are not proving sustainable. Pressures


on land and water resources have built to the point where productivity of key
agricultural systems is compromised and livelihoods are threatened.

Farming systems are becoming polarized. Large commercial holdings now dominate
agricultural land use, while fragmentation of smallholder concerns concentrates
subsistence farming on lands susceptible to degradation and water scarcity.

The challenges
Future agricultural production will depend upon managing the risks to land and
water. Land, soil and water management needs to find better synergy to keep
systems in play. This is essential to maintain the required rates of agricultural growth
without further compromising the generation of environmental services.

Land and water resources will need safeguarding. There is now only a narrow margin
for reversing trends in resource deterioration and depletion, but the complexity and
scale of the task should not be underestimated.

Responses and actions


Land and water governance has to be more inclusive and adaptive. Inclusive
governance is essential for allocating and managing natural resources. Technical
solutions to mitigate land degradation and water scarcity are unlikely to succeed
without it.

Integrated solutions need to be planned at all levels if they are to be taken to


scale. Planning can define critical thresholds in natural resource systems, leading to
the reversal of land degradation when wrapped up as packages or programmes of
technical, institutional, governance and financial support.

Technical and managerial innovation can be targeted to address priorities and


accelerate transformation. Caring for neglected soils, addressing drought and coping
with water scarcity can be addressed through the adoption of new technologies and
management approaches.

Agricultural support and investment can be redirected towards social and


environmental gains derived from land and water management. There is now scope
for progressive multiphased financing of agricultural projects that can be linked with
redirected subsidies to keep land and water systems in play.
1
© FAO/Olivier Asselin

Status of and
trends in land, soil
and water resources
Key messages
Land used for crop production increased by 208 million ha (15 percent) between 1961 and 2019. Land
used for irrigated cropping increased by 110 percent, while rainfed cropping increased by only 2.6 percent,
over the same time period. Permanent pastures for livestock rearing markedly declined from a peak area
of 3 400 million ha in 2000 to a level nearer to 3 200 million ha by 2019. This decline, together with global
population growth, reduced available agricultural land use per capita for crops and livestock rearing by
20 percent between 2000 and 2019. Agricultural land per capita is now less than 0.64 ha.

Pressures on productive land and water resources are pushing the productive capacity of agricultural
ecosystems to the limit. Land degradation, drought and related water scarcity are compromising
agricultural production and intensifying poverty and malnutrition in all regions.

The loss of soil organic carbon (SOC) is accelerating. Agriculturally managed soils contain 25 percent to 75
percent less SOC compared to soils in undisturbed or natural ecosystems. This is due to changing land use and
land management. Soils under conventional agriculture continue to be a source of carbon dioxide emissions.
Land and soils are degrading due to the spread and intensification of agriculture. Estimates suggest
human-induced degradation affects 34 percent of cropland and pasture. The demand for more calories to
satisfy population and income growth is constrained as cropping extends into marginal lands and existing
land suffers erosion and depletion of carbon, nutrients and soil biodiversity. Estimates suggest over
3.2 billion people are directly affected by soil/land degradation.

Water scarcity is becoming endemic. The local impact of physical water scarcity and freshwater pollution is
spreading and accelerating. The first sign of scarcity is increasing use and severe depletion of groundwater
– the ultimate source of water for most of the world. The global Sustainable Development Goal (SDG)
target 6.4 on water scarcity reached 18 percent in 2018, but this masks significant regional variations.
Non-conventional water use in agriculture, such as water/effluent reuse and desalination, is growing,
particularly in areas where water scarcity is most acute.

Accessible, high-quality groundwater is diminishing. Globally, groundwater accounts for over 30 percent
of freshwater withdrawals for irrigated agriculture and continues to grow at around 2.2 percent per year.
Approximately 70 percent of groundwater withdrawals are used to irrigate food, fibre and industrial
crops, and for livestock. More is used in arid and semi-arid regions. Agricultural production is constrained
where groundwater storage is depleted or degraded. Intensive exploitation in many principal continental
aquifers and saline intrusion along highly productive coastal plains are evident. This level of groundwater
exploitation is considered responsible for the loss of aquifer storage of 250 km3/year, and more importantly,
loss of aquifer function and utility to farmers as groundwater levels drop.

Water pollution is a rising global crisis that directly affects health, economic development and food
security. Agriculture is the dominant source of water pollution (mainly diffuse or non-point pollution from
agricultural land), but other human activities such as urbanization and industry are also major contributors.
Degrading water quality is a significant threat to food safety and food security.

Climate change is driving processes that cause productive land to be lost. Although anticipated
temperature changes may bring new land into production, opportunities for sustainable expansion and
intensification are severely limited. Climate change increases evapotranspiration from cropped land, and
alters the quantity and distribution of rainfall. This leads to changes in land/crop suitability and reduced
yields where temperature stresses attenuate carbon assimilation. Long-term temperature increases can be
anticipated across productive land, but rainfall intensities, duration and frequency are harder to predict.
Greater variations in river flows and groundwater recharge are expected and will adversely affect irrigated
agriculture in particular. Land-cover distribution over thermal climates and trends indicate increases in
grasslands and artificial surfaces, while tree-covered areas and bare areas show significant declines.

Land and water productivity gains over the past decade have enabled crop and livestock production
to match demand but at a cost. Land now produces more than 95 percent of the global food supply
for a human population estimated at over 7.7 billion. Unsustainable agricultural intensification has
increased environmental impacts that limit agricultural production capacity and damage a wide range of
environmental services. Intersectoral competition for land and water resources is intense, and the scope to
sustainably extend irrigation areas and convert new land to agriculture is limited.
1.1 Introduction
Pressures on land and water resources are
pushing the productive capacity of land and

©FAO/Stefanie Glinski
water systems to the limit. These concerns
are reflected in global environmental and
scientific assessments, notably the Intergov-
ernmental Panel on Climate Change (IPCC)
special report on climate change and land
(IPCC, 2019), the sixth edition of the United
is to describe the state of land and water
Nations Environment Programme (UNEP)
resources at the global level using the best
global environmental outlook (UNEP, 2019),
available global datasets to establish a base-
the Intergovernmental Science-Policy Plat-
line up to 2019 for land and 2018 for water data
form on Biodiversity and Ecosystem Services
according to the status of the FAO land and
(IPBES) assessment report on land degra-
water databases in 2021. Many of the global
dation and restoration (IPBES, 2018) and
datasets on related environmental data were
the United Nations Convention to Combat
not established in 2010 when SOLAW 2011
Desertification (UNCCD) global land outlook
was compiled, and these have been shown
(UNCCD, 2017).
as distributed data, where appropriate, to
provide a contemporary picture of land and
What are the implications for the global
water resources aggregated at the continen-
food system and the food security of the
tal regional and subregional levels used for
2.37 billion people facing moderate or severe
SOLAW 2011 (see the annex).
food insecurity? The latest report on The
state of food security and nutrition in the world
2021 (FAO et al., 2021) recognizes the sever-
ity of external drivers including conflict and
1.2 Emissions
COVID-19 containment measures, which from land and the
changing climate
constrain human engagement with produc-
tive land. This is land that produces more
than 95 percent of the global food supply
when measured in kilograms per capita per In 2019, global anthropogenic greenhouse
year (FAO, 2020a). However, the land and gas (GHG) emissions, from all economic
water systems at risk identified in the first sectors including land use, land-use change
report of the Food and Agriculture Organi- and forestry, totalled 54 billion tonnes of
zation of the United Nations (FAO) on The carbon dioxide equivalent (CO2-eq), and
state of the world’s land and water resources for emissions from agrifood systems (including
food and agriculture 2011 (SOLAW 2011; FAO, food processing and supply chain emissions)
2011) are now seeing the growth in land and amounted to some 17 billion tonnes CO2-eq
water productivity stagnate. Global datasets or 31 percent of total global emissions (FAO,
reflect a decline in per capita natural resource 2021a). Emissions from agrifood systems
availability. increased globally by 16 percent between
1990 and 2019, despite their share in total
This chapter provides a global overview of the emissions decreasing from 40 percent to
current state of land, soil and water resources 31 percent, as did the per capita emissions,
concerning agricultural production, building from 2.7 to 2.1 tonnes CO2-eq.
on the analysis in SOLAW 2011. The purpose

2 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


The 2019 total agrifood system GHG emis- emissions through changes in agricultural
sions are composed of 7.2 billion tonnes practice and land management are desir-
CO2-eq (13 percent of total global emis- able in the global effort towards achieving
sion) from activities on agricultural land (at net-zero GHG emissions.
the farm gate), 3.5 billion tonnes CO2-eq
(7 percent) from land-use change processes However, climate change is already affecting
such as deforestation and peatland degra- the human relationship with land and water.
dation, and 5.8 billion tonnes CO2-eq While some climate-induced shocks, such as
(11 percent) from pre- and post-production intense periods of heat or flood events that
processes. These processes include energy use surpass previous experience, are apparent
in fertilizer manufacturing, food processing, immediately, the slower-onset phenomena,
retail, transport and household consumption. such as elevated night-time temperatures,
Map 1.1 shows the global distribution of GHG have impacts on agricultural production that
emissions intensity from land in 2012. are more incremental in nature. The IPCC
sixth assessment report Climate change 2021:
The trends in these emissions over the past The physical science basis attributes detectable
30 years show the significance of the growth changes in the global water cycle since the
in the pre- and post-production processes middle of the twentieth century to human-
(Ffigure 1.1), while emissions from agricultural induced climate change (IPCC, 2021). Land
land and land-use change have remained and water management has played a signifi-
relatively stable. Nonetheless, agricultural cant part in triggering these changes through
land and land-use change are estimated to modified carbon and nutrient cycles, GHG
contribute 20 percent of global GHG emis- emissions, and control over the distribution of
sions. Reductions in those land-related freshwater across and within the Earth’s crust.

Map 1.1 GREENHOUSE GAS EMISSIONS INTENSITY PER UNIT OF LAND (TONNES CO2-eq PER km2)

< 10
10-25
25-50
50-100
100-250
250-500
> 500
Protein production

Dotted line represents approximately the


Source: FAO. 2013. Tackling climate change through livestock: A global assessment of emissions and Line of Control in Jammu and Kashmir
mitigation opportunities. Rome. agreed upon by India and Pakistan. The

www.fao.org/3/i3437e/i3437e.pdf final status of Jammu and Kashmir has not


yet been agreed upon by the parties.

Modified to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420 Final boundary between the Sudan and
South Sudan has not yet been determined.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 3
FIGURE 1.1 Global agrifood system GREENHOUSE GAS emissions by life-cycle stage
and per capita emissions, 1990–2019

Farm-gate emissions Land-use change Pre-and post-production Per capita emissions

18 3.0

16
2.5
14

12 2.0

Tonnes CO2eq per capita


10
Gt CO2eq

1.5
8

6 1.0

4
0.5
2

0 0.0
1990 1995 2000 2005 2010 2015 2019

Source: FAO. 2021. The share of agri-food systems in total greenhouse gas emissions: Global, regional and country trends 1990–2019. FAOSTAT
Analytical Brief Series No. 31. Rome. www.fao.org/3/cb7514en/cb7514en.pdf

The results of the IPCC sixth assessment to further land degradation and to affect
report and the special report on climate coastal forest structure and composition.
change and land point to the anticipated Sea-level rise already affects coastal erosion
evolution of the complex feedback between and salinization, leaving such areas vulnera-
the atmosphere, oceans and land (IPCC, 2019, ble to catastrophic weather events.
2021, 2022). The reports find climate change
affects the rate and magnitude of some land These short-term impacts of climate change

degradation processes and introduces new need to be considered in combination with

degradation patterns. Climate models predict long-term changes in land use and land

increasing frequency, intensity and amount management. Cropland soils are estimated

of heavy precipitation as the climate changes. to have only 20–60 percent of their potential

Rainfall that is more intense but with fewer stocks before cultivation (Lal et al., 2018),

events is combining to increase the risk of and soils under conventional agriculture

landslides, extreme erosion events and flash continue to be a source of carbon dioxide

floods. The IPCC special report on climate emissions. Peatland soil degradation and

change and land notes tropical cyclones drainage release large amounts of carbon

are already shifting towards the poles, and through decomposition, and fires in drained

the speed at which they move is slowing. peatlands accounted for about 4 percent of

Increased exposure of coastal areas to intense global fire emissions between 1997 and 2016.

and long-duration storms is expected to lead Agricultural practices also cause soils to

4 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


emit other GHGs in addition to carbon diox- requirements to complete their growth cycle,
ide, and climate change exacerbates these which can be calculated by accumulating
emissions. Soils emit nitrous oxide when temperatures above a specific threshold. The
organic and inorganic fertilizers are applied reference length of the growing season is
and when nitrogen-fixing crops are planted. therefore an important baseline to establish
They also emit methane when flooded for for crop production (Map 1.2).
rice cultivation. Hence, there is interest in
land management and conservation agricul- However, the agroclimatic context is changing

ture techniques that can halt, and, in some rapidly given the mean temperature changes

instances even reverse, the loss of SOC and observed over the past 60 years (Map 1.3).

reduce emissions of methane and nitrous Farming enterprises are adapting to new

oxide (e.g. reduced tillage with nitrogen- thermal regimes that have upset crop growth

fixing plants in crop rotations, improved stages and their supporting soil ecologies,

water management/irrigation, agrofor- with specific implications for spreading crop

estry and soil erosion control structures) disease and pests. Fundamental changes to

(IPCC, 2022). the water cycle, particularly the patterns of


rainfall and periods of drought, are forcing

1.2.1 Land and adjustment of rainfed and irrigated produc-

temperature changes
tion in particular. Under climate change,
growing periods may become longer in
At the Earth’s surface, temperatures largely boreal and arctic regions, but shorter in areas
determine what crops can be grown in any affected by extended drought periods, when
given locality. Plants have specific heat compared with current reference lengths.

Map 1.2 Reference length of the growing period , average for period 1981–2010 (days)

Days
0

+360

Dotted line represents approximately the


Line of Control in Jammu and Kashmir

Source: FAO & International Institute for Applied Systems Analysis. 2021. Global agro-ecological zones v4.0 – agreed upon by India and Pakistan. The
final status of Jammu and Kashmir has not
Model documentation. Rome. www.fao.org/nr/gaez/publications/en yet been agreed upon by the parties.

Modified to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420 Final boundary between the Sudan and
South Sudan has not yet been determined.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 5
Map 1.3 Mean temperature change, 1961–2020 (°C)

-0.02 - 0 0 - 0.7 0.7 - 1.4 1.4 - 2.1 2.1 - 2.8 2.8 - 3.5 >3.5 No data

Dotted line represents approximately the Line of Control in Jammu and Kashmir agreed upon by India and
Source: FAO. 2020. FAOSTAT. In: FAO. Pakistan. The final status of Jammu and Kashmir has not yet been agreed upon by the parties.
Rome. www.fao.org/faostat/en/#data/QC
Final boundary between the Sudan and
South Sudan has not yet been
determined.

Final status of the Abyei area is not yet determined

Modified to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420

1.2.2 Impact of weather


shorter rainfall events combined with higher
evaporation and transpiration rates will lead
events on land, soil to increased erosion from water and raindrop
and water resources impacts, and accelerated runoff and strong
winds will reduce soil moisture available for
The direct impact of weather events on
plant growth. In turn, increased incidence of
cropping, grazing and forest systems and
windstorms will accelerate soil loss. Higher
soil health is difficult to separate from the
soil surface temperatures will increase the
overall environmental outcome of land and
mineralization rate of soil organic matter
water management practices. Reduced or
(SOM) and impair the soil’s capacity to
erratic rainfall and more frequent and severe
sequester carbon and retain water and to
drought periods extend soil moisture deficits
ultimately support plant growth. Higher
on some soils but extend periods of water-
temperatures will increase evaporation and
logging on others. Heavier rains are likely to
soil salinization, particularly in arid and
increase the risk of soil erosion on cultivated
semi-arid climates.
lands, on moderate to steep slopes where
runoff rates are high, and where the land has Soils in all regions are important regulators
inadequate vegetative cover. Intensified and of climate change by virtue of their ability to

6 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


absorb and store heat, moisture and carbon.
Many soil types are affected by climate change
and influence climate change through posi-
tive feedback loops. Soil physical properties
affect how soils respond to climate change

©FAO/Lou Dematteis
and determine the soil’s capacity to main-
tain and deliver soil functions for agriculture
and sequester carbon to reduce GHGs (FAO,
2017a). Two important soil types are perma-
frost soils and peatlands. Permafrost soils,
which cover 25 percent of the northern hemi-
The long- and short-term impacts of climate
sphere, are in danger of thawing and may
change and related weather phenomena may
exacerbate warming by releasing methane,
transcend the prospects for remediation of
which is an active GHG. Thawing will increase
land and water systems that are under pres-
soil erosion, as permafrost lends stability to
sure from the level of human demand for
barren arctic slopes (Turetsky, 2019), and
food, fibre and biofuel. While the implica-
threatens industrial infrastructure, with
tions for any specific point on the Earth’s
risks of oil spills and soil contamination.
surface may be uncertain, the continuation of
Peatlands cover a modest 3 percent of the
a “no-regrets” approach to more sustainable
Earth’s ice-free landmass, yet they contain
land management and agricultural practices
30 percent of the world’s SOC. Changes in the
in the face of such uncertainty is expected to
state of peatlands resulting from fires and
be adopted at the global scale.
drainage contribute at least 5 percent of GHG
emissions (Tubiello et al., 2014).

The impacts of climate change on the water


1.3 Land-cover
cycle and renewable freshwater resources status and trends
are expected to significantly alter the agri-
cultural output and the environmental
performance of productive land and water 1.3.1 Status
systems recognized in SOLAW 2011 (FAO, The global land area, including inland waters
2011). Climate models predict decreases in and permanent snow and glaciers, amounts
renewable water resources in some regions to 14 706 million ha. Tfable 1.1 presents 11
(mid-latitude and dry subtropical regions) land-cover classes for years 2010 to 2019,
and increases in others (mainly high- together with the baseline data for 1992 using
latitude and humid mid-latitude regions). the land-cover classification of the Euro-
Even where increases are projected, there pean Space Agency Climate Change Initiative.
may be short-term shortages due to chang- The statistics do not include coastal water
ing streamflows caused by greater variabil- bodies and intertidal areas. Land cover upon
ity in rainfall. The decreases in renewable which crops are cultivated (herbaceous crops,
surface water and groundwater resources in woody crops and wetlands used for culti-
dry subtropical regions will intensify compe- vation) or available for animal husbandry
tition for water among different users. (grassland) amounted to 4 132 million ha in
2019, approximately 28 percent of the global
land area.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 7
Table 1.1 Land-use class change, 1992 and 2000–2019 (million ha)
Land-cover class 1992 2010 2011 2012 2013 2014 2015 2016 2017 2018 2019

Artificial surfaces 26 48 49 51 52 54 55 56 57 58 60
(including urban and
associated areas)

Grassland 1 773 1 796 1 799 1 800 1 801 1 802 1 802 1 801 1 801 1 810 1 813

Herbaceous crops 1 877 1 910 1 909 1 909 1 908 1 907 1 907 1 904 1 905 1 905 1 904

Woody crops 178 222 223 224 224 224 223 222 220 221 222

Shrubs and/ 202 189 189 190 190 189 189 189 189 191 193
or herbaceous
vegetation, aquatic
or regularly flooded

Shrub-covered 1 615 1 595 1 597 1 598 1 599 1 599 1 600 1 597 1 597 1 601 1 605
areas

Tree-covered areas 4 347 4 291 4 286 4 282 4 281 4 281 4 280 4 287 4 288 4 278 4 270

Sparsely natural 905 886 888 889 888 887 887 888 888 887 890
vegetated areas

Terrestrial barren 1 950 1 935 1 932 1 930 1 930 1 929 1 929 1 927 1 926 1 920 1 915
land

Inland water bodies 381 381 381 381 381 382 382 382 382 382 383

Mangroves 18 18 18 18 18 18 18 18 18 18 18

Permanent snow 1 437 1 437 1 437 1 437 1 437 1 437 1 437 1 434 1 434 1 434 1 434
and glaciers

Total land cover 14 709 14 709 14 709 14 709 14 709 14 709 14 709 14 706 14 706 14 706 14 706

Source: FAO. 2020. FAOSTAT. In: FAO. Rome. www.fao.org/faostat/en/#data/QC; using European Space Agency Climate Change Initiative Land
Cover statistics, containing annual land-cover area data for the period 1992–2019 produced by the Catholic University of Louvain Geomatics as
part of the Climate Change Initiative of the European Space Agency (version 2.0, Climate Change Initiative University of Louvin Geomatics, 2017)
and lately updated to version 2.1 under the European Copernicus programme.

Map 1.4 illustrates the global distribution of crop cover peaked in 2004, then declined
dominant land-cover classes by FAO region and plateaued from 2010 at 1 905 million ha.
using Global Land Cover Share (GLC-SHARE) Woody crops also plateaued from 2010, stabi-
data. Ffigure 1.2 shows the breakdown of these lizing at around 220 million ha. Grassland
dominant land-cover classes by SOLAW region. cover has expanded since 1992 and appears to
have stabilized at around 1 800 million ha by
1.3.2 Trends 2015, before showing a significant increase
from 2017 to around 1 813 million ha in 2019.
Since 1992, artificial surfaces (notably
urban areas and paved highways/airports) In contrast, shrub-covered areas and barren
have continued to expand, doubling from lands contracted from 2000, although
30 million ha in 2000 to almost 60 million ha in shrub-covered lands recovered from 2010.
2019 (Ffigure 1.3). Tree-covered areas declined Wetlands used for cultivation (shrubs and
significantly from almost 4 347 million ha in or herbaceous vegetation aquatic or regu-
1992 to 4 270 million ha in 2019. Herbaceous

8 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


Map 1.4 Dominant land-cover classes, GLOBAL LAND COVER SHARE data, 2010

>75% Cropland 50-75% Cropland >50% Artificial surface


>75% Tree-covered land 50-75% Tree-covered land Other land cover associations
>75% Grassland, shrubs or herbaceous cover 50-75% Grassland, shrubs or herbaceous cover Water, permanent snow, glacier
>75% Sparsely vegetated or bare 50-75% Sparsely vegetated or bare

Note: Cropland includes


herbaceous and woody crops.

Source: FAO & International Institute


for Applied Systems Analysis. 2021.
Global agro-
ecological zones v4.0 – Model
documentation. Rome. www.fao.
org/nr/gaez/publications/en

Modified to comply with UN. 2020.


Map of the World. https://www. Dotted line represents approximately the
Line of Control in Jammu and Kashmir
un.org/geospatial/file/3420
agreed upon by India and Pakistan. The
final status of Jammu and Kashmir has not
yet been agreed upon by the parties.

Final boundary between the Sudan and


South Sudan has not yet been determined.

larly flooded)1 contracted from around this trend takes account of forest expansion
203 million ha in 1992 to 190 million ha in through regeneration and afforestation (Ffig-
2019. Sparsely vegetated land and barren land ure 1.4), recent national accounts of defor-
also contracted over the same period. estation rates for conversion to grassland or
cropland are expected to reduce global cover
The global tree-covered area was estimated statistics. Indeed, most forest-cover loss is
at just over 4 269 million ha in 2019, some linked to expanding newly cultivated arable
30 percent of the total land area. The net land, while forest-cover gain is attributed to
annual forest-cover loss between 2010 and afforestation and natural forest regenera-
2020 is estimated at 4.7 million ha/year tion on abandoned arable land (FAO, 2020b).
compared with 5.2 million ha/year between More than 90 percent of the deforestation is
2000 and 2010 and 7.8 million ha/year taking place in the tropics. Between 2010 and
between 1990 and 2000 (FAO, 2020b). While 2020, of the SOLAW regions, sub-Saharan
Africa lost the largest area to deforestation,
1
FAO defines wetlands used for cultivation as areas surpassing Southern America (the previous
having free water at or on the surface for at least most
regional leader). Deforestation of primary
of the growing season. The water is sufficiently shallow
to allow the growth of a wetland crop or of natural rainforest is occurring mainly in the Amazon
vegetation rooted in the soil. This includes lowland
and Congo basins.
paddy and “bas fonds” in western Africa.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 9
FIGURE 1.2 Regional distribution of dominant land-cover classes (%)

Cultivated land Sparsely vegetated and barren land

Forestland Settlement and infrastructure

Grassland and woodland ecosystems Inland water bodies

100

90

80

70

60
Percent

50

40

30

20

10

0
a

ia

ia

ia

nd

s
a

an

nd
ic
ric

ric

ic

si

si

io
As

As

As
be

ro

la
A

tA
er
er

at
Af

Af

la
a
Eu
n

st
Am
Am

er
rib

ra

as

Is
Ze
r

ut
n

Ea
te

d
nt

c
er

ra

Ca

he

l
So

ra

Fe
es

ew

ifi
rn
rn

Ce
ha
th

ut

nt

c
d

he
he

an

Pa
or

N
Sa

an

So

Ce
ut
t
N

d
or

ss
b-

an
So

d
N

Ru
ic
Su

an
er

lia
d
rn
Am

ra
an
te

st
e
es
l

Au
ra

p
W

ro
nt

Eu
Ce

n
er
st
Ea

Sources: Based on land-cover information in FAO & International Institute for Applied Systems Analysis. 2021. Global agro-ecological zones
v4.0 – Model documentation. Rome. www.fao.org/nr/gaez/publications/en; and SOLAW regional subdivisions in FAO. 2011. The state of the world’s
land and water resources for food and agriculture: Managing systems at risk. Rome, FAO and London, Earthscan. www.fao.org/3/i1688e/i1688e.pdf

10 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


million ha million ha million ha FIGURE 1.3

0
10
20
30
40
50
60
70

0
50
100
150
200
250
1 840
1 850
1 860
1 870
1 880
1 890
1 900
1 910
1 920
1 930
19 19
9 2
19 92
92 19
19
93 19 93
93 19
19
94 19 94
94 19
19
95 19 95
95 19

Area of woody crops


19
96 19 96
96
19 19
19 97

Area of herbaceous crops


97
97 19
19 19
98 98
98
19 19
99 19 99
99 20
20 20
00 00
00
20 20
01 20 01
01 20
20 20
02 02
02
20 20
03 20 03
03
20 20 20
04 04 04
20 20 20
05 05 05
20 20 20
Artificial surfaces (including urban and associated areas)

06 06 06
20 20 20
07
Land-cover trends, 1992–2019 (million ha)

07 07
20 20 20
08 08 08
20 20 20
09 09 09
20 20 20
10 10 10
20 20 20
11 11 11
20 20 20
12 12 12
20 20 20
13 13 13
20 20 20
14 14 14
20 20 20
15 15 15
20 20 20
16 16 16

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021
20 20 20
17 17 17
20 20 20
18 18 18
20 20 20
19 19 19

11
million ha million ha million ha

12
4 220
4 240
4 260
4 280
4 300
4 320
4 340
4 360

1 575
1 580
1 585
1 590
1 595
1 600
1 605
1 610
1 615
1 620
1 750
1 760
1 770
1 780
1 790
1 800
1 810
1 820
19 19
92 92 19
19 92
19
93 93 19
19 93
19
94 94 19
19 94
19
Area of grassland

95 95 19
19 95
19
96 96 19
96
19 19
97 97 19
97

Area of tree-covered areas


19 19

Area of shrub-covered areas


98 98 19
FIGURE 1.3 (continued)

98
19 19
99 99 19
99
20 20
00 00 20
00
20 20
01 01 20
01
20 20 20
02 02 02
20 20 20
03 03 03
20 20 20
04 04 04
20 20 20
05 05 05
20 20 20
06 06 06
20 20 20
07 07 07
20 20 20
08 08 08
20 20 20
09 09 09
20 20 20
10 10 10
20 20 20
11 11 11
20 20 20
12 12 12
20 20 20
13 13 13
20 20 20
14 14 14
20 20 20
15 15 15
20 20 20
16 16 16
20 20 20
17 17 17
20 20 20
18 18 18
20 20 20
19 19 19

1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


FIGURE 1.3 (continued)

Area of shrubs and/or vegetation, aquatic or regular flooded

205

200
million ha

195

190

185

180

17
12
13

15
92
93

97

14
95
94

19
10

18
01

16
98
99
96

07
02
03

05
04

09
08
06
00

11
20

20
20
20

20
20

20
20
20
20
20
19

19
19

19
19

20
20
20

20
19
19

20
19

20
20
20
20

Area of sparsely natural vegetated areas

910

900
million ha

890

880

870

17
12
13

15
92
93

97

14
95
94

19
01

10

16

18
98
99
96

02
03

07
05
04

09
08
00

06

11
20

20
20
20

20
20

20
20
20
20

20
19

19
19

19
19

20

20
20

20
19
19

20
19

20
20
20
20

Area of terrestrial barren land

1 960
1 950
1 940
million ha

1 930
1 920
1 910
1 900
1 890
17
12
13

15
92
93

97

14
95
94

19
10

16

18
98
99

01
96

07
02
03

05
04

09
08
06
00

11
20

20
20
20

20
20

20
20
20
20
20
19

19
19

19
19

20
20
20

20
19
19

20
19

20
20
20
20

Source: FAO. 2020. FAOSTAT. In: FAO. Rome. www.fao.org/faostat/en/#data/QC

1.4 Land-use trends Table 1.2 lists the land-use categories that
are used to capture land productivity at the
global level. These categories are reported
Land used for all agricultural uses was about at the national level and compiled in the
4 752 million ha in 2019. This reflects an FAO Statistical Database (FAOSTAT) to
overall decline in land use since 2000 (Ffig- form the statistical framework for report-
ure 1.5), mainly attributed to a decline in ing agricultural statistics. Land-use classes
permanent pastures and meadows used for conform with the mapping land-use types used
livestock husbandry. in Global Agro-Ecological Zones (GAEZ) v4.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 13
FIGURE 1.4 Forest-cover trends, 1990–2020 (MILLION ha PER YEAR)

Forest expansion Deforestation

15
10
10 8
7
Million ha per year

5
5

-5

-10
-10
-12
-15
-15
-16
-20
1990–2000 2000–2010 2010–2015 2015–2020

Source: FAO. 2020. Global forest resources assessment 2020: Key findings. Rome. https://doi.org/10.4060/ca8753en

Land-cover classifications (specifically The overall distributions of land use and


GLC-SHARE) and areas equipped for irriga- farming systems identified in SOLAW 2021
tion (Global Map of Irrigated Areas; GMIA remain broadly the same as those compiled
v5) guide the distribution in order to down- for SOLAW 2011 at global level (Map 1.5),
scale the production statistics. For cropped and the land-use statistical trends to 2019
land, the production area statistics are based are given in Ffigure 1.6. Since 2000 when
on reported production volumes, cropping the original farming system descriptions
intensities and yields to derive harvested were compiled, aggregate land use for all
areas for specific crops under rainfed or irri- forms of agriculture (except aquaculture) has
gated conditions.

FIGURE 1.5 Agricultural land-use trends, 1961–2019 (Million ha)

5 000

4 900

4 800

4 700
million ha

4 600

4 500

4 400

4 300

4 200
61
63
65
67
69
71
73
75
77
79
81
83
85
87
89
91
93
95
97
99
01
03
05
07
09
11
13
15
17
19
20
19

20
20
20
19

20
19
19

20
19
19
19

19

19
19
19
19

19
19
19

19

20
19

20
20
19
19
19

20

Source: FAO. 2020. FAOSTAT. In: FAO. Rome. www.fao.org/faostat/en/#data/QC

14 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


Table 1.2 Global land-use class extent, 1992, 2000, 2010–2019 (million ha)
Land-use class 1992 2000 2010 2011 2012 2013 2014 2015 2016 2017 2018 2019

Country area 13 435 13 437 13 461 13 459 13 462 13 486 13 487 13 487 13 487 13 487 13 487 13 497

Land area 12 997 13 005 13 009 13 019 13 019 13 018 13 018 13 020 13 028 13 028 13 028 13 030

Land under permanent 3 343 3 387 3 301 3 268 3 258 3 247 3 247 3 223 3 219 3 247 3 234 3 196
meadows and pastures

Arable land 1 368 1 493 1 361 1 370 1 378 1 380 1 381 1 383 1 387 1 396 1 395 1 383

Land under permanent 118 134 158 161 163 164 164 165 166 170 170 170
crops

Cropland 1 486 13 437 1 520 1 534 1 544 1 546 1 547 1 551 1 556 1 568 1 568 1 556

Agricultural land (total of 4 829 4 880 4 820 4 802 4 801 4 793 4 795 4 774 4 775 4 815 4 801 4 752
cropland and permanent
Pasture)

Land area equipped for 264 289 322 325 329 332 333 335 337 338 339 342
irrigation

Inland waters 435 430 450 437 437 438 437 436 428 429 428 427

Forestland 4 221 4 158 4 106 4 102 4 097 4 093 4 088 4 084 4 081 4 074 4 069 4 064

Naturally regenerating 4 033 3 937 3 834 3 825 3 817 3 809 3 801 3 792 3 787 3 778 3 771 3 763
forest

Planted forest 187 220 271 275 278 282 286 290 292 294 297 299

Other land 3 947 3 968 4 060 4 102 4 102 4 103 4 098 4 138 4 138 4 111 4 133 4 188

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021
Source: FAO. 2020. FAOSTAT. In: FAO. Rome. www.fao.org/faostat/en/#data/QC

15
remained stable at around 4 800 million ha. in soil structure and fertility and affect how
But this masks a significant decline in soils respond hydrologically. Notably, the
permanent meadows and pastures since proportion of land equipped for irrigation to
2000 (net loss of 191 million ha) and the cropped land rose from 19.4 percent in 2000 to
continued increase in cropland (temporary almost 22 percent in 2018. The conversions of
and permanent crops) of some 100 million ha forested land to cropped land in the Amazon
over the same period. The net forested area and Congo basins are notable examples of the
continues to decline (by about 94 million ha scale of change. The aggregate impact of local
since 2000), although there have been slight changes in oil palm plantations or draining
increases in the planted forest. organic soils to convert wetland to cropped
land in Southeast Asia can be masked by
At the global level, changes in overall land use classification shifts. For example, plantation
appear small, but at country and local levels, development can register as a gain in forested
shifts in land use and agricultural practices land and permanent crops. Also the land
are significant. These changes trigger losses registered as equipped for irrigation does not

Map 1.5 Major agricultural systems IN 2010

Irrigated crops: other than paddy rice Rainfed agriculture: humid tropics Rainfed agriculture: highlands Rangelands: boreal Other land

Irrigated crops: paddy rice Rainfed agriculture: sub-tropics Rangelands: sub-tropics Desert

Rainfed agriculture: dry tropics Rainfed agriculture: temperate Rangelands: temperate Forest

Dotted line represents approximately the


Line of Control in Jammu and Kashmir
agreed upon by India and Pakistan. The
final status of Jammu and Kashmir has not
Source: FAO. 2011. The state of the world’s land and water resources for food and agriculture: Managing yet been agreed upon by the parties.

systems at risk. Rome, FAO and London, Earthscan. www.fao.org/3/i1688e/i1688e.pdf


Final boundary between the Sudan and
Modified to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420 South Sudan has not yet been determined.

16 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


million ha million ha million ha
FIGURE 1.6

1 200
1 220
1 240
1 260
1 280
1 300
1 320
1 340
1 360
1 380
1 400

4 200
4 300
4 400
4 500
4 600
4 700
4 800
4 900
5 000
1 200
1 250
1 300
1 350
1 400
1 450
1 500
1 550
1 600
19 19 19
61 61 61
19 19 19
63 63 63
19 19 19
Cropland area

65 65 65
19 19 19

Arable land area


67 67 67
19 19 19
69 69 69

Agricultural land area


19 19 19
71 71 71
19 19 19
73 73 73
19 19 19
75 75 75
19 19 19
77 77 77
19 19 19
79 79 79
19 19 19
81 81 81
19 19 19
Land-use trends (million ha)

83 83 83
19 19 19
85 85 85
19 19 19
87 87 87
19 19 19
89 89 89
19 19 19
91 91 91
19 19 19
93 93 93
19 19 19
95 95 95
19 19 19
97 97 97
19 19 19
99 99 99
20 20 20
0 0 0
20 1 20 1 20 1
03 03 03
20 20 20
05 05 05
20 20 20
0 0 0
20 7 20 7 20 7
09 09 09
20 20 20
11 11 11
20 20 20
13 13 13

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021
20 20 20
15 15 15
20 20 20
17 17 17
20 20 20
19 19 19

17
million ha million ha million ha

18
0
20
40
60
80
100
120
140
160
180

4 000
4 200

3 950
4 050
4 100
4 150
4 250
2 950
3 000
3 050
3 100
3 150
3 200
3 250
3 300
3 350
3 400
3 450
19 19 19
90 61 61
19 19 19
91 63 63
19 19 19
92 65

Forestland area
65
19 19 19
93 67 67
19 19 19
94 69 69
19 19 19
95 71 71
19 19 19
96 73 73
FIGURE 1.6 (continued)

19 19 19
97 75 75
19 19 19

Land under permanent crops area


98 77 77
19 19 19
99 79 79
20 19 19
00 81 81
20 19 19
0 83 83
20 1 19 19
02 85 85
20 19 19
03 87 87
Land under permanent meadows and pasture areas

20 19 19
04 89 89
20 19 19
0 91 91
20 5 19 19
06 93 93
20 19 19
0 95 95
20 7 19 19
08 97 97
20 19 19
09 99 99
20 20 20
10 0 0
20 20 1 20 1
11 03 03
20 20 20
12 05 05
20 20 20
13 0 0
20 20 7 20 7
14 09 09
20 20 20
15 11 11
20 20 20
16 13 13
20 20 20
17 15 15
20 20 20
18 17 17
20 20 20
19 19 19

1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


FIGURE 1.6 (continued)

Land area equipped for irrigation

400

350

300
million ha

250

200

150

100

50

15
13
71

91

11
73
75
77

81
61

79

83
85
87

93
95
97
89

99

20 1
63
65
67
69

20 3
05

20 7

17
09

19
0

0
0

20
20

20
19

20
19

20
19
19

20
19
19
19

19

19
19
19

19
19
19

19
19

20
19

19
19
19

Source: FAO. 2020. FAOSTAT. In: FAO. Rome. www.fao.org/faostat/en/#data/QC

imply this is the irrigated land area that will when SOC can be mineralized, although some
be recorded as the harvested area in any one land may not be permanently degraded and
calendar year as this is a function of cropping can be brought back into cultivation after long
intensity. periods of fallow. National data on the extent
of rainfed farming systems affected by land

1.4.1 Rainfed agriculture degradation are limited; it is therefore diffi-

and the impact of drought cult to estimate the precise areas involved.

Rainfed agriculture is the predominant The most productive rainfed cropping occurs

agricultural production system worldwide. in the temperate zones of Northern America

Strictly defined, it depends exclusively on and Europe, and in the subtropics and humid

rainfall for crop production, with no perma- tropics. Rainfed cropping in highland areas

nent source of irrigation. In 2018, the world and the dry tropics tends to be relatively low

cultivated area was 1 557 million ha, of which yielding, with low-input practices associated

1 221 million ha (78 percent) was rainfed, with subsistence farming. Trends in rain-

producing about 60 percent of global crop fed areas differ regionally. In sub-Saharan

output in a wide variety of production systems. Africa, where 97 percent of staple production
is rainfed, the area of cereals has doubled
The areal extent of productive rainfed crop- since 1960. In Central America and the Carib-
land has not changed significantly since bean, rainfed cultivation has expanded by
the middle of the twentieth century, but 25 percent in the last 40 years.
this masks the extent to which land newly
converted from forests and grasslands to The focus on dryland systems at the end

arable farming has replaced degraded and of Chapter 4 discusses the combination of

abandoned land. The risks of resource degra- drought impacts in dry lands where, even

dation are high during periods of drought during regular seasonal cycles, increases in

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 19
108 million ha, 33 percent of the equipped
area (Tfable 1.3). Over the same period, land
equipped for irrigation has increased from
©FAO/Giulio Napolitano

10 percent of the total cultivated land to


21 percent. Since 2010, equipped areas have
exhibited little or no growth in reported
statistics, even as the global production of
irrigated crops continues to increase. This may
be due to changes in the pattern of produc-
livestock traffic and cropping intensities can tion such as: increased cropping intensities
lead to rapid deterioration in soil fertility, and yields on existing continuously irrigated
biodiversity and soil structure, leaving large areas; infilling of gaps between equipped
swathes of semi-arid subtropical land prone areas and actually irrigated areas (areas
to degradation. harvested); and production from areas not
registered in national statistics as “equipped

1.4.2 Irrigated agriculture for irrigation”. The latter may reflect informal
and temporary irrigation systems or simply
Irrigation plays a significant role in secur- land that is equipped and not reported. This
ing food supplies and supporting economic is particularly the case in the Near East and
development in many countries. Its impor- Arabian Peninsula subregions, which have
tance is likely to grow, given the impacts experienced dramatic increases in livestock
of climate change. Irrigated production is production derived from irrigated fodder and
responsible for approximately 40 percent of the expansion of vegetable and citrus produc-
agricultural output (FAO et al., 2018). Land tion under protected cover, including tempo-
equipped for irrigation can stabilize the rary and permanent shade and greenhousing.
production of high-value crops, particularly Downscaling national statistics is improving
eliminating the risk of unreliable rainfall, due to the use of higher-resolution and cali-
but, more importantly, delivering adequate brated remote-sensing techniques, such as
soil moisture at the right time to maximize the moderate-resolution imaging spectro-
yield response. Irrigation in combination radiometer platform using the normalized
with drainage offers an important adaptation difference vegetation index (NDVI).
strategy to combat drought and flooding risk
as the climate changes. In 2018, the Asian continent had 70 percent
of the world’s area equipped for irrigation,
Land area equipped for irrigation2 (includ- mainly in the South Asia and East Asia subre-
ing all full water control irrigation systems, gions, and 32 percent of the cultivated area
equipped wetlands and spate irriga- (Tfable 1.3, Ffigure 1.7 and Ffigure 1.8). Africa,
tion) has almost doubled over the past particularly sub-Saharan Africa, had the
60 years, from 139 million ha in 1961 to smallest equipped area in 2018, accounting
over 328 million ha in 2018, with groundwa- for only 3 percent of global irrigated land.
ter-sourced irrigation accounting for some Irrigation is essential in Northern Africa,
representing 27 percent of the cultivated area.
2
The area equipped for irrigation refers to the area Countries with the largest land area equipped
equipped to provide water – via irrigation – to crops. for irrigation were China (70 million ha), India
It includes areas equipped for full control irrigation and
partially controlled irrigation (equipped lowland areas (70 million ha), the United States of America
and areas equipped for spate irrigation). (27 million ha) and Pakistan (20 million ha).

20 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


Ffigure 1.9 illustrates trends in irrigated areas responsible for a 2 percent annual increase
as a percentage of cultivated area in 2000, in the land area equipped for irrigation. By
2012 and 2018. the 1980s, this slowed to less than 1 percent.
As a percentage of the cultivated area, irriga-
In the 1960s and 1970s, the push to intensify tion increased in almost all regions, mainly
under the Green Revolution was primarily where irrigated agriculture dominated, such

Table 1.3 Areas equipped for irrigation, 1961, 2012 and 2018

Continent, region Equipped area Equipped area as Groundwater


(million ha) a percentage of irrigation (2018)
cultivated area (%)

Year 1961 2012 2018 1961 2012 2018 Area Percentage


equipped of total
(million ha) irrigated
area (%)

Africa 7.4 15.2 15.8 4.4 5.6 5.7 3.0 19.0

Northern Africa 3.9 7.3 7.6 17.1 25.5 26.3 2.3 31.0

Sub-Saharan Africa 3.5 7.9 8.2 2.4 3.2 3.3 0.7 9.0

Americas 22.6 51.3 53.6 6.7 14.0 14.5 22.0 41.0

Central America and 17.4 2.1 2.2 6.7 14.4 14.7 0.4 20.0
Caribbean

Northern America 0.6 33.6 35.2 5.5 15.2 15.9 20.2 57.0

Southern America 4.7 15.5 16.2 6.8 11.9 12.3 1.5 9.0

Asia 95.6 231.8 232.9 19.6 39.9 39.4 79.0 34.0

Central Asia 9.6 13.5 12.7 16.2 29.3 27.3 0.9 7.0

East Asia 7.2 73.8 74.8 13.4 50.5 51.4 19.7 26.0

South Asia 36.3 96.6 97.2 19.1 45.2 45.4 47.8 49.0

Southeast Asia 34.5 22.6 22.9 29.7 19.8 18.6 0.9 4.0

Western Asia 8.0 25.3 25.4 11.7 42.0 41.3 9.5 38.0

Europe 12.3 22.0 23.0 3.6 7.6 8.0 3.2 14.0

Eastern Europe and 8.7 6.3 6.3 5.8 3.9 3.8 0.5 8.0
Russian Federation

Western and Central 3.6 15.6 16.7 1.9 12.3 13.4 2.8 17.0
Europe

Oceania 1.1 3.0 3.2 3.2 9.1 9.3 1.1 6.0

Australia and New 1.1 3.0 3.0 3.2 9.3 9.5 1.1 6.0
Zealand

Pacific Islands 0.0 0.0 0.0 0.2 0.5 0.6 0.0 12.0

World 139.0 323.3 328.3 10.2 20.9 21.1 108.3 33.0

High income 26.7 53.9 55.2 6.9 15.3 15.6 22.2 40.3

Low and middle 66.6 269.4 273.2 23.6 22.6 22.7 85.1 31.2
income

Source: FAO. 2021. AQUASTAT – FAO’s Global Information System on Water and Agriculture. In: FAO. Rome.
www.fao.org/aquastat/en

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 21
FIGURE 1.7 Global distribution of irrigated surfaces by geographical region, 2018

Sub-Saharan Africa - 2%

Northern Africa - 2%
Oceania - 1%

Central America and Caribbean - 1%


Europe - 7%

Western Asia - 8% Northern America - 11%

Southern America - 5%
Southeast Asia - 7%

Central Asia - 4%

South Asia - 29% East Asia - 23%

Source: FAO. 2021. AQUASTAT – FAO’s Global Information System on Water and Agriculture. In: FAO. Rome. www.fao.org/aquastat/en

FIGURE 1.8 Area equipped for irrigation by geographical region, 2000, 2012 and 2018
(million ha)

2000 2012 2018

120.0

100.0
Million hectares

80.0

60.0

40.0

20.0

0.0
s
tio d

d
a

an a

ia

ia

ia
a

nd
ric

be ric

ic

ic

si

si

op
ra an

an
ric

As

As

As

n
lA

tA
er

er

la
Af

rib e

al
ur
Af

de e
st

n
Ca Am

Am

Am

Is
ra

as

Ze
Fe op
ut

lE
n

Ea

te
an

nt

fic
er

he
So

ra
es

ew
d al

rn

ia Eu
Ce
ar
th

ci
er

ut

nt
an ntr

W
he
ah

Pa
or

N
ss rn
th

So

ce
Ce

ut
-S
N

d
n
or

Ru ste

an
d
So
b

an
Su

Ea

lia
rn

ra
te

st
es

Au
W

Source: FAO. 2021. AQUASTAT – FAO’s Global Information System on Water and Agriculture. In: FAO. Rome.
www.fao.org/aquastat/en

22 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


FIGURE 1.9 Irrigated area as a percentage of cultivated area, 2000, 2012 and 2018 (%)

2000 2012 2018

60

50

40
Percentage (%)

30

20

10

0
ca

ric n

an a

pe

a
be ric

ic

ic

si

si

ni
Af ara

As

As

As
a
ri

ro
lA

tA
er

er

a
Af

rib e

ce
h

st

Eu
Ca Am

Am

Am

ra

as
Sa

ut

r
rn

Ea

O
te
nt

he
So
b-
e

es
d al

rn

Ce
th

er

ut
Su

an ntr

W
he
or

th

So
Ce

ut
N

or

So
N

Source: FAO. 2021. AQUASTAT – FAO’s Global Information System on Water and Agriculture. In: FAO. Rome. www.fao.org/aquastat/en

as Northern Africa, South Asia and East Asia. water demand for municipal and industrial
Globally, the annual growth rate has slowed uses, declining freshwater sources and grow-
to less than 0.5 percent, but this is based ing concerns for the aquatic environment
on reported statistics only. The develop- are also constraining growth. A contributing
ment of new irrigated areas is evident from factor in the 1980s was the loss of many
imagery and moderate-resolution imaging large schemes in Eastern Europe and the
spectroradiometer data (FAO Water Produc- former Soviet Union that proved unprofit-
tivity Open-access Portal (WaPOR); FAO, able, and which were unable to adapt and
2022a), particularly the growth of centre meet the requirements of commercial new
pivot installations, with each pivot reach market-oriented private commercial farming
averaging 50 ha. This recent expansion of (Siebert and Döll, 2007).
centre pivot installations is apparent mainly
in the Nile basin and Arabian Peninsula where Since 2000, investments have moved from

high demand for irrigated fodder is concen- developing large irrigation infrastructure,

trated (Alhumaid, 2020). including dams, reservoirs and large irriga-


tion systems, to improving on-farm irriga-
Some of the reasons for the overall decline tion systems and including microirrigation
in growth rates include increasing irriga- methods and more effective management
tion development costs, reduced government practices. The area equipped for micro-
support and financing, ageing infrastruc- irrigation (drip lines and sprinklers) in 2018
ture and lack of maintenance. But increasing covered almost 70 million ha (21 percent)

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 23
Asia. Irrigation continues to stabilize agri-
cultural output, raising cropping intensity
and encouraging farmers to grow high-value
crops (Fuglie et al., 2020).

In the baseline year of 2012, some


©FAO/Sheam Kaheel

346 million ha of irrigated production was


harvested on 261 million ha of land equipped
for irrigation, indicating a global cropping
intensity of 130 percent (FAO, 2018a). Cere-
als accounted for over 60 percent of the
harvested area, vegetables for 10 percent,
of the total equipped area. The adop-
fodder for 7 percent, oil crops for 7 percent,
tion of precision irrigation is associated
fruit for 6 percent, fibre for 5 percent and
with the uptake of protected cropping in
sugar for 4 percent.3 However, proportions
industrial-scale glasshouses and areas of
are changing in response to increasing
shade netting, particularly for high-value
demands for different products, particularly
horticultural crops. Many of these installa-
animal protein.
tions are visible on satellite imagery in the
Mediterranean basin and the Near East.
Land and water productivity under irrigation
presents a mixed picture. Ffigure 1.10 illus-
Existing irrigation schemes contend with
trates the impact of irrigation on vegetable
salinity and pollution build-up generated
yields by region based on production statis-
by decades of maladapted drainage and soil
tics compiled in 2012.
management practices. Options to manage
salinity are becoming limited in areas where
Map 1.6 shows economic water productivity
aridity is increasing (see section 1.5.3 on soil
for rice. In Asia, subtropical climates and
salinization).
short growing seasons favour higher crop-
ping intensities.
In addition to water scarcity, the availability
of suitable land for irrigation expansion is
Map 1.7 illustrates economic water produc-
another constraint as urban areas expand
tivity for wheat. For instance, productivity
and encroach on land previously dedicated to
is low in Punjab, India, because of the large
irrigated production. Fragmentation of land
volumes of water required to grow wheat.
holdings and increases in land prices also
This contrasts with high productivity in
inhibit the development of contiguous areas
Western Europe, where water consumption is
of formal production (Lowder, Sánchez and
lower, and fertilizer and pesticide inputs are
Bertini, 2019).
relatively high.

Irrigated agricultural crops typically yield


at least twice that of nearby rainfed crops. 3
Note that FAOSTAT (FAO, 2020a) data show crop
production and may differ from the FAO Global
Rainfed cereals yield on average 1.5 tonnes/ha Information System on Water and Agriculture
in the developing countries, whereas irri- (AQUASTAT) data (FAO, 2021b). In some regions, the
same area is harvested twice in the same year and so
gated cereals yield on average 3.3 tonnes/ha. some areas are counted twice. Also, the actual area
Irrigated cropping intensities are typically irrigated must tally with production data and is always
less than the area equipped for irrigation as given in
higher, with two crops per year in most of AQUASTAT.

24 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


FIGURE 1.10 Vegetable yields by region, 2012 (tonnes/ha)

Rainfed Irrigated

35

30

25
tonnes/ha

20

15

10

0
ci ia

Ch acifi ia

ia

ric d

ld
be nd
in

si

ric
Af an
Pa As

As

or
fic

in c

an

a
ng P A

lA
Ch

rib a a

Af

W
th st
e n

di e n

rn
a)

ra
th er

lu th ter

Ca ic

or a

n
he

nt

N rE

ra
e er
d st

xc nd as

ut

Ce

ha
th m
an Ea

ea
a E

So

Sa
N
d

tin
an

b-
La

Su
pe
ro
(e

Eu

Source: FAO. 2020. The state of food and agriculture 2020. Overcoming water challenges in agriculture. Rome.
https://doi.org/10.4060/cb1447en

Map 1.6 Economic water productivity for rice, average 1996–2005 (USD/m3)

≤ 0.05

0.05 - 0.075

0.075 - 0.1

0.10 - 0.15

0.15 - 0.2

0.2 - 0.25

0.25 - 0.35

> 0.35

Dotted line represents approximately the


Note: Economic water productivity is defined as crop USD value per unit of water consumed (total evapotranspiration over the crop
Line of Control in Jammu and Kashmir
growing season). Values are converted from physical water productivity (kg/m3) to economic water productivity (USD/m3) using the agreed upon by India and Pakistan. The
average global price of each crop from FAOSTAT, average for period 1981–2010 (days). final status of Jammu and Kashmir has not
yet been agreed upon by the parties.

Source: FAO. 2020. The state of food and agriculture 2020. Overcoming water challenges in agriculture. Rome.
Final boundary between the Sudan and
https://doi.org/10.4060/cb1447en. Modified to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420 South Sudan has not yet been determined.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 25
Map 1.7 Economic water productivity for wheat, average 1996–2005 (USD/m3)

≤ 0.05

0.05 - 0.075

0.075 - 0.1

0.10 - 0.15

0.15 - 0.2

0.2 - 0.25

0.25 - 0.35

> 0.35

Note: Economic water productivity is defined as crop USD value per unit of water consumed (total evapotranspiration over the crop Dotted line represents approximately the
Line of Control in Jammu and Kashmir
growing season). Values are converted from physical water productivity (kg/m3) to economic water productivity (USD/m3) using the
agreed upon by India and Pakistan. The
average global price of each crop from FAOSTAT. final status of Jammu and Kashmir has not
yet been agreed upon by the parties.

Source: FAO. 2020. The state of food and agriculture 2020. Overcoming water challenges in agriculture. Rome.
Final boundary between the Sudan and
https://doi.org/10.4060/cb1447en. Modified to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420
South Sudan has not yet been determined.

The most productive areas equipped for irri- drought in southern Australia affected large
gation are in broad alluvial plains, deltas and areas, but even with reduced water alloca-
coastal margins in subtropical climates with tions, productivity levels on irrigated land
high evaporation rates, monsoonal rainfall, were sustained (Hatfield-Dodds et al., 2018).
inundation and susceptibility to salinization.
In these irrigated systems, the annual crop 1.4.3 Land for livestock
production cycle is highly conditioned by production
climatic volatility – prolonged periods of
drought and higher-frequency intensified The rapid expansion of animal protein
rainfall and associated flooding. The land’s consumption in the latter half of the twen-
ability to recover from flooding to maintain tieth century is a feature of global food
cropping calendars is an important element production assessments (FAO, 2017a, 2020c).
of the resilience of irrigated farming systems. Global land use dedicated to livestock
In the Indus basin, the July–September flood production (permanent pastures and mead-
event in 2010 inundated at least 3.7 million ha ows) peaked around 2000 at 3.4 billion ha,
of productive irrigated floodplain, disrupting and has since declined to around 3.2 billion ha
rice food systems and industrial crops such as in 2019 (Ffigure 1.11). The increase in animal
cotton, well into 2011 (NASA, 2011). The 2018 protein consumption in the twentieth century

26 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


is reflected in the expansion of pasture places pressure on in situ water resources and
before 2000. The subsequent decline reflects soil resources for intensive feed and forage
the increase in livestock productivity and production. Some 12 percent of irrigation
stocking intensity including growth in zero- water withdrawals is attributable to fodder
grazing feedlots (FAO, 2019a). crops (FAO, 2019a). Concentrating inputs and
animal waste have resulted in energy use
Livestock production (including pasture, from fossil fuels, methane emissions and
rangeland and cropland for feed) represents point-source water pollution from nutrients
almost 80 percent of all agricultural land, and antibiotics. Livestock and forest options
with feed production taking up roughly also need analysing for a complete land-use
one-third of total cropland. Yet, the grass- scenario development for future produc-
land and shrub-covered areas used to graze tion systems.
animals or as sources of fodder have signifi-
cantly declined by 191 million ha over two Rural communities living in dry lands have
decades to an area of 3 196 million ha in developed agricultural systems and practices
2019, due to pressures of converting to crop- that are adapted to arid, semi-arid and subhumid
land. Moreover, 13 percent of the grassland conditions and drought risk over generations of
area has degraded due to high anthropo- experience. These populations that depend on
genic pressures, and 34 percent has reduced limited land potential and water resources have
biophysical status, notably due to overgraz- developed mixed crop–livestock systems based
ing and inadequate livestock mobility causing on short-season drought-resilient crops and
soil compaction and erosion, thus affecting receding floodwaters alongside wetlands and
soil function, plant growth and hydrological river plains. They can provide lessons for
services. countries recently experiencing water short-
age and drought due to climate change.
Intensive livestock production has grown
rapidly to meet expanding meat demand Overgrazing is also degrading grasslands,
in middle- to high-income countries. This and erosion rates are increasing where over-

FIGURE 1.11 Global land use under permanent meadows and pastures, 1961–2019
(Million ha)

3 450
3 400
3 350
3 300
million ha

3 250
3 200
3 150
3 100
3 050
3 000
2 950
71

91

11
73
75
77

81

93
95
97

13
15
17
61

79

83
85
87
89

99

20 1

19
63
65
67
69

03
05

20 7
09
0

20
19

20
20
20
19

20
19
19

20
19
19
19

19

19
19
19

19
19
19

19
19

20
19

20
19
19
19

Source: FAO. 2020. FAOSTAT. In: FAO. Rome. www.fao.org/faostat/en/#data/QC

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 27
grazing occurs (Pimentel and Burgess, 2013). 176 million tonnes CO2-eq of sequestered
Protecting and conserving grasslands require emissions (net of increased nitrous oxide
the fostering of crop rotations and seeded emissions) annually would be possible by
pastures in cropped grassland areas and sowing legumes in some grassland areas.
improvements in grazing management. Thus, a combined mitigation potential of
585 million tonnes CO2-eq is estimated,
The declining trend in land use for live- representing about 8 percent of livestock
stock production may flatten out as limits to supply-chain emissions (FAO, 2013a).
productivity are approached in some regions,
and demand for animal protein is saturated The current pattern of dryland management
(FAO, 2018b). Land used for livestock produc- is responsible for losing significant amounts
tion (including pastureland, rangeland and of carbon, driven mostly by increasing human
cropland) represents almost 80 percent of and livestock pressures. Dryland soils tend to
all agricultural land, with feed and fodder be low in carbon due to limited replenish-
production taking up roughly one-third of ment and loss to mineralization of humic
total cropland (FAO, 2018b). complexes when dehydrated. Thus, they are
susceptible to degradation by mechanical
Intensive livestock production is apparent erosion (wind and water), but their potential
where higher livestock densities occur and to sequester carbon may be high. Estimates
feed and water inputs are concentrated, suggest that by 2030, improved rangeland
placing pressure on in situ water resources management has the biophysical potential
(notably groundwater) and soil resources to sequester 1.3–2.0 billion tonnes CO2-eq
for higher forage production rates. Adopting worldwide (Tennigkeit and Wilkes, 2008).
zero-grazing or feedlot systems in semi-arid
and humid zones has reduced soil compac-
1.4.4 Inland fisheries
and aquaculture
tion and poaching from grazing livestock.
However, concentrated inputs and animal
waste have resulted in higher point-source The growth in freshwater fish capture and
water pollution from nutrients and antibiot- aquaculture as the dominant form of fish
ics. Additional demand for imported feed and production is significant, signalling the
forage production is significant, particularly conversion of freshwater habitats with some
for high-protein feed crops such as soya. using saline–alkaline water to raise marine
Estimates suggest that up to 12 percent of irri- species (FAO, 2020d). Significant regional
gation water withdrawals may be attributed differences in production levels reflect the
to fodder crops (Mekonnen and Hoekstra, distribution of freshwater habitats and
2012). However, in practice, national report- geographical gradients in climate, geology,
ing of fodder crop production has proved so land use, biodiversity, human population
unreliable and inconsistent that FAOSTAT no density and economic activity. Ffigure 1.12
longer reports fodder production. exhibits the growth in fish capture produc-
tion by producers in South Asia and East Asia.
Projected grassland soil carbon seques-
Map 1.8 presents the distribution of inland
tration is significant. Estimates suggest
fish capture in relation to major river basins.
improved grazing management practices
could sequester 409 million tonnes CO2-eq Inland fisheries are subject to impacts
of carbon annually on pastureland. A further from a range of human-induced drivers

28 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


FIGURE 1.12 Top five inland waters fish capture producing countries, 1999–2018
(THOUSAND tonnes)

China India Bangladesh Myanmar Cambodia

2 500

2 000
Thousand tonnes

1 500

1 000

500

0
99

00

01

02

03

04

05

06

07

08

09

10

11

12

13

14

15

16

17

18
20

20

20
20

20
20

20
20
20

20
20

20
20

20
20
19

20
20
20
20

Source: FAO. 2020. The state of world fisheries and aquaculture: Sustainability in action. Rome.
https://doi.org/10.4060/ca9229en

Map 1.8 Estimated inland fishery catch as a percentage of the global inland
2007–2016 catch WORK INSIDE 15CM

Percent
0 - 0.1
0.1 - 0.25
0.25 - 0.5
0.5 - 1
1-3
3-6
6-9
9 - 12
12 - 16

Legend: White = no significant catch; lightest green = < 0.1 percent and darkest green = 12–16 percent of the global total inland fishery catch.
Note: Retained recreational catches are not included.

Source: FAO. 2020. The state of world fisheries and aquaculture: Sustainability in action. Rome. https://doi.org/10.4060/ca9229en
Modified to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 29
Inland aquaculture increased from less than
1 million tonnes of annual live weight produc-
tion globally in 1950 to 51.3 million tonnes
in 2018. It contributes 62.5 percent of the
world fish production (FAO, 2020d). Rice–
fish culture, often operating at a family scale
with renovated paddy fields, has expanded
© L. Miuccio

rapidly among rice farmers in China in recent


decades. In 2008, an estimated 1.5 million ha
of rice fields was used for aquaculture (FAO,
including competition for freshwater from 2020a). Cage aquaculture in freshwater lakes
irrigation and agriculture and impacts on and rivers has flourished in many coun-
habitat connectivity caused by water control tries as an efficient non-consumptive use
infrastructure. This includes dams and flood of freshwater. Asia (especially China) has
protection, water regulation and pollution the greatest freshwater aquaculture produc-
caused by runoff and drainage. Pekel et al. tion in terms of land and water surface area.
(2016) note that between 1984 and 2015, Fish production in coastal and offshore
90 000 km of permanent water bodies has
2
marine environments offers alternative and
vanished altogether and over 72 000 km 2
new aquaculture opportunities when fresh-
has transitioned from permanent water bodies water and land become scarce (Kapetsky,
to seasonal waterbodies disconnected for Aguilar-Manjarrez and Jenness, 2013).
periods of the annual water cycle. This
disruption of natural surface water area and
1.4.5 Forestland
connection has diminished the potential for
fish capture and aquaculture, while pollu- The 4 060 million ha of forest cover assessed
tion from land-based source nutrients and in 2020 comprises tropical (45 percent),
pesticides has resulted in eutrophication of boreal (27 percent), temperate (16 percent)
fish habitats and high rates of fish mortality and subtropical (11 percent) forests (Map 1.9)
(Funge-Smith, 2018). (FAO, 2020c). About 1 150 million ha is
managed primarily for producing wood and
Inland capture fisheries produce some non-wood forest products. Since 1990, the
12 million tonnes of fish annually (12.5 percent area of multiple forest use has declined by
of the total global capture fishery production), 71 million ha, to 749 million ha in 2015.
which is sufficient to meet the animal protein Multiple use includes crop production,
requirements of more than 160 million people which has been relatively stable since
(FAO, 2020d). More than 40 percent of the 1990. The total carbon stock in forests
world’s inland fish capture harvest comes declined from 668 billion tonnes in 1990
from 50 low-income food-deficit countries. to 662 billion tonnes in 2020, while carbon
Inland fish provide nutritional quality to diets density increased slightly over the same
that are otherwise poor in nutrients, minerals period, from 159 tonnes/ha to 163 tonnes/ha.
and vitamins. Emissions of GHGs associated Most forest carbon is in the living biomass
with inland capture fisheries are relatively (44 percent) and SOM (45 percent), with the
small compared to livestock and rice produc- remainder in dead wood and litter.
tion, as most fish harvested are consumed
or sold locally, and fishing activities rely on Since 1990, the net decline in forest area
manual labour and low transportation costs has amounted to some 420 million ha by
(Welcomme et al., 2016). 2020. Between 2015 and 2019, forest loss
was 10 million ha/year compared with

30 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


12 million ha/year between 2010 and 611 thousand ha in 1990–1999 and 585 thou-
2014. Between 2010 and 2019, the high- sand ha in 2000–2009. The increase since
est net forest loss rates occurred in Africa 2010 is due largely to the Democratic Republic
(3.9 million ha/year) and Southern Amer- of the Congo, where the average annual rate
ica (2.6 million ha/year). Taking account of of loss was 723 thousand ha in 2010–2019, up
forest expansion through regeneration and from 442 thousand ha in 1990–2009.
afforestation, the overall net annual decline
between 2010 and 2020 was 4 million ha/year, Primary forest covers only 1 110 million ha,
compared with 5.2 million ha/year between natural regeneration covers 3 750 million ha
2000 and 2009 and 7.8 million ha/year and plantation forest covers 131 million ha.
between 1990 and 2000 (Ffigure 1.13 and Ffig- Regional variations are significant in rela-
ure 1.14). tion to the Amazon and Nile/Congo basin
hydrology and the primary forest func-
The area of primary forest decreased by tion of sequestering carbon and absorbing
81.3 million ha between 1990 and 2019. carbon dioxide. The global trend can also
The average annual rate of loss of primary mask significant increases at the country
forest was 3 410 million ha in 1990–2000 level. For instance, the slowing rate of loss
and 3 450 million ha in 2000–2010; the rate of Amazon forest cover in the Legal Amazon
dropped substantially in the latest decade, to region of Brazil observed after 1990 has been
1 270 million ha. The biggest average annual reversed, from some 4 600 km2/year in 2012 to
loss of primary forest area in 2010–2019 13 200 km2/year in 2021 (TerraBrasilis, 2022).
was in Africa, at 849 thousand ha, up from

Map 1.9 Global distribution of forests by climate domain, 2020

Boreal Temperate Subtropical Tropical

Dotted line represents approximately


the Line of Control in Jammu and
Kashmir agreed upon by India and
Pakistan. The final status of Jammu
and Kashmir has not yet been agreed
upon by the parties.

Final boundary between the Sudan


Source: FAO. 2020. Global forest resources assessment 2020: Key findings. Rome. https://doi.org/10.4060/ca8753en and South Sudan has not yet been
Modified to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420/ determined.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 31
FIGURE 1.13 Global forest areas in 2020 and net changes by decade, 1990–2020

Top five countries for forest area, Global annual forest Proportion and distribution of
2020 (million ha) area net change, by global forest area by climatic
decade, 1990–2020 (%) domain, 2020 (%)

1990 2000 2010


- 2000 -2010 -2020
2 000 0 50
45%
1 870 -1
40
1 500 -2

-3 30
27%
1 000 -4
815
20
-5 -4.7% 16%
497 -5.2%
500 -6 11%
347 310 10
200
-7
-7.8%
0 -8 0
da
tio n
or e

a
er tes

al

al

te

l
il

ca
ra sia

in
w f th

az

ic

re
na

ra
n
ld

a
Ch
a

pi
Br

op
de us

ic

Bo

pe
Am St
o

Ca

ro
R
st

Tr

m
of d

bt
Re

te

Te

Su
ni
Fe

Source: FAO. 2020. Global forest resources assessment 2020: Key findings. Rome. https://doi.org/10.4060/ca8753en

FIGURE 1.14 Annual forest area net change by decade and region, 1990–2020 (million
ha/year)

1990–2000 2000–2010 2010–2020

3
2.4
2
1.2 1.2
1 0.8
0.4 0.3
0.2 0.2
0
Million ha per year

-0.2 -0.2 -0.3 -0.1


-1

-2

-3 -2.6
-3.3 -3.4
-4 -3.9

-5
-5.1 -5.2
-6
Asia Oceania Europe North and South Africa
Central America
America

Source: FAO. 2020. Global forest resources assessment 2020: Key findings. Rome. https://doi.org/10.4060/ca8753en

32 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


1.4.6 Per capita There were slight declines in Eastern Europe
and the Russian Federation (Tfable 1.4).
land availability
Farm holdings in 179 countries are estimated
Global trends in agricultural land use (for
to total 608 million (Lowder, Sánchez and
cultivating crops and animal husbandry)
Bertini, 2021); 70 percent are smaller than
show significant variation. Between 2000 and
1 ha and account for only 7 percent of agri-
2019, agricultural land per capita declined by
cultural land. Farms with an area greater than
22 percent worldwide. Significant declines
1 thousand ha represent only 0.03 percent of
occurred in Africa, Western Asia, Australia
holdings but account for almost 40 percent
and New Zealand, and the Pacific Islands.

Agricultural lands (cropland, permanent meadows and pastures) per


Table 1.4
capita, 2000, 2010 and 2017
Region Agricultural Agricultural Agricultural Percentage
land per land per land per change 2000–
capita (ha/ capita (ha/ capita (ha/ 2017
capita) 2000 capita) 2010 capita) 2017

Northern Africa 1.33 1.12 0.92 −31

Sub-Saharan Africa 1.40 1.11 0.95 −32

Northern America 1.52 1.37 1.28 −16

Central America and 0.79 0.67 0.63 −20


Caribbean

Southern America 1.56 1.44 1.36 −13

Western Asia 1.48 1.17 1.02 −31

Central Asia 5.30 4.65 4.12 −22

South Asia 0.23 0.18 0.17 −26

East Asia 0.44 0.40 0.39 −10

Southeast Asia 0.21 0.21 0.21 0

Western and Central 0.41 0.42 0.41 −2


Europe

Eastern Europe and 1.28 1.31 1.31 2


Russian Federation

Australia and New Zealand 20.61 14.62 13.06 −37

Pacific Islands 0.25 0.23 0.20 −22

World 0.80 0.70 0.64 −20

High income 1.22 1.07 1.11 −9

Middle income 0.87 0.61 0.51 −41

Low income 0.58 0.73 0.84 45

Low income and food 0.55 0.47 0.41 −25


deficit

Least developed 1.18 0.98 0.80 −32

Source: FAO. 2018. The future of food and agriculture: Alternative pathways to 2050. Summary version. Rome.
www.fao.org/3/CA1553EN/ca1553en.pdf

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 33
of agricultural land. About 70 percent of Trends in land use and agricultural produc-
agricultural land comprises farms greater tion point to continued intensification from
than 50 ha. This skewed distribution of land a combination of high-yielding cereal grain
holdings is significant for designing and varieties, expanded irrigation systems,
deploying agricultural programmes to miti- improved management practices, and
gate land degradation processes and water increased farmer access to hybridized seeds,
scarcity. To treat 7 percent of agricultural land synthetic fertilizers and pesticides. Global
needs an outreach to more than 400 million application of fertilizer (nitrogen, phos-
farm holdings, while treating 40 percent of phorus and potassium) peaked in 2017 at

land requires outreach to some 163 thousand 192 million tonnes/year (Ffigure 1.15), and
overall pesticide use has remained stable at
holdings only.
around 4 million tonnes/year (Ffigure 1.16).
Southeast Asia has made remarkable prog- Farm mechanization is extending from
ress in terms of improving food security. In large commercial concerns to smallholder
the early 1990s, its undernourishment rates production in response to demand for higher

were the world’s highest at 31 percent, but productivity levels and as a reaction to limited
labour availability. Mechanized cultivation
these rates have now fallen below 10 percent.
for land treatment and conservation, seed
Agricultural land use has increased by more
drills and low-pressure trafficking/harvest-
than 50 percent between 1980 and 2014 in
ing vehicles is expanding into new markets
Cambodia, Indonesia, Lao People’s Demo-
that previously relied on animal traction.
cratic Republic, Myanmar and Viet Nam.
The increasing uptake of plastic cloches and
However, limits on agricultural land use are
glasshouses/polytunnels is now marked and
being reached. Agricultural production in
visible on time-lapsed satellite imagery.
Southeast Asia remains centred around rice.
But the contribution of rice to the total gross Productivity in aquaculture has increased
agricultural production value has fallen since through intensifying production methods. In
the early 1990s, from 40 percent to 30 percent Asia, farmed fish and crustacean production,
in 2013. Competition comes mainly from oil which mainly relies on additional feed, has
palm (Malaysia) and meat/fruit/vegetable replaced small-scale traditional pond aqua-
production (Myanmar). There has been a culture. Key drivers have raised land prices
significant increase in fertilizer and mecha- and the prices paid for farmed fish, which
nization use in East and South Asia, mainly in makes feed affordable.
the last 30 years.
To meet the increasing demand for food,

1.4.7 Agricultural
the Organisation for Economic Co-operation
and Development (OECD)/FAO outlook for
intensification 2021 (OECD and FAO, 2021) expects crop
yield growth to account for 88 percent of
The current level of land use is expected to
crop production increases to 2030, to come
continue with few options to set aside or
from yield improvements based on improve-
convert large areas of agricultural land other ment of genetic material, higher inputs
than through general urbanization. Although and investments in production technology.
marine food resources account for almost Some 7 percent of the projected increase
10 percent of the protein content in global could come from increasing the number
food supply, they make up less than 5 percent of harvests on the same land (cropping
of global food supply when measured in terms intensity) and 6 percent from a modest
of quantity and even less in terms of fat and expansion of cropland area.
kilocalories (FAO, 2020a).

34 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


FIGURE 1.15 Global fertilizer inputs, 2010–2019 (tonnes)

Nutrient nitrogen N (total) Agricultural use in World + (Total)

110 000 000

108 000 000

106 000 000


tonnes

104 000 000

102 000 000

100 000 000

98 000 000

96 000 000

2010 2011 2012 2013 2014 2015 2016 2017 2018 2019

Nutrient phosphate P205 (total) Agricultural use in World + (Total)

45 000 000

44 500 000

44 000 000
tonnes

43 500 000

43 000 000

42 500 000

42 000 000
2010 2011 2012 2013 2014 2015 2016 2017 2018 2019

Nutrient potash K20 (total) Agricultural use in World + (Total)

45 000 000

40 000 000

35 000 000

30 000 000
tonnes

25 000 000

20 000 000

15 000 000

10 000 000

5 000 000

0
2010 2011 2012 2013 2014 2015 2016 2017 2018 2019

Source: FAO. 2020. FAOSTAT. In: FAO. Rome. www.fao.org/faostat/en/#data/QC

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 35
FIGURE 1.16 Global crop protection inputs, 2010–2019 (tonnes/year)

Pesticides (total) Herbicides Rodenticides Fungicides and bactericides Insecticides

4 500 000

4 000 000

3 500 000

3 000 000
tonnes/yr

2 500 000

2 000 000

1 500 000

1 000 000

5 00 000

0
2010 2011 2012 2013 2014 2015 2016 2017 2018

Source: FAO. 2020. FAOSTAT. In: FAO. Rome. www.fao.org/faostat/en/#data/QC

Selective breeding of new crop varieties still varieties grown worldwide (OECD and FAO,
has the potential to boost yields. New vari- 2021). The state of the world’s biodiversity
eties that better exploit local conditions and for food and agriculture report (FAO, 2019b)
are more resistant to pests, diseases and highlights the drastic loss of biodiversity
drought, are more readily adopted than new that underpins food systems. Of some 6 000
management practices and are often cheaper plant species cultivated for food, fewer than
for farmers and extension organizations 200 contribute substantially to global food
(Fischer, Byerlee and Edmeades, 2014). Since output, and only 9 account for 66 percent
the Green Revolution, crop breeding has been of total crop production. Thus, governments
responsible for about half of all crop yield need policies that address biophysical and
gains. Progress has accelerated since 2000. structural drivers of biodiversity loss by
It will become increasingly important in the enhancing crop diversity and varieties for
future since agriculture has already exploited growers, including those developed using the
other solutions, such as adding more water, latest technologies and through participatory
using agrochemicals and introducing basic plant breeding by farmers.
machinery (Searchinger et al., 2019).
The production growth is expected mainly
Although crop breeding has contributed to in the emerging and developing economies
meeting increasing food demand, increasing where most population growth is antici-
privatization and domination of plant breed- pated. However, this raises serious concerns
ing and seed supply by a limited number of about the negative environmental and social
agribusiness multinationals has also contrib- impacts that will be created.
uted to the massive loss of crop species and

36 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


Future climate change scenarios point to the has slowed significantly, reflecting greater
need for changing cropping patterns and overall efficiency in the use of all agricultural
management practices to adapt to changes inputs and replacing intensification of inputs
in crop/land suitability. Agricultural systems on land as the primary source of growth
are already adapting, with more precise use in global agriculture output (Ffigure 1.17)
of technology and inputs, partly as a response (Fuglie, Jellife and Morgan, 2021). In 2011–
to climate change, but mainly as a response 2019, growth rates slowed to an average of
to the more sophisticated demands of the about 2 percent, while expansion of agricul-
global food system. Owing to these inno- tural land has been marked (Ffigure 1.17) and
vative measures, the traditional measures is attributed largely to slowing agricultural
of land and water productivity have seen productivity in developing countries includ-
their significance decline as more factors of ing Brazil, China and India (Fuglie, Jellife and
production are taken into account. Morgan, 2021).

Therefore, while growth in agricultural land This pattern of agricultural growth high-
use and irrigated areas has stagnated, total lights the impact of human-induced climate
factor productivity in agriculture has increased. change in slowing agricultural productivity
It grew from an average of 0.2 percent per year (Ortiz-Bobea et al., 2021) and the imperative
between 1961 and 1970 (when total agricul- for sustainable use of existing land and effi-
tural output growth peaked at 2.79 percent cient use of water resources and associated
per year) to almost 2.0 percent between 2001 nutrient and carbon cycling. While the use
and 2010 (Ffigure 1.17). This increase has of agricultural inputs has intensified to meet
occurred as growth of inputs and land factors

FIGURE 1.17 Total factor productivity growth in world agriculture, 1961–2019

Expansion of agricultural land Extension of irrigation to cropland More inputs per acre

Improvements in total factor productivity World total agricultural output growth rate

3.0
2.79
2.68

2.5 2.44
2.28 2.23
Average annual growth (percent)

2.08
2.0

1.5

1.0

0.5

0.0
1961-1970 1971-1980 1981-1990 1991-2000 2001-2010 2011-2019

Source: United States Department of Agriculture. 2021. International agriculture productivity. In: USDA Economic Research Service. Washington,
DC. www.ers.usda.gov/data-products/international-agricultural-productivity

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 37
current demand, the resulting environmen- Climate change exacerbates soil degrada-
tal impacts have accumulated to the point tion. Higher temperatures and extreme
where a wide range of environmental services weather events such as droughts, floods
are affected, limiting agriculture’s capacity and storms affect soil quantity and fertility,
to respond. At the same time, intersectoral reduce soil moisture and deplete the layers of
competition for land and water resources nutrient-rich topsoil. This section describes
is intense, so that while rainfed production the key soil processes, and section 1.6
has been able to expand since 2011 through discusses their effects on land degradation.
conversion of forested land (Ffigure 1.14), the
scope to extend irrigated has been extremely The Intergovernmental Technical Panel

constrained. on Soils defines soil health as “the abil-


ity of the soil to sustain the productivity,

1.5 Soils under


diversity, and environmental services of
terrestrial ecosystems”. Restoring soil

pressure services, of which agriculture is the primary


beneficiary, requires tackling all aspects of soil
degradation (FAO et al., 2020). Sustainable soil
Human pressures on soil resources under- management is now recognized as a funda-
pinning agricultural land use are reaching mental solution for addressing the interlinked
critical limits. Conservation agriculture problems of land degradation, climate change
techniques are now applied to conserve soil and biodiversity loss, and financing of related
structure and fertility on a large scale, such initiatives is indicated across all regions (FAO
as in southern Brazil. But the overall picture et al., 2020).
is continued land degradation through
human-induced erosion, compaction and
1.5.1 Declining soil organic
loss of structure through agricultural prac-
tice. The Status of the world’s soil resources
carbon and biodiversity
report (FAO and ITPS, 2015) found most of Soil organic carbon is the main compo-
the world’s soil resources are in only fair, nent of SOM and the principal indicator of
poor or very poor condition, with 33 percent soil health. It is responsible for many soil
of land being moderately to highly degraded functions that provide essential ecosys-
due to the erosion, salinization, compaction, tem services and plays a crucial role in the
acidification and chemical pollution of soils. global carbon balance. It regulates dynamic
biogeochemical processes and mitigates
The Status of the world’s soil resources report
the impacts of climate change by limiting
(FAO and ITPS, 2015) also identified ten main
the main carbon-based gases in the atmo-
threats to soil functions leading to soil degra-
sphere (FAO, 2019c). Soil organic carbon
dation: erosion, SOC loss, nutrient imbalance,
constitutes the largest terrestrial carbon
acidification, contamination, waterlogging,
pool, approximately 694 million tonnes of
compaction, sealing, salinization and loss of
carbon in the top 300 mm of soil (Ciais et al.,
soil biodiversity. Soil threats occur in specific
2013). Land-use change and unsustainable
soils under all types of agricultural land uses,
management practices translate into total
including wetlands and urban soils. Land-use
SOC depletion of 115–154 million tonnes of
changes, such as deforestation and urban-
carbon (Lal et al., 2018).
ization, are also increasing the rate of soil
degradation.

38 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


The Global Soil Organic Carbon Map (Map 1.10) the Congo and Indonesia have significant SOC
is the first global SOC assessment produced stock under forest-covered land.
at the national level. It enables estimates
of SOC stock in the top 300 mm of soil and Soil organic carbon stocks are susceptible

establishes a framework for compiling and and responsive to land-use changes from a

sharing georeferenced data on SOC. Continu- natural state to an agricultural ecosystem

ing improvements are expected to increase and from forestland to cropland (Map 1.11).

accuracy and resolution. Loss of SOM results in a reduction of nutrient


levels, soil structure and biodiversity, and
According to the Global Soil Organic Carbon the overall effect is to release GHGs, further
Map, 65 percent of global SOC is concen- reducing the soil’s capacity to help regulate
trated in ten countries, highlighting their climate change.
importance in managing terrestrial carbon
stocks sustainably. More than one-third of In the northern hemisphere, the permafrost

SOC stock is undercultivated land in China (permanently frozen soil) regions contain

and Kazakhstan. The Democratic Republic of twice as much carbon as in the atmosphere.
Most of the SOC stocks in Canada and the

Map 1.10 Global SOIL ORGANIC CARBON, 2019 (tonnes/ha)

0 - 20 (very low) 20 - 40 (low) 40 - 70 (medium) 70 - 90 (high) > 90 (very high)

Note: The three largest SOC stocks were found in boreal moist regions (130.5 Pg of Dotted line represents approximately the Line of Control in Jammu
and Kashmir agreed upon by India and Pakistan. The final status of
carbon) followed by cool temperate moist regions (98.8 Pg of carbon) and tropical
Jammu and Kashmir has not yet been agreed upon by the parties.
moist regions (80.4 Pg of carbon).
Final boundary between the Sudan and
Source: FAO. 2019. GLOSIS - GSOCmap (v1.5.0). Global soil organic carbon map. South Sudan has not yet been determined.

Contributing countries. In: FAO. Rome. http://54.229.242.119/GSOCmap/. Modified to


Final status of the Abyei area is not yet determined
comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 39
Map 1.11 Soil organic carbon stocks in croplands, 2019 (tonnes/ha)

SOC stocks

< 20 (Very Low)


20 - 40 (Low)
40-70 (Moderate)
70 - 90 (High)
> 90 (Very High)

Source: FAO. 2019. GLOSIS - GSOCmap (v1.5.0). Global soil organic carbon map. Contributing countries. In: FAO. Rome.
http://54.229.242.119/GSOCmap. Modified to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420

Russian Federation are under boreal forests global SOC stock: 600–644 billion tonnes of
and grasslands, which correspond to sparsely carbon (Leifeld and Menichetti, 2018). This
populated areas (Map 1.12). exceeds the carbon stored in the Earth’s
vegetation and may equal the carbon in the
Upland soils usually experience lower atmosphere (Turetsky et al., 2015). Intact
temperatures and higher precipitation and peatland ecosystems are carbon sinks, but
have higher SOC stocks than soils at lower when drained and degraded, they turn into
altitudes (FAO, 2019d). Mountain soils with long-term sources of GHGs (FAO, 2020e).
permafrost contain approximately 66 Pg of Drainage-based land-use systems have often
SOC, 4.5 percent of the global pool. High- provided short-term gains in mined organic
elevation and high-latitude soils are material (peat) or crops (from rice to oil
experiencing warmer air temperatures and palm) in exchange for long-term losses of
a thickening of the active weathering layer ecosystem services (Sumarga et al., 2016).
(Bockheim and Munroe, 2014), and are Globally, 11–15 percent of peatland has been
becoming more sensitive to change. drained for cropping, forestry, grazing and
energy use (FAO, 2020e). The largest drained
Organic soils cover only 3 percent of the global
areas are in Europe and Southeast Asia
land area but store up to 20 percent of the total
(Crump, 2017).

40 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


Map 1.12 Soil organic carbon in the circumpolar region, 2019 (tonnes/ha)

0 - 20 (very low)
20 - 40 (low)
40 - 70 (medium)
70 - 0 (high)
> 90 (very high)

Source: FAO. 2019. GLOSIS - GSOCmap (v1.5.0). Global soil organic carbon map. Contributing countries. In: FAO. Rome. http://54.229.242.119/GSOCmap
Modified to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420

Draining and clearing peat forests for plan- ecosystem service delivery. Pressures come
tations with fertilizer use and land clearing from deforestation, urbanization and agri-
by burning have led to a dramatic loss of SOC cultural intensification, loss of SOM/SOC,
(Parish et al., 2008). Constant subsidence soil compaction, surface sealing, soil acid-
combined with rising sea levels will increase ification, nutrient imbalance, pollution,
the risk of regular, and in some cases perma- salinization, sodification, desertifica-
nent, flooding in large coastal peatland areas tion, wildfires, erosion and landslides. Co-
(Sumarga et al., 2016). occurring drivers of environmental change
can have synergistic effects and may pose
Continual anthropogenic pressures are particular threats to soil organisms and
compromising the critical role of soil ecosystem functions. Deforestation and fires
biodiversity in ecosystem functioning and particularly affect soil biodiversity negatively.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 41
surface, such as the inter-Andean valleys
and the dark and deep soils characteristic
of the Argentine Pampas. However, there is
a growing weakness in the region’s ability
to produce staple foods, most of which are
©FAO/Vasily Maksimov

exported (Gardi et al., 2014).

In most African countries and other least


developed countries, continuous depletion
of soil nitrogen, phosphorus and potas-
sium, coupled with low crop production
levels, threatens agricultural sustainabil-
Changes in land use, such as forest or natural
ity and food security. Human-induced soil
grassland to pastureland or cropland, remove
fertility problems are expected to continue
biomass, disturb soils, lead to loss of SOC and
(Tan, Lal and Wiebe, 2005) including soil
other nutrients, and change soil properties
micronutrient deficiency caused by the lack
and soil biodiversity. However, the reverse
of nutrients in the geological material and
is also true. Afforestation on abandoned
poor soil management practices. For exam-
cropland can increase SOC and nutrients.
ple, boron, molybdenum and zinc deficien-
However, land use that does not change the
cies are present in 50 percent, 30 percent and
cover, such as forest harvest and regrowth or
15 percent of arable soils worldwide, respec-
increasing grazing intensity, can degrade soil
tively. Yet, soil micronutrient deficiencies can
properties (FAO and ITPS, 2015).
be closely associated to livestock and human
health issues (von Braum et al., 2021).
1.5.2 Soil nutrient loss
In some countries, nutrient removal in
The nutrient status of soil, or soil fertility,
harvested crops substantially exceeds inputs
is defined as the ability of soils to support
through natural replacement or fertilizer
and sustain plant growth, including making
application. Negative soil nutrient balances
nitrogen, phosphorus and other nutrients
have been reported in 15 agroclimatic regions
available for plant uptake. The introduction of
of India (Tandon, 1992). Agricultural inten-
mineral/synthetic fertilizers since the 1950s
sification can also deplete soil nutrients, as
has produced greater biomass production
crop residues are not returned to the soil to
and yield increases. However, SOM has still
prevent pests and weeds. Higher biomass
declined in many regions, and buffer capacity
production and yield increases have relied on
has been lost, resulting in overall nutrient
mineral/synthetic fertilizers since the 1960s
loss. In most cultivated soils, extensive use
(Ffigure 1.18). This has led to declining SOC
of mineral fertilizers has resulted in atmo-
in agricultural soils and replacing macro-
spheric pollution, GHG emissions (carbon
nutrients and micronutrients with mineral
dioxide and nitrous oxide), water eutrophi-
fertilizers. Between 1961 and 2002, global
cation and risks to human health (Galloway
nitrogen fertilizer use increased 7.4-fold and
et al., 2008).
phosphorus 2.3-fold (FAO, 2019e). Excessive

The soil atlas of Latin America and the Carib- use of mineral fertilizers can result in nutri-

bean illustrates significant soil diversity with ent loss through leaching and runoff, pollut-

contrasting situations. Naturally fertile soils ing groundwater and eutrophying surface

represent only 10 percent of the region’s water bodies. Microbial activity can also

42 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


transform excess nitrogen into nitrous oxide, demand for fertilizer has grown steadily
which is one of the GHGs responsible for by 1.6 percent each year and reached over
global warming as it contributes to deplet- 200 million tonnes in 2018. The treatment
ing the stratospheric ozone layer. Mineral of soils with inorganic fertilizers has signifi-
fertilizers are also sources of trace elements, cantly affected soil health and freshwater
increasing the soil’s toxic compound burden pollution induced by runoff and drainage.
(FAO, 2019f).
Soil nutrient budgets depend on local socio-
The sustained upward trends in nitrogen use economic conditions, market prices for farm
and phosphorus use on cropland since 1997 to inputs and government policies. In Western
counter soil nutrient loss and boost yields are Europe, rising fertilizer costs and strength-
expected to continue (FAO, 2019f). The global ening environmental policies have reduced

FIGURE 1.18 Global and continental nitrogen (Top) and phosphorus (Bottom)
fertilizer use, 1961–2019 (tonnes/year)

12 000 000

10 000 000

8 000 000

6 000 000

4 000 000

2 000 000

0
09
05
99
69

95
65

89

03
85

93
63

07
83
79

97
75

01

19
67

15
87

91
61

73

81

13
77

17
71

11
20

20
20

20
20
19

20
19

19

19

20
19

19

19
19

19

19

20
19
19

20
19

19

20
19
19

19
19

19

19

Africa Americas Asia Europe Oceania World

5 000 000

4 500 000

4 000 000

3 500 000

3 000 000

2 500 000

2 000 000

1 500 000

1 000 000

500 000

0
09
05
99
69

95
65

89

03
85

93
63

07
83
79

97
75

01

19
67

15
87

91
61

73

81

13
77

17
71

11
20

20
20

20
20
19

20
19

19

19

20
19

19

19
19

20
19

19

19
19

20
19

19

20
19
19

19
19

19

19

Source: FAO. 2020. The state of food and agriculture 2020. Overcoming water challenges in agriculture. Rome. https://doi.org/10.4060/cb1447en

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 43
nitrogen and phosphorus inputs into farm- According to the United States Department
land. This trend is expected to continue. of Agriculture, approximately 10 million ha
Dwindling phosphorus resources and climate of arable land annually drops out of agricul-
change may further affect soil nutrient tural use due to salinization, sodification and
balances in managed and natural ecosystems desertification. In 1990, the annual cost of
(FAO and ITPS, 2015). degraded salt-affected soils was estimated
at USD 264/ha. The total cost for 2013 was
Farmers practising irrigation often appreci- equivalent to USD 441/ha, adjusting for infla-
ate access to low-quality wastewater because tion (Qadir et al., 2014). Some 380 million ha
it contains macronutrients and micronutri- of salt-affected soils could be restored for
ents. However, the nutrient content is not agriculture (Lambers, 2003).
often accurately measured and balanced with
the inputs of synthetic and organic fertilizers. Human-induced salt-affected soils are
Although a benefit for some, wastewater can widespread. They exist around the Aral Sea
also negatively affect crop productivity, the and in the Islamic Republic of Iran. There is
soil, and human and environmental health. potential salinity due to permafrost melt-
ing, and saline intrusion occurs in in many
1.5.3 Soil salinization coastal aquifers and critically in small islands
with thin freshwater lenses (FAO, 2022b).
Soil salinization and sodification are major The legacy of deforestation in southeastern
soil degradation processes threatening Australia continues to compromise agricul-
ecosystem services as identified in the Status tural production as shallow groundwater
of the world’s soil resources report (FAO and rises into salt-laden aeolian soils and mobi-
ITPS, 2015). They are among the most import- lizes damaging soil water salinity levels. This
ant problems facing agricultural production, is a continuing risk to agricultural production
food security and sustainability in arid and affecting some 1 million ha in southeastern
semi-arid regions. Salt-affected soils refers Australia (Department of Primary Industries
to soils with a salt content that affects soil and Regional Development, Government of
properties, crop growth and yield (Daliako- Western Australia, 2021). Groundwater rise
poulos et al., 2016). They include saline soils, threatens an additional 2.8–4.5 million ha
saline–sodic soils and sodic soils, depending of low-lying or valley floor soils, despite
on the salt content, type(s) of salt present, recent periods of prolonged drought
amount of sodium present and soil alkalin- (Hatfield-Dodds et al., 2018).
ity/pH. Saline soils are known for containing
excessive amounts of soluble salts, mainly The distribution of salt-affected soil
calcium and magnesium. Sodic soils, with (Map 1.13) reflects a build-up of human-
abundant sodium salts such as sodium chlo- induced soil water processes. The Global
ride and sodium sulphate, usually have low Map of Salt-affected Soils represents the
permeability and a high pH of 8.2 and above spatial distribution of salt-affected soils
(FAO and ITPS, 2015). with electrical conductivity > 2 dS/m and/or
exchangeable sodium percentage > 15 percent
Soil salinization is a significant problem and/or pH > 8.2 at two depth intervals
worldwide (FAO and ITPS, 2015). It is esti- (0–30 cm and 30–100 cm). The Global Map
mated to take 0.3–1.5 million ha of farm- of Salt-affected Soils v1.0 indicates that more
land out of production each year and reduce than 424 million ha of topsoil (0–30 cm)
productivity for a further 20–46 million ha.

44 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


Map 1.13 Global Map of Salt-affected Soils (global salt-affected areas) v1.0: (Top)
topsoil (0–30 cm) and (Bottom) subsoil (30–100 cm)

Topsoil (0-30cm)

Salt-affected soils None No data

Subsoil (30-100cm)

Source: FAO. 2022. Global map of salt-affected soils (GSASmap). In: Global Soil Partnership. Rome. Dotted line represents approximately the Line of Control in Jammu and Kashmir
agreed upon by India and Pakistan. The final status of Jammu and Kashmir has not
Cited 9 February 2022. www.fao.org/global-soil-partnership/gsasmap/en. yet been agreed upon by the parties.
Modified to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420
Final boundary between the Sudan and South Sudan has not yet been determined.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 45
organisms and their biological activity, such
as reducing soil disturbance and maintaining
plant cover. Earthworms are well known for
©FAO/Christena Dowsett

enhancing soil structure and porosity, and


microorganisms stimulate nutrient cycling,
such as symbiotic nitrogen-fixing bacteria
for legumes and arbuscular mycorrhizal
fungi, which assist host plants to transport
bio-available phosphate. Yet, soil biodiver-
sity management is in its infancy; despite
and 833 million ha of subsoil (30–100 cm)
huge advances due to genetic technology,
are salt affected. These estimates, based on
knowledge remains in the research arena and
the submitted data (118 countries covering
not in the field.
73 percent of the global land area), show
that more than 4.4 percent of topsoil and Monocropping and repetitive tillage in many
more than 8.7 percent of subsoil of the total crops and environments threaten soil health
land area is salt affected. Soil salinity is esti- and biodiversity and increase pests and weed
mated to take up to 1.5 million ha of cropland infestation. In contrast, mixed-cropping
out of production each year. Higher rates of systems and agroforestry improve soil qual-
evapotranspiration are expected to exacer- ity, soil biodiversity and nutrient cycling,
bate the accumulation of salts in the surface especially when practised with conservation
horizons, but the extent of subsoil salinity agriculture to maintain cover through rota-
at the 30–100 cm depth range is much more tions and cover crops and to minimize tillage
pronounced. and traffic.

1.5.4 Soil compaction Conservation tillage, with crop rotations


that improve SOM, and organic inputs
Soil compaction is caused by livestock tram- reduce carbon dioxide losses and enhance
pling, heavy agricultural machinery and soil carbon sequestration. Even in dry lands,
inappropriate soil management practices where SOM is low, combined soil–crop–
such as tillage. It affects almost all physi- water approaches that increase biomass and
cal, chemical and biological soil properties minimize burning can sequester substantial
and functions. It also affects crop quality amounts of carbon in extensive agropastoral
and yield as plants cannot retrieve suffi- and rangelands.
cient nutrients, gases and water, and cannot
adequately develop their root systems (Soane
1.5.5 Soil erosion
and Van Ouwerkerk, 2013). Soil compaction
also increases waterlogging, soil erosion and Human activity and related land-use change
GHG emissions, reduces soil biodiversity and accelerate soil erosion. This has substantial
inhibits groundwater recharge (Radatz et al., implications for nutrient and carbon cycling,
2012). Increasing mechanization, livestock land productivity and socioeconomic condi-
and intensification exacerbate soil compac- tions. Removing forests to create cropland
tion in croplands and rangelands. and pasture is often followed by intensive
soil erosion (Pimentel and Burgess, 2013).
Soil and crop management practices can help Between 1985 and 2013, croplands and
to provide a favourable environment for soil

46 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


pasture areas increased by 279 million ha OECD, the annual soil loss rate of 11 tonnes/ha
(Borrelli et al., 2017). is considered critical for crop losses (Borrelli
et al., 2020). The ranges for soil loss and
The most likely range of annual tolerable soil loss demonstrate the need
global soil erosion (natural and human- for site-specific estimates to reflect differ-
induced) by water is now considered to be ent sensitivities to eroding surface soil
33–40 million tonnes. Given uncertainties in (FAO, 2019d).
the estimates of soil erosion rates, modelling
with a Revised Universal Soil Loss Equa- 1.5.6 Soil pollution
tion (RUSLE), Borrelli et al. estimated annual
average potential soil erosion amounts of Information about the extent of soil pollution
35 million tonnes and 36 million tonnes for at the global level is sparse, mainly due to
the 2001 and 2012 baselines, respectively the technical complexity of measuring the
(Borrelli et al., 2017). pool of contaminants present in agricultural
soils and their spatial variability (FAO and
Agriculture accounts for about three-quarters
UNEP, 2021). On agricultural land, contami-
of soil erosion globally (FAO and ITPS, 2015).
nants are derived from synthetic and mineral
This significantly affects crop yields and the
fertilizers that frequently contain significant
soil’s ability to store and cycle carbon, nutri-
concentrations of trace elements, mostly
ents and water. Annual cereal production
losses due to topsoil erosion are estimated to cadmium and radionuclides (El-Bahi et al.,

be of the order of 7.6 million tonnes. 2017). These can reduce soil organisms, plant
growth and food quality, and affect human
Upland and mountain soils are intrinsically health. Organic fertilizers such as manure or
vulnerable and sensitive to degradation sewage sludge often contain veterinary drugs
processes such as water erosion and loss of and other pharmaceuticals, trace elements,
chemical and physical quality (FAO, 2015)
and persistent organic contaminants and
(see the focus on mountain agriculture at
microplastics (Lwanga et al., 2017).
the end of this chapter). It should be noted
that annual soil erosion due to tillage alone
amounts to almost 5 million tonnes (FAO and
ITPS, 2015), indicating the relative scale of
human-induced land degradation.

Regional and global estimates of soil loss


rates differ substantially, depending on the
method used to derive them. Generally, mean
annual soil loss estimates from field plots are
considerably higher (8 tonnes/ha to almost
50 tonnes/ha) than those from regional and
global models (2–4 tonnes/ha). However,
any estimate of erosion must also be placed
in the context of the tolerable rate of loss,
which depends on the soil formation rates.
These vary significantly: early studies report
rates of 0.05–0.50 mm/year, while the most
© Ivo Balderi

referenced studies suggest 1.4–2 tonnes/ha


annually (Verheijen et al., 2009). According to

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 47
affinity with other organic compounds. Pesti-
cides deposited in soils and surface water can
be transported to places far from where they
were released. Soil organic carbon content,
©FAO/Albert Gonzalez Farran

texture, mineralogy, pH, microorganisms


and climate conditions will determine their
persistence, bioavailability and mobility.
In addition, some pesticides are sources of
trace elements, such as copper-based fungi-
cides applied in vineyards and orchards or
The legacy of polluted soils worldwide is fungicides and pesticides containing arsenic,
significant (Rodriguez-Eugenio et al., copper, manganese and zinc used in the past
2018). Thousands of chemicals are intro- to protect fruit crops.
duced into the environment every year, a
Plastic cloches, mulch film and agroplastics
high percentage of which are hazardous to
(e.g. agricultural product containers, irri-
human health and the environment (EEA,
gation hoses, bags, and fruit and vegetable
2019). Most soil contaminants come from
protection screens) are increasingly found
industrial processes, chemical production
in agricultural soils (Gao et al., 2019). Large
and waste products. These include: mineral
pieces of plastic can break into smaller pieces
extraction and processing; agrochemical
by photo-oxidation, microbial degradation
use, synthetic and organically derived from
and erosion, or can be physically damaged
manure, sewage sludge or biosolids; irriga-
by agricultural machinery, becoming micro-
tion using wastewater; transport and urban
plastics and nanoplastics. The tiny particles
activities; and military activities and armed
are incorporated into the soil structure, and,
conflicts. In Northern America, Southern
due to their hydrophobicity and lipophilic-
America, Europe, Central Asia and some East
ity, can retain other organic contaminants
Asian countries, the primary sources of soil
and form stable bindings with SOM (Boots,
pollution are intensive agriculture and live-
Russell and Green, 2019). Microplastics can
stock production. In sub-Saharan Africa and
also be ingested by soil-dwelling organisms
Pacific region countries, the main concerns
and enter the food chain.
are uncontrolled local and imported waste
accumulation. In Northern Africa and the
Near East, oil extraction and armed conflicts
are the major polluting activities (FAO and
UNEP, 2021).

Pesticide use has increased by over 34 percent


globally since 2000. In South America, it
has increased by 105 percent, in Oceania by
84 percent and in the Caribbean by 69 percent.
Many pesticides and their by-products persist
© Cristina Aldehuela

in the environment. They persist in the soil


and can be leached into groundwater and
transported by runoff to surface water bodies
due to their chemical structure, half-life and

48 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


1.5.7 Sandstorms
and dust storms
Approximately 430 million ha of dry lands,
comprising 40 percent of the Earth’s surface,
is susceptible to wind erosion (FAO and ITPS,
2015), but erosion rates are highly uncer-
tain. Estimates of the total dust mobilized
on arable land place an upper limit of about

© Roberto Faidutti
2 million tonnes/year (Yue et al., 2009).
However, wind mobilizes dust and coarser
soil particles (sand), implying much higher
total wind erosion rates.
loss of biodiversity and soil surface distur-
Sandstorms and dust storms (SDSs) are bance during cultivation. Some 40 percent of
responsible for eroding and depositing aerosols in the troposphere (the lowest layer
dryland soils. They can cause widespread of the Earth’s atmosphere) are dust particles
scouring of fine soil particles in the cold from wind erosion (Middleton and Kang,
(periglacial) and warm (desertic) climatic 2017). Storms may transport sand and dust
regimes (UNCCD, 2022). They can also cause particles hundreds to thousands of kilome-
accumulation of aeolian soils such as loess. tres. The frequency of SDSs is increasing, and
Global warming is expected to increase the climate change will be a significant potential
distribution, intensity and frequency of SDS driver of future wind erosion risk. The main
events, including local meso-climatic events areas affected are the arid regions of North-
such as tornados and local microevents such ern Africa, the Arabian Peninsula, Central
as “dust devils”. The issue is gaining atten- Asia and China. Australia, South Africa and
tion because of transboundary impacts on the United States of America are also affected
human and animal health in particular (Mu et but to a lesser extent. Global estimates of dust
al., 2013; Middleton and Kang, 2017; UNEMG, emissions, mainly derived from simulation
2022; WMO, 2022a). models, vary between 1 billion tonnes and
3 billion tonnes per year.
Sandstorms and dust storms depend on
meteorological conditions such as surface Sandstorms and dust storms adversely affect
wind speed and precipitation, and surface agriculture. They reduce crop and animal
properties including vegetation cover, sedi- production, bury crop seedlings, cause loss
ment availability and soil surface crusting. of plant tissue, reduce photosynthesis and
The main driving force is strong winds from increase soil erosion. Indirect dust depos-
thunderstorms or cyclones sweeping across its fill irrigation canals, impede transport
large areas of bare or sparsely vegetated arid routes, and affect river and stream water
and semi-arid lands that lift huge quantities quality. Sustainable agricultural and land
of soil particles into the atmosphere. Human management practices, such as conserva-
activity accelerates this process, notably land tion agriculture, agroforestry and other land
clearing, unsustainable agricultural practices management practices, can reduce the risks,
and mining, which cause vegetation deple- extent and severity of SDSs.
tion and associated hydrological changes,

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 49
1.6 Land
degradation –
human pressures
on land resources
© Daniel Hayduk

Human-induced land degradation is


affecting sustainable food production and
agriculture, livelihoods and the fight
against poverty. Degradation results
from complex local biophysical factors
Nonetheless, SDSs can bring interregional
and socioeconomic drivers, including
long-term and large-scale benefits, as
agricultural expansion, deforestation, fire,
surface dust deposits are a source of micro-
grazing density, population density and
nutrients for continental and maritime
invasive/native species ratio. The IPBES
ecosystems. For example, Saharan dust is
global land degradation and restoration
thought to fertilize the Amazon rainforest,
assessment refers to land degradation as
and dust transporting iron and phosphorus
the many processes that drive the decline
benefits marine biomass production in parts
of biodiversity, ecosystem functions or
of the oceans.
ecosystem services (Fisher, Montanarella
The impacts of SDSs on the climate, human and Scholes, 2018).
health, the environment and many socio-
Although land degradation may be appar-
economic sectors have increasingly been
ent at the field level, it cannot be measured
recognized over recent decades. The World
directly or monitored with Earth observa-
Meteorological Organization’s Sand and Dust
tion techniques because it combines local
Storm Warning Advisory and Assessment biophysical factors and socioeconomic driv-
System, launched in 2007, has three regional ers and depends on the subjective perception
centres covering: Asia; the Americas (led by of local populations and stakeholders (FAO,
the United States of America); and Northern 2013b; Fisher, Montanarella and Scholes,
Africa (led by China), the Middle East (led by 2018). Consequently, the Land Degradation
the United States of America) and Europe (led Assessment in Drylands (LADA) programme
by Spain). The centres are supported inter has defined land degradation as a process
alia by decisions of the UNCCD Conference that reduces the land’s capacity to provide

of the Parties and the UNCCD Science-policy ecosystem goods and services over time for its
beneficiaries and stakeholders (FAO, 2013b).
Interface (SPI).

In addition to the negative impacts of climate


Economic losses from single SDS events indi-
change, the human pressures on land and
cate the magnitude of their impacts. In China,
water resources are pushing the productive
between 2010 and 2013, dust storms caused
capacity of land and water systems to the limit.
losses estimated at USD 964 million. In the
These concerns are reflected in global envi-
Sistān region of the Islamic Republic of Iran, ronmental and scientific assessments, notably
between 2000 and 2005, dust storms led to the IPCC special report on climate change and
estimated losses of USD 125 million. land (IPCC, 2019), the sixth edition of the

50 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


UNEP global environmental outlook (UNEP, Additionally, the “convergence of evidence”
2019), the IPBES assessment report on land concept, developed for the World atlas
degradation and restoration (IPBES, 2018), of desertification (Cherlet et al., 2018), was
and the UNCCD global land outlook (UNCCD, adapted and applied for direct anthropogenic
2017). The land and water systems at risk drivers of degradation. The approach assumes
identified in SOLAW 2011 are now seeing their that a combination of pressures induced by
overall changing land and water productivity human activities is indicative of environ-
gains stagnate. Global datasets reflect a decline mental change. For example, when rangeland
in per capita natural resource availability. is burned to produce fresh forage for live-
stock, three direct anthropogenic drivers of

1.6.1 Biophysical status, land degradation may coincide: fire, grazing


and invasion of exotic species. The sum of
trends and pressures these pressures is referred to as the cumula-
For this report, the state of human- tive pressure by anthropogenic drivers.
induced land degradation was assessed at
the global level using an adapted Global Land Map 1.14 illustrates the biophysical status

Degradation Information System (GLADIS) of land based on nine input layers compiled
methodology (Bbox 1.1). This methodology around the year 2015. The input layers are:
compiles the changes in the biophysical soil nutrient availability, soil carbon content,
status of land elements over time at the water erosion, wind erosion, groundwater
national level and translates socioeconomic recharge, water stress, native species rich-
drivers (population density) into pressures. ness, above-ground biomass and artificial or
Biophysical status and drivers cover key built-up land cover. The grade of biophysical
environmental, social and economic vari- status, from high to low, is significant. With
ables, and the baseline is taken to represent
soil and wind erosion and water stress, high
pre-degradation conditions. As most global
erosion or stress rate implies a low status,
geospatial datasets do not date back further
while the remaining layers have a positive
than the 1980s, evaluation of the status and
impact and contribute towards a high score.
trend of land degradation and the respon-
A high score implies a high status.
sible drivers is constrained by long-term
data availability. Nevertheless, integrating
status, trends and drivers generates addi-
tional information about the distribution,
causes and land degradation processes.

A multi-index approach blends biophysical


status, trend and cumulative pressure from
anthropogenic drivers and is used together
with available high-resolution global data-
sets, to derive a global distribution of the
©FAO/Salvator Ndabirorere

extent of human-induced land degradation


(Coppus, 2022) for a 2015 baseline. The term
“degradation” is therefore used in this anal-
ysis only when associated with high pressures
from anthropogenic drivers. All other declines
in biophysical status, not related to such high
pressures, are defined as “deterioration”.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 51
Box 1.1
Global land degradation assessment using the adapted GLOBAL LAND
DEGRADATION INFORMATION SYSTEM method
Overall biophysical status and trend indices are determined using an adapted GLADIS methodology.
This applies a Geographic Information System (GIS) approach to calculate separate biophysical status
and trend indices for six components – biomass, soil health, water quantity, biodiversity, economic
services and cultural services. It combines them to give an overall status index and a trend index. Trends
refer strictly to changes over time (Table 1.5).

Table 1.5 Input layers for overall biophysical status, overall trend
and cumulative pressure by drivers
Item Soil Water Vegetation Demography

Status Nutrient availability Groundwater recharge Native species richness Built-up cover
Soil carbon content Water stress Above-ground biomass
Water erosion
Wind erosion

Trend Soil erosion change Freshwater change Change in land Population density
productivity change
Soil protection change Water stress change
Forest biomass change

Driver Agricultural expansion, deforestation, fire, grazing density, population density and ratio of invasive/native
species

The maps for overall biophysical status, trend and cumulative pressure represent three different
dimensions of land degradation. When combined, they give insight into the relationships among the
patterns, processes and their causes. Regions at risk occur when the overall status and trend are
combined. Areas with a low biophysical status and exposure to deterioration are at risk of ending in a
degraded state. Areas with high biophysical status and exposed to substantial deterioration are also
likely to be at risk. Integrating pressure from human activities with biophysical status and trends is a first
step to distinguishing natural from human-induced degradation.

Maps published in peer-reviewed journals provide the input layers. The criteria for selecting these
include availability, readiness to be used, relevance according to the literature and date of publication.

The biophysical status of land resources is based on nine input layers that reflect their present (or most
recently known) biophysical condition. These include soil nutrient availability, SOC, water erosion rate,
wind erosion, groundwater recharge, water stress, native species richness, above-ground biomass and
artificial land cover (urban and infrastructure).

The trend is based on seven input layers that indicate changes in soil, water, vegetation and population
density; they include changes in soil erosion, soil protection, freshwater, water stress, land productivity
and forest biomass. The time factor varies between 10 and 20 years.

Direct anthropogenic drivers are used to estimate pressure exerted by human activities: agricultural
expansion, deforestation, fire extent and frequency, grazing density, population density and ratio of
invasive/native species (Barger, Gardner and Mahesh, 2018).

Regions at risk are large contiguous areas with low biophysical status and subject to strong or light
deterioration. Regions with substantial deterioration and interspersed high and low biophysical status
are also at risk. Stable or improving areas are presently not at risk.

Land degradation classes are defined based on the trend of land deterioration and the presence of
anthropogenic drivers. A highly negative trend coinciding with high pressure is characteristic of substantial
human-induced land degradation. The land’s resilience (ability to withstand anthropogenic pressures)
also plays a role, for instance, when strong anthropogenic drivers do not coincide with negative trends.

Source: Nachtergaele, F., Petri, M., Biancalani, R., van Lynden, G., van Velthuizen, H. & Bloise, M. 2011. Global Land Degradation Information System (GLADIS) - An
information database for land degradation assessment at global level. LADA Technical Report 17. Rome, FAO.

52 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


Low values for biophysical status are found in southeast of the Plurinational State of Bolivia
dry lands, cold zones (with potential evapo- and northwest Paraguay. Mountainous areas
transpiration of less than 400 mm/year) with high biophysical status are situated in
and steep terrain, and appear to be related the upper Amazon basin at the foot slopes
directly to climate and geomorphology. Most of the Andes, the western Canadian Rockies,
of the land with a moderate biophysical the south coast of Quebec, the most north-
status is also situated in arid to subhumid, ern main island of Hokkaido in Japan, the
cold or mountainous environments. Flat and mountain ranges in Borneo and Papua New
humid areas with a moderate biophysical Guinea, and the Australian Alps. The highest
status are located throughout Europe, West status is located in the lowland rainforests of
Africa, the northern and southern parts of the the Amazon basin and the Guianas, the Choco
Congo basin, the Paraná basin in southwest region along the northern coast of Ecuador,
Brazil, Paraguay and north Argentina and and the Pacific coast of Colombia and eastern
the southwest coast of the United States of Panama, from the southern Gulf of Guinea
America. High biophysical status is located to the Congo basin, and in the south part of
in the remaining flat and humid regions of the Malay Peninsula, on Borneo, Papua New
the world and the dry Gran Chaco in the Guinea and Sumatra.

Map 1.14 Biophysical status of land, 2015

Status
High status

Low status

Note: Overall status, where status is defined as the capacity to provide ecosystem services and goods. Nutrient availability, soil carbon content, water erosion, wind
erosion, groundwater recharge, water stress, native species richness, above-ground biomass and built-up cover served as input layers to assess overall status.

Source: Coppus, R. 2022. Global distribution of land degradation. Thematic Background Report for SOLAW 2021. Rome, FAO.
www.fao.org/land-water/solaw2021/en. Modified to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 53
Map 1.15 presents the trend in biophysi- rainforests, decreasing forest biomass and
cal status to illustrate where there has been increasing population result in a negative
a declining status or where the status is trend, but locally decreasing soil protection
broadly stable or even improving, based on a and increasing erosion rates also contribute.
set of GIS layers in which change is detected
(soil erosion, soil protection, water stress, Map 1.16 presents drivers of land degrada-

land productivity, forest biomass and popu- tion, with the index based on six input layers

lation density). of direct anthropogenic drivers: deforesta-


tion, accessibility, agricultural expansion,
In general, an overall decline in biophysical fire, invasive species and grazing. Low values
status has resulted from various negative correspond to low intensities or pressure and
trends together with the combinations of high values correspond to high pressure.
indices varying per affected area. The trend
analysis reveals that even the few remain- At the global level, there are some regions

ing regions with large, contiguous tropical where many drivers converge, resulting in

rainforests are subject to decline. Only some extensive areas with high pressure on soil,

areas in the core of the Amazon, the eastern water and vegetation resources. These include

part of the Congo basin and isolated patches the east coast of the United States of America,

in Borneo are stable or improving. For most including the Great Lakes area and the Mexi-
can Gulf coast states; Western, Central and

Map 1.15 Trend in biophysical status, 2015

Improvement

Strong decline

Note: Overall trend, where trend is defined as a change in status (the capacity to provide ecosystem services and goods). A negative trend is
referred to as decline, a positive trend is referred to as improvement and a trend with a value near zero is referred to as stable.

Source: Coppus, R. 2022. Global distribution of land degradation. Thematic Background Report for SOLAW 2021. Rome, FAO. www.fao.
org/land-water/solaw2021/en. Modified to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420

54 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


Eastern Europe, and the adjacent Volga basin common. Australia is exposed to fire, and New
in the Russian Federation; east Pakistan, India Zealand is subject to high grazing densities.
and Bangladesh; and the Democratic People’s In Africa, fire and grazing are common, and
Republic of Korea, the Republic of Korea, in the Southern America region, grazing and
central east China and Japan. deforestation dominate.

Map 1.17 presents the dominant drivers of


1.6.2 A global perspective
land degradation. The main driver is domi-
nant relative to the other drivers in terms of At the global level, in 2015, areas affected
pressure but not necessarily responsible for by human-induced land degradation covered
human-induced land degradation because low 1 660 million ha, of which 850 million ha
and moderate pressures were also included in was subject to strong degradation, and
the analysis. Grazing and agricultural expan- 810 million ha was subject to light degrada-
sion are common in large parts of the United tion (Ttable 1.6). Degrading areas were rela-
States of America, while invasive species and tively evenly distributed over dry lands and
deforestation dominate in Alaska, northern humid areas, although humid areas had a
Canada, Northern Europe and Siberia. Inva- higher share of light degradation (Map 1.18).
sive species also dominate in Europe. In the Human-induced land degradation occurred
Asian steppe, the most frequent drivers are in 11 percent of dry lands and 15 percent of
fire and grazing, and in South and Southeast humid areas.
Asia, population density and deforestation are

Map 1.16 Land degradation pressures, 2015

Pressure
High pressure

Very low pressure

Note: The cumulative effect of direct human drivers of land degradation (deforestation, accessibility, agricultural expansion, fire, invasive species and grazing)
translated into pressure.

Source: Coppus, R. 2022. Global distribution of land degradation. Thematic Background Report for SOLAW 2021. Rome, FAO. www.fao.org/land-wa-
ter/solaw2021/en. Modified to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 55
Map 1.17 Dominant drivers of land degradation, 2015

Very low pressure Accessibility Fire Grazing


Deforestation Agricultural expansion Invasive species Various

Dotted line represents approximately the


Line of Control in Jammu and Kashmir
Note: Global distribution of the dominant direct human drivers per area. The dominant driver is defined as the driver that agreed upon by India and Pakistan. The
exerts the highest pressure in a given area. final status of Jammu and Kashmir has not
yet been agreed upon by the parties.

Source: Coppus, R. 2022. Global distribution of land degradation. Thematic Background Report for SOLAW 2021. Rome, FAO.
Final boundary between the Sudan and
www.fao.org/land-water/solaw2021/en. Modified UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420 South Sudan has not yet been determined.

High pressure does not necessarily lead to ing 71 percent is classified as deterioration
human-induced land degradation. The trend caused by natural processes or which has an
analysis shows 3 576 million ha of land was anthropogenic origin. Comparing deteriorated
under high pressure from human activi- areas with status reveals that about half have
ties in 2015, of which slightly less than half low status. Areas with low status are likely to be
(1 660 million ha) was subject to human- more sensitive to degradation processes than
induced land degradation (Tfable 1.7). This areas of high status. Moderate pressures may
implies more than half of the areas under suffice to trigger human-induced land degra-
high pressure are stable. Comparing the dation. A closer look at areas with low status
land degradation map with the status layer subject to deterioration shows 656 million ha
reveals 82 percent of these areas have high is under moderate pressure, equal to
status, suggesting favourable land conditions 12 percent of the overall global decline. Most
impede degradation processes. of these areas are probably affected by human-
induced land degradation, which means that
At the global level, the status of 5 670 million ha approximately 41 percent of global decline
of land was declining in 2015, of which can be attributed to human-induced land
1 660 million ha (29 percent) is attributed to degradation.
human-induced land degradation. The remain-

56 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


Table 1.6 Extent of human-induced land degradation, 2015 (million ha)
Degradation Global Drylands Humid areas

Total 1 660 733 927

Strong 850 418 432

Light 810 315 495

Note: Antarctica, Greenland and land with more than 90 percent bare cover (the great deserts) are excluded. For humid areas, the cold zone
where potential evapotranspiration > 400 mm is also excluded.

Source: Coppus, R. 2022. Global distribution of land degradation. Thematic Background Report for SOLAW 2021. Rome, FAO. www.fao.
org/land-water/solaw2021/en

Table 1.7 Extent of land degradation, 2015 (million ha)


Land degradation status Extent

Negative trend 5 670

High cumulative pressure 3 576

Human-induced land degradation 1 660

Source: Coppus, R. 2022. Global distribution of land degradation. Thematic Background Report for SOLAW 2021. Rome, FAO. www.fao.
org/land-water/solaw2021/en

In 2015, a fifth of human-induced degraded from the Loess Plateau to the Yellow River
land was in sub-Saharan Africa, followed by basin and the Bohai Sea region. However,
Southern America with 17 percent (Tfable 1.8). almost all inhabited parts of the world are
Northern America is about five times the size subject to some form of human-induced land
of South Asia, but both regions contributed degradation, and 52 degrading regions have
11 percent to global degradation. In rela- been identified with an optimized analysis
tive terms, South Asia was the most affected undertaken for this report (Coppus, 2022).
region, with 41 percent of its area suffering
from human-induced degradation, of which Global warming, causing ice sheet loss

70 percent was strongly degraded. South- and melting glaciers, is responsible for a

east Asia follows with 24 percent, of which substantial decrease in available freshwater

60 percent was severe, and Western Asia had in the eastern arctic region of Canada and the

20 percent, of which 75 percent was strongly stretch from southern Alaska to southwest

affected. Deserts are not included in these Yukon. Groundwater depletion and drought

estimates. have led to a substantial decrease in fresh-


water availability in California and adjacent
There are three regions severely affected Nevada (Rodell et al., 2018). In Texas, soil,
by human-induced land degradation over water and vegetation resources are in sharp
large contiguous areas: the arc in north- decline due to a combination of drought,
ern Western Asia, stretching from Israel and grazing, population pressure and expanding
Jordan to southeast Turkey into Mesopota- agriculture. Soil protection decline, high soil
mia and western Islamic Republic of Iran; the erosion rates and water stress characterize
Indo-Gangetic plain south of the Himalaya on the fragile conditions in the Sierra Madre
the Indian subcontinent; and northern China, Occidental in Mexico.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 57
Map 1.18 Land degradation classes based on the severity of human-induced
pressures and deteriorating trends, 2015

Strong human-induced land degradation Light deterioration under low pressure Bare

Light human-induced land degradation Stable or improvement under high pressure

Strong deterioration under low pressure Stable or improvement under low pressure

Note: Global distribution of land degradation. Overall trend combined with cumulative pressure by direct human drivers. Dotted line represents approximately the
Line of Control in Jammu and Kashmir
Human-induced land degradation refers to a negative trend, which is caused by human activity. Deterioration refers to a agreed upon by India and Pakistan. The
negative trend caused by natural phenomena or by humans in case status is low. final status of Jammu and Kashmir has not
yet been agreed upon by the parties.

Source: Coppus, R. 2022. Global distribution of land degradation. Thematic Background Report for SOLAW 2021. Rome, FAO.
Final boundary between the Sudan and
www.fao.org/land-water/solaw2021/en. Modified UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420 South Sudan has not yet been determined.

Parts of Spain face increasing water stress from the Loess Plateau in northern China to
and groundwater depletion. Drought in west the Bohai Sea region is coping with severe
Kazakhstan has led to deterioration of the groundwater depletion and population pres-
Caspian Sea and Aral Sea. Large parts of sure. Arid west Australia receives low rainfall,
Western Asia are at risk because of severe which is combined with a decline in land
groundwater depletion, drought and popula- productivity and large, frequent fires.
tion increase. East Pakistan and north India
are exposed to groundwater depletion and The eastern Maghreb is exposed to agri-

population pressure, whereas water stress cultural expansion and decreasing freshwa-

and agricultural expansion are the main ter availability. The northern Nile valley is

issues in southeast India. The semi-arid area subject to high-intensity grazing, population

58 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


Table 1.8 Human-induced land degradation, 2015
Region Area affected Total land Percentage Strongly Slightly
by human- area of of region degraded degraded
induced region affected (million ha) (million ha)
degradation (million ha) (%)
(million ha)

Sub-Saharan Africa 330 2 413 14 149 181

Southern America 281 1 778 16 153 128

South Asia 180 439 41 126 54

Northern America 177 2 083 8 82 95

East Asia 156 1 185 13 84 72

Western Asia 123 615 20 92 31

Southeast Asia 122 501 24 74 48

Australia and New 94 796 12 34 59


Zealand

Eastern Europe and 83 1 763 5 21 62


Russian Federation

Western and Central 56 489 11 12 44


Europe

Central Asia 31 456 7 12 19

Northern Africa 22 579 4 9 13

Central America and 11 76 14 5 5


Caribbean

Pacific Islands 0.14 7 2 0.11 0.03

World 1 660 13 178 13 850 810

High income 393 3 817 10 175 218

Upper middle income 621 5 604 11 326 295

Lower middle income 428 2 207 19 241 187

Low income 220 1 520 14 107 112

Low income and food 283 2 062 14 133 149


deficit

Least developed 288 2 097 14 134 154

Note: Percentage of region extent refers to the portion of the total regional extent that is degraded. Antarctica, Greenland and land with more
than 90 percent bare cover (the great deserts) are excluded.

Source: Coppus, R. 2022. Global distribution of land degradation. Thematic Background Report for SOLAW 2021. Rome, FAO. www.fao.
org/land-water/solaw2021/en

pressure and increasing water stress. In the Sudan is coping with agricultural expan-
Ethiopian Highlands, soil erosion caused by sion and increasing water stress. The western
intensive grazing and agricultural expansion part of South Sudan is affected by large and
is a significant issue. The southern half of recurrent fires and is also subject to massive

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 59
forest biomass loss and increasing water the year 2015. In 2015, human-induced land
stress. Southern Africa faces increasing water degradation primarily affected cropland
stress and a decline in land productivity. (FAO and IIASA, 2021). Although cropland
Eastern Brazil has been affected by a recent covered only 15 percent of the analysed area,
drought that caused increased water stress it accounted for 29 percent of all degraded
and decreased land productivity. Similar areas. Almost one-third of rainfed cropland
effects are found in central Argentina, where and nearly half of irrigated land are subject to
precipitation is decreasing and large areas are human-induced land degradation.
being burned.
In Northern Africa, Western Asia and South
Stable or improving regions with low Asia, more than 60 percent of the irrigated
biophysical status are located in the western areas are degraded. The largest areas are
arctic zones of Northern America and Eurasia in the northern hemisphere, except for
and at the edges of the great deserts, such Southeast Asia. Globally, only 38 percent of
as the Sahara (the Sahel), the Karakum in irrigated land is stable, the lowest of the land
Central Asia, the Gobi in East Asia and the covers analysed.
Kalahari in Southwest Africa. The arid and
semi-arid regions of the Taklamakan desert, In Western Asia, agricultural expansion,

the Tibetan plain, southeast Australia and the grazing and accessibility drive degradation,

Horn of Africa show low resistance to degra- while in the densely populated areas of

dation. Stable or improving regions with high East Asia and South Asia, good accessibility

biophysical status are located throughout and high grazing density are exerting high

southern Canada and the northern and central pressures on irrigated fields. Grazing,

east part of the United States of America. The accessibility and deforestation drive

stable or improving regions stretching from environmental change in irrigated cropland

Central and Southeast Europe to the Eurasian in Southeast Asia. Grazing, accessibility and

taiga and from eastern Mongolia to Manchu- agricultural expansion contribute most to the

ria also have high biophysical status. Low pressure on irrigation in the eastern United

biophysical status occurs in dry lands and States of America.

mountains and seems to be related to climate


The decline in status in East Asia and Western
and geomorphology. However, this could be
Asia is mainly due to decreasing freshwater
the result of severe degradation in historical
availability, increasing water stress, reducing
times, which has reached a new equilibrium
soil protection and increasing population.
and appears to be under natural conditions.
Similar degradation processes occur in South
Unfortunately, there is no technique to iden-
Asia. Major degradation processes in South-
tify such areas with current datasets (Fischer
east Asia are increasing erosion rates, rapidly
and van Velthuizen, 2018).
decreasing forest biomass and increasing
population. In the eastern United States of
1.6.3 Productive America, a decline in available freshwa-
areas at risk ter and loss of soil protection are the main
degradation processes. Problems are similar
Tfable 1.9 presents a summary of the spatial
in the western United States of America, but
relationship between land degradation and
rising population density brings additional
the global land-use/-cover classes derived
pressure.
from GAEZ v4 for indicators compiled around

60 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


Table 1.9 Extent of land degradation classes for global land cover, 2015
Land Total area Degradation Deterioration Stable Degraded Deteriorated Stable
cover (million ha) (million ha) (million ha) (million ha) (%) (%) (%)

Cropland 1 527 479 268 780 31 18 51

Rainfed 1 212 340 212 660 28 17 54

Irrigated 315 139 57 120 44 18 38

Grassland 1 910 246 642 1 022 13 34 54

Trees 4 335 485 1 462 2 388 11 34 55

Shrubs 1 438 218 584 636 15 41 44

Herbs 203 16 51 136 8 25 67

Sparse
1 034 85 499 450 8 48 44
vegetation
Protected
980 76 361 443 9 41 50
area

Note: The term degradation refers to high pressures from anthropogenic drivers. All other declines in biophysical status are defined as dete-
rioration.

Source: Coppus, R. 2022. Global distribution of land degradation. Thematic Background Report for SOLAW 2021. Rome, FAO. www.fao.
org/land-water/solaw2021/en

The expansion of areas at risk is indicative Africa, where decreasing land productivity
of declining ecosystem services as Map 1.19 and soil protection account for declining
illustrates. Biophysical status is given as ecosystem services. In Asia, increasing water
much importance as trend, and no distinction stress also contributes to the grasslands at
is made between anthropogenic and natural risk. In sub-Saharan Africa, grasslands are
causes. Consequently, areas with a low status prone to frequent and intense fires. Forest-
and strong decline are considered to be at risk. land at risk is prone to deforestation, and
Areas with a high status and slight decline are in sub-Saharan Africa also to frequent and
not considered to be at risk. Neither are stable severe fires. Forests at risk are affected by
nor improving areas. Based on this analysis, decreasing freshwater, loss of soil protec-
the extent of cropland at risk (Ttable 1.10) is tion and decreasing forest biomass. The
similar to the extent of degraded croplands biophysical status of most regions at risk is
in 2015. Combining status and trends indi- characterized by low SOM and low plant
cates areas at risk in 2015 amounted to some species biodiversity.
3 866 million ha. The distribution of irrigated
and rainfed cropland at risk matches the
degraded areas.

Croplands at risk tend to be areas recently


brought into production and which are
subject to limited freshwater availability and
where population density is increasing. Most
©FAO/Ivo Balderi

grasslands at risk are exposed to decreas-


ing freshwater availability. The exceptions
are in Southern America and sub-Saharan

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 61
Table 1.10 Areas of agricultural land and forest at risk, 2015
Land cover Total area (million ha) Area at risk (million ha) Area at risk (%)

Cropland 1 527 472 31

Rainfed 1 212 322 27

Irrigated 315 151 48

Grassland 1 910 660 35

Forestland 4 335 1 112 26

Source: Coppus, R. 2022. Global distribution of land degradation. Thematic Background Report for SOLAW 2021. Rome, FAO. www.fao.
org/land-water/solaw2021/en

Map 1.19 Regions at risk based on status and trends of land resources, 2015

Strong decline, low status: at risk Light decline, high status Bare
Strong decline, high status: at risk Stable or improvement, low status
Light decline, low status: at risk Stable or improvement, high status

Note: Overall biophysical risk combined with overall trend.

Source: Coppus, R. 2022. Global distribution of land degradation. Thematic Background Report for SOLAW 2021. Rome, FAO. www.fao.
org/land-water/solaw2021/en. Modified UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420

62 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


1.7 Water scarcity
Direct measurement of changes in surface
water body “cover” with high-resolution
satellite sensors have been available since

1.7.1 Structural changes 1984 (Pekel et al., 2016; EC, 2020, 2021).

within the global These data reveal that 0.9 million ha of


detectable water bodies has disappeared, and
water balance 7.3 million ha has transitioned from a perma-

The global water budget is under pressure. nent state to a seasonal state between 2000

The long-term internal renewable water and 2019. Over the same period, 18.4 million ha

resources (IRWRs) derived from rivers, lakes of new permanent water bodies was created

and shallow aquifer circulation are estimated in areas that were not previously covered.

to amount to 44 211 km3/year (Ffigure 1.19). Natural surface water bodies are expanding

Estimated withdrawals for all sectors exceeded due to accelerated runoff/snowmelt, as on

4 000 km3/year in 2018, almost 10 percent of the Tibetan Plateau. The measurable change

IRWRs. The global freshwater water balance in permanent and seasonal surface water

estimated for 2000 (Ffigure 1.19) is still valid bodies on irrigated and rainfed cropland is

for long-term means of precipitation and significant. Tfable 1.11 and Tfable 1.12 present

outflows, but structural changes to storage the regional breakdowns for water cover on

volumes in snowpacks, glaciers and aquifers irrigated land and rainfed cropland for 2019

have since occurred across many of the large and the changes established with respect to

continental river basins. a 2000–2004 baseline. For irrigated land


and rainfed cropland, the global aggregate

FIGURE 1.19 Annual global water balance, 2000

precipitation
105 316 (805 mm)
Global terrestrial annual water balance established
for the year 2000 after Hoogeveen et al. 2015.
principal hydrological circulation
incremental evaporation
rainfed evaporation
current (2018) level of withdrawals
61 106 (89 mm)
All units km3/yr

wetland
2 899 (22 mm) irrigation open water urban areas
1 268 (10 mm) 1 184 (9 mm) negligible

renewable
water resources
44 211 (338 mm)

agricultural
municipal industrial
2 950
483 646
Surface runoff : 32 615 (249 mm)
Return flows

Groundwater recharge:
11 607 (89 mm) groundwater storage/circulation

outflow to sea
38 859 (297 mm)

Source: Hoogeveen, J., Faurès, J.M., Peiser, L., Burke, J. & Van De Giesen, N. 2015. GlobWat – A global water balance model to assess water use
in irrigated agriculture. Hydrology and Earth System Sciences, 19(9): 3829–3844.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 63
increases in areas are positive, with irri- in large dam reservoir storage greater than
gated land registering nearly 5 percent gain 100 million m3 (GDW, 2022), and Ffigure 1.20
in permanent cover and rainfed cropland summarizes version 1.3 of the database.
registering 2.5 percent gain (Tfable 1.12). The Total built storage as of 2016 was estimated
largest regional gains in permanent cover to be of the order of 7 500 km3. Although
are in the South Asia and East Asia regions, large dam construction has declined over the
while the largest losses occur in Eastern past two decades, reservoir size and relative
Europe. Permanent water bodies occupy magnitude of river flows impounded have
about 10 percent of the combined areas increased. Reservoir storage is approximately
equipped for irrigation (342 million ha) and half of total freshwater withdrawals, and
rainfed (1 556 million ha) based on FAOSTAT the impact on wetlands and free-flowing
data in 2019. rivers is significant (Schneider et al., 2017;
Grill et al., 2019). Annual evaporation from
The volume of artificially stored surface impounded reservoirs is estimated to be
water is significant. The Global Reservoir and approximately 350–400 km3 (FAO, 2020a).
Dam Database monitors the rate of change In the International Commission on Large

Table 1.11 Permanent and seasonal water cover on irrigated land, 2019 and 2000–2004
Changes (ha)
Region Permanent Seasonal Total Permanent Seasonal
extent, 2019 extent, 2019 water changes since changes since
extent, 2019 2000–2004 2000–2004
baseline baseline

Australia and New 8 812 9 382 18 194 −3 425 −2 416


Zealand

Central America and 11 251 11 733 22 984 1 541 2 607


Caribbean

Central Asia 155 033 757 795 912 828 2 138 380 195

Eastern Europe 106 557 56 956 163 513 −13 373 13 233

Northern Africa 60 475 61 757 122 232 347 17 762

Northern America 238 594 1 016 245 1 254 838 15 950 319 488

Pacific Islands 0 0 0 0 0

Southern America 73 096 108 213 181 309 −1 371 −9 251

South Asia 456 204 4 358 549 4 814 754 −9 923 914 729

East Asia 1 419 312 2 305 217 3 724 529 128 047 −62 124

Southeast Asia 196 888 1 977 423 2 174 311 −7 634 −396 565

Sub-Saharan Africa 52 910 128 216 181 126 434 62 473

Western Asia 217 788 603 638 821 427 31 014 438 097

Western and Central 143 547 156 048 299 595 9 201 65 676
Europe

Total/net change on 3 140 468 11 551 172 14 691 639 152 945 1 743 903
irrigated land

Sources: Data from European Commission. 2020. Global surface water explorer. In: European Commission.
https://global-surface-water.appspot.com/#data; European Commission. 2021. Index of /ftp/jrc-opendata/GSWE.
https://jeodpp.jrc.ec.europa.eu/ftp/jrc-opendata/GSWE/

64 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


Table 1.12 Permanent and seasonal water cover on rainfed cropland, 2019 and
2000–2004 changes (ha)

Region Permanent Seasonal Total Permanent Seasonal


extent, 2019 extent, 2019 water changes since changes
extent, 2019 2000–2004 since 2000–
baseline 2004 baseline

Australia and New 191 057 440 394 631 451 −39 310 −48 526
Zealand

Central America and 185 889 164 666 350 555 8 872 32 014
Caribbean

Central Asia 705 919 1 319 492 2 025 411 19 112 702 853

Eastern Europe 1 860 830 1 359 806 3 220 636 −96 620 221 401

Northern Africa 51 381 32 954 84 334 10 281 11 623

Northern America 1 480 520 1 542 198 3 022 717 142 060 506 394

Pacific Islands 1 000 1 724 2 724 46 1 124

Southern America 3 176 202 3 372 214 6 548 416 −82 183 184 092

South Asia 726 382 2 993 642 3 720 024 61 018 1 004 399

East Asia 1 874 842 2 093 807 3 968 650 307 681 265 463

Southeast Asia 1 136 326 2 842 313 3 978 639 24 318 185 560

Sub-Saharan Africa 1 955 850 2 413 382 4 369 232 −50 706 835 436

Western Asia 345 662 323 305 668 967 47 240 158 534

Western and 702 146 232 021 934 168 13 822 46 298
Central Europe

Total/net change 14 394 006 19 131 918 33 525 924 365 632 4 106 666
on rainfed land

Sources: Data from European Commission. 2020. Global surface water explorer. In: European Commission.
https://global-surface-water.appspot.com/#data; European Commission. 2021.
Index of /ftp/jrc-opendata/GSWE. https://jeodpp.jrc.ec.europa.eu/ftp/jrc-opendata/GSWE/

Dams world register of dams, of the 58 713 The adverse impact of water storage sedi-
registered dams, 13 580 are dedicated to irri- ment flows is particularly important for the
gation as single-purpose dams and a further long-term evolution of deltas, which support
6 278 irrigation dams are registered as being irrigated production and aquaculture.
multipurpose. Reduced sediment flows into deltas combined
with land subsidence (from compaction and
The impact of small-scale hydraulic struc- groundwater withdrawals) are estimated
tures on surface storage is less certain, but to result in an average relative sea-level
the areal contribution of small reservoirs and rise of 6.8 mm/year (Tessler et al., 2018).
tanks in agricultural areas with prolonged dry Impacts of planned dams and dams under
seasons such as Eastern Africa and peninsular construction are estimated to increase the
India is indicated by surface water dynamics relative sea-level rise by up to 1 mm/year
(Pekel et al., 2016). in some deltas progressively starved of

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 65
sediment. Sediment flows have been esti- The overall change in per capita distribu-
mated to decrease by up to 60 percent in the tion of freshwater resources is significant as
Danube basin and 20 percent in the Ganges– populations grow. The decline in global per
Brahmaputra–Meghna system (Tessler et al., capita IRWRs was about 20 percent between
2018), with implications for the high popu- 2000 and 2018 (Ffigure 1.21). The change
lation concentrations associated with deltas was greater in countries with the lowest per
(Tellman et al., 2021). Higher rates of fluvial capita IRWRs, such as sub-Saharan Africa
erosion downstream of large dams releasing (41 percent), Central Asia (30 percent), West-
sediment-hungry water also threaten previ- ern Asia (29 percent) and Northern Africa
ously productive alluvial terraces (Kondolf et (26 percent). The region with the lowest
al., 2014). percentage change was Europe (3 percent).
On the demand side, the regions with the
Changes in the volumes of water withdrawn largest water withdrawals per capita were
also point to shifts in the pattern of withdraw- Northern America and Central Asia.
als. Agriculture continues to be the primary
water user at the global level, and accounted Total water withdrawals per capita remained
for some 2 950 km (72 percent) of total water
3
flat or declined from 2000 to 2018, except in
withdrawals in country reporting in 2018 Central America and the Caribbean, South-
(Ttable 1.13). This compares with an estimated ern America and Southeast Asia (Ffigure 1.22).
total of 2 703 km in 2006, which indicates
3
These trends are expected to persist as popu-
annual growth rates of about 0.8 percent per lations grow, partly due to overall increases
year. Approximately 483 km3 (12 percent) was in water productivity, including agriculture,
withdrawn for municipal use and 646 km 3
and partly due to the prevalence of water
(16 percent) for industry (Ttable 1.13). scarcity induced by extended periods of arid-
However, these figures vary significantly by ity in areas of high population density.
region. In Europe, agriculture withdraws only
30 percent, municipalities 26 percent and 1.7.2 Droughts
industry 45 percent. In South Asia, agricul-
and scarcity
ture withdraws 91 percent, municipalities
7 percent and industry 2 percent. High- Droughts are among the most complex and
income regions, such as Northern Amer- severe climate-related hazards encountered,
ica and Europe, have proportionally lower with wide-ranging and cascading impacts
withdrawals for agriculture compared across societies, ecosystems and economies.
with low-income countries. Residual flows They are recurrent, can last from a few weeks
retained in-stream or returned to shallow to several years, and affect large areas and
groundwater and draining to the marine populations around the world. Droughts have
environment still represent 88 percent of occurred throughout history, due to natural
renewable water resources at the global level, climate variability (UNDRR, 2021).
but this masks significant variation at the
Drought is a prolonged dry period in
regional level. Withdrawals and dam storage
the natural climate cycle that can occur
are estimated to account for an overall reduc-
anywhere in the world. It is a slow on-set
tion in natural pre-development flows (Pekel
phenomenon caused by a lack of rainfall.
et al., 2016; Schneider et al., 2017; Tessler et
Compounding factors, such as poverty and
al., 2018).
inappropriate land use, increase vulner-

66 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


FIGURE 1.20 Global distribution of large dams and reservoirs, 2016

GRanD v1.1 and v1.3 contain the locations and characteristics for 7,320 dams and reservoirs across the planet. GRanD
dams are snapped to the HydroSHEDS river network, which facilitates research on the size of rivers being dammed.

Total dams built Cumulative maximum Cumulative discharge of


storage in reservoirs impounded rivers
Version
GranD 1.1 GranD 1.3

600 1 500 000

Volume in Million cubic meters


90 000

Discharge in m3/s
400 1 000 000
Count

60 000

200 500 000


30 000

0 0 0

00
00

00
19 0

19 0

19 0

19 0
19 0

19 0
20 0

19 0

19 0
20 0

19 0
20 0

10
19 0
19 0
40

19 0

19 0

10

19 0
19 0
40

19 0

19 0

10

19 0
19 0
19 0
19 0

19 0
19 0

19 0

19 0
0

0
8

6
9

8
9

8
9
2
3

2
3

2
3
4
5

7
1

20
20

20
19

19
19

19

19
When focusing on reservoirs with storage greater than 100 million cubic metres (MCM), large dam and reservoir construction
peaked between 1960 and 1969. Cumulative volume of water impounded peaked later, between 1970 and 1979. Large reservoir
construction slowed considerably after these peaks. Though dam and reservoir construction has not returned to rates seen over the
middle of the 20th century, the size of rivers being dammed has increased. Fewer dams with large reservoirs were built between
2000 and 2016, but the cumulative discharge of rivers being impounded by large dams nearly reaches that of the much more active
decade between 1970 and 1979, indicating that recent dams are increasingly built on larger rivers.

GRanD Version

1.1 1.3

Reservoir capacity (km3)

<1

1-10

10-100 Number of reservoirs in GRanD

> 100 < 11 11-50 51-100 101-500 501-1000 > 1000

Source: Global Dam Watch. 2022. Research using core global dam datasets. Dotted line represents approximately the Line of Control in Jammu and Kashmir
agreed upon by India and Pakistan. The final status of Jammu and Kashmir has not
In: Global Dam Watch. http://globaldamwatch.org/our-research yet been agreed upon by the parties.

Final boundary between the Sudan and South Sudan has not yet been determined.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 67
Table 1.13 Total water and total freshwater withdrawals for human use, and
percentage of total water withdrawals, 2018
Region Agricultural Municipal Industrial Total water Total IRWR
water water water withdrawal freshwater
withdrawal withdrawal withdrawal withdrawal

km3/ % km3/ % km3/ % km3/year km3/year km3/year


year year year

Africa 186 79 36 15 15 6 237 222 3 935

Northern Africa 85 79 17 16 6 6 108 94 46

Sub-Saharan 101 78 19 15 9 7 129 128 3 889


Africa

Americas 500 56 120 13 274 31 894 896 19 673

Central America 24 63 8 20 7 17 38 35 1 209


and Caribbean

Northern 246 43 76 13 246 43 569 569 6 077


America

Southern 230 80 36 13 22 7 287 292 12 387


America

Asia 2 162 82 249 9 223 8 2 634 2 518 11 865

Central Asia 131 88 8 5 10 7 149 148 242

East Asia 462 65 102 14 150 21 714 709 3 410

South Asia 913 91 70 7 20 2 1 003 899 1 935

Southeast Asia 429 85 43 8 34 7 506 507 5 794

Western Asia 227 87 26 10 8 3 262 255 485

Europe 86 30 76 26 130 45 291 286 6 576

Eastern Europe 23 30 20 26 34 44 78 77 4 414


and Russian
Federation

Western and 63 29 55 26 96 45 213 209 2 163


Central Europe

Oceania 15 67 3 14 4 19 23 22 915

Australia and 15 68 3 14 4 19 23 21 819


New Zealand

Pacific Islands 0 59 0 30 0 11 0 0 96

Total general 2 950 72 483 12 646 16 4 079 3 944 42 964

Notes: IRWR = internal renewable water resources generated on country areas. Total water withdrawal includes use of desalinated water,
direct use of treated municipal wastewater and direct use of agricultural drainage water. Total freshwater withdrawal is defined as the sum of
surface water withdrawal extracted from rivers, lakes and reservoirs, and groundwater withdrawal extracted from aquifers. It does not include
non-conventional waters.

Source: FAO. 2021. AQUASTAT – FAO’s Global Information System on Water and Agriculture.
In: FAO. Rome. www.fao.org/aquastat/en

68 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


Total Internal Renewable Water Resources per capita [m3/inhab/year]
FIGURE 1.21 Total annual INTERNAL RENEWABLE WATER RESOURCES per capita by
geographical region, 2000, 2012 and 2018 (m3/capita)

2000 2012 2018

40 000

35 000

30 000

25 000

20 000

15 000

10 000

5 000

0
aa

rriicc nn

aann aa

aa

aa

aa

iiaa

iiaa

aa

iiaa

ppee

iiaa
rriicc

bbee rriicc

iicc

iicc

ssii

ssii
AAff aarraa

AAss

AAss

AAss

aann
aa

rroo
llAA

ttAA
eerr

eerr
AAff

rriibb mee

ccee
hh

sstt

hh

rrnn

EEuu
m

rraa
CCaa AAm

AAm

AAm

aass
SSaa

uutt
nn

EEaa

OO
ttee
nntt
eerr

hhee
SSoo
bb--

eess
dd aall

nn

rrnn

CCee
tthh

eerr

uutt
SSuu

aann nnttrr

W
hhee

W
oorr

tthh

SSoo
CCee

uutt
NN

oorr

SSoo
NN

Source: FAO. 2021. AQUASTAT – FAO’s Global Information System on Water and Agriculture. In: FAO. Rome. www.fao.org/aquastat/en

FIGURE 1.22 Total annualTotal


water withdrawals
water withdrawal per capita byper capita
geographical by[m3/inhab/year]
region geographical region, 2000,
2012 and 2018 (m3/capita)

2000 2012 2018

2 500

2 000

1 500

1 000

500

0
aa

rriicc nn

aann aa

aa

aa

aa

iiaa

iiaa

iiaa

iiaa

ppee

iiaa
rriicc

bbee rriicc

iicc

iicc

ssii
AAff aarraa

AAss

AAss

AAss

AAss

aann
aa

rroo
llAA
eerr

eerr
AAff

rriibb ee

ccee
hh

sstt

hh

tt

rrnn

EEuu
m

rraa
CCaa AAm

AAm

AAm

aass
SSaa

uutt
nn

EEaa

OO
ttee
nntt
eerr

hhee
SSoo
bb--

eess
dd aall

nn

rrnn

CCee
tthh

eerr

uutt
SSuu

aann ttrr

W
hhee

W
oorr

tthh

SSoo
n
CCee

uutt
NN

oorr

SSoo
NN

Source: FAO. 2021. AQUASTAT – FAO’s Global Information System on Water and Agriculture. In: FAO. Rome. www.fao.org/aquastat/en

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 69
ability to drought. When drought causes of extreme weather events. It changes the
water and food shortages, there can be average climate conditions and climate vari-
many impacts on the health of the popu- ability and generates new threats in regions
lation, which may increase morbidity that have little experience of dealing with
and result in death. In recent years, most drought. As with climate change, drought is
drought-related mortality has occurred slow to develop and not easily recognized at
in countries also experiencing political first but can quickly become a crisis when
and civil unrest. In the period from 1970 severe and damaging impacts emerge.
to 2012, drought caused almost 680 000
deaths, due to the severe African droughts FAO surveyed the drought characteristics and

of 1975, 1983 and 1984. (WMO, 2022b) management practices covering 2003–2013
in 48 developing countries in Latin America,
Drought needs to be distinguished from Africa and Asia, and found that agriculture
aridity. Drought is an immediate risk and can takes the brunt, absorbing over 80 percent
affect all regions and is not confined to drier of the economic losses. Crop production was
regions only. Occurrences of drought are most affected, accounting for 42 percent, and
unpredictable, but they come to an end, while livestock for 36 percent. FAO has published
aridity does not. In simple terms, a drought is detailed reports from this survey for the
a period when rainfall is less than “normal” Caribbean (FAO, 2016), Central Asia and
or “expected”, and there is not enough water Turkey (FAO, 2017b) and the Near East and
to meet the demands of human activities and North Africa (FAO, 2018c).
sustain environmental services. However,
not all droughts cause problems or become 1.7.3 Water withdrawals
crises; this depends on where and when they
flatten but consumption
increases
occur. “Agricultural drought” is usually the
first visible sign of drought. It can be short
lived, reduce crop yields, affect rangeland and Increasing global population and economic
forest productivity, and increase fire hazards. growth have been driving water withdrawals.
“Hydrological drought” follows, adversely The annual rate of increase peaked in the
affecting aquatic ecosystems, wetlands 1960s and has since been slowing, particularly
and river flows, leading to domestic water during the 2000s (Ffigure 1.23). From 2010
shortages. Finally, “socioeconomic drought” to 2018, municipal withdrawals increased
affects most aspects of life, including public by 3 percent, while agricultural withdraw-
health and economic growth, with impacts als increased by 5 percent, representing
lasting many months and even years, 72 percent of total withdrawals. Industrial
beyond the time when the meteorological withdrawals decreased by 12 percent from
drought is over and forgotten about. In rural 2010 to 2018, reflecting reductions in with-
areas, reduced crop productivity can lower drawals for thermal power production as
farm incomes and increase food prices, cooling processing has become more water
unemployment and migration. In vulnerable efficient.
communities, farm incomes can take many
years to recover after drought. In 2012, irrigation accounted for 90 percent
of all evaporation (consumptive use) induced
Climate change increases drought risk by by human activities (Hoogeveen et al., 2015).
increasing the frequency and magnitude Estimated crop water requirements in 2012

70 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


FIGURE 1.23 Evolution of global total water withdrawals, 1910–2018 (km3/year)

Municipalities Industries Agriculture

4 500

4 000

3 500

3 000
km 3/year

2 500

2 000

1 500

1 000

500

0
1900 1910 1920 1930 1940 1950 1960 1970 1980 1990 2000 2010 2017 2018

Source: FAO. 2021. AQUASTAT – FAO’s Global Information System on Water and Agriculture. In: FAO. Rome.
www.fao.org/aquastat/en

accounted for 1 507 km3 of total agricultural Non-conventional water sources (including
withdrawals (2 872 km in 2012). Evaporation
3 reclaimed wastewater and desalinated water)
from irrigated land increased from 1 268 km 3 account for only 0.12 percent of consumptive
in 2004 to 1 285 km3 in 2012 (Hoogeveen et irrigation use. The use of treated wastewater
al., 2015) and continues to place the most for irrigation is still small, but it is growing
significant pressure on river basin balances. as the marginal cost of treatment declines.
In some cases, total withdrawals and reduced The estimated volume of treated wastewa-
return flows result in basin closure (Molle ter from urban areas used for irrigation is
and Wester, 2009), indicating the sensitiv- 5 km3/year, and is concentrated in Southern
ity of hydrological circulation in subtropical America, the Near East and China. Estimates
zones in particular. Combined with antici- suggest 10 percent of the global irrigated land
pated impacts of climate change and rapid area receives untreated or partially treated
increases in demand from a predominantly wastewater, more than 30 million ha in 50
urban population, the pressure or stress countries (FAO, 2020f). Wastewater used for
on freshwater resources is set to continue. irrigation is one of the significant drivers of
Patterns of agricultural water withdrawals diffuse soil pollution. Even treated wastewa-
have changed since 2003, in response to ter still contains residues of contaminants
increased demand for calories and chang- not removed by modern technologies.
ing dietary demands, notably the growth in
consumption of animal protein (FAO, 2017a). There were approximately 18 thousand
desalination plants worldwide at the end of
2015, with a total installed annual produc-

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 71
1.7.4 Regional variations
in water stress
The SDG aggregate indicator 6.4.2 on water
©FAO/Giulio Napolitano

stress4 assesses the level of stress that human


activities are considered to exert on natural
freshwater resources (FAO and UN-Water,
2021). This is an aggregate (all-sector) indi-
cator and is taken as an overall measure of
tion capacity of 31 km3. Some 13.6 km3 physical water scarcity. At the global level,
(44 percent) was in the Near East and North SDG indicator 6.4.2 reached an average of
Africa (IWA, 2016), which is expected to grow 18 percent in 2018, but this masks substantial
by 7–9 percent annually. Desalination is also regional variations (Ffigure 1.24). In 2018,
likely to increase in Southern America, the Europe experienced a low stress level of
United States of America and Asia. Modest 8.3 percent. In comparison, the stress levels
amounts only are used directly for irrigation, in East Asia and Western Asia were about
mostly on high-value horticultural crops. 45 percent and 70 percent, respectively. In
However, desalinated water forms a high Central Asia and South Asia, they were over
percentage of urban wastewater in the Near 70 percent, while in Northern Africa, they
East and the Arabian Peninsula and is reused were above 100 percent.
for irrigation.
The SDG indicator 6.4.2 accounts for all
In 2021 the global desalination operating freshwater withdrawals relative to total
capacity was estimated at 28.6 km /year3 freshwater resources, including environ-
(78 349 678 m3/day) and 10 209 180 m3/day mental flow requirements for ecosystem
of desalination contracted capacity in 183 services. A withdrawal rate above 75 percent
countries in the world (IDA, 2021). The global of renewable water resources represents high
installed desalination capacity has been water stress, and more than 100 percent is
increasing steadily at the rate of about critical. High water stress can have devastat-
7 percent per annum since 2010 to the end ing consequences for the environment and
of 2019. Mega-plants are few in number, but hinder or even reverse economic and social
they supply most of the global desalination development.
capacity (Eke et al., 2020). The seawater and
There are concerns that SDG indicator 6.4.2,
brackish water desalination capacity has had
although useful as a broad aggregate indica-
a great increase of 6.4 million m3/day of
tor, can mask the recirculation of water use
new capacity in 2019 as a result of numer-
in river basins and aquifers (Vanham et al.,
ous mega-projects in the Gulf and Israel.
Desalination capacity is increasing across the
Middle East, but there are also a number of 4
SDG indicator 6.4.2 measures the level of water stress
large-scale projects in India. In the United and is defined as the ratio of total freshwater withdrawn
by all major sectors (agricultural, industrial and
States of America, seawater desalination municipal) to total renewable freshwater resources,
is slowly increasing (IDA, 2021). There is a after considering environmental flow requirements.
The ratio between 0 and 25 percent indicates no stress;
sharp rise in the desalination capacities in 25–50 percent indicates low stress; 50–75 percent
regions that did not have desalination in the indicates medium stress; 75–100 percent indicates
high stress; and more than 100 percent indicates
past, including Africa and Europe. critical stress.

72 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


FIGURE 1.24 Level of water stress by geographical region, 2006, 2009, 2012, 2015 and
2018

2018
World
2015
Oceania
2012
Europe 2009

2006
Southeast Asia

East Asia

Western Asia

Central Asia

South Asia

Southern America

Central America and Caribbean

Northern America

Sub-Saharan Africa

Northern Africa

0% 20% 40% 60% 80% 100% 120% 140%

Source: FAO. 2021. AQUASTAT – FAO’s Global Information System on Water and Agriculture. In: FAO. Rome. www.fao.org/aquastat/en

2018). Indices of stress that are more complex The SDG indicator 6.4.2 on water stress
are being developed to account for more vari- has been calculated at country level and
ables and sectors (e.g. the World Resources aggregated following the SOLAW regional
Institute AQUEDUCT indices; WRI, 2022) (Qin groupings. A different picture appears when
et al., 2019). The method of calculation in this aggregating the indicator by river basin
report is somewhat different from that used (Map 1.20). Water stress is high in all those
for SOLAW 2011, which expressed agricul- basins with intense irrigated agriculture, as
tural water stress as the ratio of irrigation well as in those including densely populated
consumption to river basin renewable water cities (e.g. Cape Town), which compete
resources (Hoogeveen et al., 2015) and not the with the agricultural sector for the use of
ratio of water withdrawals to renewable water water, and where there is less volume of
resources as used for SDG indicator 6.4.2. available freshwater resources due to climatic
The Hoogeveen et al. (2015) stress criterion conditions. Countries are encouraged to
considers water stress to be substantial when disaggregate at the sub-basin level to give
the incremental evaporation due to irrigation a more detailed picture of the distribution
exceeds 10 percent of the generated water of water stress. Basins affected by high or
resources in a river basin. A ratio exceeding critical water stress are located in regions of
20 percent indicates critical stress.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 73
Map 1.20 Level of water stress of all sectors by major basin, 2018

No stress (0 - 25%) Low (25% - 50%) Medium (50% - 75%) High (75% - 100%) Critical (>100%)

Dotted line represents approximately the


Line of Control in Jammu and Kashmir
agreed upon by India and Pakistan. The
final status of Jammu and Kashmir has not
yet been agreed upon by the parties.

Final boundary between the Sudan and


Source: FAO & UN-Water. 2021. Progress on level of water stress: Global status and acceleration needs for SDG indicator 6.4.2.
South Sudan has not yet been determined.
Rome. https://doi.org/10.4060/cb6241en. Modified UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420

high water stress, such as Northern America,


1.7.5 Keeping groundwater
the west coast of Central America, Northern
Africa, and South and Central Asia.
systems in play
Many countries are concerned about
Agriculture makes a significant contribu-
increasing dependence on groundwater
tion to water stress in countries with high
for domestic, industrial and agricultural
levels of water stress. Agricultural withdraw-
supply, as withdrawals are exhausting their
als account for a significant part of total
recoverable groundwater storage. Depletion
withdrawals in Northern Africa, the Middle
of non-renewable groundwater resources
East–Western Asia and Central Asia. Water
continues in the arid zones of Northern
stress due to agricultural withdrawals at the
America, Northern Africa, the Near East
water basin level shows the critical nature of
and Arabian Peninsula, and Central Asia,
the Nile River basin and river basins in the
where irrigated agriculture dominates
Arabian Peninsula and South Asia (Map 1.21).
total withdrawals. Recent reviews of large,
irrigated basins point to the growing role of

74 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


Map 1.21 Level of water stress due to the agriculture sector by basin, 2018

0 - 10% 10% - 25% 25% - 50% 50% - 75% 75% - 100%

Note: The contribution of the agriculture sector to water stress is defined as the ratio between total freshwater consumed by Dotted line represents approximately the
Line of Control in Jammu and Kashmir
the agricultural sector and total renewable freshwater resources, after considering environmental flow requirements. The
agreed upon by India and Pakistan. The
SDG water stress indicator 6.4.2 measures the contribution of the agriculture sector to water stress at the major basin level final status of Jammu and Kashmir has not
as follows: no stress – when the proportion of agricultural water withdrawal is between 0 percent and 25 percent; low stress yet been agreed upon by the parties.

– between 25 percent and 50 percent; medium stress – between 50 percent and 75 percent; high stress – between 75 percent
Final boundary between the Sudan and
and 100 percent; and critical stress – more than 100 percent.
South Sudan has not yet been determined.

Source: FAO & UN-Water. 2021. Progress on level of water stress: Global status and acceleration needs for SDG indicator 6.4.2. Rome.
https://doi.org/10.4060/cb6241en. Modified UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420

groundwater exploitation for irrigation and requirement to withdrawal ratio is assumed


the complex water management implications at 65 percent, reflecting lower conveyance
in sustaining the quantity and quality of losses associated with groundwater-sourced
groundwater (e.g. Lytton et al., 2021). irrigation (Siebert et al., 2010). This level
of withdrawal represents a 19 percent
Global groundwater withdrawals for irrigated increase relative to 2010, when an estimated
agriculture were estimated at 820 km3/year 688 km3/year was withdrawn for irrigated
based on aggregate country-level report- agriculture, and indicates annual growth rate
ing for 2018, and crop water requirements of 2.2 percent.
were over 33 percent of the total global
area equipped for groundwater irrigation Margat and van der Gun (2013) identified local
(see Tfable 1.3). For 2018, the crop water and regional aquifers with severe storage

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 75
for aquaculture (USGS, 2018). But at state
level in California, while surface water was
the primary source of irrigation water from
1950 to 2009, groundwater became the
© Prakash Singh

main source of supply between 2010 and


2015 as severe drought reduced surface water
withdrawals by 64 percent (USGS, 2018).

depletion, mainly associated with agricultural This evolution of conjunctive use may not

withdrawals for irrigation. Notable depletions be managed or planned, but it indicates a

occurred on continental aquifers associated global expansion of groundwater use in agri-

with agricultural plains and coastal margins. culture to service irrigation expansion, and,

Localized depletion in minor alluvial, coastal more significantly, intensification within

and island aquifers has been attributed existing surface command areas as in South

to agricultural withdrawals, leading to Asia (Shah, 2009). Recent updates of agri-

groundwater scarcity, pollution and saline cultural groundwater use, including those

intrusion, which threatens potable water of the United States Department of Agricul-

supply and limits agricultural production on ture (USDA National Agricultural Statistics

coastal aquifers. The impact of agricultural Service, 2019) and India’s 5th Census of Minor

withdrawals of groundwater presents a Irrigation Schemes (MoWR RD and GR, 2017),

complex picture. As more groundwater is all point to continued irrigation expansion in

pumped to surface water evaporation and which the proportion of irrigated land using

runoff to sea or saline sinks, withdrawals surface and groundwater sources conjunc-

can exceed the natural rates of groundwater tively can be expected to increase.

recharge and aquifer recovery. The result


Local and regional groundwater models
is a net gain to the atmosphere and oceans,
incorporating land-use changes and esti-
possibly accelerating sea-level rise (Wada,
mates of withdrawals and recharge can be
Van Beek and Bierkens, 2012). However,
used to track groundwater depletion (Koni-
for some aquifers, recharge patterns are
kow, 2013). However, verifying the scale and
changing as groundwater drawdown is
magnitude of depletion trends using remote
opening opportunities for higher recharge
sensing by monitoring water storage changes
rates (Konikow, 2013).
in the Earth’s crust (the National Aeronautics

Despite the range of individual aquifer and Space Administration Gravity Recovery

studies, there is no consistent reporting of and Climate Experiment satellite mission)

groundwater withdrawals and their relative has proved challenging (Famiglietti, 2014).

contribution to economic activity on land, This is largely due to the coarse resolution of

particularly for conjunctive use. Where the gravity anomalies used to infer storage

high-quality groundwater information changes (Vishwakarma et al., 2021). Modelled

is collected, the detail of these shifts estimates suggest that between 2000 and

becomes apparent. For example, across the 2009, global groundwater depletion for all

United States of America, some 117 km3 of uses was of the order of 113 km3/year (Döll

groundwater was withdrawn from principal et al., 2014), while other models suggest

aquifers in 2015. Agriculture accounted for volumes of the order of 304 km3/year for

83 km3, of which 79 km3 was withdrawn 2010, of which 228 km3/year (75 percent)

for irrigation and livestock and 4 km3 was was attributed to agriculture (Wada, Van

76 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


©FAO/Jake Salvador
1.7.6 Deteriorating
water quality
At the global level, agriculture remains the
dominant source of water pollution (mainly
Beek and Bierkens, 2012). Estimates of the diffuse or non-point pollution from agricul-
depletion attributed to irrigated production tural land), followed by human settlement
(Dalin et al., 2017) indicate this increased and industry. For 2010 it was estimated that
from almost 195 km3/year in 2000 to just over all annual non-consumed water (2 250 km3)
241 km3/year in 2010. was discharged into the environment as urban
wastewater (330 km3), industrial wastewater
In practice, quantifying aquifer storage including cooling water (660 km3) and agri-
depletion at the global scale remains conjec- cultural drainage (1 260 km3) (FAO and IWMI,
tural when pre-development states are not 2018). But agriculture is also a victim of the
documented and when aquifer systems are deterioration of water quality. Saline water
being actively pumped. Boundary and leak- significantly decreases agricultural produc-
age conditions are continually changing, and tivity, with major implications for global and
recharge capture is variable, so categoric local food security. Recent estimates indi-
distinctions between renewable and non- cate that food losses caused by the presence
renewable groundwater storage are often of saline water in soils and groundwater is
not possible without detailed hydrochemical equivalent to the annual food requirements
and isotope verification. However, models of 170 million people (Damania et al., 2019).
now include measured piezometric heads Under climate change, the consequences of
as a valuable indicator of storage deple- rainfall variability and increased temperature
tion and are available for many local and are expected to translate into further deterio-
regional aquifers (e.g. Haacker, Kendall and ration of water quality.
Hyndman, 2016).
The capacity of soils to store, buffer and
The trends in groundwater storage and the degrade water-borne contaminants is being
risks of groundwater depletion must be exceeded by anthropogenic treatment of
taken together with the build-up of aquifer soils on cropland and pasture to the point
pollution from anthropogenic sources and where elevated levels of nitrogen, salinity
the migration of geogenic pollution, nota- and biological oxygen demand in freshwater
bly arsenic and fluoride. By 2018, in China, are widespread (Map 1.22).
only 63 percent of groundwater was consid-
ered potable. In Southeast Asia, arsenic and
fluoride were the most common geogenic
pollutants. However, fertilizers and pesti-
cides remain some of the main sources of
anthropogenic pollution. In Europe, nitrate
pollution was the most common cause of
poor groundwater quality, with 23 percent
© Antonello Proto

of groundwater bodies exceeding European


Union groundwater nitrate standards (FAO
and IWMI, 2018).

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 77
Agricultural use of reactive nitrogen has communities and mammal herbivores (graz-
continued to increase since 2000, from ing animals) (Sutton et al., 2011; Stevens,
almost 81 million tonnes to a peak of David and Storkey, 2018). It was estimated
110 million tonnes in 2017, with signs of a that the annual cost of the environmental
slight decline in 2018. Industrial fertilizer impacts of nitrogen pollution in the Euro-
production and biological fixation of nitro- pean Union was between EUR 70 billion and
gen in agriculture account for 80 percent EUR 320 billion in 2012 (EC, 2013).
of anthropogenic nitrogen fixation (Stevens,
David and Storkey, 2018). In agricultural The global growth rate of phosphorus use in

systems, reactive nitrogen is a major threat agriculture is modest, from 32 million tonnes

to water quality (eutrophication of surface in 2000 to a peak of 45 million tonnes in

water), soil quality (soil acidification, changes 2016 (Ffigure 1.18 and Map 1.23). Nutrient

in SOM content and loss of soil biodiversity), phosphate is one of the essential nutrients

plant biochemistry, insects (i.e. pollina- required for plant growth and development,

tors), functional composition of vegetation but when leached from cultivated soils, it

Map 1.22 Global water quality risk for three SUSTAINABLE DEVELOPMENT GOAL 6.3.2
indicators (nitrogen, electrical conductivity and BIOLOGICAL OXYGEN
DEMAND), modelling of the Global Freshwater Quality Database data
2000–2010 at 50 km resolution

Low risk High risk

Note: This figure maps a water quality index summarizing global predictions for biological oxygen demand, electrical conductivity and nitrogen.
Each value is scaled to a common support for comparability, then summed together. Average values for 2000–2010 are displayed. Grey areas
have no data for one or more parameters.

Source: World Bank Group. 2019. Quality unknown: The invisible water crisis, R. Damania, S. Desbureaux, A.-S. Rodella, J. Russ & E. Zaveri, eds.
Washington, DC. https://openknowledge.worldbank.org/bitstream/handle/10986/32245/9781464814594.pdf?sequence=8&isAllowed=y. Modi-
fied UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420

78 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


can cause freshwater eutrophication (FAO 39 million tonnes in 2018. The impact on
and IWMI, 2018). Estimates indicate that freshwater eutrophication is not marked, as
the total phosphorus input to water bodies it is for nitrogen and phosphorus, although it
from anthropogenic sources is about contributes to runoff salinity.
1.5 million tonnes annually, with 62 percent
from point sources (domestic and industrial) The global trend in the intensification of

and 38 percent from diffuse sources agricultural production is also testing the

(agriculture) (Mekonnen and Hoekstra, 2018). capacity of the receiving freshwater to

There has also been a significant increase dilute pollutants, some of which are highly

in the annual atmospheric deposition of persistent and resistant to breakdown.

nitrogen since the 1900s, from 1.9 Tg of The global distribution of the water pollu-

nitrogen in 1900 to 3.8 Tg of nitrogen in tion threat from human activities includ-

2000, of which 63 percent was deposited on ing nitrogen loading, phosphorus loading,

agricultural land (Sutton et al., 2011). mercury deposition, pesticide loading,


organic loading, salinization, acidification
Agricultural use of potash rose from and sediment loading has been summarized
22 million tonnes in 2000 to a peak of almost by Sadoff et al. (2015). Of particular concern

Map 1.23 ANNUAL Anthropogenic phosphorus inputs into freshwater systems


from agriculture, industrial and domestic sectors, 2002–2010 (kg P/ha)

0
0-0. 5
0.5-1
1- 1. 5
1. 5 - 2
2- 2. 5
2. 5 - 5
>5

Dotted line represents approximately the


Line of Control in Jammu and Kashmir
agreed upon by India and Pakistan. The
final status of Jammu and Kashmir has not
yet been agreed upon by the parties.

Source: Mekonnen, M.M. & Hoekstra, A.Y. 2018. Global anthropogenic phosphorus loads to freshwater and asso-
Final boundary between the Sudan and
ciated grey water footprints and water pollution levels: A high-resolution global study. Water Resources Research, South Sudan has not yet been determined.
54(1): 345–358. Modified UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 79
is pollution caused by emerging chemical GIS-based modelling found that 65 percent
contaminants, including pesticides, livestock (35.9 million ha) of downstream irrigated
pharmaceuticals and plastics, for which there croplands is located in catchments with high
is currently little regulation or monitoring. levels of dependence on urban wastewater
Recent compilation of gridded data for active flows. Of these croplands, 29.3 million ha is
ingredients (Maggi et al., 2019) has allowed located in countries with low levels of waste-
the accumulation of active ingredients in water treatment exposing 885 million urban
pesticides to be mapped at the global level residents to health risks (Thebo et al., 2017).
(Tang et al., 2021) (Map 1.24).
1.7.7 Environmental
continuity at
The use of wastewater for irrigation, if not
well managed, also has the potential of caus-
ing health issues and environmental degra- breaking point
dation and groundwater pollution. Large
The global environment outlook report
areas of irrigated fields rely on the same
(UNEP, 2019) and the United Nations System
surface water sources of urban areas without
of Environmental Economic Accounting
wastewater treatment capacity. A study using

Map 1.24 Global regions of concern (global areas susceptible to pesticide


pollution), 2010

areas of high
concern

areas with no concern


no agricultural areas

water/no data

Dotted line represents approximately the Line of Control in Jammu


Sources: Tang, F.H.M., Lenzen, M., McBratney, A. & Maggi, F. 2021. Risk of pesticide pollution at the
and Kashmir agreed upon by India and Pakistan. The final status of
global scale. Nature Geoscience, 14(4): 206–210; data from Tang, F.H.M., Lenzen, M., McBratney, A. & Jammu and Kashmir has not yet been agreed upon by the parties.
Maggi, F. 2021. Global pesticide pollution risk data sets. In: figshare.
https://figshare.com/articles/dataset/Global_pesticide_pollution_risk_data_sets/10302218/1 Final boundary between the Sudan and
South Sudan has not yet been determined.
Modified UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420
Final status of the Abyei area is not yet determined

80 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


(SEEA, 2022) for natural resource accounting
confirm trends in the loss of environmental
services and biodiversity as natural land-
scapes are lost to cultivated land. Freshwater
withdrawals and drainage from agricultural
land (including irrigation return flows) place

©FAO/Giulio Napolitano
the most significant sectoral pressure on river
basin water balances (FAO and IWMI, 2018).
As this plays out in specific river basins, the
hydraulic continuity of downstream wetlands
and associated ecosystem services is compro-
mised. The patterns of surface water flows
presumptive standard proposed by Gleeson
and aquifer recharge have been disrupted to
and Richter (2018) is that a high level of
such an extent that some basin freshwater
ecological protection is provided when daily
systems are now considered “closed” (Molle
streamflow alterations are no greater than
and Wester, 2009).
10 percent.

Estimates suggest only 37 percent of rivers


Estimates suggest the world has lost
longer than 1 000 km remain free-flowing
70 percent of natural wetlands over the
over their entire length, and only 23 percent
past century, including significant losses
flow uninterrupted to oceans (Grill et al.,
of freshwater species (Gardner et al., 2015),
2019). For all other river reaches with modi-
and decline in food security and nutrition
fied magnitude, frequency and flow dura-
(e.g. Turyahabwe et al., 2013). Of the 29 500
tion can expect to adversely affect suites of
freshwater-dependent species so far assessed
aquatic habitats and ecosystem processes.
for the International Union for Conservation
The disruption to reservoir storage and
of Nature (IUCN) Red List, 27 percent are
flow diversion for agricultural purposes,
under threat of extinction (Lausche, 2019;
as opposed to hydropower, cooling water
Tickner et al., 2020a).
and municipal uses, can be assessed only at
individual catchment or basin levels. The The impact of agricultural practices on fresh-
International Commission on Large Dams water ecosystems has long been recognized
database of registered large dams confirms at all scales from the Azraq oasis in eastern
almost 50 percent of large dams (13 580 dams) Jordan to the Aral Sea in Central Asia. In
are dedicated to irrigation and 24 percent practice, the impacts on water quantity and
(6 278 dams) of multipurpose dams have water quality can be hard to untangle and
irrigation functions. attribute definitively, as each downstream
impact may have multiple causes. None-
Estimates suggest that the decline of ground-
theless, the conversion of wetlands to rice
water level (piezometric head) (de Graaf et
paddy is expected to have made a significant
al., 2019) need only to be less than 1.0 m
contribution to the 40 percent decline in
before a prescriptive limit of daily flow alter-
inland and coastal wetlands between 1970
ations of not more than 10 percent is reached
and 2008 (Leadley et al., 2014), with the
in many aquifers servicing irrigated areas.
overall decline in natural wetlands noted in
Flow disruption of more than 10 percent is
a Ramsar Convention on Wetlands briefing
presumed to remove a high level of ecological
note (Gardner et al., 2015). The impact of irri-
protection (Gleeson and Richter, 2018). The
gation technology has also been highlighted

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 81
tropics and mountain regions have exhibited
slower increases in productivity but have
proved more vulnerable to food insecurity
and poverty. Many uses of land and water
©FAO/Olivier Thuillier

systems are continuing to impose negative


impacts on ecosystem services.

World food demand is expected to increase


by 50 percent over the next 30 years, with
the greatest needs in developing countries.
with respect to maintaining desired envi- While it is expected that production will
ronmental flows (Linstead, 2018), but attrib- respond to rising demand, this will not be
uting in-stream flow volumes and timing the only measure of success. The environ-
to specific water-conservation measures is mental sustainability of the main land and
restricted to specific basins. water systems and their capacity to satisfy
the livelihood requirements of urban and
The term “environmental flows” is now
rural populations will be essential criteria.
commonly used to refer to a flow regime
designed to maintain a river or stream in Trade-offs between production and the
some agreed ecological condition (IRF, environment should be important for
2007). Recent reviews of environmental flow policymakers. Such trade-off decisions
concepts, methods and tools are available will require sound data and information
(Acreman, 2016; Poff, Tharme and Arthing- to fully understand the consequences of
ton, 2017; World Bank Group, 2018). The socioeconomic outcomes and environmental
desired environmental flows have been impacts. Decisions taken will need to
established at the country level for compiling include ways of reducing the risks and their
SDG indicator 6.4.2 on water stress (Sood et impacts to avoid further degrading natural
al., 2017). However, comprehensive analy- resources while maintaining food security
sis of environmental flow implementation and poverty targets.
and its impact across representative scales
and types of river basins still requires more
research (Tickner et al., 2020b).

1.8 Conclusions
This chapter has established the global state
of land, soil and water resources and trends
in their use, in response to the pressures
and drivers as demands change and increase.
Most of the additional growth in agricul-
tural production since SOLAW 2011 has been
© FAO/Simon Maina

derived from intensification, particularly on


prime agricultural land combined with irri-
gation. By contrast, rainfed systems in the

82 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


© FAO/Karen Minasyan
In focus: Fragile mountain agriculture

Mountains5 host important upland ecosystems that support the livelihoods of an estimated 1.1 billion
people. They are often referred to as the “water towers” of the world for their role in generating high
volumes of orographic rainfall and also providing over-year storage of freshwater in glaciers and
snowpacks. Their steep environmental gradients and climatic conditions are hosts to unique ecosys-
tems comprising a wide range of biomes. As the impacts of climate change take hold, the sensitivity
of their hydrology and related biomes has become apparent as they experience glacier retreat and
higher rates of erosion, compromising the capacity of downstream reservoirs.

Mountain agriculture is linked to water availability. In the Andes, the mountains provide water to over
75 million people in the region and a further 20 million people downstream. Most of the water is
used for agriculture, but also for hydropower and other industries. During dry periods, about 800 000
people depend on glacial water for 25 percent of their water needs (Alfthan et al., 2018).

Glacier and snowpack meltwater baseflows are increasing, yet these are essential during the
dry season for agriculture and other human needs (Biemans et al., 2019). Water availability for
1.9 billion people living in or directly downstream of mountainous areas is vulnerable to climatic and
socioeconomic changes (Immerzeel, Lutz and Andrade, 2019).

Mountain agriculture is characterized by small and fragmented plots of land with low carbon foot-
prints and time-consuming and labour-intensive cultivation. Agricultural practices aid ecosystem
conservation and restoration, and it is essential to protect soils against avalanches and floods. Farming
is predominantly carried out by families and is based on relatively high agrobiodiversity, producing
nutritious and diversified foods. In comparison to plain regions, mountains contain more diversity:
altitude changes and varied landscapes have created a multitude of ecological zones, with highly
genetically variable agricultural crops and farm animals (FAO, 2019g).

Mountain communities are preserving many of the rarest crop varieties, and have developed valuable
traditional knowledge and techniques in crop cultivation, livestock production and water harvesting that
help to sustain entire ecosystems. Terracing is widely practised; if properly planned and maintained, it
helps to stabilize the land, reduces soil erosion and prevents nutrients from being washed away.

5
For a definition of mountains, see Mountain Partnership (2015).

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 83
Upland soils are poorly developed, skeletal, shallow, acidic and relatively infertile (FAO, 2015). As
elevation increases, soils become shallower and less fertile because of soil erosion and low tempera-
tures that limit biological activities. They are often degraded due to nutrient leaching and water and
wind erosion in exposed areas. As a result, mountain soils are often less productive than lowland
soils. Globally, 45 percent of the world’s mountain areas are either unsuitable or marginally suitable
for growing crops, raising livestock or forestry (FAO, 2015). In cold mountain areas, freeze–thaw cycles
reduce the aggregation of soils and consequently affect their stability, fertility and water retention.

Production systems
Under the mountain environment, a range of farming systems have developed to cope with variations
in climate, slope and elevation, which could be classified into five systems (see the box).

Arable crops and permanent fruit crops are usually grown at low altitudes, while permanent grassland
and animal grazing are more common at higher altitudes. For example, pastoral livestock production
continues in the Tibetan steppe above 4 000 m (Sheehy, Miller and Johnson, 2006). In the Indian Garhwal
region, more than 40 crops are cultivated between 300 m and 3 000 m above sea level (FAO, 2015).

Five mountain production systems

Pastoral livestock production systems: These are grazing-based production systems whereby livestock are
fed on natural vegetation and rangelands that include grasses, legumes, shrubs and other vegetation to
provide forage throughout the year. Excessive grazing may cause degradation of rangelands, soil erosion and
loss of biodiversity. Rangeland degradation is increasing; it is crucial to halt and reverse this process globally.

Agropastoral livestock production systems: These are integrated crop–livestock–rangeland production


systems that include: different types of livestock; natural pastures and various field crops such as barley,
forage crops, shrubs and trees; and by-products of field crops, contributing to food security and nutrition in
the mountain area. These integrated systems involve a socioeconomic and policy environment in addition
to a market component that incorporates different factors to ensure an efficient and productive integrated
livestock–rangeland–crop production system.

Rainfed agriculture production systems, including fruit trees: In tropical and non-tropical areas, rainfed
agriculture occurs in areas that receive more than 400 mm of rain during the rainy season. Worldwide, rainfed
agriculture is often used as a conservation agriculture approach, meaning minimum soil disturbance or zero
tillage, stubble retention and crop rotation. Conserving soil moisture and reducing soil erosion in rainfed
agriculture production systems is crucial to ensure the sustainability of soil productivity, soil conservation and
water conservation.

Irrigated agriculture production systems, including fruit trees: Irrigated agriculture systems are practised in
arid and semi-arid mountain areas, where annual rainfall is less than 350 mm. The sources of irrigation water
are deep artesian wells, surface water from rivers or harvested rainwater in macro and micro water catchments
and dams. Farmers using irrigated mountain agriculture production systems tend to diversify production to
ensure food security with high-value crops including vegetables, fruit trees and ornamentals.

Forestry or agroforestry production systems: These are important sources of livelihoods in mountain areas
and provide essential environmental goods and services, such as timber, fuelwood, carbon storage and other
products that improve the lives of people living in mountain communities.

Source: FAO. 2019. Forests: Nature-based solutions for water. UNASYLVA, 70(251).

84 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


Mountain agriculture in Brazil, Afghanistan and Armenia
© FAO/Joao Roberto Ripper

© FAO/Giulio Napolitano

© Johan Spanner
Soil degradation
Mountain soils are intrinsically vulnerable and sensitive to degradation processes such as water
erosion and chemical and physical quality loss (FAO, 2015).

Soil erosion is common, and a destructive consequence of development. In Nepal, degraded red-soil
sites are responsible for 40 percent of the sediment load in rivers and for clogging irrigation canals and
local streams, thus increasing flood events (FAO, 2015). Agriculture is just one of many development
activities that accelerate soil erosion; road building, trail use, excavation, extractive activities and
construction also contribute (Harden, 2001).

Terracing is a frequently used means of reducing erosion. The method has been used for many
centuries across the world (Moreno-de-las-Heras et al., 2019). However, the most efficient approach
is to maintain soil cover. Annual erosion rates of less than 1 tonne/ha were recorded for rice crops
compared to over 80 tonnes/ha for cassava or bare soil terraces. Intermediate annual values, between
10 tonnes/ha and 40 tonnes/ha, were found on terraces with weeds, ginger or mixed rainfed cropping
(Arnáez et al., 2015).

In some mountain areas, mostly in marginal areas with difficult access, cultivated terraces are being
abandoned due to socioeconomic and technological changes. Although they are no longer being
maintained and are losing their soil-conservation function, they are being colonized by vegetation,
effectively controlling erosion.

Mountain systems, generally characterized by lower temperatures and higher precipitation than other
landscapes, have higher SOC stocks compared to lower-altitude systems (FAO, 2019h). Mountain soils
with permafrost contain approximately 66 Pg of SOC, which is 4.5 percent of the global pool. High-
elevation and high-latitude soils are experiencing warmer air temperatures and permafrost and a
thickening of the active layer (Bockheim and Munroe, 2014), and are highly endangered by
climate change.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 85
References
Acreman, M. 2016. Environmental flows—basics for novices. Wiley Interdisciplinary Reviews: Water,
3(5): 622–628.

Alfthan, B., Gjerdi, H., Puikkonen, L., Andresen, M., Semernya, L., Schoolmeester, T. & Jurek, M.
2018. Mountain adaptation outlook series: Synthesis report. Nairobi and Vienna, United Nations Envi-
ronment Programme and Arendal, Norway, GRID-Arendal. www.grida.no/publications/426

Alhumaid, N. 2020. KSA water report: Building a better food and water balance. Food & Agribusiness
Report. Farrelly & Mitchell.

Arnáez, J., Lana-Renault, N., Lasanta, T., Ruiz-Flaño, P. & Castroviejo, J. 2015. Effects of farming
terraces on hydrological and geomorphological processes. A review. CATENA, 128: 122–134.

Barger, N., Gardner, T. & Mahesh, S. 2018. Chapter 3: Direct and indirect drivers of land degradation and
restoration. In: A. Montanarella, L. Scholes & R. Brainich, eds. The IPBES assessment report on land
degradation and restoration. Bonn, Intergovernmental Science-Policy Platform on Biodiversity
and Ecosystem Science. https://ipbes.net/assessment-reports/ldr

Biemans, H., Siderius, C., Lutz, A.F., Nepal, S., Ahmad, B., Hassan, T., von Bloh, W. et al. 2019.
Importance of snow and glacier meltwater for agriculture on the Indo-Gangetic Plain. Nature
Sustainability, 2(7): 594–601.

Bockheim, J. & Munroe, J. 2014. Organic carbon pools and genesis of alpine soils with permafrost: A review.
Arctic, Antarctic, and Alpine Research, 46(4): 987–1006.

Boots, B., Russell, C.W. & Green, D.S. 2019. Effects of microplastics in soil ecosystems: Above and below
ground. Environmental Science and Technology, 53(19): 11496–11506.

Borrelli, P., Robinson, D.A., Fleischer, L.R., Lugato, E., Ballabio, C., Alewell, C., Meusburger, K. et al.
2017. An assessment of the global impact of 21st century land use change on soil erosion. Nature Commu-
nications, 8(2013).

Borrelli, P., Robinson, D.A., Panagos, P., Lugato, E., Yang, J.E., Alewell, C., Wuepper, D., Montanarella,
L. & Ballabio, C. 2020. Land use and climate change impacts on global soil erosion by water (2015-2070).
Proceedings of the National Academy of Sciences of the United States of America, 117(36): 21994–22001.

Cherlet, M., Hutchinson, C., Reynolds, J., Hill, J., Sommer, S. & von Maltitz, G. 2018. World atlas of
desertification. Luxembourg, Publications Office of the European Union. https://publications.jrc.ec.eu-
ropa.eu/repository/handle/JRC111155

86 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


Ciais, P., Sabine, C., Bala, G., Bop, L., Brovkin, V., Canadell, J., Chhabra, A. et al. 2013. Carbon and
other biogeochemical cycles. In: T.F. Stocker, D. Qin, G.-K. Plattner, M. Tignor, S.K. Allen, J. Boschung, A.
Nauels, Y. Xia, V. Bex & P.M. Midgley, eds. Climate change 2013: The physical science basis. Contribution
of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change.
Cambridge, UK, Cambridge University Press. www.ipcc.ch/site/assets/uploads/2018/02/WG1AR5_
Chapter06_FINAL.pdf

Coppus, R. 2022. Global distribution of land degradation. Thematic Background Report for SOLAW 2021.
Rome, FAO. www.fao.org/land-water/solaw2021/en

Crump, J. 2017. Smoke on water: Countering global threats from peatland loss and degradation. Nairobi,
United Nations Environment Programme, and Arendal, Norway, GRID-Arendal.

Daliakopoulos, I.N., Tsanis, I.K., Koutroulis, A., Kourgialas, N.N., Varouchakis, A.E., Karatzas, G.P. &
Ritsema, C.J. 2016. The threat of soil salinity: A European scale review. Science of the Total Environment,
573: 727–739.

Dalin, C., Wada, Y., Kastner, T. & Puma, M.J. 2017. Groundwater depletion embedded in international
food trade. Nature, 543(7647): 700–704.

Damania, R. Desbureaux, S., Rodella., A., Russ J. & Zaveri E. 2019. Quality unknown: The invisible
water crisis. Washington, DC, World Bank. https://elibrary.worldbank.org/doi/abs/10.1596/978-1-
4648-1459-4

de Graaf, I.E.M., Gleeson, T., (Rens) van Beek, L.P.H., Sutanudjaja, E.H. & Bierkens, M.F.P. 2019.
Environmental flow limits to global groundwater pumping. Nature, 574(7776): 90–94.

Department of Primary Industries and Regional Development, Government of Western Australia.


2021. Dryland salinity in Western Australia. In: Agriculture and Food. South Perth. Cited 9 February 2022.
www.agric.wa.gov.au/soil-salinity/dryland-salinity-western-australia-0

Döll, P., Schmied, H.M., Schuh, C., Portmann, F.T. & Eicker, A. 2014. Global-scale assessment of
groundwater depletion and related groundwater abstractions: Combining hydrological modeling with
information from well observations and GRACE satellites. Water Resources Research, 50(7): 5698–5720.

EC (European Commission). 2013. Nitrogen pollution and the European environment: Implications for air
quality policy. Science for Environment Policy Series. Brussels. https://ec.europa.eu/environment/inte-
gration/research/newsalert/pdf/IR6_en.pdf

EC. 2020. Global surface water explorer. In: EC. https://global-surface-water.appspot.com/#data

EC. 2021. Index of /ftp/jrc-opendata/GSWE. https://jeodpp.jrc.ec.europa.eu/ftp/jrc-opendata/GSWE/

EEA (European Environment Agency). 2019. The European environment — state and outlook 2020:
Knowledge for transition to a sustainable Europe. Luxembourg, Publications Office of the European Union.
https://doi.org/10.2800/96749

Eke, J., Yusuf, A., Giwa, A. & Sodiq, A. 2020. The global status of desalination: An assessment of current
desalination technologies, plants and capacity. Desalination, 495: 114633.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 87
El-Bahi, S.M., Sroor, A., Mohamed, G.Y. & El-Gendy, N.S. 2017. Radiological impact of natural radio-
activity in Egyptian phosphate rocks, phosphogypsum and phosphate fertilizers. Applied Radiation and
Isotopes, 123: 121–127.

Famiglietti, J.S. 2014. The global groundwater crisis. Nature Climate Change, 4: 945–948.

FAO. 2011. The state of the world’s land and water resources for food and agriculture: Managing systems at
risk. Rome, FAO and London, Earthscan. www.fao.org/3/i1688e/i1688e.pdf

FAO. 2013a. Tackling climate change through livestock: A global assessment of emissions and mitigation
opportunities. Rome. www.fao.org/3/i3437e/i3437e.pdf

FAO. 2013b. Land degradation assessment in drylands. Rome. www.fao.org/3/a-i3241e.pdf

FAO. 2015. Understanding mountain soils: A contribution from mountain areas to the International Year of
Soils 2015. Rome. www.fao.org/3/a-i4704e.pdf

FAO. 2016. Drought characteristics and management in the Caribbean. FAO Water Reports 42. https://relief-
web.int/sites/reliefweb.int/files/resources/Drought%20characteristics%20and%20mgmt%20in%20
the%20Caribbean.pdf.

FAO. 2017a. The future of food and agriculture: Trends and challenges. Rome. www.fao.org/3/a-i6583e.pdf

FAO. 2017b. Drought characteristics and management in Central Asia and Turkey. FAO Water Reporst 44.
Rome. www.fao.org/3/a-i6738e.pdf

FAO. 2018a. The future of food and agriculture: Alternative pathways to 2050. Summary version. Rome.
www.fao.org/3/CA1553EN/ca1553en.pdf

FAO. 2018b. World livestock: Transforming the livestock sector through the Sustainable Development Goals.
Rome. https://doi.org/10.4060/ca1201en

FAO. 2018c. Drought characteristics and management in North Africa and the Near East. FAO Water Report
45. Rome. www.fao.org/3/CA0034EN/ca0034en.pdf

FAO. 2019a. Water use in livestock production systems and supply chains: Guidelines for assessment (Version
1). Rome. www.fao.org/3/ca5685en/CA5685EN.pdf

FAO. 2019b. The state of the world’s biodiversity for food and agriculture, J. Bélanger & D. Pilling, eds.
Rome, FAO Commission on Genetic Resources for Food and Agriculture Assessments. www.fao.org/3/
CA3129EN/CA3129EN.pdf

FAO. 2019c. Measuring and modelling soil carbon stocks and stock changes in livestock production systems:
Guidelines for assessment (Version 1). Rome. www.fao.org/3/ca2934en/ca2934en.pdf

FAO. 2019d. Soil erosion: The greatest challenge for sustainable soil management. Rome. www.fao.org/3/
ca4395en/ca4395en.pdf

88 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


FAO. 2019e. Fertilizers archive. Consumption of total fertilizers between 1961 and 2002. In: FAOSTAT. Rome.
www.fao.org/faostat/en/#data/RA

FAO. 2019f. Pesticides use. Total pesticides use in agriculture from 2000 to 2017. In: FAOSTAT. Rome. www.
fao.org/faostat/en/#data/RP

FAO. 2019g. Mountain agriculture: Opportunities for harnessing zero hunger in Asia. Bangkok. www.fao.
org/3/ca5561en/CA5561EN.pdf

FAO. 2019h. GLOSIS - GSOCmap (v1.5.0). Global soil organic carbon map. Contributing countries. In: FAO.
Rome. http://54.229.242.119/GSOCmap/

FAO. 2020a. FAOSTAT. In: FAO. Rome. www.fao.org/faostat/en/#data/QC

FAO. 2020b. Global forest resources assessment 2020: Key findings. Rome. https://doi.org/10.4060/ca8753en

FAO. 2020c. The state of the world’s forests 2020: Forests, biodiversity and people. Rome. https://doi.
org/10.4060/ca8642en

FAO. 2020d. The state of world fisheries and aquaculture: Sustainability in action. Rome. https://doi.
org/10.4060/ca9229en

FAO. 2020e. Peatland mapping and monitoring: Recommendations and technical overview. Rome. https://
doi.org/10.4060/ca8200en

FAO. 2020f. The state of food and agriculture 2020. Overcoming water challenges in agriculture. Rome.
https://doi.org/10.4060/cb1447en

FAO. 2021a. The share of agri-food systems in total greenhouse gas emissions: Global, regional and country
trends 1990–2019. FAOSTAT Analytical Brief Series No. 31. Rome. www.fao.org/3/cb7514en/cb7514en.pdf

FAO. 2021b. AQUASTAT – FAO’s Global Information System on Water and Agriculture. In: FAO. Rome. www.
fao.org/aquastat/en

FAO. 2022a. WaPOR, remote sensing for water productivity. In: FAO. Rome. Cited 9 February 2022. www.fao.
org/in-action/remote-sensing-for-water-productivity/en

FAO. 2022b. International Network of Salt-affected Soils. In: Global Soil Partnership. Rome. www.fao.
org/global-soil-partnership/insas

FAO & IIASA (International Institute for Applied Systems Analysis). 2021. Global agro-ecological zones
v4.0 – Model documentation. Rome. www.fao.org/nr/gaez/publications/en

FAO & ITPS (Intergovernmental Technical Panel on Soils). 2015. Status of the world’s soil resources:
Main report. Rome. www.fao.org/3/i5199e/i5199e.pdf

FAO & IWMI (International Water Management Institute). 2018. More people, more food, worse water?
A global review of water pollution from agriculture, J. Mateo-Sagasta, S. Marjani & H. Turral, eds. Rome.
www.fao.org/3/ca0146en/ca0146en.pdf

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 89
FAO & UNEP (United Nations Environment Programme). 2021. Global assessment of soil pollution:
Report. Rome. https://doi.org/10.4060/cb4894en

FAO & UN-Water. 2021. Progress on level of water stress: Global status and acceleration needs for SDG
indicator 6.4.2. Rome. https://doi.org/10.4060/cb6241en

FAO, Intergovernmental Technical Panel on Soils, Global Soil Biodiversity Initiative, Convention on
Biological Diversity & European Commission. 2020. State of knowledge of soil biodiversity: Status,
challenges and potentialities. Report 2020. Rome, FAO. https://doi.org/10.4060/cb1928en

FAO, International Fund for Agricultural Development, United Nations Children’s Fund, World
Food Programme & World Health Organization. 2018. The state of food security and nutrition in
the world 2018. Building climate resilience for food security and nutrition. Rome. www.fao.org/3/
i9553en/i9553en.pdf

FAO, International Fund for Agricultural Development, United Nations Children’s Fund, World Food
Programme & World Health Organization. 2021. The state of food security and nutrition in the world
2021. Transforming food systems for food security, improved nutrition and affordable healthy diets for all.
Rome, FAO. https://doi.org/10.4060/cb4474en

Fischer, G. & van Velthuizen, H. 2018. Climate change impacts on suitability of main crops in the DECCMA
study areas in Ghana and in South Asia. Deltaic environments, vulnerability and climate change: The
role of mitigation as an adaptation and its policy implications (DECCMA). GCP/GLO/546/USH. Laxen-
bourg, Austria.

Fischer, T., Byerlee, D. & Edmeades, G. 2014. Crop yields and global food security. Monograph No. 158.
Canberra, Australian Centre for International Agricultural Research.

Fisher, J., Montanarella, L. & Scholes, R. 2018. Chapter 1: Benefits to people from avoiding land degrada-
tion and restoring degraded land. In: L. Montanarella, R. Scholes & A. Brainich, eds. The IPBES assessment
report on land degradation and restoration. Bonn, Intergovernmental Science-Policy Platform on Biodi-
versity and Ecosystem Services. https://ipbes.net/assessment-reports/ldr

Fuglie K., Jellife J. & Morgan, S. 2021. Slowing productivity reduces growth in global agricultural
output. In: Amber Waves. 28 December 2021. United States Department of Agriculture Economic
Research Service. www.ers.usda.gov/amber-waves/2021/december/slowing-productivity-reduc-
es-growth-in-global-agricultural-output

Fuglie, K., Gautam, M., Goyal, A. & Maloney, W.F. 2020. Harvesting prosperity: Technology and produc-
tivity growth in agriculture. Washington, DC, International Bank for Reconstruction and Development /
The World Bank. https://openknowledge.worldbank.org/bitstream/handle/10986/32350/97814648
13931.pdf?sequence=6&isAllowed=y

Funge-Smith, S.J. 2018. Review of the state of world fishery resources: Inland fisheries. FAO Fisheries and
Aquaculture Circular No. 942, Rev. 3. Rome, FAO. www.fao.org/3/ca0388en/CA0388EN.pdf

90 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


Galloway, J.N., Townsend, A.R., Erisman, J.W., Bekunda, M., Cai, Z., Freney, J.R., Martinelli, L.A.,
Seitzinger, S.P. & Sutton, M.A. 2008. Transformation of the nitrogen cycle: Recent trends, questions, and
potential solutions. Science, 320(5878): 889–892.

Gao, H., Yan, C., Liu, Q., Ding, W., Chen, B. & Li, Z. 2019. Effects of plastic mulching and plastic residue on
agricultural production: A meta-analysis. Science of the Total Environment, 651: 484–492.

Gardi, C., Angelini, M., Barceló, S., Comerma, J., Cruz Gaistardo, C., Encina Rojas, A., Jones, A. et al.
2014. Atlas de suelos de América Latina y el Caribe. Luxembourg.

Gardner, R.C., Barchiesi, S., Beltrame, C., Finlayson, C.M., Galewski, T., Harrison, I., Paganini, M.
et al. 2015. State of the world’s wetlands and their services to people: A compilation of recent anal-
yses. Ramsar Briefing Note No. 7. Gland, Ramsar Convention on Wetlands Secretariat. https://doi.
org/10.2139/ssrn.2589447

GDW (Global Dam Watch). 2022. Global Reservoir and Dam Database (GRanD). In: GDW. https://global-
damwatch.org/grand

Gleeson, T. & Richter, B. 2018. How much groundwater can we pump and protect environmental flows
through time? Presumptive standards for conjunctive management of aquifers and rivers. River Research
and Applications, 34(1): 83–92.

Grill, G., Lehner, B., Thieme, M., Geenen, B., Tickner, D., Antonelli, F., Babu, S. et al. 2019. Mapping
the world’s free-flowing rivers. Nature, 569(7755): 215–221.

Haacker, E.M.K., Kendall, A.D. & Hyndman, D.W. 2016. Water level declines in the high plains aquifer:
Predevelopment to resource senescence. Groundwater, 54(2): 231–242.

Harden, C.P. 2001. Soil erosion and sustainable mountain development: Experiments, observations and
recommendations from the Ecuadorian Andes. Mountain Research and Development, 21(1): 77–83.

Hatfield-Dodds, S., Hughes, N., Cameron, A., Miller, M. & Jackson, T. 2018. Analysis of 2018 drought:
26 October 2018. ABARES Insights 2. Canberra, Australian Bureau of Agricultural and Resource Economics
and Sciences. https://doi.org/10.25814/5bceb32574707

Hoogeveen, J., Faurès, J.M., Peiser, L., Burke, J. & Van De Giesen, N. 2015. GlobWat – A global water
balance model to assess water use in irrigated agriculture. Hydrology and Earth System Sciences, 19(9):
3829–3844.

IDA (International Desalination Association). 2021. The IDA – Desalination & reuse handbook
2021-2022. Oxford. www.globalwaterintel.com/products-and-services/market-research-re-
ports/ida-desalination-reuse-handbook

Immerzeel, W., Lutz, A. & Andrade, M. 2019. Importance and vulnerability of the world’s water towers.
Nature, 577: 364–369.

IPBES (Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services). 2018.


The IPBES assessment report on land degradation and restoration, L. Montanarella, R. Scholes & A.
Brainich, eds. Bonn. https://ipbes.net/assessment-reports/ld

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 91
IPCC (Intergovernmental Panel on Climate Change). 2019. 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. Geneva. www.ipcc.ch/srccl

IPCC. 2021. Climate change 2021: The physical science basis. Contribution of Working Group I to the Sixth
Assessment Report of the Intergovernmental Panel on Climate Change, V. Masson-Delmotte, P. Zhai, A.
Pirani, S. L. Connors, C. Péan, S. Berger, N. Caud et al., eds. Cambridge, UK, Cambridge University Press.
www.ipcc.ch/report/ar6/wg1/#TS

IPCC. 2022. Climate change 2022: Mitigation of climate change. Contribution of Working Group III to the
Sixth Assessment Report of the Intergovernmental Panel on Climate Change, P.R. Shukla, J. Skea, R. Slade,
A. Al Khourdajie, R. van Diemen, D. McCollum, M. Pathak, Set al., eds. Cambridge, UK and New York, USA.
Cambridge University Press.

IRF (International River Foundation). 2007. The Brisbane Declaration. Conference presentation at 10th
International River symposium and International Environmental Flows Conference. http://riverfoun-
dation.org.au/wp-content/uploads/2017/02/THE-BRISBANE-DECLARATION.pdf

IWA (International Water Association). 2016. Desalination – past present and future. In: IWA. London.
https://iwa-network.org/desalination-past-present-future

Kapetsky, J., Aguilar-Manjarrez, J. & Jenness, J., eds. 2013. A global assessment of offshore mariculture
potential from a spatial perspective. Rome, FAO. www.fao.org/3/i3100e/i3100e.pdf

Kondolf, G.M., Gao, Y., Annandale, G.W., Morris, G.L., Jiang, E., Zhang, J., Cao, Y. et al. 2014. Sustain-
able sediment management in reservoirs and regulated rivers: Experiences from five continents. Earth’s
Future, 2(5): 256–280.

Konikow, L.F. 2013. Groundwater depletion in the United States (1900–2008). United States Geological Survey
Scientific Investigations Report 2013–5079. https://pubs.usgs.gov/sir/2013/5079/SIR2013-5079.pdf

Lal, R., Smith, P., Jungkunst, H., Mitsch, W., Lehmann, J., Nair, P.K., McBratney, A. et al. 2018. The
carbon sequestration potential of terrestrial ecosystems. Journal of Soil and Water Conservation, 73(6):
145A–152A.

Lambers, H. 2003. Introduction: Dryland salinity: A key environmental issue in Southern Australia. Plant
and Soil, 257: 5–7.

Lausche, B. 2019. Integrated planning: Policy and law tools for biodiversity conservation and climate
change. Bonn, International Union for Conservation of Nature. https://doi.org/10.2305/IUCN.CH.2019.
EPLP.88.en

Leadley, P., Krug, C., Alkemade, R., Pereira, H., Sumaila, U.R., Walpole, M., Marques, A. et al. 2014.
Progress towards the AICHI biodiversity targets: An assessment of biodiversity trends, policy scenarios
and key actions. Technical Series No. 78. Montreal, Canada, Secretariat of the Convention on Biological
Diversity. www.cbd.int/doc/publications/cbd-ts-78-en.pdf

Leifeld, J. & Menichetti, L. 2018. The underappreciated potential of peatlands in global climate change
mitigation strategies. Nature Communications, 9(1).

92 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


Linstead, C. 2018. The contribution of improvements in irrigation efficiency to environmental flows. Fron-
tiers in Environmental Science, 6: 1–6.

Lowder, S.K., Sánchez, M. V & Bertini, R. 2019. Farms, family farms, farmland distribution and farm
labour: What do we know today? FAO Agricultural Development Economics. Rome, FAO. www.fao.org/3/
ca7036en/ca7036en.pdf

Lowder, S.K., Sánchez, M. V. & Bertini, R. 2021. Which farms feed the world and has farmland become
more concentrated? World Development, 142: 105455.

Lwanga, E.H., Vega, J.M., Quej, V.K., de los Angeles Chi, J., del Cid, L.S., Chi, C., Segura, G.E. et al. 2017.
Field evidence for transfer of plastic debris along a terrestrial food chain. Scientific Reports, 7(1).

Lytton, L., Ali, A., Garthwaite, B., Punthakey, J.F. & Basharat, S. 2021. Groundwater in Pakistan’s Indus
basin: Present and future prospects. Water Global Practice. International Bank for Reconstruction and
Development / The World Bank. https://openknowledge.worldbank.org/handle/10986/35065

Maggi, F., Tang, F.H.M., la Cecilia, D. & McBratney, A. 2019. PEST-CHEMGRIDS, global gridded maps of
the top 20 crop-specific pesticide application rates from 2015 to 2025. Scientific Data, 6(1): 1–20.

Margat, J. & van der Gun, J. 2013. Groundwater around the world: A geographic SYNOPSIS. Boca Raton,
USA, Taylor & Francis Group. www.un-igrac.org/sites/default/files/resources/files/Groundwater_
around_world.pdf

Mekonnen, M.M. & Hoekstra, A.Y. 2012. A global assessment of the water footprint of farm animal prod-
ucts. Ecosystems, 15(3): 401–415.

Mekonnen, M.M. & Hoekstra, A.Y. 2018. Global anthropogenic phosphorus loads to freshwater and asso-
ciated grey water footprints and water pollution levels: A high-resolution global study. Water Resources
Research, 54(1): 345–358.

Middleton, N. & Kang, U. 2017. Sand and dust storms: Impact mitigation. Sustainability, 9(6): 1053.

Molle, F. & Wester, P., eds. 2009. River basin trajectories: Societies, environments and development.
Comprehensive Assessment of Water Management in Agriculture Series. Wallingford, UK, CAB Inter-
national. www.iwmi.cgiar.org/Publications/CABI_Publications/CA_CABI_Series/River_Basin_
Trajectories/9781845935382.pdf

Moreno-de-las-Heras, M., Lindenberger, F., Latron, J., Lana-Renault, N., Llorens, P., Arnáez, J.,
Romero-Díaz, A. & Gallart, F. 2019. Hydro-geomorphological consequences of the abandonment of
agricultural terraces in the Mediterranean region: Key controlling factors and landscape stability patterns.
Geomorphology, 333: 73–91.

Mountain Partnership. 2015. Mountains. In: Definitions. Rome. www.fao.org/mountain-partner-


ship/about/definitions/en

MoWR RD (Ministry of Water Resources, River Development) & GR (Ganja Rejuvenation). 2017.
Report of 5th Census of Minor Irrigation Schemes, pp. 120–159.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 93
Mu, H., Otani, S., Shinoda, M., Yokoyama, Y., Onishi, K., Hosoda, T., Okamoto, M. & Kurozawa, Y.
2013. Long-term effects of livestock loss caused by dust storm on Mongolian inhabitants: A survey 1 year
after the dust storm. Yonago Acta Medica, 56(1): 39–42.

NASA (National Aeronautics and Space Administration). 2011. Heavy rains and dry lands don’t
mix: Reflections on the 2010 Pakistan flood. In: Earth Observatory. https://earthobservatory.nasa.
gov/features/PakistanFloods

OECD (Organisation for Economic Co-operation and Development) & FAO. 2021. OECD-FAO agricul-
tural outlook 2021-2030. Paris, OECD Publishing. https://doi.org/10.1787/19428846-en

Ortiz-Bobea, A., Ault, T.R., Carrillo, C.M., Chambers, R.G. & Lobell, D.B. 2021. Anthropogenic climate
change has slowed global agricultural productivity growth. Nature Climate Change, 11(4): 306–312.

Parish, R., Sirin, F., Charman, A., Joosten, D., Minayeva, H., Silvius, T. & Stringer, M. 2008. Assessment
on peatlands, biodiversity and climate change: Main report. Kuala Lumpur, Global Environment Centre
and Wageningen, Wetlands International. www.imcg.net/media/download_gallery/books/assess-
ment_peatland.pdf

Pekel, J.F., Cottam, A., Gorelick, N. & Belward, A.S. 2016. High-resolution mapping of global surface
water and its long-term changes. Nature, 540(7633): 418–422.

Pimentel, D. & Burgess, M. 2013. Soil erosion threatens food production. Agriculture, 3(3): 443–463.

Poff, N.L.R., Tharme, R.E. & Arthington, A.H. 2017. Evolution of environmental flows assessment science,
principles, and methodologies. In: Water for the environment: From policy and science to implementation
and management, pp. 203–236. https://doi.org/10.1016/B978-0-12-803907-6.00011-5

Qadir, M., Quillérou, E., Nangia, V., Murtaza, G., Singh, M., Thomas, R.J., Drechsel, P. & Noble,
A.D. 2014. Economics of salt-induced land degradation and restoration. Natural Resources Forum,
38(4): 282–295.

Qin, Y., Mueller, N.D., Siebert, S., Jackson, R.B., AghaKouchak, A., Zimmerman, J.B., Tong, D., Hong,
C. & Davis, S.J. 2019. Flexibility and intensity of global water use. Nature Sustainability, 2(6): 515–523.

Radatz, A., Lowery, B., Bland, W. & Hartemink, A. 2012. Groundwater recharge under compacted agricul-
tural soils, pine and prairie in central Wisconsin, USA. Agrociencia Uruguay Special issue, 16(3): 235–240.

Rodell, M., Famiglietti, J.S., Wiese, D.N., Reager, J.T., Beaudoing, H.K., Landerer, F.W. & Lo, M.H.
2018. Emerging trends in global freshwater availability. Nature, 557(7707): 651–659.

Rodriguez-Eugenio, N.R., McLaughlin, M.J. & Pennock, D.J. 2018. Soil pollution: A hidden reality. Rome,
FAO. www.fao.org/3/i9183en/I9183EN.pdf

Sadoff, C., Hall, J., Grey, D., Aerts, J.C.J., Ait-Kadi, M., Brown, C., Cox, A. et al. 2015. Securing water,
sustaining growth. Report of the GWP/OECD Task Force on Water Security and Sustainable Growth. Oxford,
UK, University of Oxford. www.gwp.org/globalassets/global/about-gwp/publications/the-glob-
al-dialogue/securing-water-sustaining-growth.pdf

94 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


Schneider, C., Flörke, M., De Stefano, L. & Petersen-Perlman, J.D. 2017. Hydrological threats to riparian
wetlands of international importance – A global quantitative and qualitative analysis. Hydrology and
Earth System Sciences, 21(6): 2799–2815.

Searchinger, T., Waite, R., Hanson, C., Ranganathan, J. & Dumas, P. 2019. Creating a sustainable food
future: A menu of solutions to feed nearly 10 billion people by 2050. Washington, DC, World Resources
Institute. https://research.wri.org/sites/default/files/2019-07/WRR_Food_Full_Report_0.pdf

SEEA (System of Environmental Economic Accounting). 2022. What is the SEEA? In: SEEA. New York.
Cited 9 February 2022. https://seea.un.org

Shah, T. 2009. Taming the anarchy: Groundwater governance in South Asia. Washington, DC, Resources for
the Future.

Sheehy, D., Miller, D. & Johnson, D. 2006. Transformation of traditional pastoral livestock systems on the
Tibetan steppe. Sécheresse, 17(1–2): 142–151.

Siebert, S. & Döll, P. 2007. Irrigation water use – A global perspective. In: J.L. Lozán, H. Graßl, P. Hupfer,
L. Menzel & C. Schönwiese, eds. Global change: Enough water for all?, pp. 104–107. University of
Hamburg/GEO.

Siebert, S., Burke, J., Faures, J.M., Frenken, K., Hoogeveen, J., Döll, P. & Portmann, F.T. 2010. Ground-
water use for irrigation – A global inventory. Hydrology and Earth System Sciences, 14(10): 1863–1880.

Soane, B. & Van Ouwerkerk, C. 2013. Soil compaction in crop production. Amsterdam, Elsevier.

Sood, A., Smakhtin, V., Eriyagama, N., Villholth, K., Liyanage, N., Wada, Y., Ebrahim, G. & Dickens,
C. 2017. Global environmental flow information for the Sustainable Development Goals. Colombo, IWMI.

Stevens, C.J., David, T.I. & Storkey, J. 2018. Atmospheric nitrogen deposition in terrestrial ecosystems:
Its impact on plant communities and consequences across trophic levels. Functional Ecology, 32(7):
1757–1769.

Sumarga, E., Hein, L., Hooijer, A. & Vernimmen, R. 2016. Hydrological and economic effects of oil palm
cultivation in Indonesian peatlands. Ecology and Society, 21(2): 52.

Sutton, M.A., Billen, G., Bleeker, A., Erisman, J.W., Grennfelt, P., van Grinsven, H., Grizzetti, B.,
Howard, C.M. & Leip, A. 2011. Technical summary. The European Nitrogen Assessment, pp. xxxv–lii.
Cambridge, UK, Cambridge University Press. https://doi.org/10.1017/CBO9780511976988.003

Tan, Z.X., Lal, R. & Wiebe, K.D. 2005. Global soil nutrient depletion and yield reduction. Journal of Sustain-
able Agriculture, 26(1): 123–146.

Tandon, H. 1992. Assessment of soil nutrient depletion. FADINAP seminar: Fertilization and the environ-
ment. Thailand, Chiang Mai.

Tang, F.H.M., Lenzen, M., McBratney, A. & Maggi, F. 2021. Risk of pesticide pollution at the global scale.
Nature Geoscience, 14(4): 206–210.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 95
Tellman, B., Sullivan, J.A., Kuhn, C., Kettner, A.J., Doyle, C.S., Brakenridge, G.R., Erickson, T.A. &
Slayback, D.A. 2021. Satellite imaging reveals increased proportion of population exposed to floods.
Nature, 596: 80–86.

Tennigkeit, T. & Wilkes, A. 2008. An assessment of the potential for carbon finance in range-
lands. ICRAF Working Paper No 68. Beijing. http://apps.worldagroforestry.org/downloads/Publica-
tions/PDFS/WP15892.pdf

TerraBrasilis. 2022. Amazon forest cover. In: TerraBrasilis. Cited 9 February 2022. http://terrabrasilis.dpi.
inpe.br/app/dashboard/deforestation/biomes/legal_amazon/rates

Tessler, Z.D., Vörösmarty, C.J., Overeem, I. & Syvitski, J.P.M. 2018. A model of water and sediment
balance as determinants of relative sea level rise in contemporary and future deltas. Geomorphology,
305: 209–220.

Thebo, A.L., Drechsel, P., Lambin, E.F. & Nelson, K.L. 2017. A global, spatially-explicit assessment of irri-
gated croplands influenced by urban wastewater flows. Environmental Research Letters, 12(7): 074008.

Tickner, D., Opperman, J.J., Abell, R., Acreman, M., Arthington, A.H., Bunn, S.E., Cooke, S.J. et al.
2020a. Bending the curve of global freshwater biodiversity loss: An emergency recovery plan. BioScience,
70(4): 330–342.

Tickner, D., Kaushal, N., Speed, R. & Tharme, R. 2020b. Editorial: Implementing environmental flows:
Lessons for policy and practice. Frontiers in Environmental Science, 8(August): 1–3.

Tubiello, F., Salvatore, M., Cóndor Golec, R., Ferrara, A., Rossi, S., Biancalani, R., Federici, S., Jacobs,
H. & Flammini, A. 2014. Agriculture, forestry and other land-use emissions by sources and removals by
sinks 1990-2011 analysis. FAO Statistics Division Working Paper Series ESS/14-02. Rome, FAO. www.fao.
org/3/i3671e/i3671e.pdf

Turyahabwe, N., Kakuru, W., Tweheyo, M. & Tumusiime, D.M. 2013. Contribution of wetland resources
to household food security in Uganda. Agriculture & Food Security, 2: 5.

Turetsky, M.R. 2019. Permafrost collapse is accelerating carbon release. Nature, 569: 32–34.

Turetsky, M.R., Benscoter, B., Page, S., Rein, G., van der Werf, G.R. & Watts, A. 2015. Global vulnera-
bility of peatlands to fire and carbon loss. Nature Geoscience, 8(1): 11–14.

UNCCD (United Nations Convention to Combat Desertification). 2017. Global land outlook. Bonn.
https://knowledge.unccd.int/glo/GLO_first_edition

UNCCD. 2022. Sand and dust storms. In: UNCCD. Bonn. www.unccd.int/actions/sand-and-dust-storms

UNDRR (United Nations Office for Disaster Risk Reduction). 2021. GAR special report on drought 2021.
Geneva. www.undrr.org/publication/gar-special-report-drought-2021

UNEMG (United Nations Environment Management Group). 2022. UN coalition to combat sand and dust
storms. In: UNEMG. Geneva. https://unemg.org/our-work/emerging-issues/sand-and-dust-storms

96 1. STATUS OF AND TRENDS IN LAND, SOIL AND WATER RESOURCES


UNEP (United Nations Environment Programme). 2019. Global environment outlook: GEO-6: Healthy
planet, healthy people. Nairobi. www.unep.org/resources/global-environment-outlook-6

USDA (United States Department of Agriculture) National Agricultural Statistics Service. 2019. 2017
census of agriculture. 2018 irrigation and water management survey. Volume 3, Special Studies, Part 1.
www.nass.usda.gov/Publications/AgCensus/2017/Online_Resources/Farm_and_Ranch_Irriga-
tion_Survey/fris.pdf

USGS (United States Geological Survey). 2018. Estimated use of water in the United States in 2015. Circular
1441, Water Availability and Use Science Program. Virginia. https://pubs.usgs.gov/circ/1441/circ1441.pdf

Vanham, D., Hoekstra, A.Y., Wada, Y., Bouraoui, F., de Roo, A., Mekonnen, M.M., van de Bund, W.J.
et al. 2018. Physical water scarcity metrics for monitoring progress towards SDG target 6.4: An evaluation
of indicator 6.4.2 “Level of water stress”. Science of the Total Environment, 613–614 (February): 218–232.

Verheijen, F.G.A., Jones, R.J.A., Rickson, R.J. & Smith, C.J. 2009. Tolerable versus actual soil erosion rates
in Europe. Earth-Science Reviews, 94(1–4): 23–38.

Vishwakarma, B.D., Bates, P., Sneeuw, N., Westaway, R.M. & Bamber, J.L. 2021. Re-assessing global
water storage trends from GRACE time series. Environmental Research Letters, 16(3).

von Braum, J., Afsana, K., Fresco, L.O. & Hassan, M., eds. 2021. Science and Innovation for Food
Systems Transformation and Summit Actions. Papers by the Scientific Group and its partners in
support of the UN Food Systems Summit. Science Group of the UNFSS. https://sc-fss2021.org/wp-con-
tent/uploads/2021/07/Scientific-Group-Strategic-Paper-Science-for-Transformation-of-Food-Sys-
tems_August-2.pdf

Wada, Y., Van Beek, L.P.H. & Bierkens, M.F.P. 2012. Non-sustainable groundwater sustaining irrigation:
A global assessment. Water Resources Research, 48(6).

Welcomme, R.L., Baird, I.G., Dudgeon, D., Halls, A., Lamberts, D. & Mustafa, M.G. 2016. Fisheries of the
rivers of Southeast Asia. Freshwater Fisheries Ecology (December 2017): 363–376.

WMO (World Meteorological Organization). 2022a. Sand and dust storms. In: WMO. Geneva. https://
public.wmo.int/en/our-mandate/focus-areas/environment/SDS

WMO. 2022b. Drought. In: WMO. Geneva. https://public.wmo.int/en/our-mandate/water/drought

World Bank Group. 2018. Good practice handbook: Environmental flows for hydropower projects – Guid-
ance for the private sector in emerging markets. Washington, DC. https://doi.org/10.1596/29541

WRI (World Resources Institute). 2022. Aqueduct: Using cutting-edge data to identify and evaluate water
risks around the world. In: WRI. Washington, DC. www.wri.org/aqueduct

Yue, X., Wang, H., Wang, Z. & Fan, K. 2009. Simulation of dust aerosol radiative feedback using the
Global Transport Model of Dust: 1. Dust cycle and validation. Journal of Geophysical Research Atmo-
spheres, 114(D10).

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 97
2
Socioeconomic
settings

Key messages
Demographic growth, economic growth and urbanization are changing patterns in food demand. These
changes are placing unprecedented pressures on ecosystems and limited renewable land, soil and water
resources. Higher incomes and urban lifestyles are changing food demand towards more resource-intensive
consumption of animal proteins, fruits and vegetables. The world’s population is expected to grow from
7.7 billion in 2019 to 9.7 billion in 2050 (26 percent). The fastest growth will be in the poorest countries,
such as those in sub-Saharan Africa, where the population is expected to double by 2050.

Globally, 80 percent of the extreme poor live in rural areas, mostly in the developing world, and 64
percent of the extreme poor are employed in agriculture. It is key to increasing food security, reducing
poverty and achieving multiple SDGs, but is highly exposed to current and future climate risks.

Uncontrolled urbanization threatens sustainable resources management. By 2050, two out


of three people will be living in towns and cities, with most urban growth occurring in the less-
developed regions of Africa and Asia. Urban dwellers consume 80 percent of all food produced globally.
“Industrial” foods dominate and bring alarming health consequences – the triple burden of malnutrition
comprising undernutrition, overweight and obesity – and micronutrient deficiencies.

Increasing population reduces the natural resources available per capita. In sub-Saharan Africa, which
has the fastest demographic growth, water availability per capita declined by 40 percent over the past two
decades, and agricultural land declined from 0.80 ha/capita to 0.64 ha/capita between 2000 and 2017.

Increasing concentration of farmland among larger farms as economies grow brings increasing inequality
in agriculture. All types of farms and the entire value chain, from producers to consumers, need to consider
ways of transforming food systems to address SDGs.
©FAO/Giulio Napolitano

Ensuring equitable access to land and water resources is key for promoting inclusive rural transformation.
The lack of adequate access and user rights and increasing disparities in capacities to take advantage of
natural capital are underlying drivers of overuse of resources to meet short-term needs.

Social, agricultural and environmental policies can be mutually reinforcing in order to reconcile
competition over land and water. Opportunities to change or modify national policies exist, supported by
United Nations Decades that focus on ecosystems, water, SDGs and family farming, and which encourage
agroecological approaches and harmonized decisions by parties to the multilateral environmental
conventions.
2.1 Introduction

© Pep Bonet/NOOR for FAO


Chapter 1 established the challenges of
water scarcity, land and soil degradation,
and the uncertainties from climate change
that reinforce the need to adapt and inte-
grate sectoral policies to ensure wise use of
limited resources for people and the envi- ing poverty and food insecurity and achieving
ronment. The management of land and water SDGs. However, agriculture is highly exposed
resources for agriculture, forestry and other to land degradation and water scarcity, in
uses is driven as much by socioeconomics and addition to climate risk. Responding to these
governance as by biophysical and technical risks by improving land and water access
factors. Trade-offs between competing social and management is an essential part of
and economic demands and desired envi- enhancing the resilience of rural livelihoods
ronmental outcomes become inevitable in and sustaining ecosystem services. Access to
setting a path towards sustainable land and land, water and associated biological diver-
water management. sity is essential for most rural households,
but it is vital for the rural poor.
Land and water are crucial assets for
livelihoods and well-being among rural This chapter describes the main socioeco-
communities and farming households. nomic and governance drivers that affect
The multidimensional role of these assets land and water availability and use. It also
and their sustainable management require discusses the socioeconomic trends that will
different kinds of institutions. Rural influence future land and water strategies
communities are the custodians of natural and planning, and how they are expected to
resources. Their local decision-making is affect the state of land and water resources
affected by decision-making at national, and their associated farming systems.
regional and international levels. Such
decision-making influences the governance
of natural resources in rural settings. Local 2.2 Socioeconomic
transitions –
decision-making is also affected by markets
and linkages of rural areas to urban centres,
which creates employment but additional
pressures on limited resources. The scale of
implications for
farms, whether smallholdings less than 2 ha land and water
management
in extent or large-scale commercial concerns,
determines the type and effectiveness of land
and water governance needed.
Population growth, economic growth and
In 2013, household surveys indicated that
urbanization are the principal socioeco-
80 percent of the extreme poor living on less
nomic variables driving demand for land
than USD 1.90 per day is in rural areas (World
and water resources. In addition, migration
Bank, 2016; Castañeda et al., 2018). Most live
of human populations, adoption of new
in low-income food-deficit countries, and
forms of communication and exchange of
their livelihoods are highly dependent on
information are transforming the charac-
agriculture. The performance of all types of
ter of rural economies and the mobility of
agriculture in such countries is key to reduc-
the rural workforce. At the same time there

100 2. SOCIOECONOMIC SETTINGS


is continuous adaptation to climate change and 811 million people faced hunger in 2020,
in all regional and agricultural subsectors. about 118 million more than in 2019. More
Pressures placed on limited land, soil and than half were in Asia (418 million) and more
renewable water resources are there- than a third were in Africa (282 million).
fore unprecedented. Higher incomes and Projections suggest that by 2030 there will
urban lifestyles are also steering food
be 660 million people undernourished, in
demand towards more resource-intensive
part due to the lasting effects of COVID-19 on
consumption.
global food security. The COVID-19 pandemic
has therefore exposed the vulnerabilities
2.2.1 Population growth within poor populations and the global food

Demographic growth drives the demand system as a whole. The World Bank estimates

for food and agricultural products, putting the COVID-19 pandemic pushed an additional

unprecedented pressure on renewable but 119 million to 124 million people into extreme

limited water and land resources. Projec- poverty in 2020 (Lakner et al., 2021).

tions suggest the world’s population will


Climate variability and extreme weather
grow from 7.7 billion in 2019 to 8.5 billion
events, such as droughts, floods and extremes
by 2030 (10 percent increase), to 9.7 billion
of heat, together with territorial conflicts and
by 2050 (26 percent) and to 10.9 billion by
economic downturns are crucial drivers of
2100 (42 percent) (United Nations, 2019)
food insecurity and malnutrition (FAO et al.,
(Figure 2.1). The fastest growth is in the poor-
2021) and have highlighted the close rela-
est regions, including sub-Saharan Africa
tionship with land and water. The number of
where the population will double by 2050,
low- and middle-income countries exposed
thus creating immense challenges to achiev-
to climate extremes increased from 76 coun-
ing SDGs, in particular, SDG 1 (no poverty),
tries in 2000–2004 to 98 in 2015–2020 (FAO
SDG 2 (zero hunger), SDG 6 (clean water and
et al., 2021).
sanitation) and SDG 15 (life on land).

Climate extremes can also intensify other


2.2.2 Food insecurity, drivers of food insecurity, such as conflicts,
malnutrition and poverty loss of livelihoods, poverty and increased
inequality. According to the World Bank,
Agricultural production more than trebled
extreme poverty increased in 2020 for the
between 1960 and 2015 (HLPE, 2017). Diets
first time in over 20 years (World Bank,
are changing, and numerous local, national
2020). The poor are predominantly rural,
and multinational food-related enterprises
young and undereducated. Half of those in
have emerged, providing livelihoods for
poverty are children and women. In 2018,
millions. However, global hunger is rising
four out of five people below the inter-
again, and climate change and the effects of
national poverty line lived in rural areas,
the COVID-19 pandemic are making matters
although the rural population accounted for
worse. According to The state of food secu-
only 48 percent of the global population.
rity and nutrition in the world 2021 report
Indeed, poverty became more rural between
(FAO et al., 2021), the world is not on track
2015 and 2018, and increased by more than
to eradicate hunger, food insecurity and
2 percent (World Bank, 2020). Climate change
all forms of malnutrition by 2030. Efforts
is expected to drive 68 million to 132 million
will need redoubling, given the challenges
into poverty by 2030.
brought by COVID-19. Between 720 million

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 101
FIGURE 2.1 Population by SUSTAINABLE DEVELOMENT GOAL region: estimates, 1950–2020 and
medium-variant projection with 95 percent prediction intervals, 2020–2100

5
Projection

Sub-Saharan Africa

3
Total population (billions)

Central and Southern Asia

2
Eastern and South-Eastern Asia

Europe and North America


1
Northern Africa and Western Asia

Latin America and Caribbean

Australia/New Zealand
0 Oceania (excluding
Australia and New Zealand)
1950 2000 2050 2100

Year

Source: United Nations. 2019. World population prospects 2019: Highlights. ST/ESA/SER.A/423. New York, Department of Economic and Social Affairs,
Population Division. https://population.un.org/wpp/Publications/Files/WPP2019_Highlights.pdf

102 2. SOCIOECONOMIC SETTINGS


Dry lands often include a disproportionate
share of poor people. This is especially the
case in Africa, where dry lands are home to

© FAO/Richard Trenchard
50 percent of the population and account
for 43 percent of the region’s land surface,
of which 75 percent is used for agriculture.
In 2010, the World Bank reported that about
171 million people living in African dry lands
depended on agriculture, including 26 million
pastoralists and 105 million agropastoral- Suggested interventions include policy
ists. They were exposed to weather shocks, support and investment to improve natural
especially drought, due mainly to the poor resources management practices and tech-
performance of the agriculture sector. More- nologies for pastoral and agropastoral live-
over, dry lands could expand by 20 percent stock keepers and crop producers in rainfed
under some climate scenarios, with popu- and irrigated systems. Other options include
lation growth bringing even more people reducing trade barriers to make food more
into a challenging environment. The negative available and affordable, and strengthening
trends identified in human-induced land integrated landscape management to reverse
degradation raise concerns about the adverse degradation trends and enhance ecosystem
impacts on land productivity, reducing farm health and function (Cervigni et al., 2016).
incomes and increasing vulnerability and Chapter 5 develops these response options in
stress (Cervigni et al., 2016). more detail.

Political economy factors affecting resilience, The FAO framework on extreme rural poverty
especially the uneven distribution of wealth recognizes that conservation and restoring
and power, can marginalize many dryland natural resources should directly benefit the
groups. This can skew the distribution of rural extreme poor, particularly those living
social services for human health and educa- in remote marginal areas. This is linked to
tion. Targeted adaptive interventions could promoting responsible governance of the
help reduce the impact of droughts by about tenure of resources. Recognizing the legiti-
half, keeping 5 million people each year out mate tenure rights of people to use, manage
of danger in some of Africa’s poorest zones and control land, water, biodiversity, forests
(Cervigni et al., 2016). and fisheries is fundamental to helping the
rural extreme poor adapt to climate change
In June 2021, the World Bank reported that (FAO, 2019).
COVID-19 plunged sub-Saharan Africa into
its first recession in over 25 years (with activ- As identified in the Action Tracks for the
ity contracting by nearly 5 percent on a per 2021 United Nations Food Systems Summit,
capita basis, exacerbating high public debt). “advancing equitable livelihoods and value
This disproportionately affects vulnerable distribution” and “building resilience to
groups, such as the poor, informal sector vulnerabilities, shocks and stresses” are
workers, women and youth, thus reducing essential and interlinked components of
opportunities and access to social safety nets. shifting to sustainable production and
Up to 40 million people could be pushed into consumption patterns at scale, and ensuring
extreme poverty, erasing at least five years of access to safe and nutritious food for all
progress. (Ffigure 2.2).

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 103
FIGURE 2.2 Action tracks in a food system

1. Ensuring access to safe and nutritious food for all


(Enabling all people to be well nourished and healthy)

4. Advancing equitable
livelihoods and value distribution
(Raising incomes, distributing
risk, expanding inclusion, creating jobs)

3. Boosting nature-positive
2. Shifting to sustainable production at sufficient scale
consumption patterns (Acting on climate change, reducing
SEEKING SYNERGIES
(Promoting and creating demands for emissions, regenerating/protecting
healthy and sustainable diets, ecosystems, reducing food loss/energy
reducing waste) usage, without undermining health or
nutritious diets)

5. Building resilience to
vulnerabilities, shocks, stresses
(Ensuring continued functionality
of healthy and sustainable
food systems)

Source: von Braun, J., Afsana, K., Fresco, L., Hassan, M. & Torero, M. 2021. Food systems – definition, concept and application for the UN Food
Systems Summit. A paper from the Scientific Group of the UN Food Systems Summit. www.un.org/sites/un2.un.org/files/scgroup_food_
systems_paper_march-5-2021.pdf

2.2.3 Urbanization Higher incomes and urban lifestyles are


changing food demand towards more
and changing resource-intensive animal proteins, fruits
consumption patterns and vegetables. Animal production relies
mainly on rainfed systems (Heinke et al.,
It is estimated that 55 percent of the world’s
2020), while intensive and concentrated
population now lives in urban areas (United
industrial production units represent the
Nations, 2018a). By 2050, two out of three
largest share of resource consumption (HLPE,
people are expected to live in towns and cities,
2016). The rising consumption of processed
with most growth in the less-developed
foods, particularly meat products, is associ-
regions of Africa and Asia. Urban dwellers
ated with urbanization. Meat consumption is
consume 80 percent of all food produced.
increasing annually by 1.4 percent globally,
and grew by 58 percent over 20 years to 2018,
Food demand is expected to increase by
reaching 360 million tonnes. Per capita meat
50 percent between 2013 and 2050 (FAO,
consumption has increased as consumer
2017). The main drivers are population
preferences change, particularly for poultry.
growth, urbanization and rising incomes.

104 2. SOCIOECONOMIC SETTINGS


Consumption is growing fast in Africa and
the Near East, but China is the world’s largest
meat consumer. Since 2009, China has turned
from a net exporter of corn to a net importer

© FAO/Giuseppe Bizzarri
to meet the demand for feed grain.

Managing the rural–urban transition is a


significant challenge for large cities and the
growing network of small- and medium-sized
emerging towns competing for limited land
and water resources. It requires paying Urban-based farming accounts for a limited
attention to the types of farms and the types land area worldwide, but it involves a
of food supply chains, and providing support significant share of the agricultural popu-
through rural–urban public policies that are lation: 3–7 percent. Urban farming remains
inclusive in terms of avoiding concentration marginal in terms of overall production, but
and large-scale industrial value chains. the COVID-19 crisis has revealed the impor-
tance of local systems, close to consumers
Urbanization concentrates food demands and based on trust regarding food quality and
and trends; it requires renewed attention safety. The linkages with organic agriculture
to land planning to manage the demand and water waste recycling are assets expected
for green spaces and quality water (greener to contribute to sustainable development in
cities) and to bring about a transition to the coming decades.
more sustainable patterns of food produc-
tion and distribution. Addressing spatial and Cities and municipal governance are becom-
social inequalities will require differentiated ing increasingly important agents of change
policies focusing on strengthening regional for sustainable rural development, includ-
and local food systems and changing waste ing policies and actions to improve natural
management towards a circular economy resources sustainability and access to healthy
(Chapter 5). diets (Neufield, Hendriks and Hugas, 2021).
The FAO Green Cities Initiative focuses on
In cities, processed foods are part of a broad improving the urban environment, strength-
dietary transition and bring alarming health ening rural–urban linkages, increasing the
consequences: the triple burden of malnutri- availability of green spaces through urban
tion, comprising undernutrition, overweight and peri-urban forestry and improving
and obesity, and micronutrient deficiencies. access to healthy diets.
The spread of small- or medium-sized food
distribution stores supplying processed In Africa, urbanization is increas-
“fast” food with low nutritive value and its ing the population density in small- and
spin-off in reduced capacity and interest to medium-sized towns, thus creating conur-
cook at home are a new threat to public health. bations in many subregions, such as the
These trends compete with locally produced, Mediterranean coast, from the Gulf of
good-quality and diversified food that Guinea coast to the Sahel region, in the East
provides rural livelihoods, including income African Highlands and in the Great Lakes
for many women engaged in food crafting. In region (Chatel, Imbernon and Moriconi-Eb-
some countries, small- and medium-sized rard, 2016). This brings opportunities for
stores are increasingly spreading industrially rural family farms by developing local food
processed food in rural areas, thus affecting markets for millions of consumers and devel-
local markets, diets and health. oping alternative and complementary income

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 105
sources as local economies diversify (Mainet Migration is closely related to the challenges
and Edouard, 2017; Agergaard et al., 2019). affecting rural communities including food
Population growth in Africa calls for renewed insecurity, limited income-generating activ-
land-use policies to enable sustainable agri- ities, and lack of employment and decent
cultural development to boost productivity working conditions. Rural–urban inequality
and diversification and innovative marketing has also pushed people to migrate to cities to
channels to reach consumers and encourage find better jobs and living conditions and to
sustainable diets. In restoring 100 million ha access education, health services and social
of degraded lands, the Great Green Wall for protection. Migrants have diverse socioeco-
the Sahel and Sahara Initiative (see Chap- nomic profiles and expectations according to
ter 4) includes efforts to promote urban and their economic, political and cultural circum-
peri-urban agriculture, to facilitate market stances, which change over time. Migration
development for commodities at national may be temporary or permanent. Family
and regional levels and create 10 million mobility, including the “walking assets” of
“green” jobs in rural areas (Cunningham and herders and pastoralists, is part of a liveli-
Abasse, 2005). hood and way of living.

2.2.4 Migration There is growing evidence that environmental


change and disasters aggravated by climate
Seasonal migration for pastoral communities change amplify international and national
and farm labour has always been part of human movements and displacement (FAO,
the evolution and development of agrarian 2018a). Climate- and water-related disas-
societies. However the acceleration of migra- ters were responsible for displacing over
tion of internally displaced communities 23 million people in 2016. In Somalia, drought
and transboundary migration has increased is causing increased malnutrition, food inse-
significantly over the past two decades. It has curity and competition for limited resources,
now become a global issue of the twenty-first especially among farmers and pastoralists.
century and is high on the political agendas Some 25 thousand people were displaced
of many countries. Estimates indicate that in because of drought in 2018 (IOM, 2019). The
2020, the number of international migrants World Bank estimated water shortages were
increased to 281 million (3.5 percent of the linked to a 10 percent increase in total migra-
global population) from 173 million in 2000 tion within countries between 1970 and 2000,
(United Nations, 2020). Between 2000 and and adversely affected the skills and educa-
2019, the number of refugees increased tional levels of migrants (World Bank, 2020).
from 14 million to nearly 26 million, and Those who leave countries because of drought
the number of internally displaced persons usually have lower academic grades and skills
increased from 21 million to just over than other migrant workers, implying lower
41 million (IOM, 2019). wages and poorer access to essential services
at their destination.

Migrants affect land and water use in rural


©FAO/IFAD/WFP/Michael Tewelde

areas in their homeland, during transit and at


their destination. Migration among youth and
young men can lead to poor land manage-
ment as responsibilities are left to female
heads of family who face depleted family
and community labour. Communities that
generate a flow of migrants limit the develop-

106 2. SOCIOECONOMIC SETTINGS


© FAO/Lekha Edirisinghe
2.2.5 Innovation through
digital technologies
Digital technologies are transforming the
way land and water resources are managed,
making agriculture more resilient, innovative
and efficient. They are changing the way
ment or maintenance of water harvesting and farmers work in the field and in the market
improved land management practices. The for agricultural produce. The advent of
use of child labour may also increase. Water
affordable mobile phones and the expansion
insecurity is one of the main vulnerabili-
of mobile networks have spurred adoption
ties suffered by forcibly displaced people and
of just-in-time application of field inputs,
their host communities. However, migration
including machinery, and the operation
also has positive impacts. Sending remit-
markets and crop/livestock processing
tances home to rural communities enables
families to secure their basic social needs and beyond the farm gate. The volume and

invest in better equipment to reduce drudg- quality of customizable farming information


ery, improve productivity and create better available to farmers now allows near
links to markets. Remittances increased from real-time adjustment of farming practices
USD 126 billion in 2000 to USD 689 billion from large-scale commercial operations to
in 2018 (IOM, 2019), with about 40 percent those operating small hectare-level plots for
going to rural areas (IFAD, 2017). high-value crops. Mobile internet and Global
Positioning System services for tractors
Migration policies and strategies vary
and harvesters are now standard, giving
according to their context. Some policies
much more precision to land treatment,
try to reduce the risk of water insecurity
including automated laser levelling, fertilizer
in the countries of origin by investing in
application and seed drilling.
water storage infrastructure, irrigation, early
warning systems and agricultural adaptation.
The Internet of Things, data analytics, cloud
This approach can alleviate economic imper-
computing and blockchain systems offer new
atives to make migration a choice. Other
policies focus on promoting the integration capabilities to analyse, manage in real time,

of migrants at their destination, promoting predict and minimize risks. Freely available
growth in cities and incentivizing mobility. satellite imagery can now reduce the cost of
monitoring agricultural activities, including
There have been substantial developments in land and water management. For pastoral-
global migration governance in recent years, ists, mobile phones and information systems
but there is not yet an overarching framework can help to better manage natural resources,
that provides policy guidance on migration
such as access to pastures and related water
and environmental stressors including
sources, to better manage animal move-
climate change. Policy developments include
ment/transhumance routes, and to provide
the United Nations Framework Convention
information on market prices to get a better
on Climate Change (UNFCCC) climate
value for products (Lawali and Idrissa, 2015).
negotiations and the Global Compact for Safe,
Orderly and Regular Migration that addresses While facilitating trading, information
with varying emphasis the mobility aspects technologies can also increase security by
of environmental degradation and climate connecting people during emergencies and
change (IOM, 2019). build social capital among farmer groups.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 107
and agriculture absorbs a disproportionate
63 percent share of their impacts, compared
to other sectors such as tourism, commerce
© FAO/GMB Akash

and industry (FAO, 2021).

Important impacts of climate change are


increasing rainfall variability and water scar-
city. Rainfall variability is responsible for an
However, the adoption of digital technolo- annual net loss of food production equivalent
gies and communications faces challenges, to feeding 81 million people every day. Wors-
including obsolete infrastructure and lack of ening droughts are projected to affect about
investment in operation and maintenance. 700 million people by the end of this century,
Some 4.5 billion people had internet access and to disproportionately affect developing
in 2020 (ITU, 2022), although there are still countries. In addition to affecting agricultural
significant differences in connectivity, access yields, rainfall variability is also responsible
and broadband speed between the wealthi- for cropland expansion, thus placing much
est countries and the rest of the world, and pressure on forested areas (Damania et al.,
between urban and rural areas. The “digital 2017). However, not all the changes are detri-
divide” or the technological gap between mental. In the Sahel, the wetter trend leading
countries and between rural and urban areas to increased intensity of rain events in recent
can widen the development gap within and years (Fiondella, 2013) has enhanced natural
among countries. Some 40 percent of the regeneration of critical tree cover to protect
global population remains unconnected, and and restore cultivated soils. This process was
the latest data indicate women are more likely subsequently scaled out to reverse degrada-
to be unconnected than men. Addressing this tion and desertification trends, and to adapt
unbalanced distribution and lack of access to climate change.
is critically important in achieving food
systems that are more sustainable, resilient The projected impacts of climate change
and inclusive. are diverse, among and within countries.
Policymaking often struggles to integrate
2.2.6 Impacts of farm diversity, especially for family farms

climate change in (Kansiime, van Asten and Sneyers, 2018).


Climate action and the SDGs of the 2030
rural communities Agenda for Sustainable Development (2030
Agenda) are closely connected. It will be diffi-
Climate change is an important global trend
cult, if not impossible, to eradicate poverty,
that has key implications for land and water
end hunger or ensure access to water with-
resources. The effects of climate change on
out building resilience and mitigating the
land and water resources (Chapter 1) exac-
impacts of climate change in smallholder
erbate the pressures arising from population
agricultural production systems (Poláková
growth, urbanization and dietary changes.
et al., 2013) and even beyond sector policies
Climate change will increase the risks to
(Alpha and Fouilleux, 2018) through more
livelihoods of rural communities, and to
integrated strategies.
food security and nutrition among rural and
urban populations. The rural poor are the
Policy innovation for climate change and,
most vulnerable and are likely to be dispro-
in particular, dealing with land and water
portionately affected. A recent FAO report
management in agriculture has so far been
suggests disasters happen three times more
limited. Many agriculture and land and water
often today than in the 1970s and 1980s,
planning documents do not yet consider the

108 2. SOCIOECONOMIC SETTINGS


anticipated effects of climate change; even ing systems (Robinson et al., 2014) to assess
if they do, there is a lack of commitment, and understand resource consumption and
capacity, financing and tools for effective its impact on climate change.
implementation. In many cases, conventional
policy instruments have been relabelled as Climate change adaptation needs to include
instruments for climate change adaptation, the “farming system dimension” at the
but need additional policy support to scale up territorial level. This also requires a “food
adoption and enhance resilience. system” perspective to consider the impacts
of climate change, including the value chain
From the mid-2000s onwards, under down to the consumers. Understanding
UNFCCC, countries began formulating the timeline for interventions is crucial
National Adaptation Plans and nationally to avoid the inherent mismatch between
determined contributions (NDCs) that include short-term and long-term benefits. For a
efforts in the agriculture sector to identify farmer, the financial and economic horizon
medium- and long-term adaptation needs is generally short (seasonal) to medium
and to develop strategies. By 2020, 125 of term (one to five years), or longer (up to
the 154 developing countries had undertaken ten years) if tree crops or large investments
activities related to formulating and imple- are concerned, allowing time to adapt to
menting National Adaptation Plans, of which market fluctuations and climate variability
55 were supported by the Green Climate Fund to some extent. However, climate action
(UNFCCC, 2020). FAO analysis of the intended requires a larger-scale landscape and
NDCs in 2016 noted agricultural sectors are ecosystems perspective and a longer-term
still not developed and prioritized enough collective adaptation process to cope with
in NDCs. Nevertheless, agricultural sectors forecasted changes over two or more decades
and land and water management will need and to achieve results in terms of improving
to and are expected to play a significant food security and reducing malnutrition
role in national responses to climate change. and poverty. Farmers therefore need greater
This is particularly important for develop- tenure security and intergenerational
ing countries where agriculture and natural planning at the territorial level for sustainable
resources management are critical for rural development under climate change.
livelihoods, food security and nutrition (FAO,
2016a). Responses are needed at a national
level, strategies need to be implemented at
provincial and municipal levels, and plans
and actions need to be implemented by farm-
ers and rural communities. This requires
intersectoral coordination and support for
land and water management and agriculture
and food systems.

Many initiatives are guiding climate change


adaptation. These include supporting
research and innovation, setting up informa-
© FAO/Giulio Napolitano

tion systems, and fostering and publishing


regular reporting processes at different levels.
However, implementation requires adjust-
ment to specific contexts. For example, in the
livestock sector, policies and instruments are
needed to differentiate among diverse farm-

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 109
demand for food and from other sectors. An
annual level of 500–1 000 m3/capita denotes
“water scarcity” and an annual level less than
500 m3/capita denotes “absolute water stress”,
© FAO/Andrew Esiebo

with serious impacts for the environment and


socioeconomic development (Falkenmark,
Lundqvist and Widstrand, 1989).

In many countries, surface and groundwa-


ter resources may be plentiful but unevenly
distributed and difficult to access. The meta-
2.3 Diminishing data for SDG target 6.4.2 give classes for water

per capita water


stress based on water availability, where low
stress is 25–50 percent, medium stress is

resources availability 50–75 percent, high stress is 75–100 percent


and critical stress is >100 percent of freshwa-
ter withdrawals relative to total freshwater
The reduction in per capita water resource resources, taking account of environmental
availability (supply) and its relation to flow requirements.
withdrawals (demand) is evaluated at the
In 2018, Central Asia had significant annual
macro level with the derivation of SDG
water withdrawals reaching 1 370 m3/capita
indicator 6.4.2 on water stress (Chapter 1).
and Northern America had 1 159 m3/capita.
Arguably, this is the most salient consequence
Since 2000, the global population has
of economic growth and climate change, and
increased eight times faster than per capita
is felt by all types of farming communities
water withdrawals. But, withdrawals per
as large swathes of productive land are capita have mostly declined as the global
subject to drought and reduced surface population has continued to grow. Since
water allocations. The reaction of many 2000, only Central America and the Carib-
smallholders and commercial farmers has bean, Southern America and Southeast Asia
been to turn to groundwater as a “lender of have seen increases in per capita withdrawals
last resort”, which, in turn, has unleashed (Chapter 1).
another set of social and environmental
More than 733 million people live in coun-
externalities.
tries with high (70 percent) and critical
(100 percent) water stress areas, accounting
2.3.1 Water scarcity for almost 10 percent of the global popu-
as a driver lation in 2018. Between 2018 and 2020, the
number of people living in areas under crit-
As population increases, IRWRs per capita
ical water scarcity increased from 6 percent
are declining (Chapter 1). At a regional level,
to 7 percent, but in high water scarcity areas,
Southern America has the most water avail-
numbers have decreased from 4 percent to
able annually, with 29 357 m3/capita, followed
2 percent (Ffigure 2.3). Population density is
by Oceania with 27 903 m3/capita. In contrast,
higher in critical and high water-stressed
Northern Africa has only 237 m3/capita and
basins (Ffigure 2.4).
Western Asia 1 379 m3/capita. These are both less
than the annual level of 1 700 m3/capita that is The evolution of large-scale irrigation
considered to reflect “water stress” and which schemes, such as those found in the Indus,
compromises a nation’s ability to meet water Mekong and Nile basins, has seen the

110 2. SOCIOECONOMIC SETTINGS


FIGURE 2.3 Population distribution according to country threshold
water stress, 2000 (left) and 2018 (right)
2000 2018
4% 6% 2%
2000 2018 7%
4% 6% 2%
20% 7%
20%
21%
21%

30%
30% 31%
31%

39%
39%
39%
39%

No stress Low stress Medium stress High stress Critical stress


No stress Low stress Medium stress High stress Critical stress
Source: FAO & UN-Water. 2021. Progress on level of water stress: Global status and acceleration needs for SDG indicator 6.4.2. Rome.
https://doi.org/10.4060/cb6241en

FIGURE 2.4 Population density mean (people/km2) by water stress class at major
basin level, 2018 (%)

100

75
Percent

50

25

0
No stress Low stress Medium stress High stress Critical stress

Source: FAO & UN-Water. 2021. Progress on level of water stress: Global status and acceleration needs for SDG indicator 6.4.2. Rome.
https://doi.org/10.4060/cb6241en

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 111
© FAO/Isak Amin/WFP as solar-powered pumps increase in popu-
larity and high-value crops offer sufficient
financial rewards. Meanwhile, unsustainable
levels of pumping continue and are increas-
ing (FAO, 2018b).

development of irrigated areas outstrip the These global trends have been documented
available supply from surface flows. The in many semi-arid and arid regions, such as
environmental externalities resulting from in Algeria, Australia, China, India, Mexico,
the scale of demand on river flows and built Morocco, Spain, Tunisia and the United States
storage are all too apparent (Molle and of America, where scarcity of renewable
Wester, 2009). The hydraulic adjustment to surface water supplies has made groundwater
changing demands for irrigation and drainage a strategic resource for irrigation (Margat and
services has lagged behind rising demand van der Gun, 2013). Increases in groundwater
or failed to adjust through institutional use are anticipated in sub-Saharan Africa,
rigidity. The resulting inequity in timing and where food insecurity is a principal driver,
duration of surface water allocations and the and easier access to pumping technology
political economy of farmer and irrigation combined with the occurrence of shallow
organization adaptation has been well groundwater circulation offer expanded
recognized (Chambers, 1988). opportunities for smallholder farming.

2.3.2 Entrenched
Access to land and water relies mainly on
informal arrangements in which access by
groundwater dependency farmers to production factors and markets
can be highly informal (López-Gunn, Rica
The social reaction to apparent water scarcity
and van Cauwenbergh, 2012; Kuper et al.,
or lack of water service has been profound.
2016). Formalizing tenure and responsibility
Many countries are now concerned about
for groundwater abstraction and aquifer
increasing dependence on groundwater for
pollution has proved challenging (FAO,
domestic, industrial and agricultural use.
2020a), primarily due to the wide range of
Scarcity of surface water resources and the
local groundwater governance arrangements
availability of affordable pumping equip-
practised among competing users
ment combined with energy subsidies have
(Blomquist, 1992).
driven demand from smallholder irrigators
for groundwater as an alternative to unre- Despite the concerns of overexploitation,
liable surface supplies, primarily because of many countries continue to use public
the convenience and control over on-farm policies to subsidize wells and boreholes,
abstractions. Groundwater irrigation has energy costs and land policies allowing the
been triggered in many semi-arid regions development of newly irrigated areas. While
since the 1960s by a combination of easily the negative impacts of groundwater-based
accessible pumping and irrigation technol- irrigation have been apparent in all econ-
ogies, promoted by public policies through omies (Steenvorden and Endreny, 2004),
subsidies for equipment and energy inputs many countries remain tolerant because of
(Molle, Shah and Barker, 2003). This was the political and economic stability offered by
referred to as the “silent revolution”, which continued access and the compelling nature
quickly turned into a form of “anarchy” that of groundwater tenure (FAO, 2020a).
threatened long-term groundwater access
and water quality (Shah, 2009). This pattern The impact of soil and land management on
of exploitation is becoming more complex groundwater quality through application of
nutrients and pesticides is pervasive, and

112 2. SOCIOECONOMIC SETTINGS


the tainting of water recharge to shallow also indicates a trend towards large farms
aquifers affects all users and communities. in higher-income regions compared with
Groundwater pollutants are easily embedded low- and middle-income countries (Lowder,
in aquifer fabrics and are extremely difficult Sánchez and Bertini, 2021). In other regions,
or impossible to remove. Competition for the 2 ha share decreases as average income
groundwater quality can be as intense as levels rise (Lowder, Sánchez and Bertini,
competition for groundwater quantity among 2019). In Central America and the Caribbean,
urban and rural communities at all scales small farms represent only about 35 percent
(Barraqué, 2011). Serious impacts on human of holdings. Much of the land (about
health from the use of polluted groundwa- 90 percent) is operated by 8 percent of farms
ter resources contaminated with arsenic or larger than 50 ha. In Europe, Northern Africa,
other heavy metals and chemical or biolog- East Africa, the Near East and Central Asia,
ical contaminants are increasingly being 60–70 percent of all farms are smaller than
reported. 2 ha, but more than half of the land is farmed
by holdings larger than 10 ha.

2.4 Patterns of Between 1960 and 2010, the average farm

landholding
size decreased in nearly all low and lower
middle-income countries with data avail-
able, although in some countries, there was a

2.4.1 Landholdings
slight increase in average farm size between
2000 and 2010. Farm size increased in a third
and farm size of middle-income countries and nearly all
high-income countries. Food system trans-
Statistical analysis of agricultural census data
formation may affect farm size and income.
from 129 countries estimates that there are
For example, the number of small farms
now over 608 million farm holdings on agri-
producing food consumed close to the source
cultural land (Lowder, Sánchez and Bertini,
and the expansion of organic agriculture
2019; Ffigure 2.5). Some 43 percent of the
increased as local famers used local markets
holdings are located in East Asia and Oceania,
supplying the urban populations (Lowder,
including China (34 percent) and South Asia
Sánchez and Bertini, 2019).
(30 percent). The size of individual holdings
is highly skewed. Farms smaller than 1 ha An upper limit of 2 ha is typically identi-
account for 70 percent of all farm holdings fied as the cropland area of a smallholding.
but operate on only 7 percent of agricul- Another measure is the number of livestock
tural land, while land holdings larger than
50 ha operate on more than 70 percent of the
world’s farmland.

Approximately 80 percent of farms smaller


than 2 ha (nominally “smallholders”)
are in low- and middle-income countries
mainly in sub-Saharan Africa, South Asia,
© FAO/Richard Trenchard

Southeast Asia and East Asia (Ffigure 2.6).


Only 40–50 percent of farms smaller than
2 ha are in upper middle-income and
higher-income countries (mainly in Central
America and the Caribbean and the Near
East and North Africa), and the research

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 113
FIGURE 2.5 Worldwide distribution of farms and farmland, by land size class, 2010
census data for 129 countries and territories

Share of farms Share of agricultural area

80%

70%

60%

50%
Percent

40%

30%

20%

10%

0%
<1 1-2 2-5 5-10 10-20 20-50 50-100 100-200 200-500 500-1000 > 1000

Hectares

Source: Lowder, S.K., Sánchez, M.V. & Bertini, R. 2021. Which farms feed the world and has farmland become more concentrated?
World Development, 142: 105455.

FIGURE 2.6 Share of value of food production from smallholders (<2 ha), by region
and income grouping, 2010 census data for 129 countries and territories

All income groups and regions (112) 36%


Low-income (18) 44%
Lower-middle-income (includes Nigeria and India) (25) 41%
Upper-middle-income (includes Brazil and China) (29) 51%
High-income (40) 3%
East Asia and the Pacific without China (9) 44%
China 80%
Europe and Central Asia (11) 8%
Latin America and the Caribbean without Brazil (15) 4%
Brazil 1%
Middle East and North Africa (10) 20%
Asia without India (6) 40%
India 47%
Sub-Saharan Africa without Nigeria (18) 35%
Nigeria 5%

0% 10% 20% 30% 40% 50% 60% 70% 80% 90%

114 2. SOCIOECONOMIC SETTINGS


operated or owned by individual farmers
and their families (Thapa, 2009). However,
the land area threshold varies greatly across

© FAO/GMB Akash
national statistical authorities, and small-
holdings can be defined according to various
criteria: endowment of land, labour and tech-
nology; type of management of the holding
and degree of family involvement; market
orientation; and/or economic terms such as pastureland for grazing, while mixed crop–
value of production. Some smallholders may livestock (agropastoral) and agroforestry
specialize in one activity but not to the exclu- systems are mutually supporting in terms of
sion of other options for food and income. The inputs and resources management.
tendency is to diversify and develop complex
This “macro approach” to farming systems
livelihood systems using family labour and
helps analyse resource use and impacts. Some
available resources.
systems use large quantities of resources
(including non-renewable fossil fuels) and
At the other extreme, large commercial land-
contribute to land and water pollution and
holdings dedicated to agribusiness present a
GHG emissions. This is particularly the case
distinct governance target upon which regu-
for intensive livestock-raising (Herrero et al.,
lation of land and water management can
2009), where industrial-scale animal-raising
achieve impacts at scale. However, the spatial
concentrates most of the negative impacts on
arrangement may be complex, with small- the environment and resource use.
holder subsistence agriculture practised side
by side with plantation operations. Smallholder farming systems coexist with
off-farm work and migration. Although this
2.4.2 Farming systems trend is not new, over the past 20 years,
off-farm work and remittances from family
Despite the polarization of land holdings, the migrants (not limited to cash transfer) have
global farming systems described by Dixon, grown in importance. According to the World
Gulliver and Gibbon (2001) are becoming Bank database, remittances represent up to
more diverse in response to changing market a third of their external inflows. In Africa,
conditions and climate change. Farming they represent up to 22 percent of gross
systems can still be classified generally domestic product (GDP) (Dridi et al., 2019).
between rainfed and irrigated systems The COVID-19 crisis has demonstrated how
(FAO, 2020b). The application of freshwater dependent some highly productive intensi-
over and above naturally occurring rainfall fied farming systems are on migrant workers
or occult precipitation still marks a sharp and how rural households rely on remittances
dividing line between methods of cultivation from migrants.
and application of inputs. Within those broad
categories, individual farming systems are The FAO Globally Important Agricultural
demonstrating a high degree of diversity as Heritage Systems (GIAHS) programme aims
they adapt to changing market conditions to identify and promote outstanding agricul-
and changing climates. The wide-ranging tural systems that have evolved over gener-
irrigated systems have diverse sources of ations in specific sites in all ecoregions to
water, equipment and infrastructure. Rainfed provide aesthetic landscapes that combine
systems distinguish between cropland, agricultural biodiversity, resilient ecosys-
including tree and fodder crops, and tems and valuable cultural heritage (Bfox 2.1).

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 115
Box 2.1
Globally Important Agricultural Heritage Systems
Through conservation, ecological knowledge systems and adapted biodiverse agricultural practices,
GIAHS generate food and livelihoods in rural areas and also deliver public goods by shaping and
modelling biocultural landscapes. FAO has designated 62 systems in 22 countries in all regions as
GIAHS since 2005, with most being in Asia and the Pacific. Building on local knowledge and experiences
and the profound relationship between people and nature, they sustainably provide multiple goods
and services and ensure food and livelihood security for millions of small-scale farmers and local
communities. However, they represent only a limited share of all the families involved in farming
systems that are able to adapt to continuous change.

Extensive knowledge of the environment and biodiversity allows these farming populations to farm
and manage territories with strong environmental constraints and risks such as mountain areas (see
the in-focus section at the end of Chapter 1) and dry lands (see the in-focus on dryland systems at
the end of Chapter 4), where corporate farming is unlikely to substantively invest. Unfortunately, many
factors threaten these agricultural systems, including climate change and increased competition for
natural resources. They are also dealing with migration due to low economic viability, remoteness and
lack of adequate support and investment, which has, in some cases, resulted in traditional farming
practices being abandoned and the loss of endemic species and breeds. Increasing recognition of
the small-scale and family farmers’ roles in maintaining the landscape and managing land and water
resources is expected to bring in more support for the sustainable evolution of such heritage systems
and their biodiversity, landscapes and cultures.

Source: FAO & Globally Important Agricultural Heritage Systems. 2022. Globally important agricultural heritage systems. In: FAO. Rome.
www.fao.org/giahs/en

2.4.3 The agencies focus on downstream markets that


serve as outlets for smallholders’ products
smallholder challenge and services.

Smallholders with land holdings less than


Nevertheless, smallholders also engage in
2 ha may occupy only 12 percent of agricul-
upstream markets where they can acquire
tural land but are nevertheless responsi-
specific inputs (and technologies). They use
ble for diverse and unique farming systems.
family labour to engage in diverse labour
Smallholders seek security in a diverse
markets locally or through migration. Small-
set of activities, including harvesting wild
holders are also part of the general market for
plants, hunting and fishing, exploiting non-
consumer goods. In addition, land markets
renewable natural resources and off-farm
are part of the daily life of smallholders,
activities. Specialization is a risky way to
where they may rent, buy or sell land. They are
develop and is not usually the dominant
also engaged in financial markets (including
pathway for smallholders (Bonnal et al.,
informal lenders) to acquire capital to cover
2018). Smallholders no longer live in a world
operations costs and investments. Small-
where they are isolated, producing for their
holders are fully part of the market economy,
subsistence and looking reluctantly at the
even in remote places. The challenge is about
monetary economy. Most are fully involved
improving the conditions that govern their
in market-driven activities even if providing
participation in markets, which are rarely
enough food for the family remains a proven
favourable (HLPE, 2013).
and efficient safety net. Usually, development

116 2. SOCIOECONOMIC SETTINGS


Although size is a common feature, small- 2015; Wada et al., 2016; Guiomar et al., 2018).
holders face different challenges in differ- The sustainable rural livelihoods (SRL)
ent socioeconomic settings, have different framework is suitable for analysing the live-
farm sizes, and manage land and water lihoods of smallholders and family farmers.
resources in different ways. Smallholders It relies on access to a number of liveli-
produce a significant share of food and traded hood resources (e.g. human, natural, social
commodities in world markets (e.g. coffee, and economic assets). This framework also
cocoa, natural rubber and rice). Yet, they are considers a range of formal and informal
rarely recognized as the backbone of agri- institutions that influence the livelihoods of
cultural development (Rafflegeau et al., 2015; smallholders and family farmers (Bbox 2.2).
Samberg et al., 2016). Most manage their
natural resources and have limited resources Smallholders are part of a multiple and

to invest in resource improvement, except complex set of local solutions available to

through their labour. Several assessments restore and improve the quality of the diverse

have demonstrated how smallholders use ecosystem in which they live and work. The

their skills, knowledge and family labour to challenge lies in mainstreaming policies,

produce food and manage natural resources, programmes and support to enhance their

often in harsh environments (Sourisseau, productive and entrepreneurial capacities

Box 2.2
The SUSTAINABLE RURAL LIVELIHOODS framework
The SRL framework (Scoones, 2009) describes the situation of most smallholders and family farmers
(Bosc et al., 2014; Bonnal et al., 2018). It relies on human, natural, social, physical and financial assets,
and accounts for different types of capital and the institutions and organizations that make these
investments possible. Activities can be socially or market oriented, which corresponds to the situation
of most smallholders.

The SRL framework considers natural resources as natural capital given by local resources endowment
that is also a product of human actions (Figure 2.7). Thus, investments in natural capital imply access
and security conditions that do not necessarily mean private property (Ciriacy-Wantrup and Bishop,
1975; Oakerson, 1992; Ostrom, 1992). Investment in natural capital may also involve collective action.
In such cases, this will depend on coordinating capacities at the territorial level involving individual,
customary and local authorities and collective stakeholders. Inequality of access or lack of access for
the most vulnerable may require public action to redistribute or allocate land through agrarian reforms.

The SRL framework helps to identify the different levels of investments that directly influence or im-
prove the capacity of smallholders to invest, increasing their capabilities through social relations,
institutions and organizations that provide increased opportunities to individuals. This includes several
types of collective investments: (i) in landscapes and resources management, (ii) in improving access
to markets through cooperatives and associations, (iii) in self-help groups (socially oriented), (iv) in cor-
porate and private stakeholders upstream and downstream and (v) in public goods. This framework is
valid for households and families and takes account of organizations and institutions. It also replaces the
household/family level in an environment that considers national and international trends, shocks and
other unpredictable events.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 117
118
Box 2.2 (continued)

FIGURE 2.7 Analysing smallholdings and family farms through the SUSTAINABLE RURAL LIVELIHOODS framework

Context / trends Structure Activities Performance/sustainability


Environment/determinants Access permitted by... Livelihood platform Composed of and resulting in With effects on
Composed of

National and international CAPABILITIES / FUNCTIONINGS Natural resources Strategies Social and human
trends and contexts based activities sustainability
Specialization
(Ensuring continued Social relations Assets Crops (food and markets) (individual/family, Improved education
functionality of healthy and Livestock agricultural/non- Health situation
Gender/social class Natural capital
sustainable food systems) Other non-agricultural agricultural) Social and political
Age Physical capital
activities based on Diversification participation
Ethnicity Human capital
natural resources (agricultural/non- Collective dynamics
Urban/rural Financial capital
Local trends and context agricultural) Concentration, fragmentation
Participation in social & Social capital
political life Intensification
Population, migration, local
Non natural resources Extensification
policies, local economic
based activities Migration Economical sustainability
trends (including supply
Income strategies
chains), collective Organizations Institutions Wages Production & income
(rents)
dynamics, access to public Rural trade Degree of stability
goods, access to natural Associations, Rules and customs Other services and rural
NGOs, Land tenure Seasonality
resources, etc. manufacture Degrees of risk
Local administration, Markets in practice Remittances
State agencies, Food security Integration
Other transfers by markets
Supply chains
Shocks
Value chains
Drought, floods, pests,
Environmental sustainability
diseases, civil war
Soil and land quality
Water
Rangeland
Forests
Biodiversity
Energy balance
Carbon balance

2. SOCIOECONOMIC SETTINGS
Source: Bosc, P.-M., Sourisseau, J.-M., Bonnal, Gasselin, P., Valette, E. & Bélières, J.-F. 2014. Diversity in family farming : Theoretical and empirical approaches to its many forms. Rethinking Development Working Paper Series No.
2014/2. University of Pretoria. https://doi.org/10.13140/RG.2.1.4444.8409
reports that women and youth often have
limited access to smaller and poor-quality
plots of land and insecure tenure rights
© FAO/Horst Wagner

(Oxfam, 2016). Indigenous people often


have difficulties obtaining recognition
and registration of their legitimate tenure
rights at the same pace as private operators.
Even when their rights are acknowledged
and granted, they may face difficulties in
and result in positive potential implications preventing encroachment. There is also an
on the state of the resources they manage. imbalance between the rights of first settlers
granted unsecured access rights and new
Despite the important role of smallholders in groups willing to settle or use part of the
food production and landscape management, resources controlled by the early comers. The
they tend not to be the ones receiving atten- same applies with pasture rights threatened
tion as part of ongoing investments in land by farmers extending and enclosing their
cropped area and questioning the customary
and water, nor do they receive policy support
post-harvest pasture rights when their own
to formalize the allocation of their land-
needs for fodder develop.
and water-use rights. The current context of
climate change even further accentuates the The recent Land Inequalities Initiative
importance of access to water for smallholder (Anseeuw and Baldinelli, 2020) suggests
and family farmers. traditional land inequalities analysis, based
on the Gini coefficient for land distribution,

2.5 Access to
underestimated inequalities by omitting
criteria such as multiple ownership of

land and water plots, land values and landless population.


In 17 countries, analysis showed 10 percent
of the largest landowners accumulated
The rules of access, regulation, competition 40–60 percent of the land value while the
and conflict resolution defined by gover- bottom 50 percent of small landowners
nance shape how natural resources are accumulated only 6–10 percent (Bauluz,
managed over time and define the condi- Govind and Novokmet, 2020). The analysis
tions for restoring and improving renewable also showed land inequalities increased when
resources, including investments, incentives landless persons engaged in agriculture
and disincentives. These require understand- were considered in countries with landless

ing and capacities within communities, local populations.

authorities and higher levels of governance to


Increased concentration of farmland among
interact effectively to address issues affecting
larger farms in countries with higher
sustainable resource use. income levels is occurring in most of the
larger European countries (except Spain), in
2.5.1 Power asymmetries Brazil and in the United States of America.

over land There is increased inequality with an appar-


ent re-emergence of small farms, while the
Local authorities and communities face power share of farmland on the largest holdings
asymmetries, either within communities has increased. In 2010, the average farm size
or with external entities or individuals. In was 1.3 ha in low-income countries, 17 ha in
Central America and the Caribbean, Oxfam lower middle-income countries, 23.8 ha in

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 119
upper middle-income countries (excluding In Africa, large- and medium-scale
China) and 53.7 ha in high-income countries acquisitions have been facilitated by
(Lowder, Sánchez and Bertini, 2019). commodifying rural land markets, leading
to the privatization of public and customary
Investments in large-scale land acquisition land. In Central America and the Caribbean,
can bring prosperity, but may adversely there are documented trends in land
affect local communities and smallholders concentration (Baquero and Gómez, 2014)
if not adequately planned. Investors target due to growing interest in large-scale lease
irrigated lands for their potential as they seek schemes for the annual production of soya
to secure their capital and get a good return and corn in the Southern Cone and through
on investment. This is a complex activity large-scale land ownership for all type of
with diverse actors, contracts and practices production in the rest of the region (Bres,
around land tenure and water rights, often 2017). In many countries across the world,
within incomplete and inconsistent legal these recent trends reinforce historical
frameworks. dualistic polarized agrarian structures.

Since the first land matrix report, large-scale


land acquisitions continue to be a significant 2.5.2 Inequalities in
issue globally, with differentiated impacts access to water resources
across regions and countries (Anseeuw et al.,
Equitable access to water and land will
2012). Africa ranks first with 10 million ha
worsen with the increasing scarcity of land
acquired, representing 37 percent of the
and water and the impacts of climate change.
global acquired area, principally along the
Estimates suggest 77 percent of small-scale
main rivers in East Africa. The picture varies
farms in low- and middle-income countries
among other target countries, with the top
are located in water-scarce regions, and less
five being Indonesia, Ukraine, Russian Feder-
than a third of these have access to irrigation
ation, Papua New Guinea and Brazil (Nolte,
systems. The greatest disparities in irrigation
Chamberlain and Giger, 2018).
coverage between small-scale and medium-
Acquisitions are increasingly becoming oper- and large-scale farms are in Central America
ational and can threaten the livelihoods of and the Caribbean, sub-Saharan Africa and
rural people, who are not always part of the South Asia (Ricciardi et al., 2020). In view
negotiation process (only 60 percent of 180 of climate change impacts, poor access to
recorded cases). More than half deal with irrigation can become a major constraint
lands already under cultivation in regions for rural livelihoods, particularly in arid
with relatively high population densities, regions. Sustainable resources management
forest zones used by communities and and improved access to natural resources and
marginal lands with less-populated areas services can mitigate the adverse effects.
but crucial for pastoralism. More than half of
Irrigation investments have increased
acquisitions belonged to communities, and
crop yields and irrigated areas and induced
only 30 percent received compensation.
changes in cropping patterns, including from
Since 2000, African investors from non- single-cropping to double-cropping systems.
agricultural backgrounds and farmers Studies have found crops show significant
wishing to increase their holdings have yield increase when cultivation shifts from
accounted for a significant number of rainfed to surface irrigation in semi-arid
medium-scale acquisitions (5–100 ha). In regions. Notably, studies generally illustrate
Ghana, they accounted for about 50 percent the potential for infrastructure investments
of national cropped land (Jayne et al., 2019). to decouple economic growth from rain-

120 2. SOCIOECONOMIC SETTINGS


fall variability. Irrigation can benefit poor
consumers by reducing food prices. Ground-
water irrigation has fostered a groundwater
economy. It has helped to alleviate poverty,
boost economic growth and transform rural

© FAO/Olivier Asselin
economies in many countries in the Americas
and Europe, in Asia through the Green Revo-
lution, and more recently, in Northern Africa.
Irrigation also brings many indirect bene-
fits such as: (i) increased labour demand,
particularly during planting and harvest
periods, (ii) improved nutrition and health
farmers. Irrigation may increase poverty if
and (iii) economy-wide multiplier effects.
all legitimate tenure rights are not recog-
Smallholders that irrigate can increase their
nized, particularly those affecting the most
farm income by growing higher-value crops
vulnerable. This can lead to poor households
and increasing the availability of vegeta-
losing rights and converting marginal and
bles, fruits and cash crops. Examples include
poor farmers to landless labourers. Mecha-
Ethiopia and the United Republic of Tanza-
nization and the use of herbicides can also
nia (Passarelli et al., 2018) and countries in
replace labour on big production units.
Northern Africa (Dugué et al., 2014).
Irrigation development can have off-site
Smallholder and family farmers in many
environmental impacts that are particu-
countries have developed systems of informal
larly important for the poor. Irrigation is
irrigation. Only a minority of the world’s
often associated with soil salinization and
small-scale users of water hold a legally
heavy uses of fertilizers and pesticides; it
sanctioned water right. Small-scale irrigation
may reduce environmental service provision
is not accounted for in official national
upon which the poor rely, such as inland and
statistics, and water users are often reluctant
marine fisheries. There is growing evidence
to register their water use due to fear of water
that links groundwater irrigation expansion
fees being imposed. Yet, the informality of
with increasing socioeconomic inequalities
small-scale irrigation may increase the
as water tables decline and cropping changes.
risk of water insecurity (United Nations,

2.5.3 Gender inequalities


2018b). In addition, the apparent success of
groundwater irrigation for smallholders is
creating resources problems as local aquifers Access and management of land and water
are drawn down and degraded by migration of have strong gender and equality dimensions.
low-quality groundwater or saline intrusion. Women play a key role in ensuring food secu-
rity and managing natural resources. They
Irrigation impacts on poverty vary greatly
often have responsible roles across all agri-
depending on farm size and location within
cultural subsectors including forestry and
an irrigation system and structural issues
fisheries, while providing water, food and
related to the overall institutional and socio-
energy to their households. Globally, women
economic environment related to gender,
comprise over 37 percent of the world’s rural
caste and class. For example, irrigation
agricultural workforce, rising to 48 percent
effects on employment may benefit the land-
in low-income countries. They account for
less poor, but inequalities may also widen
about 50 percent of the world’s small-scale
as the increase in production may depress
livestock managers and about half of the
market prices, which disadvantages rainfed
labour force in small-scale fisheries. These

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 121
nexus persists and has important impacts on
realizing and securing water rights, partic-
ularly traditional customary water tenure
among rural communities. A more inte-
grated rights-based approach to tenure could
© FAO/Giulio Napolitano

unpack the relationship between rights to


water, land and other terrestrial resources,
and help to identify gaps and synergies across
sectoral legislation (FAO, 2020a).

Human rights mechanisms and United


Nations entities recognize that ensuring
percentages are likely to underestimate
women’s land and water rights are essential
women’s full contribution to agriculture
for achieving substantive equality and eradi-
as their work, often unpaid, is not always
cating many forms of discrimination against
adequately captured in official statistics.
women. This is a fundamental precondition
Women are often marginalized and vulnera- to realizing rights to an adequate stan-
ble to land tenure and water insecurity, and dard of living, including food and housing,
are less resilient to climate shocks. Compared health, well-being, work, cultural identity
to men, they have less access and control of and participation in civil and political life,
assets to increase their resilience to climate particularly when land and water manage-
and economic shocks (e.g. infrastructure, and ment infrastructure is being planned and
land and water rights), and less access to irri- developed (United Nations OHCHR and
gation and opportunities. Women often have Heinrich-Böll-Stiftung, 2018). The 2030
limited participation in decision-making and Agenda recognizes women’s land and water
do not usually benefit from land and water rights as an explicit cross-cutting cata-
investments. lyst to ending poverty (SDG 1), seeking to
achieve food security and improved nutrition
Land and water management investments (SDG 2), ensuring clean water and sanitation
can be gender blind, and fail to respond to for all (SDG 6) and achieving gender equality
the specific needs of women or address the and women’s empowerment (SDG 5).
issues constraining women farmers’ empow-
erment. Land and water tenure insecurity can Investment in women and girls can be a
come from discriminatory laws and practices catalyst to accelerate progress in agricul-
at national, community and family levels. For ture, rural development and, ultimately, food
example, fewer than 50 countries have laws security and nutrition. This should integrate
or policies that specifically mention women’s the gendered implications of land and water
participation in rural sanitation and water investments, such as the appropriateness of
resources management (UN-Water, 2021). technologies, governance arrangements and
financing mechanisms, to ensure effective
National and global data on women’s access targeting of poor and vulnerable groups.
to land and water that would enable them to Women are not a homogeneous group, and
monitor and enforce their rights are lacking. social dimensions such as age and ethnicity
Land and water tenure are often linked. While play a role in determining multiple forms of
contemporary water laws tend to decouple marginalization and exclusion.
water rights from land tenure, the land–water

122 2. SOCIOECONOMIC SETTINGS


©FAO/ Nikos Economopoulos
2.6 Competition for
land and water – an
issue of governance
FAO defines governance as formal and
complex policy bottlenecks, political conflicts
informal rules, organizations and processes
and local organizational realities that impede
through which public and private actors
effective decision-making and land and water
articulate their interests and make and
governance in practice (FAO, 2021).
implement decisions. Governance issues
arise in various public and private settings,
The focus has also shifted towards a more
from local communities, farms and coopera-
pragmatic agenda committed to bottom-up
tives, business organizations and large-scale
problem-solving approaches that recognize
enterprises, to local, regional, national and
the development process as being deeply
international contexts. It includes social,
rooted in established socioeconomic, cultural
political and economic dimensions, tradi-
and political relationships at national and
tional authorities, and customary laws and
local levels (Gonzalez Fischer and Garnett,
norms (UNDP and EC, 2007; Bruch, Muffett
2016; FAO, 2021). Indeed, land and water
and Nichols, 2016; FAO, forthcoming).
resources generate revenues and drive
economic growth. Yet, the way institutions
Governance over land and water resources
operate and cooperate, and the relative power
relates to the enabling environment in which
and capabilities of different actors, strongly
land and water management actions take
shape outcomes and welfare distribution.
place at multiple levels of decision-making:
Strengthening governance is therefore about
the overarching policies, strategies,
enabling effective and efficient problem
plans, finances and incentive structures
solving and decision-making in ways that
that concern or influence land and water
stakeholders regard as legitimate.
resources; the relevant legal and regulatory
frameworks and institutions; and planning,
decision-making and monitoring processes. 2.6.1 Adoption of land
Effective governance promotes responsible tenure guidelines
actions and measures to protect and ensure
the sustainability of resources for current and In building the capacities of countries to
future generations and optimize the services improve the governance of land tenure, FAO
and benefits obtained from those resources. has developed technical guides for a range
of actors that provide practical mechanisms,
There has been a shift from focusing on processes, good practices and tools for the
promoting “good governance” principles – design of policy and reform processes, for
from participation (including the rule of law, investment projects and for guiding imple-
transparency, responsiveness, consensus mentation of the Voluntary guidelines on
orientation, equity, inclusion, effectiveness, the responsible governance of tenure of land,
efficiency and accountability (United Nations fisheries and forests in the context of national
ESCAP, 2009)), to establishing a formal food security (VGGT; CFS and FAO, 2012). The
normative set of institutional, financial and technical guide on governing tenure rights
organizational procedures for regulating to commons (FAO, 2016b) provides recom-
natural resources. These include informal mendations and calls on States to meet their
and operational approaches that address the obligations to secure legitimate tenure rights.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 123
The Committee on World Food Security (CFS)

© Benedicte Kurzen/NOOR for FAO


country members and stakeholders negotiated
and adopted the VGGT. The guidelines
promote secure legitimate tenure rights and
equitable access to land, fisheries and forests
as a means of eradicating hunger and poverty,
supporting sustainable development and
enhancing natural resources management.
They set out principles and internationally
accepted standards or practices for the nance to address: (i) the linkages among
responsible governance of tenure that agriculture, water and related key sectors and
States can use when developing their own elements such as food, land, energy, natural
strategies, policies, legislation, programmes resources, societal goals and major drivers
and activities. They allow governments, civil of change; (ii) a moving intervention scale
society, the private sector and citizens to from management to governance of water
judge whether their proposed actions and in agriculture, pointing to underlying issues
the actions of others constitute acceptable that management approaches alone cannot
practices. solve; and (iii) governance issues of access,
rights and tenure from the perspective of
2.6.2 Emergence sustainability, inclusiveness and efficiency.
of water tenure The FAO Council and committees on food
security and agriculture are kept informed
The VGGT focus on land tenure, and did
of advances in land tenure and water tenure,
not initially include water tenure because of
and address the crucial links with agriculture
its complexity as a transboundary resource
and food security.
beyond the scope of national sovereignty
alone. However, because of its fundamental Agriculture, which accounts for 70 percent
importance for effective water allocation and of all water withdrawals, is increasingly
management and for agriculture and food required to “make its case” for its share of
security, in 2014, FAO adopted a framework water to enable food production and ensure
for systematic engagement in water gover- food security. The sustainability of agricul-
tural water use is increasingly under scrutiny.
There is an urgent need to consider how best
to address control over, competition for and
access to water resources while also ensur-
ing efficient and effective management. In
many places, water governance regimes have
not kept pace with growing competition for
water and are not conducive to its efficient
and equitable management. Mechanisms to
reflect values in conditions of resource scar-
© FAO/Simon Maina

city and increase resource-use efficiency are


generally lacking. Moreover, the water-use
rights held by farmers are often not protected
by law or formally registered.

124 2. SOCIOECONOMIC SETTINGS


FAO is developing the concept of water tenure
to better assist Members in making the

© FAO/Noah Seelam
institutional, legal and political adjustments
needed for successful water management
(FAO, 2016b, 2020a). This includes governance
of water in river basins and watersheds,
groundwater governance, governance of
irrigation, governance of water for pollution
control and water quality management, and on Access and Benefit-sharing from the utili-
management and putting food security at the zation of genetic resources.
centre of the international water debate. In
view of the fundamental importance also for 2.6.3 Conflicts over land
human health, increasing emphasis is needed
and water resources
on governance for sustaining water quantity
and quality. The potential for conflict is increasing as
populations compete for resources in land-
Efforts have intensified to support effective
and water-stressed areas. Conflicts continue
water governance through international
to grow in areas such as: the Horn of Africa
initiatives such as the OECD water
and the Near East, where conflict exists
governance principles (OECD, 2015) and
between Israel and Palestine in Gaza and
the report on implementation, which took
the Golan Heights; Eritrea and Ethiopia over
stock of progress and proposed two tools –
land; and Egypt and Ethiopia over River
an indicator framework, and 50+ concrete
Nile water resources. Water conflicts have
practices (OECD and FAO, 2018) and World
intensified over the past decade, according
Water Week – which have helped advance
to the Pacific Institute. However, tensions
knowledge and promote more effective
do not always lead to conflicts and wars;
governance. Target 6.5 of SDG 6 provides an
they can lead to negotiations and improved
agreed global target on water governance:
cooperation (Yoffe, Wolf and Giordano, 2003;
the implementation of integrated water
Michel, 2020).
resources management (IWRM) at all levels
based on principles of social equity, economic Disagreements over water allocation among
efficiency and environmental sustainability. countries that share river systems are a
common source of political conflict, espe-
For the land and water governance agenda,
cially where demands outgrow the available
lessons can also be learned from measures
resources. Water scarcity may well increase
and best practices in other sectors, nota-
transboundary conflicts if political discussion
bly the development of farmers’ rights16 over
and appropriate governance arrangements
genetic resources (Lowder, Sánchez and
fail to prevent them. About 40 percent of the
Bertini, 2019) and an inventory of national
world’s population lives in transboundary
measures, best practices and experiences of
river basins, which cover about half of the
the United Nations Convention on Biological
Earth’s land surface. Globally, more than 300
Diversity (CBD) through the Nagoya Protocol
watersheds and over 360 aquifers cross the
political boundaries of two or more countries,
6
Farmers’ rights developed under the International highlighting the need for effective trans-
Treaty on Plant Genetic Resources for Food and
Agriculture and the Commission on Genetic Resources boundary governance arrangements among
for Food and Agriculture, which covers biological countries and local populations.
diversity for food and agriculture.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 125
Concerns over national sovereignty pervade

© FAO/Jon Spaull
shared resources management, but this does
not stop international cooperation on planned
development of large dams on transboundary
rivers, even when there is a long-standing
history of mistrust among neighbouring
States. The experience of international Conflict sensitivity should be integral to
cooperation over shared aquifers is more any interventions involving land and water.
limited and finds more application in national Participatory and inclusive approaches should
jurisdictions with federated territories. In be central to strengthen buy-in, transpar-
all transboundary resource negotiations, ency and sustainability. Despite agreement
the burden of proof on land-use impacts on on important component principles, there is
water quantity and quality is perhaps the no consolidated and agreed set of principles
most demanding. This is particularly the case for joint land and water management nor an
in shared aquifers where the effects of land agreed international integrated framework.
management practices are highly distributed Cooperation should be a means to sustainable
and are expensive to monitor and control use of transboundary water resources.
over the required periods of time.

The tension between farming communities 2.7 Conclusions


and pastoralists has become marked. Pasto-
ralism has often developed in territories Demographic and economic growth are driving
where land is marginal or not suitable for food demand, placing unprecedented pressures
crop production, as the best land-use option. on ecosystems and limited renewable land, soil
and water resources. Socioeconomic trends,
Another perspective is that pastoralism is
including urbanization, migration and
also a means of taking the best of harsh envi-
technological change, will continue to drive
ronments marked by a high climatic vari-
the distribution of these pressures on available
ability, by managing resources (fodder and
natural resources. Higher incomes and urban
water) that are randomly spread over vast
lifestyles are changing food demand towards
territories. Pastoralism is often combined more resource-intensive consumption of
with some type of seasonal cropping, which animal proteins, fruits and vegetables. At the
generally has “very low productivity” from same time, malnutrition persists among the
an agronomic perspective but is important urban and rural poor who are disconnected
for nutrition and diet. from markets or access to productive land
through poverty or geography.

Globally, 80 percent of the extreme poor live in


rural areas. Most live in the developing world,
and their livelihoods are disproportionally
dependent on agriculture, which is highly
exposed to current and future climate risks.
Ensuring equitable access to land and water
resources is key for promoting inclusive rural
transformation. The lack of adequate access
and user rights and increasing disparities in
© FAO

capacities to take advantage of natural capital

126 2. SOCIOECONOMIC SETTINGS


are underlying drivers of overuse of resources Governance over land and water resources
to meet short-term needs. requires an enabling environment in which
land and water management actions take
Underneath the patterns of economic growth, place at multiple levels of decision-making.
competition for land and water resources Social, agricultural and environmental
is intensifying. Increasing population is policies need to be harmonized mutually
reducing the natural resources available per reinforcing if they are to reconcile competition
capita. More than 733 million people live in over land and water. There is progress in
countries with high (70 percent) and critical land tenure initiatives, but land and water
(100 percent) water stress areas, accounting allocation adjustments will be possible only
for almost 10 percent of the global population when explicit instruments are joined up and
in 2018. Over the past decade, sub-Saharan resource management decision-making
Africa experienced a 40 percent reduction in become inclusive. Integrated land water
water availability per capita, and a decline resources planning is urgently needed to
in agricultural land from 0.80 ha/capita guide land and water use, not just to promote
to 0.64 ha/capita between 2000 and 2017. sustainable resource management but to
Northern, Western and Southern Africa establish the realistic scope for reducing
have less than 1 700 m3/capita, a level that emissions.
compromises a nation’s ability to meet water
demand for food and for other sectors.

The social and economic structure of most


populations is still finely tuned to natu-
ral resources access, even as populations
concentrate in urban areas. Large-scale
commercial holdings dominate agricultural

© FAO/Noah Seelam
land use to supply global food systems, while
land tenure patterns restrict and concentrate
up to 500 million smallholdings of less than
2 ha in subsistence farming on lands suscep-
tible to degradation and water scarcity.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 127
© Soliman Ahmed

Case study: Gender empowerment


resolves water-related conflict in Yemen

In Yemen, disputes over land and water are endemic and often violent, claiming thousands of lives
each year, destroying valuable crops, delaying investment and inhibiting social and economic devel-
opment. FAO is helping farmers resolve such conflicts and gain access to water resources by improv-
ing opportunities for women and youth, who represent over 60 percent of the agricultural workforce,
to play an influential role in decision-making.

Yemen is one of the Arab world’s poorest nations and is among the world’s most water-stressed coun-
tries (SDG indicator 6.4.2, level of water stress, is 170 percent). Yemeni farmers have long coped with
their harsh environment and water scarcity, and developed indigenous water management practices
to regulate water allocation. However, because
of the unequal power relations between the
Woman providing water for cattle in a Yemeni
genders, women are not usually involved in
smallholding – Al Hudaydah, Yemen
managing natural resources. Indeed, Yemen
ranks last out of the 144 countries listed in the
2016 World Economic Forum’s Global Gender
Gap Index, a position it has held for the last
ten years.

The war in Yemen began in 2014; it is a complex


mix of politics, socioeconomics and history
connected to resource scarcity and has inflicted
© Chedly Kayouli

severe damage to the country’s water infra-


structure. As men were drawn into the conflict,
many women became heads of households and
key family decision-makers.

128 2. SOCIOECONOMIC SETTINGS


In response, FAO developed a project to reha-
Women learn about the many bilitate community water supplies. This included
interconnections influencing water
encouraging and training women to take a more
sustainability – Al Hudaydah, Yemen
prominent role in water management, facilitating
and mitigating resource conflicts through medi-
ation, and discussions at the household level.

FAO helped to establish water user associations


(WUAs), comprising women’s water user groups
and also conflict resolution committees (CRCs), in
which women would participate on equal terms
with men. Women were empowered to lead
in conflict resolution and mobilize community
members to support agreed resolutions. Under
© Soliman Ahmed

the current management system, all WUAs


choose their board members through elections,
and 30 percent of the seats are for women.

As a result, 27 165 farmers improved their access


to irrigation water. Some 1 083 people (294 women; 789 men) from low-income and vulnerable
households increased income through cash for work because of the increased water availability. This
improved local food production and increased the economic prospects for vulnerable families when
the population was facing severe food insecurity.

Conflict over water supplies from the Sana’a Queen Dam storage is an example of farmers’ issues. Built
in 2002 and designed to benefit 350 farmers, a dispute led to a ban on using 170 000 m3 of stored
water annually to irrigate 34 ha. FAO worked with the community to resolve the dispute and empow-
ered women to take an active role. In partnership with Al-Malakah WUA in Bani Al-Harith District, a CRC
was formed (two women and two men) to analyse the dispute, the reasons behind it and its impact.
FAO and Al-Malakah CRC were supported by village youth and women to dispel deep-seated mistrust
and misunderstandings among the communities. They then agreed to construct shallow wells that
would connect with the water stored in the dam. Participatory negotiations resolved a 17-year-old
dispute that had prevented farmers from using water rather than leaving it to evaporate.

This project demonstrated that when women and youth play prominent roles in WUAs, they can bring
innovative ideas to mitigate many resource-based disputes. It was possible to resolve water conflicts
in a peaceful, participatory and equitable manner and improve secure access to natural resources.
Training in conflict management offers a significant opportunity to develop human and social capital.
However, training alone is unlikely to address all societal disputes. Socioeconomic and political
factors may require appropriate reform of policy, legislation and institutions to provide an enabling
environment.

Sources: FAO. 2017. FAO support provides water sustainability for farmers while empowering women. In: Resilience.
www.fao.org/resilience/news-events/detail/en/c/1045903; FAO. 2018. FAO innovative approach to resolve land and water conflict in Yemen.
In: La resiliencia. Rome. www.fao.org/resilience/multimedia/videos/video-detalle/es/c/1151119

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 129
References
Agergaard, J., Tacoli, C., Steel, G. & Ørtenblad, S.B. 2019. Revisiting rural–urban transforma-
tions and small town development in sub-Saharan Africa. European Journal of Development
Research, 31(1): 2–11.

Alpha, A. & Fouilleux, E. 2018. How to diagnose institutional conditions conducive to


inter-sectoral food security policies? The example of Burkina Faso. NJAS: Wageningen Journal
of Life Sciences, 84(1): 114–122.

Anseeuw, W. & Baldinelli, G. 2020. Uneven ground. Land inequality at the heart of unequal societ-
ies. Rome, International Land Coalition. https://oi-files-d8-prod.s3.eu-west-2.amazonaws.
com/s3fs-public/2020-11/uneven-ground-land-inequality-unequal-societies.pdf

Anseeuw, W., Boche, M., Breu, T., Giger, M., Lay, J., Messerli, P. & Nolte, K. 2012. Transnational
land deals for agriculture in the Global South: Analytical report based on the Land Matrix Database.
Bern, Centre for Development and Environment, Montpelier, University of Bern/Centre
de coopération Internationale en Recherche Agronomique pour le Développement and
Hamburg, German Institute of Global and Area Studies. https://landmatrix.org/docu-
ments/82/Analytical_Report_I_LMI_English_2012_xP1xzqu.pdf

Baquero, F. & Gómez, S., eds. 2014. Reflexiones sobre la concentración y extranjerización de la
tierra en América Latina y el Caribe. Rome, FAO. www.fao.org/3/i3075s/i3075s.pdf

Barraqué, B., ed. 2011. Urban water conflicts. United Nations Educational, Scientific and Cultural
Organization International Hydrological Programme and CRC Press.

Bauluz, L., Govind, Y. & Novokmet, F. 2020. Global land inequality. WID.world Working
Paper No. 2020/10. https://wid.world/document/global-land-inequality-world-inequali-
ty-lab-wp-2020-10

Blomquist, W. 1992. Dividing the waters: Governing groundwater in southern California. San Fran-
cisco, ICS Press Institute for Contemporary Studies.

Bonnal, P., Sourisseau, J.M., Bosc, P.M., Gasselin, P., Bélières, J.F. & Valette, É., eds. 2018.
Diversity of family farming around the world. Existence, transformations and possible futures of
family farms. Cham, Springer.

Bosc, P.-M., Sourisseau, J.-M., Bonnal, Gasselin, P., Valette, E. & Bélières, J.-F. 2014. Diversity
in family farming: Theoretical and empirical approaches to its many forms. Rethinking Develop-
ment Working Paper Series No. 2014/2. University of Pretoria.
https://doi.org/10.13140/RG.2.1.4444.8409

Bres, A. 2017. Challenge and opportunities for the VGGT implementation in Latin America and
the Caribbean. Paper presented at 2017 World Bank conference on land and poverty, Wash-
ington, DC.

130 2. SOCIOECONOMIC SETTINGS


Bruch, C., Muffet, C. & Nichols, S.S., eds. 2016. Governance, natural resources, and post-conflict
peacebuilding. Routledge.

Castañeda, A., Doan, D., Newhouse, D., Nguyen, M.C., Uematsu, H. & Azevedo, J.P. 2018. A new
profile of the global poor. World Development, 101: 250–267.

Cervigni, R., Morris, M., Scandizzo, P. & Savastano, S. 2016. Vulnerability in drylands today.
In: R. Cervigni & M. Morris, eds. Confronting drought in Africa’s drylands: Opportunities for
enhancing resilience. Washington, DC, World Bank.
https://openknowledge.worldbank.org/handle/10986/23576

CFS (Committee on World Food Security) & FAO. 2012. Voluntary guidelines on the responsible
governance of tenure of land, fisheries and forests in the context of national food security. Rome,
FAO. www.fao.org/3/i2801e/i2801e.pdf

Chambers, R. 1988. Managing canal irrigation. Practical analysis from South Asia. Cambridge, UK,
Cambridge University Press.

Chatel, C., Imbernon, J. & Moriconi-Ebrard, F. 2016. Megacities and archipelagos: An emerging
urban framework. HAL open science.
https://hal-univ-paris.archives-ouvertes.fr/hal-01713844/document

Ciriacy-Wantrup, S.V. & Bishop, R.C. 1975. Common property as a concept in natural resources
policy. Natural Resource Journal, 15(4): 713–727.

Cunningham, P.J. & Abasse, T. 2005. Reforesting the Sahel: Farmer managed regeneration.
In: A. Kalinganire, A. Niang & A. Kone, eds. Domestication des especes agroforestieres au Sahel:
Situation actuelle et perspectives. ICRAF Working Paper 5, pp. 75–80. Nairobi, World Agrofor-
estry Centre.
https://fmnrhub.com.au/wp-content/uploads/2013/09/Cunnigham-Abasse-2005-Reforest-
ing-the-Sahel.pdf

Damania, R., Desbureaux, S., Hyland, M., Islam, A., Moore, S., Rodella, A., Russ, J. & Zaveri,
E. 2017. Uncharted waters: The new economics of water scarcity and variability. Washington, DC,
International Bank for Reconstruction / The World Bank.
https://openknowledge.worldbank.org/handle/10986/28096

Dixon, J.A., Gibbon, D.P. & Gulliver, A., eds. 2001. Farming systems and poverty: Improving farm-
ers’ livelihoods in a changing world. Rome, FAO.
www.fao.org/3/y1860e/y1860e.pdf

Dridi, J., Gursoy, T., Perez-Saiz, H., Bari, M., Berg, C., Chami, R., Kebhaj, S. et al. 2019. The
impact of remittances on economic activity: The importance of sectoral linkages. IMF Working
Paper WP/19/175. International Monetary Fund.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 131
Dugué, P., Lejars, C., Ameur, F., Amichi, F., Braiki, H., Burte, J., Errahj, M., Hamamouche,
M. & Kuper, M. 2014. Recompositions des agricultures familiales au Maghreb: Une analyse
comparative dans trois situations d’irrigation avec les eaux souterraines. Revue Tiers Monde,
220(4): 99–118.

Falkenmark, M., Lundqvist, J. & Widstrand, C. 1989. Macro-scale water scarcity requires
micro-scale approaches: Aspects of vulnerability in semi-arid development. Natural
Resources Forum, 13(4): 258–267.

FAO. 2015. Mapping the vulnerability of mountain peoples to food insecurity. Rome.
www.fao.org/3/i5175e/i5175e.pdf

FAO. 2016a. The agriculture sectors in the intended nationally determined contributions: Analysis.
Environmental and Natural Resources Management Working Paper 62. Rome.
www.fao.org/3/i5687e/i5687e.pdf

FAO. 2016b. Governing tenure rights to commons. Rome. www.fao.org/3/a-i6381e.pdf

FAO. 2017. The future of food and agriculture: Trends and challenges. Rome.
www.fao.org/3/i6583e/i6583e.pdf

FAO. 2018a. The state of food and agriculture: Migration, agriculture and rural development. Rome.
www.fao.org/3/I9549EN/i9549en.pdf

FAO. 2018b. Progress on water-use efficiency: Global baseline for SDG indicator 6.4.1. Rome, FAO
and UN-Water. www.fao.org/3/CA1588EN/ca1588en.pdf

FAO. 2019. FAO framework on rural extreme poverty: Towards reaching target 1.1 of the Sustainable
Development Goals. Rome. www.fao.org/3/ca4811en/ca4811en.pdf

FAO. 2020a. Unpacking water tenure for improved food security and sustainable development. Land
and Water Discussion Paper 15. Rome. https://doi.org/10.4060/cb1230en

FAO. 2020b. The state of food and agriculture 2020. Overcoming water challenges in agriculture.
Rome. https://doi.org/10.4060/cb1447en

FAO. 2021. The impact of disasters and crises on agriculture and food security: 2021. Rome. https://
doi.org/10.4060/cb3673en

FAO. (forthcoming). Governance assessment guidance. Rome.

FAO, International Fund for Agricultural Development, United Nations Children’s Fund,
World Food Programme & World Health Organization. 2021. The state of food security and
nutrition in the world 2021. Transforming food systems for food security, improved nutrition and
affordable healthy diets for all. Rome, FAO. https://doi.org/10.4060/cb4474en

132 2. SOCIOECONOMIC SETTINGS


Fiondella, F. 2013. A wetter Sahel, but will it last? In: Columbia Climate School International
Research Institute for Climate and Society.
https://iri.columbia.edu/news/a-wetter-sahel-but-will-it-last

Gonzalez Fischer, C. & Garnett, T. 2016. Plates, pyramids and planets. Developments in national
healthy and sustainable dietary guidelines: A state of play assessment. Rome, FAO and Oxford,
University of Oxford. www.fao.org/3/I5640E/i5640e.pdf

Guiomar, N., Godinho, S., Pinto-Correia, T., Almeida, M., Bartolini, F., Bezák, P., Biró, M. et
al. 2018. Typology and distribution of small farms in Europe: Towards a better picture. Land
Use Policy, 75: 784–798.

Heinke, J., Lannerstad, M., Gerten, D., Havlík, P., Herrero, M., Notenbaert, A.M.O., Hoff, H. &
Müller, C. 2020. Water use in global livestock production—Opportunities and constraints for
increasing water productivity. Water Resources Research, 56(12).

Herrero, M., Thornton, P.K., Gerber, P. & Reid, R.S. 2009. Livestock, livelihoods and the
environment: Understanding the trade-offs. Current Opinion in Environmental Sustainability,
1(2): 111–120.

HLPE (High Level Panel of Experts on Food Security and Nutrition). 2013. Investing in small-
holder agriculture for food security. A report by the High Level Panel of Experts on Food Secu-
rity and Nutrition of the Committee on World Food Security. HLPE Report No. 6. Rome, FAO.
www.fao.org/3/i2953e/i2953e.pdf

HLPE. 2016. Sustainable agricultural development for food security and nutrition: What roles for
livestock? A report by the High Level Panel of Experts on Food Security and Nutrition of the
Committee on World Food Security. HLPE Report No. 10. Rome, FAO.
www.fao.org/3/i5795e/i5795e.pdf

HLPE. 2017. 2nd note on critical and emerging issues for food security and nutrition. A note by the
High Level Panel of Experts on Food Security and Nutrition of the Committee on World Food
Security. Rome, FAO. www.fao.org/fileadmin/user_upload/hlpe/hlpe_documents/Criti-
cal-Emerging-Issues-2016/HLPE_Note-to-CFS_Critical-and-Emerging-Issues-2nd-Edi-
tion__27-April-2017_.pdf

IFAD (International Fund for Agricultural Development). 2017. Sending money home: Contrib-
uting to the SDGs. Rome. www.ifad.org/documents/38714170/39135645/Sending+Mon-
ey+Home+-+Contributing+to+the+SDGs%2C+one+family+at+a+time.pdf/c207b5f1-9fef-4
877-9315-75463fccfaa7

IOM (International Organization for Migration). 2019. World migration report 2020. Geneva.
https://publications.iom.int/system/files/pdf/wmr_2020.pdf

ITU (International Telecommunications Union). 2022. ITU. https://www.itu.int

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 133
Jayne, T.S., Muyanga, M., Wineman, A., Ghebru, H., Stevens, C., Stickler, M., Chapoto, A.,
Anseeuw, W., Westhuizen, D. van der & Nyange, D. 2019. Are medium-scale farms driving
agricultural transformation in sub-Saharan Africa? Agricultural Economics, 50(S1): 75–95.

Kansiime, M.K., van Asten, P. & Sneyers, K. 2018. Farm diversity and resource use efficiency:
Targeting agricultural policy interventions in East Africa farming systems. NJAS: Wageningen
Journal of Life Sciences, 85(1): 32–41.

Kuper, M., Faysse, N., Hammani, A., Hartani, M., Marlet, S., Hamamouche, M.F. & Ameur,
F. 2016. Liberation or anarchy? The janus nature of groundwater use on North Africa’s new
irrigation frontiers. In: A.J. Jakeman, O. Barreteua, R.J. Hunt, J.-D. Rinaudo & A. Ross, eds.
Integrated groundwater management concepts: Approaches and challenges, pp. 583–615.

Lakner, C., Yonzan, N., Gerszon Mahler, D., Castaneda Aguilar, R.A. & Wu, H. 2021. Updated
estimates of the impact of COVID-19 on global poverty: Looking back at 2020 and the
outlook for 2021. In: World Bank. Washington, DC. Cited 17 September 2021. https://blogs.
worldbank.org/opendata/updated-estimates-impact-covid-19-global-poverty-look-
ing-back-2020-and-outlook-2021

Lawali, S. & Idrissa, I. 2015. Roles and impacts of ICT in the reduction of vulnerability in rural
households in Niger: The case of Maradi. ICTs for Agriculture Series. CTA Working Paper 15/10.
https://cgspace.cgiar.org/bitstream/handle/10568/89993/1873_PDF.pdf

López-Gunn, E., Rica, M. & van Cauwenbergh, N. 2012. Taming the groundwater chaos. In: L.
De Stefano & M. Famón Llamas, eds. Water, agriculture and the environment in Spain: Can we
square the circle?, pp. 227–240. Boca Raton, CRC Press. https://rac.es/ficheros/doc/00935.pdf

Lowder, S.K., Sánchez, M. V & Bertini, R. 2019. Farms, family farms, farmland distribution and
farm labour: What do we know today? FAO Agriculture Development Economics Working Paper
19-08. Rome, FAO. www.fao.org/3/ca7036en/ca7036en.pdf

Lowder, S.K., Sánchez, M. V. & Bertini, R. 2021. Which farms feed the world and has farmland
become more concentrated? World Development, 142: 105455.

Mainet, H. & Edouard, J.-C. 2017. The role of small and medium sized towns in Eastern Africa
mountains: New opportunities or challenged position? In: Rural-urban dynamics in the East
African mountains, pp. 27–45. Dar es Salaam, Mkuki Na Nyota and Nairobi, IFRA Publishers.

Margat, J. & van der Gun, J. 2013. Groundwater around the world: A geographic synopsis. Leiden,
CRC Press/Balkema.
www.un-igrac.org/sites/default/files/resources/files/Groundwater_around_world.pdf

Michel, D. 2020. Water conflict pathways and peacebuilding strategies. Peaceworks, 14.
United States Institute of Peace. www.usip.org/publications/2020/08/water-conflict-path-
ways-and-peacebuilding-strategies

134 2. SOCIOECONOMIC SETTINGS


Molle, F. & Wester, P., eds. 2009. River basin trajectories: Societies, environments and develop-
ment. Comprehensive Assessment of Water Management in Agriculture Series, vol. 8. Wall-
ingford, CAB International. www.iwmi.cgiar.org/Publications/CABI_Publications/CA_CABI_
Series/River_Basin_Trajectories/9781845935382.pdf

Molle, F., Shah, T. & Barker, R. 2003. The groundswell of pumps: Multilevel impacts of a silent
revolution. Paper prepared for the International Commission on Irrigation and Drainage
Asia meeting.

Neufield, L., Hendriks, S. & Hugas, M. 2021. Healthy diet: A definition for the United Nations Food
Systems Summit 2021. A paper from the Scientific Group of the UN Food Systems Summit.
https://sc-fss2021.org/wp-content/uploads/2021/04/Healthy_Diet.pdf

Nolte, K., Chamberlain, W. & Giger, M. 2018. Transactions foncières internationales dans le
domaine de l’agriculture. Nouvelles perspectives offertes par Land Matrix. Bern, Centre for Devel-
opment and Environment, Montpelier, Centre de coopération Internationale en Recherche
Agronomique pour le Développement, Hamburg, German Institute of Global and Area Studies
and Pretoria, University of Pretoria. https://doi.org/10.7892/boris.107739

Oakerson, R.J. 1992. Analyzing the commons: A framework. In: D.W. Bromley, ed. Making the
commons work. Theory, practice and policy, pp. 41–59. San Francisco, Institute for Contempo-
rary Studies.

OECD (Organisation for Economic Co-operation and Development). 2015. OECD principles on
water governance. Paris.
www.oecd.org/cfe/regionaldevelopment/OECD-Principles-on-Water-Governance.pdf

OECD & FAO. 2018. OECD‑FAO agricultural outlook 2018‑2027. Paris and Rome.
https://doi.org/10.1787/agr_outlook-2018-en

Ostrom, E. 1992. Crafting institutions for self-governing irrigation systems. San Francisco, Insti-
tute for Contemporary Studies Press.

Oxfam. 2016. Desterrados: Tierra, poder y desigualdad en América Latina. Oxfam Inter-
national. www.oxfam.org/sites/www.oxfam.org/files/file_attachments/desterra-
dos-full-es-29nov-web_0.pdf

Passarelli, S., Mekonnen, D., Bryan, E. & Ringler, C. 2018. Evaluating the pathways from
small-scale irrigation to dietary diversity: Evidence from Ethiopia and Tanzania. Food Secu-
rity, 10(4): 981–997.

Poláková, J., Farmer, A., Berman, S., Naumann, S., Frelih-Larsen, A. & von Toggenburg, J.
2013. Sustainable management of natural resources with a focus on water and agriculture. Science
and Technology Options Assessment. Brussels, European Union.
https://doi.org/10.2861/27324

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 135
Rafflegeau, S., Losch, B., Daviron, B., Bastide, P., Charmetant, P., Lescot, T., Prades, A. &
Sainte-Beuve, J. 2015. Contributing to production and to international markets. In: Family
farming and the worlds to come, pp. 129–144. Dordrecht, Springer.
https://doi.org/10.1007/978-94-017-9358-2_8

Ricciardi, V., Wane, A., Sidhu, B.S., Godde, C., Solomon, D., McCullough, E., Diekmann, F. et al.
2020. A scoping review of research funding for small-scale farmers in water scarce regions.
Nature Sustainability, 3(10): 836–844.

Robinson, T.P., Wint, G.R.W., Conchedda, G., Van Boeckel, T.P., Ercoli, V., Palamara, E.,
Cinardi, G., D’Aietti, L., Hay, S.I. & Gilbert, M. 2014. Mapping the global distribution of live-
stock. PLoS ONE, 9(5): 13 (e96084).

Samberg, L.H., Gerber, J.S., Ramankutty, N., Herrero, M. & West, P.C. 2016. Subnational
distribution of average farm size and smallholder contributions to global food production.
Environmental Research Letters, 11(12): 124010.

Scoones, I. 2009. Livelihoods perspectives and rural development. The Journal of Peasant Stud-
ies, 36(1): 171–196.

Shah, T. 2009. Taming the anarchy: Groundwater governance in South Asia. Washington, DC,
Resources for the Future and Colombo, International Water Management Institute.

Sourisseau, J.-M., ed. 2015. Family farming and the worlds to come. Dordrecht, Springer.

Steenvorden, J. & Endreny, T., eds. 2004. Wastewater re-use and groundwater quality. IAHS
Publication No. 285. Wallingford, International Association of Hydrological Sciences.

Thapa, G. 2009. Smallholder farming in transforming economies of Asia and the Pacific: Challenges
and opportunities. Discussion paper prepared for the side event organized during the thir-
ty-third session of IFAD’s Governing Council, 18 February 2009. Rome, International Fund
for Agricultural Development.

UNDP (United Nations Development Programme) & EC (European Commission). 2007. Gover-
nance indicators: A user’s guide. New York and Luxembourg.
www.un.org/ruleoflaw/files/Governance%20Indicators_A%20Users%20Guide.pdf

UNFCCC (United Nations Framework Convention on Climate Change). 2020. Progress in the
process to formulate and implement national adaptation plans. FCCC/SBI/2020/INF.13.
https://unfccc.int/sites/default/files/resource/sbi2020_inf13.pdf

United Nations. 2018a. World urbanization prospects: The 2018 revision. ST/ESA/SER.A/420. New
York, Department of Economic and Social Affairs, Population Division.
https://population.un.org/wup/Publications/Files/WUP2018-Report.pdf

136 2. SOCIOECONOMIC SETTINGS


United Nations. 2018b. Sustainable Development Goal 6: Synthesis report 2018 on water and
sanitation. New York. https://www.unwater.org/publications/sdg-6-synthesis-re-
port-2018-on-water-and-sanitation/

United Nations. 2019. World population prospects 2019: Highlights. ST/ESA/SER.A/423. New York,
Department of Economic and Social Affairs, Population Division.
https://population.un.org/wpp/Publications/Files/WPP2019_Highlights.pdf

United Nations. 2020. International migration 2020 highlights. ST/ESA/SER.A/452. New York,
Department of Economic and Social Affairs, Population Division. www.un.org/develop-
ment/desa/pd/sites/www.un.org.development.desa.pd/files/undesa_pd_2020_interna-
tional_migration_highlights.pdf

United Nations ESCAP (Economic and Social Commission for Asia and the Pacific). 2009. What
is good governance?
www.unescap.org/sites/default/d8files/knowledge-products/good-governance.pdf

United Nations OHCHR (Human Rights Office of the High Commissioner) & Hein-
rich-Böll-Stiftung. 2018. The other infrastructure gap: Sustainability. Human Rights and Envi-
ronmental Perspectives. Geneva and Berlin.
www.ohchr.org/Documents/Publications/TheOtherInfrastructureGap_FullLength.pdf

UN-Water. 2021. The United Nations world water development report 2021: Valuing water. United
Nations Educational, Scientific and Cultural Organization.
https://unesdoc.unesco.org/ark:/48223/pf0000375724

Wada, Y., Flörke, M., Hanasaki, N., Eisner, S., Fischer, G., Tramberend, S., Satoh, Y. et al. 2016.
Modeling global water use for the 21st century: The Water Futures and Solutions (WFaS)
initiative and its approaches. Geoscientific Model Development, 9(1): 175–222.

World Bank. 2016. Poverty and shared prosperity 2016: Taking on inequality. Washington, DC.
https://openknowledge.worldbank.org/handle/10986/25078

World Bank. 2020. Poverty and shared prosperity 2020: Reversals of fortune. Washington, DC.
https://openknowledge.worldbank.org/bitstream/handle/10986/34496/9781464816024.pdf

Yoffe, S., Wolf, A.T. & Giordano, M. 2003. Conflict and cooperation over international freshwa-
ter resources: Indicators of basins at risk. Journal of the American Water Resources Association,
39(5): 1109–1126.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 137
RISKS TO LAND AND
WATER RESOURCES
RUN DEEP
Key messages
3
Pressures on land and water systems risk compromising agricultural productivity. This is occurring
precisely at times and in places where growth is most needed to meet global food security targets.
Human-induced land degradation and water scarcity are increasing the risk levels for agricultural
production and ecosystem services.

By 2050, FAO estimates agriculture will need to produce almost 50 percent more food, fibre, livestock feed
and biofuel than in 2012. Agricultural production in sub-Saharan Africa and South Asia will need to at least
double (increase of 112 percent) to meet estimated calorific requirements. The rest of the world will need to
produce at least 30 percent more.

Meeting future demand will require support measures and interventions that complement the sustainable
intensification of agriculture. These include substantially improving productivity along the food value chain,
reducing food loss and waste (FLW), and addressing human dietary health.

Climate change adds uncertainty to the agroclimatic risks that producers are facing, particularly those
who are least able to buffer shocks and are food insecure. Climate volatility and extreme hydrological and
thermal events will affect all producers, but risks are greater in areas with minimal resource endowments,
growing populations and limited economic powers to adapt local food systems or find substitutes.

Increasing competition for land, soil and water for agriculture and food production adds to the pressures
on limited resources. It increases the risks to sustainable agriculture and food production and broader
goals such as zero hunger and eliminating poverty. Annual cereal production growth rates remain below
1 percent. Limits on the global food system must be recognized, and alternative approaches planned and
implemented, to avoid, mitigate or manage risks.

The land degradation risk caused by agricultural production is significant. However, it is rarely considered
until cropland soils and pastures are significantly depleted or lost because of human-induced erosion,
salinization and pollution. Climate change is expected to adversely affect growing conditions for crops and
natural ecosystems in subtropical developing countries. In contrast, warming in temperate latitudes could
extend growing seasons for some cereals. Sustainable land and water management across all agroclimatic
zones will become a priority if GHG emissions are to be controlled and food production increased.
©FAO/Yasuyoshi Chiba
Water scarcity increases agricultural production risks as water availability, storage and conveyance
systems reach their design limits. In many areas of high water stress, farmers manage their production
risks by abstracting shallow groundwater for irrigation, and in some cases, by using non-renewable
groundwater. Competition for diminishing quantities of high-quality groundwater is intensifying as many
aquifers suffer from overabstraction and saline intrusion plus a combination of agricultural and industrial
pollution.

Water pollution from agriculture is proliferating. New and emerging pollutants and antimicrobial resistance
in the environment are adding to clean-up costs and challenging technological and management solutions
on land, water, and lacustrine and nearshore marine environments. Although plastic pollution is primarily
land based, a significant amount of it travels via rivers and threatens marine life and human health.

The operational question for agriculture is complex. The sector needs to review if the overall risk to food
production can be avoided or mitigated by changing agricultural and land management practices while
reducing impacts on livelihoods, human health and ecosystem services.
freshwater scarcity, including groundwater
depletion and deteriorating water quality.
The purpose is to assist in identifying the
©FAO/Giulio Napolitano

preventive measures necessary to protect


stable systems and the proactive measures
required to reverse negative trends and move
towards sustainable food and agricultural
production as discussed in Chapters 4

3.1 Introduction
and 5. The central concern is the risk to
agricultural production presented by internal
factors (e.g. the environmental impact of

The risk to agricultural production is rooted agricultural practice on land and water) and

in the state of land, soil and water resources, external factors (including the intersectoral

the effects of human use and their systemic competition for land and water resources).

interaction with the climate. The climatic


The main socioeconomic drivers of demand
regime in each GAEZ primarily sets the risk
for land and water resources are population
and determines the specific frequency and
growth, rapid urbanization, increased mobil-
magnitude of temperature and precipitation
ity, and the effects of increased income level
events. Competition for land and access to
dietary changes – leading to higher demand
water compound the natural resource risk
for meat, dairy products and ultraprocessed
as freshwater is depleted and soil resources
foods. However, poverty, unplanned urban
degraded. These affect impoverished commu-
expansion, civil strife, migration and inse-
nities, whose food security and livelihoods
curity of tenure lead to localized pressures on
depend directly upon land and water (Chapter
resources as current agricultural production
2). The scale and intensity of current land and
systems adapt, and, in some cases, result in
water use for agriculture are not sustainable
unsustainable resource management prac-
at the global level (Gerten et al., 2020). In
tices. Over the past decade, land-use changes
addition, climate change introduces uncer-
and pressures on productive land and water
tainty and takes the agroclimatic context into
systems have steadily increased to the point
ranges beyond standard probabilities. Climate
where some systems can no longer provide
change projections illustrate how tempera-
or maintain former ecological function levels
ture changes can exacerbate production risks
and agricultural production. Some evidence
by extending the magnitude of climate events
is immediately visible, such as erosion in
beyond the normal or historical distributions
deforested land, but many issues are less
of agroclimatic regimes (Kummu et al., 2021).
apparent. One example is increased water and

This chapter assesses land, soil and water soil pollution levels from agricultural land,

resources as systems under human pressure leading to land salinization and eutrophica-

and risks to future agricultural production tion of surface waters.

as the demand for limited natural resources


Climate change is generating additional
intensifies. The previous edition of this report
pressures, compounding agroclimatic risks
(SOLAW 2011) identified a wide range of risks
associated with agricultural systems and
to production, but this chapter now focuses
introducing a level of uncertainty at a global
on those that are most prominent: climate
scale, with faster warming translating into
change, land and soil degradation, and overall
specific impacts across all regions (IPCC,

140 3. RISKS TO LAND AND WATER RESOURCES RUN DEEP


2021). Rising temperatures and changes in
the hydrological cycle amplify the frequency
and severity of extreme flood and drought

© Soliman Ahmed
events. There is evidence that hydrologic
regimes and weather systems are causing
significant shifts in agricultural production
zones and cropping patterns. Agricultural
practices, such as draining organic soils, are
ing declining structure and fertility. These
accelerating GHG emissions, and the shifting
changes are less apparent than the immedi-
seasonal availability of local water resources
ately visible impacts of mechanical erosion
affects rural livelihoods, particularly those of from rainfall, runoff and aeolian processes,
smallholder families with no access to water but are nevertheless significant.
storage or irrigation services.
It has proved difficult to classify, with any
This chapter offers a brief analysis of risks degree of consistency, the extent and sever-
generated by current patterns and practices ity of human-induced land degradation in
of land and water management. Unlike total the past (Gibbs and Salmon, 2015). But the
crop water requirements, which are predict- compilation of contemporary datasets illus-
able to 2050 with and without climate change, trates converging evidence to set a baseline
it is not yet possible to predict the impacts of for 2010–2012 (Coppus, 2022; see Chapter 1).
climate change on land degradation. Thus,
In 2018, IPBES published a thematic
land degradation risk can be assessed only
assessment of land degradation and
in the broader multi-index approach used
restoration to establish the effect of
in Chapter 1. These issues set the agenda for
degradation on biodiversity, ecosystem
the policy and management responses in
services and human welfare (IPBES, 2018).
Chapters 4 and 5.
The findings were elaborated in the 2019
IPCC special report on climate change and

3.2 Looking into land (IPCC, 2019). Both reports found land
degradation affects people and ecosystems

the future in all regions. Patterns of land use and


management are estimated to produce almost
23 percent of global GHG emissions (IPCC,
3.2.1 Climate change 2019). Taking into account post-harvest

and land degradation processing, storage and transport increased


the overall contribution of the global food
The risks to agricultural production from system to 35 percent in 2015 (Muntean et
the impacts of climate change (drought, al., 2018). Agriculture absorbs 26 percent of
rainfall events and temperature extremes) the economic impact of climate disasters,
are already being experienced across rainfed rising to 83 percent for drought in developing
agricultural land. Historically, irrigation has countries. With climate change expected
been the prime adaptation to variations in to push 122 million more people, mainly
climate, deployed when soil moisture defi- farmers, into extreme poverty by 2030,
cits in rainfed land have become intolerable. leveraging the adaptation and mitigation
However, within the current distribution of potential simultaneously across agricultural
rainfed and irrigated land (Chapter 1), land landscapes is vital in meeting emissions
degradation processes have intensified to the reduction targets.
point where major soil groups are exhibit-

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 141
fire regimes and produce changes in vegeta-
tion and soil properties that ultimately affect
©FAO/Asim Hafeez

biodiversity, carbon stocks, albedo and fire–


atmosphere–vegetation feedbacks, among
other impacts (IPCC, 2019).

The impacts of climate change on land degra-


dation are difficult to separate from other
Climate change and land degradation are impacts given the diversity of social, economic
interrelated. Extremes of heat and high rain- and political settings, but there is little doubt
fall intensities exacerbate land degradation. about the risk to production. As risk multi-
Land degradation processes accelerate GHG pliers, climate change and land degradation
emissions and reduce carbon sequestration in will interact to affect poverty, food security,
ecosystems. Although carbon cycle stability conflict and migration, with the main burden
is crucial for soil fertility and biodiversity, falling on communities whose access to land
estimates suggest that between 20 and 60 and water is limited or excluded.
percent of carbon stocks, historically stored
in SOM as active components, has been lost
since soil cultivation began (IPCC, 2019). This
3.2.2 Future of food
is attributed to changes in land use and land and agriculture
management practices (see Chapter 1). foresight scenarios
Climate change will affect the rate and Anticipating the future of agriculture is a
magnitude of some degradation processes central concern for FAO Members, particu-
and introduce new degradation patterns larly those experiencing high levels of food
(IPCC, 2019). Climate models predict increas- insecurity. The FAO global outlook exercise
ing frequency, intensity and amount of heavy “The future of food and agriculture: Trends
rainfall, raising the risk of landslides, erosion and challenges” (FOFA) (FAO, 2018) produced
events and flash floods. Tropical cyclones are a single projection of the future but without
already shifting towards the poles, and the explicit consideration of climate change or
speed at which they move is slowing. Increas- possible mitigation pathways. To test possi-
ing exposure of coastal areas to intense ble climate futures and the implications for
and long-duration storms will lead to land land and water resource availability under
degradation and affect coastal forest struc- rising demand for food and fibre, FAO devel-
ture, composition and resilience. Sea-level oped three climate-based scenarios (Bb
ox 3.1)
rise will increase coastal erosion and saline based on a range of assumptions about the
intrusion, leaving coastal areas vulnerable to future to 2050 (FAO, 2018). They build upon
catastrophic weather events. the Shared Socioeconomic Pathways (SSPs)
established for the IPCC fifth assessment
Heat stress from rising global temperatures report (IPCC, 2014; O’Neill et al., 2017).
is already affecting agricultural productivity,
the suitability of some areas for commod- Two economic models provide quantitative
ity crops (see Chapter 4, section 4.2.7) and projections for the scenarios: the FAO Global
livestock rearing. Prolonged droughts reduce Agriculture Perspectives System, which
vegetation cover and make soil prone to focuses on the relationships between produc-
erosion, nutrient depletion and biodiver- tion and consumption of food and agricultural
sity loss. Heat stress is also altering wildfire goods, and food security and nutrition, and
regimes in many parts of the world. Some the Environmental Impact and Sustainability
ecosystems have already adapted to specific Applied General Equilibrium model.

142 3. RISKS TO LAND AND WATER RESOURCES RUN DEEP


Box 3.1
FAO FUTURE OF FOOD AND AGRICULTURE scenarios from a land-and water-use
perspective

Business as usual (BAU): Climate futures, Representative Concentration Pathway


(RCP) 6.0 and SSPs 2/3 (“middle of the road”)
Arable land (the physical area under temporary and permanent agricultural crops) expands at faster
annual rates than in the last decades, and land degradation is only partially addressed. Land intensity,
the quantity of land per unit of output, decreases as crop and animal yields increase, but these
achievements require the progressive use of chemicals. Deforestation and unsustainable raw material
extraction continue while water efficiency improves, but the lack of significant changes in technology
leads to the emergence of more water-stressed countries.

Towards sustainability (TSS): Climate futures, RCP 4.5 and SSP 1 (“the green road”)
Low-input processes lead water intensity to decrease substantially and energy intensity to improve
substantially against the levels seen under the BAU scenario. Land-use intensity, the quantity of land
per unit of output, drops compared to current levels, thanks to sustainable agricultural intensification
and other practices to improve resource efficiency. This helps to preserve soil quality and restore
degraded and eroded land. Agricultural land is no longer substantially expanded, and land degradation
is addressed. Water abstraction is limited to a smaller fraction of available water resources.

Stratified societies (SSS): Climate futures, RCP 8.5 and SSP 4 (“a road divided”)
The world suffers further deforestation. New agricultural land is used to compensate for increased
degradation and satisfy additional agricultural demand, which is left unmanaged. The quantity of land
per unit of output decreases for commercial agriculture but remains stable or increases for family
farmers, who increasingly suffer from crop losses fuelled by extreme climate events. Water use is not
sustainable in many regions, and there is little investment towards water-use efficiency. Climate change
exacerbates water and land constraints.

Notes
Harvested areas and yield differentials for each cropping system (irrigated and rainfed)
Data on harvested areas are used to calculate the shares of irrigated and rainfed production systems
by crop and yield differentials between the two systems in the base year. The FAO and the International
Institute for Applied Systems Analysis (IIASA) GAEZ data portal includes geospatial datasets consistent
with country-level FAOSTAT data on harvested areas, yields and crop production. These are derived
by disaggregating (“downscaling”) country-level FAOSTAT production data for the period 2009–2011
to pixel level using an iterative rebalancing approach that ensures matching country totals. The
assignment of crops and crop systems to each pixel is based on FAO GLC-SHARE (Latham et al., 2014),
which provides high-resolution land-cover data, geospatial data on land equipped for irrigation (GMIA,
available at https://www.fao.org/aquastat/en/geospatial-information/global-maps-irrigated-areas
(Siebert et al., 2013)) and other datasets.

Land areas
Data on land cover are used to estimate the amount of suitable land available in the future under
alternative climate scenarios. The GAEZ data portal includes pixel-level data on protected areas, based
on a recent version of the World Database of Protected Areas (available at https://www.unep-wcmc.
org/resources-and-data/wdpa) a comprehensive global dataset of marine and terrestrial protected
areas that includes those under IUCN, such as nature reserves and national parks, protected areas with
an international designation status, such as World Heritage and Ramsar Wetland areas, and those with
national protection status. The land-suitability assessment does not account for land productivity changing
over time due to natural or human-induced degradation and may overestimate potential land availability.
Source: Adapted from FAO. 2018. The future of food and agriculture: Alternative pathways to 2050.
Summary version. Rome. www.fao.org/3/CA1553EN/ca1553en.pdf

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 143
©FAO/Giulio Napolitano intensification is factored in through crop-
ping intensity assumptions, but conjectural
projections such as land required to substi-
tute animal protein with plant-based protein
are not included.

In 2017, FAO projected the global area


equipped for irrigation might increase at a
3.2.3 Implications low annual growth rate of 0.1 percent, reach-
of the scenarios for ing 337 million ha by 2050 (FAO, 2017). In

land and water 2012, the area equipped for irrigation was
323.3 million ha, and in 2018, the reported
The FOFA foresight scenarios for cropland area had reached 328.3 million ha, indicat-
(arable land and land under permanent crops) ing annual growth rates of the order of 0.3
apply a set of technical improvements (yield percent. The FOFA foresight BAU scenario
growth and cropping intensities) and climate expects the global land area equipped for irri-
change drivers to arrive at harvested areas gation to expand to 498 million ha by 2050
of crop production to satisfy food balance (FAO, 2018), indicating an annual growth rate
sheets in 2030 and 2050. The projections for of only 0.14 percent. This represents a signif-
harvested areas on rainfed and irrigated land icant slowdown, compared to that for 1961 to
generate demand for land and water resources 2009, when the global area under irrigation
under the three FOFA scenarios (FAO, 2018). grew at an annual rate of 1.6 percent and
Under the BAU scenario, irrigated areas would more than 2 percent in the poorest countries.
need to increase their contribution to total Most expansion of irrigated land is likely to
production value from 42 percent in 2012 to take place in low-income countries.
46 percent by 2050 (FAO, 2018).
The water resource implications for this
When harvested area projections for irri- growth in irrigated harvested areas were
gated and rainfed production are converted modelled with an FAO global water balance
to arable land requirements, globally the model, GlobWat (Hoogeveen et al., 2015),
cultivated area under the BAU scenario would for the three FOFA climate change scenarios
need to grow from 1 567 million ha in 2012 to (FAO, 2018). Keeping the same set of crop-
1 690 million ha by 2030 and 1 732 million ha ping calendars (seasonality), the changes
by 2050 (FAO, 2018). This growth projection in temperature and precipitation under the
is based on expected yield growth and higher respective RCP scenarios drive crop water
cropping intensities required to meet the requirements on irrigated land purely through
anticipated demand in 2050. incremental evapotranspiration due to the
import of irrigation water into each irrigated
While it is possible to project land suitability cell in the model. In addition, the specific
under climate change scenarios (Chapter 4, water requirements for land preparation in
section 4.2.7), it is not possible to predict paddy irrigated areas are held constant since
how production scenario projections will be residual soil moisture that is not evaporated
distributed in detail across agroecological is assumed to drain to groundwater.
zones. Some expansion to new agricultural
land can be expected, together with conver- From a 2012 baseline used in the FOFA analysis
sion from non-agricultural uses, preparation (FAO, 2018) in which some 407 million ha of
of fallow land and restoration of abandoned irrigated land was harvested (on an equipped
land through consolidation/land bank- area of approximately 305 million ha), the
ing (FAO, 2017). Land substitution through

144 3. RISKS TO LAND AND WATER RESOURCES RUN DEEP


growth in harvested areas was calculated 12–16 percent by 2050. When taken with
for the three scenarios with and without the additional water volumes required for
climate change. The model indicates that land preparation and maintenance of rice
total crop water requirements (incremental paddy, total crop water requirements are
evapotranspiration due to irrigation plus even higher (Ttable 3.1).
land preparation, leaching requirements and
maintenance of rice paddy) would increase The expected agricultural future hinges on
from 1 507 km /year in 2012 to almost 1
3 the continued availability of suitable land,
761 km3/year by 2050 without considering soil and water resources. However, the risks
climate change and to almost 1 952 km /year
3 of failure of current production patterns
with climate change under the BAU scenario are immediate. Thresholds are already
(Ttable 3.1). The spread between the with exceeded for some land and water systems,
climate scenarios is not that wide, from 1 816 where land degradation and water scarcity
km /year under “sustainability” assumptions
3 combine to affect food security and associ-
to 1 993 km3/year under “stratified society” ated livelihoods (Gerten et al., 2020). Abrupt
assumptions. socioeconomic transitions are taking place
in degraded landscapes that need targeted
Taking account of conveyance losses from policy and management interventions.
the point of withdrawal to the point of
consumption, the BAU assumptions would Climate change increases drought risk by
push annual gross agricultural withdrawals increasing the frequency and magnitude of
from 2 673 km /year in 2012 towards 3 500
3 extreme weather events. It changes the aver-
km /year in 2050 on the basis of current
3 age climate conditions and climate variability
crop water requirement to withdrawal ratios and generates new threats in regions with
(FAO, 2021a) assumed at a global average of little experience of dealing with drought.
0.56. In general, these withdrawal ratios can Droughts develop slowly and are not easily
be expected to improve (i.e. become more recognized at first, but can have deep and
“efficient”) as the proportion of pressurized widespread impacts on societies, ecosystems
irrigation systems increases with the adop- and economies (UNDRR, 2021) (Chapter 4).
tion of more precision agriculture. Section 3.4
examines the risks associated with this level Rainfed systems
of withdrawal and consequent soil/water
Seasonal rains and temperature progressions
pollution impacts in more detail.
have immediate impacts on rainfed farming
systems (Map 3.1). The drought frequency
3.2.4 Scenario areas in main cereal-producing regions, such as
of concern central United States of America, the Punjab
state in India, Ukraine and eastern Austra-
The foresight scenarios set out possible
lia, is cause for concern, particularly where
food and agricultural production futures in
irrigation is not an option. The FAO FOFA
broad macroeconomic terms (FAO, 2017) and
explore alternative pathways (FAO, 2018). scenarios, which exclude protected areas

They indicate arable land availability and (615 million ha) and land-cover classes
expected crop yield growth and cropping used for other purposes, limit suitable areas
intensities under the climate scenarios. The for rainfed crop production expansion to
increases in evapotranspiration with the approximately 400 million ha. More than
climate futures factored in are striking. For two-thirds of this suitable land is in low- and
all scenarios, the increase in water consump- middle-income countries, half in sub-Saharan
tion due to climate change above the 2012 Africa (29 percent) and Latin American coun-
baseline reaches 8–9 percent by 2030 and tries (21 percent).

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 145
146
Table 3.1 Anticipated increases in crop water requirements induced by demand and climate forcing to 2050
Baseline BAU TSS SSS

Total Total Total Total Total Total


ET Total
Region Area CWR CWR Δ CWR CWR Δ CWR CWR Δ
irrigation CWR Δ Δ Δ
irrigated 2050 2050 cc cc 2050 2050 cc cc 2050 2050 cc cc
(million (million (%) (%) (%)
(ha) (million (million (%) (million (million (%) (million (million (%)
m3) m3)
m3) m3) m3) m3) m3) m3)

Northern Africa 6 336 576 61 750 62 969 60 682 66 194 −4 5 50 404 55 396 −20 −12 61 099 70 089 −3 11

Sudano−Sahel 2 607 514 17 297 19 193 19 888 23 127 4 20 21 407 25 147 12 31 19 612 23 736 2 24

Gulf of Guinea 585 938 2 303 3 015 5 078 4 851 68 61 6 688 6 359 122 111 4 888 4 918 62 63

Central Africa 121 634 221 319 488 497 53 56 635 647 99 103 479 498 50 56

Eastern Africa 619 385 3 367 3 603 6 211 5 640 72 57 8 332 7 452 131 107 5 869 5 378 63 49

Southern Africa 2 060 468 8 871 9 046 17 530 19 804 94 119 15 643 17 938 73 98 16 795 20 525 86 127

Indian Ocean
1 074 645 2 233 4 046 6 507 7 069 61 75 8 263 9 287 104 130 6 374 7 096 58 75
islands

Northern America 29 470 830 99 007 101 101 140 448 131 006 39 30 132 546 124 269 31 23 168 901 168 774 67 67

Mexico 6 778 510 29 260 29 306 44 815 44 249 53 51 44 063 44 065 50 50 42 557 44 567 45 52

Central America 509 032 4 525 4 742 6 455 6 906 36 46 7 610 8 162 60 72 6 056 6 654 28 40

Antilles 1 295 871 2 503 3 565 4 663 4 991 31 40 4 276 4 774 20 34 4 374 5 021 23 41

Guyanas 190 128 428 708 859 1 067 21 51 971 1 201 37 70 843 1 020 19 44

Andes 4 330 560 22 938 26 039 37 786 37 090 45 42 44 399 43 186 71 66 36 084 36 544 39 40

Brazil 4 431 697 27 568 30 336 40 405 47 277 33 56 47 588 55 025 57 81 37 735 49 810 24 64

Southern America 4 022 088 23 845 24 912 38 109 44 722 53 80 43 578 52 085 75 109 38 671 48 983 55 97

Arabian Peninsula 2 458 827 17 455 17 455 20 257 20 796 16 19 17 771 18 294 2 5 21 127 22 337 21 28

Caucasus 2 130 901 3 512 3 516 4 457 5 487 27 56 3 802 4 791 8 36 4 269 5 703 21 62

3. RISKS TO LAND AND WATER RESOURCES RUN DEEP


Table 3.1 (CONTINUED)
Anticipated increases in crop water requirements induced by demand and climate forcing to 2050
Baseline BAU TSS SSS

Total Total Total Total Total Total


ET Total
Region Area CWR CWR Δ CWR CWR Δ CWR CWR Δ
irrigation CWR Δ Δ Δ
irrigated 2050 2050 cc cc 2050 2050 cc cc 2050 2050 cc cc
(million (million (%) (%) (%)
(ha) (million (million (%) (million (million (%) (million (million (%)
m3) m3)
m3) m3) m3) m3) m3) m3)

Iran (Islamic
8 823 642 46 738 47 849 46 959 49 479 −2 3 38 711 41 214 −19 −14 46 959 51 944 −2 9
Republic of)

Near East 10 568 152 44 242 44 623 53 323 61 705 19 38 50 292 59 078 13 32 53 381 65 994 20 48

Central Asia 13 677 878 55 664 56 451 56 439 62 246 0 10 52 389 58 631 −7 4 55 602 65 032 −2 15

South Asia 85 245 570 528 886 617 833 687 800 801 483 11 30 595 247 677 020 −4 10 656 718 762 395 6 23

East Asia 66 392 461 169 812 235 821 278 689 308 548 18 31 263 193 300 516 12 27 279 428 318 185 18 35

Mainland Southeast
13 672 858 42 456 70 500 76 625 82 045 9 16 81 433 86 456 16 23 74 106 80 461 5 14
Asia

Maritime Southeast
8 564 294 24 932 44 065 48 982 46 812 11 6 54 612 52 062 24 18 47 568 45 636 8 4
Asia

Northern Europe 837 947 504 504 615 823 22 63 580 805 15 60 890 1 273 77 153

Western Europe 4 300 929 3 765 3 808 5 202 6 920 37 82 4 855 6 561 28 72 6 308 9 095 66 139

Central Europe 3 575 931 1 100 1 155 1 409 1 884 22 63 1 386 2 015 20 74 1 578 2 408 37 108

Mediterranean
10 085 452 24 086 24 914 35 223 40 881 41 64 28 091 34 059 13 37 38 331 49 223 54 98
Europe

Russian Federation 2 378 206 2 520 2 912 3 720 4 307 28 48 4 381 5 098 50 75 3 738 4 691 28 61

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021
Eastern Europe 2 812 284 1 193 1 246 1 592 1 978 28 59 1 847 2 362 48 90 1 600 2 192 28 76

Australia and New


4 618 645 11 661 11 856 10 681 12 161 −10 3 10 879 12 197 −8 3 10 941 13 125 −8 11
Zealand

World 304 578 642 1 284 642 1 507 408 1 761 898 1 952 046 17 29 1 645 873 1 816 153 9 20 1 752 882 1 993 308 16 32

Note: cc = climate change; CWR = crop water requirements; ET = evapotranspiration.


Source: Adapted from FAO. 2022. The state of the world’s land and water resources for food and agriculture 2021: Systems at breaking point.

147
SOLAW 2021 background studies. In: Land & Water. Rome. www.fao.org/land-water/solaw2021/en
Map 3.1 Drought frequency on rainfed farming systems, 1984–2018

≤ 10 10-20 20-30 > 30 No data No rainfed cropland

Source: FAO. 2020. The state of


food and agriculture 2020.
Overcoming water challenges
in agriculture. Rome.
https://doi.org/10.4060/cb1447en
Dotted line represents approximately the Line of Control in Jammu and Kashmir agreed upon by India
Modified to comply with to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420 and Pakistan. The final status of Jammu and Kashmir has not yet been agreed upon by the parties.
Final boundary between the Sudan and South Sudan has not yet been determined.

Map 3.1 shows the historical frequency of the drought frequency is greater than 30
severe drought in relation to the distribution percent, particularly where carrying capacity
of rainfed cropland between 1984 and 2018. is frequently exceeded with the consequent
Drought frequency exceeding 30 percent is breakdown in soil structure and loss of soil
considered to amplify production risks. Map through wind and water erosion.
3.2 can be interpreted to indicate the high
level of drought risk on soils subject to over- Irrigation systems
grazing where reduced SOM and increased
Concerns over irrigated areas are significant.
soil compaction combine with pressures on
The areas equipped for irrigation that are most
local groundwater resources to meet irriga-
productive are broad alluvial plains, deltas
tion and livestock watering demands.
and coastal margins in subtropical climates
with high evaporation rates but subject to
Pastoral systems monsoonal rainfall, inundation and suscep-
Pastoral systems also mostly depend on tibility to salinization. In 2012, irrigated areas
seasonal rainfall for forage, even if access to accounted for 42 percent of total produc-
stored water in dams or aquifers mitigates tion value using base-year commodity prices.
the risk of dehydration (Map 3.2). The risk is This reflects higher land productivity (yield),
high where there is low rainfall, soil desicca- greater cropping intensities and higher-value
tion, high temperature and limited or saline crops (FAO, 2018). Irrigated agriculture is
groundwater, as is the case in central Sudan concentrated on just 22 percent of cropland,
and the Horn of Africa. Pastoral systems and, together with hydrological variability,
for dairy and meat products, including the risks from water stress and flood damage
small ruminants, have most to lose when are relatively high (Map 3.3).

148 3. RISKS TO LAND AND WATER RESOURCES RUN DEEP


Map 3.2 Drought risk on pastoral farming systems, 1984–2018

≤ 10 10-20 20-30 > 30 No data No pastureland

Source: FAO. 2020. The state of


food and agriculture 2020.
Overcoming water challenges
in agriculture. Rome.
https://doi.org/10.4060/cb1447en
Dotted line represents approximately the Line of Control in Jammu and Kashmir agreed upon by India
Modified to comply with to comply with UN. 2020. and Pakistan. The final status of Jammu and Kashmir has not yet been agreed upon by the parties.
Final boundary between the Sudan and South Sudan has not yet been determined.
Map of the World. https://www.un.org/geospatial/file/3420

Map 3.3 Levels of water stress on irrigated areas, 2015


Extent (ha) of irrigated cropland by SDG indicator 6.4.2 level of water stress
Extent (ha) of irrigated cropland by SDG indicator 6.4.2 level of water stress
<5 5-10 10-25 25-50 50-100 >100
<5 5-10
>1 000 10-25
>500 25-50
>1 000 50-100
>3 000 >100
>3 000 >3 000
>1 000 >500 >1 000 >3 000 >3 000 >3 000
No irrigated
0 0 0 0 0 0 No irrigatedcropland
0 0 0 0 0 0 cropland

Source: FAO. 2020. The state of food and agriculture 2020. Overcoming water challenges in agriculture. Rome. Dotted line represents approximately the Line of Control in Jammu and Kashmir
agreed upon by India and Pakistan. The final status of Jammu and Kashmir has
https://doi.org/10.4060/cb1447en. Modified to comply with to comply with UN. 2020. Map of the World. not yet been agreed upon by the parties.
https://www.un.org/geospatial/file/3420 Final boundary between the Sudan and
South Sudan has not yet been determined.

Final status of the Abyei area is not yet determined

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 149
The gradual loss of soil structure and fertility,
and salt accumulation, multiply the produc-
tion risks. The land’s ability to recover from

© Giulio Napolitano
early frosts, heat flux and flooding to main-
tain cropping calendars is another crucial
element of the resilience of irrigated farming
systems and food security.

Areas needing urgent attention provision of ecosystem services will continue


The scale of risks to sustain production from to dominate a global debate on the future of
land degradation and water scarcity requires food and agriculture. Agricultural and envi-
urgent attention for improving agricultural ronmental practices will need to continuously
practices and applying nature-based solu- improve to reverse trends and spread benefits
tions (NbSs) to sustain productive systems, where there continues to be unmet demand
specifically on: among vulnerable populations. Generating
environmental benefits through agricultural
ƒƒ Clusters of severely degraded land
practices that also sustain these vital agri-
in coastal zones where depleted aqui-
cultural systems is therefore critical. This
fers supply local and international food
aligns with SDG aspirations and associated
systems. For example, soil and water
targets, specifically SDG targets 15.3 and 6.4,
pressures are high where pollution from
which implicate land and water resources
nutrient overenrichment produces dead
management in generating and spreading
zones, such as in the Mediterranean basin
long-term benefits.
and coastal Southeast Asia.
Land and water policy responses can focus
ƒƒ Broad alluvial plains dedicated to irri-
on protecting and restoring locations where
gated cereal production that lack adequate
land and water resources are degrading and
drainage and where soils have become
limiting agriculture’s contribution to food
saline and sodic, such as in the Indus basin.
production and global efforts to reduce
poverty. Chapters 4 and 5 propose responses
ƒƒ Semi-arid rangelands with limited aqui-
that align with global frameworks for assess-
fer storage supporting agropastoral
ing and monitoring risks.
systems on fragile soils experiencing high
rates of pluvial and aeolian erosion and
deposition, such as in the Sahel and East
Africa uplands.
3.3 Land
ƒƒ Humid uplands experiencing high defor-
degradation risk
estation and soil erosion rates, such as
in Central America and Caribbean, and The FOFA scenarios (section 3.2.2) assume

Southern America regions. that the projected growth in rainfed and irri-
gated harvested areas will be on existing and
Land and water management continually available arable land. However, land degra-
adapts to changing agroclimatic conditions dation is expected to constrain anticipated
and market demand to keep pace with an growth in areas currently identified as at risk
expanding global food system. The central (Ttable 3.2). For this reason, most produc-
concern remains. Maintaining sustainable tive cropland and permanent pastures will
levels of production while avoiding further require soil-conservation measures, and this
damage to the natural resources and the section examines the risk. Chapter 4 (section

150 3. RISKS TO LAND AND WATER RESOURCES RUN DEEP


4.2.7) examines the change in land suit- not at risk. In large parts of the United States
ability under climate change in relation to of America, the main drivers are overgraz-
land planning and management responses. ing and agricultural expansion, while invasive
If there is no action to reduce erosion, by species and deforestation dominate Alaska,
2050, cereal losses are expected to exceed northern Canada, Northern Europe and Sibe-
253 million tonnes (FAO and ITPS, 2015). This ria. In the Asian steppe, significant risks come
is equivalent to removing 1.5 million km of 2
from fire and overgrazing, and in the south
land – equal to the total area of arable land in and southeast of the steppe, increasing popu-
India – from crop production. lation density and deforestation dominate.
Australia faces fire risk, and New Zealand
Estimates of land degradation (Coppus, 2022), has high grazing densities. In Africa, fire and
applied to the GAEZ v4 cropland distribu- overgrazing are common, and grazing and
tion, are expected to constrain anticipated deforestation dominate in the Central America
(modelled) yield growth and harvested areas and Caribbean, and Southern America regions.
where land has been left uncultivated or even
abandoned. Combining status and land degra- The extent and impact of land degradation
dation trends (Chapter 1) indicates areas at cannot be overemphasized. A combination of
risk (Map 3.4 and Ttable 3.2). Regions at risk physical and chemical pressures from culti-
are large contiguous areas with low “status” vation practices can reduce or eliminate soil
and subject to light or strong deterioration. functions and their ability to support sustain-
Regions with substantial degradation and able production. Degraded soils have proved
interspersed high and low status are also at challenging to restore without comprehen-
risk. Stable or improving areas are presently sive land management measures (Chapter 4).

Map 3.4 Regions at risk based on status and trends of land degradation, 2021

Strong human-induced land degradation Light human-induced land degradation Strong deterioration under low pressure

Light deterioration under low pressure Strong human-induced land degradation Light human-induced land degradation

Strong deterioration under low pressure

Source: Coppus, R. 2022.


Global distribution of land
degradation. Thematic Background
Report for SOLAW 2021. Rome, FAO.
www.fao.org/land-water/solaw2021/en
Modified to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 151
Table 3.2 Productive land at risk from land degradation, 2021

Total area Area at risk Area at risk


Land COVER
(million ha) (million ha) (%)

Cropland 1 527 472 31

Rainfed 1 212 322 27

Irrigated 315 151 48

Grassland 1 910 660 35

Forestland 4 335 1 112 26

Source: Coppus, R. 2022. Global distribution of land degradation. Thematic Background Report for SOLAW 2021.
Rome, FAO. www.fao.org/land-water/solaw2021/en

3.3.1 Progressive version of RUSLE, called the RUSLE2015 model


(Panagos et al., 2016), which draws upon the
soil erosion extensive, harmonized datasets amassed by
The Status of the world’s soil resources report the Joint Research Centre (JRC) of the Euro-
(FAO and ITPS, 2015) ranks soil erosion as pean Commission (FAO, 2019).
the most critical risk to agricultural produc-
In collaboration with several worldwide
tion, because the process of erosion implies
research institutes, JRC has established the
significant SOC, soil biodiversity and nutrient
Global Soil Erosion Modelling platform. A
losses in addition to mechanical disturbances.
study assessed global soil erosion using a
However, the slow onset of soil resource
combination of remote sensing, GIS model-
depletion on non-erodible soils poses a more
ling and census data, and estimated the
extensive set of risks to agricultural produc-
amount of soil eroded in 2012 to be 35.9
tivity and the stability of global food systems.
million tonnes/year (Borrelli et al., 2017). The
The effort required for remediation is signifi-
study also assessed the spatial and temporal
cant and possibly goes unrecognized. The
effects of land-use change between 2001 and
anticipated impact of higher-intensity and
2012 and the potential offset of applying
longer-duration rainfall events combined
conservation practices. It indicated a poten-
with extended dry periods is expected to
tial global increase in soil erosion driven by
exacerbate this risk.
cropland expansion, with the most significant
Numerous models are available to provide increases occurring in sub-Saharan Africa,
estimates of erosion at broader scales. These Central America and Caribbean, Southern
are essential for evaluating the extent of America and Southeast Asia. The least devel-
erosion and assessing its importance relative oped economies would experience the high-
to other land degradation processes. However, est soil erosion rates (Map 3.4).
field-based researchers often criticize model-
Overall, JRC found an area-specific annual
ling efforts for their simplified view of the
soil erosion average of 2.8 tonnes/ha for 2001.
complex nature of erosion and its controlling
This increased by 2.5 percent between 2001
factors (Evans and Boardman, 2016). This
and 2012, driven primarily by global land-use
is best illustrated by the erosion modelling
change. Some 6.1 percent of the global land-
efforts in Europe using a Europe-specific

152 3. RISKS TO LAND AND WATER RESOURCES RUN DEEP


© Chris Steele-Perkins/Magnum Photos for FAO Monocultures and the use of fewer varieties
reduce local variety traits, which can lead to
soil biodiversity loss, though the magnitude
is not quantified. The intensive use of inor-
ganic fertilizers and pesticides will affect water
quality and above- and below-ground biodi-
versity. Synthetic nitrogen fertilizers can affect
microbial biomass, and arbuscular mycorrhi-
zal fungal and faunal diversity. Tillage can
reduce soil faunal and bacterial diversity.
mass experiences annual erosion rates above
10 tonnes/ha, which is used to establish the Soil biodiversity loss reduces soil carbon
tolerable soil loss value. Areas exceeding this sequestration, raising the risk of soil erosion,
level are lowest in Oceania (0.8 percent) and compaction and salinization. Adopting
highest in Central America and Caribbean sustainable practices, such as promoting SOM
together with Southern America (both 8.3 accumulation and retention, can enhance
percent). The global annual cropland rate soil biodiversity and improve soil health (de
is 12.7 tonnes/ha, which is 79 times higher Graaff et al., 2019).
than forest (0.16 tonnes/ha) and nearly seven
times higher than other natural vegetation A related concern under climate change is

(1.84 tonnes/ha). These rates are expected the increasing risk of SDSs, which involve a

to accelerate under a combination of more greater rate of aeolian erosion on susceptible

intensive land use, higher rainfall intensities soils, and higher rates and wider dispersion

and extended dry periods. of aeolian deposits.

3.3.2 Potential soil 3.3.3 Soil nutrient loss


organic carbon and Soil nutrient mining is the most common
biodiversity loss form of soil degradation. Adverse impacts
of soil nutrient loss on nutrient cycling and
Agricultural intensification threatens ecosys-
productivity result in less biomass and less
tem functioning and land degradation.
soil cover, thereby exacerbating other soil
Unsustainable farming practices change soil
degradation processes, such as SOM loss,
environmental properties and disturb soil
soil erosion, acidification and the formation
structure, leading to loss of SOM and soil
of hardpans.
organism habitats. Estimates indicate the
annual global potential for SOC sequestration
is 1.45–3.44 million tonnes of carbon (5.3–
12.6 million tonnes CO2-eq) (Lal et al., 2018).
In 2017, this represented 38–91 percent of
global power industry fossil fuel emissions,
© FAO/Christena Dowsett

67–100 percent of global transport fossil


fuel emissions (Muntean et al., 2018) and
9–23 percent of the total global emissions
(53 million tonnes CO2-eq) from all sectors in
that year (UNEP, 2018).

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 153
Low soil nutrient levels may result from poor
3.3.4 Soil pollution
farming practices by households with insuf-
ficient resources. In sub-Saharan Africa, Growth in the use of alternative nutrient
increasing population densities and demand sources such as biosolids, sewage sludge
for land affect nutrient availability in soils. and animal manure can present pollution
The traditional practice of leaving land fallow risks. These organic fertilizers benefit soil
is no longer an option without sufficient health, but are also a source of contaminants,
external nutrient input (Vanlauwe et al., such as trace elements, heavy metals,
2015). Increasing micronutrient depletion pharmaceuticals, microplastics, organic
rates inadvertently occur through increasing contaminants and other toxic substances.
crop yields with nitrogen fertilizer applica- Some are not easily removed during waste
tions. Long-term depletion of micronutrients treatment or pass from livestock to their
presents a slow-onset risk. faeces and manure (Chen et al., 2019).

Human migration has links to soil nutrient Contaminants can also enter the food chain
loss. Soil degradation, including nutrient loss when crops and pastures absorb them from
and other forms of environmental change, has the soil and accumulate them in edible parts of
displaced millions of people (Warner, 2010). plants. They can reduce crop yields and induce
health problems in vulnerable communities
Recent analyses suggest that, in some regions, unable to migrate to uncontaminated areas.
increased annual additions of nitrogen in agri- In China, some 10 million tonnes of crops are
cultural systems cannot occur without causing lost annually because contamination reduces
significant environmental harm. Phosphorus yields or renders crops and food products
additions have exceeded safe boundaries in unmarketable (Wu et al., 2010).
several major agricultural regions (Bijay-Singh
and Craswell, 2021), while nutrient mining In turn, polluted soils affect aquatic ecosys-
still occurs in those areas lacking fertilizer tems. Contaminants leach into groundwater
supply. Irrespective of the application method, and pollute surface water and marine envi-
nitrogen recovery efficiencies rarely exceed 50 ronments. Rainfall, flooding, snowmelt and
percent (Delgado and Follett, 2010). Much irrigation increase the soil water content and
of the unrecovered nitrogen accumulates in encourage runoff and flooding, which trans-
groundwater, wetlands and the atmosphere. port contaminants to nearby wetlands, rivers
This contributes to climate change and is an and lakes, causing eutrophication and eventu-
immediate risk that will require high levels of ally contaminants to reach coastal zones and
mitigation to reduce. the oceans. This reduces water quality, affects
the effective functions of aquatic ecosystems,
washes out soil particles that cause turbidity,
and reduces the depth of watercourses and
reservoirs (FAO and UNEP, 2021).

Managing SOC in urban soils offers an oppor-


tunity to improve soil ecosystem services
within the urban fabric (Jansson, 2013). The
© Eddie Gerald

spread of urban areas is now significant.


Land-use patterns within urban environ-
ments include urban agriculture, forestry and

154 3. RISKS TO LAND AND WATER RESOURCES RUN DEEP


caused a 50 percent decline in SOC (Govern-
ment of Canada, 2003). In Brazil, deforestation
and subsequent cultivation depleted SOC by
60–85 percent (Rezapour and Alipour, 2017).
© Hkun Lat

In Argentina, SOC decreased by 36–53 percent


after a long cropping period. Conservation
practices are needed to reduce further loss and
green infrastructure. Urban soils are subject deterioration of soil quality (Liu et al., 2012).
to strong anthropogenic influences, altering
FAO established the International Network
biogeochemical cycling, particularly carbon
of Black Soils in 2017 as a platform to focus
cycling and accumulation. Urban soils have
attention on black soil global importance
substantial SOC storage potential and may
in supporting food security and climate
accumulate SOC at high rates. Data from 116
change mitigation. This network aims to
cities worldwide showed the total carbon
bring together member countries to provide
content of urban soils is 1.5–3 times higher
a scientific platform to discuss and contrib-
and stored at a greater depth than rural soils
ute to improving management, conservation,
(Vasenev and Kuzyakov, 2018).
mapping and monitoring (FAO, 2021b).

3.3.5 Regional soil Permafrost soils


groups at risk Permafrost soils, which cover 25 percent of
the northern hemisphere and contain high
Black soils levels of SOC, are in danger of thawing. This
Reports have highlighted the crucial role would exacerbate global warming, worsen
black soils play in food security and climate soil erosion and threaten industrial infra-
change mitigation, yet they are sensitive to structure (see Chapter 1).
anthropogenic intervention. They are prone
The constant increase in temperature and
to severe SOC loss, erosion, compaction, sali-
land-use change in permafrost regions could
nization and sodification, and can suffer from
lead to carbon dioxide and methane being
anthropogenic soil acidity (FAO and ITPS,
released into the atmosphere, with poten-
2015; see also Chapter 1). Further pressure on
tially devastating ecological and economic
these soils is anticipated, particularly where
costs. Various scenarios show a possible
changing land use and maladapted manage-
release of 92 million tonnes of carbon (on
ment is leading to a significant decline in SOC
average) by 2100 under the current climate
content in the weak (15 percent), medium (25
warming trajectory (RCP 8.5); this will have
percent) and severely eroded (40 percent)
impacts for centuries. Given the magni-
black soils of the Russian Federation.
tude of carbon stocks and the high release
Studies show that 30 percent of SOC was lost potential, improved analyses that are more
in Ukrainian black soils (Balyuk and Medve- accurate using Earth systems models are
dev, 2012). Chinese black soils experienced an crucial for assessing the physical and biolog-
average annual rate of decline in SOC in the ical processes that control the dynamics of
top 900 mm soil profile of 0.91 percent and permafrost distribution and soil’s thermal
0.48 percent under monocropping systems regimes (Schuur et al., 2015). Conserva-
(Liu et al., 2005). Excessive cultivation and tion policies that transcend administrative
summer fallowing in the Canadian prairie borders will be essential.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 155
3.4 Water scarcity
risk to land
productivity
3.4.1 Changing
© Giulio Napolitano

hydrological baseline
Immediate risks to agricultural production
will persist where surface water is scarce,
and groundwater is exploited intensively (see
Chapter 1). Higher evaporative demand is
Peatlands (organic soils) expected to increase irrigation withdrawals

Although peatlands represent only 3 percent and water stress at the local and basin levels.

of the Earth’s surface, they provide important The FOFA projections (section 3.2.3) indicate

ecosystem services, such as regulating the that by 2050, crop water requirements will

hydrological cycle, conserving biodiversity, increase from the 2012 baseline by 17 percent

providing forest products and recreation, under BAU assumptions and by almost 30

and storing information about past environ- percent with climate forcing an additional

ments. They store significant amounts of 445 km3 of evaporation in existing irrigated

carbon (644 million tonnes of carbon to a 3 m areas when temperature and precipitation

depth). However, these are rapidly lost when changes are combined with the projected

the peatlands are drained for agriculture and increase in harvested areas. This will double

commercial forestry (Hooijer et al., 2010). agricultural withdrawal volumes, assuming


the current ratio of the global average for
The FAO emissions database estimates there crop water requirement to withdrawals.
are 250 thousand km of drained organic soils
2

under cropland and grassland globally, with Countries with high groundwater dependency

total GHG emissions of 0.9 million tonnes will experience greater stress from incremen-

CO2-eq/year in 2010. Significant contribu- tal evaporation (consumption) of 2–5 percent

tions come from Asia (0.44 million tonnes under BAU and of 5 percent in a worst-case

CO2-eq/year) and Europe (0.18 million tonnes (SSS) climate change scenario. This will

CO2-eq/year). Global estimates indicate more significantly affect existing groundwater flow

than 500 thousand km2 of drained peatlands, and storage, and diminish the chances of

including under forests, with carbon dioxide recharge, particularly in arid landscapes.

annual emissions increasing from 1.06 Pg of


Longer-term water scarcity risk will become
carbon dioxide in 1990 to 1.30 Pg of carbon
apparent as the interannual storage in snow-
dioxide in 2008 (FAO and ITPS, 2015).
pack and glaciers diminishes, affecting large,

Preserving, rewetting and managing peat- irrigated plains in all northern-hemisphere

lands sustainably (i.e. paludiculture) may continents, but notably in the western United

offer a practical strategy for maintaining SOC States of America (Lovelace et al., 2020) and

and mitigating global warming (Leifeld and the Indus systems (Yu et al., 2013).

Menichetti, 2018).

156 3. RISKS TO LAND AND WATER RESOURCES RUN DEEP


The magnitude and frequency of flood requirements with and without climate
events affecting agricultural production are change was carried out to identify the level
becoming less predictable as climate change of future risk to irrigated agriculture (Ttable
forces higher intensity, longer duration 3.1). In all SOLAW subregions except North-
and increased frequency of rainfall. Open- ern Africa, Islamic Republic of Iran, Central
channel irrigation systems could become Asia and Australia/New Zealand, crop water
prone to higher levels of flow perturbation requirements show growth for all scenarios
and water control infrastructure failure. with and without climate change (cc). At the
global level, the TSS + cc scenario results in
Sea-level rise will increase the risk of saline a flattened rate of increase compared with
intrusion into coastal aquifers, and attenuated BAU + cc, down from 29 percent under BAU +
crop yield growth can be expected as a result. cc to 20 percent under TSS + cc. The SSS + cc
scenario produces only a marginal increase
The FOFA projected growth in irrigation
in total crop water requirements, from 29
harvested areas is 92 million ha by 2050. As
percent to 32 percent under BAU + cc. The
areas actually irrigated are typically about 15
influence of climate change under each of the
percent less than the areas equipped for irri-
scenarios is striking. Climate change almost
gation (FAO, 2021a), a significant proportion
doubles the crop water requirements for each
of that growth will occur within the extent of
scenario at the global level, largely because
the areas equipped for irrigation (GMIA v5;
of increased evapotranspiration and dimin-
FAO, 2021c).
ished rainfall projected by the climate models

The FOFA irrigation projections for the three on irrigated cropland. The increases may be

scenarios indicate increases in harvested higher or lower for individual subregions.

areas and proportionally more irrigation The change compared with the 2012 baseline

consumption per cubic metre of water under BAU + cc ranges from 4 percent for

withdrawn as yields improve and cropping Maritime Southeast Asia to 153 percent for

intensities increase in line with overall Northern Europe under the SSS + cc scenario.

productivity gains in agriculture (FAO, 2017).


The decline in growth in total crop water
The regional- and country-level picture is
requirements for Northern Africa, Islamic
variable, as increasing water scarcity will
Republic of Iran, Central Asia and Austra-
attenuate growth in some of the main centres
lia/New Zealand is attributed to increased
of irrigation production (section 3.2.3).
water scarcity and reduced growing seasons,
as low rainfalls and high temperatures
3.4.2 Increasing combine to reduce harvested areas.
agricultural withdrawals
and water scarcity
Water scarcity trends (Chapter 1) are based
on the requirements for SDG indicator 6.4.2,
predicated on fixed environmental flow
requirements and aggregated water with-
© Giulio Napolitano

drawals for all sectors. The limits of this


approach notwithstanding (Vanham et al.,
2018), an analysis of projected crop water

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 157
The GlobWat model also offers an analysis 3.5), the combination of water scarcity
of risk based on agricultural water stress induced by irrigation and land degradation
expressed as the ratio of irrigation consump- is a reminder that efforts directed at soil
tion to water withdrawals (Hoogeveen et al., and water conservation across these regions
2015). The model considers water stress to be will need to intensify if the land and water
substantial when the incremental evapora- systems are to remain in play.
tion for irrigation exceeds 10 percent of the
generated water resources in a river basin. A 3.4.3 Diminishing
ratio exceeding 20 percent indicates critical
groundwater availability
stress. In 2012, the cluster of arid zone coun-
tries in Northern Africa, the Near East, the The prospect of reducing abstraction from
Arabian Peninsula, South Asia and Central aquifers to sustainable levels is not prom-
Asia all lay well above the 10 percent limit. ising, particularly given the scale of the
By 2050, incipient scarcity in the Mediterra- projected increase in crop water require-

nean, Sudano-Sahel, Caucasus and East Asia ments under the three FOFA scenarios. All

will fall below the critical limit. signs point to intensifying groundwater use
for irrigation as farmers switch away from
Given that 48 percent of some of the most reduced or regulated surface supplies (Dieter
productive irrigated cropland is at risk (Map et al., 2018). The subsequent risk for users

Map 3.5 Irrigated cropland subject to human-induced land degradation in 2014:


(a) Africa and Western Asia, (b) South Asia, (c) East Asia, (d) Southeast Asia
and (e) PARTS OF AMERICA

a b c

Degraded irrigated cropland

0 250 500 1 000 1 500 2 000


km

Note: Areas with more than 10 percent irrigated cropland cover are shown.
Source: Coppus, R. 2022. Global distribution of land degradation. Thematic Background Report for SOLAW 2021. Rome, FAO.
www.fao.org/land-water/solaw2021/en. Modified to comply with to comply with UN. 2020. Map of the World.
https://www.un.org/geospatial/file/3420

158 3. RISKS TO LAND AND WATER RESOURCES RUN DEEP


of high-quality water for potable supply is Poor water quality limits options to increase
magnified. In India, for example, approxi- groundwater use in many accessible aqui-
mately 90 million rural households depend fers. In China, even under natural conditions,
directly on groundwater irrigation (Shah, only 63 percent of groundwater is potable.
2009). Productive coastal margins present a Arsenic and fluoride are the most common
formidable challenge as upstream inflows are geogenic pollutants in Southeast Asia. Fertil-
curtailed, available land is more intensively izers and pesticides in agriculture are the
cultivated, and urban pollution and saline primary sources of anthropogenic pollution
intrusion threaten groundwater quality. The globally (OECD and FAO, 2020). In Europe,
groundwater account is significantly over- nitrate pollution is the most frequent cause
drawn, leaving little recoverable freshwater of poor groundwater quality, with 23 percent
margin in place. This produces a range of of groundwater bodies exceeding European
drawdown “externalities”, such as ingress of Union standards (Kløve et al., 2017). Pesti-
low-quality groundwater and reduced leak- cides and volatile organic compounds are
age to adjacent aquifers. commonly found across the United States of
America (Toccalino et al., 2014). Pollution due
Some large continental aquifers exploited for to mining (e.g. leakage of acidic leachate) and
irrigation and stock watering cross national urbanization (e.g. wastewater, salinization in
borders. There are more than 350 trans- coastal cities or leakage from urban landfills)
boundary aquifers worldwide that have been are additional global concerns.
delineated and described (UNESCO, 2021;
TWAP, 2022). But much less is known about Human activities and climate change
the impact of groundwater abstraction and are significantly increasing pressure on
pollution across borders, and few interna- groundwater resources. Participatory
tional water agreements refer to conjunctive watershed management and closer attention
management of shared surface water and to the monitoring and assessment of
groundwater (Chapter 5). groundwater resources will be vital to inform
aquifer management and governance.
Current patterns of exploitation present There are some encouraging examples
long-term risks for sustained agricultural of stakeholder participation in agriculture
production where transboundary aquifers leading to reduced irrigation demand and
are decoupled from contemporary recharge, providing an environment for decisive joint
such as in the northwest Saharan, Nubian management (Govardhan Das and Burke,
and Arabian aquifer systems. As the climate 2013; Deines et al., 2019).
changes and affects recharge regimes, it will
not always be possible to distinguish between
renewable and non-renewable groundwater
resources. Evidence from long-term
aquifer monitoring in intensively irrigated
areas indicates that long-term production
©FAO/Benedicte Kurzen/NOOR

risks can be expected to increase in terms


of economically recoverable groundwater
storage and associated groundwater quality
(Shamsudduha et al., 2011; Konikow, 2013;
MacDonald et al., 2016; see also the case study
on understanding how groundwater responds
to climate and anthropogenic abstraction).

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 159
some highly persistent pollutants active for
extended periods (Chapter 1). Although a
source of pollution, agriculture is threat-
ened by poor water quality, particularly from
aquifers used for irrigation. Tainted ground-
water circulation is expected to persist, given
©FAO/Michael Tewelde

that many organic and synthetic pollut-


ants imprint themselves into the fabric of
aquifers. A long-term impact on agricul-
tural productivity and profitability can be
anticipated for crops grown with contami-
nated groundwater, as biosafety provisions,
brought into increasingly sophisticated food
3.4.4 Emerging water markets, catch up with food producers.
quality risks
Contaminants of concern
Freshwater pollution attributable to agricul-
ture is an emerging global crisis with direct Emerging pollutants, or contaminants of

impacts on health, economic development emerging concern, are “new” substances

and food security. Although other anthropo- being used and discharged into the freshwa-

genic activities, such as human settlements ter systems for which there are no regula-

(urbanization) and industries, are major tions in place and little monitoring. Most are

contributors to water quality degradation, organic compounds and are present as phar-

agriculture has become the main source of maceuticals, antibiotics, personal care prod-

pollution in many countries. Agriculture ucts, hormones, food additives, pesticides,

intensification to provide more food to a plasticizers, wood preservatives, laundry

growing population has also increased the detergents, disinfectants, surfactants, flame

use of inputs such as fertilizers, pesticides retardants and other organic compounds. For

and antimicrobials. When not managed example, every year, the Government of the

properly, agricultural practices can increase United States of America receives notices for

pollutant loads (nutrients, salts, sediments, the discharge of more than a thousand new

agrochemicals and pathogens) into ground- chemicals into the environment.

water and surface water, making the water


Wastewater treatment plants, where they
unfit for some other users.
exist, cannot entirely remove all chemical

Nutrients, particularly phosphorus and nitro- and biological contaminants. Countries like

gen, are exported from agricultural activities Iraq, Israel, Mexico or Pakistan primar-

to the environment either through diffuse ily rely on wastewater for irrigation (FAO,

pollution or through emission into the atmo- 2010; Reznik, Dinar and Francesc Hernán-

sphere (i.e. reactive nitrogen). Phosphorus is dez-Sancho, 2019). Several contaminants are

a limiting element for surface water eutro- added to agricultural soils with wastewa-

phication, and nitrogen is a more significant ter, such as trace elements, polychlorinated

threat to the environment, human health and dibenzo-p-dioxins and dibenzofurans, poly-

urban infrastructure. chlorinated biphenyls, chlorinated paraffin


and perfluorinated alkylated substances like
The capacity of receiving freshwater to perfluorooctane sulfonate or perfluoroocta-
dilute pollutants is decreasing rapidly, with

160 3. RISKS TO LAND AND WATER RESOURCES RUN DEEP


noic acid, resulting in the pollution of agri-
cultural soils. Crops treated with wastewater
can absorb contaminants from the soil solu-
tion and accumulate them in above- and
below-ground tissues. The contaminants
thus enter the food chain.

Plastic waste is emerging as a significant

© Claudia Amico
global pollution problem on agriculture land
and in rivers worldwide. Plastics are highly
visible in waterways and oceans, unlike other
pollutants. Estimates from sampled rivers
between 2010 and 2014 indicated freshwater and coastal waterways towards reducing the
systems transported 1.15–2.41 million tonnes large quantities of plastic being produced
of plastic into the oceans. Asia accounted for and disposed of on land. Modelling indicates
67 percent of this (Lebreton et al., 2017). A that reducing the global annual target to less
resolution at the fifth session of the United than 8 million tonnes by 2030 would require
Nations Environment Assembly was endorsed a fundamental transformation in the plastics
to end plastic pollution and forge an interna- economy where end-of-life plastic products
tional legally binding agreement by 2024. are valued rather than discarded as waste.
This would involve a 25–40 percent reduc-
Looking at 192 coastal countries, Jambeck tion in plastic waste and an increase in plastic
et al. (2015) estimated that 275 million tonnes waste management from 6 to 60 percent in
of plastic wasted was produced in 2010, with low-income economies.
between 4.8 and 12.7 million tonnes enter-

3.5 Conclusions
ing the oceans. A recent study indicated this
volume entering the oceans might rise to 53
million tonnes annually by 2030 (Borrelle
et al., 2020). Attention is expected to shift from Land and water systems face significant and
focusing on cleaning up plastic waste in inland interconnected biophysical risks related to
the increasing frequency and magnitude of
agroclimatic events, including droughts and
floods, and the slow onset of human-induced
land and soil degradation.

The immediate risks to global food systems


will remain associated with water scarcity.
Irrespective of the long-term shift of climatic
zones, the impacts of rainfall volatility and
temperature events on rainfed agriculture will
affect the main cereal production centres in
the northern hemisphere. The pattern of irri-
© Rosetta Messori

gated production is expected to remain simi-


lar to the current GMIA (Chapter 1), except for
lateral extension and new development where
deeper groundwater can be exploited.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 161
ners therefore respond to the challenge using
remote-sensing, big data and innovative
analytical methods that are revolutionizing
approaches to resources planning.
© FAO/Christena Dowsett

A converging range of economic driv-


ers and climate variability are affecting the
long-term viability of global food systems.
Climate change alone is expected to result
in at least a 30 percent increase in total crop
water requirements for irrigated production
by 2050. When translated into withdraw-
A less-apparent risk is slow-onset land als for irrigation, an additional ~600 km3
degradation. Soils traditionally used for of freshwater is needed on top of current
continuous crop production, such as black withdrawals of 4 thousand km3 with IRWRs
soils, will experience declining soil health as of only 44 thousand km3.
they reach critical thresholds in soil structure
and water chemistry and in their inherent Global food systems are transforming and
productive capacity. becoming more productive, but largely at
the expense of long-term sustainability of
The current rate of land and water resource land and water systems, which rely on avail-
exploitation for agricultural production is able, yet limited and finite, land and water
compromising the levels of land produc- resources. Unsustainable agricultural inten-
tivity needed to meet long-term dietary sification brings long-term environmental
requirements. The impact of land degrada- and economic challenges that affect the
tion and water scarcity on the productivity integrity and productive capacity of existing
of agricultural land at a specific location is land and water systems and heighten the
not predictable, but outcomes of climate risk to production and the production growth
projections become apparent at regional needed to feed a global population.
aggregated levels.

As climate zones shift, water scarcity and


land quality risk already apparent today
will escalate by the end of the twenty-first
century. Extended growing seasons on
northern temperate soils may alleviate
concerns over limited harvested areas, but
only if accompanied by adequate rainfall to
maintain acceptable levels of soil moisture.
The effects of extending the thermal growing
season by more than 10 days for crops have
been assessed in Europe. Chapter 4 details
projections of climate futures for land suit-
ability as part of the planning process.
©FAO/Truls Brekke

Pressures on land and water systems risk


compromising agricultural productivity in
places where growth is most needed to meet
global food security targets. Resource plan-

162 3. RISKS TO LAND AND WATER RESOURCES RUN DEEP


© FAO/Giuseppe Bizzarri
Case study: Understanding how
groundwater responds to climate
and anthropogenic abstraction
The Indus–Ganges–Brahmaputra–Meghna mega river basin is one of the world’s largest transboundary
aquifer systems and accounts for about 25 percent of global groundwater abstraction. Recharge from
rainfall is delivered from June to September during the summer monsoon. This densely populated area
is home to many thousands of smallholder farmers who rely on abstraction from shallow unconfined
and deep confined aquifers across the basin for irrigated food production. This is thought to be one
of the primary contributors to groundwater storage variability. However, the influence of the current
climate is unknown, and climate change is expected to increase temporal and spatial rain variability
throughout South Asia, thus affecting water resources and groundwater recharge.

The lack of evidence and understanding of the relative influence of climate and abstraction on the
aquifer led to a study of the aquifer’s response to climate (rainfall, and global climate cycles including
the El Niño–Southern Oscillation, Indian Ocean Dipole, North Atlantic Oscillation and Pacific Decadal
Oscillation) and human influence (mainly abstraction for irrigation and rural water supply). The analysis
used observations from 6 753 wells over a period of 30 years (1985–2015) to highlight the variable
patterns of phase lags between multidepth groundwater levels and rainfall depending on the differ-
ent nature of climate and abstraction in various parts of the basin.

Some observations were intuitive, such as the rapid response in shallow groundwater and the relatively
delayed response to the global climate patterns with increasing depth. Variations in influence were
observed across the mega basin. Groundwater abstraction dominated the Indus and Meghna basins,
while rainfall was more influential in the Brahmaputra and Meghna basins. In the Ganges basin, the
influences of rainfall and abstraction were moderate. In the most exploited areas, such as the Indus
basin, groundwater abstraction overwhelmed the hydrological processes. The influence of abstraction
on groundwater levels in the deeper observation wells was stronger than the shallow observation wells.
There was a rapid response in shallow groundwater and relatively delayed responses to climate patterns
with increasing depth, leading to enhanced recharge of shallow unconfined groundwater aquifers.

Overall, the results suggested that groundwater abstraction was the dominant influence in most of the
basin, particularly at the greater aquifer depths, highlighting the importance of understanding multidepth
groundwater dynamics for future groundwater management and policy interventions. Recommenda-
tions included increasing monitoring of deep groundwater levels to enhance understanding of aquifer
performance. In areas of overabstraction, effort priority should focus on regulating withdrawals from
deep aquifers.

Source: Malakar, P., Mukherjee, A., Bhanja, S.N., Ganguly, A.R., Ray, R.K., Zahid, A., Sarkar, S., Saha, D. & Chattopadhyay, S. 2021. Three decades of
depth-dependent groundwater response to climate variability and human regime in the transboundary
Indus-Ganges-Brahmaputra-Meghna mega river basin aquifers. Advances in Water Resources, 149: 103856.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 163
© FAO/Alessandro Penso

Case study: Farmers and water utilities


voluntarily cooperating to reduce nitrate
concentrations in Germany

In Germany in the 1980s, water utilities set up a voluntary scheme with farmers to lower nitrates in
drinking water from 90 mg/litre to 50 mg/litre, in line with government requirements. In some areas
where agriculture was intensive and nitrate-laden soil water was slowly percolating into deeper
groundwater, meeting this threshold would be a long-term challenge given the amount of fertilizer use.

The option to impose restrictions meant high administrative and control efforts to enforce them at a
time when authorities did not have the capacity to do this. In water protection areas, standard ordi-
nances would be required that determine restrictions on fertilizer practices. For example, to conduct
nitrogen balances on plots and farms, limits would be imposed on the amount of fertilizers to be
applied and lock-up periods set when manure application was prohibited. These measures would
require farmers to shift to new, and unproven farming practices that would incur extra costs (e.g. for
labour, machinery and manure storage facilities). Many farmers opposed this approach as the restric-
tions would limit their autonomy. Significant numbers of farmers were therefore unlikely to comply.

In contrast to prescriptive, rigid ordinances, voluntary cooperation allows farmers to take part in
decision-making and to develop site-specific measures supported by agricultural advisers and fund-
ing. The cooperating parties agreed to work for common objectives endorsed in binding agreements,
which had several common characteristics:

ƒƒ voluntary establishment and membership;

ƒƒ legal recognition;

ƒƒ benefits for members only;

ƒƒ free-of-charge advisory services for fertilizer practices;

ƒƒ costs passed onto consumers via water utility charges offset against water abstraction charges; and

ƒƒ payments to farmers for efforts that go beyond good agricultural practices.

164 3. RISKS TO LAND AND WATER RESOURCES RUN DEEP


Assessment studies indicate that cooperation successfully reduced nitrate concentrations in soil and
untreated water. Agriculture and water administrations confirm that targeted advisory services were
key to solving the nitrate problem. In 2021, most cooperation agreements still existed in Bavaria (>200),
North Rhine-Westphalia (~113), Lower Saxony (~112) and Hesse (>70), and continue to have proactive
support from regional governments.

Cooperation is effective when most farmers in a water protection area participate, and particularly
in the areas most at risk. In view of the success, regional governments are now calling to establish
cooperation agreements in other nitrate vulnerable areas.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 165
© Rosetta Messori

Case study: Land and water systems at risk


in the Arab region due to climate change

Owing to its unique and complex geopolitical and socioeconomic settings, the Arab region is facing
land and water management challenges, evolving demographics and pressures on ecosystems.
Climate change is expected to add to this complexity, affecting two-thirds of croplands and half of
livestock areas within the region by 2050, with adverse impacts on freshwater quality and quantity,
food security, rural livelihoods and biodiversity.

The Regional Initiative for the Assessment of Climate Change Impacts on Water Resources and
Socio-Economic Vulnerability in the Arab Region, which includes FAO and the United Nations Economic
and Social Commission for Western Asia, aims to strengthen the science–policy interface by addressing
climate change and sectoral vulnerabilities based on specific regional issues. Scientific methods are
applied together with consultations to enhance access to knowledge, build capacity and strengthen
institutions for climate change assessment in the Arab region. The initiative also provides a common
platform for assessing, addressing and identifying regional climate change challenges, which, in turn,
inform dialogue, priority setting, policy formulation and responses to climate change at the regional level.

The assessment methodology includes regional climate and hydrological modelling to inform
sectoral vulnerability assessments through integrated mapping. Regional climate models better
portray smaller-scale atmospheric processes than global climate models by focusing on specific
geographical domains.

The Regional Initiative for the Assessment of Climate Change Impacts on Water Resources and
Socio-Economic Vulnerability in the Arab Region presents the RCP 8.5 scenario, regarded as a high-
emissions BAU scenario. Regional climate modelling projects a general increase in average temperature
of 1.7–2.6 °C by mid-century (2046–2065) compared to the reference period (1986–2005) (Map A). Higher
temperature increases (> 3 °C) are projected in non-coastal areas, including the Sahara Desert.

166 3. RISKS TO LAND AND WATER RESOURCES RUN DEEP


MAP A Temperature for the period 1986–2005
and projected change in temperature for the period 2046–2065

°C

6 9 12 15 18 21 24 27 30 33 36

°C

1 1,5 2 2,5 3 3,5 4 4,5 5

Source: United Nations Economic and Social Commission for Western Asia. 2017. Arab climate change assessment report –
Main report. E/ESCWA/SDPD/2017/RICCAR/Report. Beirut.
https://www.unescwa.org/sites/default/files/pubs/pdf/riccar-main-report-2017-english_0.pdf
Modified to comply with to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 167
Map B indicates that 67 percent of croplands will be highly vulnerable to climate change by mid-century,
with the remaining areas moderately vulnerable. Hotspots include the croplands of sub-Saharan Africa,
the Horn of Africa and the southwestern Arabian Peninsula. These are largely rainfed and are thus
vulnerable to increasing rainfall variability. The most productive farming systems are irrigated agricul-
ture and dry savanna, and 85–90 percent of their combined areas fall within high vulnerability classes.

MAP B Predicted effects of climate change


on water availability for crops, 2046–2065

Algiers Tunis
! !

Rabat! Beirut! Baghdad


Tripoli ! !
!
Damascus
Cairo !

Amman
!
! Kuwait City

Manama
Doha
!
!
! ! Abu Dhabi
Riyadh ! Muscat

!
Nouakchott
Khartoum
!
Sana'a !

! Djibouti

!
Mogadishu

AGRICULTURE:
Lakes WATER AVAILABLE FOR CROPS
Rivers AGRICULTURE: WATER AVAILABLE FOR CROPS
VULNERABILITY: RCP8.5 MID-CENTURY (2046-2065)
Reservoirs Intermittent rivers
VULNERABILITY:Low Vulnerability
RCP8.5 High Vulnerability
MID-CENTURY (2046-2065)
Area not relevant
Legend
to subsector Major cities
Lakes Rivers !(
Major cities

Source: United Nations Economic and Social Commission for Western Asia. 2017. Arab climate change assessment report –
Main report. E/ESCWA/SDPD/2017/RICCAR/Report. Beirut.
https://www.unescwa.org/sites/default/files/pubs/pdf/riccar-main-report-2017-english_0.pdf
Modified to comply with to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420

168 3. RISKS TO LAND AND WATER RESOURCES RUN DEEP


Water availability for livestock will experience high vulnerability (49 percent of livestock areas). The
impacts are concentrated in the region’s least developed countries, where incomes depend upon
livestock production. Hotspots are dispersed in eastern sub-Saharan Africa, the southwestern Arabian
Peninsula and southern Mauritania (Map C), where estimates suggest that 94 percent of available water
is used for agriculture. High vulnerability will significantly affect rural livelihoods unless strong adapta-
tion strategies are adopted. Options include switching from crop to mixed crop–livestock or livestock
only systems. In dry and semi-dry lands, livestock systems based on grassland grazing will be more
prone to climate shocks compared to mixed systems. Vulnerability can be reduced by adjusting animal
movement cycles, modifying feed compositions and appropriate animal health interventions.

Sources: United Nations Economic and Social Commission for Western Asia. 2017. Arab climate change assessment report –
Main report. E/ESCWA/SDPD/2017/RICCAR/Report. Beirut.
https://www.unescwa.org/sites/default/files/pubs/pdf/riccar-main-report-2017-english_0.pdf; FAO, Deutsche Gesellschaft für Internationale
Zusammenarbeit, the Arab Center for the Studies of Arid Zones and Dry Lands and the United Nations Economic and Social Commission
for Western Asia. 2018. Climate change and adaptation solutions for the green sectors in the Arab region. E/ESCWA/SDPD/2017/RICCAR/
TechnicalReport.2. Beirut.
https://riccar.org/sites/default/files/2020-01/Technical%20Report2_Green%20Sectors_Final.pdf

MAP C Predicted effects of climate change on water availability for


livestock, 2046–2065

Algiers Tunis
! !

Rabat! Beirut! Baghdad


Tripoli ! !
!
Damascus
Cairo !

Amman
!
! Kuwait City

Manama
Doha
!
!
! ! Abu Dhabi
Riyadh ! Muscat

!
Nouakchott
Khartoum
!
Sana'a !

! Djibouti

!
Mogadishu

AGRICULTURE: WATER AVAILABLE FOR LIVESTOCK


Lakes Rivers
VULNERABILITY: RCP8.5 MID-CENTURY (2046-2065)
Reservoirs Low Vulnerability High Vulnerability
Intermittent rivers
Legend
Area not relevant
to subsector
Lakes Rivers
Major
!( cities
Major cities

Source: United Nations Economic and Social Commission for Western Asia. 2017. Arab climate change assessment report –
Main report. E/ESCWA/SDPD/2017/RICCAR/Report. Beirut.
https://www.unescwa.org/sites/default/files/pubs/pdf/riccar-main-report-2017-english_0.pdf
Modified to comply with to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 169
References
Balyuk, S.A. & Medvedev, V.V. 2012. Strategy of balanced use, reproduction and management of soil
resources of Ukraine. Kiev, Agrarian Science.

Bijay-Singh & Craswell, E. 2021. Fertilizers and nitrate pollution of surface and ground water:
An increasingly pervasive global problem. SN Applied Sciences, 3: 518.

Borrelle, S.B., Ringma, J., Law, K.L., Monnahan, C.C., Lebreton, L., McGivern, A., Murphy, E.
et al. 2020. Predicted growth in plastic waste exceeds efforts to mitigate plastic pollution.
Science (New York, N.Y.), 369(6510): 1515–1518.

Borrelli, P., Robinson, D.A., Fleischer, L.R., Lugato, E., Ballabio, C., Alewell, C., Meusburger, K.
et al. 2017. An assessment of the global impact of 21st century land use change on soil ero-
sion. Nature Communications, 8: 2013.

Chen, Z., Zhang, W., Yang, L., Stedtfeld, R.D., Peng, A., Gu, C., Boyd, S.A. & Li, H. 2019. Anti-
biotic resistance genes and bacterial communities in cornfield and pasture soils receiving
swine and dairy manures. Environmental Pollution, 248: 947–957.

Coppus, R. 2022. Global distribution of land degradation. Thematic Background Report for SOLAW
2021. Rome, FAO. www.fao.org/land-water/solaw2021/en

Deines, J.M., Kendall, A.D., Crowley, M.A., Rapp, J., Cardille, J.A. & Hyndman, D.W. 2019. Map-
ping three decades of annual irrigation across the US High Plains Aquifer using Landsat and
Google Earth Engine. Remote Sensing of Environment, 233: 111400.

de Graaff, M.A., Hornslein, N., Throop, H.L., Kardol, P. & van Diepen, L.T.A. 2019. Effects of
agricultural intensification on soil biodiversity and implications for ecosystem functioning:
A meta-analysis. Advances in Agronomy, 155: 1–44.

Delgado, J. & Follett, R. 2010. Advances in nitrogen management for water quality. United States
Department of Agriculture Agricultural Research Service.
www.ars.usda.gov/research/publications/publication/?seqNo115=258447

Dieter, C.A., Maupin, M.A., Caldwell, R.R., Harris, M.A., Ivahnenko, T.I., Lovelace, J.K., Barber,
N.L. & Linsey, K.S. 2018. Estimated use of water in the United States in 2015. United States Geo-
logical Survey Circular 1441, Water Availability and Use Science Program.

Evans, R. & Boardman, J. 2016. A reply to Panagos et al., 2016 (Environmental science & policy
59 (2016) 53–57. Environmental Science & Policy, 60: 63–68.

FAO. 2010. The wealth of waste: The economics of wastewater use in agriculture. Rome.
www.fao.org/3/i1629e/i1629e.pdf

FAO. 2017. The future of food and agriculture: Trends and challenges. Rome.
www.fao.org/3/a-i6583e.pdf

FAO. 2018. The future of food and agriculture: Alternative pathways to 2050. Summary version.
Rome. www.fao.org/3/CA1553EN/ca1553en.pdf

FAO. 2019. Soil erosion: The greatest challenge for sustainable soil management, D. Pennock, C.
Lefèvre, R. Vargas, L. Pennock & M. Sala, eds. Rome. www.fao.org/3/ca4395en/ca4395en.pdf

FAO. 2021a. AQUASTAT – FAO’s Global Information System on Water and Agriculture. In: FAO.
Rome. www.fao.org/aquastat/en

170 3. RISKS TO LAND AND WATER RESOURCES RUN DEEP


FAO. 2021b. International Network of Black Soils. In: Global Soil Partnership. Rome.
www.fao.org/global-soil-partnership/inbs/en

FAO. 2021c. AQUASTAT – FAO’s Global Information System on Water and Agriculture. In: FAO.
Rome. www.fao.org/aquastat/en/geospatial-information/global-maps-irrigated-areas/
latest-version

FAO & ITPS (Intergovernmental Technical Panel on Soils). 2015. Status of the world’s soil re-
sources: Main report. Rome. www.fao.org/3/i5199e/i5199e.pdf

FAO & UNEP (United Nations Environment Programme). 2021. Global assessment of soil pollu-
tion: Report. Rome. https://doi.org/10.4060/cb4894en

Gerten, D., Heck, V., Jägermeyr, J., Bodirsky, B.L., Fetzer, I., Jalava, M., Kummu, M. et al.
2020. Feeding ten billion people is possible within four terrestrial planetary boundaries.
Nature Sustainability, 3(3): 200–208.

Gibbs, H.K. & Salmon, J.M. 2015. Mapping the world’s degraded lands. Applied Geography, 57: 12–21.

Govardhan Das, S.V. & Burke, J. 2013. Smallholders and sustainable wells. A retrospect: Participa-
tory groundwater management in Andhra Pradesh (India). Rome, FAO.
www.fao.org/3/i3320e/i3320e.pdf

Government of Canada. 2003. Agriculture and Agri-food Canada. In: Government of Canada.
https://publications.gc.ca/site/eng/245998/publication.html?wbdisable=true

Hoogeveen, J., Faurès, J.M., Peiser, L., Burke, J. & Van De Giesen, N. 2015. GlobWat – A global
water balance model to assess water use in irrigated agriculture. Hydrology and Earth System
Sciences, 19(9): 3829–3844.

Hooijer, A., Page, S., Canadell, J.G., Silvius, M., Kwadijk, J., Wösten, H. & Jauhiainen, J.
2010. Current and future CO2 emissions from drained peatlands in Southeast Asia. Biogeosci-
ences, 7(5): 1505–1514.

IPBES (Intergovernmental Science-Policy Platform on Biodiversity). 2018. The IPBES assessment


report on land degradation and restoration, L. Montanarella, R. Scholes & A. Brainich, eds. Bonn.

IPCC (Intergovernmental Panel on Climate Change). 2014. Climate change 2014: Synthesis report.
Contribution of Working Groups I, II and III to the Fifth Assessment Report of the Intergovernmen-
tal Panel on Climate Change. Geneva.
www.ipcc.ch/site/assets/uploads/2018/05/SYR_AR5_FINAL_full_wcover.pdf

IPCC. 2019. 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. Geneva. www.ipcc.ch/srccl/

IPCC. 2021. Climate change 2021: The physical science basis. Contribution of Working Group I to the
Sixth Assessment Report of the Intergovernmental Panel on Climate Change, V. Masson-Del-
motte, P. Zhai, A. Pirani, S.L. Connors, C. Péan, S. Berger, N. Caud et al., eds. Cambridge, UK,
Cambridge University Press. www.ipcc.ch/report/ar6/wg1/#TS

Jambeck, J.R., Geyer, R., Wilcox, C., Siegler, T.R., Perryman, M., Andrady, A., Narayan, R.
& Law, K.L. 2015. Plastic waste inputs from land into the ocean. Science, 347(6223): 768–771.

Jansson, Å. 2013. Reaching for a sustainable, resilient urban future using the lens of ecosystem
services. Ecological Economics, 86: 285–291.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 171
Kløve, B., Kvitsand, H.M.L., Pitkänen, T., Gunnarsdottir, M.J., Gaut, S., Gardarsson, S.M., Ros-
si, P.M. & Miettinen, I. 2017. Overview of groundwater sources and water-supply systems,
and associated microbial pollution, in Finland, Norway and Iceland. Hydrogeology Journal,
25(4): 1033–1044.

Konikow, L.F. 2013. Groundwater depletion in the United States (1900–2008). United States Geo-
logical Survey Scientific Investigations Report 2013–5079.
http://pubs.usgs.gov/sir/2013/5079

Kummu, M., Heino, M., Taka, M., Varis, O. & Viviroli, D. 2021. Climate change risks to push
one-third of global food production outside safe climatic space. One Earth, 4(5): 720–729.

Lal, R., Smith, P., Jungkunst, H., Mitsch, W., Lehmann, J., Nair, P.K., McBratney, A. et al.
2018. The carbon sequestration potential of terrestrial ecosystems. Journal of Soil and Water
Conservation, 73(6): 145A–152A.

Latham, J., Cumani, R., Rosati, I. & Bloise, M. 2014. Global Land Cover SHARE (GLC-SHARE) data-
base. Rome. www.fao.org/uploads/media/glc-share-doc.pdf

Lebreton, L.C.M., Van Der Zwet, J., Damsteeg, J.W., Slat, B., Andrady, A. & Reisser, J. 2017. River
plastic emissions to the world’s oceans. Nature Communications, 8: 15611.

Leifeld, J. & Menichetti, L. 2018. The underappreciated potential of peatlands in global climate
change mitigation strategies. Nature Communications, 9(1): 1071.

Liu, X., Liu, J., Xing, B., Herbert, S.J., Meng, K., Han, X. & Zhang, X. 2005. Effects of long-term
continuous cropping, tillage, and fertilization on soil organic carbon and nitrogen of black
soils in China. Communications in Soil Science and Plant Analysis, 36(9–10): 1229–1239.

Liu, X., Burras, C.L., Kravchenko, Y.S., Duran, A., Huffman, T., Morras, H., Studdert, G., Zhang,
X., Cruse, R.M. & Yuan, X. 2012. Overview of Mollisols in the world: Distribution, land use
and management. Canadian Journal of Soil Science, 92(3): 383–402.

Lovelace, J.K., Nielsen, M.G., Read, A.L., Murphy, C.J. & Maupin, M.A. 2020. Estimated ground-
water withdrawals from principal aquifers in the United States, 2015. United States Geological
Survey Circular 144, Version 1.2. https://doi.org/10.3133/cir1464

MacDonald, A.M., Bonsor, H.C., Ahmed, K.M., Burgess, W.G., Basharat, M., Calow, R.C., Dixit,
A. et al. 2016. Groundwater quality and depletion in the Indo-Gangetic Basin mapped from in
situ observations. Nature Geoscience, 9(10): 762–766.

Muntean, M., Guizzardi, D., Crippa, E., Solazzo, M., Olivier & Vignati, J.G.J. 2018. Fossil CO2
emissions of all world countries: 2018 report. JRC Science for Policy Report. Luxembourg, Publi-
cations Office of the European Union. https://doi.org/10.2760/30158

OECD (Organisation for Economic Co-operation and Development) & FAO. 2020. OECD-FAO
agricultural outlook 2020-2029. Rome and Geneva. https://doi.org/10.1787/1112c23b-en

O’Neill, B.C., Kriegler, E., Ebi, K.L., Kemp-Benedict, E., Riahi, K., Rothman, D.S., van Ruijven,
B.J. et al. 2017. The roads ahead: Narratives for shared socioeconomic pathways describing
world futures in the 21st century. Global Environmental Change, 42: 169–180.

Panagos, P., Imeson, A., Meusburger, K., Borrelli, P., Poesen, J. & Alewell, C. 2016. Soil conser-
vation in Europe: Wish or reality? Land Degradation & Development, 27(6): 1547–1551.

Rezapour, S. & Alipour, O. 2017. Degradation of Mollisols quality after deforestation and culti-
vation on a transect with Mediterranean condition. Environmental Earth Sciences, 76(22): 755.

172 3. RISKS TO LAND AND WATER RESOURCES RUN DEEP


Reznik, A., Dinar, A. & Hernández-Sancho, F. 2019. Treated wastewater reuse: An efficient and
sustainable solution for water resource scarcity. Environmental and Resource Economics, 74:
1647–1685.

Schuur, E.A.G., McGuire, A.D., Schädel, C., Grosse, G., Harden, J.W., Hayes, D.J., Hugelius, G.
et al. 2015. Climate change and the permafrost carbon feedback. Nature, 520(7546): 171–179.

Shah, T. 2009. Taming the anarchy: Groundwater governance in South Asia. Washington, DC, Re-
sources for the Future.

Shamsudduha, M., Taylor, R.G., Ahmed, K.M. & Zahid, A. 2011. The impact of intensive ground-
water abstraction on recharge to a shallow regional aquifer system: Evidence from Bangla-
desh. Hydrogeology Journal, 19(4): 901–916.

Siebert, S., Henrich, V., Frenken, K. & Burke, J. 2013. Update of the Digital Global Map of Irrigation
Areas to version 5. Rome, FAO.

Toccalino, P.L., Gilliom, R.J., Lindsey, B.D. & Rupert, M.G. 2014. Pesticides in groundwater of
the United States: Decadal-scale changes, 1993–2011. Groundwater, 52(S1): 112–125.

TWAP (Transboundary Waters Assessment Programme). 2022. Transboundary Waters Assess-


ment Programme. geftwap.org

UNDRR (United Nations Office for Disaster Risk Reduction). 2021. GAR special report on drought
2021. Geneva. www.undrr.org/publication/gar-special-report-drought-2021

UNEP (United Nations Environment Programme). 2018. Emissions gap report 2018. Nairobi.
https://wedocs.unep.org/bitstream/handle/20.500.11822/26895/EGR2018_FullReport_
EN.pdf?sequence=1&isAllowed=y

UNESCO (United Nations Educational, Scientific and Cultural Organization). 2021. ISARM - In-
ternational Shared Aquifer Resources Management. In: ISARM/UNESCO. https://isarm.org

Vanham, D., Hoekstra, A.Y., Wada, Y., Bouraoui, F., de Roo, A., Mekonnen, M.M., van de Bund,
W.J. et al. 2018. Physical water scarcity metrics for monitoring progress towards SDG target
6.4: An evaluation of indicator 6.4.2 “Level of water stress”. Science of the Total Environment,
613–614(February): 218–232.

Vanlauwe, B., Six, J., Sanginga, N. & Adesina, A.A. 2015. Soil fertility decline at the base of rural
poverty in sub-Saharan Africa. Nature Plants, 1: 15101.

Vasenev, V. & Kuzyakov, Y. 2018. Urban soils as hot spots of anthropogenic carbon accumula-
tion: Review of stocks, mechanisms and driving factors. Land Degradation and Development,
29(6): 1607–1622.

Warner, K. 2010. Global environmental change and migration: Governance challenges.


Global Environmental Change, 20(3): 402–413.

Wu, G., Kang, H., Zhang, X., Shao, H., Chu, L. & Ruan, C. 2010. A critical review on the bio-
removal of hazardous heavy metals from contaminated soils: Issues, progress, eco-
environmental concerns and opportunities. Journal of Hazardous Materials, 174(1–3): 1–8.

Yu, W., Yang, Y.-C., Savitsky, A., Alford, D., Brown, C., Wescoat, J., Debowicz, D. & Robin-
son, S. 2013. The Indus Basin of Pakistan: The impacts of climate risks on water and agriculture.
World Bank Group. https://doi.org/10.1596/978-0-8213-9874-6

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 173
Sustainable
resources planning
and management
Key messages
4
Land-use planning and land resources planning (LRP) are essential for managing limited
resources for all agroclimatic zones and in crop, livestock, forest and mixed land-use systems.
They are used to guide sustainable management of land and water resources and anticipate the
challenges that come from population growth and increasing demand. Global assessments
of land, soil, water, biodiversity, climate and ecosystems are now providing data and information.
A wide range of resource planning tools and approaches are available to support decision-
makers, planners and practitioners to take informed actions and promote the scaling out of sustainable and
resilient options.

Lower rainfed crop yields and shifts in land suitability are anticipated in the future, in many regions,
as the climate changes. Innovative tools are now available to support decision-makers in understanding
the extent and location of existing yield gaps17 and to anticipate shifts in areas suitable for different crops
and to identify potential impacts on productivity; complementary options include breeding and selecting
suitable crops, changing land use and switching to crops, including trees and livestock, more suited to the
changing climate. Together, these offer the means of turning opportunities into realistic adaptations to
climate change, local biophysical conditions and socioeconomic circumstances. All are vital elements for
planning a sustainable future.

Reversing the trends in human-induced land degradation will be essential to meet global food
security objectives. Preventing land degradation costs much less than restoration. Yet few
countries have a specific competent environmental judicial body to enforce their national land
protection legislation. Coordinated action and political will are needed to overcome long-
entrenched degrading practices. The concept of land degradation neutrality (LDN) will become fundamental
in planning interventions.

7
Yield gap refers to the difference between actual yields and yields expected under optimum growing conditions for particular soils
and climate.
©FAO/Giuseppe Bizzarri

Most countries need to move from crisis to risk-based management to lessen drought risks and impacts.
Many countries still put drought in the same category of natural hazards as floods and earthquakes. This
wastes valuable resources and does not help to build resilience for future events. A “three-pillar” approach
that requires investment in monitoring and early warning systems, studies to assess vulnerability to drought
and actions to reduce adverse impacts is now being deployed. A proactive drought risk management policy
with strengthened institutional capacities would lead to more robust planning and investment decisions,
with early intervention and mitigation and less costly damage due to drought.
A special focus study at the end of the chapter
is devoted to dryland systems. It describes
the status and trends, risks and threats,
and discusses the responses and manage-
ment pathways for these unique and fragile
landscapes.
©FAO/Antonello Proto

4.2 Sustainable land


resources planning
4.1 Introduction The increasing challenges of population
growth and demands on limited resources
by diverse actors, land degradation, biodi-
This chapter responds to the risks, issues and versity loss and climate change require the
emerging challenges identified in Chapter 3. rational use of resources to sustain and
It focuses on land-use planning that informs enhance productivity and maintain resilient
interventions and behaviour to face the chal- ecosystems.
lenge of climate change impacts on land suit-
ability for agricultural production and the Land-use planning and, more broadly,
threat to land, soil and water resources from LRP are tools for achieving sustainable and
human-induced land degradation. efficient use of resources, considering
biophysical and socioeconomic dimensions.
Innovations in land-use and land resources Land resources planning encompasses land
management and planning are described evaluation and land-use planning. It is
below. New tools are available that enable the systematic assessment of land poten-
policymakers and planners to help prac- tial and alternatives for optimal land use,
titioners tackle resource management and improved economic and social condi-
challenges, make the best use of available tions through participatory processes that
resources, prepare for future climate change, are multisectoral, multistakeholder and
and adapt agricultural resource use to sustain scale dependent. It relies on an iterative
livelihoods and contribute to development process of implementing, refining, adapt-
goals. The latest climate models provide ing and improving land-use systems and
insights into climate change impacts on management practices based on results and
agricultural resource distribution, such as experiences.
changes in productivity and geographic shifts
in crop suitability. This allows the best future Land-suitability assessment provides
use of land resources to be identified for decision-makers with viable land-use
rainfed and irrigated production in terms options based on the biophysical potential
of appropriate agronomic management, of resources and socioeconomic condi-
inputs and water supply systems according to tions. These options support land-use deci-
land/soil potential and water resource avail- sion-making processes in fulfilling the needs
ability. Scenario development offers options of different sectors operating in a landscape
to help reverse human-induced land degra- while optimizing and sustaining resource use.
dation in crop, livestock, aquaculture and Land resources planning plays an important
mixed systems. role in integrating the various sociocultural

176 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


and biophysical elements of landscapes and
land-use dynamics, including responsi-
ble governance of tenure, to ensure stake-
holders are not marginalized (FAO, 2020a).
It provides tools for using land and water
resources most efficiently, and promotes
options to maintain sustainable, productive

©Pixabay/Sasint
landscapes, ecosystems and food systems.
This generates multiple benefits and invest-
ment opportunities for local and national
economies and private/public investors.

The tools and knowledge required vary


4.2.1 Resource planning depending on the scale, purpose and nature
tools and approaches of the planning process (Ziadat, Bunning and
De Pauw, 2017). They also incorporate the
Open information exchange underpins
socioeconomic circumstances of those who
all aspects of natural resources planning,
rely mainly on natural resources for their
management and good governance. Global
livelihoods, notably farmers, pastoralists
assessments of land, soil, water, biodiversity,
and fishers, whose interests are increasingly
climate and ecosystems are now providing
managed through formalized participatory
essential data and information for planning
negotiation processes that have become a
and managing natural resources, inform-
significant element of LRP (Tarrason, Andrian
ing global and national decision-makers and
and Groppo, 2017; FAO, 2022a, 2022b).
practitioners, and increasingly a wide range
of stakeholders who participate in planning Bbox 4.1 describes current LRP planning tools
processes. and approaches, and provides some import-
ant definitions of terms used in this and
Resource planning tools and approaches are
subsequent chapters of this report.
available to support decision-makers, plan-
ners and practitioners, working at different The LRP Toolbox (Ziadat, Bunning and De
decision-making levels to design appropriate Pauw, 2017) contains summary descriptions
policies and plans, take informed actions and and links to a comprehensive list of LRP tools
promote the scaling out of sustainable and and approaches developed by FAO and other
resilient options. The tools and approaches institutions, including:
can help to: (i) identify additional areas
suitable for sustainable agricultural use and ƒƒ Biophysical approaches/tools giving
inform sustainable land-use and food system prominence to biophysical attributes
changes; (ii) create links among actors (climate, soil, terrain and water) and
involved in land, soil and water resources methods that guide users towards suit-
to ensure effective use for agriculture and able land-use options and alternatives to
food production; (iii) locate and assess areas sustain resources quality and quantity and
to enhance productivity to close yield gaps, ecosystem functions and services, based
and increase food and livelihood security; mainly on these attributes and land-use
and (iv) pinpoint areas that are overex- and climate change impacts. Land suit-
ploited, hotspots for immediate restoration, ability and similarity analysis are typical
and bright spots for future investment and examples.
management.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 177
Box 4.1
Innovative tools and approaches for land-use planning
Agroecological zoning (AEZ) and land-suitability analysis, developed by FAO and IIASA, can help to
identify areas for implementing land-use planning and management programmes based on specific
crop/land-use/land management practices. It offers global-, regional- and national-level assessments
of potential agricultural production options considering historical and future climate conditions, soil
and terrain resources, land cover, land protection status and biodiversity, under three distinct levels of
inputs and management for rainfed and irrigated water supply systems. It includes a spatial inventory of
downscaled actual area, yield and production of the main agricultural commodities and the occurrence
and significance of apparent yield gaps.

Sustainable land management (SLM) is defined as “the use of land resources, including soils,
water, animals and plants, for the production of goods to meet changing human needs, while
simultaneously ensuring the long-term productive potential of these resources and the maintenance
of their environmental functions” (FAO, 2022c). It includes a range of complementary measures
(policy, legislation, institutional reform and technologies) that are adapted to the biophysical and
socioeconomic context for the protection, conservation and sustainable use of resources (soil, water
and biodiversity), restoration or rehabilitation of degraded natural resources, and maintenance of the
ecosystem functions and services that support the livelihood and well-being of people. Integrated LRP
tools are needed to enhance the scaling out of SLM options.

Land-use planning is the systematic assessment of land potential and alternatives for optimal land
use and improved economic, environmental and social conditions through participatory processes
involving multisector, multistakeholder and scale-dependent processes (FAO, 1993). Land-use planning
helps decision-makers to adopt appropriate options for the use of land and water resources based on
their natural potential and hence avoid unsustainable exploitation and prevent further degradation.
Proper planning should avoid detrimental land-use change and help land users to select and put SLM
options into practice that support land/soil restoration in already degraded areas and sustain resources
(soil, water and biodiversity) and ecosystem services.

Land resources planning is an overarching approach and set of tools for various land users to plan
and manage land resources. Rather than a top-down process, participatory LRP involves the multiple
sectors and stakeholders concerned in a given land area or territory (from the local community to the
river basin, provincial, national or transboundary level). Land resources planning offers a set of tools
– procedures, guidelines, methods and datasets, covering biophysical, economic, sociocultural and
governance dimensions – that guide the design of implementation plans and decision-making for
SLM and restoration and the delivery of ecosystem services. Land resources planning encompasses
land evaluation and land-use planning, and addresses the biophysical, socioeconomic and negotiatory
domains.

Integrated land-use planning can be used to support transformative change in land use and
management so as to deliver a range of ecosystem services that support human well-being
and livelihoods in line with SDGs. This can help to sustain or improve productivity, achieve land
degradation targets, enhance climate change resilience and strengthen land-based mitigation,
and address trade-offs in land use, taking into account national policies, priorities and regulations.
Integrated land-use planning requires a participatory approach to ensure local communities and all
stakeholders, including marginalized or vulnerable groups and the private sector, engage in consensual
decision-making and conflict resolution. FAO offers technical support to Members to develop
country-specific integrated land-use planning approaches that account for national land governance
strategies and laws as well as diverse socioeconomic contexts, to enhance implementation, inter alia,
through decentralized governance mechanisms, negotiated territorial development, tenure security,
and access and user rights.

178 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


A land resources information management system (FAO, 2022d) is a tool designed to create a secure,
reliable, efficient, accountable and equitable system of land resources management for agriculture. It
comprises a comprehensive set of GIS-based tools (e.g. land suitability module and map generator)
and a central spatial database, and provides an evaluation of land-use suitability based on modular
multiple criteria analysis-based assessment, including a socioagricultural vulnerability analysis. It
allows assessments of physical and socioeconomic conditions of the land and evaluation of benefits
and constraints of different options by simulating various impact scenarios.

The Hand-In-Hand Initiative Data Platform (FAO, 2022e) is an evidence-based country-led FAO
initiative to accelerate agricultural transformation and sustainable rural development to support SDGs.
The platform guides action among partners and in keeping with national sustainable development
priorities. Tools, such as geospatial modelling and analytics, are available to identify the best
opportunities to raise incomes and reduce inequities and vulnerabilities among the rural poor and
present an evidence-based view of economic opportunities to improve targeting and tailoring of policy
interventions, innovation, finance and investment, and institutional reform.

The LRP Toolbox was developed by FAO in response to demand from a range of stakeholders
(planners, policymakers, governments, institutions, communities, technical specialists, etc.) for a
resource that supports participatory LRP. The toolbox provides information and an inventory of tools
and approaches to support the planning requirements of different stakeholders working at different
levels in different regions and sectors (Ziadat et al., 2021). It is web based and freely available, and
is regularly updated with summary descriptions and links to a comprehensive number of LRP tools
and approaches developed by FAO and other institutions. In 2021, the toolbox comprised 157 tools
grouped in five thematic domains in the land-use planning process: (i) biophysical approaches/tools,
(ii) socioeconomic and negotiation approaches/tools, (iii) integrated biophysical, socioeconomic and
negotiation approaches/tools, (iv) databases/information systems and (v) support tools. The tools
are further characterized in terms of thematic area, type of tool, scale of applicability and user (see
Figure 4.1).

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 179
Figure 4.1 Search criteria and options for the LAND RESOURCES
PLANNING Toolbox

MAIN CATEGORIES SUB CATEGORIES


Biophysical approaches/tools Land evaluation
Agroecological zoning and derived tools
Socioeconomic/negotiated approaches tools
Soil productivity Indecies
Integrated biophysical and socioeconomic/ Software/applications land resources planning
negotiated approaches/tools
Farm systems
Databases/information systems Gender
Governance/tenure
Support tools
Household surveys
Participatory/negotiated approaches

Rural appraisal
Spatial planning (urban/rural)
Territorial development/sustainable land management

Soil databases
Land degradation databases
Climate data bases
Statistics data bases
Crop databases

Assessment and mapping tools: land, soil, crop, water


Assessment and mapping tools: climate
Other support tools

THEMATIC AREAS TYPE OF TOOL


Agriculture, statistics Data
Agriculture, productivity Documentation/manuals
Cadaster Educational materials
Climate Framework/guidelines
Crops, distribution Maps/GIS
Crops, productivity Model
Crops, suitability Questionnaire/survey
Economy, statistics Software
Environment, the distichs
Farming systems SCALE OF APPLICABILITY
Food, statistics
Global
forestry, statistics
Regional
General
National
Land degradation
Subnational/province/district
Land evaluation
Watershed/basin/landscape
Land management/planning
Locality/farm/site
Land/water rights
Land/cover
Population, distribution
Population, statistics
Remote sensing
Social participatory approaches
Social, statistics
Soils, distribution and properties
soils, management and conservation
Water, productivity
Water, statistics

Source: FAO. 2022. Land Resources Planning Toolbox. In: Land & Water. Rome. www.fao.org/land-water/land/land-governance/land-resources-planning-tool-
box/en; adapted from Ziadat, F., De Pauw, E., Nachtergaele, F. & Fetsi, T. 2021. A land resources planning toolbox to promote sustainable land management.
Sustainable Agriculture Research, 10(1): 73.

180 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


ƒƒ Socioeconomic and negotiation

©FAO/Lekha Edirisinghe
approaches/tools covering aspects of
the human environment (e.g. farming
systems, tenure, gender, participatory
planning and governance). These
tools give prominence to social and
economic settings required for land-use
planning and include the participatory flagged in the SDG 6 synthesis report on
decision-making approaches and methods water and sanitation (United Nations, 2018).
of those institutions and actors involved Such deficiencies hinder sound planning and
in land management and governance. efforts to guide interventions and invest-
ments for sustainable land planning and
ƒƒ Integrated biophysical, socioeconomic management, such as locating and remedy-
and negotiation approaches/tools are used ing hotspots. Improving monitoring and data
to process information on biophysical collection must not be ignored; they should
characteristics and social and economic be harmonized and coordinated, drawing on
conditions, to consider access, user public and private sector investments.
rights, competition and conflict over
resources, and for managing trade-offs. 4.2.2 Land suitability
They incorporate principles, approaches
for crop production
and methods of participatory land-use
planning or LRP, with the overall Suitability analysis and land-suitability maps
objective of reaching mutually beneficial are important foundations for sound agri-
outcomes for all stakeholders, including cultural development planning. They provide
socioeconomic and environmental information on potential land suitability
benefits in line with the SDG framework. and limitations. They also help planners and
decision-makers to identify optimum land
The quality and availability of natural uses for current and potential agricultural
resources data at national, regional and global lands while supporting the protection and
levels are increasing rapidly as new data sustainable use and restoration of land and
sources come on stream. The climate crisis water resources.
has substantially improved and increased
climate resources data. Current terrain and Recent developments in approaches to LRP
land-cover data are detailed and reliable. for sustainable use and management of land
However, the lack of spatial distribution and and water resources exploit well-established
quality of soil resources data is constraining databases on climate, soil, terrain, land cover,
advances in land-use planning. The Global land use and crop requirements (Fischer et
Soil Partnership (GSP) is improving data al., 2021). They also exploit climate change
through its Global Soil Information System modelling to assess anticipated changes
and building country capacities in soil data in land suitability resulting from expected
and mapping through the International rising temperatures and changes in rainfall
Network of Soil Information Institutions. distribution. They aim to make the best use
Similarly, there is room for improvement of limited land (and water) resources, to
in acquiring surface water and groundwater yield optimal benefits of rational land use
data, particularly water quality data, an issue while avoiding conflicts over how and who

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 181
©FAO/Believe Nyakudjara production, protected areas, renewable water
resources, and climatic conditions for the
period 1961–2010, and a selection of future
climate simulations using the IPCC fifth
assessment report climate model outputs
for four RCPs. The analysis uses the latest
version of the Harmonized World Soil Data-
uses tracts of land (FAO, 1993). This process
base (HWSD; FAO et al., 2012).
turns promising land-use types, defined
in terms of products, agricultural inputs,
The GAEZ method is a global AEZ method and
management practices and water resources
is not designed for local-level use. However,
availability, into feasible production systems,
a case study using the Land Potential Knowl-
such as in rainfed and irrigated agriculture,
edge System (LandPKS) mobile application
forestry and ruminant livestock production.
illustrates how GAEZ could be integrated with
field data to downscale at local/farm level and
4.2.3 The Global Agro- benefit from the information in the GAEZ
Ecological Zones data portal (Bb
ox 4.2).
methodology
The GAEZ methodology, developed by FAO
4.2.4 Land suitability
and IIASA (Fischer et al., 2021), assesses
for rainfed crops
the potential for growing crops in terms of In the analysis of this report, GAEZ data are
the maximum potential and agronomically used to illustrate a range of options available
attainable crop yields for land resource for land units now and in the future for
units under different land-use types. The ten widely grown crops (wheat, maize, rice,
methodology uses agroclimatic, soil and sorghum, citrus, tomato, alfalfa, chickpea,
terrain data and levels of agricultural inputs olives and coffee), to guide in the selection of
and management to establish areas suitable promising land-use types based on climate
for sustainable agricultural use. These are data between 1981 and 2010.
generic agricultural production systems
defined by crop parameters, such as harvest The results for rainfed wheat provide an
index, maximum rate of photosynthesis, example of the mapping potential using the
maximum leaf area index, water supply GAEZ methodology. Map 4.1 shows current
systems in rainfed and irrigated systems, and potential production for rainfed wheat,
levels of inputs and management ranging assuming high input. The analysis uses data
from low to high. from the background report of Tuan et al.
(2022). Access to georeferenced results and
The first global AEZ assessment was in 2000. the GAEZ v4 data portal is also available at the
Since then, GAEZ assessments have been SOLAW 2021 website (FAO, 2022f; FAO and
updated continuously and published through IIASA, 2021).
data portals, in 2000 (GAEZ v1), 2002
(GAEZ v2) and 2012 (GAEZ v3). The latest Results for other crops can be processed using
data portal for GAEZ v4 and the database are data available on the GAEZ v4 data portal.
fully accessible to the public (FAO and IIASA, The procedure and results for selected crops
2021). In this analysis, GAEZ v4 uses 2010 are available in the current GAEZ v4 model
baseline data that include land cover, crop documentation (Fischer et al., 2021).

182 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


Box 4.2
Localizing/increasing accuracy of GLOBAL AGRO-ECOLOGICAL ZONES
predictions with site-specific soil data collected using a mobile application
A major obstacle to selecting the most appropriate crops and closing the yield gap is a lack of
site-specific soil information. This adds a high level of uncertainty to valuable land-use and management
planning tools, such as the FAO GAEZ soil suitability modelling framework.

The GAEZ framework uses soil data and detailed agronomic knowledge to predict crop-specific
agronomic potential. It accomplishes this by calculating seven crop-specific soil quality indices used to
generate crop-specific soil suitability ratings (Figure 4.2, item 3).

Figure 4.2 The GLOBAL AGRO-ECOLOGICAL ZONES soil suitability


downscaling framework

1. Site-based soil data 2. Soil map data

SoilID Traditional Digital


Soil Map Soil Map

Site-based soil
map data
LandPKS

Agro-Ecological Zone
Soil Suitability

3. Crop -specific soil recommendations


Localized Crop-specific
Soil Management
Recommendations
GAEZ Input Data GAEZ Soil
Quality Indices
• Site-Specific Soil Data
• Nutrient Availability
• Crop Type (54 crops)
• Nutrient Retention Capacity Crop-Specific GAEZ
• Farm Input Level (low,
• Rooting Conditions Soil Suitability
intermediate, high)
• Oxygen Availability
• Water Supply System
(rain -fed, rain-fed water
• Excess Salts
conservation, irrigated • Toxicities
• Workability

Source: Adapted from Grameen Foundation, University of Colorado Boulder & United States Department of Agriculture, Agricultural Research
Service. 2020. Map the the future (M2F): Integrating soil mapping into coca farm development plans in Ghana.
https://pubdocs.worldbank.org/en/613921612401424054/Grameen-Map2Future-Final-Report-low-res.pdf

Accurate soil information is critical for identifying limitations and management practices to improve crop
yields. However, this can be difficult and costly to obtain. Recent advances in information technologies
have made it possible to create mobile decision-support tools that can assist users in acquiring accurate
site-specific soil data (Figure 4.2). The LandPKS application is one such example. It provides a complete
mobile computing platform that allows non-soil scientists to describe and identify the soil at a location using
limited, simple soil observations. The application offers a digital interface for collecting and recording soil
profile data and a global soil identification tool (SoilID) that leverages user-recorded soil data, existing soil
maps and cloud-based computing to determine the most probable soil type at a location (Figure 4.2, item 2).

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 183
Information on static soil properties can be used directly to inform farmer decisions on various
management practices, such as irrigation frequency, the need for organic amendments and the
likelihood of erosion. Soil identification with direct links to FAO and other soil survey information via
LandPKS-SoilID can further tailor soil management decisions.

The AEZ modelling framework can translate site-specific soil information from LandPKS and other
applications into crop-specific soil suitability ratings. Recent work by the Unites States Department
of Agriculture’s Agricultural Research Service and University of Colorado scientists working on
LandPKS have taken the AEZ methodology and localized the soil suitability calculations by leveraging
site-specific soil property data and the LandPKS-SoilID algorithm to identify the most likely soil and/or
soil component at a sampling location from commonly used soil map products (Figure 4.2).

The AEZ downscaling framework (Figure 4.2) was evaluated at 6 065 LandPKS sampling sites in Ghana
using the scenario of rainfed, low-input maize production systems (Figure 4.3). This analysis compares
the soil suitability for maize based on the dominant mapped soil type from HWSD versus site-specific
soil data measured using the LandPKS application, combined with the HWSD soil map data selected
using the SoilID matching algorithm. This analysis shows that relying on the dominant soil mapped
at a location will often lead to an under or overestimation of soil suitability due to the inability of soil
maps to accurately characterize the variation of soil conditions across the sampling sites. These results
demonstrate the importance of site-specific soil data for understanding a soil’s agronomic limitations
and the feasibility of soil management interventions for improving crop yields. When smallholder
farmers have limited resources, these differences could mean success or failure or limited impact of
the investments made.

Sources: Herrick, J.E., Urama, K.C., Karl, J.W., Boos, J., Johnson, M.V.V., Shepherd, K.D., Hempel, J. et al. 2013. The global Land-Potential Knowledge System (LandPKS):
Supporting evidence-based, site-specific land use and management through cloud computing, mobile applications, and crowdsourcing. Journal of Soil and Water
Conservation, 68(1); Maynard, J.J., Salley, S.W., Beaudette, D.E. & Herrick, J.E. 2020. Numerical soil classification supports soil identification by citizen scientists using
limited, simple soil observations. Soil Science Society of America Journal, 84(5): 1675–1692.

Figure 4.3 Maize soil suitability at 6 065 LAND POTENTIAL KNOWLEDGE


SYSTEM’ sampling sites in Ghana based on low-input rainfed
farming systems

100

Soil Suitablity
80
Slight constraint
Percentage of Sites

60 Moderate constraint

Severe constraint
40
Very severe constaint

Not suitable
20

0
HWSD Dom LandPKS-SoilID

Soil Data Source


Note: HWSDDom = dominant soil type mapped at a location.

Source: Jonathan Maynard, University of Colorado, Boulder

184 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


MAP 4.1 Potential production for rainfed wheat at high inputs (tonnes/ha), based
on average climate from 1981 to 2010 and global soil and terrain information

79

Dotted line represents approximately the Line of Control in Jammu and Kashmir agreed upon by India and
Source: FAO & International Institute for Applied Systems Analysis. 2021. Global Agro-Ecological Zoning
Pakistan. The final status of Jammu and Kashmir has not yet been agreed upon by the parties.
version 4 (GAEZ v4). In: FAO. Rome. www.fao.org/gaez/entps://gaez-data-portal-hqfao.hub.arcgis.com
Modified to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420 Final boundary between the Sudan and South Sudan has not yet been determined.

This analysis addresses land suitability for rights over land and other resources. The
crop production. However, land-suitability impact of market reform policies on marginal
maps for livestock, forestry mixed agrofor- areas has often been detrimental to the poor.
estry and agropastoral systems are equally
important. Marginal lands for cropland may 4.2.5 Mapping yields
be suitable for livestock and forestry enter-
and production
prises (Bbox 4.3).
The GAEZ methodology produces the most
Policies for marginal environments should likely distribution (pixel level) of crops
encourage the use of ecological processes within cultivated land, their yields (mass per
instead of relying entirely on external inputs unit area) and production, by downscaling
for crop production. Future technologies national and subnational land-use data from
should account for and must be suited to the FAOSTAT, Agro-MAPS and national statis-
high degree of diversity in biophysical and tics, complemented with information on
socioeconomic conditions typical of marginal land cover and land suitability. Details of the
areas. Farm policies intended for marginal procedure are available in the GAEZ v4 model
agriculture must therefore encourage prop- documentation (Fischer et al., 2021).
erty rights systems to secure the ownership

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 185
Box 4.3
Marginal lands for crop production
“Marginal land” refers to land that is no longer economical for crop production. However, it could still
be important for grazing. Marginal land has little potential for profit, and often has poor soil or other
undesirable characteristics. FAO and UNEP classified land supporting a yield of up to 40 percent of
the crop potential as marginal. Such land is identified as areas where cost-effective production is not
possible under given conditions, cultivation techniques, agriculture policies, and macroeconomic and
legal settings.

Marginal areas are perceived to have low crop production potential, which has led to bias in
policymaking to support the development of agriculture in marginal areas. However, marginality is not
a static and permanent condition, and marginal lands are subject to change in land use, agricultural
technologies and socioeconomic environment (Map 4.2). Investments in technologies and applying
good management practices and tools could reverse this situation. Thus, unproductive and marginal
lands could be transformed into productive agricultural lands. Marginal areas present opportunities
for alternative models of development. Research and development and public policies towards these
marginal lands need to be revised to target marginal producers, especially the extremely poor, to
provide incentives to maintain and improve the natural resource base for production without further
land degradation.

In 2010, about 1.75 billion people worldwide (38 percent of the rural populations) lived in remote
less-favoured agricultural areas, up from 1.56 billion people in 2002, and the majority of them (1.6 billion
out of 1.75 billion) were in developing countries.

Map 4.2 Marginality hotspots – overlapping dimensions of


marginality

Marginality
0

Source: V. Graw, personal communication (2022), based on Ahmadzai, H., Tutundjian, S. & Elouafi, I. 2021. Policies for sustainable agriculture and
livelihood in marginal lands: A review. Sustainability (Switzerland), 13(16): 1–18.
Modified to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420

186 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


MAP 4.3 Distribution and intensity of cropland, 2010 (% of 30 arcsecond grid cell)

7.9

Source: FAO & International Institute for Applied Systems Analysis. 2021. Global Agro-Ecological- Dotted line represents approximately the Line of Control in Jammu and Kashmir agreed upon by India and
Pakistan. The final status of Jammu and Kashmir has not yet been agreed upon by the parties.
Zoning version 4 (GAEZ v4). In: FAO. Rome. www.fao.org/gaez/en
Modified to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420 Final boundary between the Sudan and South Sudan has not yet been determined.

Map 4.3 shows the crop cover rate, between 0 and location of current and potential yield
and 100 percent, of cropland in a grid cell for and production gaps is essential to exploit
2010. This illustrates the heavy concentration investment opportunities and enhance food
of cropland in temperate and subtropical production.
zones with the highest concentrations
around the Great Lakes in Canada and the Comparing actual crop yields and potential

United States of America, Central and Eastern attainable crop yields identifies areas where

Europe, China, northern India and Pakistan. increases in food production are achievable

Map 4.4 illustrates georeferenced rainfed by improving management practices.28 More

maize yields in 2010. details on yield-gap analysis and calculations


are available in GAEZ v4 model documenta-

4.2.6 Mapping yield gaps tion (Fischer et al., 2021).

Current rainfed crop yields and production Map 4.5 illustrates yield-gap ratios for maize
fall short of what is potentially achievable in 2010. The most significant gaps occur in
in many regions. For example, in sub-
Saharan Africa, yields are only 24 percent of 8
Actual georeferenced crop distribution and yields, from
downscaling 2009–2011 statistics, were compared with
what is possible with higher levels of input
corresponding anticipated yields obtained using AEZ
and management. Understanding the extent crop modelling (estimated in GAEZ v4).

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 187
MAP 4.4 Downscaled yield of rainfed maize, 2010 (tonnes/ha)

10

Source: FAO & International Institute for Applied Systems Analysis. 2021. Global Agro-Ecological Dotted line represents approximately the Line of Control in Jammu and Kashmir agreed upon by India and
Pakistan. The final status of Jammu and Kashmir has not yet been agreed upon by the parties.
Zoning version 4 (GAEZ v4). In: FAO. Rome. www.fao.org/gaez/en
Modified to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420 Final boundary between the Sudan and South Sudan has not yet been determined.

India and most countries in Africa, signalling China and the southeast coast of Australia.
much lower yields than potentially achiev- Significant yield gaps in most of Africa, partic-
able in these areas. Similar maps are available ularly in the Sahel, reflect current low levels of
for other crops. inputs and management. The substantial yield
gaps in Central America, India and the Russian
Map 4.6 illustrates the yield-gap occur- Federation are partly attributed to lower inputs
rences based on 26 main crops.39 The small and partly to suboptimum management.
gaps reflect high levels of manage-
ment and inputs in Canada, the midwest Regions where the yields are high and the
of the United States of America, parts yield gaps are small (green) have the most
of Brazil, Western Europe, southern significant land degradation risks due to
unsustainable intensification. Sustainable
9
Wheat, rice, maize, sorghum, millet, barley, other
cereals (buckwheat, oats, rye, upland rice), tubers management methods are needed to counter
(sweet and white potato combined), cassava, yams and soil pollution, compaction and sealing, sali-
other roots, sugar beet, sugar cane, pulses (Phaseolus
beans, chickpeas, cow peas, dry peas, pigeon peas, nization, acidification, erosion, carbon and
gram), soybean, rapeseed, sunflower, groundnut, oil biodiversity loss and soil sealing (Chapter 5).
palm fruit, olives, cotton, banana, tobacco, vegetables
(cabbages, carrots, onions, tomatoes), stimulants Areas with low yields and significant yield
(cocoa, coffee, tea), fodder crops and all other crops
from FAOSTAT.

188 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


MAP 4.5 Yield achievement ratio (100 × actual/potential) for maize under rainfed
water supply conditions, 2010

< 10%
10%-25%
25%-40%
40%-55%
55%-70%
70%-85%
> 85%
Dotted line represents approximately the Line of Control
in Jammu and Kashmir agreed upon by India and Pakistan.
The final status of Jammu and Kashmir has not yet been agreed upon by the parties.
Source: FAO & International Institute for Applied Systems Analysis. 2021. Global Agro-Ecological
Zoning version 4 (GAEZ v4). In: FAO. Rome. www.fao.org/gaez/en. Modified to comply with Final boundary between the Sudan and South Sudan has not yet been determined.

UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420


Final status of the Abyei area is not yet determined

MAP 4.6 Yield achievement ratio (100 × actual/potential) for 26 crops in current
rainfed cropland, 2009–2011

< 10%
10%-25%
25%-40%
40%-55%
55%-70%
70%-85%
> 85%

Source: FAO & International Institute for Applied Systems Analysis. 2021. Global Agro-Ecological Dotted line represents approximately the Line of Control in Jammu and Kashmir agreed upon by India and
Pakistan. The final status of Jammu and Kashmir has not yet been agreed upon by the parties.
Zoning version 4 (GAEZ v4). In: FAO. Rome. www.fao.org/gaez/en. Modified to comply with
UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420 Final boundary between the Sudan and South Sudan has not yet been determined.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 189
are also areas where opportunities exist to
increase production and productivity by
selecting crops according to the current and
©FAO/Vasily Maximov

near-future land suitability and by improv-


ing on-farm and territorial or watershed
management practices. This analysis and
mapping help identify yield-gap hotspots
where future investment in sustainable land
gaps (yellow) reflect mainly soil nutrient and water management is likely to succeed.
deficiencies. They are at risk from different Bbox 4.4 describes another approach based on
forms of land degradation resulting from real-time remote-sensing data.
nutrient mining, overgrazing and defor-
estation, including large-scale desertifica- Unlocking this yield potential is difficult.

tion, soil erosion and biodiversity loss. This Several socioeconomic and ecological condi-

points to the need to promote sustainable tions determine whether farmers are willing

intensification in areas with small and large to apply management practices and higher

yield gaps. inputs that are affordable, desirable and feasi-


ble, and support the adoption of improved
The areas where significant yield gaps farming systems and management practices
exist (yellow) are usually associated with (SLM). Much depends on good governance at
subsistence and low-input farming. They local and municipal levels (Chapter 5).

Box 4.4
Forecast crop yields informing the European Union’s Common Agricultural
Policy and drought management
The European Commission’s JRC in Ispra, Italy, houses the European Union’s Food Security Unit,
whose role is to forecast crop production during the current growing season to inform the European
Commission’s Common Agricultural Policy.

To do this, JRC developed the Monitoring Agricultural Resources (MARS) Crop Yield Forecasting
System. This provided timely forecasts of crop production, including biofuel crops, for Europe and
other strategic areas of the world, including Africa, China, India, Kazakhstan, the Russian Federation and
South America since 1992. The system monitors crop vegetation growth (cereal, oil seed crops, protein
crops, sugar beet, potatoes, pastures and rice), including the short-term effects of meteorological
events on crop production. It also provides seasonal yield forecasts of key European crops, thereby
contributing to evaluating global production estimates of crops such as wheat and maize to support
Common Agricultural Policy management decisions.

Software tools are available to access the data for analysis to support decision-makers and are
invaluable for informing users about the potential impacts of agricultural drought. The JRC MARS
Explorer displays current weather conditions and progress in crop growth based on meteorological
station data, crop growth simulations and remote-sensing observations originating from the MARS
Crop Yield Forecasting System. An analysis of weather, crop conditions and quantitative crop yield
forecasts for Europe is published monthly in the JRC MARS bulletins on crop monitoring in Europe.

Source: European Commission. 2022. Monitoring agricultural resources (MARS). In: EU Science Hub.
https://ec.europa.eu/jrc/en/mars

190 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


4.2.7 Planning for land
suitability under a
changing climate

©FAO/Soliman Ahmed
Climate change is likely to change land
suitability and productivity in the future.
Using the most advanced tools available,
such as general circulation models, it is now
possible to compare current land suitability
and productivity at baseline climate (1981– some crops, which could mean higher or
2010) with anticipated changes in the 2080s lower yields, shifts in suitable geographical
(2070–2099). However, the resolution is areas expanding some and shrinking others,
coarse and the predicted rainfall distribution and opportunities for multiple cropping.
is less reliable than that of temperature. Note that the model assumption is that soil
Several general circulation models exist with conditions remain unchanged over time, and
advantages and disadvantages, so an average current fragile permafrost areas are assumed
(“ensemble”) of results was used for five to remain permanently protected and are not
main models (Bindoff et al., 2013). The most included in agricultural projections.
realistic pathway for SOLAW 2021 analysis
Shifts in land suitability. A common rainfed
is a middle-of-the-road RCP (RCP 4.5). This
cereal crop (wheat) and a cash crop (coffee)
scenario leads to an expected temperature
illustrate the potential impacts of shifts in
increase of 2.0 °C by 2100. A high-end
land suitability. Map 4.7 shows shifts in land
scenario was also used with RCP 8.5, resulting
suitability for wheat based on high inputs
in a temperature increase of 4.2 °C by 2100.
and RCP 4.5. Areas marked green show land-
Several maps illustrate various options for suitability increases, while those marked red
the ten crops (wheat, maize, rice, sorghum, show decreases. Thus, Argentina, Canada,
citrus, tomato, alfalfa, chickpea, olives Northern Eurasia, South Africa and the United
and coffee). These show three important States of America would see the areas of suit-
changes: changes in land suitability for able land increasing (green) and northern
Brazil, Central Africa and Eastern Europe
would see areas decreasing (red). This does
not mean that red areas would be unsuit-
able for wheat; instead, alternative crop
types/improved species and varieties with
adapted tolerance traits and crop manage-
ment may be needed in the future.

Map 4.8 illustrates shifts in land suitabil-


ity for rainfed coffee grown under RCP 4.5.
©FAO/Hoang Dinh Nam

Large areas presently suitable for growing


coffee would decline. A significant decline is
expected in Brazil and West Africa. Gains in
land suitability for coffee are likely in East
Africa and parts of China. Tbable 4.1 illus-

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 191
MAP 4.7 Difference in land suitability for rainfed wheat with high inputs on
actual cultivated land between baseline climate (1981–2010) and the
projected climate for the period 2070–2099, ensemble mean for
REPRESENTATIVE CONCENTRATION PATHWAY 4.5 scenario

Value
Positive

Negative

Source: Tuan, H., Nachtergaele, F., Chiozza, F. & Ziadat, F. 2022. Land suitability for crop production in the Dotted line represents approximately the Line of Control in Jammu and Kashmir agreed upon by India and
future. Thematic Background Report for SOLAW 2021. Rome, FAO. www.fao.org/land-water/solaw2021/en; Pakistan. The final status of Jammu and Kashmir has not yet been agreed upon by the parties.

based on GAEZ v4 data from FAO & International Institute for Applied Systems Analysis. 2021. Final boundary between the Sudan and South Sudan has not yet been determined.
Global Agro-Ecological Zoning version 4 (GAEZ v4). In: FAO. Rome. www.fao.org/gaez/en.
Modified to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420

trates how the total area of land-use suit-


ability for coffee410 may decline by the 2080s
based on a conservative emission scenario of
RCP 4.5. Of the areas currently most suited
to growing coffee (880 thousand km2), about
335 thousand km2 (38 percent) would remain
unchanged, but yields would decline on
545 thousand km2 (62 percent), and 300 thou-
©FAO/Isaac Kasamani

sand km2 (34 percent) would no longer be

10
The coffee scenario assesses land suitability for arabica
and robusta varieties, and assumes an agronomic
adaptation based on temperature. However, the
overall message is that coffee growing will be seriously
affected by climate change.

192 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


MAP 4.8 Difference in land suitability for rainfed coffee between baseline climate
(1981–2010) and the climate in the period 2070–2099, ENSEMBLE MEAN FOR
REPRESENTATIVE CONCENTRATION PATHWAY 4.5 scenario

Value
Positive

Negative

Source: Tuan, H., Nachtergaele, F., Chiozza, F. & Ziadat, F. 2022. Land suitability for crop production in the Dotted line represents approximately the Line of Control in Jammu and Kashmir agreed upon by India and
Pakistan. The final status of Jammu and Kashmir has not yet been agreed upon by the parties.
future. Thematic Background Report for SOLAW 2021. Rome, FAO. www.fao.org/land-water/solaw2021/en;
based on GAEZ v4 data from FAO & International Institute for Applied Systems Analysis. 2021. Final boundary between the Sudan and South Sudan has not yet been determined.
Global Agro-Ecological Zoning version 4 (GAEZ v4). In: FAO. Rome. www.fao.org/gaez/en.
Modified to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420

suitable. Tbable 4.1 also shows a range of ipatory land-use planning processes would
suitability classes using the area suitability provide the means of encouraging land users
index (SI). These results emphasize the need to consider changes backed up as needed by
to adjust coffee management practices or to an appropriate enabling environment such
shift locations of coffee production. as incentives, financing, enhancing capac-
ity, policies, tenure security and market-
Other options include breeding and selecting ing support.
crop resources, changing land use by switch-
ing to crops, including trees and livestock, Map 4.9 illustrates shifts in land
more suited to the changing climate. This areas suitable for rainfed wheat for a
would involve changing land management high-emission/high-temperature scenario
practices and adapting food systems to turn (RCP 8.5), leading to a 4.2 °C temperature
these opportunities into realistic adap- increase. Wheat production would increase
tations to climate change. However, there in Argentina, Australia, Canada, Chile and
are many factors other than suitable land Northern Eurasia, and decline in most of
use to consider, such as local biophysical Central Africa and parts of Brazil, Central
conditions and socioeconomic issues. Partic- Asia and India. Other crop results are mixed,

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 193
Table 4.1 Change in the extent (km2) of land suitability classes for rainfed coffee
between baseline climate (1981–2010) and the climate in the period
2070–2099 (2080s), for ENSEMBLE MEAN FOR REPRESENTATIVE CONCENTRATION
PATHWAY 4.5 scenario

Suitability Area extent Unchanged Enhanced Decreased Changed to


class (thousand suitability suitability suitability not suitable
km2) (thousand (thousand (thousand (thousand
km2) km2) km2) km2)

1980–2010 2080 2080 2080 2080

SI > 85 880 335 0 545 300

SI > 70 2 920 830 330 1 760 1 465

SI > 55 4 990 1 290 365 3 335 2 860

SI > 40 6 180 1 105 440 4 635 4 045

SI > 25 4 825 1 060 300 3 465 3 015

SI > 10 3 870 1 335 265 2 270 2 030

SI > 0 2 420 1 560 170 690 690

SI = 0 107 310 106 185 1 125 0 0

Total 133 395 113 700 2 995 16 700 14 405

Note: SI > 85 indicates very high suitability; SI > 70 indicates high suitability; SI > 55 indicates good suitability; SI > 40 indicates medium suitability;
SI > 25 indicates moderate suitability; SI > 10 indicates marginal suitability; SI > 0 indicates very marginal suitability; SI = 0 indicates not suitable.

Source: Tuan, H., Nachtergaele, F., Chiozza, F. & Ziadat, F. 2022. Land suitability for crop production in the future. Thematic Background Report for
SOLAW 2021. Rome, FAO. www.fao.org/land-water/solaw2021/en; based on GAEZ v4 data from FAO & International Institute for Applied Systems
Analysis. 2021. Global Agro-Ecological Zoning version 4 (GAEZ v4). In: FAO. Rome. www.fao.org/gaez/en

with some predicted to increase and others to Shifts in opportunities for multiple crop-
reduce potential cropped areas. ping. Single crop yields do not reflect the
full potential of the land for rainfed agricul-
ture in areas where growing periods allow
more than one crop to be grown annually or
seasonally on the same tract of land. Several
zones are defined by matching growth cycle
and temperature requirements of individual
crops with the time available for crop growth
to assess multiple cropping potential. Param-
eters used were the number of days during
which temperature and moisture conditions
permit crop growth and the accumulated
© Olivier Asselin

temperatures (degree day) required to meet


heat unit requirements of individual crops or
sequential crop combinations.

194 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


MAP 4.9 Difference in land suitability for rainfed wheat between baseline climate
(1981–2010) and the climate in the period 2070–2099, ENSEMBLE MEAN FOR
REPRESENTATIVE CONCENTRATION PATHWAY 8.5 scenario

Value
Positive

Negative

Source: Tuan, H., Nachtergaele, F., Chiozza, F. & Ziadat, F. 2022. Land suitability for crop production in the future. Dotted line represents approximately the Line of Control in Jammu
and Kashmir agreed upon by India and Pakistan. The final status of
Thematic Background Report for SOLAW 2021. Rome, FAO. www.fao.org/land-water/solaw2021/en; based on
Jammu and Kashmir has not yet been agreed upon by the parties.
GAEZ v4 data from FAO & International Institute for Applied Systems Analysis. 2021. Global Agro-Ecological
Zoning version 4 (GAEZ v4). In: FAO. Rome. www.fao.org/gaez/en. Modified to comply with UN. 2020. Map of Final boundary between the Sudan and South Sudan has not yet
been determined.
the World. https://www.un.org/geospatial/file/3420

Map 4.10 illustrates the extent of multiple hemisphere and anticipated higher rainfall
cropping zones for baseline climate (1981– in some areas, the single-cropped area could
2010). Map 4.11 illustrates multiple cropping increase by 9 751 thousand km2 (20 percent)
zones for the 2080s (2070–2099), showing (from no cropping). Double cropping with
the effects of climate change. Supplemen- rice could increase by 601 thousand km2
tary irrigation could also extend the growing (27 percent), and the potential for triple rice
season and add value, but introducing irriga- cropping would be 910 thousand km2 (34.3
tion brings another set of problems, such as percent).
access to equipment and water, cost and the
required skills to practice efficient irrigation Apart from the adverse impact of climate

practices. change on current crop production systems,


results indicate significant potential oppor-
Tb
able 4.2 lists the absolute and percentage tunities to increase crop production using
changes in rainfed multiple cropping poten- alternative crops. Several land-use options
tial between baseline climate (1981–2010) and are available to enhance farmers’ resilience
the 2080s (ENS-RCP 4.5 scenario). Selected and adaptation to climate change. Realiz-
significant changes are highlighted in red. ing these “benefits” of climate change will
Due to higher temperatures in the northern largely depend on the ability of farmers to

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 195
MAP 4.10 Multiple cropping zone classes for rainfed conditions, climate of 1981–2010

no cropping
single cropping
limited double cropping
double cropping
double cropping with rice
double rice cropping
triple cropping
triple rice cropping

Source: FAO & International Institute for Applied Systems Analysis. 2021. Global Agro-Ecological Zoning Dotted line represents approximately the Line of Control in Jammu and Kashmir agreed upon by India and
Pakistan. The final status of Jammu and Kashmir has not yet been agreed upon by the parties.
version 4 (GAEZ v4). In: FAO. Rome. www.fao.org/gaez/en. Modified to comply with UN. 2020.
Map of the World. https://www.un.org/geospatial/file/3420 Final boundary between the Sudan and South Sudan has not yet been determined.

MAP 4.11 Multiple cropping zone classes for rainfed conditions, and the climate in
the period 2070–2099, ENSEMBLE MEAN FOR REPRESENTATIVE CONCENTRATION PATHWAY 4.5 scenario

no cropping
single cropping
limited double cropping
double cropping
double cropping with rice
double rice cropping
triple cropping
triple rice cropping

Source: FAO & International Institute for Applied Systems Analysis. 2021. Global Agro-Ecological Zoning version 4 Dotted line represents approximately the Line of Control in Jammu and Kashmir agreed upon by India and
Pakistan. The final status of Jammu and Kashmir has not yet been agreed upon by the parties.
(GAEZ v4). In: FAO. Rome. www.fao.org/gaez/en. Modified to comply with UN. 2020. Map of the World.
https://www.un.org/geospatial/file/3420 Final boundary between the Sudan and South Sudan has not yet been determined.

196 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


Table 4.2 changes of rainfed multiple cropping potentials between baseline climate
(1981–2010) and the 2080 climate (ENSEMBLE MEAN FOR REPRESENTATIVE CONCENTRATION
PATHWAY 4.5)
Future climate (2080s ENS-RCP 4.5) Change

Total baseline
cropping with

Difference (%)
climate (1981–
wetland rice
No cropping

(ENS-RCP 4.5)
Total 2080s

Difference
Triple rice
Rainfed
cropping

cropping

cropping

cropping

cropping

cropping
multiple

double

Double

Double

Double
Limited

(000 ha)
Single

cropping

Triple

2010)
zones (000

rice
ha)

No 38 628 100 9 751 000 38 800 500 0 0 0 0 48 418 400 39 817 000 −8 601 400 −18
Baseline climate (1981–2010)

cropping

Single 1 188 900 40 582 700 3 674 400 187 200 3 000 0 0 0 45 636 200 52 233 500 6 597 300 14
cropping

Limited 0 1 659 500 3 325 500 1 352 700 424 900 10 700 300 0 6 773 600 8 897 800 2 124 200 31
double
cropping

Double 0 224 900 1 811 800 9 485 100 696 400 447 700 20 100 0 12 686 000 13 710 800 1 024 800 8
cropping

Double 0 15 400 46 800 538 200 534 100 601 700 486 500 0 2 222 700 1 857 500 −365 200 −16
cropping with
wetland rice

Double 0 0 500 2 057 600 140 500 3 328 500 849 500 36 400 6 413 000 6 781 900 368 900 6
wetland rice
cropping

Triple 0 0 0 21 200 58 600 367 600 1293 400 910 600 2 651 400 2 756 400 105 000 4
cropping

Triple 0 0 0 68 300 0 2 025 700 106 600 7 325 900 9 526 500 8 272 900 −1 253 600 −13
wetland rice
cropping

Total 2080s 3 981 7000 5 223 3500 889 7800 1 371 0800 185 7500 678 1900 275 6400 827 2900
(RCP 4.5)

Note: Green indicates no change.

Source: Tuan, H., Nachtergaele, F., Chiozza, F. & Ziadat, F. 2022. Land suitability for crop production in the future. Thematic Background Report for SOLAW 2021. Rome, FAO.
www.fao.org/land-water/solaw2021/en; based on GAEZ v4 data from FAO & International Institute for Applied Systems Analysis. 2021. Global Agro-Ecological Zoning version 4
(GAEZ v4). In: FAO. Rome. www.fao.org/gaez/en

select suitable land uses and implement the period 2070–2099 (ENS-RCP 4.5). Addi-
sustainable crop, land and water manage- tional maps and results are available on the
ment practices. SOLAW 2021 website (FAO, 2022f).

Overall crop suitability under present and Results show that more than one crop type
future climates. Ten crops (wheat, maize, is suitable for some locations, indicating a
rice, sorghum, citrus, tomato, alfalfa, chick- range of options available for future land use.
pea, olives and coffee) were assessed for However, these results are derived at a global
their suitability and subsequently mapped level with limited crop selection and using
when their SI was greater than 40 (medium globally available climate, soil and terrain
suitability). Map 4.12 shows the locations datasets. This would support decision-
of those crops attributed with the highest making at global and possibly national levels,
suitability under baseline climate conditions but it would be of limited value at a local
(1981–2010). Map 4.13 presents the number level. To overcome this, FAO has developed
of crops that can be grown (SI > 40) in high-resolution national AEZ (NAEZ) studies

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 197
MAP 4.12 Most suitable crops (SUITABILITY INDEX > 40) for 1981–2010 climate conditions
based on an analysis of ten crops

Limited Chickpea Maize Tomato Multi crop


agricultural use Citrus Olive Wetland rice Alfalfa
Protected area Coffee Sorgum Wheat

Source: Tuan, H., Nachtergaele, F., Chiozza, F. & Ziadat, F. 2022. Land suitability for crop production in the future. Dotted line represents approximately the Line of Control in Jammu and Kashmir agreed
upon by India and Pakistan. The final status of Jammu and Kashmir has not yet been
Thematic Background Report for SOLAW 2021. Rome, FAO. www.fao.org/land-water/solaw2021/en; based on agreed upon by the parties.
GAEZ v4 data from FAO & International Institute for Applied Systems Analysis. 2021. Global Agro-Ecological
Zoning version 4 (GAEZ v4). In: FAO. Rome. www.fao.org/gaez/en. Modified to comply with UN. 2020. Final boundary between the Sudan and South Sudan has not yet been determined.

Map of the World. https://www.un.org/geospatial/file/3420

4.2.8 Prospects for land-


to support subnational decisions on crop type
and management practices. Several NAEZ
studies are available, including for Ghana suitability analysis
(Bbox 4.5) and the case study using LandPKS
The results from this land-suitability analysis
(Bbox 4.2). A study in North Macedonia
provide general guidance for a range of
providing crop suitability maps at 100 m
options for land use and future crop selection
resolution is published as an Agro-ecological
(summarized in Bbox 4.6). However, climate
atlas of the Republic of North Macedonia (Aksoy
modelling has limitations and inherent
et al., 2020). The NAEZ studies are available
uncertainties for simulating the effects on
for Afghanistan, Ghana, Lao People’s
land suitability and crops. The methodology
Democratic Republic, Pakistan, Thailand and
does not account for changes in soil carbon,
Turkey (FAO, 2022g).
soil erosion (Borrelli et al., 2020), land
degradation, sea-level rise and anticipated
changes in extreme weather events, nor

198 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


MAP 4.13 Number of different crops possible to be grown (SUITABILITY INDEX > 40) for
the period 2070–2099 climate conditions (ENSEMBLE MEAN FOR REPRESENTATIVE
CONCENTRATION PATHWAY 4.5) based on analysis of ten crops

SI > 35
Number of crops
1
2
3
4
5
6
7
8
Not suitable
Limited agricultural use
Protected area

Source: Tuan, H., Nachtergaele, F., Chiozza, F. & Ziadat, F. 2022. Land suitability for crop production in the future. Thematic Dotted line represents approximately the Line of Control in Jammu and Kashmir
agreed upon by India and Pakistan. The final status of Jammu and Kashmir has not
Background Report for SOLAW 2021. Rome, FAO. www.fao.org/land-water/solaw2021/en; based on GAEZ v4 data from FAO
yet been agreed upon by the parties.
& International Institute for Applied Systems Analysis. 2021. Global Agro-Ecological Zoning version 4 (GAEZ v4). In: FAO.
Rome. www.fao.org/gaez/en. Modified to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420 Final boundary between the Sudan and South Sudan has not yet been determined.

Box 4.5
High-end climate change impact on rainfed crops in Ghana
Climate change threatens rainfed production systems in sub-Saharan Africa. In Ghana, an NAEZ was
developed to assess the impacts of high-end RCP 8.5 global warming on agricultural production until
the end of this century.

Results highlight different potential impacts across the country, mainly due to significant increases in
the number of days exceeding high-temperature thresholds. Rainfed production of several food and
export crops could be significantly reduced compared to the historical 30 year average (1981–2010) (see
Table 4.3). By the 2050s, plantain production (an important food crop) would be less than half of current
levels, and fall by more than 90 percent by the 2080s. Suitable areas for cocoa production (an important
cash crop) would be only one-third of current levels. Production of other crops, such as oil palm, sugar
cane, robusta coffee and rubber, would also suffer. Maize, sorghum and millet production would cope
much better in the warmer climate.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 199
Table 4.3 Changes in production for baseline climate and climate
scenario ensembles for the 2050s and 2080s with (+)
and without carbon dioxide fertilization on very suitable
(VS), suitable (S) and moderately suitable (MS) land
Rainfed crops VS+S+MS Change in production
– high inputs Area
and advanced
management production ENS+ ENS Change ENS+ ENS Change

(thousand Base 2050s 2050s 2080s 2080s


tonnes)

Banana/plantain 17 071 100 54 43 ↓↓ 21 8 ↓↓


(perennial C3)

Beans (annual C3) 14 532 100 106 93 ↓ 111 89 ↓

Cashew (perennial 11 657 100 104 92 ↓ 82 65 ↓


C3)

Cassava (perennial 42 709 100 104 91 ↓ 100 80 ↓


C3)

Cocoa (perennial 6 685 100 72 62 ↓ 35 24 ↓↓


C3)

Coconut 12 655 100 98 84 ↓ 97 76 ↓


(perennial C3)

Coffee (perennial 7 967 100 82 70 ↓ 62 42 ↓↓


C3)

Cotton (annual C3) 3 131 100 123 103 ↔ 129 95 ↔

Groundnut (annual 12 880 100 107 94 ↓ 108 85 ↓


C3)

Maize (annual C4) 32 088 100 116 109 ↑ 123 111 ↑

Mango (perennial 24 143 100 92 81 ↓ 72 54 ↓


C3)

Oil palm (perennial 10 761 100 72 59 ↓ 73 51 ↓


C3)

Pearl millet 7 059 100 164 141 ↑ 192 149 ↑


(annual C4)

Rubber (perennial 2 912 100 67 53 ↓ 65 36 ↓↓


C3)

Sorghum (annual 20 238 100 134 126 ↑ 143 129 ↑


C4)

Sugarcane 26 936 100 79 72 ↓ 78 67 ↓


(perennial C4)

Sweet potato 38 855 100 109 96 ↔ 109 88 ↓


(annual C3)

Yam (perennial C3) 34 129 100 101 89 ↓ 96 75 ↓

Note: Arrows refer to results without carbon dioxide fertilization effects and indicate changes of less than 5 percent
(↔), 5–25 percent (↓↑), 25–50 percent (↓↑) and losses of more than 50 percent (↓↓) compared to baseline
conditions.

Source: Fischer, G. & van Velthuizen, H. 2018. High-end climate change impacts on rain-fed crops in Ghana. Rome, FAO. www.fao.org/3/cb5581en/cb5581en.pdf

200 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


Box 4.6
Summary of anticipated shifts in land suitability
Indications are that climate change will bring shifts in land suitability. Some cropped areas will increase,
while others will reduce or deteriorate, requiring changes in crop selection and management. Areas
where suitable land for current crops decreases will require changes in crop variety (selection and
breeding) or a switch to other crops better adapted to the changed conditions. Other options may
include changes in water management, such as dryland farming options or irrigation with attention to
maximizing water-use efficiency when water resources are available and/or a shift to more resilient
mixed agroforestry or agropastoral systems.

Higher levels of carbon dioxide concentrations (RCP 8.5 compared to RCP 4.5) suggest a greater shift
in the current land-use pattern, and more intensive land management and land-use changes may be
needed in the future to maintain/enhance crop productivity.

Increasing temperatures would improve options for expanding cereal production to higher latitudes,
benefiting especially Canada and Northern Eurasia. However, in other areas, such as the highly
productive wheat areas in Central and Eastern Europe, it is likely to decline.

Moreover, increasing temperatures would reduce traditional cash crops, such as coffee in Brazil and
West Africa and olives in the Maghreb. But better growing conditions for coffee may occur in other
areas such as East Africa.

Alternative crops (adaptation) and adjustments in management practices, including technology transfer
programmes, will be needed in some regions where farmers must change their traditional cropping
patterns.

There are large areas where crop production would benefit from adopting higher inputs and improved
crop management.

Climate change may bring opportunities for increasing multiple rainfed cropping, particularly in the
tropics and parts of the subtropics.

Increasing investment in germplasm and seed exchange among ecoregions and crop breeding for
tolerant traits will be crucial in developing crops and varieties that can withstand future changes in
temperature, soil moisture supply, salinity, wind speed and evaporation.

For those areas where the climate becomes marginal for current staple and niche crops, there are
alternative annual and perennial tree crops, livestock, and soil and water management options
available. Experiences from similar ecoregions and other socioeconomic contexts should be analysed
to guide how the land is best used in the future.

Socioeconomic and ecological conditions will essentially determine the feasibility and justify investing
in the most appropriate adaptations. Such analysis and scenario development are essential elements
of land-use planning, as are participatory approaches that involve all stakeholders, notably farmers,
pastoralists, and fishers and foresters and their rural communities, and other users of the land and water
resources (in aquaculture, beekeeping, greenhouse use, carbon manufacture and sand mining).

Whatever the choices, future cropping should avoid protected areas and fragile lands, such as
land under permafrost, peatland, steep lands and rainforests. Measures should be taken to ensure
appropriate soil and water conservation and restoration in accordance with country LDN targets and
SLM strategies.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 201
4.3 Reversing
human-induced
degradation
©FAO/Giulio Napolitano

Chapter 3 described the significant risks to


agricultural production and food security
from human-induced land degradation. It
highlighted that these are rarely considered
until cropland soils and pastures are lost
or productivity severely compromised due
does it include future water availability for
to human-induced erosion, salinization and
irrigation. Although this analysis has focused
pollution or other degradation processes.
on specific crop options, alternative land-use
Human-induced land degradation constrains
and diversification options could be explored
anticipated growth, particularly on cultivated
in any specific national or territorial context.
and productive land where soil and water
Nevertheless, the analysis provides a good conservation measures are lacking or
indication of future hotspots and bright spots inadequate. Climate change is expected to
at the global and regional scales for growing further affect growing conditions for crops
specific crops, and guides expected shifts and associated livestock and forest systems,
in land suitability. The likely severe socio- and natural ecosystems, particularly in
economic stress resulting from the need to subtropical developing countries. However,
adapt land use and the changes in cropping preventing and reversing degradation will
systems, including knowledge, access to help to build resilience, in line with the
inputs and marketing perspectives, cannot LDN hierarchy, through protection (Avoid),
be underestimated, and so any anticipated conservation (Reduce) and restoration
shift would depend as much on the enabling (Reverse) measures (section 4.3.4).
environment for technology transfer and
Avoiding and reducing degradation, restor-
prevailing socioeconomic circumstances as
ing degraded lands and avoiding associated
on the environmental conditions.
biodiversity loss are crucial to meeting global
This analysis of transitioning to sustainable aspirations for achieving SDGs, including
and diversified land-use systems to address SDG 1 (no poverty), SDG 2 (zero hunger),
degradation and climate trends aligns with SDG 6 (clean water and sanitation), SDG 13
the findings of recent flagship reports such as (climate action) and SDG 15 (life on land). This
The state of food and agriculture (FAO, 2020b) is particularly central to achieving SDG 15:
and The state of food insecurity and nutrition “Protect, restore and promote sustainable
in the world (FAO et al., 2021). In particu- use of terrestrial ecosystems, sustainably
lar, this aligns with the vision of the United manage forests, combat desertification, and
Nations Food Systems Summit, calling for halt and reverse land degradation and halt
joint action for transforming and rebuilding biodiversity loss” (United Nations, 2015).
food systems worldwide to make progress
towards all 17 SDGs.

202 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


4.3.1 Initiatives to
address degradation

© FAO/Giulio Napolitano
International attention has focused on
sustainable land resources management
over the past three decades. This began
with the 1992 United Nations Conference
on Environment and Development and the
Rio multilateral environmental conventions,
Other global initiatives, endorsed by govern-
and the 2012 United Nations Conference on
ments to address degradation, and support
Sustainable Development outcome document
the conventions on biological diversity (CBD)
The future we want (United Nations, 2012).
and climate change (UNFCCC), include:
This focused on achieving a world that is
land degradation neutral. The 2030 Agenda ƒƒ Reducing Emissions from Deforestation
followed in 2015, with the SDG framework and Forest Degradation in Developing
including a dedicated SDG on land. Countries (REDD+), a UNFCCC mechanism
initiated in 2005;
An important global initiative calling for the
restoration of degraded lands worldwide is ƒƒ Aichi Biodiversity Target 15 of the Stra-
detailed in the second edition of the UNCCD tegic Plan for Biodiversity 2011–2020,
global land outlook (UNCCD, 2022a). This adopted under CBD in 2010;511
focuses on conservation, rehabilitation and
sustainable management of land and water ƒƒ the Bonn Challenge on forests, climate
resources in dry lands prone to desertifi- change and biodiversity, launched by
cation. It acknowledges the importance of the Government of Germany and IUCN
land-use planning and secure tenure for in 2011 that focuses on forest landscape
successful implementation. restoration;

ƒƒ the New York Declaration on Forests in


2014, aiming to halve the loss of natural
forests by 2020, and striving to end it by
2030, and restore degraded forests and
reduce carbon losses; and

ƒƒ World Soil Day, held annually on 5 Decem-


ber, raising awareness of the need for
effective partnership in implementing
plans towards sustainable soil manage-
ment and applying voluntary guidelines
for sustainable soil management.
©FAO/Giulio Napolitano

11
“By 2020, ecosystem resilience and the contribution
of biodiversity to carbon stocks has been enhanced,
through conservation and restoration, including
restoration of at least 15 per cent of degraded
ecosystems, thereby contributing to climate change
mitigation and adaptation and to combating
desertification.”

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 203
ƒƒ the Consultative Group on International
Agricultural Research Program on Dryland
Systems, 2007–2017.

Global and national statistics are


©FAO/Benedicte Kurzen/NOOR

being compiled by GSP on the status of


salt-affected soils to develop the first global
soil salinity map directly involving countries
in developing their national maps. The
global map will provide the foundation for
regular monitoring. In 2019, GSP and the
International Center for Biosaline Agriculture

These initiatives are designed to generate established the International Network of

additional benefits through effective partici- Salt-Affected Soils to address soil salinity

pation of the rural poor, women, Indigenous issues at global, regional and national levels

and local communities, civil society orga- and control the increase of salt-affected soils

nizations, and stakeholders from multiple in agricultural areas.

sectors and the private sector.


4.3.2 Sustainable
The United Nations Decade on Ecosystem
land management
Restoration, 2021–2030, is a broad-based
global movement, led by UNEP and FAO, Sustainable land management refers to
to ramp up restoration efforts as a basis the land-use and management actions and
for enhancing livelihoods, counteracting complementary measures (policy, legisla-
climate change and stopping the collapse of tion, institutional reform and technologies)
biodiversity. adapted to the biophysical and socioeco-
nomic contexts to maintain and restore
Several global initiatives focus specifically ecosystem functions and services that land
on dryland sustainable management (see the resources provide for people’s livelihoods
in-focus study on dryland systems at the end and well-being.
of this chapter) and include:
Sustainable land management encompasses
ƒƒ the CBD Programme of Work on the land-use systems and the management prac-
Biological Diversity of Dry and Sub-humid tices of soils, water and biological diversity
Lands, initiated in 2000; by land users for sustained production of
goods to meet changing human needs while
ƒƒ the African Union’s Great Green Wall
ensuring the long-term productive potential
for the Sahara and Sahel Initiative,
of these resources and their environmental
launched in 2007;
functions. Understanding and managing the
interrelations among soil, water, biological
ƒƒ the FAO Dryland Restoration Initiative
resources and the atmosphere are crucial for
Platform, initiated in 2015 (FAO, 2016a);
sustaining the capacity to mitigate and adapt
ƒƒ the IUCN Global Drylands Initiative to climate change.
(drynet, 2022); and

204 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


The focus on SLM for sustainable agriculture Interventions to promote SLM should
and food systems gives due attention to the enable land and water users to enhance and
need for: sustain productivity to meet the increasing
demands of rural and urban populations and
ƒƒ efficient, resilient, inclusive and climate- optimize the economic and social benefits
smart management practices and agri- from the land. This involves: (i) protecting or
food systems; conserving the resource base and ecological
functions, (ii) using the resources in a
ƒƒ balancing the interconnected economic, sustainable manner (reducing or minimizing
social and environmental dimensions of degradation risks) and (iii) restoring or
sustainable development; and rehabilitating degraded resources and
thereby (iv) maintaining or enhancing
ƒƒ minimizing risk and uncertainty in the ecosystem services (Bbox 4.7). The FAO
face of climate change and variability, and strategic framework sums up this response
other shocks. in the agrifood sectors as “better production,
better nutrition, a better environment, and a
better life” (FAO, 2021).

Box 4.7
Sustainable land management objectives
Sustainable land management is key for implementing SDG 15: to “Protect, restore and promote
sustainable use of terrestrial ecosystems, sustainably manage forests, combat desertification, and halt
and reverse land degradation and biodiversity loss”. Target 15.3 aims, by 2030, to combat desertification,
and restore degraded land and soil, including land affected by desertification, drought and floods,
and strives to achieve a world that is land degradation neutral, sustainably manage forests, combat
desertification, and halt and reverse land degradation and halt biodiversity loss.

In considering the linkages among SLM practices to address degraded land, desertification and
drought, climate change adaptation and mitigation, and resulting synergies and trade-offs, SPI specifies:
“SLM represents a holistic approach to achieving long-term productive ecosystems by integrating
biophysical, sociocultural and economic needs and values. SLM is one of the main mechanisms to
achieve Land Degradation Neutrality (LDN).” In its key terms, SPI cites the framework for evaluating
SLM (FAO, 1993).

Sustainable land management combines technologies, policies and activities, and aims to integrate
socioeconomic principles with environmental concerns, to simultaneously: maintain or enhance
production/services; reduce the level of production risk; protect the potential of natural resources, and
prevent soil and water quality degradation; be economically viable; and be socially acceptable.

Ecosystem services include: the supply of nutritious food, fibre, raw materials, energy and drinking water;
water supply regulation; soil formation and nutrient cycling; carbon cycle regulation (sequestration and
emissions); reduction of natural hazards; pest and disease control; and conservation of biodiversity,
cultural heritage, and spiritual and recreational benefits.

Sustainable land management thus contributes directly to SDG 15 (life on land), SDG 2 (zero hunger),
SDG 6 (clean water and sanitation), SDG 7 (affordable and clean energy), SDG 12 (responsible
consumption and production), SDG 13 (climate action) and SDG 3 (good health and well-being), which
is intrinsically linked to the other SDGs.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 205
documenting the experiences and potential
for NbSs for water.

Nature-based solutions offer multiple


benefits (FAO et al., 2020). Analysis by FAO
and The Nature Conservancy (Iseman and
Miralles-Wilhelm, 2021) in agricultural
landscapes includes:

ƒƒ enhancing farmer resilience to increase


©FAO

food production and improve rural liveli-


hoods, through restoring soil health and

4.3.3 Nature-based soil moisture, downstream water supply

solutions and quality, and nutritious food;

Nature-based solutions address societal ƒƒ mitigating and adapting to climate change

challenges through working with nature and through soil, wetlands and forest carbon

biodiversity. First used by the World Bank in sequestration;

2008, NbSs have been highlighted in recent


ƒƒ improving ecosystems and increasing
global assessment reports (IPBES, 2018;
biodiversity and associated benefits; and
IPCC, 2019), and were high on the agenda at
the World Economic Forum and the Climate ƒƒ achieving net-zero environmental
Adaptation Summit in January 2021. impacts in agricultural production and
supply chains.
Nature-based solutions are defined as
“actions to protect, sustainably manage, Policymakers need to address the poten-
and restore natural or modified ecosystems, tial synergies and trade-offs associated with
that address societal challenges effectively NbSs. Concerns about the focus on large-scale
and adaptively, simultaneously providing internationally supported tree planting and
human well-being and biodiversity bene- forest systems as a primary climate miti-
fits” (Cohen-Shacham et al., 2016). They gation solution distract from the need to
are restorative and regenerative by design protect and sustainably manage a range of
and aim to increase productivity and reduce terrestrial and aquatic ecosystems. Atten-
waste, aligning with the principles of the tion is needed to protect resource rights and
circular economy. implement NbSs in ways that respect cultural
and ecological rights (Seddon et al., 2021).
Using NbSs is a potentially valuable strategy
for transforming the agricultural sector into
a beneficiary and a custodian of ecosystems
4.3.4 Land degradation
(FAO, 2018b). Nature-based solutions repre- neutrality
sent effective, long-term and cost-effective
The concept of LDN, introduced at the 2012
interventions to address water management,
United Nations Conference on Sustainable
ecosystem services and soil restoration. The
Development, is designed to support SDG
United Nations world water development report
target 15.3 and avoid loss of natural capital by
2018 (WWAP/UN-Water, 2018) focused on
restoring and rehabilitating degraded lands.

206 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


Parties to UNCCD agreed to “formulate volun- reflecting land-based natural capital. The
tary targets to achieve LDN following their arrow in Fbigure 4.4 indicates that neutral-
specific national circumstances and devel- ity needs maintenance over time through
opment priorities” and to “integrate such land-use planning that predicts losses and
targets in their National Action Programme”. gains. This requires adaptive learning and
The UNCCD SPI provides scientifically based tracking impacts and achievements to enable
guidance for understanding, planning, plans to be continually adjusted and updated.
implementing and monitoring LDN (Cowie
and Orr, 2017). Land degradation neutrality The LDN conceptual framework is appli-

is defined by UNCCD as “a state whereby the cable across all land types, land uses and

amount and quality of land resources neces- ecosystem services. It is implemented at the

sary to support ecosystem functions and landscape scale, considering all land units

services and enhance food security remain of each land type and their interactions and

stable or increase within specified temporal ecological trajectories. This allows for opti-

and spatial scales and ecosystems.” mizing LDN interventions among those land
units to maintain or exceed no net loss at
Fbigure 4.4 captures the LDN vision and how the land-type level (Cowie and Orr, 2017).
best to achieve this by assessing land degra- By 2022, 128 countries had committed to
dation, identifying appropriate management setting LDN targets, more than 100 had set
actions and reporting progress. The objec- them, and many had secured high-level
tive is to maintain and enhance the land government commitment to achieving LDN
resource base, including the stocks of natural (UNCCD, 2022b).
capital associated with land resources and
the ecosystem services that flow from them, A minimum set of three global indicators and

to ensure healthy linkages between human associated metrics are proxies for changes

prosperity and land-based natural capital. in land-based natural capital: land cover

The balance scale illustrates the mechanism (physical land-cover class), land productivity

for achieving neutrality: counterbalancing (net primary productivity) and carbon stocks

future land degradation (losses) with planned (SOC) (Cowie et al., 2018). These are comple-

positive actions elsewhere (gains) within the mentary and universally applicable to allow

same land type. global tracking of progress.

The fulcrum in Fbigure 4.4 illustrates the


hierarchy of responses to Avoid > Reduce
> Reverse land degradation when planning
LDN interventions at landscape level. This
recognizes that prevention is better than
cure, as avoiding land degradation is usually
more cost-effective than efforts to restore
moderately to severely degraded lands.

Neutrality is assessed by monitoring LDN


© Giulio Napolitano

indicators relative to a fixed baseline at the


national level. These indicators include land
productivity, carbon stocks and land cover,

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 207
Figure 4.4 Land degradation neutrality conceptual framework

Source: Orr, B.J., Cowie, A.L., Castillo Sanchez, V.M., Chasek, P., Crossman, N.D., Erlewein, A., Louwagie, G. et al. 2017. Scientific conceptual framework
for land degradation neutrality. A report of the Science-Policy Interface. Bonn, United Nations Convention to Combat Desertification.
https://www.unccd.int/sites/default/files/2018-09/LDN_CF_report_web-english.pdf

208 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


IPCC reports establish the symbiotic relation-
ship between land degradation and climate
change. Solving one problem contributes to
© Giulio Napolitano

solving the other. The reports by IPBES and


IPCC lay out the policy imperative and show
institutional reform and adapted policies can
change incentives that would go a long way
towards making land part of the solution
The precautionary principle of “one-out, rather than part of the problem.
all-out” is applied. If one of the three key
Although some signs of degradation are
indicators shows a negative change, LDN is
easily recognizable (in the field through
not achieved, even if the others are substan-
evidence of erosion and silting, and through
tially positive.
productivity decline, often compensated
These biophysical indicators should be through increased nutrient inputs). Unlike
supplemented by national (or subnational) climate change, there are many reasons
indicators, according to the context, to capture why land degradation has failed to attract
land-based ecosystem services including adequate global attention. Land degradation
indicators of the social and economic impact is a slow-onset process, and people perceive
of LDN on human well-being, such as safe- degradation differently depending on their
guarding land tenure rights, and impacts on relationship with the land. Some see it as
local communities (Cowie and Orr, 2017). an unavoidable side effect of development,
Indicators could measure progress in estab- and others see little urgency when bene-
lishing LDN enabling policies and monitoring fiting economically from exploiting the
systems, and LDN field interventions such as land. Usually, they are not those suffering
the areas of SLM and restoration and rehabil- the consequences of degradation. A simple
itation activities. cause–effect relationship does not exist,
and this makes the issue easy to dismiss.
The LDN concept is ground-breaking in estab- There is also a disconnect between degrada-
lishing an agreed mechanism and mobilizing tion and remedial action. Policymakers and
country commitments to establish a baseline consumers are often unaware or do not feel
and set targets for implementation to protect responsible for land degradation (Willemen
the land from degradation, reduce degra- et al., 2020).
dation processes and rehabilitate degraded
lands under SDG target 15.3.

4.3.5 Reversing the


degradation trend
The IPBES assessment report on land degra-
dation and restoration provides evidence that
© FAO/Marco Longari

land degradation is avoidable, and in many


instances, reversible (IPBES, 2018). This
SOLAW 2021 report, the IPCC special report
on climate and land (IPCC, 2019) and previous

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 209
pilot activities means transitioning from a
time-bound project focus to a long-term
sustained landscape and ecosystem focus.
This requires good governance and competent
institutions for planning at landscape and
regional scales. Scaling up restoration also
requires capacity building supported by
© Seyllou Diallo

appropriate technology, knowledge-sharing,


continual refinement and improvement,
building on progress and experiences, and
infrastructure and sustainable financing.

The analysis in this chapter uses Restoration should be a sustainable

land-suitability assessment and LDN to economic activity, and building confidence

help understand the complex factors that for the long-term requires accountability

drive degradation, from a highly visible and transparency about who is paying

phenomenon that is difficult to measure what costs and who is receiving the various

directly, to one that can be assessed, socioeconomic and environmental benefits.

classified and mapped to identify areas for


remedial action. 4.3.6 Identifying
restoration areas,
mobilizing investments
Sustainable Development Goals set the basis
for creating well-defined and measurable
metrics to guide policy. However, a chal- and strategic vision
lenge remains on integrating the assessment,
One approach to restoration is to identify
monitoring and decision-making processes
target areas where SLM options have a high
for the different SDGs, as the responsibility
potential for success and guide the implemen-
is fragmented among various institutions in
tation and scaling out programme supported
each country.
by appropriate policies and financial mech-
anisms (Vlek, Khamzina and Tamene, 2017).
More information and data, and specific
national and subnational analyses,
The Global Environment Facility (GEF) is
are required to guide national policy
supporting UNCCD implementation through
development and investments to implement
mobilizing investments at country and
effective, cost-efficient and equitable
regional levels to improve data acquisi-
outcomes.
tion and understanding, to develop tools
and strengthen institutional capacities for
Conservation, sustainable use and
planning and policy development, includ-
restoration are best achieved by engaging a
ing extension services for promoting SLM
broad coalition of stakeholders with shared
actions on the ground.
interests that facilitate collective action at
appropriate territorial scales. A transparent
Lessons learned on scaling up policies,
legal environment, a coherent agenda,
investments and actions from the TerrAfrica
sufficient finance and effective incentives
Strategic Investment Programme (SIP) on
are essential to promote action at scale and
SLM in sub-Saharan Africa are available from
to ensure equitable sharing of costs and
the portfolio of 36 projects in 26 countries
benefits. Scaling up restoration from current
(Bb
ox 4.8).

210 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


Box 4.8
Lessons learned from the TerrAfrica STRATEGIC INVESTMENT PROGRAMME
The TerrAfrica SIP was the first opportunity to give a high profile and visibility to the importance of
promoting SLM in Africa. Some USD 150 million of GEF grants mobilized an estimated USD 800 million
of cofinancing for 36 projects in 26 countries between 2010 and 2015, including four transboundary river
basin/watershed projects and four regional thematic projects.

Lessons learned from SIP demonstrate that landscapes may be the most appropriate geographic areas
or territorial units for SLM interventions and investment projects. However, local circumstances should
determine the most appropriate scale, approach and required support mechanisms. The SIP portfolio
highlights include:

• The importance of mainstreaming SLM for food security, poverty reduction and climate change.

• Prospects for SLM are increased when measures are mainstreamed in national policies and laws,
by-laws and regulations enforceable at local level.

• SLM scaling up needs to be flexible and able to react to change from local to global levels.

• Blanket approaches and top-down processes should be avoided.

• People and their actions cause land degradation and need to be at the centre of SLM programmes.
Women represent a large share of direct and indirect beneficiaries and need to be formally
recognized.

• Most SLM technologies in crop and grazing lands contribute to climate-smart agriculture.

• More success is achieved by combining technologies on large areas.

Source: FAO. 2016. Informing future interventions for scaling-up sustainable land management. Rome. www.fao.org/3/i5621e/i5621e.pdf

Forest and landscape restoration received The Great Green Wall for the Sahel and
renewed attention through the global Bonn Sahara Initiative began in 2007 to restore
Challenge, launched by the Government 100 million ha of degraded arid and semi-arid
of Germany and IUCN in 2011, to bring land, sequester 250 million tonnes of carbon
150 million ha of degraded and deforested and create 10 million green jobs by 2030
landscapes into restoration by 2020 and across the Horn of Africa, North Africa and
350 million ha by 2030. Such restoration the Sahel, through a mosaic of green and
seeks sustainability in all land uses in a productive landscape spanning over 8 thou-
given landscape and prioritizes biodiversity sand km2 from Senegal to Djibouti. It supports
conservation and human livelihoods. The communities to expand fertile land, economic
150 million ha milestone for pledges was opportunities for the world’s youngest popu-
surpassed in 2017, through regional initia- lation, food security for millions and climate
tives in Central America and the Caribbean, resilience. Implementation has begun in
Europe, the Caucasus and Central Asia, and more than 20 countries across Africa, with
Asia and the Pacific. support from many partners with pledges
of more than USD 8 billion. Reports indi-
cate that 20 million ha has been restored.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 211
Achieving the 2030 goal will require a faster 400 communities, improving livelihoods for
pace to restore 8.2 million ha annually at an close to 1 million people. A comprehensive
annual cost of USD 3.6 billion (Bb
ox 4.9). restoration approach provides a guide for
scaling up (FAO, 2016b).
Restoration needs and opportunities for the
Great Green Wall for the Sahel and Sahara Numerous studies have attempted to esti-
Initiative were mapped and quantified by mate sustainable use and restoration costs
the global drylands assessment conducted and benefits to ensure viable interventions,
by FAO and partners (FAO, 2016b). FAO has but they have tended to focus on specific
supported field projects through the FAO regions or ecosystems. One study suggests
Action Against Desertification programme. the restoration cost was only 34 percent of
Based on experiences, 50 thousand ha of the cost of inaction (Nkonya et al., 2016).
barren lands has been restored in more than A field study in Madhya Pradesh, India,

Box 4.9
Restoration interventions in the Great Green Wall for the Sahara and Sahel
Initiative
Restoration connects plant science to communities, supplementing tree planting with the cultivation
of fodder for livestock, and deploying mechanization, where appropriate, for water harvesting. It
emphasizes the link between ecology and economics, through developing value chains for non-timber
forest products to generate income for vulnerable rural communities, particularly women, to improve
their livelihoods and resilience. A toolkit supports capacity development for national experts in modern
geospatial technologies for innovative monitoring and evaluation of operations.

Actions include:

• Promoting natural regeneration, in which farmers protect and manage the natural regeneration of
native species in forests, croplands and grasslands (most effective in dry subhumid and semi-arid
zones).

• Investing in large-scale land preparation and enrichment planting where degradation is so severe
that natural vegetation will not regenerate on its own; communities select the native woody and
grass species to be used (mostly arid and semi-arid zones).

• Fighting sand encroachment by establishing and protecting native woody and grassy vegetation
adapted to sandy and arid environments (mostly in the hyper-arid zone).

• Mobilizing high-quality seeds and planting materials of well-adapted native species to build
ecological and social resilience.

• Developing comprehensive value chains that benefit local communities and countries and enable
green economies and enterprises to flourish.

• Building inexpensive, participatory information systems to support baseline assessments, identify


interventions, track progress, inform stakeholders and investors, and aid learning and adaptive
management.

Sources: Liniger, H.P., Mekdaschi Studer, R., Hauert, C. & Gurtne, M. 2011. Sustainable land management in practice: Guidelines and best practice for sub-Saharan Africa.
TerrAfrica, World Overview of Conservation Approaches and Technologies and FAO. www.fao.org/3/i1861e/i1861e.pdf; FAO. 2016. Building Africa’s great green wall:
Restoring degraded drylands for stronger and more resilient communities. Rome. www.fao.org/3/i6476e/i6476e.pdf

212 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


suggests that interventions to build local to communicate a new and transformative
community capacity to implement watershed vision for land management policy, planning
development with climate adaptive measures and practice at global and national scales.
systematically and to maintain the struc-
tures is economically viable and protects the The World Overview of Conservation

ecosystem regenerated for periods of normal Approaches and Technologies (WOCAT) is a

rainfall and extreme events (Das et al., 2020). well-regarded global network among scien-
tists, technical experts and practitioners that
Analysing the economic potential for coastal promotes sharing and use of knowledge to
zone restoration suggests this is expensive support SLM adaptation, innovation and
and, in many situations, not cost-effective decision-making. The global SLM database
in strictly financial terms. However, coastal is updated regularly to support SLM best
mangrove restoration is among the more practices on conservation and restoration.
cost-effective options (Bayraktarov et al., The 1 500 technologies and approaches in
2015; Jakovac et al., 2020). Restoring salt- the multilingual database are supported by
affected soils is economically feasible under a quality control process and tagged to the
some conditions (Qadir et al., 2014). Public LDN hierarchy (Liniger and Studer, 2019;
investments will be required where there is a WOCAT, 2022). Eight consortium partners
public good, particularly for projects initiated (Deutsche Gesellschaft für Internationale
by the private sector and where there are Zusammenarbeit, FAO, International Center
public and private benefits. for Agricultural Research in the Dry Areas,
International Centre for Integrated Moun-
4.3.7 Tools for tain Development, International Centre for

implementation Tropical Agriculture, Swiss Agency for Devel-


opment and Cooperation, the University of
Many resources exist to support countries Bern’s Centre for Development and Environ-
to develop locally adapted SLM and land ment and World Soil Information), WOCAT
restoration programmes at different scales. regional and WOCAT national institutions
Experiences are documented on multicountry and individual members support tool devel-
and transboundary responses, through the opment and piloting, and contribute to the
TerrAfrica and the Great Green Wall for the WOCAT knowledge products.
Sahara and Sahel Initiative, supported under
the GEF land degradation portfolio, the GEF The global GEF/FAO project Decision Support
International Waters programme, and asso- for Mainstreaming and Scaling out Sustain-
ciated transboundary river basin and source able Land Management provides a knowledge
to sea projects. management and decision-support system
and tools from local to national levels. Under
The Global Land Outlook is a UNCCD Secre- this project, the lessons learned in main-
tariat strategic communications platform streaming and scaling up are available in the
with associated publications to demonstrate WOCAT knowledge system to inform wider
the central importance of land quality to SLM and LDN implementation.
human well-being. It focuses on land degra-
dation and land-use change, the driving The FAO Sustainable Forest Management
factors and human impacts, and scenarios for Toolbox provides resources for planning
future challenges and opportunities. It aims forest and landscape restoration, including

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 213
decisions on the appropriate types of ical areas and millions of people. They bring
interventions, institutional arrangements, lost revenue and slow growth, and exacer-
financial considerations and more. The bate long-term food insecurity, poverty and
toolbox also provides case studies of inequality. In developing countries, the cost
successful restoration and rehabilitation of droughts is borne disproportionately by
efforts (FAO, 2022h). the most vulnerable people, which can mean
famine and death. The GAR special report on
The Restoration Opportunities Assessment drought 2021 (UNDRR, 2021) estimated direct
Methodology, produced by IUCN and the annual costs of droughts in the United States
World Resources Institute, provides a flexible of America of USD 6.4 billion, in the Euro-
framework for countries to rapidly identify pean Union the figure is EUR 9 billion, and
and analyse priority areas for forest land- the agricultural productivity in Australia fell
scape restoration at national and subnational by 18 percent in the period 2002–2010 due
levels (IUCN, 2022). to the Australian Millennium Drought. In
India, drought is estimated to cost as much as
The farmer field school (FFS) approach
2–5 percent of the country’s GDP.
has been successful in building capacity to
enable land users to adapt to land manage- Most countries still deal with droughts as
ment practices and SLM. The approach crises in much the same way they approach
combines local and traditional knowledge other natural hazards such as floods and
with modern science and shares experiences earthquakes, and often with little done in
farmer to farmer through improved farmer– the aftermath to prepare for the next one.
extension–research interaction (Bb
ox 4.10). However, emergency action treats only
the symptoms of drought, hunger, famine
These are just a few available tools to help
and water shortages, not the root causes of
plan and implement SLM and restoration
drought impacts. The High-Level Meeting
initiatives for large and small schemes. The
on National Drought Policy in 2013 (WMO,
list will continue to grow as stakeholders
2013) initiated a dialogue on the need for
learn and share experiences.
governments to shift from crisis manage-
ment to drought risk management. This

4.4 Planning approach seeks mitigation and adaptation

for drought
measures that lessen the risks of drought
impacts through planning and by improv-
ing a nation’s resilience and coping capacity

4.4.1 From crisis to (WMO and GWP, 2014). Rather than recovery

risk management
alone, which re-establishes the status quo,
this requires a complete “disaster manage-
Responding to the impacts of drought and ment” cycle (recovery plus protection) as
providing relief and recovery are more shown in Fb
igure 4.6.
complex than other natural hazards (Wilhite,
The High-Level Meeting on National
2011). Droughts are consistently under-
Drought Policy also established the elements
reported (Gall, Borden and Cutter, 2009),
of a national drought management policy
and indirect losses can dwarf direct losses.
to include:
Droughts do not usually affect infrastructure
but they can adversely affect large geograph-

214 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


Box 4.10
Farmers field schools develop SUSTAINABLE LAND MANAGEMENT capacity
In the Kagera basin in East Africa, the FFS approach has been a successful strategy to increase farmer
capacities to manage SLM and water management as part of small-scale watershed management.

A capacity needs assessment established the baseline for training and the knowledge gaps. During
this phase, the causes of land degradation and other production constraints were identified and
documented, and solutions were identified and prioritized with local actors (using WOCAT and LADA1
tools). This defined the FFS learning curriculum and the opportunities and good practices.

A development phase followed to establish FFS sites in microcatchments, select potential facilitators,
train trainers, and develop the curriculum and action plans, including participatory monitoring and
evaluation.

The implementation phase built farmer capacity through the growing season and provided year-long
learning groups and backstopping by facilitators and extension officers/service providers, including
exchanging experiences among districts/provinces and countries. Monitoring and evaluation entailed
follow-up activities, monitoring and fostering adoption, documentation of FFS activities at all levels and
reporting. The monitoring and evaluation phase was continuous during all project phases.

In the final action stage, the FFS development process followed a sequence of activities for
mainstreaming and scaling up, including policy support on territorial planning and tenure security and
resource mobilization through FAO, the International Fund for Agricultural Development and the World
Bank Lake Victoria development programme (Figure 4.5).

Figure 4.5 Farmer field school generic sequence of activities

Participatory diagnosis of constraints and opportunities

Training of trainers/facilitators

Ground working – identification of participants, groups formation etc.

Establishment and running of FFS – season-long farmer learning sessions

Field days and farmer to farmer visits

Graduations

Follow-up activities – farmer to farmer FFS, FFS networks (alumni associations)

Source: FAO. 2017. Sustainable land management (SLM) in practice in the Kagera basin. Lessons learned for scaling up at landscape level. Results of the Kagera Trans-
boundary Agro-ecosystem Management Project (Kagera TAMP). Rome. www.fao.org/3/i6085e/i6085e.pdf

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 215
Figure 4.6 Cycle of disaster management

Risk Management

Planning Monitoring and prediction

Mitigation
Disaster

PROTECTION
RECOVERY

Impact assessment
Reconstruction

Recovery Response

Crisis Management

Source: World Meteorological Organization & Global Water Partnership. 2014. National drought management policy guidelines: A template for action.
Integrated Drought Management Programme Tools and Guidelines Series 1. Geneva and Stockholm. www.droughtmanagement.info/literature/IDMP_
NDMPG_en.pdf

ƒƒ establishing a clear set of princi- The World Meteorological Organization and


ples or operating guidelines to govern the Global Water Partnership initiated the
the management of droughts and Integrated Drought Management Programme
their impacts; in 2014. This proposes a generic ten-step
process to support governments in develop-
ƒƒ promoting standard approaches to
ing national drought management policies
vulnerability and impact assessment;
and plans. Its three pillars – monitoring
and early warning; vulnerability and impact
ƒƒ implementing effective drought monitor-
assessment; and preparedness, mitigation
ing and early warning systems (MEWS);
and response – provide a foundation for

ƒƒ enhancing preparedness and mitiga- planning and management (WMO and GWP,

tion actions; 2014) (Fbigure 4.7).

ƒƒ implementing mitigation, emergency Monitoring and early warning systems

response and recovery measures that (pillar 1) provide a repository for climate

reinforce national drought management data and drought indicators and the capac-

policy goals; and ity to analyse data, assess information, and


communicate it promptly and effectively to
ƒƒ developing a drought plan – the instru- those exposed to drought and who need to
ment through which the above policy take action and reduce risk. Most countries
principles are executed. have systems to monitor hazards like earth-

216 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


Figure 4.7 (Top) Ten steps and (Bottom) three pillars of drought policy and preparedness

Evaluate and revise policy


and supporting plans 10 1 Appoint a national drought
management policy
commission

Develop education
programmes for all age
and stakeholder groups
9 2 State or define the goals
and objectives of a
risk-based national
drought management
Drought policy

3
policy and
8
Publicize the policy
and preparedness
plans; build public
preparedness Seek shareholder
participation; define
awareness and process and resolve conflicts

10 steps
consensus between key water

7
users for sectors

4
Integrate science and
policy aspects of Inventory data and
drought management financial resources
available and broadly
identify groups at risk

Identify research needs


and fill institutional gaps 6 5 Prepare/write the key tenets of
the national drought management
policy and preparedness plans,
following the three-pillar approach

Preparedness, mitigation
and response
Fe
ck

e
a

db
db

a ck
Fee

Drought
policy
Monitoring and Vulnerability and
early warning impact assessment

F ee db a c k

Source: United Nations Office for Disaster Risk Reduction. 2021. GAR special report on drought 2021. Geneva. www.undrr.org/publication/gar-spe-
cial-report-drought-2021; adapted from World Meteorological Organization & Global Water Partnership. 2014. National drought management policy
guidelines: A template for action. Integrated Drought Management Programme Tools and Guidelines Series 1. Geneva and Stockholm. www.droughtman-
agement.info/literature/IDMP_NDMPG_en.pdf and Pischke, F. & Stefanski, R. 2018. Integrated drought management initiatives. In: Drought and water
crises: Integrating science, management and policy, pp. 39–55. CRC Press.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 217
quakes, flooding, storms and forest fires, but for effective and timely intervention. Media
few can detect the early signs of drought and and internet communications can play an
how the event will unfold, to trigger actions important part in bridging this information
and improve proactive responses. gap, but they can do this only when provided
with reliable and timely information.
Most countries lack capacity to monitor and
rapidly communicate real-time conditions, Bbox 4.11 outlines some global and regional
which are essential for dealing with the MEWS and initiatives (in the Caribbean,
impacts on agriculture and food systems. Horn of Africa, Sahel and United States of
They lack data and the capacity to collect and America). However, much more is needed at
process information and communicate this national and local levels to cope with local
circumstances.

Box 4.11
Monitoring and early warning systems
The FAO Global Information and Early Warning System on Food and Agriculture monitors the condition
of major food crops across the globe to assess production prospects. It uses remote-sensing data to
provide valuable insights on water availability and vegetation health during the cropping seasons to
support the analysis and supplement ground-based information. In addition to rainfall estimates and
NDVI, the Global Information and Early Warning System and FAO have developed an agricultural stress
index, a quick-look indicator for early identification of agricultural areas probably affected by dry spells
or drought in extreme cases (see Map 4.14). This map represents one date, but multi-temporal changes
are better to understand agricultural stress areas.

MAP 4.14 Agricultural stress index for 1 May 2021 (%)


ASI (%)

< 10 10 - 25 25 - 40 40 - 55 55 - 70 70 - 85 >=85

off season no data no seasons no cropland

Source: FAO. 2022. GIEWS - Global Information and Early Warning Dotted line represents approximately the Line of Control in Jammu and Kashmir agreed upon by India and Pakistan.
The final status of Jammu and Kashmir has not yet been agreed upon by the parties.
System on Food and Agriculture. In: FAO. Rome.
www.fao.org/giews/en. Modified to comply with UN. 2020. Final boundary between the Sudan and South Sudan has not yet been determined.
Map of the World. https://www.un.org/geospatial/file/3420

218 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


Caribbean
The Caribbean MEWS involves several institutes across the region that collaborate to monitor and
attempt to assess drought severity using indicators and by making short-term rainfall predictions. As
the Caribbean comprises many small islands surrounded by large ocean areas, the region lends itself
to meteorological monitoring and early warning. The Caribbean Drought and Precipitation Monitoring
Network launched in haste in 2009, during a severe drought, and which has proved effective, continues
to make regional rainfall predictions three months and six months ahead using a consensus among 20
organizations from across the region. Such regional forecasts are helpful, but island forecasts would be
ideal as the individual small islands differ in their drought risks.

Source: FAO. 2016. Drought characteristics and management in the Caribbean. FAO Water Reports 42. Rome. www.fao.org/3/i5695e/i5695e.pdf

Horn of Africa
In the Horn of Africa, the Intergovernmental Authority on Development’s Climate Prediction and
Applications Centre was tasked in 2003 with monitoring, predicting and providing early warning of
climate-related disasters, including droughts. The centre is responsible for regional climate outlook
forums to provide consensus early warning seasonal climate information to support the regional
disaster resilience and sustainability strategy frameworks. Generally, however, existing meteorological
stations are far from adequate, their numbers are declining and none exist in Somalia and South Sudan.
Satellite observations complement ground-based systems, but human and institutional capacity to
support this initiative is insufficient.

Source: FAO. 2018. Drought characteristics and management in North Africa and the Near East. FAO Water Reports 45. Rome. www.fao.org/3/CA0034EN/ca0034en.pdf

Sahel
Sahel countries have established a network of national and regional institutions to avert ecological
disasters such as the tragic deaths in the drought of 1968–1973. Central to this is the Permanent
Interstate Committee for Drought Control in the Sahel, which collects and analyses natural resource
data and operates a MEWS to provide alerts of potential drought and locust outbreaks and conducts
socioeconomic research.

Source: FAO. 2018. Drought characteristics and management in North Africa and the Near East. FAO Water Reports 45. Rome. www.fao.org/3/CA0034EN/ca0034en.pdf

United States of America


In the United States of America, a multiagency and cross-ministerial coordination mechanism facilitates
data sharing in real-time informatics related to drought for all sectors under the national drought
policy and supported by the national integrated drought information system. This coordinates drought
monitoring, forecasting, planning and information at national, tribal, state and local levels. Updated
weekly, it shows the location and intensity of drought across the country giving expert assessments
of conditions related to dryness and drought including observations of how much water is available in
streams, lakes and soils compared to usual for the same time of year.

Source: National Integrated Drought Information System. 2022. Advancing drought science and preparedness across the nation. www.drought.gov

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 219
Vulnerability and impact assessments resources. Bbox 4.12 illustrates the case of the
(pillar 2) are essential in guiding MEWS and Horn of Africa, where food security depends
investment in mitigation and adaptation. on smallholder farming and pastoralism,
They address key questions such as: and severe droughts and floods have life-
threatening consequences.
ƒƒ Who is affected by drought?
Preparedness, mitigation and response
ƒƒ What is at risk and why? (pillar 3) comprise measures taken to reduce
adverse drought impacts and respond to
ƒƒ What are the priorities/ranking for deal-
drought emergencies informed by MEWS
ing with them?
and vulnerability and impact assessments.
In turn, mitigation and response determine
Most developing countries in the FAO drought
the critical indicators for MEWS and affect
survey (FAO, 2018b) listed agriculture and
impacts and vulnerability.
smallholder subsistence farming families as
most at risk because they depend on the
Thus, coping with drought relies on all
uncertainties of meagre seasonal rainfall and
three pillars, across which collaboration
rainfed farming for their livelihood.
and continuous information feedback are
essential. Deficiencies in any pillar, as with
The dryland corridor in Central America, the
weakness in one leg of a three-legged stool,
Andean region and Southern Africa are prone
will inhibit the effectiveness of drought
to severe drought, even though the subre-
planning and management.
gions as a whole are well endowed with water

Box 4.12
Responding to crises in the Horn of Africa
About 80 percent of people in the Horn of Africa rely on agriculture and pastoralism as their primary
source of food and income. In 2011, this subregion faced one of the driest years in 60 years, causing
a food crisis that escalated into famine in places, such as southern Somalia. Some 12.4 million people
were in need of urgent assistance. This number nearly doubled in subsequent months. FAO assisted
local populations and governments to respond to the crisis and ensured communities were better
equipped to cope with future droughts.

Heavy rainfall caused severe flooding in Sudan in July 2020, leading to displacement, destruction of
homes, loss of more than 1 thousand ha of agricultural land in the harvest season, and human and
livestock deaths. Hundreds of thousands of people were affected in 17 of the country’s 18 states. The
Nile River reached its highest level in a century.

In Sudan, in response to the COVID-19 pandemic, FAO updated its humanitarian response plan for
2020, working with partners to improve the availability and access to quality and nutritious food to
enhance the resilience of vulnerable people. It continued to provide agricultural and livestock inputs
and animal health support to enable smallholder farmers and pastoralists to maintain their production
and livelihood activities.

FAO provides long-term support to Somalia through the Water and Land Information Management
project to strengthen community resilience using FAO early warning information to improve flood and
drought risk reduction, preparedness and mitigation.

Source: FAO. 2022. FAO in emergencies. In: FAO. Rome. www.fao.org/emergencies/fao-in-action/stories/en

220 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


4.4.2 Formulating a make and implement plans to reduce drought
risks. This process should be an integral part
national drought policy of national water policies, IWRM and the drive
Formulating, developing and implementing to improve water resources management and
a national drought policy requires harmo- increase water security under SDG 6.
nizing policies and legal and institutional
The 2030 Agenda, which calls for an integrated
frameworks and strengthening multisectoral
approach to water resources management,
coordination. Governments need updated
offers opportunities to integrate effective
water laws and guidelines on possible courses
drought preparedness and management.
of action, established contingency plans and
operational modalities. A proactive drought
risk management policy with strengthened 4.4.3 Drought
institutional capacities would lead to more management and
robust planning and investment decisions, its development
with early intervention and mitigation and
There is often a tacit assumption that miti-
less costly damage due to drought.
gation and adaptation are synonymous
Although most countries recognize the need with long-term economic development and
to prepare a national drought policy and investment in water infrastructure and water
national and subnational preparedness plans, security. This is true to some extent. But
many continue to deal with drought as an if economic development were the answer,
emergency response to a meteorological event the United States of America and countries
rather than focusing on subsequent impacts. across Europe would be relatively free of
drought impacts. Recent experiences in both
Few have embarked on the first steps iden- regions demonstrate this is not the case, as
tified in the ten-step process, and even fewer they continue to experience the effects of
have reached a point of putting policy and severe droughts. An encouraging example
plans into practice. One of the main constraints is an initiative taken in 2013 by ten Central
is complacency and the apathy that so often and Eastern European countries to establish
sets in during periods of “normal” rainfall. an integrated drought management plan to
Urgency and action for drought planning are combat severe threats to the region’s agri-
strongest during drought, but when the rains culture (Bb
ox 4.13).
begin again, interest wanes as other, more
immediate issues push drought down the
political agenda. There is often little appe-
tite and funding for gathering and analysing
data, which may take many years and untan-
gling of the effects of drought from other
socioeconomic events. Persistence among
organizations involved will be essential for
effective drought preparedness.
© Giulio Napolitano

Risk management requires all those involved,


from government departments to communi-
ties, to work together, solve problems, and

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 221
Box 4.13
An integrated drought management plan in Central and Eastern Europe
All Central and Eastern European countries have well-developed meteorological and hydrological
monitoring systems, but many do not yet have systems to make good use of the information to support
decision-making in areas like agriculture and energy production. Drought also does not recognize
administrative borders, which adds to the complexity of managing shared water resources and drought
in the region. Most Central and Eastern European countries share water in river basins, such as the
Danube, Sava and Tisza. Several platforms are now in place to encourage information sharing: Sava GIS,
a river commission platform for data sharing, the Drought Management Centre for Southeastern Europe
and Drought Watch (Danube Interreg Programme, 2022).

In 2013, ten Central and Eastern European countries made the first steps towards an integrated approach
to drought management and launched an integrated drought management programme to combat the
growing threat. The first phase (2015–2017) brought together policymakers and stakeholders, including
farmers, from over 40 organizations across the Central and Eastern European countries, to identify
strong and weak areas and examine how they could make plans to improve drought management. The
main achievements of the first phase were (WMO and GWP, 2014):

• a concise overview of the current approaches to drought management in Central and Eastern
European countries;

• a guideline published for preparing a drought management plan that complements the European
Union Water Framework Directive;

• improved communication links among experts and policymakers at the country level;

• increased capacity to implement national drought management plans;

• a collection of existing drought monitoring indices, methods and approaches from the Central and
Eastern Europe region, and the establishment of a link and integration of data into the European
database and monitoring service (European Drought Observatory);

• demonstrated innovative approaches to drought management; and

• exchanges of information and results with organizations in the region that deal with similar issues.

Phase II (2017–2019) focused on building capacity to change ad hoc drought responses into proactive
drought management

Developing drought management plans and putting them into practice is still in its infancy. Clearly, this
is a marathon rather than a sprint. It is also a process and not a project; it has milestones, but there is no
“completion” date. It will be a process of collaboration and continually improving facilities and services
to reduce risks and tackle emergency droughts as they occur. It is about moving from recovery to
protection, from crisis management to risk management.

Source: Bokal, S. & Müller, R. 2018. Integrated drought management in central and eastern Europe. WMO Bulletin, 67(1).

In Southern Africa, concerns about climate for their livelihoods. The Caribbean Disas-
change drive the drought risk agenda, as ter Emergency Management Agency seeks
long-term development plans are put into to integrate disaster management with
place to help protect the many millions of development planning, with clear linkages
smallholders who rely on rainfed farming with planning for climate change. Although

222 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


primarily concerned with responding to
cyclones and flood risks, droughts are now a
recognized hazard requiring more strategic
management in the Caribbean (FAO, 2016c).

4.5 Conclusions

© Joan Manuel Baliellas


This chapter has anticipated that climate
change will bring shifts in land suitability
for all types of cultivation, notably the key
staple crops, and that with the combina-
To address the range of adaptive manage-
tion of changing economic and ecological
ment and attain national emission targets for
conditions, better-informed land-use plan-
agricultural land, it is important to take stock
ning will become the first line of adaptive
of land and water assets and develop realistic
management response.
forms of spatial planning for agricultural
Although climate change has many adverse land use, for which economic trade-offs can
impacts, the analysis also shows opportuni- be evaluated and policies in public subsidies
ties to maintain or increase crop production developed.
and diversify farming systems. Alternative
On-farm operational decisions for agricul-
crops and land-use options are available
ture are manifold if the overall risk to food
to enhance the resilience and adaptation of
production is to be avoided or mitigated by
farmers. Realizing these benefits will largely
transforming agricultural and land manage-
depend on the capacity of supporting services
ment practices. Decisions become more
to guide an informed adaptation process and
complex when calculating how to reduce
farmers’ ability to select suitable options and
negative impacts on livelihoods, human
implement sustainable crop, land and water
health and the delivery of ecosystem services.
management practices. This includes the
Land-use planning and, more broadly, land
conducive enabling environment to support
resources planning are therefore needed
the shift or transformation.
at different levels of decision-making to
address challenges set by changing human
demands. When matched with conformable
SLM options and financing mechanisms, land
resources planning can provide the essential
impetus to reverse trends in land degradation.

The tools for sustainable land-use planning


and management are available to assess the
potential impacts of climate change on crop
production and tackle the growing pressures
©FAO/Asim Hafeez

on freshwater ecosystems and degrading


land, soil and water quality. Monitoring
the accumulated impact of climate change
in relation to agroecological suitability will

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 223
risk management approach can significantly
lessen drought risks and impacts.

Integrated multisectoral approaches need not


be complex, they can be intuitive. However,
solutions require close collaboration across
sectoral boundaries where interests align.
©Picabay/Kangbch

Planning and implementing measures that


sustain productivity, reduce pollution,
sequester carbon and mitigate emissions can
be straightforward, and tested technologies
in SLM can be married with inclusive plan-
prove essential for planning resource use ning approaches at scale when good land and
along the entire food value and supply chains. water governance is in place.
Planning tools can define critical thresh-
olds in natural resource systems, leading Resource planners can now respond to

to the reversal of land degradation when the challenge using remote sensing, big

wrapped up as packages or programmes data and innovative analytical methods

of technical, institutional, governance and that are revolutionizing approaches to

financial support. In this respect, LDN can resources planning. A wide range of resource

help governments set targets and plan inter- planning tools and approaches support

ventions based on the principle of Avoid > decision-makers, planners and practitioners,

Reduce > Reverse land degradation. working at global, national and local levels
to plan, take actions and scale out SLM
Models are now essential tools for land-use options. However, integrated solutions need
planning and LRP, and are increasingly used to be planned at all levels if they are to be
together with participatory approaches to taken to scale.
develop better adapted food and agricultural
systems. Combining LRP tools, including
GAEZ methods, with the latest climate
models provides invaluable insights into
how these changes will redistribute land
available for agricultural production and
affect water availability. This includes shifts
in areas suitable for different crop and
© Daniel Hayduk

livestock species and farming systems, and


identifying potential impacts on productivity
and yield gaps. In particular, shifting to a

224 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


© Jeanette Van Acker
Case study: Participatory land resources
planning to promote sustainable land
management in Morocco

Land degradation poses many challenges on the livelihoods of rural communities in Morocco.
An assessment of land degradation at subnational and landscape/local levels at pilot sites in the
Souss-Massa region has initiated a participatory territorial planning process and action plans to
promote and scale out SLM across the region/country.

Morocco is characterized by scarcity of land and water resources. The agricultural sector is therefore
vulnerable to climate change and impacts on the livelihoods of rural communities and the national
economy. Morocco was selected as one of 14 countries to participate in a project to provide decision
support for mainstreaming and scaling up SLM (FAO, 2018a). The aim was to enhance knowledge and
understanding of land degradation, strengthen institutional capacities and generate decision-support
tools to mainstream and scale out SLM nationally.

Pilot sites were selected at Ameskroud, Aziar and Tamri in the Souss-Massa region according to the
severity of land degradation (low, medium and high). The LADA tools (FAO, 2013a) were used for the
assessment at both levels and enabled the identification and analysis of different forms and severity
of land degradation. The LADA–WOCAT methodology (WOCAT, 2022) at subnational level included
the development of regional maps for land-cover types, land-use systems, main types of degradation,
severity of degradation and identification of good agricultural practices.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 225
At the landscape/local level, the methodology included the identification and mapping of land-use
systems, socioeconomic and biophysical assessments and mapping of good practices in each of
the three pilot sites. Qualitative and quantitative data from the landscape-level assessment were
reviewed with several stakeholders (institutional parties, local authorities and local development asso-
ciations) during a regional consultation workshop to identify, negotiate and select territorial responses,
considering existing plans and implementation mechanisms.

Following the consultations, a participatory SLM territorial planning pact was developed and the
actions agreed between the stakeholders. These comprised: (i) local demonstration areas to test
good practices in each of the three rural communes, (ii) a list of good practices to be implemented
to commit financial and technical input of stakeholders, (iii) territorial watershed or community
approaches developed by FAO and WOCAT and synthesis of the main results, products and lessons
learned to support integration and scaling up of SLM and (iv) simple, measurable indicators for moni-
toring the impacts from implementing good practices, and their degree of adoption.

Capacity was built throughout the project to allow partners to use these tools and approaches to
facilitate scaling up and cultivate ownership of the process and ensure sustainable management of
natural resources.

Following the pilot studies, a three-year action plan was developed according to the LADA–WOCAT
approach involving stakeholder participation from the beginning to scale out SLM in the Prefecture
of Agadir-Ida-Ou Tanane. The plan was aligned with the development plans for the prefecture and
the regional development strategy for the Souss-Mass region and comprised: (i) mitigating the
effects of water erosion, (ii) improving vegetation cover and management, (iii) building capacity and
creating awareness among stakeholders and (iv) promoting SLM. Sustainable land management good
practices (eight practices) were promoted in nine villages, with an overall cost of MAD 180 565 000
(MAD 1 = USD 0.11).

Source: Rouchdi, M., Sabir, M., Qarro, M. & Chattou, Z. 2018. Degradation assessments and good sustainable land management practices within and through their
systems of use, “Souss-Massa region / permanent ecological monitoring and surveillance observatories”. Project Report. Rabat.

226 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


© FAO/Manan Vatsyayana
Case study: Information-based climate-
proof land management in the Lao
People’s Democratic Republic

Agricultural planning in the face of climate change presents a unique challenge because it involves
assessing trade-offs between different land-use strategies now and in the future, based on uncertain
and incomplete information about the nature of the future climate and the state of land resources.

Developing materials to inform planning processes under these conditions is complicated, requiring
a mix of historical data, integrated modelling and scenario building. Developing countries often lack
the human and technical capacity to develop the national-level data, and undertake modelling exer-
cises required to inform sophisticated scenario development exercises and government responses
to climate change impacts on agriculture. Scenario-based analysis is important in contexts involving
uncertainty and complexity to allow for consideration of a wide range of potential future changes in
drivers such as climate change, human population, demand for food and possible trade-offs between
different responses (van Soesbergen et al., 2016; van den Ende et al., 2021). It helps to assess potential
alternative futures (Habegger, 2010; Bourgeois et al., 2012) without giving a false impression of confi-
dence to information users (Nissan et al., 2019).

Tools drawing on widely accepted technical standards and global data and information, such as GAEZ,
have been developed to address the gaps. However, institutional issues can limit the integration and
adoption of such globally oriented tools and data outputs into national decision-making processes.
Many countries will not use global modelling data and outputs in national planning and policy docu-
ments, even if the local capacity to produce similar analysis is available. As a result, despite efforts by
a range of technical advisory agencies to improve agricultural land-use planning processes and tools,
the capacity gaps remain while the risks posed by climate change to agriculture continue to grow and
are poorly understood.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 227
The project Strengthening Agro-climatic Monitoring and Information Systems (SAMIS) to improve
adaptation to climate change and food security in the Lao People’s Democratic Republic is increasing
decision-making and planning capacity for the agricultural sector at national and decentralized levels
in the country. Its objective is to enhance capacities to gather, process, analyse and share climatic
and geospatial information so these can be applied to planning and decision-making (FAO and GEF,
2022). Under SAMIS, the Government of the Lao People’s Democratic Republic has developed land-
suitability and land-use models coupled with climate change projections to produce scenarios
informing decision-making processes.

The process developed through SAMIS is centred around the efforts of local agencies to develop
needed national-level datasets to inform modelling and scenario-building exercises. An annual
agricultural map was prepared using machine learning. The soil map was updated using FAO World
Reference Base classification. Participatory data-collection exercises were conducted at district and
province levels to collect information on land utilization type, crop calendar database and livelihoods.
National climate observation was used to dynamically downscale daily climatic data for the last
30 years and produce statistically downscaled future scenarios. These data inputs are now being used
to drive the development of suitability maps for six crops under current and future climate conditions
using a tailored software, called pyAEZ, developed by FAO and the Asian Institute of Technology
based on the FAO AEZ approach. This effort represents the first nationally led exercise to produce AEZ
analysis using national data by national experts.

The SAMIS project has demonstrated that fast progress in developing land management planning
exercises to address climate change impacts is possible, even in countries with limited technical
capacity. The pyAEZ software has been instrumental in achieving this quick success, and has enabled
the algorithms underpinning the FAO AEZ approach to be openly accessible and run by local oper-
ators with minimal additional technical input or guidance. The SAMIS project has also empowered
technical staff to lead scenario development and modelling exercises by assigning clear roles to staff,
recognizing success and establishing a process of rewarding for technical staff capacities.

From an institutional perspective, the Government of the Lao People’s Democratic Republic has
worked to address barriers to data sharing among agencies. The scenarios developed by SAMIS
required inputs from several sources. The scenario-building exercises developed involved multiple
data producers and coordination, and in some cases, negotiation, among entities at different scales
and levels. The usefulness of most of the data shared was dependent upon the availability of related
datasets held by other agencies. Open and transparent scenario-building exercises that recognize
the power implicit in data management and different data users’ needs helped address sensitive
data-sharing issues.

The datasets and information products developed by SAMIS include policy processes. They regularly
inform planning and policy processes at national and subnational levels. Anticipatory governance for
climate adaptation has been tested (Vervoort and Gupta, 2018). A machine-learning crop monitoring
procedure pioneered by SAMIS is used to validate progress against the National Socio-Economic
Development Plan. The SAMIS tools are also being used to inform crop yield estimation exercises at
village levels to inform prioritization of different investments.

228 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


© GMB Akash
Case study: Restoring degraded
land in Rohingya refugee camp
in Cox’s Bazar, Bangladesh

New geospatial technologies are providing timely and detailed information on natural resources and
SLM in a complex refugee camp setting as part of a humanitarian response. One outcome over the
past three years is the participation of the refugee community in restoring over 350 ha of degraded
land inside the refugee camp in Cox’s Bazar, Bangladesh. This is reducing the risks from natural disas-
ters, and improving ecosystem services and general living conditions inside and around the camps.

Since 2017, there has been a huge increase in the number of Rohingya people displaced, and 742 000
refugees have fled to Bangladesh, which has led to the development of the world’s largest refugee
camp in Cox’s Bazar. This has put intolerable pressure on the regional landscape and is posing chal-
lenges to sustaining human health, food security, nutrition, water supply and sanitation, providing
shelter, education, environmental services and energy, not just for the refugees but also for the host
communities.

Trees have provided fuelwood, and grass covering and soils have been excavated to level the land for
building shelters (UNDP and UN WOMEN, 2018). Land degradation was severe, forests and topsoils
were lost, which intensified surface water runoff, increasing the risk of landslides and flash floods,
putting thousands of people at risk, and provoking conflict between host communities and the
Rohingya refugees.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 229
The initial humanitarian response to land degradation had mixed results, mainly due to lack of
informed decision-making and collaboration. To overcome this, FAO introduced geospatial technol-
ogies and remote sensing to provide information to enable planners to assess land use and inform
resources planning. An example was remote sensing used to illustrate and measure the changes in
land use and vegetative cover on 7 220 ha of degraded forestland in and around the camp area (see
map, which illustrates the changes between February 2017 and February 2018). Experts predicted
that the entire forest area of Cox’s Bazar was likely to disappear by 2019 if the rate of deforestation
continued unabated.

Change of vegetation between February 2017 and February 2018, as depicted


by decreased NDVI (a lower NDVI means less vegetation cover)

February 2017 February 2018

km

Rohingya Refugee Camp

NDVI
High

Low

Location of Cox’s Bazar district

Cox’s Bazar

Source: Mahamud, R., Tanjim, A., Ritu, S., Mondal, F.K. & Arafat, F. 2021.

Since 2018, an integrated approach has evolved that is helping to reverse the degradation, reduce
the risk of natural hazards, and improve the living conditions among refugees and local communities.
FAO, in close coordination with the Energy and Environment Technical Working group and United
Nations organizations (International Organization for Migration, United Nations High Commissioner for
Refugees and World Food Programme), international and national partners, such as the Bangladesh
Ministry of Environment, Forest and Climate Change, local communities and Rohingya refugees have
brought the degraded lands together under a land restoration programme.

This programme has: (i) engaged a range of partners for coordinated planning, implementing and
monitoring land restoration activities, (ii) used evidence-based information to assess gaps and needs,
in particular subsistence issues such as energy supply access and demand, (iii) prepared technical
guidance for land restoration activities including increasing the supply of fuelwood and (iv) used
advanced geospatial technologies and remote sensing to conduct analysis for planning, coordination
and monitoring of land restoration activities.

230 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


The information derived from using the geospatial technologies and understanding the drivers of
land degradation from the beginning of the crisis was key to integrating evidence-based ecosystem
restoration into the humanitarian response plan. The successful implementation and sustainability
of restoration activities in a displacement and emergency setting are highly challenging. After three
years of raising trees in nurseries, stabilizing land and planting trees inside and outside the camps, the
benefits can be seen, on satellite images and on the ground. The WOCAT network also proved to be
an invaluable resource in providing overall guidance for landscape restoration (WOCAT, 2022).

Though every refugee crisis has its own challenges, the approach taken in Cox’s Bazar has potential
for wider application. It must be flexible enough to adapt to rapid changes, collaborative to engage
various stakeholders, coordinated to maximize synergies, and based on robust and documented
evidence for adequate resource allocation.

Watering by a Rohingya refugee at a reforestation site in the camp

© FAO/Saikat Majumuder

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 231
© FAO/Richard Trenchard
In focus: Dryland systems

This focus study highlights the issues facing dryland systems. It looks at their status and trends and
their role in supporting food and nutrition security for billions of people, with attention to the drivers
and pressures, risks and responses. Despite their importance, dry lands are at particular risk. They
face complex challenges of population pressures, unsustainable farming methods, overgrazing and
climate change, leading to land and soil degradation and water scarcity. The required responses and
actions presented here aim to stop and reverse land degradation, and also to sustain and increase
agricultural production, close yield gaps, capture atmospheric carbon in soils, and increase the overall
resilience of communities and ecosystems throughout the dry lands. Many of these issues are not
unique to dry lands, so references here complement those raised in the chapters of this report.

Status of dry lands


The United Nations defines dry lands as lands where the ratio of annual precipitation to mean
annual potential evapotranspiration is less than 0.65 (United Nations, 1992). Dry lands occur on most
continents (Map A). They cover more than 47 percent of the global land surface (6.1 billion ha), with
the largest areas in Australia, China, Kazakhstan, the Russian Federation and the United States of
America. Six countries have at least 99 percent of their area classified as dry and subhumid lands:
Botswana, Burkina Faso, Iraq, Kazakhstan, the Republic of Moldova and Turkmenistan (Mortimore,
2009). A common misperception is that dry lands are “economic wastelands” with low productivity and
are unworthy of investment. They account for about 44 percent of cultivated land and more than half
of the world’s livestock (UNCCD, 2017). A global assessment (FAO, 2019) revealed that dry lands have
diverse land cover and land use (Figure A). They include 27 percent of the world’s forests (1.1 billion
ha), 25 percent of the grasslands and croplands and 28 percent of the barren lands (FAO, 2019). Some
16 percent of dry lands are the “hyper-arid zone”, comprising mainly desert sandy and rocky land-
scapes and hence not suitable for agricultural or forest production.

232 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


MAP A The world’s dry lands

Hyperarid
(P/PET < 0.05)
Arid
(P/PET = 0.05 - 0.20)
Semiarid
(P/PET = 0.20 - 0.50)
Dry subhumid
(P/PET = 0.5 - 0.65)

Excluded presumed drylands

Source: United Nations Environment Programme World Conservation Monitoring Centre. 2007. A spatial analysis approach to the global delineation of
dryland areas of relevance to the CBD Programme of Work on Dry and Subhumid Lands. Cambridge. https://www.unep-wcmc.org/system/dataset_file_
fields/files/000/000/323/original/dryland_report_final_HR.pdf?1439378321. Dataset based on spatial analysis between WWF terrestrial ecoregions (WWF-US,
2004) and aridity zones (CRU/UEA; UNEPGRID, 1991). Dataset checked and refined to remove many gaps, overlaps and slivers (July 2014); based on Miles, L.,
Newton, A.C., DeFries, R.S. Ravilious, C., May, I., Blyth, S., Kapos, V. & Gordon, J.E. 2006. A global overview of the conservation status of tropical dry forests. Journal
of Biogeography, 33: 491–505; and Sörensen, L. 2007. A spatial analysis approach to the global delineation of dryland areas of relevance to the CBD Programme
of Work on Dry and Subhumid Lands. United Nations Environment Programme. www.unep-wcmc.org/system/dataset_file_fields/files/000/000/323/origi-
nal/dryland_report_final_HR.pdf?1439378321. Modified to comply with UN. 2020. Map of the World. https://www.un.org/geospatial/file/3420

Figure A Distribution of land uses in dry lands (thousand ha)

841 271
(14%)

Forest 1 549 706 1 719 004


(25%) (28%)
Other wooded land
Inland water bodies

Other land:
Grassland
71 962
Crops (1%) 68 839
Barren land (1%)
Built up
Other/not identified 208 764
(3%)
582 738
(10%)
1 090 507
(18%)

Source: FAO. 2019. Trees, forests and land use in drylands: The first global assessment. FAO Forestry Paper 184. Rome.
www.fao.org/3/ca7148en/ca7148en.pdf

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 233
Dry lands are characterized by aridity (Figure B), yet they support rich biodiversity and are home to
diverse human cultures, including some of the world’s largest cities (UNCCD, 2017). Some 2.1 billion
people live in dry lands, most of whom depend on forests, grasslands and agricultural areas for their
livelihoods and food security, including income, food, shelter and fuelwood for cooking and heating.
Rural communities in dry lands are often more impoverished than elsewhere, and the land is more
vulnerable to human-induced degradation.

Figure B Distribution of dry lands among aridity zones (Thousand ha)

Hyperarid Arid Semi-arid Dry subhumid

500 000
450 000
400 000
350 000
300 000
250 000
200 000
150 000
100 000
50 000
0 000
ca

a
ca

ia

ia

pe
ric

si

ni

ic

ic
As

As

As
fri
ri

ro
tA

er

er
a
Af
Af

lA

ce
n
n

Eu
Am

m
as
er
rn

er

er
n

ra

lA
er

he
he

st
t

h
nt

es

h
ut

ra
Ea
th

ut
Ce

ut

ut
W

So

nt
or

So

So
So
nd
d

Ce
N

an

la

d
ra
rn

an
nt
te

th
es

Ce

or
W

Source: FAO. 2019. Trees, forests and land use in drylands: The first global assessment. FAO Forestry Paper 184. Rome.
www.fao.org/3/ca7148en/ca7148en.pdf

Dry lands are often considered marginal lands (Bb


ox 4.3), yet this is in terms of mainstream crop
production only. They are home to important rangeland and grazing systems and mixed crop–
livestock systems that rely on short-season drought-resilient crops and receding floodwaters along-
side wetlands and river plains. These have evolved to cope with aridity and drought and provide
invaluable and resilient livelihood systems.

Despite their name, dry lands include globally important watersheds that supply clean water to millions
of people, regulate water flows and mitigate the risks of floods and droughts. Some 15 percent of the
world’s major river basins fall within dry lands (Davies, 2017; Cowie et al., 2018).

The highly variable and unpredictable weather events prevailing in dry lands, including droughts and
floods, have shaped the strong resilience of dryland systems and driven species adaptation. Dryland
capacity to capture and store water, minimize evaporation and increase transpiration determines how
well they function. Examples include how termites in savannahs help maintain soil porosity and recycle
organic matter in the driest and nutrient-poor soils (Davies, 2017). Bacteria in the guts of large-hoofed
herbivores support soil fertility by digesting vegetation and providing manure that accelerates nutrient
cycling on grass growth in the African Serengeti reserve and the Asian Steppes.

234 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


Flora and fauna in dry lands have developed a notable capacity for adapting to periods of water
stress. Many of the grasses, shrubs and trees have acquired deep roots that enhance access to water.
Their leaf form reduces evapotranspiration, and others can store water in their roots and leaves or rely
on dormancy during the dry season. Animals minimize their water loss either through physiological
adaptation or migration to regions that are more humid. Some plants rely on fires (a common hazard
in dry lands) for their reproduction and growth (Davies, 2017).

Risks and emerging issues


Dry lands face numerous challenges linked to population pressures, climate change, land degradation
and desertification, overharvesting, overgrazing and mismanagement of land, soil and water resources.

Approximately one-third of global conservation biologically diverse and seriously threatened hotspots
lie in dry lands (Davies, 2017). The biodiversity in dry lands is well adapted to the harsh conditions
typified by inconsistent rainfall patterns. In many cases, high temperatures and dry lands have fragile
environments that warrant priority attention to avoid irreversible loss of biological diversity.

Population growth
About one-third of the global population inhabits dry lands (UNCCD, 2022b), with 90 percent living
in developing countries (Figure C). The population growth rate is about 18.5 percent, which is faster
than in any other ecological zone. Population density decreases as aridity increases, ranging from
10 people/km2 in deserts to 71 people/km2 in dry subhumid rangeland areas (Lambin et al., 2001;
Mortimore, 2009), including in rural and urban areas. Indeed, some of the world’s largest cities, such
as Cairo, Los Angeles, Mexico City and New Delhi, are located in dry lands, and cities now occupy
about 10 percent of dry lands. As urban growth continues, the land and water available for crop
production will decrease with the likelihood of increasing environmental and socioeconomic stresses
(UNCCD, 2022b).

Figure C Total dryland population per ecosystem type

1 000 000
909 972
900 000 855 333

800 000

700 000
Number of people

600 000

500 000

400 000

300 000 242 780


200 000
101 336
100 000

0 000
Desert Semi-desert Grassland Rangelands

Source: United Nations Decade for Deserts and the Fight Against Desertification. 2020. Why now? In: United Nations 2010-2020 Decade for Deserts and the
Fight Against Desertification. www.un.org/en/events/desertification_decade/whynow.shtml

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 235
Population growth is driving land degradation and desertification, and is increasing demands on
dryland ecosystems to produce food, fuel and fibre. This adds to the general decrease in area of
agricultural land available and compounds the problems of reduced land productivity from declining
soil fertility and water availability.

Climate change
Societies have adapted and prospered in dry lands for centuries by implementing a multitude of
SLM methods. Many can no longer cope with the speed of change, especially changes in the climate
(Mortimore, 2009). Changing land use and practices have led to land degradation and desertification,
water shortages and significant losses in environmental and ecosystem services, as in the extreme
case of the Aral Sea illustrated in Box A.

Dry lands are vulnerable to climatic variability and change, mainly due to rainfall scarcity, where small
changes can have significant impacts. The drying trend in the 1950s is attributed to a 10 percent
increase in global dryland areas, mainly in East Asia, the Sahel in West Africa, Southern Africa and east-
ern Australia. Recent research indicates that the widespread drying trend over the global land surface
from 1950 to 2015 can be partially attributed to global warming. It is estimated that dry areas increased
by about 10.4 percent of global land area between 1950 and 2008. But large areas with a weak wetting
trend still exist, especially at high latitudes in the northern hemisphere. The increasing magnitude and
spatial extent of aridity affect dryland ecosystems’ functional performance (Huang et al., 2017).

Model climate projections illustrate that drying is much more widespread than wetting in the tropics,
subtropics and mid-latitudes, because of increases in evapotranspiration (Map B). Observations and
model simulations have also indicated that rainfall and temperature changes play an important role
as the climate shifts (Huang et al., 2017). The frequency and probability of climatic variables, such as
precipitation and temperature extremes, and their long-term historical trends must be monitored and
evaluated at local scales to help identify suitable types of agricultural practice and crop cultivars. These
measures are essential to mitigate the impacts of climate change and for sustainable development.

MAP B Aridity projections – drier types

Humid to subhumid Semiarid to arid Subhumid/humid to semiarid Subhumid/humid to semiarid Humid to subhumid

Dotted line represents approximately the Line of Control in Jammu and Kashmir
Source: European Commission Joint Research Centre. 2018. Aridity projections – agreed upon by India and Pakistan. The final status of Jammu and Kashmir has
drier types. In: World atlas of desertification. https://wad.jrc.ec.europa.eu/aridityprojections. not yet been agreed upon by the parties.
Modified to comply with UN. 2020. Map of the World.
Final boundary between the Sudan and South Sudan has not yet been determined.
https://www.un.org/geospatial/file/3420

236 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


Box A
The Aral Sea: impacts of land degradation on ecosystems and human health

One of the best examples of environmental degradation with multiple impacts on ecosystems and human
health is the Aral Sea. This is the site of one of the most significant ecological disasters in the world, covering
the five states of Central Asia and affecting almost 50 million people. In the 1900s, the Aral Sea was the world’s
fourth-largest inland lake and an important ecosystem providing natural resources to many communities with
good access to fishing, water and land.

The water level and salinity of the Aral Sea remained stable by inflows of freshwater from two rivers: the Syr
Darya in the east and the Amu Darya in the south. After 1918, policymakers from the former Soviet Union
decided to divert the rivers to irrigate cotton for export. Millions of people from the region were employed, and
the production area was raised from 2.5 million ha to 6.25 million ha within two decades.

In the early 1960s, the Aral Sea began shrinking, and an environmental crisis ensued. It had lost half of its
surface area by 2005. Impacts on the ecosystem within the region included a collapsed fishing industry,
with 60 thousand fishing-related jobs lost, and dust storms from the dried sea bed carrying chemicals and
pesticides originating from the intensive monoculture agriculture occurring along the two rivers leading to
toxic air and water pollution (Akramkhanov et al., 2021).

Moreover, many health impacts emerged: cancers, respiratory diseases, anaemia, miscarriages, maternal
and infant mortality, maternal milk toxicity, kidney and liver diseases, and infectious diseases. The average
life expectancy declined from 64 years to 51 years, and almost half of the population reported emotional
stress. Furthermore, people were forced to migrate due to damaged livelihoods, health and well-being and
increasing unfavourable living conditions. Some remedial measures are under way with good results, but a
large area of the Aral Sea is still disappearing.

The main revival in the North Aral Sea ecosystem, realized between 2011 and 2020, followed completion of
the Kokaral Dam project with the assistance of the World Bank. This made significant improvements to the Syr
Darya River and increased water flow into the Aral Sea.

The second phase of the Kokaral Dam project is anticipated to bring further improvements in employment and
poverty levels, health of locals, environmental quality and overall living standards.

Source: Aladin, N., Chida, T., Cretaux, J.F., Ermakhanov, Z., Jollibekov, B., Karimov, B. & Toman, M. 2017. Current status of
Lake Aral – challenges and future opportunities. Lake ecosystem health and its resilience: Diversity and risks of extinction.
Proceedings of the 16th World Lake Conference, pp. 448–457.
www.zin.ru/labs/brackish/presentations/Current_status_of_Lake_Aral_%E2%80%93_challenges_and_future_opportunities.pdf

Overharvesting and natural resources mismanagement


Mismanagement of natural resources includes transforming rangelands and other silvopastoral
systems into cultivated croplands, inappropriate cultivation and grazing practices, unsustainable
allocation of water resources and wasteful water use, introduction of non-native plants, overharvesting
of fuelwood and wild species, and excessive use of chemical fertilizers for artificial enrichment of
grassland fertility (UN DPA and UNEP, 2015). Such land-use changes and inappropriate practices have
led to widespread human-induced degradation of dryland ecosystems, especially forests, that affects
biodiversity, soil fertility and water availability, and consequently, the livelihoods of local communities
(Mortimore, 2009).

Natural regeneration of soils and vegetation cover in arid areas takes five to ten times longer than
in areas with more regular rainfall. Desertification hotspots, identified by a significant decline in

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 237
vegetation productivity from the 1980s to the 2000s, included 9.2 percent of dry lands and affected
500 million people in 2015 (Mirzabaev et al., 2019). Dry lands cover one-third of the Mediterranean
region, where poor land management, deforestation, overgrazing, natural hazards and resulting
desertification threaten 30 percent of the semi-arid lands (Zdruli, 2014; Ziadat et al., 2022).

Traditional biomass used for cooking and heating by some 2.8 billion people in non-OECD countries
(38 percent of the global population), accounts for more than half of all bioenergy used worldwide
and contributes to land degradation, losses in biodiversity and reduced ecosystem services (REN21,
2018) and 1.9–2.3 percent of global GHG emissions. In hotspots in East Africa and South Asia, land
degradation and deforestation are mainly driven through reliance on open fires, inefficient stoves and
overharvesting fuelwood (Bailis et al., 2015). Excessive removal and use of agricultural wastes and
residues in South and Southeast Asia are due to woody biomass scarcity. Overharvesting wood for
charcoal is fuelling severe deforestation in sub-Saharan Africa (five times the world average).

Responding to the risks


Like in other agroclimatic zones, simple and fragmented responses, such as planting trees, are insuf-
ficient to resolve dryland challenges. A wide range of options is needed to avoid, reduce and reverse
degradation across dryland areas. Many of the actions required for SLM also contribute to climate
change adaptation and mitigation, with further sustainable development cobenefits in poverty alle-
viation and food security (Mirzabaev et al., 2019). Holistic SLM approaches, planned at the landscape
level and tailored to specific socioeconomic and environmental conditions, are needed to stop and
reverse land degradation, close the yield gaps in agricultural production, capture atmospheric carbon
in soils, and increase the overall resilience of communities and ecosystems throughout the dry lands
(UNCCD, 2017). These approaches include soil- and water-conservation measures and sustainable
crop production, livestock-raising practices, and land-use and food system diversification through
adapted species and varieties, agroforestry and agropastoral systems, and associated marketing and
value chain support.

At an international level, global frameworks recognize and understand the urgent need for action.
However, responses are essential at the national and subnational levels to develop policies and
institutional structures founded on integrated, intersectoral land-use planning and SLM. At the
local community, landscape and municipal levels, responses are to implement technical options
such as integrated crop–soil–water management and grazing and fire management that consider
socioeconomic circumstances.

Analysing the interlinkages between land degradation, resource base management and food security
in the Near East and North Africa region offers mitigation and remediation options. These include
knowledge management and sharing, establishment of a regional platform to facilitate dialogue,
public and private investment opportunities, provision of tools to scale out sustainable land and water
management options, and creation of a conducive enabling environment supported by policies and
strategies. This provides policy and decision-makers with priority actions and options to enhance
productivity, and combat land degradation to improve food security in the Near East and North Africa
region (Ziadat et al., 2022).

238 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


The CBD programme of work on dry lands and subhumid lands, initiated in 2000, focuses on the
biodiversity of dryland, Mediterranean, arid, semi-arid, grassland and savannah ecosystems. It applies
an ecosystem approach with attention to water resources management and climate change. There
is interaction with UNCCD for synergy and to avoid duplication of efforts. Activities are carried out
through capacity building, particularly at national and local levels, establishing an international
network of designated demonstration sites and case studies on successful management and partner-
ships among relevant stakeholders.

The FAO Dryland Restoration Initiative Platform aims to enhance measurement of restoration efforts
and country reporting, project analysis, sharing of best practices and successful approaches, and
improvement of efforts of practitioners and decision-makers to address challenges and scaling out.
This builds from the Rome Promise on Monitoring and Assessing Drylands for Sustainable Manage-
ment and Restoration, agreed by FAO and partners in 2015, as a basis for informing sustainable
management and restoration (Box B).

Box B
Rome Promise on Monitoring and Assessment of Drylands for Sustainable
Management and Restoration

In 2014, the FAO Committee on Forestry called for action and investment in dryland assessment,
monitoring, sustainable management and restoration. It requested FAO to undertake a global
assessment of the extent and status of dryland forests, rangelands and agro-silvopastoral systems to
prioritize and target the investments needed for dryland restoration and management.

A workshop in 2015 called for developing more comprehensive and cost-effective methods, including
using existing methods and tools as building blocks and developing new methods integrating
remote sensing and local participation. Through the Rome Promise, participants agreed to: (i) form an
open-ended collaborative network or community of practice to advance monitoring and assessment
of dry lands, including an understanding of their users; (ii) communicate the value and importance
of dryland monitoring to relevant stakeholders, including policymakers and resource partners; and
(iii) develop a dynamic road map for collaborative action.

The first FAO global assessment of dry lands was among the initial steps in implementing the Rome
Promise, building a robust baseline for future monitoring, which will support countries in their efforts
to develop needed strategies and identify appropriate investments for sustainable management of
dry lands. Results demonstrate that dry lands are productive landscapes with considerable economic
potential and environmental value.

This is a step towards regular monitoring of changes in dry landscapes, which is vital to evaluate the
impact of climate change, human activities and the results of adaptation and mitigation measures and
progress towards meeting regional LDN targets. The process should include assessing the effects
of different governance frameworks, policies and legislation related to land use for more effective
support in improving the livelihoods and climate change resilience of dryland populations.

Source: FAO. 2015. Drylands monitoring week 2015: The Rome Promise on Monitoring and Assessment of Drylands for Sustainable Management and Restoration. Rome.
www.fao.org/3/a-i5600e.pdf

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 239
The IUCN Global Drylands Initiative612 supports adapted ecological assessments for targeted monitor-
ing of dryland conditions and trends, strengthening sustainable land and ecosystem management
governance. It established an agreement with UNCCD in 2015 to support progress towards policies
and programmes that deliver LDN by applying NbSs at national and subnational levels.

The Consultative Group on International Agricultural Research Program on Dryland Systems was
a global agricultural research partnership from 2007 to 2017 to reduce the vulnerability of poor,
marginalized, dryland communities, to sustainably intensify agriculture for improved food security and
income, and to develop more equitable and sustainable management of land and natural resources.
It targeted 1.6 million smallholders in dry lands in the Sahel and dryland savannahs of West Africa,
North Africa, East and Southern Africa, West Asia, South Asia and Central Asia. Research efforts in
dry lands continue through restructured global agrifood systems research programmes, notably
Water, Land and Ecosystems, Climate Change Agriculture and Food Security and specific-crop-based
programmes (CGIAR, 2022).

Implementing land degradation neutrality


Public policy response
Initially, the LDN concept targeted dry lands as the most vulnerable to land degradation. However,
during the twelfth session of the UNCCD Conference of the Parties in 2015, LDN was adopted as a
global concept. In 2015, it became SDG target 15.3, which states that the global community shall “By
2030, combat desertification, restore degraded land and soil, including land affected by desertification,
drought and floods, and strive to achieve a land degradation-neutral world” (United Nations, 2015).

Pursuing LDN will be essential to achieving the SDGs related to food security, environmental protec-
tion and sustainable natural resources (Gilbey, 2018). Dry lands can greatly benefit from this global
drive to achieve the LDN target of “no net loss” through a dual-pronged approach to avoid or reduce
land degradation, combined with measures to reverse already existing desertification and land degra-
dation via restoration, such that losses are balanced by gains (Safriel, 2017).

Successful SLM and achieving LDN targets in each country must be founded on integrated and
intersectoral land-use planning. Integrated land-use planning can reconcile LDN and other targets
through a policy process that leads to desirable land use (UNCCD, 2019).

Inclusive and responsible land governance is required to ensure the development of effective policy,
legal and organizational frameworks that secure land tenure and foster sustainability, and improve
livelihoods and well-being. This should include governance over water resources, agricultural land,
rangeland, forestland and other uses of fragile dryland resources such as mining, urban expansion
and tourism. Ensuring that dry lands are under secure tenure, with recognized and safeguarded rights
to use and manage land, provides land managers with the freedom and legitimacy to implement SLM
strategies (Davies, 2017). Policies that promote land tenure security for all legitimate land tenure rights
should be designed to support increasing agricultural productivity and incomes, including sedentary
and nomadic livestock production, as well as minimize random appropriation of land, for example, for
large-scale commercial farms, especially in communally managed lands.

12
A comprehensive list of regional and national programmes on dryland related issues is available (drynet, 2022) .

240 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


There is a need to improve understanding of formal and informal legitimate land tenure rights and
develop innovative solutions that bring together statutory law and customary rights or practices. In
many countries, land tenure and planning processes are administered locally to enable greater partic-
ipation in local-level decision-making and to enable more respect for local rights and responsibilities.
Strong local institutions are a vital interface between statutory and customary legal systems, and are
key to combining improved local governance and access to markets and other services (UNCCD, 2019).

Policies to improve access to markets help farmers and livestock producers to increase profits and
encourage the adoption of SLM practices. Policies that promote payments for ecosystem services
(PES) or other incentive measures for investing in SLM provide incentives to restore degraded land or
increase ecosystem services (Garrett and Neves, 2016). However, individual landowners usually under-
invest in SLM as they are unable to reap the full benefits. Incentives for ecosystem services provide a
mechanism to transfer some of these benefits to landowners and stimulate further investment in SLM
(UNCCD, 2019). Effective incentives for ecosystem services/PES schemes (FAO, 2022i) depend on land
tenure security and appropriate policy design that considers specific local conditions, with equity and
justice in distributing the benefits. Decentralized approaches that provide local communities with a
larger role in the decision-making process can also improve the impact of PES (Mirzabaev et al., 2019).

Promoting renewable energy resources can help reduce deforestation by populations in developing
countries dependent on traditional biomass, especially fuelwood and charcoal, for their energy needs.

Policies that empower women and secure their land rights also enhance SLM. They target changes in
customary norms and practices that undervalue women and efforts to safeguard women’s equitable
access to knowledge, support services, markets and resources. Policies that promote education and
capacity building and expand access to information and agricultural services with attention to gender
tend to accelerate the adoption of SLM practices

The private sector’s role


The increase of large- and medium-scale commercial farms (see Chapter 2) with stronger bonds to an
integrated value chain has amplified the role of the private sector and corporate organizations in land
management and agricultural production (UNCCD, 2019; Debonne et al., 2021). Corporate investments
are channelled towards large-scale land acquisition. There are concerns about the potentially nega-
tive impacts on the environment, national economies, social welfare and human right to land tenure
(Baker-Smith and Attila, 2016). Smaller-scale land transfers also lead to unplanned and unregulated
changes in land use (Davies, 2017). Such investment and land governance arrangements can threaten
global and regional LDN targets when using highly intensive and unsustainable modes of production.
Private sector involvement must consider LDN initiatives to ensure progress and success at scale in
implementing SLM and land restoration. Greater attention is needed to ensure the right investors are
attracted to suitable investments and respect the principles for agricultural investments (CFS, 2014),
building on VGGT through a consensual approach, respecting rights, livelihoods and the environment
(see Chapter 5).

Although there are concerns over private sector actions, investment is also flowing from the business
sector to support land restoration and funding from national governments, international organizations
and local communities. More corporations are moving towards SLM practices, using agricultural train-

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 241
ing, ecocertification and other instruments (FAO, 2015). Although welcome, investments are far short
of what is needed. Governments must increase efforts to mobilize private investment that supports
existing land users to improve their land management and develop public land-use plans effective
at a landscape level. This can support integrated crop farming, grazing, forest management, wildlife
management and wetlands protection (Davies, 2017).

Making dry lands more attractive for private sector investment can be achieved through effective
policies, regulation, incentives and technical measures. Developing innovative and productive part-
nerships between the private sector and local communities can also help create an environment that
fosters private sector investment (FAO, 2015). Non-governmental organizations (NGOs) also have an
important role in accessing innovative financing sources seeking attractive and diverse returns. An
example addresses the gap in national and international forest restoration financing and translating
investments into practical action that safeguards the ecosystem and people’s livelihoods (Gutierrez
and Keijzer, 2015).

Examples of private sector participation in LDN initiatives include the Great Green Wall for the Sahara
and Sahel Initiative, involving working with national governments, civil society and development orga-
nizations under pan-African coordination to halt land degradation. The UNCCD Global Mechanism is
supporting the development of sustainable value chains. It works with the private sector in the Sahel
and guarantees dryland products, leading to thousands of new land-based jobs for rural women in
the region.

Smallholder farmers are also active members of the private sector, and their small-scale investments
are central to achieving sustainable farming practices. Farmers are the largest investors in developing
country agriculture and their capacity to invest needs to be strengthened (CFS, 2014). Dryland farmers
and pastoralists invest in many ways on a small scale, but this is multiplied thousands of times across
a landscape. These investments can be difficult to evaluate, but they represent a diverse financial,
labour and social capital portfolio that generates a wide array of revenues in food, insurance and
ecosystem services (Davies, 2017). A particular effort is needed to mobilize local entrepreneurs and
develop small- and medium-sized enterprises on farms and along the supply chains to strengthen
and diversify rural livelihoods (Davies, 2017).

Local-level sustainable land management actions


Proven nature-based approaches at the landscape and farm levels exist that encourage sustainable
agriculture, enhance ecosystem services, decrease land degradation and increase resilience among
vulnerable communities, some of which are outlined below.

Integrated crop–soil–water management


Integrated cropland management is a long-established and continuing practice in dry lands (Mirz-
abaev et al., 2019). Actions include diversifying crop species and drought-resilient and ecologically
appropriate plants, reducing or avoiding tillage, maintaining healthy vegetation and mulch cover,
applying organic compost and fertilizers, and adopting water-conserving irrigation practices.

Changing agronomic practices, such as adopting intercropping and relay cropping, using drought-
tolerant species and varieties, and minimizing tillage, help to reduce soil loss, maintain soil cover and

242 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


improve soil health. Different forms of agroforestry and shelterbelts help to reduce erosion, improve
soil conditions and maintain SOC (FAO, 2013b).

Rainwater harvesting is receiving increasing attention for bridging short dry spells, and thus decreas-
ing risk in rainfed agriculture (Wani, Rockström and Oweis, 2008). However, such techniques generally
do not protect crops from the long dry spells that lead to crop failure. A global assessment suggested
that rainwater harvesting increased crop production by an average of 78 percent. Although care is
needed to avoid erosion and impacts downstream, capturing runoff may reduce the amount of water
available to those farmers who traditionally rely on the flow downstream.

Sustainable rangeland management


Rangelands account for 69 percent of dry lands or some 4.2 billion ha. In sub-Saharan Africa, range-
lands are estimated to feed over 55 percent of Africa’s livestock and provide a major source of income
to 268 million pastoralists and agropastoralists. However, rangelands face complex challenges of
degradation, mobility, conflict, access to markets and so forth.

The report Sustainable rangeland management in sub-Saharan Africa: Guidelines to good practice
(Liniger and Studer, 2019) was prepared for the TerrAfrica partnership and documents 30 case studies.
They cover a diverse range of practices and systems from small-scale settled pasture to bounded
rangelands with wildlife management and pastoral rangelands (Box C). The research and experiences
throughout the region highlight the importance of integrated land and water management.

Box C
Sustainable land management practices in rangelands in
sub-Saharan Africa

The following practices were identified:

Enabled mobility, including improved access, involves practices that assist grazing over large/diverse
areas to seek forage and water using traditional knowledge and innovations and new technologies
(e.g. satellite image analysis and early warning systems at large scales).

Controlled grazing, including seasonal grazing, involves enclosures, physical or social fencing,
rotations, grazing reserves (fodder banks), regulating grazing and mobility.

Range improvement involves management of fire, grazing quality, soil fertility and moisture.

Supplementary feeding for increased milk and meat production and as a life-saving strategy during
an emergency such as drought may involve fodder collection within or outside the rangeland areas
including production or buying of processed or compound feed.

Infrastructure improvement includes water points and macrocatchments, floodwater spreading, soil-
and water-conservation trenches, protecting drinking water quality, livestock corridors, access roads
and transport routes of animals and feed.

Source: Liniger, H.P. & Studer, M. 2019. Sustainable rangeland management in sub-Saharan Africa: Guidelines to good practice. TerrAfrica, World Bank, World
Overview of Conservation Approaches and Technologies, World Bank Group and Centre for Development and Environment and University of Bern. www.wocat.
net/library/media/174

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 243
Sustainable grazing approaches and revegetation increase rangeland productivity. However, they
require pastoralists to carefully manage rangelands to avoid overgrazing and fire by changing
frequency and intensity of use. Controlled fire is an essential component of rangeland management.
It encourages fresh growth of pastures and removes waning and inedible forage, exotic weeds
and woody species that harbour parasites (Davies, 2017). Grazing and fire regimes determine the
relative abundance of trees versus grasses and the health of species richness and basal cover within
grasslands, savannah and woodland areas. This affects levels of soil erosion, soil nutrients, secondary
production and additional ecosystem services. Although fire has a lower impact on SOC and soil
nutrients than grazing, elevated fire frequency does increase SOC and nitrogen loss (Mirzabaev et al.,
2019). A context-specific evaluation of grazing and fire influences on particular species ensures the
persistence of target species over time.

Proactive management to prevent land degradation by changing grazing systems or clearing bush
encroachment can be more cost-effective than restoring already degraded land. Drought forecasting
and contingency planning can also help to reduce land degradation. Intensive bush encroachment is
a form of desertification, but some levels of encroachment may lead to a net increase in ecosystem
services, preserve fodder production, and increase wood production and associated products (Mirz-
abaev et al., 2019).

Reviving traditional Indigenous practices


Indigenous and local knowledge can enhance the success of SLM and address desertification and
land degradation. Building Indigenous knowledge among dryland communities and combining it with
modern scientific knowledge and methods can bring significant benefits.

FAO global guidelines for restoring degraded forests and landscapes in dry lands are starting
points. Such guidance can help to adapt and develop, rather than replace, local communities’ tested
management strategies and enhance knowledge on how to adapt to changing and unpredictable
climates (Mortimore, 2009).

Other practices in dry lands


Dune stabilization techniques and building palisades to prevent the movement of sand and reduce
sand deposits on infrastructure can reduce SDSs. Calcium bentonite or silica gel can stabilize mobile
sand, and permanent plant cover using pasture species can improve grazing at the same time. When
dunes are stabilized, suitable woody perennials can be planted to stabilize the soils (Mirzabaev et
al., 2019).

Halophytes or salt-tolerant crops offer an alternative with high economic potential on land where
salinity is a problem. The use of saline land and water in biosaline agricultural production may provide
an attractive alternative. The biomass can be useful for forage, food, feed, essential oils, timber, fuel-
wood and biofuel, and can enhance terrestrial carbon stocks (ICBA, 2021).

Catalysing sustainable land management adoption


As part of a land and water planning process, socioeconomic and policy responses can be a catalyst
for adopting SLM practices that interact to achieve LDN. Collective community action for natural

244 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


resources management at territorial level supported by reliable networks and mutual trust and
cooperation in and among communities can also favour SLM practices. Encouraging farmers and
pastoralists to adopt SLM practices can be more successful than introducing external technologies.
Peer-to-peer mutual learning and sharing expertise and experience can positively contribute to better
technology adoption (Mirzabaev et al., 2019).

Global evidence based on the past 30 years suggests that USD 1 invested in restoring degraded lands
yields USD 3–6 in social returns, including ecosystem services. Despite these returns, the take-up of
SLM practices remains relatively low, as many social benefits are intangible. Economic and institu-
tional barriers also exist that seriously limit SLM strategies.

Agricultural communities in dry lands also depend on diversification into non-farm employment,
including through migration and improved marketing and alternative incomes. Such activities can
improve livelihoods and provide the finance for investment in SLM (Mirzabaev et al., 2019). Wildlife
management and tourism are opportunities for the conservation and sustainable use of biodiversity
and income generation. Investment in infrastructure may be needed to improve access to water
resources (surface and groundwater) and markets (Liniger and Studer, 2019).

Looking forward
Enhancing land and water productivity, reversing land degradation, and coping with water scarcity
and drought are all crucial for achieving food security, sustainable agriculture and SDGs in dry
lands. Technical options for SLM provide promising solutions for various land users to reduce and
reverse degradation and enhance productivity and livelihoods through improved water, land and soil
resources, and ecosystem management. Options are also available to advance land restoration in
high-potential areas.

However, participatory resources planning at community level supported by technical services is also
needed to identify potential practices to suit the prevailing socioeconomic and biophysical conditions
and adapt to climate change, coupled with a favourable enabling environment through policy support
and financial/investment mechanisms, including private sector partnerships, to enhance uptake and
continuous adaptation of improved practices and strengthen preparedness to current and future
challenges.

Finally, land tenure security, including rights of access to water, through improved governance and
land administration is also a critical part of the enabling environment, as a lack of clear land rights
hinders all public and private investments in sustainable water, land and soil management, as elabo-
rated in Chapters 2 and 5.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 245
References
Akramkhanov, A., Strohmeier, S., Yigezu, Y.A., Haddad, M., Smeets, T., Sterk, G., Zucca,
C. et al. 2021. The value of landscape restoration in Uzbekistan to reduce sand and dust
storms from the Aral seabed. Washington, DC, World Bank. https://openknowledge.worldbank.
org/handle/10986/36461?locale-attribute=fr

Aksoy, E., Arsov, S., Mincev, I. & Fang, C. 2020. Agro-ecological atlas of the Republic of North
Macedonia. Rome, FAO. www.fao.org/publications/card/en/c/CA7519EN

Bailis, R., Drigo, R., Ghilardi, A. & Masera, O. 2015. The carbon footprint of traditional woodfu-
els. Nature Climate Change, 5: 266–272.

Baker-Smith, K. & Attila, S.B.M. 2016. What is land grabbing? A critical review of existing definitions.
Romania, Eco Ruralis. www.farmlandgrab.org/uploads/attachment/EcoRuralis_WhatIsLand-
Grabbing_2016.pdf

Bayraktarov, E., Saunders, M.I., Abdullah, S., Mills, M., Beher, J., Possingham, H.P., Mumby,
P.J. & Lovelock, C.E. 2015. The cost and feasibility of marine coastal restoration. Ecological
Applications, 26(4): 1055–1074.

Bindoff, N.L., Stott, P.A., AchutaRao, K.M., Allen, M.R., Gillett, N., Gutzler, D., Hansingo, K. et
al. 2013. Detection and attribution of climate change: From global to regional – Supplementary
material. In: T.F. Stocker, D. Qin, G.-K. Plattner, M. Tignor, S.K. Allen, J. Boschung, A. Nauels
et al., eds. Climate change 2013: The physical science basis. Contribution of Working Group I to the
Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Geneva, Intergov-
ernmental Panel on Climate Change. www.ipcc.ch/site/assets/uploads/2018/07/WGI_AR5_
Chap.10_SM.pdf

Borrelli, P., Robinson, D.A., Panagos, P., Lugato, E., Yang, J.E., Alewell, C., Wuepper, D.,
Montanarella, L. & Ballabio, C. 2020. Land use and climate change impacts on global soil
erosion by water (2015-2070). Proceedings of the National Academy of Sciences of the United
States of America, 117(36): 21994–22001.

Bourgeois, R., Ekboir, J., Sette, C. & Egal, C. 2012. The state of foresight in food and agriculture
and the roads toward improvement. Conference presentation at Second Global Conference
on Agricultural Research for Development, 2012. https://agritrop.cirad.fr/570486/1/docu-
ment_570486.pdf

CFS (Committee on World Food Security). 2014. Principles for responsible investment in agriculture
and food systems. Rome, FAO. www.fao.org/3/au866e/au866e.pdf

CGIAR. 2022. Research portfolio 2017–2021. In: CGIAR. www.cgiar.org/research/research-portfolio

Cohen-Shacham, E., Walters, G., Janzen, C. & Maginnis, S., eds. 2016. Nature-based solutions
to address global societal challenges. Gland, International Union for Conservation of Nature.
https://portals.iucn.org/library/sites/library/files/documents/2016-036.pdf

246 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


Cowie, A. & Orr, B. 2017. The scientific conceptual framework for land degradation neutrality. Bonn,
United Nations Convention to Combat Desertification. www.unccd.int/sites/default/files/rele-
vant-links/2017-09/CST13_Item2a_Cowie_Orr_LDN conceptual framework_0.pdf

Cowie, A.L., Orr, B.J., Castillo Sanchez, V.M., Chasek, P., Crossman, N.D., Erlewein, A., Louwa-
gie, G. et al. 2018. Land in balance: The scientific conceptual framework for land degradation
neutrality. Environmental Science & Policy, 79: 25–35.

Danube Interreg Programme. 2022. Drought Watch. https://droughtwatch.eu/

Das, S., Duraisamy, V., Yaduvansh, Ia., Shinde, A., Ankita, Y., Solanky, V., D’Souza, M., Jatav, Y. &
Garg, R. 2020. Economic valuation of reducing land degradation through watershed development in
east Madhya Pradesh under risks of climate extremes. Pune, Watershed Organisation Trust. www.
eld-initiative.org/fileadmin/pdf/WOTR_2020_ELD_Madhya_Pradesh_Report_final.pdf

Davies, J. 2017. The land in drylands: Thriving in uncertainty through diversity. Global Land Outlook
Working Paper. Bonn, United Nations Convention to Combat Desertification. https://knowl-
edge.unccd.int/publication/land-drylands-thriving-uncertainty-through-diversity

Debonne, N., van Vliet, J., Metternicht, G. & Verburg, P. 2021. Agency shifts in agricultural
land governance and their implications for land degradation neutrality. Global Environmental
Change, 66: 102221.

drynet. 2022. Useful links. In: drynet: A global initiative, giving future to drylands. https://dry-net.
org/about-us/useful-links

FAO. 1993. Guidelines for land-use planning. FAO Development Series 1. Rome. www.fao.org/3/
t0715e/t0715e00.htm

FAO. 2013a. Land degradation assessment in drylands: Methodology and results. Rome. www.fao.
org/3/a-i3241e.pdf

FAO. 2013b. Climate-smart agriculture source book. Rome. www.fao.org/3/i3325e/i3325e.pdf

FAO. 2015. Natural capital impacts in agriculture: Supporting better business decision-making. Rome.
www.fao.org/fileadmin/templates/nr/sustainability_pathways/docs/2015-11-19_Natural_
Capital_Impacts_in_Agriculture-Supporting_Better_Business_Descision-Making_v8.pdf

FAO. 2016a. Dryland Restoration Initiative Platform. Rome. www.fao.org/3/i6509e/i6509e.pdf

FAO. 2016b. Building Africa’s great green wall: Restoring degraded drylands for stronger and more
resilient communities. Rome. www.fao.org/3/i6476e/i6476e.pdf

FAO. 2016c. Drought characteristics and management in the Caribbean. FAO Water Reports 42.
Rome. www.fao.org/3/i5695e/i5695e.pdf

FAO. 2018a. Decision support for mainstreaming and scaling up of sustainable land management
(DS-SLM). Rome. www.wocat.net/documents/462/DS-SLM_fact_sheet_pr00ihQ.pdf

FAO. 2018b. Drought characteristics and management in North Africa and the Near East. FAO Water
Reports 45. Rome. www.fao.org/3/CA0034EN/ca0034en.pdf

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 247
FAO. 2019. Trees, forests and land use in drylands: The first global assessment. FAO Forestry Paper
184. Rome. www.fao.org/3/ca7148en/ca7148en.pdf

FAO. 2020a. Strengthening civic spaces in spatial planning processes. Governance of Tenure Techni-
cal Guide 12. Rome. https://doi.org/10.4060/cb0422en

FAO. 2020b. The state of food and agriculture 2020. Overcoming water challenges in agriculture.
Rome. https://doi.org/10.4060/cb1447en

FAO. 2021. FAO strategic framework. In: Office of Strategy, Programme and Budget (OSP). Rome.
www.fao.org/about/strategy-programme-budget/strategic-framework/en

FAO. 2022a. Toolkit for the application of Green Negotiated Territorial Development (GreeNTD).
In: Land & Water. Rome. www.fao.org/land-water/land/land-governance/land-resourc-
es-planning-toolbox/category/details/en/c/1047635

FAO. 2022b. Participatory and Negotiated Territorial Development (PNTD). In: Land & Water.
Rome. www.fao.org/land-water/land/land-governance/land-resources-planning-tool-
box/category/details/en/c/1043145

FAO. 2022c. Sustainable land management. Rome. www.fao.org/3/i4593e/i4593e.pdf

FAO. 2022d. Land Resources Information Management Systems (LRIMS). In: Geospatial infor-
mation for sustainable food systems. Rome. www.fao.org/geospatial/resources/tools/lrims/en

FAO. 2022e. The Hand-In-Hand Initiative Data Platform. In: FAO. Rome. https://data.apps.fao.org

FAO. 2022f. The state of the world’s land and water resources for food and agriculture 2021:
Systems at breaking point. In: Land & Water. Rome. www.fao.org/land-water/solaw2021/en

FAO. 2022g. Projects. In: Geospatial information for sustainable food systems. Rome. https://www.
fao.org/geospatial/projects/en

FAO. 2022h. Sustainable Forest Management (SFM) toolbox. In: FAO. Rome. www.fao.org/sustain-
able-forest-management/toolbox/en

FAO. 2022i. Incentives for ecosystem services. In: Land & Water. Rome. www.fao.org/land-wa-
ter/overview/integrated-landscape-management/incentives-for-ecosystem-services/en

FAO & GEF (Global Environment Facility). 2022. Strengthening Agro-climatic Monitoring and
Information System (SAMIS). In: FAO. Rome. www.fao.org/in-action/samis/en

FAO & IIASA (International Institute for Applied Systems Analysis). 2021. Global Agro-
Ecological Zoning version 4 (GAEZ v4). In: FAO. Rome. www.fao.org/gaez/en

FAO, IIASA, International Soil Reference and Information Centre, Institute of Soil Science –
Chinese Academy of Sciences & JRC. 2012. Harmonized world soil database (version 1.2). https://
esdac.jrc.ec.europa.eu/ESDB_Archive/Soil_Data/Docs_GlobalData/Harmonized_World_
Soi_Database_v1.2.pdf

248 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


FAO, International Fund for Agricultural Development, United Nations Children’s Fund &
World Health Organization. 2021. The state of food security and nutrition in the world 2021.
Transforming food systems for food security, improved nutrition and affordable healthy diets for all.
Rome, FAO. https://doi.org/10.4060/cb4474en

FAO, Intergovernmental Technical Panel on Soils, Global Soil Biodiversity Initiative, Conven-
tion on Biological Diversity & European Commission. 2020. State of knowledge of soil biodiver-
sity: Status, challenges and potentialities. Rome, FAO. https://doi.org/10.4060/cb1928en

Fischer, G., Nachtergaele, F.O., van Velthuizen, H.T., Chiozza, F., Franceschini, G., Henry, M.,
Muchoney, D. & Tramberend, S. 2021. Global Agro-Ecological Zones v4 – Model documentation.
Rome, FAO. https://doi.org/10.4060/cb4744en

Gall, M., Borden, K.A. & Cutter, S.L. 2009. When do losses count? Six fallacies of natural hazards
loss data. Bulletin of the American Meteorological Society, 90(6): 799–810.

Garrett, L. & Neves, B. 2016. Incentives for ecosystem services: Spectrum. Rome, FAO. https://www.
slideshare.net/FAOoftheUN/incentives-for-ecosystem-services-spectrum

Gilbey, B. 2018. A review of the land degradation neutrality process. Gland, International Union
for Conservation of Nature. http://comunidadpnacc.com/wp-content/uploads/2020/01/ldn_
process_review_formatted-1.pdf

Gutierrez, V. & Keijzer, M.-N. 2015. Funding forest landscape restoration using a business-
centred approach: An NGO’s perspective. Unasylva 245, 66: 99–106.

Habegger, B. 2010. Strategic foresight in public policy: Reviewing the experiences of the UK,
Singapore, and the Netherlands. Futures, 42(1): 49–58.

Huang, J., Li, Y., Fu, C., Chen, F., Fu, Q., Dai, A., Shinoda, M. et al. 2017. Dryland climate change:
Recent progress and challenges. Reviews of Geophysics, 55(3): 719–778.

ICBA (International Center for Biosaline Agriculture). 2021. ICBA. Dubai. www.biosaline.org

IPBES (Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services).


2018. The IPBES assessment report on land degradation and restoration, L. Montanarella, R.
Scholes & A. Brainich, eds. Bonn. https://ipbes.net/assessment-reports/ld

IPCC (Intergovernmental Panel on Climate Change). 2019. 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. Geneva. www.ipcc.ch/srccl/

Iseman, T. & Miralles-Wilhelm, F. 2021. Nature-based solutions in agriculture: The case and path-
way for adoption. Virginia, FAO and The Nature Conservancy. https://doi.org/10.4060/cb3141en

IUCN (International Union for Conservation of Nature). 2022. Restoration Oppor-


tunities Assessment Methodology (ROAM). In: IUCN. Gland. www.iucn.
org/theme/forests/our-work/forest-landscape-restoration/restoration-opportunities-as-
sessment-methodology-roam

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 249
Jakovac, C.C., Latawiec, A.E., Lacerda, E., Leite Lucas, I., Korys, K.A., Iribarrem, A., Malaguti,
G.A. et al. 2020. Costs and carbon benefits of mangrove conservation and restoration: A global
analysis. Ecological Economics, 176: 106758.

Lambin, E.F., Turner, B.L., Geist, H.J., Agbola, S.B., Angelsen, A., Bruce, J.W., Coomes, O.T. et
al. 2001. The causes of land-use and land-cover change: Moving beyond the myths. Global
Environmental Change, 11(4): 261–269.

Liniger, H.P. & Studer, M. 2019. Sustainable rangeland management in sub-Saharan Africa: Guide-
lines to good practice. TerrAfrica, World Bank, World Overview of Conservation Approaches and
Technologies, World Bank Group and Centre for Development and Environment and Univer-
sity of Bern. www.wocat.net/library/media/174

Mirzabaev, A., Wu, J., Evans, J., García-Oliva, F., Hussein, I., Iqbal, M., Kimutai, J. et al. 2019.
Desertification. In: P.R. Shukla, J. Skea, E. Calvo Buendia, V. Masson-Delmotte, H.-O. Pört-
ner, D.C. Roberts, P. Zhai et al. eds. Climate change and land: An IPCC special report on climate
change, desertification, land degradation, sustainable land management, food security, and green-
house gas fluxes in terrestrial ecosystems. Geneva, IPCC.
www.ipcc.ch/site/assets/uploads/sites/4/2019/11/06_Chapter-3.pdf

Mortimore, M. 2009. Dryland opportunities: A new paradigm for people, ecosystems and develop-
ment. Gland, IUCN. www.cbd.int/doc/case-studies/inc/cs-inc-iucn-dryland-en.pdf

Nissan, H., Goddard, L., de Perez, E.C., Furlow, J., Baethgen, W., Thomson, M.C. & Mason, S.J.
2019. On the use and misuse of climate change projections in international development. Wiley
Interdisciplinary Reviews: Climate Change, 10(3): e579.

Nkonya, E., Anderson, W., Kato, E., Koo, J., Mirzabaev, A., von Braun, J. & Meyer, S. 2016. Global
cost of land degradation. In: E. Nkonya, A. Mirzabaev & J. von Braun, eds. Economics of land
degradation and improvement: A global assessment for sustainable development, pp. 117–165.
Cham, Springer International Publishing.

Qadir, M., Quillérou, E., Nangia, V., Murtaza, G., Singh, M., Thomas, R.J., Drechsel, P. & Noble,
A.D. 2014. Economics of salt-induced land degradation and restoration. Natural Resources
Forum, 38(4): 282–295.

REN21 (Renewable Energy Policy Network for the 21st Century). 2018. Renewables 2018:
Global status report. www.ren21.net/wp-content/uploads/2019/05/GSR2018_Full-Report_
English.pdf

Safriel, U. 2017. Land degradation neutrality (LDN) in drylands and beyond – where has it come
from and where does it go. Silva Fennica, 51(1B): 1650.

Seddon, N., Smith, A., Smith, P., Key, I., Chausson, A., Girardin, C., House, J., Srivastava, S. &
Turner, B. 2021. Getting the message right on nature-based solutions to climate change. Global
Change Biology, 27(8): 1518–1546.

250 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


Tarrason, D., Andrian, G. & Groppo, P. 2017. Toolkit for the application of green negotiated
territorial development (GreeNTD). Land and Water Division Working Paper 16b. Rome, FAO.
www.fao.org/3/a-i6591e.pdf

Tuan, H., Nachtergaele, F., Chiozza, F. & Ziadat, F. 2022. Land suitability for crop production in
the future. Thematic Background Report for SOLAW 2021. Rome, FAO.
www.fao.org/land-water/solaw2021/en

UNCCD (United Nations Convention to Combat Desertification). 2017. Global land outlook: First
edition. Bonn. https://knowledge.unccd.int/sites/default/files/2018-06/GLO English_Full_
Report_rev1.pdf

UNCCD. 2019. Land under pressure – Health under stress. Global Land Outlook Working Paper.
https://catalogue.unccd.int/1463_Land_under_pressure_Health_under_stress.pdf

UNCCD. 2022a. Global land outlook: Second edition. Bonn.


https://www.unccd.int/resources/global-land-outlook/global-land-outlook-2nd-edition

UNCCD. 2022b. The LDN target setting programme. In: UNCCD.


www.unccd.int/actions/ldn-target-setting-programme

UNDP (United Nations Development Programme) & UN WOMEN. 2018. Report on envi-
ronmental impact of Rohingya influx. Bangladesh. www.bd.undp.org/content/bangla-
desh/en/home/library/environment_energy/a-new-land-use-model--forest-fruit-fish1.
html?utm_source=EN&utm_medium=GSR&utm_content=US_UNDP_PaidSearch_Brand_
English&utm_campaign=CENTRAL&c_src=CENTRAL&c_src2=GSR&gclid=CjwKCAiAyPy-
QBhB6EiwAFUuakszdxjx2-Qi51AMINuyb85A9pRngfDU6K6D53YW5Y8iG7pjTCjbEBhoCq-
fYQAvD_BwE

UN DPA (United Nations Department of Political Affairs) & UNEP (United Nations Environment
Programme). 2015. Natural resources and conflict: A guide for mediation practitioners. Nairobi.
https://wedocs.unep.org/bitstream/handle/20.500.11822/9294/-Natural_resources_and_
conflic.pdf?sequence=2&amp%3BisAllowed=

UNDRR (United Nations Office for Disaster Risk Reduction). 2021. GAR special report on drought
2021. Geneva. www.undrr.org/publication/gar-special-report-drought-2021

United Nations. 1992. United Nations Conference on Environment & Development Agenda 21.
https://sustainabledevelopment.un.org

United Nations. 2012. The future we want: Outcome document of the United Nations Conference on
Sustainable Development. Rio de Janeiro.
https://sustainabledevelopment.un.org/content/documents/733FutureWeWant.pdf

United Nations. 2015. Transforming our world: The 2030 agenda for sustainable development. New
York. A/RES/70/1. www.un.org/ga/search/view_doc.asp?symbol=A/RES/70/1&Lang=E

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 251
United Nations. 2018. Sustainable Development Goal 6: Synthesis report 2018 on water and
sanitation. New York. www.unwater.org/publications/sdg-6-synthesis-report-2018-on-wa-
ter-and-sanitation

van den Ende, M., Wardekker, A., Mees, H. & Hegger, D. 2021. Towards a climate-resilient future
together: A toolbox with participatory foresight methods, tools and examples from climate
and food governance. Utrecht University. www.researchgate.net/publication/345733771_
Towards_a_climate-resilient_future_together_A_toolbox_with_participatory_fore-
sight_methods_tools_and_examples_from_climate_and_food_governance

van Soesbergen, A., Arnell, A.P., Sassen, M., Stuch, B., Schaldach, R., Göpel, J., Vervoort, J.,
Mason-D’Croz, D., Islam, S. & Palazzo, A. 2016. Exploring future agricultural development
and biodiversity in Uganda, Rwanda and Burundi: A spatially explicit scenario-based assess-
ment. Regional Environmental Change, 17(5): 1409–1420.

Vervoort, J. & Gupta, A. 2018. Anticipating climate futures in a 1.5 °C era: The link between fore-
sight and governance. Current Opinion in Environmental Sustainability, 31: 104–111.

Vlek, P.L.G., Khamzina, A. & Tamene, L., eds. 2017. Land degradation and the Sustainable Devel-
opment Goals: Threats and potential remedies. Publication No. 440. Nairobi, International Center
for Tropical Agriculture. https://cgspace.cgiar.org/handle/10568/81313

Wani, S., Rockström, J. & Oweis, T., eds. 2008. Rainfed agriculture: Unlocking the potential.
Comprehensive Assessment of Water Management in Agriculture Series No. 7. CAB Inter-
national. www.iwmi.cgiar.org/Publications/CABI_Publications/CA_CABI_Series/Rainfed_
Agriculture/Protected/Rainfed_Agriculture_Unlocking_the_Potential.pdf

Wilhite, D. 2011. Quantification of agricultural drought for effective drought mitigation and
preparedness: Key issues and challenges. Drought Mitigation Centre Faculty Publications Paper
82. https://digitalcommons.unl.edu/droughtfacpub/82/?utm_source=digitalcommons.unl.
edu%2Fdroughtfacpub%2F82&utm_medium=PDF&utm_campaign=PDFCoverPages

Willemen, L., Barger, N.N., ten Brink, B., Cantele, M., Erasmus, B.F.N., Fisher, J.L., Gardner, T.
et al. 2020. How to halt the global decline of lands. Nature Sustainability, 3(3): 164–166.

WMO (World Meteorological Organization). 2013. High-level Meeting on National Drought


Policy (HMNDP). In: WMO. Geneva. https://community.wmo.int/meetings/high-level-meet-
ing-national-drought-policy-hmndp

WMO & GWP (Global Water Partnership). 2014. National drought management policy guidelines: A
template for action. Integrated Drought Management Programme Tools and Guidelines Series
1. Geneva and Stockholm. www.droughtmanagement.info/literature/IDMP_NDMPG_en.pdf

WOCAT (World Overview of Conservation Approaches and Technologies). 2022. WOCAT:


Welcome to WOCAT. In: WOCAT. www.wocat.net/en

WWAP (United Nations World Water Assessment Programme)/UN-Water. 2018. The United
Nations world water development report 2018: Nature-based solutions for water. Paris, United

252 4. SUSTAINABLE LAND RESOURCES PLANNING AND MANAGEMENT


Nations Educational, Scientific and Cultural Organization. https://www.unwater.org/publica-
tions/world-water-development-report-2018/

Zdruli, P. 2014. Land resources of the Mediterranean: Status, pressures, trends and impacts on
future regional development. Land Degradation & Development, 25(4): 373–384.

Ziadat, F., Bunning, S. & De Pauw, E. 2017. Land resource planning for sustainable land manage-
ment: Current and emerging needs in land resource planning for food security, sustainable live-
lihoods, integrated landscape management and restoration. Land and Water Division Working
Paper 14. Rome, FAO. www.fao.org/3/a-i5937e.pdf

Ziadat, F., De Pauw, E., Nachtergaele, F. & Fetsi, T. 2021. A land resources planning toolbox to
promote sustainable land management. Sustainable Agriculture Research, 10(1): 73.

Ziadat, F.M., Zdruli, P., Christiansen, S., Caon, L., Abdel Monem, M. & Fetsi, T. 2022. An over-
view of land degradation and sustainable land management in the Near East and North Africa.
Sustainable Agriculture Research, 11(1): 11–24.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 253
Responses
and action areas
Key messages
5
Four key action areas emerge from this report to facilitate a programmatic approach to land, soil and water
resources planning and management to turn natural processes and human action towards a desired state
or new equilibrium. They do not lead to prescriptive single-purpose “solutions”, as the operational issues
facing today’s agriculture are multifaceted. But as agriculture plays a large role in the processes that cause
natural resource scarcity and degradation, it must also be part of any solution.
Land and water governance needs to be more inclusive and adaptive. Inclusive governance is essential for
allocating, planning and managing natural resources to continue to meet increasing demands. Technical
solutions to mitigate land degradation and water scarcity are unlikely to succeed without adaptive gover-
nance with all concerned institutions and actors.

Integrated solutions need planning with stakeholders and need to be mainstreamed to take them to scale.
Planning is essential for best and optimum solutions with multiple actors that maintain resource use below
critical thresholds in natural resource systems and lead to restoration of resources and ecosystem services
when supported by appropriate technical, institutional, governance and financial packages or programmes.

Technical and managerial innovation needs to be targeted to address priorities, reduce risks and enhance
resilience of people and ecosystems. Caring for neglected soils, addressing drought and coping with water
scarcity will need special measures for incentivizing local adaptation and wide adoption of new technolo-
gies and management approaches.

Agricultural support and investment should be redirected towards social and environmental gains
derived from the range of land and water management solutions available, leaving no one behind. There
is now scope for progressive multiphased financing of agrienvironmental interventions linked with redi-
rected subsidies to keep land and water systems in play and to contribute towards multiple SDGs, notably
those on food security and poverty alleviation.
5.1 Introduction

© FAO/Olympia de Maismont
The responses in this chapter build on and
complement the land planning and integrated
management options presented in Chapter
4 and add to the analysis in Chapters 1–3.
Responses in policy, institutions and techni-
exposed and exploited governance weak-
cal domains can be applied to create positive
nesses and inequalities in the global food
and transformative changes that keep land,
system, including among and within coun-
soil and water systems in play and mitigate
tries and population groups.
the further build-up of pressures.

This SOLAW 2021 report comes at a time


The basis for resource management deci-
when driving forces are accelerating change.
sions is established in broad terms by taking
Urgent, coordinated action across many
stock of the trends in land and water resource
sectors is required to meet the food demand
use (Chapter 1). However, in practice, such
of a global population that continues to
decisions will occur in socioeconomic
grow, although at a decelerated rate, within
settings and under governance regimes that
the carrying capacity of natural resources
may facilitate or limit societal capacity to
and without compromising the ecosystem
implement them (Chapter 2). Socioeconomic
services on which all life depends. Recent
trade-offs become increasingly complex to
assessments and global reviews related to
evaluate as resource use intensifies, and the
SDGs confirm this urgency.
broader goals of eliminating hunger and
sustaining the natural environment need to
This chapter therefore offers a structured
be reconciled. They extend across social and
selection of key response options that
economic sectors, water-cycle components,
policymakers can put together to find
stages of the food value chain and supply
the best combination according to their
chains above and below the land. They bring
circumstances, needs and capabilities. These
about higher-level interdependencies and
are grouped under four action areas:
establish the basis for cobenefits in deal-
ing with the multiple aspects of natural I. Adopting inclusive land and water
resources management and food and agri-
governance (section 5.3).
cultural systems.
II. Implementing integrated solutions at
Since the COVID-19 pandemic that began in
scale (section 5.4).
2020, the development priority has focused
on tackling health issues, the economic III. Embracing innovative technologies
downturn and impacts on food security and management (section 5.5).
(FAO et al., 2021). Yet, other serious chal-
lenges are emerging, driven by poverty and IV. Investing in long-term sustainability
climate change, including natural disasters (section 5.6).
and migration trends. Severe floods and
droughts have caused loss of life, damage to But first, it is important to take stock of the
infrastructure, significant agricultural losses various platforms for responding to land and
and food security impacts. The pandemic water degradation and water scarcity, which
and associated policy responses have also is briefly reviewed in the next section.

256 5. RESPONSES AND ACTION AREAS


5.2 Response
The decisions and priorities of the three
multilateral Rio Conventions – addressing

platform: from global biodiversity (CBD), desertification, land

to individual efforts
degradation and drought (UNCCD) and
climate change (UNFCCC) – the Ramsar
Convention on Wetlands of International
At the global level, concerns over the state of Importance Especially as Waterfowl Habitat
land and water resources were embedded in (Ramsar Convention, see below), and the water
the United Nations Rio Conventions arising and watercourse conventions (see below)
from the 1992 United Nations Conference provide scope for solutions that concern land
on Environment and Development (the Rio and water for food and agriculture (Bbox 5.1).
Earth Summit) and their financing instru- The UNCCD LDN target-setting programme
ments (including the Green Climate Fund and explicitly includes water and food security
GEF). Many regional and local initiatives for while aiming to avoid, reduce and reverse land
land and water resources management are degradation. The CBD (Bbox 5.1) promotes the
now guided by the 2030 Agenda. In addition, restoration and maintenance of biologically
regional economic initiatives will remain diverse ecosystems through thematic
important. For instance, the European Green programmes (inland waters, marine and
Deal is expected to mainstream sustainable coastal, agriculture, forests and dry lands)
development and land and water resources and an integrated ecosystem approach. The
management in Europe and beyond, through decisions and commitments under CBD,
policy initiatives and focused investments SDG 15 targets and UNCCD contribute to
and incentives by the European Commission improving land and water management as
towards a “climate-neutral continent” by a basis for conservation and sustainable use
2050 (EC, 2019). of above- and below-ground biodiversity, to
protection and restoration of soil and water
Land and water resources are implicated in resources and ecosystem services, and to
the 2030 Agenda, for example in: SDG 2, improved availability and access to clean
ending hunger and achieving sustainable drinking water.
food and agriculture systems (via increas-
ing the proportion of agricultural area under The 2015 Paris Agreement under the UNFCCC
productive and sustainable agriculture); (Bb
ox 5.1) was a milestone in making agricul-
SDG 6, securing water and sanitation for all; ture part of the solution rather than a primary
SDG 13, combating climate change and its cause of climate change. Under the agree-
impacts; and SDG 15, sustainably managing ment, countries agreed to work together to
terrestrial ecosystems. Target-level linkages ensure agricultural development increases
extend to most other SDGs: SDG 1, ending food security in the face of climate change
poverty; SDG 5, achieving gender equality; and also reduces GHG emissions. The land-
SDG 7, ensuring access to sustainable energy; mark Koronivia Joint Work on Agriculture
SDG 12, ensuring responsible consumption; (KJWA) mainstreams agriculture’s role in
SDG 14, land–freshwater–oceans interlink- tackling climate change through request-
ages and the impact on food security; and ing two subsidiary bodies under UNFCCC to
SDG 17, revitalizing the role of partnerships. address issues related to agriculture (section
The analyses and state indicators of SOLAW 5.4.2). It provides a platform for land and
2021 align well with the various SDG targets water policy coherence in climate adaptation
and their indicators. and mitigation across agricultural sectors.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 257
The Ramsar Convention (Bb
ox 5.1) was signed
in 1971 and entered into force in 1975. Almost
90 percent of United Nations Member States

©FAO/Roberto Faidutti
have become contracting partners since
then. This treaty provides a framework for
national action and international cooperation
for conserving and wise use of wetlands and
their resources.

The international community agreed to nizations, international organizations, civil

establish two United Nations global water society organizations and NGOs.

conventions for transboundary watercourses:


At its forty-second session in 2015, CFS
stressed the key role of water in achieving
ƒƒ The Convention on the Protection and
the 2030 Agenda, and encouraged stakehold-
Use of Transboundary Watercourses and
ers to join forces to address the challenges
International Lakes (Water Conven-
related to water’s contribution to food secu-
tion) was adopted in 1992 in Helsinki
rity and nutrition through ecosystem and
and entered into force in 1996 by and for people-centred approaches. It recalled that
the Member States of the United Nations water is essential for realizing the right to
Economic Commission for Europe. The adequate food in the context of national food
Water Convention was opened to global security, and the right to safe drinking water
accession in 2016 (Bb
ox 5.1). and sanitation. It offered eight specific policy
recommendations on water for food security
ƒƒ The Convention on the Law of the and nutrition (CFS, 2015).
Non-Navigational Uses of International
Watercourses (Watercourses Convention) The state of food security and nutrition in the

was adopted by the United Nations in 1997 world 2021 (FAO et al., 2021) and the multi-
actor consultative dialogues and prioritized
in New York, and entered into force in
work under five interlinked Action Tracks
2014 (Bb
ox 5.1).
through the 2021 United Nations Food

The CFS is a multistakeholder, inclu- Systems Summit process were guided by CFS,
and focused on transforming food systems
sive international and intergovernmen-
for food security, improved nutrition and
tal platform that develops and endorses
affordable healthy diets for all.
policy recommendations and guidance
for the United Nations system on a wide
Endorsed by its member governments, OECD
range of food security and nutrition topics
has launched a water governance initiative
to ensure food security and nutrition for that provides policy guidance and recom-
all (CFS, 2015). It is supported by: scien- mendations on rules, processes and institu-
tific and evidence-based reports produced tions involved in sound water governance,
by a high-level panel of experts; techni- with specific focus on stakeholder involve-
cal support from the Rome-based United ment. The OECD also provides support for
Nations agencies (FAO, International Fund sustainable use and management of natu-
for Agricultural Development and World Food ral resources, including land and water
Programme); the CFS Bureau, composed of (OECD, 2022).
governments; and the CFS Advisory Group,
comprising relevant United Nations orga-

258 5. RESPONSES AND ACTION AREAS


The private sector’s engagement with land izer and pesticide use and retaining natu-
and water will remain fundamental at all ral habitats, and at intermediary level to
stages of the food value chain. The choice reduce externalities of food waste, transport
of technology and site selection for opera- and energy use. Behaviour change at the
tions, environmental stewardship and social consumer level may not be sending a strong
responsibility practices are under a spot- enough signal to producers, due to the large
light, and offer more initiatives and exam- amount of value added in supply chains.
ples of best practices, including certification
and corporate disclosure schemes. There is The broad and circumstance-specific actions
a need to foster greater public and private presented in this chapter are neither prescrip-
sector collaborative engagement to finance tions nor templates for action, but are intended
governance, and systemic and technological to inform and guide stakeholders, from leaders
innovations for sustainable and resilient land to individuals, and from producers to consum-
and water management. These could involve ers, in their decision-making to produce
diverse private financiers and development actionable ways forward for their circum-
financing institutions, bilateral donors, stances and blend of issues. These actions
international organizations, research by combined provide a sound framework to miti-
academia and implementation by civil society gate the risks identified in Chapter 3, and
organizations. enhance sustainable development, economic
growth and food security, with attention to
The investment cost needed at the farm smallholders, women farmers and vulnerable
level towards sustainability is typically large groups. Policymakers are urged to incorpo-
in relation to the incremental price that rate these options into strategic actions that
consumers are willing to pay. This makes protect and enhance efficiency, productiv-
it necessary for governments to introduce ity and resource availability, and ensure food
regulations that directly operate at the farm security and nutrition for all. Some options
and territorial scales, such as limiting fertil- are already proving successful in practice and
can be replicated and scaled up at all levels –

Box 5.1
International frameworks: convergence around integrated, sustainable
and equitable land and water governance

In addition to the 2030 Agenda, many international legal agreements and high-level international
political commitments form a strong mandate for promoting multisectoral and integrated approaches
to land and water governance. These frameworks shift the international development agenda focus
to inclusion, equity and ecosystem integrity, and resilience as essential foundations for sustainable
development.

Convention on Biological Diversity: This is in a critical phase of post-2020 planning to reinvigorate


action to achieve the Aichi Targets most closely related to land and water governance. Under sustainable
agriculture, the Global Biodiversity Outlook 5 calls for integrated land and water policies to support
reduced pollution, increased irrigation efficiency and redirection of perverse subsidies and incentives.
In the first draft of the Post-2020 Global Biodiversity Framework, 4 of the 21 action-oriented targets
are directly relevant to the land and water agenda (UNEP, 2021) for “Reducing threats to biodiversity”
(targets 1–3) and for providing “Tools and solutions for implementation and mainstreaming” (target 21):

“Target 1. Ensure that all land and sea areas globally are under integrated biodiversity-inclusive spatial
planning addressing land- and sea-use change, retaining existing intact and wilderness areas.”

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 259
Box 5.1 (continued)

“Target 2. Ensure that at least 20 per cent of degraded freshwater, marine and terrestrial ecosystems are
under restoration, ensuring connectivity among them and focusing on priority ecosystems.”

“Target 3. Ensure that at least 30 per cent globally of land areas and of sea areas, especially areas of
particular importance for biodiversity and its contributions to people, are conserved through effectively
and equitably managed, ecologically representative and well-connected systems of protected areas
and other effective area-based conservation measures, and integrated into the wider landscapes and
seascapes.”

“Target 21. Ensure equitable and effective participation in decision-making related to biodiversity by
indigenous peoples and local communities, respect their rights over lands, territories and resources, as
well as by women and girls, and youth.”

Paris Agreement: This emphasizes the “intrinsic relationship that climate change actions, responses and
impacts have with equitable access to sustainable development and eradication of poverty” and the
“fundamental priority of safeguarding food security”. Integrated, holistic and balanced approaches that
aim to enable opportunities for coordination across instruments and relevant institutional arrangements
are also emphasized (Article 6), and in taking these measures, countries shall cooperate in enhancing
access to information, public awareness and public participation (Article 12). This is reflected in the
IPCC special report on climate change and land (IPCC, 2019), which demonstrates that land is part of
the climate solution and how managing land resources sustainably can help address climate change
with attention to desertification, land degradation, SLM, food security and GHG fluxes in terrestrial
ecosystems.

Ramsar Convention: This is an intergovernmental treaty that provides the framework for national action
and international cooperation for the conservation and wise use of wetlands and their resources.

Sendai Framework for Disaster Risk Reduction 2015–2030: This emphasizes reduced exposure and
vulnerability to disaster risks through more people-centred, inclusive and coordinated intersectoral
approaches that address the underlying drivers of those risks. The framework calls explicitly for
implementing integrated and inclusive legal, political and institutional measures that mainstream
disaster risk reduction into land and water policies, implement and enforce land-use and resource
regulatory mechanisms to ensure ecosystem health, and support intersectoral coordination appropriate
to national systems of governance while empowering effective engagement of women, youth, the
elderly, migrants, people with disabilities and indigenous people.

United Nations Convention to Combat Desertification: As the legally binding international agreement
linking environment and development to SLM, this convention complements biodiversity, climate and
land dynamics, and is aligned in its objectives to tackle desertification, land degradation and the effects
of droughts in the SDG suite, with particular emphasis on SDG 15. Land degradation neutrality has
gained momentum in recent years. By 2022, 128 countries had committed to setting targets and more
than 100 countries had agreed targets to avoid and reduce degradation and restore degraded lands.
It promises to be a high-priority agenda item for governments to support UNCCD objectives, the 2030
Agenda, the United Nations Watercourses Convention and the United Nations Economic Commission
for Europe Water Convention.

United Nations Watercourses Convention and United Nations Economic Commission for Europe
Water Convention: These two conventions address transboundary watercourses, covering 85 percent
of all river basins and some 40 percent of the global population (UNECE, 2021).

260 5. RESPONSES AND ACTION AREAS


community, subnational, national, regional ƒƒ social and adaptive learning that builds on
and global – for wide-scale and long-term multistakeholder collaboration (Reed and
implementation and continuous adaptation Massie, 2013).
towards sustainable land and water manage-
ment and sustainable agrifood systems. Increasing uncertainties and complexi-
ties surrounding land and water governance

5.3 Action area I:


require an adaptive approach to policies and
management systems (Pahl-Wostl, 2015).

Adopting inclusive Efforts to enhance adaptive capacity have

and adaptive
attempted to distil and operationalize “adap-
tive governance” approaches that build on

land and water adaptive management and comanagement,


in which governments share powers with

governance resource users (Folke et al., 2005; Ostrom, 2005;


Chaffin, Gosnell and Cosens, 2014). Many of
the core components of adaptive governance
The variety of governance approaches to land
and water identified in Chapter 2 point to an across countries and local contexts closely

overall rigidity in tenure systems. This can lead align with other governance approaches

to missed opportunities to be more inclusive explored in this chapter (Pahl-Wostl et al.,


and more adaptive towards environmental 2007; Plummer et al., 2012; Chaffin, Gosnell
change. The governance and management and Cosens, 2014). For adaptive governance
of land, soil and water resources require to work in practice, the increased need for
transformative changes that enable actions flexibility should be balanced with consider-
towards sustainable agriculture and food ations of legitimacy and stability that prevent
systems, from production to consumption. arbitrariness (Cosens et al., 2017).
Changes in policy, institutional and technical
domains that disrupt BAU responses may Investments can help to strengthen gover-
be required. nance arrangements and policy change. These
include planning and management actions
Inclusive governance approaches for policy at farm and landscape levels and associated
action in land and water include: technological innovation. However, behav-
ioural change is also required. People-centred
ƒƒ multistakeholder collaboration to draw
on various knowledge systems, values and
experiences (Pahl-Wostl, Mostert and
Tàbara, 2008);

ƒƒ polycentric governance systems with


shared governance responsibilities across
decision-making at various levels of
© FAO/Stefanie Glinski

governance (Ostrom, 2010; Pahl-Wostl


et al., 2013);

ƒƒ experimentation and flexibility in testing


policy interventions; and

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 261
governance approaches are needed to and wide adoption of sustainable and
include all stakeholders, and collaborative transformative land-use and food
decision-making and learning require delib- systems and practices adapted to specific
erate linkages across institutions, scales and socioeconomic and ecological settings.
sectors to capitalize on stakeholder interests.
The need to recognize and protect land and
There is increasing recognition in interna- water tenure rights (particularly among rural
tional frameworks and national governance communities, indigenous people, women
mechanisms of the crucial role of land and and other vulnerable groups) underpins food

water management in climate change (Bbox security and nutrition, sustainable livelihoods

5.2). The IPCC special report on climate and climate resilience. Harmonizing land and
water governance systems is essential and
change and land (IPCC, 2019) highlights that
should build on experiences in addressing
land must remain productive to maintain food
specific land and water governance chal-
security as the population increases and the
lenges. Addressing the needs of vulnerable
impacts of climate change on soils and crops
and marginalized populations requires an
are felt. The roles of soil and land management
understanding of power structures and
in carbon sequestration and emission mitiga-
incentives within society that govern natural
tion are particularly important. They require
resources access, incorporating their specific
recognition in policies and governance instru-
needs into policy, planning and investments.
ments backed up by land-use and resource
evaluation and vulnerability risk assessments. Previous chapters have highlighted the chal-
lenges facing land and water resources as
Inclusive governance is key to building
increasing demands for food and agricul-
capability and competent and informed
ture, energy, industries and municipalities
institutions and organizations at all
compete with conserving and enhancing
levels of decision-making. This enables
ecosystems and their services. Food systems
mainstreaming and mobilizing effective
drive climate change and contribute to land

Box 5.2
Facilitating policy coherence and integrated land and water governance
through climate responses

The ability of socioecological systems to respond, recover and adapt to climate impacts is closely linked
to how well climate mitigation can be achieved. Integrated approaches to adaptation and mitigation can
reduce risks and identify synergies that mitigate threats to food security (Di Gregorio et al., 2016).

The REDD+ initiative, originally conceived as a mitigation solution, has evolved to include conservation
and sustainable forest management and enhanced forest carbon stocks through interventions to address
a suite of forest governance issues, including tenure, gender equality and stakeholder participation.
For example, the Lower Zambezi REDD+ project focuses on establishing community-based forest
mitigation through conservation farming and tree nursery development to create sustainable
alternatives to deforestation, thus increasing communities’ resilience while preventing emissions
(Munroe and Mant, 2014).

The KJWA provides a platform for strengthening land and water governance by integrating climate
adaptation and mitigation policies across agricultural sectors (see section 5.4.2).

262 5. RESPONSES AND ACTION AREAS


and soil degradation, water scarcity and

© FAO/Alessandra Benedetti
biodiversity loss. The COVID-19 pandemic
has exacerbated these challenges, which
disproportionately affect vulnerable and
marginalized populations.

Building on the land and water governance


issues raised in Chapter 2, five governance
responses promise effective transformation:
can be established as a basis for monitor-
ing progress, assessing risks and developing
ƒƒ developing coordinated and coherent
investment plans. Desirable outcomes will
policies and approaches;
depend on locally appropriate policies and
governance systems and competent insti-
ƒƒ strengthening and harmonizing land and
tutions for their implementation. Interna-
water tenure systems;
ox 5.1), national poli-
tional instruments (Bb
ƒƒ effectively engaging actors in negotiation; cies and legislative frameworks recognize the
interconnectedness of development issues
ƒƒ improving employment, livelihoods and and promote integrated and intersectoral
gender equity; and approaches.

ƒƒ undertaking governance analysis. Improved intersectoral coordination (land,


water, agriculture, environment, finance

5.3.1 Developing and planning) can help identify and address


overlaps and trade-offs, improve perfor-
coordinated and coherent mance across multiple levels of government,
policies and approaches reduce costs and identify areas where lines
of authority can be better delineated should
Most countries have adopted policy and
a conflict arise, and clarify to stakeholders
legal frameworks governing land and water
who is accountable for decisions and actions
resources. However, in many cases, such
(Kristensen, 2004; Tripathi et al., 2019; UNEP,
frameworks are fragmented and lack effec-
2019). Improved coordination is necessary to
tive implementation in practice. This is
equitably distribute cobenefits from poli-
because traditional siloed land and water
cies and decisions, especially for vulnerable
management and sectoral agricultural
populations.
(crops, livestock and forestry) approaches
persist, and effective links among levels of Finding appropriate pathways to make inte-
decision-making are weak. A focus on tech- grated approaches work is critical to stem-
nical solutions has frequently resulted in ming overexploitation, degradation of land
fragmented jurisdiction over ecologically and soil resources, and water scarcity, which
interconnected resources. could undermine resilient ecosystems and
sustainable development. This requires inno-
Power imbalances inhibit coordination
vative governance responses and enhanced
among ministries and technical sectors, limit
institutional and societal capacities to build
competition over budgetary resources and
on synergies, address trade-offs and manage
foster mistrust among agencies. To counter
processes that may involve (re)allocating
this, competent institutions need to promote
limited resources, addressing inequalities
participatory planning and administration
and changing the way of empowering actors
processes in land and water sectors, through
at different levels of decision-making.
which “red line” targets and thresholds

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 263
5.3.2 Strengthening
and harmonizing land
and water tenure systems
Land and water tenure systems determine how
individuals, communities and others acquire
rights and associated duties to use, manage

©FAO/Olivier Asselin
and benefit from land and water resources.
Although data on tenure security are incom-
plete (to be improved under SDG indicator
1.4.2), insecurity of tenure rights continues
to threaten the livelihoods and well-being of
a significant share of the population depen-
Most countries fail to recognize the inter-
dent on land and water resources for their
relations among land and water tenure rights,
livelihoods. This weakens incentives for
legally and in practice (RRI and ELI, 2020;
farmers and other rural land users to invest
FAO, 2020a). Yet, recent work demonstrates
in improving their land and water resources
that it is possible to articulate, conceptually
(RRI, 2017). Insecurity also reduces access to
and based on legislative practice, a core set
credit, further undermining the capacity to
of water-related rights that comprise the
sustain and improve agricultural productivity.
diverse water tenure regimes found world-
Over the past decade, a key milestone has been wide. Taking a “bundle of rights” approach
achieved to support countries in improving enables countries to identify areas for harmo-
tenure security for all, in particular the most nization across water, land, forest, fisher-
vulnerable, with negotiation and adoption ies and other key resource sector legislation
by government, civil society, private sector for improved and integrated land and water
and academia representatives, under CFS, of governance (FAO, 2020b; RRI and ELI, 2020).
the VGGT (CFS and FAO, 2012). As a common FAO is now facilitating international debate
international standard of responsible gover- to develop further the concept of water tenure
nance of tenure, the guidelines are being and guidance for countries and to address
implemented in over 100 countries (Global water tenure reforms to support food secu-
Donor Platform for Rural Development, rity, sustainable livelihoods, climate resil-
2022), supporting tackling specific tenure ience and development goals (FAO, 2020c).
issues or supporting broad land governance
FAO work on water governance and tenure
programmes providing multiple benefits.
includes raising awareness and developing
Recent progress has secured communities’ tools and capacity for integrating tenure
land and forest tenure rights, attributable in assessment to strengthen water governance.
part to developing and implementing prin- This forms the basis for sustainable and
ciples and tools to guide policy and legal equitable water management, recognizing
reform. Of note in this regard are the VGGT the legitimate tenure rights of pastoralists
and concerted advocacy efforts on behalf and pastoralism and formalizing women’s
of rural communities (CFS and FAO, 2012). land and water rights for gender-equitable
However, a significant gap remains between ox 5.3).
outcomes (Bb
commitments and practice.

264 5. RESPONSES AND ACTION AREAS


Box 5.3
Strengthening water governance and water tenure rights

Water governance through tenure assessment and governance analysis


Under the Knowing Water Better project, FAO is working in Rwanda, Senegal and Sri Lanka to strengthen
water governance processes through water accounting, governance and tenure assessment. The
project is piloting a national water tenure assessment methodology that can map water tenure regimes
(including those of individuals, the private sector and communities) and identify gaps and overlaps in
sectoral legislation, where customary practices present threats to water tenure security.

Additional work under the water efficiency, productivity and sustainability project (which focuses on
implementing the 2030 Agenda for water efficiency and productivity in eight Near East and North
Africa countries) provides data and information for sustainable water management that balances
environmental, economic and social sustainability to improve rural livelihoods, especially smallholder
farming. This project is also piloting a combined water accounting and auditing/governance analysis to
help policymakers achieve sustainable and equitable water management and use.
Sources: FAO. 2021. Methodology. In: Knowing water better: Towards fairer and more sustainable access to natural resources - KnoWat. Cited 3 March
2022. www.fao.org/in-action/knowat/wt-assessment/methodology/en; FAO. 2021. Water efficiency, productivity and sustainability in the NENA regions
(WEPS-NENA). In: FAO. www.fao.org/in-action/water-efficiency-nena/en

Recognizing customary land and water tenure in pastoralism


Customary laws and practices often determine how rural communities access, use and govern their
land and water resources. Where customary rights are not legally recognized, they may be ignored
or manipulated when competing claims for resources arise, thus increasing the vulnerability of
communities. Insecure tenure also reduces communities’ ability to invest in maintaining and sustaining
local agriculture and forest-based food systems. This undermines their key role in storing and managing
forest carbon (Byagmugisha, 2013; Oxfam and ILC, 2016).

Globally, 500 million people rely on pastoralism for livelihoods. Water tenure rights are critical, as
communities are often organized around access to various grazing lands and their limited water
supplies. Yet, in many countries, particularly where pastoral communities cross national boundaries,
overlapping or competing customary and formal governance systems do not recognize resource rights
(De Haan et al., 2016; Davies et al., 2018).

Governments and regional organizations are beginning to recognize the legitimate tenure rights of
pastoralists and pastoralism as an important and appropriate use of land and water resources. Burkina
Faso, Mail and Mauritania have passed legislation protecting grazing land and granting herders rights to
land and water resources, recognizing existing access and sharing arrangements and livestock corridors
as a critical tool to protect customary pastoral water tenure rights.
Sources: République du Mali. 2001. Loi n° 01-004 du 27 fevrier 2001, portant charte pastorale du Mali. Mali; Behnke, R. & Freudenberger, D. 2013.
Pastoral land rights and resource governance. In: LandLinks. Cited 2 March 2022. https://land-links.org/issue-brief/pastoral-land-rights-and-resource-
governance; Davies, J., Herrera, P., Ruiz-Mirazo, J., Mohamed-Katerere, J., Hannam, I. & Nuesiri, E. 2016. Improving governance of pastoral lands:
Implementing the voluntary guidelines on the responsible governance of tenure of land, fisheries and forests in the context of national food security.
Governance of Tenure Technical Guide 6. Rome, FAO. www.fao.org/3/i5771e/I5771E.pdf

Women’s land and water tenure rights drive social development and growth
Women’s land and water resource rights are important drivers of social development and economic
growth. Yet, in many instances, legislative and customary systems fail to promote women’s secure
land and water tenure (Keene, Troell and Ginsburg, 2020). Even where women are landholders, they
frequently face challenges in accessing productive resources, hold land of lower quality and have
lower levels of agricultural productivity than men. Women’s water tenure also depends on their legally
recognized land rights, further emphasizing the importance of strengthening their land and forest
tenure (Keene, Troell and Ginsburg, 2020).

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 265
Box 5.3 (continued
Where women do have secure tenure rights, they tend to invest in improving land, participate in land
rental markets, and contribute to family food security, children’s health and sustainable agricultural
productivity (USAID, 2016). In Rwanda, women with formalized land rights were 19 percent more likely
to invest in soil conservation, compared to only 10 percent of men. Globally, children whose mothers
own land are 33 percent less likely to be undernourished (Viña, 2020). Secure land rights also improve
women’s participation and leadership in community governance institutions.

Formalizing women’s land and water rights is crucial in analysing gender inequities in legal frameworks
(including marriage and inheritance laws), and also in the political economy of land and water resource
governance at the local level to facilitate women’s equal participation and achieve truly gender-equitable
outcomes (Doss and Meinzen-Dick, 2020).
Source: FAO. 2020. The state of food and agriculture 2020. Overcoming water challenges in agriculture.
Rome. https://doi.org/10.4060/cb1447en

5.3.3 Effectively engaging Green Negotiated Territorial Development

actors in negotiation
approach adapted such a methodology to
safeguard ecological integrity (Bb
ox 5.4).
National governance should help to secure
Engaging diverse stakeholders in policy
tenure rights, and recognize and protect
decisions about land and water governance
local land rights that people consider socially
brings multiple sources of knowledge, values
legitimate, including customary rights where
and information to the table, contribut-
relevant. It should also tackle competition
ing to building trust, social cohesion and
over limited land resources, in law and in
the rule of law. Participatory policymak-
practice. This can help to avoid the risk of
ing and decision-making also help defuse
inequalities due to social differentiation (e.g.
conflict and reframe issues holistically by
depriving local communities from access
identifying trade-offs and synergies across
to natural resources on which they depend)
constituencies.
or expropriation of marginalized groups
with limited rights (e.g. women, youth and
At the national level, some countries include
migrants) through investments in land.
legal requirements for civic engagement in
Enabling legal frameworks and financing for
land and water decision-making in their
implementation is crucial to effective civic
framework environmental laws, water and
engagement and rights-based approaches.
land sectoral laws, and planning laws, as
part of impact assessment requirements.
Participatory negotiated territorial develop-
Impact assessment laws can ensure proposed
ment approaches can promote multistake-
projects and activities are subject to public
holder dialogues on territorial development
consultation if implemented and enforced
opportunities to address competition over
appropriately (UNEP, 2019). Civic engage-
land and water resources. Such approaches
ment in permitting processes is also a critical
have been developed through practice and
means for individuals and communities to
successfully applied in many recovery
have notice of potential infringements on
situations, after conflict and in complex
their land and water rights. Some countries
emergencies (e.g. in Angola, Democratic
require environmental and social impact
Republic of the Congo and Mozambique). The

266 5. RESPONSES AND ACTION AREAS


Box 5.4
Green Negotiated Territorial Development and its contribution to
improving livelihoods resilience
The Green Negotiated Territorial Development approach facilitates interaction among stakeholders
involving land disputes to find solutions to competition and other problems of accessing land and
limited natural resources. It is adaptable to different scales of intervention and various stakeholders,
namely policymakers, tenants, communities, entrepreneurs and NGOs, and those who provide
expertise and economic resources. The process consists of five phases and concludes by the signing
of a Socio-Ecological Territorial Agreement:

ƒƒ Preparatory work: Identifying the territorial perimeter, and the stakeholders and their motivations
to intervene.

ƒƒ Phase I. Views: Understanding the territory as a socioecological system; preparing a first analytical
framework of concerned stakeholders, differentiating their positions, interests and strategies, and
creating an information system (socioeconomic, productive and ecosystemic) to better understand
the impacts, risks and conflicts.

ƒƒ Phase II. Horizons: Outlining coherent and feasible proposals for territorial development; setting
scenarios to facilitate consensus; and identifying the issues that adversely affect the territory.

ƒƒ Phase III. Negotiation: Seeking consensus for territorial development; creating round-table negotia-
tions involving all stakeholders; and analytical work for coherent, feasible, efficient and sustainable
interventions.

ƒƒ Phase IV. Stakeholders sign the Socio-Ecological Territorial Agreement: The fundamental basis for
implementing short-, medium- and long-term business plans, formalizing rights and duties, and
creating an implementation stakeholder platform.

ƒƒ Phase V. Monitoring and evaluating the activities.


Source: FAO. 2017. Toolkit for the application of green negotiated territorial development (GreeNTD).
Land and Water Working Paper 16b. Rome. www.fao.org/3/i6591e/i6591e.pdf

assessments in determining the validity of Increased production stimulates demand for


permit applications. Meaningful engagement labour during the primary cropping seasons
is often challenging, requiring substan- and the dry season (livestock, processing
tial efforts to build relationships, facilitate and marketing), increasing the number of
collaboration, build the capacity of margin- workers required and extending periods
alized stakeholders to participate and link of employment. Introducing circular food
these processes to policy outcomes. systems can also increase employment
opportunities through green jobs (FAO,
5.3.4 Improving 2014a), the range of labour-intensive

employment, livelihoods activities (UNEP, 2015) and payment or other

and gender equity


incentives for the provision or restoration of
ecosystem services. High-value production
Sustainable land and water management and inclusive value chain development
offers promising pathways to improve models can create additional value and jobs
livelihoods, create new employment (Pfitzer and Krishnaswamy, 2007). Reusing
opportunities, close gender gaps and enhance resources from waste in agriculture also offers
the resilience of people and ecosystems. various opportunities to reduce pollution,

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 267
Box 5.5
Role of women in water resources management in agriculture

In Northern Africa, a study in Algeria and Tunisia on the role of women in water resources management
and water in agriculture concluded that:

ƒƒ Women play a crucial role and participate actively in irrigated agriculture management.

ƒƒ The level of education of women producers is low and their poor financial situation is due to their
social position, which limits their participation in remunerated work and decision-making.

ƒƒ There is a gap between women’s workload in agriculture and their access to land, credit and
organizations. Women are not usually members of water user/farmer associations.

ƒƒ Development programmes do not benefit everyone equally, especially within the family.

ƒƒ Participation of women in agriculture is different among countries. They participate in decisions


about crops and livestock, but less so in decisions about investment and equipment.

ƒƒ There is a lack of data at the national level on agriculture and gender, and more specifically,
agricultural water and gender.

A further study to develop and propose several gender-differentiated water indicators for integration in
AQUASTAT, as a basis to address men’s and women’s differential situations included:

ƒƒ access to drinking water;

ƒƒ economic contribution to irrigated agriculture;

ƒƒ access to economic resources in irrigated area;

ƒƒ competency in water resources management/irrigation;

ƒƒ technical management capacity;

ƒƒ participation in water governance; and

ƒƒ perceptions and practice of roles in water management.

Sources: FAO. 2014. Le rôle des femmes dans la gestion des ressources en eau en général et de l’eau agricole en particulier.
Rome. www.fao.org/3/a-bc820f.pdf; FAO. 2016. Le rôle des femmes dans la gestion des ressources en eau agricole – Phase 2. Rome. www.fao.org/3/
i5680f/i5680f.pdf

improve sanitation, create additional cal capacity of women. New technological


value and increase employment (Otoo and solutions within a circular food system can
Drechsel, 2005). open up new income-generating employ-
ment opportunities for women. Empowering
Sustainable land and water management women through education, income, organi-
can help to close the gender gap, given that zation and recognition, updating legislation
women comprise 43 percent of the agri- and ensuring representation enables them
cultural labour force globally, and over 50 to seek and acquire formal rights to land and
percent in many African and Asian countries water and to participate in future planning
(FAO, 2019a). Enhanced and stable income and decision-making (FAO, 2016a).
can improve the education and techni-

268 5. RESPONSES AND ACTION AREAS


To improve women producers’ access to Water governance has lagged behind land
water and economic resources, their partici- governance. However, OECD developed 12
pation in water management decisions and principles for efficient, effective and inclusive
establishing working conditions requires water governance (OECD, 2015), and a water
systematic disaggregation of gender data at governance indicator framework (OECD,
the national level. This requires appropriate 2018) to assess and guide better water policies
coordination among agriculture and water and reform. Recent research (FAO, forthcom-
sectors and national statistical services (FAO, ing) has led to identifying the core elements
2016a). Bb
ox 5.5 provides key indicators for of water tenure, based on how water tenure
improving knowledge and thereby address- systems function in practice and how they are
ing gender bias. legally recognized at the national level. This
analysis highlights the legal interdependen-
In this context, municipalities can also play cies across water, land and forest resource
a critical role in planning and implementing tenure systems that shape equitable and
the shift towards sustainable food systems. sustainable use, management and develop-
An example is the Milan Urban Food Policy ment of terrestrial and freshwater resources.
Pact, signed by over 200 cities, to develop Water and other sectoral policies and legisla-
sustainable food systems that are inclusive, tion need to better reflect the practical needs
resilient, safe and diverse, which provide and realities of governments and the users
healthy and affordable food to all people in a of water and land resources. The need is to
human-rights-based framework, and which secure access and user rights, as a basis for
minimize waste and conserve biodiversity sustainable livelihoods, equitable develop-
while adapting to and mitigating the impacts ment and climate resilience, particularly for
of climate change. Another example is the indigenous people, rural communities and
Quito Agri-Food Strategy, which is a multi- other vulnerable populations.
stakeholder engagement that addresses the
limited availability of fresh and nutritious FAO is supporting water governance analysis
foods in vulnerable neighbourhoods. This in several regions under its Water Scarcity
concerns the development of urban gardens, Initiative including Southern America, the
84 percent of which are led by women, for Near East and North Africa, and Asia and
home consumption and supply to local food the Pacific (Bb
ox 5.6) as a basis for improved
markets, thus creating lower costs due to water resource management, access and use
shorter supply chains and enhancing resil- by multiple stakeholders.
ience during the COVID-19 pandemic.

5.3.5 Undertaking 5.4 Action area II:


governance analysis Implementing
Simultaneously achieving diverse but inter- integrated solutions
at scale
related goals requires that sectors understand
the root causes of problems and related socio-
economic and political dynamics. Pragmatic
governance analysis facilitates understanding Evidence shows that restoring degraded
of existing institutions, how they have evolved resources, sustainable intensification and
and how the relative power and capacities of resilience can be achieved through planning
different actors influence the work of those
and implementing integrated and multi-
institutions in practice (FAO, forthcoming).
stakeholder initiatives at scale. This requires

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 269
Box 5.6
Supporting water governance analysis

Politically smart, locally led development in Western Odisha, India

The Western Odisha Livelihoods Project was a ten-year initiative (supported by the Department for
International Development, United Kingdom of Great Britain and Northern Ireland) in India to reduce
poverty by improving community water infrastructure for irrigation and flood control, and by improving
agricultural practices. Tangible achievements included improved agricultural productivity in over 70
percent of the watersheds, and reduced poverty levels in up to 75 percent. Key factors related to
success include:

ƒƒ Locally negotiated and delivered processes. Local leadership was prioritized as local experts
provided motivation, credibility, knowledge and networks.

ƒƒ Effective partnerships, based on investment in strategic relationship building that created mutual
accountability.

ƒƒ Iterative problem solving. Project development and design was undertaken with the state govern-
ment and with project beneficiaries, allowing a strategy of piloting approaches that provided for
ongoing learning and adjustment.

ƒƒ Long-term commitment. The programme enabled up-front investment in relationship building and
allowed for an adaptive approach to testing realistic solutions within the political environment.

ƒƒ An integrated, anticorruption approach emphasized community-level accountability and transpar-


ency to beneficiaries in villages.

Sources: Independent Commission for Aid Impact. 2013. DFID’s livelihoods work in Western Odisha.Report 18. https://icai.independent.gov.uk/wp-content/
uploads/ICAI-Report-DFIDs-Livelihoods-Work-in-Western-Odisha.pdf; Booth, D. & Unsworth, S. 2014. Politically smart, locally led development. Discussion
Paper. London, Overseas Development Institute. https://cdn.odi.org/media/documents/9204.pdf

Countries facing water shortage in Central and Southern America

Case studies on water governance in selected agricultural territories and river basins in Andean
countries (Plurinational State of Bolivia, Chile and Peru) and in the dry corridor of Meso-America (El
Salvador, Guatemala and Panama), and consultations with government and non-state actors, identified
and analysed challenges and gaps. This allowed the development of recommendations for addressing
water scarcity, food insecurity and resilience to climate change, and led to development of policy briefs.
The findings varied within and among river basins, but four main recommendation areas were identified,
in line with OECD water governance principles, for effective governance towards sustainable policy
goals, efficient governance for maximizing benefits of sustainable water management, and trust and
engagement of stakeholders for legitimacy and equity:

ƒƒ reform and strengthen the water-related institutional framework (sectoral and territorial);

ƒƒ improve the efficient and equitable use of water in agricultural territories for productivity and climate
resilience;

ƒƒ promote watershed management to improve water availability for production, consumption and
climate resilience; and

ƒƒ integrated management of groundwater and surface water.

Sources: Organisation for Economic Co-operation and Development. 2015. OECD principles on water governance. Paris.
www.oecd.org/cfe/regionaldevelopment/OECD-Principles-on-Water-Governance.pdf; FAO. 2021. Abordando la gobernanza del agua en territorios
agrícolas de países andinas con escasez hídrica. Policy Brief. Rome. www.fao.org/3/cb5938es/cb5938es.pdf; FAO. 2021. FAO publica estudios de gestión
del agua en países de Centro y Sudamérica. In: FAO Regional Office for Latin America and the Caribbean. www.fao.org/americas/noticias/ver/en/c/1382637

270 5. RESPONSES AND ACTION AREAS


long-term strategies, investments and inno- In practice, such approaches often cut across
vative financing, as well as partnerships to existing institutional and jurisdictional
sustain initiatives and improve livelihoods. boundaries, frequently resulting in an insti-
tutional ownership vacuum or overlapping
The wide range of agricultural and environ- authority (Ros-Tonen, Reed and Sunder-
mental policies can change and shift over land, 2018).
time, and can influence agricultural produc-
tion beyond the farm gate. Public subsidy Integrated landscape
and agricultural tariffs to promote domes- management
tic production and food security remain
Experiences on the ground show that inte-
the policy instruments of choice for many
grated landscape management, including local
developing countries. However, many other
territorial or catchment planning and appro-
policy options that have a direct bearing on
priate governance approaches, can effectively
land and water management are now main-
promote sustainable land resources and land-
stream. This section describes some impor-
scape management (Bbox 5.7).
tant options and their influences. It includes
current approaches to reconciling agricul-
Effective integrated landscape management
tural production and ecosystem services,
requires strategic legal and policy tools,
supporting agricultural productivity growth,
particularly land-use planning, and
reducing FLW, changing food consump-
appropriate incentive structures (Clinton et al.,
tion patterns and promoting sustainable
2018). Integrated land-use planning involves
diets, and the advent of circular food system
several legal frameworks, including land-use
approaches that address resource-use effi-
planning laws, zoning laws and planning
ciency. These reflect the potential benefits of
provisions within relevant sectoral legislation
adopting advanced forms of sustainable agri-
to mainstream sustainable use across these
culture that generate employment and secure
diverse frameworks at national and landscape
livelihoods and contribute to food security
levels (Lausche, 2019). National support is
and nutrition across diverse landscapes and
needed for developing capacities of national
social settings.
and decentralized institutions (including
provincial and local planning bodies and
5.4.1 Applying integrated municipalities) in integrated spatial and
approaches participatory planning tools. This includes
the use of remote sensing and diagnostic tools
Various approaches to intersectoral coor-
on the ground, and stakeholder analysis to
dination and integrated natural resources
integrate environmental and socioeconomic
planning and management have emerged
development goals and address rural–urban
that recognize the complex and intersectoral
interactions.
nature of land and water ecosystems. They
share many elements, including empha- Implementation of plans also requires robust
sizing adaptive and collaborative learning, systems for permitting environmental impact
stakeholder participation, focusing on assessments and aligning incentives through
community-based management nested subsidies and PES. Strategic environmen-
in accountable multilevel governance, and tal assessment can facilitate intersectoral
increased policy, legal and institutional and cumulative impacts on a landscape and
coherence, and coordination across sectors. provide opportunities for public participation

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 271
Box 5.7
Watershed management for resilience and SUSTAINABLE LAND MANAGEMENT
scaling out

In 2017, FAO conducted a comparative study of watershed management projects in 12 countries in


Latin America, Africa and Asia to bring together and disseminate the lessons learned and to provide
recommendations for use by practitioners in watershed-related initiatives at national, subnational and
local levels. The study concluded watershed management promotes the transition to more sustainable
production systems and practices in the crop, livestock, forestry and fisheries sectors, while enhancing
access to nutritious food for all and maintaining ecosystem services, functions and biodiversity to support
current and future human needs. It was confirmed as an effective approach for responding to global
challenges of water supply, land restoration, climate change adaptation, disaster risk management and
fighting hunger. However, the study highlighted that to meet these challenges, watershed management
initiatives must be implemented over longer time frames, and require sustained and coordinated
investment from the public and private sectors. In particular, the review identified five areas for building
robust cooperative approaches in watershed management (FAO, 2017a):

ƒƒ Institutional strengthening for improved watershed governance. Based on sound analysis of


underlying policy and institutional challenges and the causes of resource competition, interventions
should support strategic planning and institutional coordination processes and create incentives for
multistakeholder dialogue and action platforms.

ƒƒ Watershed monitoring. Priority must be given to systematic and regular collection and analysis
of data on conditions in the watersheds. Technical guidance and tools are needed to support the
selection of appropriate indicators and develop stakeholders’ capacities to monitor processes in
watersheds. Capitalizing on increased data availability and more systematic use of increasingly
available geospatial data and tools may complement on-the-ground assessments and contribute
to improved quality of environmental information while reducing time and costs.

ƒƒ Increased data availability. A more systematic use of satellite and mobile data tools in watershed
management may complement on-the-ground assessments and contribute to the improved quality
of environmental information.

ƒƒ Knowledge-sharing and learning. A platform for systematic sharing of watershed management


experiences, approaches and tools among development partners and research organizations could
avoid duplication of effort, help future programmes take advantage of the latest knowledge and
contribute to harmonization of approaches.

ƒƒ Strategic partnerships for joint action on the ground. Technical assistance projects by FAO can be
associated with larger investment programmes by the International Fund for Agricultural Develop-
ment, the World Bank and regional development banks for guidance on responsible investments
and greater impacts, as well as other international organizations working on broader landscape
management and restoration initiatives.

As an example, in the Transboundary Agro-ecosystem Management Project in the Kagera River


basin, supported by FAO, GEF and participating governments, catchment planning and management
approaches were integrated into local governance strategies to promote participatory and sustainable
land, water and biodiversity management. In Burundi and the United Republic of Tanzania, watershed
management groups were established to prioritize and oversee implementation, resulting in
improvements in food security and resolving resource conflicts. In Uganda and the United Republic of
Tanzania, participatory land-use planning enabled communities and the government to endorse the
results of catchment planning and integrated agroecosystem management for achieving agricultural
productivity, natural resources, climate, biodiversity, food security and livelihood benefits. Benefits
included building community capacity using FFSs to improve practices at farm and catchment scales,
and collaboration and exchange among local, provincial and national government bodies and among
the four riparian countries enabled identification of policy support for managing transboundary land
and water resources and livestock (FAO, 2017b, 2017c).

272 5. RESPONSES AND ACTION AREAS


in strategic decision-making (OECD, 2017;
Whitehead, Kujala and Wintle, 2017). Compe-
tent institutions and adequate financing

© FAO/John Isaac
mechanisms are needed to support a dynamic
and participatory land planning process with
regular assessments of implementation and
results by the range of stakeholders, to adjust
and update plans, and revise human and
financial allocations to meet goals and address tably accrued. Reviews of successful forest
emerging issues. The effective engagement of landscape restoration highlight engaging
all land users and other non-state actors is also private landowners and well-enforced legal
essential to ensure their specific challenges protection of forests (Mansourian, Dudley
and uncertainties are addressed, including and Vallauri, 2017). Valuation of resources
those of vulnerable groups and indigenous and ecosystem services can also be critical to
people (Ziadat, Bunning and De Pauw, 2017). identify optimal use scenarios (FAO, 2017d).

Payment for ecosystem services and other The high economic value of ecosystems
regulatory incentives can also distribute rarely translates into monetary benefits for
benefits fairly across a landscape to compen- users, perversely incentivizing activities that
sate for trade-offs. Each tool needs to be well result in resource degradation or destruction.
calibrated to the social, economic, cultural Payment for ecosystem services aligns incen-
and ecological status and goals. They are often tives and generates revenue for conservation
most successful when local authorities and through payments from ecosystem service
stakeholders take a leadership role in every beneficiaries (e.g. users of clean water or bulk
design and implementation stage. Reviews water service providers) to the service provid-
of experiences worldwide in scaling up SLM ers (e.g. upstream communities responsible
and restoration through large-scale initia- for watershed stewardship). Thus, PES can
tives demonstrate the need for substantial, provide a framework for integrated land and
long-term and targeted incentives to engage water management approaches (Bbox 5.8; Fb
ig-
the various stakeholders from the design ure 5.1 and Fbigure 5.2). Other incentives range
stage and through planning, implementation from policy measures (e.g. rights, regulations,
and monitoring, as well as clear land tenure subsidies or taxes) to green bonds or conces-
and use rights to ensure the benefits are equi- sions and marketing labels or certificates.

Box 5.8
Payment for ecosystem services: investing
in nature, investing in people

In 2010, Viet Nam adopted a PES system for forest ecosystem services that provides funding for
landscape management, generating revenue comprising about 22 percent of the overall investment
in the forestry sector (Pham et al., 2018). Under the Viet Nam Law on Forestry (Forest Environmental
Services), hydroelectric facilities, water utilities, industrial water users and aquaculture operators pay
those with legitimate forest tenure rights for ecosystem services, including erosion protection and
water quality maintenance. The government sets the payment amounts, channels them through
forest protection and development funds, and reaches over 500 000 households. In some cases, this
represents 80 percent of the annual household cash income and contributes to a 75 percent reduction in
the degraded forest area (Pham et al., 2018; Duong and Groot, 2020; McElwee, Huber and Nguyễn, 2020).

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 273
Box 5.8 (continued)

FIGURE 5.1 Ecosystem services from land, soil and water

Source: FAO. 2021. Incentives for ecosystem services. In: Land & Water. Rome.
www.fao.org/land-water/overview/integrated-landscape-management/incentives-for-ecosystem-services/en

FIGURE 5.2 Sources of incentives

Policy-driven Voluntary
Investments Investments

Prohibition of use Subsidies Green Bonds Rewards for Ecosystem


Services (RES)
Property use rights Conservation easements Voluntary farm set-asides
Marketing labels
(without certificates
Taxes/charges Permits and quotas Conservation concessions or standards)

Mandatory farm set-asides Direct Payments for Cultural and social norms
Ecosystem Services (PES)

Marketing labels (certificates/sustainability standards)

Offsets

Impact funds

Responsible sourcing of agriculture products and services

Corporate Social Responsibility (CSR)

Farmers and companies Pre-compilance to save Voluntary action with direct Voluntary action de-linked
fulfilling government costs or position private return on investment: from environmental
regulations actors on new emerging • Insetting outcomes
market • Impact marketing

Source: FAO. 2021. Incentives for ecosystem services. In: Land & Water. Rome.
www.fao.org/land-water/overview/integrated-landscape-management/incentives-for-ecosystem-services/en

274 5. RESPONSES AND ACTION AREAS


An adapted resilient watershed management

© FAO/Giuseppe Bizzarri
approach is being applied in several FAO
projects, including in Peru and the
Philippines, which includes climate change
and disaster risk management in the
overall integrated watershed management
approach. Projects also incorporate a
complexity but also present opportunities for
landscape approach, where planning, design
more comprehensive policy solutions. Despite
and implementation are carried out based
the challenges, implementing this approach
on specific areas affected by a particular
at various scales of governance reveals
hazard, including microwatersheds, multiple
lessons concerning different intersectoral
watersheds or risk reduction opportunity
assessment and analytical tools that
areas. The overall aim is to strengthen
require tailoring to produce policy-relevant
the resilience of communities and ensure
outcomes (Allouche, Middleton and Gyawali,
sustainable ecosystem services, while
2015; Albrecht, Crootof and Scott, 2018)
reducing existing disaster and climate risks
(Fb
igure 5.3).
and preventing new ones.

The Economic Commission for Latin America


Interlinkages: from integrated and the Caribbean has published a method-
water resources management ological guide on the design of actions with
to nexus approaches a focus on the nexus between water, energy
Integrated water resources management is and food for Latin American and Caribbean
now widely endorsed as the dominant global countries (Naranjo and Willaarts, 2020). It
approach to water management, supported has also developed training videos to support
by SDG target 6.5: “By 2030, implement countries to evaluate and adopt the nexus
integrated water resources management at approach in policies, plans, programmes and
all levels, including through transbound- projects in the water, energy, agriculture and
ary cooperation as appropriate”. Integrated environment sectors.
water resources management uses water
Case studies applying the nexus approach in
as an entry point to stress the need for
the Central and Southern America subregions
coordinated development and management
(irrigation policy in the Plurinational State of
of water, land and related resources, to resolve
Bolivia and Chile, multipurpose dams in the
trade-offs across multiple water users,
Plurinational State of Bolivia and Ecuador, and
acknowledging the interconnected nature of
IWRM in the Plurinational State of Bolivia)
hydrological resources, and balancing social,
demonstrate the benefits and synergies of
economic and environmental goals.
integrated policies in achieving food, water
Nexus approaches attempt to tackle the and energy security and more efficient use
challenges facing IWRM by bringing sectors of resources contributing to SDGs 2, 6, 7 and
together as an interrelated system; the 13. They require policy leadership and align-
most well-developed such approach is the ment, participation and consensus-building
water–food–energy nexus. FAO adopted this across actors and sectors, coherent planning
to develop sustainable food and agriculture and finance (Economic Commission for Latin
based on integrated land, water and America and the Caribbean, 2021). Bb
ox 5.9
ecosystems (FAO, 2014b). Nexus approaches illustrates the nexus experience in irrigated
inevitably introduce higher levels of agriculture in Asia.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 275
FIGURE 5.3 The nexus approach

Goals and Interests

Different social, economic and environmental


goals and interests related to:

Food Water Energy

Governance
Population
growth and mobility Sectoral policies and
Managing
Scenario
vested interests
Diversifying Evidence the Nexus
Development
and changing International
DRIVERS

diets

DRIVERS
STAKEHOLDER and regional
DIALOGUE trade, markets
and prices
Cultural and
societal beliefs Industrial
and behaviours Response
development
Options

Technology and Agricultural


innovation transformations

Urbanization Climate change


Resource base

Land Water Energy

Capital Labour

Source: FAO. 2014. The water-energy-food nexus: A new approach in support of food security and sustainable agriculture. Rome.
www.fao.org/3/bl496e/bl496e.pdf

5.4.2 Initiatives to address tivity and incomes, adapting and building

climate change impacts resilience to climate change, and reducing and


removing GHGs where possible. FAO mobi-
Climate-smart agriculture now helps guide lizes resources for climate-smart agriculture
actions needed to transform and reorient implementation and knowledge-sharing
agricultural systems to effectively support to contribute to FAO strategic objectives to
development and ensure food security in a make agriculture, forestry and fisheries more
changing climate. It has three main pillars: productive and sustainable. Knowledge has
sustainably increasing agricultural produc- been synthesized and updated from applica-

276 5. RESPONSES AND ACTION AREAS


Box 5.9
The nexus approach

Breaking the deadlock for irrigated agriculture in India

Gujarat is one of the driest states in India. In the 1980s, electricity subsidies encouraged private
investment in tube wells to facilitate groundwater irrigation and boost rural food and livelihood security.
Unfortunately, this led to heavily depleted aquifers.

In 2003, the state government initiated Jyotigram Yojana, which is a policy for “intelligent power
rationing” that separates electricity lines for agricultural and non-agricultural users. This limited power
to farms while allowing continuous supply for domestic and industrial users. Farmers accepted the
rationed supply because the reduced supply enabled uninterrupted service, reduced the aggregate
subsidy burden and capped groundwater withdrawals without hurting farmer welfare. The campaign
affected more than 40 million people over 3.5–4.0 million ha of irrigated agriculture, reorganized,
modernized and increased power generation capacity, and raised agricultural GDP by nearly 10 percent
while restoring groundwater levels.

The state government later introduced solar-powered irrigation pumps to explore whether farmers
would use their land to increase solar power for irrigation and earn income by selling surplus solar
energy. Over 45 months, members of the cooperative sold over 250 thousand kWh of electricity worth
USD 22 000. In 2018, this approach reached up to 33 districts. Rather than focusing on sector-based
processes, the political will to seek optimized solutions was critical in breaking the deadlock among the
sectoral stakeholders (Bird et al., 2014; Shah, 2022).

The nexus approach in the Red River basin in Viet Nam

Reservoirs in the upstream reaches of the Red River in northern Viet Nam regulate flows and generate
much of the electricity needed for the modernization and industrialization strategies of Viet Nam. The
same system supplies water for domestic use for irrigating 750 000 ha of rice in the Red River delta,
which is critical to social stability and food security. Most irrigation systems use electric pumps with
energy supplied from upstream hydropower schemes.

As water becomes scarce and competition between the energy and agricultural sectors increases,
there is still a lack of reliable and policy-relevant data and information to guide water allocation choices.
Effective intersectoral consultation is needed to address this problem and to ensure decisions on water
release and allocation are taken as part of an integrated, long-term and multisectoral strategy.

tions across the regions in a climate-smart Measures to adapt to and mitigate the impacts
agriculture sourcebook (FAO, 2017e). Success- of climate change in agriculture are part of a
ful case studies have been analysed and docu- continuum ranging from addressing the drivers
mented to show how the management of of vulnerability to those explicitly targeting the
farms, crops, livestock and aquaculture can impacts of climate change. The landmark KJWA
reduce climate risks/impacts, and balance (Bb
ox 5.10) places soil and water management
short- and long-term food security needs with practices within a systems approach for tack-
priorities to enable farmers to adapt to and ling mitigation and adaptation in agriculture.
mitigate GHGs (FAO, 2018a). Specific issues addressed under KJWA include

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 277
Soil-centred initiatives
Sustainable soil management helps to mini-
© FAO/Luis Tato

mize GHG emissions, including reducing


nitrous oxide emissions (primarily from fertil-
izer misuse), reducing methane emissions (e.g.
from paddy rice systems, and draining peat-
lands and wetlands) and reducing carbon diox-
methods and approaches for assessing adap-
ide emissions (e.g. from burning and tillage).
tation, adaptation cobenefits, mitigation,
improved soil carbon, health and fertility in Conservation agriculture is a movement that
grasslands and croplands and improved live- has been expanding worldwide and is now
stock management, including agropastoral practised on about 180 million ha of crop-
production and water management. Through land, corresponding to about 12.5 percent
their work, two UNFCCC subsidiary bodies of the total global cropland. It has increased
– the Subsidiary Body for Scientific and Tech- by some 69 percent globally since 2008. Its
nological Advice and the Subsidiary Body for adoption has been reported by 78 countries,
Implementation – emphasize agriculture and with largest extents in Southern and North-
food system vulnerabilities to climate change, ern America, followed by Australia and New
drive transformation, and identify the syner- Zealand, Asia, Russian Federation, Ukraine,
gies and trade-offs among adaptation, miti- Europe and Africa. The combined application
gation and agricultural productivity, explic- of no or minimum mechanical soil distur-
itly referencing soil and water management. bance, crop rotations that improve SOM and
the use of cover crops or mulch contribute
Adaptation requires a focus on irrigated and to reduced carbon dioxide emissions and
rainfed systems. Changes in water availability enhance soil carbon sequestration (Kassam,
and seasonal distribution driven by climate Friedrich and Derpsch, 2018).
change amplify the pressures and competition
The Global Peatlands Initiative is an effort to
for water among all water-using sectors.
improve the conservation, restoration and
Soil- and water-conservation measures,
sustainable management of peatlands in
rainwater harvesting and increasing water
over 180 countries worldwide (Chapter 1). In
storage reduce the risks of floods and
2012, FAO, the Migration of Climate Change
droughts. Sustainable improvements in
in Agriculture programme and Wetlands
land and water productivity for irrigated
International launched the organic soils and
crops under conditions of scarcity align
peatlands climate change mitigation initia-
well with adaptation. So does conservation
tive, in which ten institutions were involved.
agriculture for rainfed farming. Improving
water-conservation measures and soil health
enables farmers to diversify their systems,
amend cropping patterns and introduce
aquaculture. More attention to climate
forecasting and early warning systems will
© FAO/Giuseppe Bizzarri

also support adaptation. Fbigure 5.4 offers


a logical framework for planning and
implementing changes in land and water
management for adapting to climate change
in Eastern Africa.

278 5. RESPONSES AND ACTION AREAS


Box 5.10
Koronivia Joint Work on Agriculture

The Conference of the Parties to UNFCCC, at its twelfth plenary meeting (17 November 2017):

“Recalling decision 2/CP.17, particularly paragraphs 75–77,

Having considered the reports to the Subsidiary Body for Scientific and Technological Advice on the five
in-session workshops on issues related to agriculture,

1. Requests the Subsidiary Body for Scientific and Technological Advice and the Subsidiary Body
for Implementation to jointly address issues related to agriculture, including through workshops
and expert meetings, working with constituted bodies under the Convention and taking into
consideration the vulnerabilities of agriculture to climate change and approaches to addressing
food security;

2. Invites Parties and observers to submit, by 31 March 2018, their views on elements to be included
in the work referred to in paragraph 1 above for consideration at the forty-eighth sessions of the
subsidiary bodies (April–May 2018), starting with but not limited to the following:

(a) Modalities for implementation of the outcomes of the five in-session workshops on issues
related to agriculture and other future topics that may arise from this work;

(b) Methods and approaches for assessing adaptation, adaptation co-benefits and resilience;

(c) Improved soil carbon, soil health and soil fertility under grassland and cropland as well
as integrated systems, including water management;

(d) Improved nutrient use and manure management towards sustainable and resilient
agricultural systems;

(e) Improved livestock management systems;

(f) Socioeconomic and food security dimensions of climate change in the agricultural sector;

3. Requests that any actions of the secretariat resulting from the provisions in paragraph 1 above
be undertaken subject to the availability of financial resources;

4. Also requests the subsidiary bodies to report to the Conference of the Parties on the progress and
outcomes of the work referred to in paragraph 1 above at its twenty-sixth session (November 2020).”

In June 2019, the two subsidiary bodies also requested the secretariat to organize an intersessional
workshop to take into account two further topics:

ƒƒ sustainable land and water management, including integrated watershed management strate-
gies, to ensure food security; and

ƒƒ strategies and modalities to scale up implementation of best practices, innovations and tech-
nologies that increase resilience and sustainable production in agricultural systems according
to national circumstances.

Source: United Nations Framework Convention on Climate Change. 2018. Decision 4/CP.23. In: Report of the Conference of the Parties on its twenty-third
session, held in Bonn from 6 to 18 November 2017. FCCC/CP/2017/11/Add.1.
https://undocs.org/en/FCCC/CP/2017/11/Add.1

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 279
FIGURE 5.4 Logical framework for adapting to climate change through land and
water management in Eastern Africa

Climate change
Frequency and severity of extreme events (wind, rain, hail, frost, etc.)
Erratic rainfall distribution, unseasonal prolonged wet/dry periods.
Diagnostic: Evaluate vulnerability/

(biophysical, social, economic)


risks and apaptive capacity

Exposure Sensitivity Adaptive capacity

Potential impact
Human-managed ecosystem
(biophysical and human dimensions)
Adaptation planning

Identify biophysical Identify socioeconomic


adaptation actions adaptation actions

Implementation of adaptation plan


Implementation

Soil Water Farming system/ Strengthen


health/ management/ livelihood local
fertility use efficiency diversification institutions
evaluation & feedback
Monitoring

Improved resilience and reduced climate impact

Source: FAO. 2014. Adapting to climate change through land and water management in Eastern Africa: Results of pilot projects
in Ethiopia, Kenya, and Tanzania. Rome. www.fao.org/publications/card/en/c/96164f0a-c3dc-422d-afc3-1b3f605aefd3

280 5. RESPONSES AND ACTION AREAS


FAO is also supporting countries in peatland

© FAO/Luis Tato
mapping and monitoring (FAO, 2020d).

Recarbonization of Global Soils is a GSP


initiative for recarbonizing agricultural soils
worldwide through SOC sequestration. The
Agroecological approaches
GSP has the tools to assess and map SOC
stocks and their potential, and measure, Innovative approaches that target transitioning
report and verify SOC sequestrations and the to sustainable food systems that serve food
impact on GHG emissions (FAO, 2022a). From security and nutrition have emerged and found
2021, it has supported countries to establish application in specific land and water settings
agreements with farmer organizations and in all regions, but to different extents according
provide technical support for adopting a set to enabling policies, technical skills and market
of soil management practices and incentives forces. These include agroecology, conservation
through green benefits and carbon credits. agriculture, organic agriculture, agroforestry,
integrated crop–livestock systems and
Nature-based solutions innovations for sustainable rangeland
The NbS approach can help mitigate drought management in dry lands (Chapter 4).
and floods, notably through watershed
Agroecological approaches combine modern
and river basin management, increas-
ing and maximizing water storage capaci- agronomic methods with traditional knowl-

ties upstream that slow the release of water, edge and local food production practices, and

providing flood protection, and increasing focus on conserving water resources and soil
preparedness in low-lying lands and urban biodiversity. They can close the nitrogen cycle,
areas (FAO, 2018b) (Bbox 5.11). Such inte- improve overall productivity and provide
grated approaches need to be supported with environmental cobenefits, including reduced
land-use planning and regulations, early waste and pollution at the landscape level and
warning systems, and emergency response increased economic efficiency on farms (FAO,
and recovery plans (WMO and GWP, 2017). 2017f). Agroecology can play an important role
Even though soils constitute one of the in building resilience and adapting to climate
main reservoirs of biodiversity at the global change by building ecological buffers, SOM
level and host more than 25 percent of the and soil moisture retention.
world’s biological diversity, soil biodiversity
and overall sustainable soil management Conservation agriculture is an alternative
are neglected. The economic implications to conventional tillage; it seeks to conserve
of biodiversity loss have been profound natural resources while increasing crop
(Dasgupta, 2021). yields. It promotes minimum soil distur-
bance through direct seeding or planting and
reduced farm machinery traffic, maintaining
soil cover and using diverse plant species to
enhance biomass, SOM and soil structure.
In particular, it enhances biodiversity and
© FAO/Marco Longari

natural biological processes above and below


ground, contributing to increased water and
nutrient use efficiency and improved and
sustained crop production.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 281
Box 5.11
Nature-based solutions help mitigate droughts and floods

An example of drought adaptation is the sand dams in seasonal river beds in Southern Africa that store
increasing amounts of water as the sediments build up and the dam height is raised. Solar and treadle
pumps lift water to irrigate a second cash crop and water livestock. This cost-effective technology
deserves to be scaled out to build resilience to drought and climate variability.

A second example is catchment management in Rajasthan, India, which combines small-scale water
harvesting structures with regenerating forest cover, soils and farmland, to help improve groundwater
recharge. This has had significant impacts on water availability for a thousand villages across the
state. Flow has returned and fisheries have resumed in five rivers that used to run dry after the annual
monsoon season, groundwater levels have risen by some 6 m, productive farmland has increased
from 20 percent to 80 percent of the catchment, and forest cover in the upper catchments has
increased by 33 percent.

Examples of NbSs to reduce flooding include the construction of artificial reefs such as oyster
beds to prevent sea surges, using inland low-lying delta areas for flood prevention while cultivating
salt-tolerant rice varieties, and retaining flood water in coastal reservoirs for storage and cultivating
halophytes (salt-tolerant plants) and salt-tolerant crops. Natural and nature-based flood management:
A green guide (or Flood Green Guide) is a holistic NbS framework to support communities (WWF,
2017). The Global Water Partnership and the World Meteorological Organization have set out a range
of approaches to decrease flood risk in watershed, river and coastal area management, and in urban
areas. These approaches are combined with land-use planning, regulations, early warning systems,
evacuation plans, emergency responses and recovery plans (WMO and GWP, 2017).

Labour requirements for land preparation


Agroforestry systems
and planting are minimal, and the reduced
application of synthetic fertilizers, pesticides Agroforestry is a land-use system that inte-

and fossil fuels makes conservation agri- grates woody perennial crops and livestock

culture a practice with a low carbon foot- to balance agricultural production with

print. Simultaneous use of these techniques sustainable harvesting of forest resources.

has synergetic effects that allow sustain- Agroforestry includes forest farming, alley

able improvements in productivity and the cropping, and the use of riparian forest

environment. Conservation agriculture is buffers and windbreaks. Many of the prac-

suited to small- and large-scale farming, tices are part of traditional land-use systems,

and is appropriate where labour is in short which can benefit from introducing new

supply and agricultural input costs are high technologies to enhance synergetic effects

(FAO and ITPS, 2015). However, it requires (pest and disease control or nutrient uptake)

research–extension–farmer collaboration, and productivity. Agroforestry systems can

adaptation and fine-tuning to each context significantly improve soil fertility, especially

to develop appropriate rotations, mixes when practised with conservation agricul-

of cover crops, and practices and tools for ture. Yields of grain crops are usually higher

management to maximize protective cover under specific trees such as leguminous

and minimize use of herbicides. It also takes species than in open fields. This is attributed

time, maybe several years, to restore soil to higher SOM and the fertilizing effect of

biological functions after transitioning from decomposing foliage and dung droppings

conventional tillage. of animals grazing in the shade of trees in


agroforestry settings (Bb
ox 5.12).

282 5. RESPONSES AND ACTION AREAS


Integrated
crop–livestock systems

© FAO/Roberto Faidutti
Integrated crop–livestock systems benefit
from the synergies of crop rotations and
animal wastes to restore soil nutrients
and produce fodder crops and residues to
enhance animal productivity. They include
agropastoral systems that control grazing to
improve biomass production and livestock Mitigation and adaptation are also central
quality and productivity, and short-season concerns for sustainable and resilient live-
cropping. Examples include sedentary farm- stock systems. The livestock sector is a major
ers, who raise livestock herds, the size of user of land and water resources. Efforts
which varies according to farmland area and need to be made where possible to reduce soil
access to grazing land or fodder within the degradation, consumptive use of water and
vicinity of the farm, and transhumant pasto- pollution from intensive systems, in response
ralists who move from lowlands to highlands to water scarcity and climate change. Solu-
or may plant a crop on their seasonal migra- tions include soil and water conservation,
tion to wet-season pastures (up to 100 km improved water storage and delivery to
away) and harvest upon their return. There reduce losses, improved water productivity
are different degrees of transhumance, size in feed crops, use of manure for cropland, and
and composition of livestock herds, and use of wastewater for grazing land, buffer
types of cropping systems associated with strips and wetland management to reduce
such mixed systems (Bbox 5.13). runoff and nutrient entry into waterways.

Box 5.12
Agroforestry can enhance soil fertility

Combining agroforestry with conservation farming is emerging as an affordable and accessible


science-based solution to reduce soil depletion and increase smallholder food production. Millions
of farmers in Burkina Faso, Malawi, Niger and Zambia are restoring depleted soils and increasing crop
yields and incomes using this approach. The most promising results come from integrating “fertilizer
trees”, such as Sesbania, Gliricidia, Tephrosia and the indigenous African Acacia (Faidherbia albida), into
cropping systems. These improve soil fertility by facilitating nitrogen uptake from the air and transferring
it to the soil through their roots and leaf litter.

In Zambia, 160 000 farmers grow food crops within the agroforests of Faidherbia covering 300 000
ha. The Conservation Farming Unit observed that unfertilized maize yields in the vicinity of Faidherbia
trees averaged 4.1 tonnes/ha, compared with 1.3 tonnes/ha beyond the tree canopy. Similar promising
results have emerged from Malawi, where maize yields increased by up to 280 percent under the
canopy of Faidherbia trees compared with the zones outside. In Niger, there are now more than 4.8
million ha of Faidherbia-dominated agroforests, enhancing millet and sorghum production.
Source: Garrity, D.P., Akinnifesi, F.K., Ajayi, O.C., Weldesemayat, S.G., Mowo, J.G., Kalinganire, A., Larwanou, M. & Bayala, J. 2010. Evergreen agriculture:
A robust approach to sustainable food security in Africa. Food Security, 2(3): 197–214.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 283
A wide range of land and water management grating livestock into NDCs under the Paris
practices exist for sustainable pastoral and Agreement (reported in 92 countries), quan-
agropastoral systems in rangelands, which titative assessment and improved manage-
have evolved over generations to support ment practices to reduce emissions from
the livelihoods of sedentary and nomadic livestock systems by about 30 percent. Guide-
communities (see the focus on dryland lines include measuring and modelling soil
systems at the end of Chapter 4). carbon stocks and stock changes in livestock
production systems (FAO, 2019b), assessing
Since 2012, the FAO-led Livestock Environ- water use in livestock production systems
mental Assessment and Performance multi- and supply chains (Davies et al., 2018) and
stakeholder global initiative has aimed to environmental performance of feed additives
accelerate sustainable development of the in livestock supply chains (FAO, 2020e).
livestock supply chain. Support includes inte-

Box 5.13
Reducing risks, addressing vulnerability and enhancing
pastoralist resilience in Africa

Pastoralism is the main livelihood for about 268 million people across Africa’s dry lands, from the
Sahelian West to the rangelands of Eastern Africa and the Horn of Africa, and the nomadic populations
of Southern Africa. It represents one of the most viable, and sometimes the only suitable, livelihood
options in dry lands. It makes enormous contributions to social, environmental and economic
well-being in dryland areas and beyond. The mobility of pastoralists exploiting animal feed resources
along different ecological zones represents a flexible response to a dry and increasingly variable
environment. Pastoralism ensures livestock access sufficient high-quality grazing, and creates
economic value by converting scarce natural resources into meat, milk, income and livelihoods.

Yet, pastoral livelihoods have been severely undermined by decades of neglect (with as low as 1
percent of government budget allocation), violence and displacement, insecure land rights and access,
deteriorating natural resources, climate variability and change, and growing risk of animal and zoonotic
diseases. The pastoral system is increasingly threatened despite demonstrated remarkable resilience
and being well adapted to manage the risks and uncertainties faced in Africa’s dry lands. Pastoralist
populations are increasingly vulnerable to malnutrition and food insecurity as their capacity to adapt to
and recover from crises declines in the face of recurrent and often overlapping shocks.

FAO advocates for enhanced efforts and more robust partnerships among all actors to strengthen the
resilience of pastoral livelihoods through a deliberate mix of short-, medium- and long-term actions
across the humanitarian–development–peace nexus. Exchange of experiences and analysis by experts
and partners in Western and Eastern Africa and consultation under FAO resilience hubs in Kenya and
Senegal in 2018 led to several recommendations, including engaging pastoralists in policymaking and
decision-making, engaging local, national and regional partners to address the cross-border dimension
of pastoralism, and developing livelihood-based information and monitoring systems.

Moreover, the development of an enabling policy environment for sustainable pastoral and agropastoral
systems in marginal and fragile ecosystems should also consider incentives for the private sector to
flourish and continued investments in innovation and technologies.
Source: FAO. 2018. Pastoralism in Africa’s drylands: Reducing risks, addressing vulnerability and enhancing resilience.
Rome. www.fao.org/3/CA1312EN/ca1312en.pdf

284 5. RESPONSES AND ACTION AREAS


© FAO/Giulio Napolitano The management of dry lands is an important
aspect of sustainable agriculture land and
water management and restoring degraded
lands, given their extent, populations affected
in terms of poverty and climate change and
variability, and opportunities for investment.
5.5 Action area III: The focus on dryland systems at the end of

Embracing Chapter 4 analyses drivers and pressures and


appropriate responses for sustainable land

innovative and water management including dryland

technologies and
cropping, and livestock and rangeland
management.

management 5.5.1 Sustainable


A wide range of innovative technological and soil management
management responses are now available and soil health
within the immediate farming domain for
Led by FAO through GSP, there is an iterative
rainfed (dry lands) and irrigated farming.
process of country to global assessments,
They include practices for achieving
international symposia and outcome
sustainable soil management, restoring soil
documents outlining status, threats and
health, improving soil water management,
responses. This has led to data collection,
accessing non-conventional water resources,
raised awareness, action plans, solutions and
adopting NbSs, managing environmental
guidance for addressing loss of SOM and SOC
risks, coping with climate change, reducing
(issued in 2017), soil erosion (2019), nutrient
carbon emissions, and using information and
imbalance and use of fertilizers (2019), loss of
communications technology (ICT) and big
soil biodiversity (2020), soil pollution (2021)
data. They also include prospects for scaling
and salt-affected soils (2021).
up and implementing technical solutions.

A wide range of proven soil- and water-


Agricultural agencies need to update the
conservation technologies are available to
capacity and tools for managing GISs,
reduce runoff, tackle soil erosion, restore
developing and using maps and plans, and
SOM and SOC, and improve soil fertility. These
monitoring trends and impacts. These tools
include regenerative agriculture practices to
are now required to manage agricultural
build soil health and reverse adverse effects
production and mitigate GHG emissions
of tillage, such as conservation agriculture,
from land.
intercropping, agroforestry and sustainable

For irrigated agriculture, more capital- rangeland management. Successful interven-

intensive options are available for augment- tion approaches include FFS approaches for

ing water resources and modernizing capacity-building and information exchange,

irrigation systems. Improving crop water and watershed or other territorial planning

productivity and water-use efficiency and management.

and investing in non-conventional water


resources are among the technical options
for improving irrigation production systems.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 285
The Voluntary guidelines for sustainable soil The Global Soil Doctors programme launched
management (FAO and ITPS, 2017) guide stra- in 2020 is a farmer-to-farmer training
tegic and context-specific decision-making initiative to enhance farmer capacities and
at all levels to promote practices that address knowledge in sustainable soil management
soil threats and the means to restore and at the farm level. The programme has been
maintain soil health (Bbox 5.14). The GSP successful in some countries in Asia and
secretariat and the Intergovernmental particularly useful in locations where soil
Technical Panel on Soils have established a extension services are weak or absent. It aims
protocol to assess the voluntary guidelines’ to empower farmers within a community
interventions and to ensure improvements by training a lead farmer (soil doctor) in
in production systems, ecosystem restora- diffusing methods and tools to detect and
tion and carbon sequestration are sustainable provide practical solutions to soil degrada-
(FAO and ITPS, 2020) and to address the tion. It provides educational materials and a
interlinked problems of land degradation, soil testing kit for assessing soil conditions
climate change and biodiversity loss. and a set of good practices under the sustain-
able soil management voluntary guidelines.
The WOCAT database, endorsed by UNCCD for
country sharing of best practices, provides Despite significant growth in the use of
many SLM practices and experiences. There is chemical fertilizers in some countries, such
a need to encourage further sharing of tech- as in sub-Saharan Africa, soil testing and
nologies, innovation and results from differ- fertilizer use are low due to high costs,
ent ecological and socioeconomic contexts, weak supply chains and lack of extension to
and across actors and institutions, for support their wise use on farms. Fertilizers
example, to reduce soil and water contami- require tailoring to site-specific ecological
nation, ameliorate soil salinization, restore and socioeconomic conditions. The Interna-
soil biodiversity, and improve water use and tional Code of Conduct for the Sustainable
reuse in rainfed and irrigated systems. Use and Management of Fertilizers (Fertil-
izer Code) offers guidance to tackle misuse,
Soil and crop management practices should underuse and overuse of fertilizers, bearing
provide a favourable environment for soil in mind nutrient imbalances and soil and
organisms and their biological activity, such water pollution (FAO, 2019c).
as reducing soil disturbance, maintaining soil
cover and rotating crops. Inoculating selected
Bradyrhizobium bacterial strains in soybean
production is a successful and cost-effective
biotechnology used in Argentina, Brazil and
Uruguay to replace mineral nitrogen fertiliz-
ers and to avoid leaching and volatilization of
©FAO/Olivier Asselin

nitrogen compounds (Franco, 2009). Wider


uptake of such practices requires filling
knowledge gaps, good research–extension–
farmer links and supportive policy (Bbox 5.14).

286 5. RESPONSES AND ACTION AREAS


Box 5.14
Soil biological diversity at the heart of sustainable soil management

The Post-2020 Global Biodiversity Framework for soil biodiversity and ecosystem services is critical to
the success of the recently declared United Nations Decade on Ecosystem Restoration (2021–2030).
Maintaining soil biodiversity is an effective NbS to address degradation, food insecurity, climate change
and poverty-related problems facing humanity from the field scale to the global scale.

This requires increased attention to the conservation and sustainable use of biodiversity as part of
sustainable soil management practices to restore SOM and the substrate for soil organisms, and to
increase favourable conditions for soil biological activity. This includes the vital role of soil organisms in
plant growth and nutrition (on which crop and livestock productivity depend), and mitigating processes
of land degradation.

Soil biodiversity is the large reservoir of organisms in the soil responsible for a multitude of soil functions
from microbial bacteria, fungi and microfauna (nematodes and protozoa) that are invisible to the naked
eye, to mesofauna (mites and springtails) to macrofauna (centipedes, millipedes, ants, ground beetles,
spiders and earthworms) and megafauna (moles and other vertebrates) that live on and in the soil. More
than 40 percent of living organisms in terrestrial ecosystems are directly associated with soils during
their life cycle.

These soil organisms are largely invisible yet make a vital contribution to agricultural production. They
make macro and micro nutrients available for growth and nutrition, and minimize cost and dependence
on synthetic fertilizers in agriculture by:

ƒƒ providing the nutrients in soils that plants need to fix carbon from the atmosphere and create biomass;

ƒƒ playing a vital role in the physical breakdown of plant residues and allowing soil microorganisms to
liberate nutrients and energy bound up in the organic plant material;

ƒƒ participating in filtering, degrading and immobilizing contaminants in water and soil; and

ƒƒ including “ecosystem engineers” that modify soil porosity, water and gas transport, and bind
together soil particles into stable aggregates that hold the soil in place, reducing soil erosion, and
retaining soil moisture and nutrients.

Source: FAO, Intergovernmental Technical Panel on Soils, Global Soil Biodiversity Initiative, Convention on Biological Diversity & European
Commission. 2020. State of knowledge of soil biodiversity: Status, challenges and potentialities. Rome, FAO.
https://doi.org/10.4060/cb1928en

Integrated soil fertility management is a tive interactions. However, for lasting effects
strategy for combining organic and inor- on soil health, it is essential to avoid soil
ganic mineral nutrients. It relies on nutri- pollution and soil tillage. The integrated soil
ent application from organic inputs such as fertility management framework has proved
compost, manure and inorganic fertilizers, its suitability for an extensive range of soil
together with growing nutrient-fixing crops fertility conditions, agroecological zones and
in rotations, growing cover crops and mini- cropping systems (Roobroek et al., 2015).
mum tillage. Mixing organic and inorganic
(mineral) fertilizers can optimize nutrient Combining organic and mineral fertilizers

availability according to soil deficits, crop and implementing sustainable soil manage-

type and growth stage, and has many posi- ment practices can also support nutrition-
sensitive agriculture (FAO, 2014c). Research

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 287
Soil water is usually the only source avail-
able for producing biomass, and this depends
mainly on the soil’s capacity to store water
during the dry seasons. Water availability
correlates directly with SOC, soil structure
and nutrient availability (FAO, 2020f). Deep,
non-stony, non-saline, fine-textured and
© FAO/Marco Longari

organic carbon-rich soils have the highest


available water-holding capacity and so are
more resilient to droughts.

A package of incentives for sustaining


ecosystem services can overcome barriers

data confirm the micronutrient superiority of to more sustainable agricultural systems,

some lesser-known cultivars and wild vari- combining public policies to improve

eties over other, more extensively utilized farm productivity with those that reward

cultivars. For example, sweet potato culti- conservation practices and partnering

vars differ in their carotenoid content by two with green business strategies. Case

orders of magnitude or more, the protein studies illustrate the value of incentives

content of rice varieties can range from for groundwater recharge and watershed

5 percent to 13 percent, and the provitamin A management for water supply and quality

carotenoid content of bananas can vary from (FAO, 2022b).

1 μg/100 g to 8 500 μg/100 g among culti-


vars. This shows the importance of selecting Enhancing water productivity
crop varieties for food composition and not in rainfed systems
just for yield (Burlingame, Charrondiere and Recent productivity measures have focused
Mouille, 2009). on nutritional water productivity, “better
nutrition per drop” (Renault and Wallender,
5.5.2 Addressing drought 2000), linking water, agriculture and nutri-

in rainfed systems tion (Lundqvist and Unver, 2018). FAO is


exploring the concept of nutritional water
Improving rainfed agricultural production productivity and expanding the methodol-
requires optimizing soil moisture and making ogy to include crop production, nutrient
the best use of rainwater by maximizing content and economic value (Lundqvist et al.,
infiltration and minimizing runoff and 2021). Another approach from an ecological
evaporation. Soils provide the buffer between perspective suggests using “less drop per
rainfall events and crop water demands, and crop” as a measure of wise water use.
help protect crops against drought, flood
and climate variability. Sustainable soil The scope for improving water productivity
management practices, including integrated varies with production systems and regions.
soil fertility management, complement and Water productivity is higher in Australia,
enhance buffering for nutrients and pH Europe, Northern America and the Yellow
through building up SOM and improving River basin in China. Areas with the highest
cation exchange capacity. They offer the potential for water productivity gains are
potential to increase production and provide sub-Saharan Africa and South, Southeast and
reliable yields. Central Asia.

288 5. RESPONSES AND ACTION AREAS


Box 5.15
Rainwater harvesting serves protected cropping in Lebanon

Lebanese growers using protected cropping under glass were concerned about the reliability of
groundwater and its overexploitation, and have turned to rainwater harvesting from microcatchments
as an alternative water source. National guidelines for greenhouse rainwater harvesting systems offer
information to growers on all aspects of design and installation. They provide a brief overview of
greenhouse types used in Lebanon, irrigation scheduling, crop water requirements and main crops
grown in protected environments. They focus on microcatchment rainwater harvesting systems
(direct/indirect pumping and gravity fed) and describe the main system components that follow the
water flow, starting with the catchment area, collection and conveyance system, rainwater quality, and
pretreatment, storage and pumping and distribution systems.

FAO has developed a multicriteria assessment method for selecting water harvesting methods. Each
criterion pools several dedicated indicators to appraise the suitability of techniques and support
decision-making.
Source: United Nations Development Programme & Lebanese Ministry of Energy and Water. 2016.
National guideline for rainwater harvesting systems. Beirut.

The water productivity index (WPI) was Harvesting rainwater


introduced as a tool to support policymak-
Water harvesting offers opportunities for
ers in making more informed decisions on
improving agricultural productivity in dry
water resources management and allocating
regions. It can boost yields two to threefold
scarce water resources. However, changes
over rainfed production, especially when
in WPI can be attributed to factors other
combined with minimum-tillage methods
than water, such as national macroeconomy
that enhance water conservation (Oweis,
structure, applied technologies and manage-
2016). Water harvesting can also augment
ment practices, and climatic conditions
available irrigation water supplies (Bb
ox 5.15)
that short-term policy measures cannot
and improve household access to water.
modify. Furthermore, WPI does not reflect
the uneven spatial distribution of resources
The WOCAT water harvesting guidelines for
or geographic conditions and may mask
good practice (WOCAT, 2013) provide compre-
local differences. However, WPI is useful for
hensive and practical advice covering a wide
assessing incremental benefits at different
range of flood, macro and micro catchments,
scales, from individual crops, farms and irri-
and rooftop/courtyard water harvesting tech-
gation systems to basins and regions.
niques. Although rainwater harvesting is a
common practice, even for household use, it
is illegal for households to capture rainfall
in the arid states of Colorado and Utah in the
United States of America. Likewise, in Chile,
the Water Code enables farmers to abstract
© FAO/Giulio Napolitano

surface water and groundwater, but does not


allow them to harvest surface water runoff
from the land. Uneven water access, due to
the allocation of rights of water access to
initiatives with the greatest market value, is

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 289
withdrawal for irrigation (Hoogeveen et al.,
2015). The implication is that much of the
© FAO/John Isaac

water diverted for irrigation never reaches


the crops and is lost through seepage in canal
systems and poor on-farm water manage-
ment, creating further problems such as
waterlogging, salinity and pollution.
a critical issue as large-scale farmers with
inherited or traded rights are able to extract As water demand for agriculture increases,
disproportionate amounts of water (a large the 2018 United Nations SDG 6 synthesis
share for export crops), which can compro- report on water and sanitation suggested
mise smallholder and rural community access that agriculture (mainly irrigation), as the
(OECD, 2017; Lobos, 2021). largest user of water, offers the most signifi-
cant potential for saving water: “Saving just
The conjunctive management of groundwater a fraction can significantly alleviate water
and surface water needs to be enhanced in stress in other sectors, particularly in arid
countries through integrated watershed/basin countries where agriculture consumes a
management with all stakeholders. It also considerable amount of the available water
needs to take into account the current water resources” (United Nations, 2018). However,
rights and climate change context, to although it appears that significant water
safeguard the interests of smallholders and savings are possible, recent research shows
other water users reliant on ecosystems that in many instances, “real” water savings
services for their livelihoods. are much less than expected (see the follow-
ing section on water-use efficiency).
5.5.3 Coping with
water scarcity in Meeting the increasing demand for food from

irrigated systems
limited land will lead to increases in irrigation
on current rainfed croplands, where suffi-
Agriculture dominates freshwater with- cient water resources are available (see FOFA
drawals, mainly through irrigated agricul- scenarios in Chapter 3). There is renewed
ture, which accounts for almost 70 percent investment interest in irrigation, but devel-
of all freshwater withdrawals to produce oping new irrigation schemes will present
40 percent of the world’s food, fibre and significant challenges, and so will modern-
fuel needs. Irrigation (see Chapter 1) can izing existing systems that have long been
remove the uncertainties of inadequate and criticized for their poor overall performance,
unreliable rainfall and significantly increase not just in terms of water-use efficiency. The
crop production and water productivity lack of institutional and economic capacity
when adequate water resources are avail- may constrain development. In addition, the
able. However, irrigation has a reputation location and productive potential of econom-
for inefficiency; in many instances, this is ically water-scarce croplands are unknown
undeserved. The global average agricultural (Rosa et al., 2020).
water-use efficiency is estimated to be 55
percent, with national figures ranging from Improving water-use efficiency
40 percent to 60 percent, measured as a ratio Terminology around water-use efficiency is
of crop water evapotranspiration to water confusing; different definitions exist across

290 5. RESPONSES AND ACTION AREAS


disciplines and scales. For example, the clas- in water-saving technologies did not return
sic definition (ratio of water consumed by a surplus water for others to use, as might be
crop to the amount of water withdrawn from assumed (FAO, 2017g). Instead, farmers used
a river or groundwater) is useful for planning the extra water to expand irrigated areas or
and designing irrigation schemes. However, switch to crops that were more water inten-
when it is used to evaluate performance, it sive, thereby increasing water productivity to
assumes that any excess water applied is lost improve farm incomes, but not resulting in
and ignores that those “losses” may be used water resource conservation or redistribution.
elsewhere by others downstream. The multi-
plicity of definitions across disciplines and the Similarly, on irrigation schemes described

lack of agreed terminology can lead to serious as “inefficient”, seepage from irrigation

misunderstandings at technical and policy distribution channels or excessive applica-

levels (Balasubramanya and Stifel, 2020). tion returns to the river through soil drain-
age or recharge to shallow groundwater, and
Reducing water losses is never easy; confu- provides a source of water for irrigators and
sion over efficiency measurements adds other users downstream. Thus, improving
to the problem. This has led to traditional “efficiency” on upstream farms can reduce
approaches to improving water use in irriga- water available to others downstream. As
tion, such as lining canals and switching to such, irrigation’s reputation for using too
trickle irrigation, being challenged. Studies much water is not always justified (Kay,
show that what appear to be more “effi- 2020). Bbox 5.16 illustrates the complexity of
cient” technologies can increase water use investing in technologies to save water and
rather than reduce it. FAO has published a the unexpected consequences.
review demonstrating that farmers investing

Box 5.16
Montana versus Wyoming: sprinklers, irrigation efficiency
and recapturing return flows

In 2012, a legal case in the United States of America demonstrated the serious and unexpected impacts
of increasing irrigation efficiencies to reduce water losses (called “return flows”). The Yellowstone River
basin is nearly equally divided between the states of Montana and Wyoming. In 1950, the two states
agreed to apportion the available water for irrigation and other purposes. However, in 2007, following a
severe drought between 2000 and 2006, Wyoming invested in sprinkler and trickle irrigation to increase
irrigation efficiency to use its limited water allocation better. But Montana had long benefited from the
inefficiencies in Wyoming. The impact of increasing efficiency in Wyoming was to reduce the return
flows to the detriment of Montana. Montana alleged sprinklers increased water consumption from 65
percent to 90 percent, reducing return flows from 35 percent to only 10 percent. Montana argued that
Wyoming should have imposed administrative requirements to offset these adverse effects on Montana.

This was a complex legal case, and dealt with the laws of the doctrine of recapture. Can farmers
recapture their water losses by increasing their irrigation efficiency when others downstream have
long benefited from those losses? The court held that such improvements were permitted under the
Yellowstone River agreement. However, this may not be the case for irrigation schemes in other parts
of the world, where legislation is unclear or non-existent.
Source: MacDonnell, L. 2012. Montana v. Wyoming: Sprinklers, irrigation water use efficiency and the doctrine of recapture.
Golden Gate University Environmental Law Journal, 5(2).

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 291
The FAO Water Scarcity Programme devel- climate would score high efficiency values,
oped the Real Water Savings (REWAS) tool whereas a developing country in a semi-arid
to assess “real” water savings in irrigation, climate, dependent on irrigation for staple
rather than what is described as “dry” water food crops, would score low values. Thus,
savings (Seckler, 1996). The guiding principle results published for SDG target 6.4 require
is to “follow the water” (Kaune et al., 2020). careful interpretation in context to be help-
Bbox 5.17 illustrates this tool, based on the ful in decision-making. Steps are underway
principles of water accounting. in many countries to clarify what efficiency
means in a local context.
To add to the confusion over efficiency defini-
tions, SDG target 6.4 requires the increase in The confusion over the meaning of effi-
water-use efficiency, which can be described ciency highlights the complex relationships
as the ratio of the gross value added per among water, agriculture and poverty, and
unit of water, measured in United States the essential need for a common language
dollars per cubic metre. This assesses the among multiple disciplines to inform
economic and social use of water resources decision-makers on water resources plan-
in terms of the value added when using ning and management. Policymakers must be
water in different sectors of the economy clear in the terminology they use, and under-
(United Nations, 2018). Using this metric in a stand the misconceptions in common use.
highly industrialized country in a temperate

Box 5.17
“Following the water” to assess “real” water savings

A river basin study in Nepal reported irrigation water savings of 75 percent. However, the study failed
to adhere to the “follow the water” principle as it assumed that all return flows were losses. Fully
accounting for all the water flows found that 80 percent of the “losses” were return flows, which were
recovered and used by irrigators downstream.
The original study focused only on the amount of water diverted for irrigation and the amount used by
crops. The REWAS analysis focused on the return flows and non-beneficial consumption (dotted yellow
boxes in Figure 5.5) as these were recoverable and could be available for others to use.
The results showed “real” water saving in the river basin was only 6 percent.

FIGURE 5.5 Sources of incentives

Beneficial
(irrigation + precipitation)

Consumption
Water use

Non-beneficial

Recoverable

Return Flows

Note: dS refers Non-recoverable


to water storage dS
in the soil.

Sources: Droogers, P., Kaune, A., Opstal, J. Van, Perry, C. & Steduto, P. 2020. Training manual: Crop water productivity options to achieve real water savings.
FutureWater Report 199. Wageningen, FutureWater.
www.futurewater.nl/wp-content/uploads/2020/05/FAO_Training_v11.pdf;
Kaune, A., Droogers, P., Van Opstal, J., Steduto, P. & Perry, C. 2020. REWAS: REal WAter Savings tool: Technical Document. FutureWater Report 200. Rome.
www.futurewater.nl/wp-content/uploads/2020/06/FAO_REWAS_v08.pdf

292 5. RESPONSES AND ACTION AREAS


Water accounting and auditing when water is scarce and risks and uncertain-
ties over water availability increase. The aim
A growing number of international orga-
is to use water-related information better
nizations are promoting water accounting
when matching and adapting coping strat-
and auditing as an invaluable tool for water
egies to different biophysical and societal
resources planning and management, mainly
contexts (Bb
ox 5.18).

Box 5.18
Water accounting and auditing

When there is competition for scarce water resources, any analysis must go beyond a simple water
balance and account for proper comparison and assessment of resources and all water uses. Water
accounting and auditing provides the framework. Water accounting brings together the hydrological
cycle water balance with assessments of spatial and seasonal variations in the climate and medium-
and long-term changes in demand across all water users (Figure 5.6). It also informs water resources
planning and infrastructure investment. Water auditing provides a connection between water accounting
and effective water governance by providing sound evidence for decision-making. It offers qualitative
judgments to the water account and puts the recommendations of water accounting into the broader
societal context of water management (Karimi, Bastiaanssen and Molden, 2013; FAO, 2016b, 2018c).

FIGURE 5.6 Water accounting brings together all water flows and uses

Process
Gross inflow

Net inflow

Available

Beneficial
Depleted
Surface and subsurface
flows precipitation

Non-process
g/+S8$+$M.)"4&

INFLOW Non-beneficial
g/+S%$./<$%"84$&

Utilizable
3$./<$%"84$&

Uncommitted Non-utilizable
T+./??)f$(&
Outflow

Committed
@/??)f$(&

Change in
storage

Source: Kaune, A., Droogers, P., Van Opstal, J., Steduto, P. & Perry, C. 2020. REWAS: REal WAter Savings tool: Technical Document. FutureWater
Report 200. Rome. www.futurewater.nl/wp-content/uploads/2020/06/FAO_REWAS_v08.pdf

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 293
Box 5.18 (continued)

Combining water accounting with data from WaPOR

Water accounting is a powerful tool for accurately assessing crop water consumption on irrigation
schemes. Many countries still do not have the capability to measure the amount of water they use
for irrigation. Some rely on measuring the volume of water diverted, but this is usually significantly
greater than the amount consumed by the crops. Measuring crop water use rather than depending on
irrigation diversion data is vital to producing an accurate water budget for a scheme or basin, mainly
when irrigation takes a large percentage of the blue water resource. Water accounting and remote
sensing offer a solution.

An example is the Litani River basin in Lebanon. Water accounting used the FAO WaPOR data portal,
which uses remote-sensing technologies to monitor and report agricultural productivity over Africa and
the Near East to overcome limited data availability. The system measures irrigated crop areas and water
consumed by crops, thus providing a more accurate picture of water use; rather than relying on patchy
water withdrawal data (FAO and IHE Delft Institute for Water Education, 2019).

The WaPOR data portal also provides gross biomass water productivity data across Africa and the Near
East (Figure 5.7).

FIGURE 5.7 Gross biomass water productivity, 2020

0 - 0.1 kg/m3

0.1 - 0.5 kg/m3

0.5 - 1 kg/m3

1 - 2 kg/m3

2 - 3 kg/m3

>3 kg/m3
Dotted line represents approximately the Line of Control in Jammu
and Kashmir agreed upon by India and Pakistan. The final status of
Jammu and Kashmir has not yet been agreed upon by the parties.

Final boundary between the Sudan and


South Sudan has not yet been determined.

Final status of the Abyei area is not yet determined

Source: FAO. 2020. WaPOR: The FAO portal to monitor WAter Productivity
through Open access of Remotely sensed derived data. In: FAO. Rome.
https://wapor.apps.fao.org/home/WAPOR_2/1

294 5. RESPONSES AND ACTION AREAS


Modernizing medium- and

© FAO/Sergey Kozmin
large-scale irrigation schemes
Medium- and large-scale irrigation schemes
are generally owned and operated by govern-
ment agencies that supply water and services
to individuals and groups of smallholder irrigation performance by upgrading and
farmers. Although over the past 50 years, improving all aspects of an irrigation scheme
large-scale canal irrigation has made a to respond to modern farming requirements.
significant contribution to increasing food It is driven partly by farmers who want more
production, reducing hunger and poverty, flexible and reliable water delivery and partly
increasing employment and securing rural by governments concerned about making the
livelihoods for many millions of smallholder best use of available water resources and the
farmers, critics have suggested that the plan- rising costs of scheme construction, opera-
ning and design have remained technically tion and maintenance.
stagnant (Plusquellec, 2014). Canal irriga-
tion continues to suffer from problems of Modernization usually requires upgrad-

poor flow regulation to farmers, and there ing technologies – the “hardware” – which

have long been significant discrepancies goes beyond rehabilitation, and which

between design assumptions and actual replaces only what is already there. This is

performance – hydraulically, economically the more visible part of a system. Options for

and socially. Water scarcity exacerbates this improvement include installing networks and

situation, which is now the main driver to control structures, automation, lining canals,

improve performance by modernizing exist- constructing reservoirs and installing modern

ing schemes and designing new schemes to information systems to improve manage-

overcome past problems. ment and control. As more than 90 percent


of irrigation globally uses surface irrigation
Modernization is a complex process and methods, most technology upgrades must
is not just about saving water. It requires focus on simplifying canal management and
significant changes in the way schemes are surface irrigation performance. Moderniza-
planned, designed and managed. In the 1990s, tion also requires hardware improvements on
FAO coined modernization as “a process farms, such as control systems that simplify
of technical and managerial upgrading (as canal management and provide farmers with
opposed to mere rehabilitation) of irrigation flexible and reliable water supplies. Reli-
schemes with the objective to improve ability creates confidence in managers and
resource utilization (water, labour, economic, farmers, enabling them to switch off water
and environmental) and water delivery to supplies when irrigation ends. Where appro-
farms”. Implicit in modernization is a shift priate, farmers can also consider switching
from traditional supply-driven irrigation to from gravity-fed to pressurized sprinkler
demand-driven irrigation and introducing and trickle irrigation to improve control over
the concept of providing irrigation services to water application. Installing drainage can help
farmers (FAO, 2007a). to remove excess water and control salinity.

Modernizing irrigation is a means of rectify- Modernization does not refer only to


ing past mistakes by taking a more holis- high-technology solutions such as canal
tic and coordinated approach to improving automation and pressurized sprinkler and

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 295
trickle irrigation systems. Such technolo- large irrigation schemes elsewhere. The
gies have a role to play, but significant methodology seeks to stimulate a critical
improvements are also possible using simple sense among scheme managers to diagnose
gravity-fed technologies, such as night stor- and evaluate obstacles, constraints and
age to balance supply and demand at farm opportunities, and develop a consistent
level and fixed broad weirs to simplify water modernization strategy. A step-by-step
level and discharge control (Horst, 1998). approach is offered to convert a complex
set of circumstances into simple elements
Equally important is upgrading the that can be explored and improved. FAO
management and institutional structures is developing a similar methodology for
that govern irrigation, the “software”, pressurized systems, Mapping System and
which is much less visible than the hardware. Services for Pressurized Irrigation, to enable
This includes increasing the capacity scheme managers to optimize sprinkler
and capability of organizations to provide and trickle systems designed to respond to
services to farmers appropriate to modern irrigation on demand.
irrigation farming (Kay and Renault, 2004;
Kay, 2020). Improvements include changing Enhancing smallholder
the traditional “top-down” approaches to irrigation
scheme management to ones that accept
Irrigation is an integral part of smallholder
farmer participation in management
farming for many millions of smallhold-
decision-making at all levels. This may
ers across the Near East and North Africa,
involve transferring scheme management
sub-Saharan Africa and Asia. Smallholder
and maintenance at the tertiary level to
irrigation is usually farmer led, and refers
farmer organizations and providing a reliable
to individuals or small groups of farmers
water delivery service for which farmers are
who own and operate their systems inde-
willing to pay. Above all, these changes need
pendent of government control (Ffigure 5.8).
strong political support at the highest level
Individual farms are small, 2–5 ha in size,
and an enabling environment that provides
and farmers exploit water resources in many
farmers with incentives, manageable
ways, including using surface water and shal-
risks and uninterrupted access to markets
low groundwater, water harvesting, natural
(FAO, 2007b).
springs and wetlands, spate flows in rivers
FAO developed the Mapping Systems and and recession flows in flooded areas (Izzi,
Services for Canal Operating Techniques Denison and Veldwisch, 2021).
methodology in 2007, designed to assist
Smallholder irrigation systems exist in almost
technical experts, irrigation professionals
all agroecological zones. However, they are
and scheme managers in modernizing
particularly important in arid and semi-arid
schemes (FAO, 2007a). The entry point is
areas where subsistence farming prevails
canal operation, but the focus is on identifying
on marginal lands and where unpredictable
targets, including finance and water use,
and inadequate rainfall limits crop produc-
and meeting environmental requirements.
tion. Productivity is typically well below
Although based mainly on FAO experiences
that of medium- and large-scale irrigation
in Asia, the Mapping Systems and Services
schemes due to the lack of modern water
for Canal Operating Techniques is a generic
control technologies, agronomic practices,
methodology that applies to medium and
farm inputs, access to markets and economies

296 5. RESPONSES AND ACTION AREAS


FIGURE 5.8 Farmer-led irrigation development

The most familiar contexts of FLID

Mountains & Wetlands Urban outflows Near large-scale


schemes Ponds, rivers
hillsides (streams (dambos & (effluent & & groundwater
& springs) fadama) drainage flows) (tail & drainage
water)
In wetter areas, farmers In wetland areas, bunds In urban and peri-urban Along canals, drains and In floodplains and flat
often build diversion and drains are settings farmers use a tailwater outlets of many areas where groundwater
structures on mountain constructed to control variety of wastewater major irrigation schemes, is well below the root
streams leading water shallow groundwater sources, such as the individuals divert or zone, but still shallow
into gravity irrigation levels just below the outflows from pump water to land on enough to access with
systems using canals root zone to enable wastewater treatment the periphery. Soils are open wells (typically
and flood irrigation. plant growth through plants and open roadside often marginal and water < 15 m), petrol and diesel
Where the topography capillary action. drains. Water-quality unreliable. Technologies pumps, bucket-and-rope
and financial means issues from sewage and are similar to open water systems, and solar-
allow, plastic pipelines urban pollutants are bodies. electric pumps are used.
feeding hoses or potentially serious. Similar technologies are
sprinklers are used. used alongside rivers and
within reservoirs of dams.

mostly farmers’ groups mostly individual farmers

Note: FLID = farmer-led irrigation development.

Source: Izzi, G., Denison, J. & Veldwisch, G. 2021. The farmer-led irrigation development guide: A what, why and how-to for intervention design. Washington,
DC, International Bank for Reconstruction and Development/The World Bank.
https://pubdocs.worldbank.org/en/751751616427201865/FLID-Guide-March-2021-Final.pdf

of scale. However, the systems benefit from as de facto landowners and displace women
deep-rooted indigenous knowledge, good soil farmers from seasonal farming in wetlands.
and water management practices, and reliable
local social networks that support subsistence In general, farmers are more interested in

farming. Simple measures that do not change saving money than water, but adopting the

local management practices can improve best water management and agronomic

schemes, such as lining canals for local practices can benefit both. Best practices

spring-fed schemes. Localized approaches to include: ranking irrigation highly within

the transfer of knowledge and technology that farm management activities; understand-

benefit from indigenous experience are more ing the interactions among soils, crops and

likely to secure investment and long-term water; scheduling irrigation; using objec-

support of engaged communities. tive monitoring tools where possible; and


remaining open to new ideas such as solar
Special care is needed to ensure introduc- pumps for renewable energy. Benchmark-
ing change is gender sensitive, to avoid ing also helps farmers improve performance,
disadvantaging women farmers and to avoid and, together with WUAs, can provide oppor-
compromising existing sustainable practices tunities for farmers to work together to share
or adversely affecting land and water tenure ideas, compare performance and transfer
arrangements. For example, planting tree knowledge. Understanding and applying best
crops could affect land tenure security in practices can help to ensure farmers become
specific tenure regimes, and improving irri- agricultural water stewards.
gation technology could favour male farmers

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 297
© FAO/Patrick Meinhardt Storage has many forms, in natural and built
infrastructure. Nature has always supplied
the bulk of water storage (GWP and IWMI,
2021). People have long relied on natural
storage in ponds, lakes, wetlands and rivers

Optimizing water storage (Ffigure 5.9). Groundwater is a significant


store of water exploited for irrigated agricul-
Water storage provides a buffer for manag-
ture, as is soil water storage, which farmers
ing climate uncertainty and variability, and
are encouraged to increase using conserva-
is essential in building resilience to climate
tion agriculture practices for rainfed crops.
change. It can help water managers cope
with changing societal priorities and water Although the number of large built-storage
demand patterns. It can balance supply and dams increased significantly over the
demand to mitigate shocks such as drought twentieth century (Ffigure 5.10), the Global
and floods. Storing water during the wet Reservoir and Dam Database recorded a
season enables farmers to grow crops and significant decrease in the number of large
provide water for livestock during the dry dams completed since the 1990s (Lehner et
season. For irrigation schemes, overnight al., 2011). Lower investment requirements
storage allows farmers to continuously take make smaller storage facilities more
water from a canal system and irrigate crops justifiable, especially for irrigation. Myriad
according to their water needs rather than a small storage reservoirs are serving small
fixed water schedule determined by scheme irrigation schemes, but data are sparse or
managers. Conjunctive use in irrigation non-existent. Ultimately, the scale of surface
using natural groundwater storage and built water resources development depends on the
surface water storage is another example of scale of water allocation for irrigation. Such
balancing water supply with variable daily decisions are usually taken at the basin level
and seasonal irrigation demand. or within a national IWRM plan.

FIGURE 5.9 The water storage continuum

SUBSURFACE SURFACE Access


Dam outlets,
Increasing capital, environmental and social costs

Reservoirs pumps,
off-take towers
small large
and management complexity

Direct, buckets,
Ponds and tanks pumps

Aquifers Boreholes, deep /


shallow wells, etc.
deep shallow

Soil moisture Planting crops

Natural wetlands (lakes, swamps, etc.) All of the above

Increasing storage reliability Increasing storage reliability

Source: McCartney, M. & Smakhtin, V. 2010. Water storage in an era of climate change: Addressing
the challenge of increasing rainfall variability. Blue Paper. International Water Management Institute.
www.iwmi.cgiar.org/Publications/Blue_Papers/PDF/Blue_Paper_2010-final.pdf

298 5. RESPONSES AND ACTION AREAS


FIGURE 5.10 Number of large built-storage dams, 1900–2010

200
Number of completed large dams

150

100

50

0
1900 1910 1920 1930 1940 1950 1960 1970 1980 1990 2000 2010

Source: FAO. 2022. Geo-referenced database on dams. In: AQUASTAT - FAO's Global Information System on Water and Agriculture. https://www.fao.org/
aquastat/en/databases/dams

Just how effective storage can be depends on storage. A review of water storage (GWP
catchment characteristics such as vegetation, and IWMI, 2021) suggests there is already
soils, rainfall runoff response, and land cover a gap between available storage and the
and use. Changes in catchment parameters amount needed, and it is widening. There
will affect the amount and quality of storage. are variations, and countries have different
Investment in watershed management may priorities, but the storage gap threatens
benefit afforestation, reforestation, soil sustainable development for many. The
conservation, soil moisture retention and economic cost of an increasing storage gap is
groundwater recharge. However, it may significant. Benefits from agricultural water
reduce the capacity to harvest runoff in storage come from extending the area under
reservoirs, affect the hydrological flow regime irrigation and increasing the reliability of
in streams and rivers, and reduce surface supply to farmers, consequently reducing
storage reliability. But soil-conservation rural poverty and hunger, and promoting
measures may reduce suspended sediment growth. More storage and storage types are
and increase dry-season river flows that urgently needed, and existing storage needs
benefit storage and irrigation farming managing better.
(McCartney et al., 2019).
The availability of buffering capacity in natural
Despite the decline in large built-storage and built storage systems can significantly
facilities in recent years, the global need reduce drought impacts. Most developing
for more water storage is growing as water countries suffer from “difficult” hydrology,1113
demand increases across all sectors. However,
even the available built storage decreases
13
Europe, in contrast, has mostly “easy hydrology”,
due to sedimentation resulting from soil which lacks the extremes seen in developing countries.
erosion and the effects of environmental This is much simpler to deal with technologically and
institutionally, and the countries involved are usually
degradation and climate change on natural wealthy enough to invest in well-designed and robust
water infrastructure and strong institutions.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 299
© FAO/Patrick Meinhardt
Protecting groundwater
There has been consistent growth in ground-
water use for irrigation, livestock water-
which produces extreme drought and flood ing and agricultural processing, despite the
events that are difficult and costly to control, widespread problems of aquifer depletion
and funding is limited to mitigating the and pollution associated with agricultural
impacts. The strong correlation between land management for crop production and
drought events and low GDP aptly demon-
livestock grazing. It has increased in absolute
strates the need for more storage in Ethiopia
terms and as a percentage of total irrigation
and the United Republic of Tanzania (Fb
igure
(section 1.7.4). Limiting groundwater exploi-
5.11), to decouple climate and water security.
tation to maintain a wide range of water

Investment in storage tends not to be high- supply and environmental services is well
lighted in infrastructure studies. Instead, recognized as being essential (Foster and
storage is sector driven and is an integral Loucks, 2006). However, efforts to manage
part of water supply, irrigation and flood demand across the large continental aquifers
control. However, this can inhibit investment in the United States of America are still being
in multipurpose storage that could be effec- evaluated (Haacker, Kendall and Hyndman,
tive in meeting several sector objectives. 2016; Lubell, Blomquist and Beutler, 2020).
Adopting irrigation technologies, designed to
The review of water storage recommends a
make better use of available water resources,
new agenda that changes current silo think-
has not proven effective in reducing overall
ing about water storage to one that addresses
demand (Batchelor et al., 2014).
all the many different kinds of storage, natu-
ral and built, in an integrated system that
Groundwater use in irrigation has been
provides multiple benefits (GWP and IWMI,
increasing in absolute terms and as a percent-
2021). This includes assessing the socioeco-
age of total irrigation, despite the problems
nomic costs and benefits of integrated storage
systems, developing innovative approaches of depletion and pollution. Thus, limiting
to water storage, and optimizing integrated groundwater exploitation to sustainable
storage planning and operations. levels is desirable and essential.

FIGURE 5.11 Relationship between rainfall variability and GROSS DOMESTIC PRODUCT, 1990–2016

GDP growth (%) Rainfall variability (%)

United Republic of Tanzania Ethiopia

8 8
Rainfall variability (%)

Rainfall variability (%)

20 20
GDP growth (%)

GDP growth (%)

10 6 10 6

0 0
4 4
-10 -10
2 2
-20 -20
20 0

20 0
20 6

20 6
20 8

20 8
04

04
20 2

20 2
19 0

19 0
19 6

19 6
20 8

20 8
20 0

20 0
16

16
94

94
19 2

19 2
20 4

14
20 2

20 2
0

0
0

0
0

0
0

0
9

9
1

1
1
9

9
9

9
9

9
1

20
20

20
19

19
19

19

Year Year

Source: Global Water Partnership & International Water Management Institute. 2021. Storing water: A new integrated approach for resilient development, W.
Yu, W. Rex, M. McCartney, S. Uhlenbrook, R. Von Gnechten & J.D. Priscoli, eds. Stockholm, Global Water Partnership.
www.gwp.org/globalassets/global/toolbox/publications/perspective-papers/perspectives-paper-on-water-storage.pdf

300 5. RESPONSES AND ACTION AREAS


Groundwater pumping technologies continue is associated with land and water pollution
to improve, along with borehole technolo- due to inadequate management of livestock
gies adopted from petroleum and mining waste, especially from intensive feedlots.
industries, giving more range to recover
groundwater from energy-efficient vari- Attenuating contamination
able drive pumps to solar-powered pumps by nutrients and fertilizers
(FAO and IWMI, 2018). The demand pull from
Various nutrients and organic amendments
irrigated agriculture and livestock water-
to enhance soil fertility, stability and func-
ing is thus expected to increase but with
tion all require proper management to avoid
more accurate targeting of shallow and deep
or mitigate soil contamination and associ-
aquifers. Attempts to moderate groundwater
ated processes. These include surface water
withdrawals through energy pricing is one
and groundwater pollution, nitrous oxide
approach that has met with a degree of success
and methane emissions, eutrophication and
in various states in India (Shah, 2009).
acidification.

Land management can play a crucial role


Responsible use and management of
in maintaining patterns of aquifer recharge.
fertilizers is vital for sustainable intensive
Large-scale managed aquifer recharge must
agriculture. The FAO Fertilizer Code guides
be part of a landscape approach to improving
their use (FAO, 2019c) (Bbox 5.19). Chemical
land and water quality (Dillon et al., 2020).
use must be considered at the landscape,
Maintaining healthy soils, free of contami-
regional and global levels. This requires a
nation, should be an essential starting point,
holistic approach to using nutrients and their
bearing in mind that conserving natural
cycles in soils, plants, animals, humans,
biodiversity in protected areas will be as
water and the environment.
important as adopting conservation agricul-
ture techniques on cultivated land. Nitrogen and phosphorus compounds are
essential for crop growth. However, nitro-
5.5.4 Managing gen fertilizers are highly water soluble and

environmental risks rapidly cycled in the soil. Unused nitrogen can


find its way into drainage systems, water-
The agricultural sector is responsible for courses and groundwater. Reducing nitrogen
managing environmental risks by reducing losses requires better fertilizer application
pollution and the harmful effects of fertiliz- practices, improved water management, and
ers, pesticides, herbicides and livestock waste, maintenance of healthy plants with good
by minimizing antibiotic use and avoiding nitrogen uptake and healthy soil that holds
secondary health issues such as microbial and transforms the nitrogen.
resistance, and by reducing GHG emissions.
Slow-release nitrogen compounds can help
There are also risks to health from chemical reduce the risk of leaching, and biological
pollution and water-related diseases, such additives can enhance nitrogen-use effi-
as diarrhoea, and water-borne diseases, ciency by inducing more robust root growth
such as bilharzia and malaria, often called and more active uptake. But farmers need
the diseases of irrigation because of their training to encourage their use, regulation
prevalence in stagnant water in poorly and incentives.
maintained schemes. The livestock sector

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 301
Box 5.19
The Fertilizer Code

The Fertilizer Code was developed in response to the request of the Committee on Agriculture to
increase food safety and the safe use of fertilizers. It is also a response to the declaration of the third
United Nations Environment Assembly on soil pollution, which aims to ensure broader support for
implementing the Voluntary guidelines for sustainable soil management (FAO and ITPS, 2017).

The Fertilizer Code provides a locally adaptable framework to avoid misuse, overuse and underuse
of fertilizers and a set of voluntary practices for stakeholders involved with fertilizers. Adhering to the
principles of the Fertilizer Code contributes to sustainable agriculture and food security from a nutrient
management perspective. It aims to assist countries in addressing the multiple and complex issues
related to responsible use and management of fertilizers at farm, ecosystem and national levels. The
Fertilizer Code helps stakeholders establish systems for monitoring production, distribution (including
sale), quality, management and use of fertilizers to achieve sustainable agriculture and SDGs by
promoting integrated, efficient and effective use of quality fertilizers.
Source: FAO. 2019. The international code of conduct for the sustainable use and management of fertilizers.
Rome. www.fao.org/3/ca5253en/ca5253en.pdf

Minimum-tillage or no-tillage agriculture costs and respond to environmental


and other practices that restore SOM help awareness. In developing countries, uptake
maintain healthy soils and reduce nitrate and of integrated pest management is much
phosphate pollution in linked water bodies. slower, although FFSs effectively improve
Unlike nitrogen, phosphorous is generally farmer knowledge (Settle and Garba, 2011).
bound to soil particles and is released slowly The slow progress in establishing regulatory
to plants. It is therefore less likely to find its and legislative frameworks for approval and
way into groundwater or drainage systems. safe use of pesticides is a cause for concern.
However, runoff from farms represents a This is especially true for cheap generic
significant risk of phosphorus entering brands of harmful pesticides, which some
rivers, lakes and coastal systems. countries still produce and use, although they
are banned in international markets.
Reducing pesticides and
other contaminants Some agricultural practices release other
contaminants, such as trace elements, micro-
A range of integrated pest manage-
plastics, antibiotics, antimicrobial resistant
ment methods are available to help reduce
bacteria and pathogens, into soils. Examples
chemical pesticide use, and associated soil
include irrigating with untreated wastewater,
and water pollution and health risks. FAO
applying fresh manure from animals treated
supports countries to apply the International
with high-dose antibiotics (Zhang et al.,
Code of Conduct on Pesticide Management
2016) or fed with food rich in trace elements,
for distributing pesticides and their use, and
using sewage sludge as organic fertilizer and
for ensuring the safe storage of pesticides
abandoning agricultural plastics in the field
and safe disposal of obsolete pesticides.
(Nizzetto, Futter and Langaas, 2016).

Large-scale farms in developed industrial


Simple alternatives are available to avoid
countries have adopted integrated pest
soil pollution at the field level and to avoid
management methods to reduce production
contaminants entering the food chain.

302 5. RESPONSES AND ACTION AREAS


On-farm waste-management plans include
complete removal of packaging waste
and other plastics from the soil, proper
management of animal faeces and urine, and
establishing controlled and impermeable

© FAO/Giuseppe Bizzarri
collection areas to prevent leakage. Beyond
the farm gate, vegetative and physical
barriers and drainage improvements in
agricultural areas close to potential sources
of contaminants, such as heavy-traffic
roads, mines and industries, offer low-cost
solutions to prevent pollutants from reaching Risks of soil salinization have long been a
agricultural soils (Kibblewhite, 2018). problem in irrigation, particularly in arid and
semi-arid areas, where salts build up in the
Wastewater use, particularly from densely
surface soil through evaporation and waste-
populated and industrial areas, requires at
water is reused for irrigation (Sjoerd et al.,
least secondary treatment. Selecting cultivars
2017). The traditional solution for remov-
with lower contaminant uptake capacity or
ing salts in soils with shallow groundwater
cultivating industrial and bioenergy crops
is to leach excess water through the soil
on farms are other options. Treating soils
profile into underground tile drains and open
with inorganic soil amendments, such as
ditches. Plastic soil mulching is also used for
lime and iron oxides, and improving SOC
improving water and salt balances, but this
content can help to immobilize contaminants.
may have environmental impacts. Managing
Soil biodiversity has an important role in
soil salinity involves reducing evaporation
the bioremediation of contaminated soils
from the soil surface through controlling
as certain bacteria and fungi can degrade
water applications to meet crop demand and
and immobilize specific environmental
providing a leaching fraction to maintain an
contaminants such as aromatic hydrocarbons
acceptable salt balance in the soil.
(FAO et al., 2020).

One option is to accept saline drainage water


Managing soil and adopt biosaline agriculture by selecting
and groundwater salinity salt-tolerant crops and appropriate crop-
Over 1 100 million ha of soils are affected by ping patterns and management practices.
salinity and sodicity, of which 60 percent are If planned at the watershed or landscape
saline, 26 percent are sodic and the remain- level, this adaptive approach can reduce
ing 14 percent are saline–sodic. The regions environmental degradation and contrib-
most affected are arid or semi-arid zones in ute to ecosystem restoration in dry lands. A
Australia, Central Asia, Near East and North- handbook for saline soil management (FAO,
ern Africa. Estimates of irrigated salt-affected 2018d) provides innovative methods and
soils vary widely between 20 percent and 50 technologies for ameliorating salt-affected
percent of the irrigated area (FAO, 2022a). soils, including a proximal technique of
Thus, GSP prioritized soil salinity mapping electro-melioration, precision agriculture,
to identify the scale of the problem in each diversification of salt-resistant crops and the
region and the required investment in reme- use of halophytes.
dial measures (Chapter 4).

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 303
must speed up, particularly in developing
countries where programmes can take up to
ten years or more to develop new generations
of seeds, plus many more years to disseminate
them (Atlin, Cairns and Das, 2017).

Improving crop nutritional value is impor-


© FAO/Fahad Kaizer.

tant given that over 2 billion people world-


wide are food insecure and 688 million are
undernourished (Ziadat, Bunning and De
Pauw, 2017; FAO, 2020b). Future breeding
programmes must also focus beyond tradi-
In Central Asia, 40–60 percent of irrigated
tional staple crops, such as maize, wheat,
land is salt affected or waterlogged.
rice and soybeans, to include neglected crops
Countries in Central Asia and Turkey have
been supported to develop integrated natural vital for nutrition and resilience, such as

resources management in drought-prone cassava, millet, peas and sorghum. There are

and salt-affected agricultural production myriad traditional varieties of crops and live-
landscapes. Activities include soil mapping, stock breeds, and also crop wild relatives that
applying innovative approaches and may show beneficial nutritional and climate
biotechnologies to restore soil fertility, and resilience traits that require identification,
incentives for adoption. This is supported by safeguarding, improvement and prioritiza-
the Central Asian Countries Initiative for Land tion. Policies that are biodiversity friendly
Management, FAO, GEF, the International and participatory breeding efforts are needed
Center for Biosaline Agriculture and country that recognize the enormous contribution of
partners (FAO, 2018e). How countries use indigenous people and smallholder farmers
their salt-affected soils will be important for as custodians of the world’s food crops and
their future food security, but will require domesticated animals and farmers’ rights.
strong political support and funding.
Future efforts can take advantage of

5.5.5 Going beyond biotechnologies that can reduce the plant

the farm breeding cycle from ten to two years, such


as marker-assisted and genomics-assisted
Research is vital breeding that uses molecular biology tools
and information technology to identify
New crop varieties will need to adapt to a
promising crop traits (Varshney et al., 2012).
wide range of rapidly changing climate
and socioeconomic conditions. They must
cope with increasing frequency of damag-
ing high-temperature events, new pest and
disease pressures, increasing weed competi-
tion, more frequent and prolonged extreme
© FAO/Roberto Faidutti

weather events, increasing water scarcity and


quality deterioration, and decreasing use of
agrochemicals. Productivity must also increase
faster than historical trends and include
improved nutritional values. Breeding cycles

304 5. RESPONSES AND ACTION AREAS


Genetically modified crops continue to be Biofortification is a promising, cost-effective
the subject of a long-running debate. One and sustainable technology that can improve
unfortunate consequence of this is that other nutritional quality through agronomic prac-
successful biotechnologies have been over- tices, conventional plant breeding or modern

shadowed (FAO, 2017f). New Rice for Africa biotechnology (Garg et al., 2018). It offers a
way to reach the rural poor who rarely have
varieties, which are now widely distributed
access to commercially fortified foods. More
in sub-Saharan Africa, have been devel-
than 20 million smallholders in develop-
oped using biotechnologies that combine
ing countries grow and consume biofortified
high-yielding Asian rice with the robustness
crops (Venkatesh and Hurrell, 2018) such as
of African rice. Adopting “biotech crops” is
vitamin A in sweet potatoes, zinc in rice and
the most pronounced crop technology trend iron in beans.
(James, 2014). Genetically modified crops
are likely to become much more widely used A comprehensive inventory is available of
when tested in new forms to counterbal- near-ready and future technologies that
ance the various associated risks. But new can increase food production while reduc-
techno-bio-socio-cultural solutions will be ing pressure on land and water resources
(Herrero et al., 2020). Among these, biorefin-
required, and which also need to be accepted
eries already exist to manufacture meat and
by people (IIASA, 2019). Gene-editing
vegetable substitutes, but overcoming public
technologies can avoid the drawbacks of
perceptions of quality and health risks will be
traditional genetic engineering technolo-
challenging (IIASA, 2019).
gies. They could transform conventional
agriculture, create new laboratory farming Urban farming is emerging as a means of
practices and help find new ways to leverage enhancing food security within cities (Bb
ox
complementary agroecological approaches 5.20). Vertical farms grow produce inside or
(Zhang et al., 2018). on top of buildings, and hydroponic agri-
culture grows plants without soil with plant
roots in a water solution of mineral nutri-

Box 5.20
Urban farming: a solution to enhance food security in cities

Urban farming is a form of natural capital for growing food and other crops within cities. It offers the
potential to ameliorate urban environmental problems by increasing vegetation cover and contributing
to a decrease in the urban heat island intensity, improving the liveability of cities and providing enhanced
food security. A global assessment of urban farming ecosystem services indicates a potential annual
food production of 100–180 million tonnes, energy savings of 14–15 billion kWh, nitrogen sequestration
of between 100 000 and 170 000 tonnes, and avoided stormwater runoff of between 45 and 57 billion
m3 annually. The value of the ecosystem services provided by urban farming could be worth as much
as USD 80–160 billion annually.

High- and low-technology solutions exist for urban farming. One company introduced a low-cost
aquaponics system combining fish farming with vegetable cultivation in closed-loop water circulation
to smallholder farmers in tropical areas, where it increases food and nutrition security in the dry season
and contributes to generating additional income. Another company in Dubai, United Arab Emirates, uses
hydroponics technology to grow vegetables for top restaurants and caterers.
Source: Clinton, N., Stuhlmacher, M., Miles, A., Uludere Aragon, N., Wagner, M., Georgescu, M., Herwig, C. & Gong, P. 2018. A global geospatial
ecosystem services estimate of urban agriculture. Earth’s Future, 6(1): 40–60.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 305
Using information
and communications
technology and big data
© FAO/Riccardo De Luca

Opportunities are emerging from advances


in ICT. The application of ICT to agriculture
can also help improve productivity, manage
associated environmental risks, and ensure
sustainable land and water management.

Recent advances in ICT, big data science,


ents. Aquaponic farms leverage the symbiosis
Earth observation systems, open access,
between hydroponic agriculture and aqua-
artificial intelligence, machine learning
culture: plants absorb fish excretions as
and cloud computing platforms, along with
nutrients and clean water returns to the fish
smartphone-enabled citizen science, have
basins. Such systems operate in controlled
increasingly made big data analytics much
environments, enabling faster crop cycles
smarter and more useful for agricultural
and more crop rotations each year. They use
planning and management. They have
70–90 percent less fertilizer and water by
also created baseline information for
capturing and condensing evapotranspira-
tion and recirculating them within the system better-informed decision-making and

(Crawford, 2018). However, not all crops can opened up opportunities to fill knowledge
be grown in a controlled environment. They gaps at multiple levels (e.g. data, yield,
are currently limited mainly to vegetables ecology, economy and resilience) and scales
and herbs, and there are challenges in scaling (e.g. space, time and package) to target
up these solutions (Foley, 2018). demand-driven interventions for sustainable
land and water management.
Several constraints still impede the uptake
of near-future technologies. For example, Bb
ox 5.21 illustrates the potential for big data
inadequate market infrastructure has limited to benefit smallholder rice growers, enabling
fertilizer adoption by African smallholders them to increase their cropping intensity. This
(FAO, 2011a). Uptake requires investment in is an example of a multicriteria assessment
research and establishing regulatory frame- of farming systems and resources that allows
works to ensure that innovations meet accept- upscaling from farm to national and regional
able human health, social and environmental levels (Löw et al., 2017; Biradar et al., 2020).
standards, that commercial interests do not
monopolize technologies, and that there is Critical questions require soil data and infor-
increasing awareness of the potential benefits mation at the global scale to understand
as well as risks (Searchinger et al., 2019). Also, Earth processes and to provide the context
tenure security and farmers’ rights must be for national to local decision-making. To
recognized and applied, to reduce inequity achieve this, GSP and the International
in access to natural resources. It is essential Network of Soil Information Institutions are
to acknowledge the vital role of smallholder developing GLOSIS – a federation of soil
farmers and indigenous people in conserv- information systems that shares soil datasets
ing, using, exchanging and improving genetic via web services. This aims to empower coun-
resources for food and agriculture, the impor- tries to build their national soil information
tance of safeguarding indigenous knowledge, systems as reference centres. Its architecture
and the importance of their participation in allows holders of soil data to engage at differ-
decision-making and benefit sharing. ent levels, according to technical skills and

306 5. RESPONSES AND ACTION AREAS


Box 5.21
Intensifying production using rice fallows to grow pulses and vegetables

As the area of arable land is not expected to increase significantly, agricultural fallow areas offer
opportunities for growing additional food and nutrition provided their production potential can be
unlocked (Biradar et al., 2019).
An example is the potential use of rice fallows. A digital platform was developed to provide
near-real-time information that identifies “hotspots” of suitable areas for specific crops, lengths of
crop fallows, soil moisture and water harvesting potential for supplementary irrigation. Among other
opportunities, this system was used to identify rice fallows suitable for growing food legumes in the
Eastern Gangetic Plains.
Fine spatial resolution data from the Copernicus Sentinel series of satellites have enabled rice fallow
areas on smallholdings of less than 2 ha to be mapped and assessed as suitable for growing pulses
using conservation agriculture to increase farm income and supporting marginalized farmers. With a
temporal resolution of 10 m and a frequency of 3–5 days between mapping, this system enabled small
parcels of land to be monitored for sustainable agricultural practices, specifically pulse intensification
in rice fallow areas.
Sources: Biradar, C., Sarker, A., Krishna, G., Kumar, S. & Wery, J. 2020. Assessing farming systems and resources for sustainable pulses intensifica-
tion. Conference presentation at Pulses the Climate Smart Crops: Challenges and Opportunities (ICPulse2020); International Center for Agricultural
Research in the Dry Areas. 2022. Agricultural intensification and crop diversification.
http://geoagro.icarda.org/intensification

ambitions, to set up and maintain national The State of food and agriculture 2019 report
soil information systems. A series of thematic (FAO, 2019d) distinguishes between food
soil assessments feed into GLOSIS to improve “loss”, which occurs post-harvest, but not
understanding for informed responses. including the retail level, and food “waste”,
which refers to the decrease in the quantity or
Agencies involved in land and water planning
quality of food resulting from decisions and
and management need to update the capac-
actions by retailers, food service providers and
ity and tools for managing GISs, develop and
consumers. This aligns with the distinction
use maps and plans, and monitor trends and
implicit in SDG target 12.3.
impacts. This is often a critical capacity and
investment gap that takes time to fill and can Food loss and waste represents an inefficient
limit progress on planning. use of valuable agricultural resources, and
causes avoidable environmental degradation
Reducing food loss and waste (HLPE, 2014). Globally, FLW accounts for 24

Food loss and waste is a function of marketing percent of total freshwater used in food crop

and distribution that ultimately influences land production, 23 percent of cropland area and

use. Reducing FLW is one measure to improve 23 percent of fertilizer use (Kummu et al.,

food security, lower production costs, reduce 2012). Halving FLW would provide enough

pressures on natural resources and improve food for approximately 1 billion people. Alter-

environmental sustainability. The SDG target natively, resources used to grow FLW could

12.3 calls for halving per capita global food be redirected to higher-value use or support

waste at the retail and consumer levels and more environmentally sustainable agricultural

reducing food losses along production and production and consumption.

supply chains by 2030 (United Nations, 2015).

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 307
© FAO/Miguel Schincariol
Promoting sustainable diets
and consumer options
Rapidly rising incomes and urbanization are
driving a global change in lifestyle and food
Measures for reducing FLW for different consumption patterns, in which traditional
production stages vary along the food supply diets are being replaced by diets higher in
chain (Searchinger et al., 2019). They need animal-based foods, refined sugar and fat
adapting to local conditions and targeting (FAO et al., 2018).
towards critical loss points to cope with
Dietary shifts have traditionally sought to
the various barriers. They vary by region,
promote health and well-being but are now
food supply chain stage and supply chain
linked to reducing the environmental impacts
actors, and include poor institutional regu-
of food production (Springmann et al., 2018;
lations, limited financial sources, resources
IPCC, 2019). Dietary patterns with low envi-
constraints, information gaps and consumer
ronmental impacts can also be consistent
behaviour (Shafiee-Jood and Cai, 2016). Bb
ox
with good health (Gonzalez Fischer and
5.22 illustrates measures to substantially
Garnett, 2016). However, researchers have
reduce food losses in two example countries.
not yet calculated the adjusted land and
Reducing FLW will require broadly shared water resource requirements to service the
commitments to quantitative goals, careful change in crop production to substitute for
measurement and persistent action. In terms animal protein.
of policies and infrastructure investments,
public interventions may create an enabling With rising urbanization, interest in
environment that allows private actors to peri-urban and urban farming to meet
invest in reducing FLW (FAO, 2019d) (see also the increasing demand for local, fresh and
the section on harnessing circular economies relatively unprocessed food is growing.
for natural resources). Organizing short supply chains between local

Box 5.22
Reducing FOOD LOSS AND WASTE in Senegal and the United Kingdom

In Senegal in the early 1990s, hand threshing led to losses of 35 percent of harvested rice. Researchers
worked with farmers to modify a mechanized threshing tool for local conditions that harvested 6 tonnes
of rice per day and captured 99 percent of grains. Despite a cost of USD 5 000, the benefits were
sufficiently high that the technology is used to harvest about half of rice production in Senegal (Diagnea,
Demonta and Diagneb, 2009).
The United Kingdom achieved a 21 percent reduction in household food waste between 2007 and
2012, mainly through various labelling and public relations efforts. Supermarket chains printed tips for
improving food storage and lengthening shelf life for fruits and vegetables directly onto the plastic
produce bags in which customers place their purchases. Some chains shifted away from “buy one
get one free” promotions for perishable goods towards using price promotions. The government
revised guidance on food date labels, suggesting retailers remove “sell by” dates as many consumers
mistakenly interpreted this as meaning food was unfit to eat after that date. Instead, they displayed “use
by” dates, which more clearly communicate when food is no longer fit for consumption. Also, many
food manufacturers, food retailers and local government authorities participated in the Love Food Hate
Waste campaign, which raised public awareness and provided practical waste reduction tips through
in-store displays, pamphlets and the media (Searchinger et al., 2019).

308 5. RESPONSES AND ACTION AREAS


farms and retailers or consumers in nearby
cities reduces food transport.

New digital marketplace platforms connect

© Pixabay
farmers to food purchasers to provide food
traceability, greater price transparency
and faster, round-the-clock access to
information. Nanotechnology has proven are consistent with environmental sustain-
capabilities that are valuable in packaging ability. This means “closing the loop” of
food, including improved mechanical, resource use to decouple economic activity
thermal and biodegradable barriers. from consuming finite and limited resources
Intelligent food packaging technologies (Ellen MacArthur Foundation, 2015). It
(e.g. microchipping) that contain sourcing, includes searching for resource-efficient
safety and traceability information on food agricultural practices, encouraging regen-
production, processing and environmental erative agriculture, prioritizing renewable
footprint are becoming available (Herrero et energy, preventing resource leakages (e.g.
al., 2020). carbon, nitrogen, phosphorus and water),
and stimulating reuse and recycling resource
Harnessing circular economies losses in a way that adds the highest value to
for natural resources the food system (Jurgilevich et al., 2016).
Current food production and consumption
patterns are primarily built around a linear The European Commission has already

economic model involving extracting natural developed a circular economy action plan
resources to make products, using them for that includes specific measures for the food
a limited period and discarding them into system (Bb
ox 5.23).
landfill as waste. This is an inefficient way
Innovative agricultural practices are improv-
of using natural resources; in 2011, it had
an estimated annual cost to the global food ing resource-use efficiency on farms. Preci-

system of USD 1 trillion (FAO, 2011b). sion agriculture combines geomorphol-


ogy, satellite imagery, global positioning
The global food system already generates and smart sensors to provide farmers with
significant environmental impacts decision-support systems based in real
(Springmann et al., 2018). It is vulnerable time for whole-farm management (Lowen-
to environmental changes (e.g. severe berg-DeBoer and Erickson, 2019). Global
droughts), floods and diseases, and land positioning system enabled autonomous
degradation caused, in part, by climate farm machinery can operate continuously,
change. The idea of a circular economy is reduce labour inputs, and minimize planting
receiving increasing attention worldwide and harvesting costs. Smart sensors using
to promote sustainable consumption and drone technology can measure soil and plant
production patterns, including food systems characteristics, thus enabling efficient use of
(Ghisellini, Cialani and Ulgiati, 2016). fertilizers, pesticides and water. By combining
precision agriculture with no-tillage farming,
A circular economy brings together all the farmers reported a 10–20 percent reduction in
issues of waste and inefficiency. It encour- fertilizer and pesticide use and as much as a
ages businesses and households to change 75 percent reduction in machinery and input
practices so production and consumption costs (Ellen MacArthur Foundation, 2015).

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 309
Box 5.23
European Commission’s circular economy action plan and FOOD LOSS AND
WASTE

In 2015, the European Commission adopted an ambitious circular economy action plan that
included measures to help stimulate Europe’s transition towards a circular economy, boost global
competitiveness, foster sustainable economic growth and generate new jobs. The proposed actions
aimed to “close the loop” of product life cycles through more recycling and reuse, thus bringing benefits
for the environment and the economy.
Food waste prevention was identified as a priority. The European Union Platform on Food Losses and
Food Waste, established in 2016, brings together all key actors representing public and private interests
from farm to fork to advance European Union progress towards SDG target 12.3. Members include
international organizations (FAO, OECD and UNEP), European Union institutions, experts from European
Union member states and stakeholders from the food supply chain, including food banks and other NGOs.
The platform aims to support all actors in defining measures to prevent food waste, including possible
recommendations for action at the European Union level, sharing best practices and evaluating
progress. The European Commission has adopted European Union guidelines to facilitate food
donations and redirecting food no longer fit for human consumption into feed.
Measurement is critical to food waste prevention. Revised European Union waste legislation adopted
in 2018 has introduced specific measures, which provide the European Union with new and consistent
data on food waste levels. In 2019, the European Commission adopted a delegated act laying down
a common food waste measurement methodology to help member states quantify food waste at
each stage of the food supply chain and ensure coherent food waste monitoring at all levels across
the European Union.
Source: Adapted from FAO. 2019. The state of food and agriculture 2019. Moving forward on food loss and waste reduction. Rome. www.fao.org/3/
ca6030en/ca6030en.pdf

The potential to recover resources from waste brackish water, agricultural drainage, water
streams along the entire agrifood chain can containing toxic elements and sediments,
be significant. By-products from produc- and wastewater effluents. All are of poor
tion and consumption include crop residues, quality and unsuitable for most purposes, but
coproducts from industrial food processing, may be acceptable in some circumstances for
food waste, and animal and human excreta. agricultural use.

Closing resource loops requires new interac- Wastewater remains a largely untapped
tions among food system components, such resource; the treatment capacity for wastes
as between cities and rural food-producing generated by growing cities is inadequate
areas. Cities are sources of large amounts of in most countries. Most wastewater is
food waste and human excreta, which could discharged without treatment into the envi-
provide valuable nutrients for food produc- ronment. It either runs to waste, or is diluted
tion in farming systems that combine plant, in the region’s waterways and reused down-
insect and fish production. stream in some countries to irrigate millions
of hectares of cropland, often unintentionally
The benefits of a circular economy are just as posing serious risks to the health of farmers
applicable to agricultural water management and consumers and the environment. The
as to the broader land-use and food systems. SDG target 6.3 for water quality, wastewater
For water, this approach offers opportunities treatment and safe reuse requires halving
to use non-conventional waters that might the proportion of untreated wastewater and
otherwise go to waste, such as saline and

310 5. RESPONSES AND ACTION AREAS


Box 5.24
Wastewater: a potential water resource in the Central America
and Caribbean region

Estimates indicate that the Central America and Caribbean region generates some 30 km3 of
municipal wastewater annually, but has the capacity to treat only 40 percent. However, the proportion
actually treated is even lower because of inadequate maintenance. Pollution, including faecal matter,
is causing serious degradation in 25 percent of the region’s rivers. Only a marginal amount of treated
water is directly reused for agriculture in a planned, productive and safe manner. Concerted action
is needed to mitigate the health and environmental risks in the region’s peri-urban hotspots and to
capitalize on the opportunities that reuse brings. An analysis by FAO and the International Water
Management Institute to assess the potential for water reuse in agriculture in the region, based on
country experiences, demonstrated the opportunities to consider urban effluents as a resource and
set out the principles and the stringent management required to evaluate and mitigate the risks.
While the region has made substantial investments in wastewater treatment in recent years,
their effectiveness, efficiency and sustainability are far from guaranteed. Challenges include an
excessive emphasis on developing new infrastructure, poorly developed legislation, lack of policy
and regulatory mechanisms to allow gradual improvement, regulations that limit or forbid resource
recovery, technology selection criteria biased towards expensive technologies, lack of adequate
control of industrial discharges, and reliance on conventional financing.
Wastewater reuse can transform wastewater treatment plants from cost into profit centres. Using
marginal quality water for irrigation liberates better quality water for higher-value uses and creates
value beyond that due to its direct use. Creating such value is significant in dry areas that suffer from
chronic water scarcity. Wastewater management represents the largest market for clean technologies
in the region, with an estimated size of USD 160 billion in the decade to 2023.
Sources: Martin-Hurtado, R. & Nolasco, D. 2016. Managing wastewater as a resource in Latin America and the Caribbean: Towards a circular economy
approach. Washington, DC.
https://programme.worldwaterweek.org/Content/ProposalResources/allfile/managing_wastewater_as_a_resource_in_lac.pdf FAO. 2017. Reutilización
de aguas para agricultura en América Latina y el Caribe: Estado, principios y necesidades. Santiago.
www.fao.org/3/i7748s/i7748s.pdf

substantially increasing recycling and safe Wastewater reuse in agriculture can be attrac-
reuse. There is great interest in the safe reuse tive to farmers because the nitrogen and
of wastewater, and many countries are now phosphorus contents in sewage can reduce
working to improve data collection to under- the need for chemical fertilizers (Bb
ox 5.25).
stand how best to make use of it (Bb
ox 5.24). With increasing urbanization, larger volumes
of municipal wastewater become available
The International Water Management Insti- for peri-urban agriculture. However, waste-
tute and the Near East and North Africa water requires treatment appropriate to its
ReWater programme (whose partners use to avoid posing environmental or public
include FAO, the International Centre for health risks. Moreover, strict and enforceable
Advanced Mediterranean Agronomic Studies rules are essential when adapting cropping
and the International Centre for Agricultural patterns to effluent use based on safe waste-
Research in the Dry Areas) support capacity water reuse and controls on contaminants at
development on water reuse in agriculture, the water source.
addressing barriers to reuse and promoting
safe reuse practices that improve food safety,
health and livelihoods (IWMI, 2021).

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 311
Box 5.25
Valuing wastewater as a source of nutrients for agriculture

Interest in recovering nutrients from streams of wastewater is increasing as municipal wastewater


volumes increase and methods of nutrient recovery are developed. A global assessment suggests
that wastewater annually contains 16.6 million tonnes of nitrogen, 3.0 million tonnes of phosphorus
and 6.3 million tonnes of potassium. If fully recovered, this could offset some 13 percent of the global
demand for these nutrients in agriculture and generate revenues of USD 13.6 billion. An environmental
benefit from reducing the pollution of municipal effluents is reduced eutrophication in water bodies.
Source: Qadir, M., Drechsel, P., Jiménez Cisneros, B., Kim, Y., Pramanik, A., Mehta, P. & Olaniyan, O. 2020. Global and regional potential of wastewater as
a water, nutrient and energy source. Natural Resources Forum, 44(1): 40–51.

Water extracted from saline aquifers or funding from the International Bank for
captured from agricultural drainage offers Reconstruction and Development and the
options for irrigation when mixed/diluted International Development Association.
with freshwater for traditional crops like rice.
There is also potential to diversify cropping The OECD Development Assistance

into marine plants, such as seaweed, with the Committee classifies the control of soil

potential for gains in land and water. Such degradation, salinization, erosion and

practices can offset water scarcity in some desertification, as well as soil improvement,

areas, but there are risks of further soil salini- drainage, land surveys and land reclamation,

zation and degrading drainage water quality. under “agricultural land resources”. This
includes irrigation, reservoirs, hydraulic
Integrating wastewater reuse with other structures and groundwater exploitation.
options in the farming system can bring However, within the committed funds for
additional benefits. These include nutrient 2010–2018, less funding went to agricultural
recycling, regenerating soil health, and land and more to agricultural water.
reducing non-renewable energy and Specifically:
materials used in irrigation. This requires
a multisector approach to agricultural ƒƒ investments in irrigation infrastructure

ecosystems, as recognized by the water– ranked highest in terms of number of

food–energy nexus approach. projects and level of investment;

ƒƒ there was an increase in projects that

5.6 Action area IV: address climate change; and

Investing in long- ƒƒ the level of investment in ecosystems and

term sustainability land/landscape management was rela-


tively low but was gradually increasing.

5.6.1 Trends from The main scope of international investment


in agriculture sectors has included
2010 to 2018 agricultural development and governance,

Trends in investments in agricultural land irrigation and drainage improvement, water

and water resources in the period 2002– resources management, climate change and,

2010 relative to 2010–2018 broadly parallel to a lesser extent, land and soil resources

the growth in GDP in countries eligible for management. Many projects also seek to

312 5. RESPONSES AND ACTION AREAS


improve agribusiness, have an ecological or
environmental focus, or focus on poverty
alleviation and community development.
Conventional funding has aimed to maximize

© FAO/Ishara Kodikara
agricultural efficiency and find competitive
advantage, which has meant that in land-
and water-scarce areas in particular, food
self-sufficiency has been given a lower
priority than that of producing exports of
high-value crops.
ƒƒ low investment in agriculture is problem-
atic according to the 2018 Asian Develop-
Against this trend, land-based subsidies
ment Bank evaluation (ADB, 2018), but
to agriculture are generating undesirable
is even more so with the emergence of
externalities (FAO, UNDP and UNEP, 2021).
COVID-19 in 2020;
Removing distorting support measures and
decoupling subsidies and production to direct
ƒƒ there is scope for improving land and
subsidies toward public goods and services
water productivity in rainfed areas to
is a trend observed in developed economies
moderate the need for irrigation invest-
but less so in developing economies where
ment while contributing to smallholder
emissions from land are accelerating (Crippa
livelihoods;
et al., 2021). There is still time to “repurpose
agricultural support to drive a transforma- ƒƒ environmental benefits and natu-
tion towards healthier, more sustainable, ral resource protection are emphasized
equitable and efficient food systems” (FAO, to varying degrees by different IFIs in
UNDP and UNEP, 2021). their impact evaluations, particularly
in relation to distorting effects of direct
There are three broad categories of financing
payments to land productivity; and
instruments common to most international
financing institutions (IFIs): investment lend- ƒƒ the unequal distribution of benefits and
ing, results-based lending and policy-based costs of irrigation and drainage invest-
lending. While IFIs use a broad spectrum ments exacerbates inequities.
of financing instruments for public sector
projects, they primarily choose some form of Bb
ox 5.26 illustrates various approaches to
investment lending (debts, grants, loans, etc.) financing NbSs, and Tb
able 5.1 offers several
for agricultural land and water projects. case studies of investment in NbSs. Examples
include the Nairobi Water Fund investment
Five key points emerge from an assessment to improve the sustainability of small- and
of the performance of IFIs relative to FAO large-scale farming in the Upper Tana River
objectives: basin, an Ecosystem Service Marketplace
Consortium in the United States of America
ƒƒ there is a need to understand the inter-
and the Qiandao Water Fund addressing
dependence of urban and rural water
non-point-source pollution in Qiandao Lake,
requirements to achieve resilient water,
which is an important drinking water source
food and land security;

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 313
Box 5.26
Financing NATURE-BASED SOLUTIONS

Nature-based solutions are receiving increasing attention as an alternative solution to grey infrastructure.
International financing institutions have shown interest in funding NbSs as part of climate financing,
with varying degrees of success. The lack of a standardized methodology has slowed progress, but
IUCN has introduced a global NbS standard (IUCN, 2020). Together with the European Green Deal (EC,
2019), it can be a game changer in making NbS investments attractive for private institutions and IFIs.
The European Commission’s definition of NbSs builds on and supports other closely related concepts,
such as ecosystem approaches, ecosystem services, ecosystem-based adaptation/mitigation, and
rainfed and irrigated infrastructure. For agriculture, this is in line with FAO policy recommendations for
NbSs (FAO, 2018b), CFS (HLPE, 2015) and collaboration between FAO and IUCN (IUCN and FAO, 2020)
for developing agroecological practices such as NbSs.

As a think tank based in the United Kingdom, E3G provides an account of the IFI alignment of NbS
investments with climate financing (E3G, 2020). In its 2020 assessment, the picture looks promising
but needs attention. Out of the nine multilateral development banks, only the Asian Development
Bank has aligned NbS frameworks with the Paris Agreement. Six others (African Development Bank,
Asian Infrastructure Investment Bank, European Investment Bank, Inter-American Development Bank,
International Bank for Reconstruction and Development/International Development Association
and International Finance Corporation) have made partial progress by including some biodiversity
commitments and declaring the intention to scale up NbSs but without firm strategies. The three
remaining (Asian Infrastructure Investment Bank, European Bank for Reconstruction and Development,
and Islamic Development Bank) do not have explicit NbS policies.

The World Bank Group argues the financial sector has a crucial function in addressing the global
biodiversity crisis, and that governments and regulators must mobilize private finance at scale to
protect nature (World Bank Group, 2020). Its report presents the “Big Five”: five ideas for actions that
would help integrate biodiversity risk and opportunities into private sector decisions. These range from
environment fiscal reform and better data collection to broad support of the recently announced Task
Force on Nature-related Financial Disclosures.

in the Yangtze River delta in China (Hallstein ment financing institutions; bilateral donors;
and Iseman, 2021). international organizations; academia; and
civil society organizations. It is focused on
The Roundtable on Financing Water is a finding novel ideas and solutions. A brochure
global public–private platform established on financing a water secure future outlines
by OECD, the Netherlands, the World Bank OECD work in this area (OECD, 2021).
and the World Water Council. It draws upon
political leadership and technical expertise Farmer-led irrigation is a welcome initiative,
to facilitate and increase investments that pioneered by the World Bank (Izzi, Denison
contribute to water security and sustainable and Veldwisch, 2021). It aims to overcome the
growth. The round table engages a diver- inability of the financial system, government
sity of actors: governments and regulators schemes and market arrangements to enable
in developed, emerging and developing smallholder farmers to establish their own
economies; private financiers (e.g. institu- irrigation systems. Scaling out farmer-led
tional investors, commercial banks, asset irrigation will unleash the entrepreneur-
managers and impact investors); develop- ial power of a large number of farmers to

314 5. RESPONSES AND ACTION AREAS


Table 5.1 Selected NATURE-BASED SOLUTION investment case studies
Project/location Practices Scale Benefits Replicability

Nairobi Water Riparian One million ha A USD 10 million There are 41 water
Fund: Watershed management/buffer watershed that investment over ten funds in 13 countries,
management for zones; agroforestry supplies 95% of years would yield and over 80% of
healthy forests, adoption; terracing Nairobi’s drinking USD 21.5 million in cities globally can
agriculture, of hill slopes; water, food for millions economic benefits, meaningfully reduce
water quality and reforestation for of Kenyans and 65% including up to USD sediment or nutrient
hydropower degraded lands; of the hydropower 3 million/year in pollution through
grass strips in of Kenya increased yield for agriculture NbSs
farmlands; road farmers, over USD
erosion mitigation; 600 thousand/
soil conservation and year increase
water harvesting in hydropower
revenue and a 50%
reduction in sediment
concentration

Colombia Scattered trees in Developed in 87 Twenty percent These practices


Silvopasture: pasturelands; timber municipalities (12 increase in milk could be deployed in
Using silvopastoral plantations with states) in Colombia and/or beef cattle ranching across
practices to help livestock grazing; covering a total area production; improved Colombia with scaling
ranching and pastures between tree of 159 811 ha management on up to 1 million ha by
ecosystems alleys, windbreaks, live over 20 thousand 2030; they could also
fences and shrubs; ha and protection of reduce grazed area by
fodder banks almost 18 thousand 30% for conservation or
ha; reduction of 1.5 other purposes
million tonnes of GHG
emissions

Ecosystem Service Developing markets Market value of The goal is to launch


Marketplace to enable farmer quantified ecosystem a fully functioning
Consortium adoption of the benefits could be national-scale
NbS Ecosystem as high as USD 13.9 ecosystem services
Service Marketplace billion, by reducing market to sell carbon
Consortium, currently carbon emissions by and water quality and
conducting pilots 190 million tonnes, quantity credits for
in key agricultural nitrogen runoff by 700 agriculture by 2022
regions, including thousand tonnes and
the great plains, corn phosphorus runoff by
and soy belts, and 400 thousand tonnes
California fruit and nut
areas

Qiandao Water Cooperative Qiandao Lake Reduced loss Currently expanding


Fund: Innovation application of watershed is a of nitrogen and best management
plus tradition fertilizers and key drinking water phosphorus by practices to a broader
to engage pesticides; mulching source in the Yangtze 35–40%; increased scale in the watershed
smallholder and burying fertilizer; River delta and farmer income by and exploring other
farmers planting cover crops; for the Hangzhou 30–40% for green tea opportunities for a
planting nectar source metropolitan Water Fund model in
plants area; targeted China
subwatersheds
to deploy best
management
practices on 333 ha in
2020

Source: Adapted from Iseman, T. & Miralles-Wilhelm, F. 2021. Nature-based solutions in agriculture: The case and pathway for adoption. Rome, FAO and Virginia,
The Nature Conservancy. https://doi.org/10.4060/cb3141en

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 315
nance, integrated interventions at scale and
innovation in management and technology.

Governance is receiving increased attention,


© FAO/Giulio Napolitano

recognizing that an infrastructure focus


is insufficient to address poverty, equity
and sustainability. In view of increasing
food demands, decreasing land and
water availability, and environmental
pollution challenges, the next generation
improve their livelihoods, build resilience, of investments will need to focus on
create employment, increase access to food sustainably intensifying rainfed and
and support the food supply chain against irrigated agricultural production through
external shocks such as pandemics. improved data, technology, innovation,
management and governance, and integrated
5.6.2 Innovation in interventions at scale.

agricultural support Future investments are beginning to focus


to land and water on increasing resilience, reducing risk, and
enhancing connectivity and communication
Several innovations relate to financial instru-
through better mechanisms for collecting and
ments, emerging financing technologies
disseminating information, using modern
and governance of investments. Interna-
technologies for improved production and
tionally, there is a shift towards multigoal
inputs, and improving institutional capacity
frameworks such as the 2030 Agenda and
and governance. Improved connectivity is
observing effective practices. International
allowing smallholder farmers, for example, to
financing institutions have identified strate-
use mobile phones to enhance connections to
gies for more closely tracking performance in
other farmers and other actors with a range of
complex projects. Two specific instruments
benefits. Like the African Development Bank,
have gained momentum:
IFIs have noted that accelerating digitization
ƒƒ a multiphase programmatic approach, across the agricultural sector is a long-term
also called a “multitranche” approach, goal and will enable smallholders to access
which seeks to reduce the complex- market information and transactions. More
ity of implementing large, complex and immediately, it has been invaluable during
long-duration projects by splitting inter- the COVID-19 pandemic when face-to-face
ventions into multiple phases; and interactions were limited. However, even
today only 50 percent of the global population
ƒƒ a performance-based lending approach, has access to the internet.
which holds funds back until performance
criteria are met. There are many efforts to improve the data
value chain in the agricultural sector. This
Investments are needed to move from infra- includes collecting and analysing produc-
structure solutions and increasing produc- tion data on a large scale and then using
tion to sustaining productivity of rainfed and the processed data to improve agricultural
irrigated systems through improved gover- productivity and reduce land-related impacts.

316 5. RESPONSES AND ACTION AREAS


For example, older technologies such as and incentive mechanisms. Public invest-
remote sensors are being deployed along- ment can help to develop capacities across
side new technologies like drones to collect producer associations, regulators and applied
more data faster. Machine learning helps research. An effective land and water gover-
process big data in a relatively short time to nance framework that mobilizes responsible
support decision-making. Sensors connected investments and promotes the adoption of
to controllers as well as mobile phones and innovative management and technology
apps enable a two-way flow of information. in concert with sustainable land and water
This helps government agencies to better practices is a realizable goal. It requires
understand what is happening in the field understanding the trade-offs among sectors,
and also to improve their decision-making. the conflicts between land and water use for
agriculture, forests and urban needs, and the
To ensure reliable harvests and livestock urgent need to curb GHG emissions, through
productivity and to respond to climate avoiding deforestation and enhancing carbon
change and agricultural land expansion sequestration.
constraints, researchers are continually
finding new and innovative methods to Investment from the private sector needs to
improve agricultural yields and product complement investment from development
quality. Biotechnology continues to support banks and environmental funds. Govern-
yield increases and create crops that are ments can encourage consumers, NGOs and
more resilient. To complement advances in businesses to adopt responsible investments
irrigated systems, innovations in land, soil towards land and water management and
and water management are focusing more on sustainable food and agriculture systems.
rainfed agriculture and SLM methods. These
all deserve greater investment. Farmers and local communities are also
key investors when productivity gains help
Rethinking investment in agriculture is to sustain livelihoods and improve income
needed to support integrated land and water
resources management in rainfed and irri-
gated agriculture and to focus on policy
coherence. The high costs of degradation and
inaction highlight the urgency to increase
investments in sustainable land, soil and
water management and in restoring degraded
ecosystems, including viable land and water
management technologies, integrated land-
scape approaches in priority river basins and
ecosystems at risk. Emerging events follow-
ing the advent of COVID-19 in early 2020 also
need to be part of future investments, as they
have exposed vulnerabilities in global supply
chains that are still playing out.

Investment in integrated interventions at


scale shows great promise, and can be
© FAO

supported through innovative financing

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 317
levels. Incentivizing farmers to become inves-
tors in sustainable land and water manage-

© FAO/Khaled Desouki
ment can bring all-round environmental
benefits. However, they will need support
from innovative financing and instruments
that reconcile production and environmen-
tal management. Instruments that support
community-based land and water productiv-
ity improvements and adaptive management, Trust Fund is a successful mechanism for
capacity-building of producers’ associations, mobilizing investment for technical support
small-scale infrastructure and access to to the global and regional initiatives devel-
microcredit are all likely to be effective. oped and supported through GSP. Several
Green Climate Fund and GEF projects include
5.6.3 Prospects for land soils to some extent. But it was only in 2019
and water investment that the GEF Council endorsed its first project
concept focusing on soils – the Caribbean
Meeting global food demand will continue to small island developing States multicoun-
be a challenge, especially with growing social try soil management initiative for integrated
inequalities, conflicts, climate shocks and landscape restoration and climate-resilient
economic instabilities that affect food supply food systems – which is under development.
and distribution. Crises such as the COVID-19
pandemic will exacerbate this situation, as Promoting innovative technologies can
governments may reprioritize national funds accelerate achieving SDGs related to land
towards more immediate economic recovery. and water. These include genetic research
International financing institutions can help and trials, precision agriculture, biotechnol-
ensure funds continue to flow towards SDG ogy, soil carbon sequestration and renewable
efforts in food security, resource management energy systems in rural settings. Comple-
and rural livelihoods, so that investments mentary investment is needed in data
are realized in terms of sustained social, and information management to improve
economic and environmental benefits. connectivity among all producers, markets
and regulators. Early warning systems and
There is a continued need to engage farm- performance monitoring will also improve
ers as investors in sustainable land, soil and on-farm decision-making, while information
water management, rather than them being on adverse environmental and social impacts
passive beneficiaries, to help in enhancing will help guide responsible investment.
productivity and sustainability. Investments
in integrated landscape management show Future investments are expected to improve
great promise in the form of ecocompen- resilience, thus reducing risk and enhancing
sation (China), water funds (Africa and connectivity and communication through
Central, Northern and Southern America) and better mechanisms for collecting and
PES (globally). These are critical to ensure disseminating information, using modern
sustainability and to achieve environmen- technologies for improved production
tal benefits and natural resource protection. and efficient use of inputs and resources,
Soils have been seriously neglected in terms and improving institutional capacity and
of investments. However, the Healthy Soils governance.

318 5. RESPONSES AND ACTION AREAS


5.7 Conclusions

© FAO/Marco Longari
Over the past century, the world has largely
met the increasing demand for food, feed and
fibre by expanding the cultivated area and
intensifying the use of land, soil and water mented, but above all, better governed. With-
resources. However, increasing resource out doubt, governance is the most important
scarcity and inequalities of access are chang- element in successfully putting technolo-
ing the global dynamics around food, climate, gies into practice at territorial scale with
energy and allocation of financial resources to all stakeholders for achieving significant
solve social and environmental problems. The social, economic and environmental benefits.
ensuing economic tension has exposed the Technologies will have little impact without
reality of shared dependency. The COVID-19 strong and coherent governance at all levels
pandemic is another aggravating factor. of decision-making.
Ensuring that land, soil and water resources
are used in a sustainable manner requires Collaborative decision-making and learn-
careful balancing of competing goals such ing require deliberate linkages across
as economic growth, equity and a sustain- institutions, scales and sectors to capital-
able environment. These involve significant ize on stakeholders’ diverse knowledge,
trade-offs as well as opportunities. experiences and values to ensure negotiated
trade-offs are realistic, innovative and equi-
Land, soils and water feature in all five table. Actions will also need to be inclusive
Action Tracks prioritized through the United across physical and economic landscapes.
Nations Food Systems Summit process and
coalitions. Their instrumental role in access Current levels of financing remain substan-
to food, sustainable consumption patterns, tially inadequate to reach the international
nature-positive and resilient production, and community’s goal for life on land (SDG
advancing equitable livelihoods must not be 15) and sustainable management of water
underplayed. The four action areas in this (SDG 6). International funding and public
chapter complement the interlinked Food and private investments are encouraged
Systems Summit tracks and are an integral to improve the enabling environment and
part of the FAO strategic framework. These explore new approaches for investment in
signal the much-needed transformation of environmentally sustainable land, soil and
agricultural production from current perfor- water resources.
mance, which is stagnant or experiencing low
growth and still generating rising GHG emis- With well-adapted investments and actions

sions, to improved productivity, better nutri- by all stakeholders, unpredictable climate

tion and sustainable livelihoods in concert and socioeconomic shocks can be mitigated,

with positive environmental outcomes. and better food security, nutrition and
environmental health achieved as a result.
Land and water management can respond Taken together, the responses and actions
to meet the challenges of climate change outlined in this chapter can be expected to
based on “no regrets” investment in actions make positive contributions to the achieve-
that are better planned, informed and imple- ment of SDGs.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 319
© FAO/Marco Longari

Case study: Water accounting


and auditing in the West Bank

There has been significant economic and agricultural development in the Al Moqatta sub-basin
in northern West Bank in the last decade. This includes investment in greenhouses, drip irrigation,
improved cropping systems, irrigation of fodder crops with treated wastewater, and the development
of value chains that serve local and external markets. It is notable also that women farmers have
invested in and are managing greenhouses, for example, to grow strawberries. These activities have
benefited from FFSs and extension services supported by the Ministry of Agriculture and FAO, as part
of a project funded by the Swedish International Development Cooperation Agency.

Water scarcity constrains agricultural production and economic development in the Al Moqatta
sub-basin. Climate change and deteriorating water quality, linked to urbanization and agricultural
intensification, add further constraints. These prompted the Palestinian Water Authority to adopt
water accounting to assess and monitor trends in water supply, demand and consumptive use in
the sub-basin, and to identify and quantify significant return flows at different scales to manage and
improve water reuse.

The Palestinian Water Authority coordinates with the Ministry of Agriculture and other technical
institutions to use water accounting and auditing to identify potential trade-offs and unintended
consequences of investing in water management, such as constructing wastewater treatment plants
that reduce water access for some farmers. Water auditing goes one step further than water account-
ing by placing trends in water supply in the broader context of governance, institutions, public and
private expenditure, legislation and the Moqatta sub-basin’s political economy within the West Bank.

Source: FAO. 2022. Water accounting and auditing. In: Land & Water. Cited 7 March 2022.
www.fao.org/land-water/water/water-managment/water-accounting/en

320 5. RESPONSES AND ACTION AREAS


© Giulio Napolitano
Case study: Technology impacts on
traditional water rights systems in the
Near East and North Africa region

In the Near East and North Africa, the sustainability and equity of resources have long guided tradi-
tional water management systems, such as terracing, springs, aflaj/qanats and spate irrigation and
their complementary rules and administrative procedures. However, the increasing demand for water
for food production in the region has introduced new technologies, such as tube wells, spate irrigation
and permanent surface water diversion structures. But engineering solutions are often inappropriate,
participation is poor and the capacity to enforce new water regulations is weak. All these disturb
traditional water rights and threaten resource sustainability.

Springwater rights under threat from groundwater pumping


Irrigation using groundwater pumped from tube wells expanded in many countries across the region
during the 1960s and 1970s. However, legal and institutional frameworks set up to manage modern
groundwater abstraction have rarely successfully incorporated traditional springs and oasis water
rights and management systems. Many governments have been unable to confine agriculture within
sustainable water resources limits, as groundwater abstractions were driven by individual interests
that proved difficult for States to control. Clashes with traditional systems as springs ran dry have
forced those who lost water rights to follow the trend and invest in wells.

A new power structure and consequently a de facto water rights system have emerged, reducing
traditional spring and oasis water ownership and water rights in favour of open access to aquifers. This
has limited groundwater abstraction to those who can afford it.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 321
Most springs in the oases in Palmyra (Syrian Arab Republic), south Algeria, the Western Desert (Egypt),
Al Kufrah (Libya) and Al Ahsa (Saudi Arabia), and the natural springs in Bahrain, Tozeur and Kebili
(Tunisia) and Al Ahjar in Yemen, have all been lost or affected through excessive upstream pumping,
which has lowered groundwater levels

Source: United Nations Development Programme. 2013. Water governance in the Arab region: Managing scarcity and securing a future. New York.
www.undp.org/content/dam/rbas/doc/Energy and Environment/Arab_Water_Gov_Report/Arab_Water_Gov_Report_Full_Final_Nov_27.pdf

Modernizing traditional spate irrigation and social equity


In many countries in the region, spate irrigation has long been a vital method of exploiting flood flows
in seasonal rivers and diverting water into fields for cropping. In Yemen, written records dating back
some 600 years describe the complex arrangements for allocating water along the rivers. In Egypt,
spate water from 26 wadis in the northwest coastal region has been used for irrigation since Roman
times, and farmers in central Tunisia have practised this technique since the late 1800s. Other, more
permanent water diversion structures have been introduced to modernize these traditional systems.
In many cases, this shift has led to detrimental impacts on the original water allocation arrangements,
especially for farmers in the middle and tail sections of the schemes. In Yemen, water abstractions
upstream from some diversion sites have substantially increased, reducing wadi base flows to down-
stream users.

In Wadi Mawr in Yemen, a large-scale spate diversion system was constructed in the 1980s to enhance
water-use efficiency and improve water supply for irrigation. However, following the construction of
intake structures, sluices and canals to help manage the flood flows, upstream wealthy landowners
prevented sluice and sediment-flushing facilities from working properly. Moreover, a new, and unau-
thorized, canal was constructed to divert and sell water to farmers outside the boundaries of the Wadi
Mawr system. As a result, farmers downstream with original water rights entitlements lost access to
their traditional water supply. Some adapted by investing in wells and exploiting groundwater, but
those who could not afford to do this had to cope with the uncertainty of excess water from large
floods that overflowed the diversion weir. This case highlights how public investment to improve a
water diversion system for the benefit of all farmers can lead to changes in traditional water allocation
arrangements for the benefit of a few.

Source: FAO. 2010. Guidelines on spate irrigation. FAO Irrigation and Drainage Paper 65. Rome.
www.hydrology.nl/images/docs/dutch/key/2010_Guidelines_on_spate_irrigation.pdf

These examples demonstrate that introducing technologies without effective stakeholder participa-
tion and in the absence of suitable legal and institutional frameworks can lead to inequity in access to
water and threaten the sustainability of water resources.

322 5. RESPONSES AND ACTION AREAS


© FAO/Luis Tato
Case study: Restoring rangeland
productivity, biodiversity and
ecosystems in Ethiopia and Jordan

Methods to protect degraded land from overgrazing and enabling grass and fodder crops to recover
have evolved based on indigenous knowledge among farmers. In East Africa, they are called “area
closures” and are widely used to rehabilitate millions of hectares of degraded lands. In Arab countries,
they are called “Al Hima”, which refers to enclosures to protect rangelands. The following examples in
Ethiopia and Jordan highlight the benefits of protecting rangelands that are cost-effective, environ-
mentally beneficial and widely accepted among local communities.

In Ethiopia, “area closures” describe areas of degraded land excluded from human activities and
livestock grazing. Protection encourages natural regeneration through rich and diverse plant cover,
including trees and shrubs, it improves soil health and reduces erosion, it increases productivity and it
enhances economic and ecological benefits to local communities.

Many communities and institutions have reported that lost trees and shrubs species have re-emerged
after two to three years from when areas were closed. Development begins with demarcating and
fencing areas where productivity is poor, based on participatory decision-making involving men and
women from beneficiary communities and local institutions. People and animals are excluded for
three to five years, but this can increase to seven to ten years depending on the degree of degrada-
tion. In some circumstances, additional SLM measures such as terracing, enrichment plantation and
oversowing of grass help accelerate restoration.

Maintenance activities include replanting, maintaining fences, pruning and weeding. Plant materials
are prepared in nearby nurseries and local by-laws are used to regulate and protect enclosures from
trespassers, livestock encroachment and deforestation. Violating the protection rules can result in
punishment by the local authorities and confiscation of materials removed from the protected area.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 323
Examples of degraded and restored landscapes in Ethiopia

© FAO/D. Dale

© FAO/D. Dale
Medium- to long-term benefits include increased fuelwood, vegetation cover, availability of fodder
for livestock feed, medicinal plants and bee forage, thus providing additional income sources and
savings. Cash crops, trees and fodder bushes can also be grown on terraces. Wood for construction
would be available after about seven years. Wider benefits can come from improving the productivity
of downstream farmlands and protecting farmland and communities from flooding.

A cost–benefit analysis of area closures shows the practice has a positive net present value. The
benefit–cost ratio varies between 4.6 and 54.3; that is, a USD 1 investment will bring at least USD
4 through carbon credits. These economic benefits are in addition to carbon dioxide sequestration
benefits that accrue as the land becomes covered in vegetation.

Environmentally, area closures can significantly reduce sediment loads from upstream croplands
and rangelands, reduce runoff coefficients and increase soil moisture. Highly erosive peak flows from
steep slopes are reduced by area closures, and biomass increases carbon stocks. There are positive
impacts on biodiversity, wildlife habitat, floral and faunal diversity, and natural regeneration through
improved seed dispersal. Previously degraded farmlands or grazing lands have regenerated to either
dense or open woodlands, with substantial improvement in the vegetation cover. Springs are also
re-emerging after running dry two or three decades ago. The rise in groundwater tables has made
irrigation more accessible. Farmers in some microwatersheds have dug wells and started small-scale
irrigated cropping.

Involving farmers and communities in area closures and demonstrating the multiple benefits that
come at low cost have encouraged farmers to implement closures on their own initiative and are
helping to ensure sustainability.

Sources: FAO. 2022. Rangeland restoration and sustainable pastoralism go hand in hand. In: Pastoralist knowledge hub. Cited 7 March 2022. Rome. www.
fao.org/pastoralist-knowledge-hub/news/detail/en/c/1044677; Gebrehiwot, T. & Veen, A. 2014. The effect of enclosures in rehabilitating degraded
vegetation: A case of Enderta District, northern Ethiopia. Forest Research, 3: 128; Kasim, M., Assafaw, Z., Deraro, D., Melkato, M. & Mamo, Y. 2016. The role
of area closure in the recovery of woody species composition in degraded land and its socio-economic importance in central rift valley area of Ethiopia.
International Journal of Development Research, 5: 3348–3358; United Nations Convention to Combat Desertification. 2015. Ethiopia - Land degradation
neutrality national report. https://knowledge.unccd.int/sites/default/files/inline-files/ethiopia-ldn-country-report-final.pdf

In Jordan, Al Hima or protected area is a traditional rangeland management system, similar to area
closures, but developed by tribal peoples in Arab countries to survive under harsh climate conditions
and scarce natural resources. Hima provides quick economic and environmental benefits liked by
farmers, pastoralists and herders as the most preferred approach to support livelihoods. “HIMA” is
also an abbreviation for the human integrated management approach, which emphasizes the role of
human activities within this nature conservation system.

324 5. RESPONSES AND ACTION AREAS


Harvesting and processing herbs before and after restoration using Hima

© FAO/D. Dale

© FAO/D. Dale

© FAO/D. Dale
Studies have shown that rangeland productivity in the Jordanian Badia has halved since the beginning
of this century, and many indigenous plant species have disappeared. Hima was adopted in the Zarqa
River basin in the Badia region to help restore the productivity of rangeland where rainfall is low and
land is used mainly for domestic animal grazing. This included fencing selected areas of rangeland,
participatory planning involving local communities, gender mainstreaming and developing alternative
income-generating opportunities such as producing herbal medicines and making soap.

Sustainable land management was a central feature of restoring the rangelands. However, the legal
framework for land tenure and land- and water-use rights initially hindered implementation. Local
SLM knowledge was limited, and local communities were not usually involved in decision-making. All
these issues changed as the project progressed, and communities became interested in SLM as they
gained access to technical support from specialist advisers and began to participate in the project.
Tenure issues were resolved by reclassifying barren and degraded land to rangeland and allocating
land to the care of a cooperative.

Men and women were well represented in the project. Training topics included marketing, processing,
packaging and collecting herbal/medicinal plants and grass for rotational grazing. Local institutions
were established and strengthened. Women, in particular, were increasingly involved in training,
engaging in income-generating activities and decision-making.

Following project completion, land users have taken responsibility for sustaining the gains made
and monitored by the Ministry of Agriculture. A particular benefit is reducing soil erosion and the
sediment-free runoff stored in downstream dams for domestic and farm use.

Hima has highlighted the importance of good governance at community, state and international levels
in preventing and restoring rangelands. The project has also demonstrated the difference between
nature reserves, which exclude people, and community-based natural resources management, which
encourages participation and active involvement. The motivating force was increased profits and a
growing awareness of the value of environmental sustainability.

Sources: WANA Institute. 2018. The concept of Al Hima. In: WANA Institute. Cited 7 March 2022.
http://wanainstitute.org/en/fact_sheet/concept-al-hima; Westerberg, V. & Myint, M. 2015. An economic valuation of a large-scale rangeland restoration
project through the Hima system in Jordan. Report for the ELD Initiative by International Union for Conservation of Nature. Nairobi. https://inweh.unu.
edu/wp-content/uploads/2015/03/ELD_IUCN_Case_Study_Jordan.pdf

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 325
© FAO

Case study: Unconventional


farming in marginal areas

Although some 83 percent of the global population will be living in developing countries by 2025,
the capacity of available land resources and technologies to satisfy the growing demand for food
and other agricultural commodities is far from certain. Exploiting marginal lands could provide a
viable option.

Most agricultural policies favour agricultural lands with high potential, leading to a bias in
policymaking that avoids marginal areas. A fresh policy outlook is needed that investigates options
and innovative technological solutions to make the best use of marginal areas could be vital to meet
future food demands.

Governments usually avoid marginal lands because of biophysical constraints, such as extreme
weather, drought, salinity and socioeconomic conditions. Also, traditional agricultural cropping and
practices may not be productive or economically feasible. Despite these problems, marginal areas
offer territorial advantages and present an opportunity for alternative development models. The
diverse and heterogeneous conditions, including spatial diversity and territorial capital, in marginal
areas have a comparative advantage that can benefit the extreme poor, who are often overlooked
and left behind.

Options in marginal areas include alternative crops that are resource efficient and climate smart.
Regenerative technologies and practices best suited to areas affected by salinity, water scarcity and
drought are options for sustaining marginal and salt-affected lands. Agricultural research has already
documented how marginal lands can be sustainably cultivated with heat-/drought- and salt-tolerant
crops such as barley, amaranth, types of millet, forages and halophyte (salt-tolerant) plants, mainly for
human consumption and animal feed.

326 5. RESPONSES AND ACTION AREAS


In areas where marginality is driven by salinity and waterlogging, agricultural planning involves
a combination of salt-resistant crops and best irrigation management practices and methods to
manage irrigation-induced salinity under the context of biosaline agriculture. Salt-friendly agriculture
represents an opportunity to practice a new type of farming unconventionally by growing salt-tolerant
varieties of conventional crops and halophytes using marginal water, such as drainage water,
produced water and different types of saline water, including rejected brine and seawater. Improved
climate-smart irrigation systems, applying models to improve agricultural water productivity, water
accounting, air-to-water technologies, and water and crop modelling are innovative tools to respond
to water scarcity in marginal environments. Additionally, climate-resilient, nutrient-dense agricultural
schemes that combine fish and crop farming in a saline farming context can support sustainable food
production in increasingly saline environments while contributing to the restoration and protection of
productive natural capital affected by salinity and water scarcity. Such farming approaches benefit the
“circular agriculture economy” models because they minimize the number of external inputs, close
nutrient loops and reduce negative environmental impacts by eliminating discharges.

Quinoa (left) and salicornia (right) growing in a desert environment

Sources: Shahid, S.A. & Al-Shankiti, A. 2013. Sustainable food production in marginal lands—Case of GDLA member countries. International Soil and Water
Conservation Research, 1(1): 24–38; Ahmadzai, H., Tutundjian, S. & Elouafi, I. 2021. Policies for sustainable agriculture and livelihood in marginal lands: A
review. Sustainability, 13(16): 8692; FAO. 2022. WASAG – The global framework on water scarcity in agriculture. In: FAO. Rome. www.fao.org/wasag/work-
ing-groups/saline-agriculture/en

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 327
References
ADB (Asian Development Bank). 2018. 2018 Annual evaluation review: The quality of project design
and preparation for efficiency and sustainability. Manila.
www.adb.org/sites/default/files/evaluation-document/388366/files/2018-aer.pdf

Albrecht, T.R., Crootof, A. & Scott, C.A. 2018. The water-energy-food nexus: A systematic re-
view of methods for nexus assessment. Environmental Research Letters, 13(4): 043002.

Allouche, J., Middleton, C. & Gyawali, D. 2015. Technical veil, hidden politics: Interrogating the
power linkages behind the nexus. Water Alternatives, 8(1): 610–626.

Atlin, G.N., Cairns, J.E. & Das, B. 2017. Rapid breeding and varietal replacement are critical to
adaptation of cropping systems in the developing world to climate change. Global Food Secu-
rity, 12: 31–37.

Balasubramanya, S. & Stifel, D. 2020. Viewpoint: Water, agriculture & poverty in an era of cli-
mate change: Why do we know so little? Food Policy, 93: 101905.

Batchelor, C., Reddy, V.R., Linstead, C., Dhar, M., Roy, S. & May, R. 2014. Do water-saving tech-
nologies improve environmental flows? Journal of Hydrology, 518(A): 140–149.

Biradar, C., Sarker, A., Krishna, G., Kumar, S. & Wery, J. 2020. Assessing farming systems and
resources for sustainable pulses intensification. Conference presentation at Pulses the Cli-
mate Smart Crops: Challenges and Opportunities (ICPulse2020).

Biradar, C.M., Ghosh, S., Löw, F., Singh, R., Chandna, P., Sarker, A., Sahoo, R.N. et al. 2019. Geo
big data and digital augmentation for accelerating agroecological intensification in drylands.
International Archives of the Photogrammetry, Remote Sensing and Spatial Information Sciences,
42(3/W6): 545–548.

Bird, J., Dodds, F., McCornick, P.G. & Shah, T. 2014. Water-food-energy nexus. In: On target
for people and planet: Setting and achieving water-related Sustainable Development Goals, pp.
9–12. Colombo, International Water Management Institute.
https://digitalcommons.unl.edu/wffdocs/4

Burlingame, B., Charrondiere, R. & Mouille, B. 2009. Food composition is fundamental to the
cross-cutting initiative on biodiversity for food and nutrition. Journal of Food Composition
and Analysis, 22(5): 361–365.

Byagmugisha, F. 2013. Securing Africa’s land for shared prosperity: A program to scale up reforms
and investments. Washington, DC, International Bank for Reconstruction and Development/
The World Bank. https://openknowledge.worldbank.org/handle/10986/13837

CFS (Committee on World Food Security). 2015. Water for food security and nutrition: Policy rec-
ommendations. Rome. http://www.fao.org/3/av046e/av046e.pdf

CFS & FAO. 2012. Voluntary guidelines on the responsible governance of tenure of land, fisheries and
forests in the context of national food security. Rome, FAO.
http://www.fao.org/3/i2801e/i2801e. pdf

328 5. RESPONSES AND ACTION AREAS


Chaffin, B.C., Gosnell, H. & Cosens, B.A. 2014. A decade of adaptive governance scholarship:
Synthesis and future directions. Ecology and Society, 19(3): 56.

Clinton, N., Stuhlmacher, M., Miles, A., Uludere Aragon, N., Wagner, M., Georgescu, M., Her-
wig, C. & Gong, P. 2018. A global geospatial ecosystem services estimate of urban agriculture.
Earth’s Future, 6(1): 40–60.

Cosens, B.A., Craig, R.K., Hirsch, S.L., Arnold, C.A., Benson, M.H., Decaro, D.A., Garmestani,
A.S. et al. 2017. The role of law in adaptive governance. Ecology and Society, 22(1): 30.

Crawford, A. 2018. Big data suggests big potential for urban farming. In: WIRED.
www.wired.com/story/big-data-suggests-big-potential-for-urban-farming

Crippa, M., Solazzo, E., Guizzardi, D., Monforti-Ferrario, F., Tubiello, F.N. & Leip, A.
2021. Food systems are responsible for a third of global anthropogenic GHG emissions.
Nature Food, 2(3): 198–209.

Dasgupta, P. 2021. The economics of biodiversity: The Dasgupta review. London, HM Treasury.
https://assets.publishing.service.gov.uk/government/uploads/system/uploads/attach-
ment_data/file/962785/The_Economics_of_Biodiversity_The_Dasgupta_Review_Full_
Report.pdf

Davies, J., Ogali, C., Slobodian, L., Roba, G. & Ouedraogo, R. 2018. Crossing boundaries: Legal and
policy arrangements for cross-border pastoralism. Rome, FAO and Gland, International Union
for Conservation of Nature. www.fao.org/3/ca2383en/CA2383EN.pdf

De Haan, C., Dubern, E., Garancher, B. & Quintero, C. 2016. Pastoralism Development in the
Sahel: A Road to Stability? Washington, DC, International Bank for Reconstruction and
Development/The World Bank. https://openknowledge.worldbank.org/bitstream/han-
dle/10986/24228/K8813.pdf?sequence=2&isAllowed=y

Di Gregorio, M., Fatorelli, L., Pramova, E., May, P., Locatelli, B. & Brockhaus, M. 2016. Inte-
grating mitigation and adaptation in climate and land use policies in Brazil: A policy document
analysis. Centre for Climate Change Economics and Policy Working Paper No. 257. Sustain-
ability Research Institute Paper No. 94. CIFOR Working Paper No. 194. www.cccep.ac.uk/
wp-content/uploads/2016/02/Working-Paper-257-Di-Gregorio-et-al-2016.pdf

Diagnea, M., Demonta, M. & Diagneb, A. 2009. Adoption and impact of an award winning
post-harvest technology: The ASI rice thresher in the Senegal River Valley. Conference pre-
sentation at International Association of Agricultural Economists Conference, 16–22 August
2009. Beijing, International Association of Agricultural Economists.

Dillon, P., Escalante, E.F., Megdal, S.B. & Massmann, G. 2020. Managed aquifer recharge for
water resilience. Water 2020, 12(7): 1846.

Doss, C. & Meinzen-Dick, R. 2020. Land tenure security for women: A conceptual framework.
Land Use Policy, 99: 105080.

Duong, N.T.B. & De Groot, W.T. 2020. The impact of payment for forest environmental services
(PFES) on community-level forest management in Vietnam. Forest Policy and Economics, 113:
102135.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 329
EC (European Commission). 2019. A European green deal. In: EC. Cited 13 July 2021.
https://ec.europa.eu/info/strategy/priorities-2019-2024/european-green-deal_en

Economic Commission for Latin America and the Caribbean. 2021. Natural resources and de-
velopment. In: Economic Commission for Latin America and the Caribbean.
www.cepal.org/en/publications/type/natural-resources-and-development

E3G. 2020. Public bank climate tracker matrix. In: E3G. Cited 13 July 2021. www.e3g.org/matrix

Ellen MacArthur Foundation. 2015. Growth within: A circular economy vision for a competitive Europe.
https://ellenmacarthurfoundation.org/growth-within-a-circular-economy-vision-for-a-
competitive-europe

FAO. 2007a. Modernizing irrigation management – the MASSCOTE approach: Mapping system and
services for canal operation techniques. FAO Irrigation and Drainage Paper 63. Rome.
www.fao.org/3/a1114e/a1114e.pdf

FAO. 2007b. Irrigation management transfer: Worldwide efforts and results. FAO Water Reports 32.
Rome. www.fao.org/3/a1520e/a1520e00.pdf

FAO. 2011a. Biotechnologies for agricultural development. Conference presentation at Agricul-


tural Biotechnologies in Developing Countries: Options and Opportunities in Crops, For-
estry, Livestock, Fisheries and Agro-industry to Face the Challenges of Food Insecurity and
Climate Change (ABDC-10). www.fao.org/biotech/abdc/backdocs/en

FAO. 2011b. Global food losses and food waste: Extent, causes and prevention. Rome.
www.fao.org/3/i2697e/i2697e.pdf

FAO. 2014a. Water and the rural poor: Interventions for improving livelihoods in Asia. Bangkok.
www.fao.org/3/i3705e/i3705e.pdf

FAO. 2014b. The water-energy-food nexus: A new approach in support of food security and sustain-
able agriculture. Rome. www.fao.org/3/bl496e/bl496e.pdf

FAO. 2014c. Turning family farm activity into decent work. Rome.
www.fao.org/fileadmin/user_upload/fao_ilo/pdf/FF_DRE.pdf

FAO. 2016a. Le rôle des femmes dans la gestion des ressources en eau agricole – Phase 2. Rome.
www.fao.org/3/i5680f/i5680f.pdf

FAO. 2016b. Water accounting and auditing: A source book. FAO Water Reports 43. Rome.
www.fao.org/3/i5923e/i5923e.pdf

FAO. 2017a. Watershed management in action: Lessons learned from FAO field projects. Rome.
www.fao.org/3/i8087e/i8087e.pdf

FAO. 2017b. Sustainable land management (SLM) in practice in the Kagera Basin: Lessons
learned for scaling up at landscape level. In: Kagera Agro-ecosystems. Rome. Cited 3 March
2022. www.fao.org/in-action/kagera/news-archive/news-detail/en/c/901665

FAO. 2017c. Sustainable land management (SLM) in practice in the Kagera basin: Lessons learned for
scaling up at landscape level – Results of the Kagera Transboundary Agro-ecosystem Man-
agement Project (Kagera TAMP). Rome. www.fao.org/3/i6085e/i6085e.pdf

330 5. RESPONSES AND ACTION AREAS


FAO. 2017d. Land resource planning for sustainable land management. Land and Water Division
Working Paper 14. Rome. www.fao.org/3/i5937e/i5937e.pdf

FAO. 2017e. Climate-smart agriculture sourcebook: Summary. 2nd edition. Rome.


www.fao.org/3/i7994e/i7994e.pdf

FAO. 2017f. The future of food and agriculture: Trends and challenges. Rome.
www.fao.org/3/a-i6583e.pdf

FAO. 2017g. Does improved irrigation technology save water? A review of the evidence. Rome.
www.fao.org/3/I7090EN/i7090en.pdf

FAO. 2018a. Climate-smart agriculture: Case studies 2018. Successful approaches from different regions.
Rome. www.fao.org/policy-support/tools-and-publications/resources-details/en/c/1177071

FAO. 2018b. Nature-based solutions for agricultural water management and food security. Land and
Water Discussion Paper 12. Rome. www.fao.org/3/ca2525en/ca2525en.pdf

FAO. 2018c. Water accounting for water governance and sustainable development. Rome.
www.fao.org/3/i8868en/i8868en.pdf

FAO. 2018d. Handbook for saline soil management. Rome. www.fao.org/3/i7318en/I7318EN.pdf

FAO. 2018e. Integrated natural resources management in drought-prone and salt-affected agri-
cultural production systems in Central Asia and Turkey. FAO/GEF Regional Project CACILM-2.
www.fao.org/3/CA1423EN/ca1423en.pdf

FAO. 2019a. Water and gender. In: Land & Water. Cited 3 March 2022.
www.fao.org/land-water/water/watergovernance/water-gender/en

FAO. 2019b. Water use in livestock production systems and supply chains: Guidelines for assessment.
Rome. www.fao.org/3/ca5685en/ca5685en.pdf

FAO. 2019c. The international code of conduct for the sustainable use and management of fertilizers.
Rome. www.fao.org/3/ca5253en/ca5253en.pdf

FAO. 2019d. The state of food and agriculture 2019. Moving forward on food loss and waste reduc-
tion. Rome. www.fao.org/3/ca6030en/ca6030en.pdf

FAO. 2020a. Unpacking water tenure for improved food security and sustainable development.
Land and Water Discussion Paper 15. Rome. https://doi.org/10.4060/cb1230en

FAO. 2020b. The state of food security and nutrition in the world 2020. Transforming food systems
for affordable healthy diets. Rome. https://doi.org/10.4060/ca9692en

FAO. 2020c. The state of food and agriculture 2020. Overcoming water challenges in agriculture.
Rome. https://doi.org/10.4060/cb1447en

FAO. 2020d. Peatland mapping and monitoring: Recommendations and technical overview.
Rome. https://doi.org/10.4060/ca8200en

FAO. 2020e. Environmental performance of feed additives in livestock supply chains: Guidelines for
assessment, Version 1. Rome. https://doi.org/10.4060/ca9744en

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 331
FAO. 2020f. A protocol for measurement, monitoring, reporting and verification of soil organic car-
bon in agricultural landscapes: GSOC MRV protocol. Rome. https://doi.org/10.4060/cb0509en

FAO. 2022a. Global Soil Partnership. In: FAO. Rome. www.fao.org/global-soil-partnership/en

FAO. 2022b. Incentives for ecosystem services . In: FAO. Rome.


www.fao.org/in-action/incentives-for-ecosystem-services/resources/case-studies/en

FAO. (forthcoming). Water auditing/water governance analysis: A methodological guide. Rome.

FAO & IHE Delft Institute for Water Education. 2019. Water accounting in the Litani River basin:
Remote sensing for water productivity. Water Accounting Series. Rome.
www.fao.org/3/ca6679en/CA6679EN.pdf

FAO & ITPS (Intergovernmental Technical Panel on Soils). 2015. Status of the world’s soil re-
sources: Main report. Rome. www.fao.org/3/i5199e/i5199e.pdf

FAO & ITPS. 2017. Voluntary guidelines for sustainable soil management. Rome, FAO.
www.fao.org/3/a-bl813e.pdf

FAO & ITPS. 2020. Protocol for the assessment of sustainable soil management. Rome, FAO.
www.fao.org/fileadmin/user_upload//GSP/SSM/SSM_Protocol_EN_006.pdf

FAO & IWMI (International Water Management Institute). 2018. More people, more food, worse
water? A global review of water pollution from agriculture, J. Mateo-Sagasta, S. Marjani & H.
Turral, eds. Rome. www.fao.org/3/ca0146en/ca0146en.pdf

FAO, UNDP (United Nations Development Programme) & UNEP (United Nations Environment
Programme). 2021. A multi-billion-dollar opportunity: Repurposing agricultural support to
transform food systems. Rome. https://doi.org/10.4060/cb6562en

FAO, International Fund for Agricultural Development, United Nations Children’s Fund, World
Food Programme & World Health Organization. 2018. The state of food security and nutrition
in the world 2018. Building climate resilience for food security and nutrition. Rome, FAO.
www.fao.org/3/i9553en/i9553en.pdf

FAO, International Fund for Agricultural Development, United Nations Children’s Fund, World
Food Programme & World Health Organization. 2021. The state of food security and nutrition
in the world 2021. Transforming food systems for food security, improved nutrition and affordable
healthy diets for all. Rome, FAO. https://doi.org/10.4060/cb4474en

FAO, ITPS, Global Soil Biodiversity Initiative, Convention on Biological Diversity & EC.
2020. State of knowledge of soil biodiversity: Status, challenges and potentialities. Rome, FAO.
https://doi.org/10.4060/cb1928en

Foley, J.A. 2018. No, vertical farms won’t feed the world. In: GlobalEcoGuy.org.
https://globalecoguy.org/no-vertical-farms-wont-feed-the-world-5313e3e961c0

Folke, C., Hahn, T., Olsson, P. & Norberg, J. 2005. Adaptive governance of social-ecological sys-
tems. Annual Review of Environment and Resources, 30: 441–473.

332 5. RESPONSES AND ACTION AREAS


Foster, S. & Loucks, D., eds. 2006. Non-renewable groundwater resources: A guidebook on
socially-sustainable management for water-policy makers. Paris, United Nations Educational,
Scientific and Cultural Organization.
https://fr.ircwash.org/sites/default/files/Foster-2006-Nonrenewable.pdf

Franco, A. 2009. Fixação biológica de nitrogênio na cultura da soja no Brasil: Uma lição para
o futuro. (Biological nitrogen fixation on soybean crop in Brazil: A lesson for the future.).
Jan–Apr: 23–24.

Garg, M., Sharma, N., Sharma, S., Kapoor, P., Kumar, A., Chunduri, V. & Arora, P. 2018. Bio-
fortified crops generated by breeding, agronomy, and transgenic approaches are improving
lives of millions of people around the world. Frontiers in Nutrition, 5: 12.

Ghisellini, P., Cialani, C. & Ulgiati, S. 2016. A review on circular economy: The expected transi-
tion to a balanced interplay of environmental and economic systems. Journal of Cleaner Pro-
duction, 114: 11–32.

Global Donor Platform for Rural Development. 2022. Land governance programme map & da-
tabase. https://landgov.donorplatform.org

Gonzalez Fischer, C. & Garnett, T. 2016. Plates, pyramids and planets. Developments in national
healthy and sustainable dietary guidelines: A state of play assessment. Rome, FAO and Oxford,
University of Oxford. www.fao.org/3/I5640E/i5640e.pdf

GWP (Global Water Partnership) & IWMI. 2021. Storing water: A new integrated approach for
resilient development, W. Yu, W. Rex, M. McCartney, S. Uhlenbrook, R. Von Gnechten & J.D.
Priscoli, eds. Stockholm, GWP. www.gwp.org/globalassets/global/toolbox/publications/
perspective-papers/perspectives-paper-on-water-storage.pdf

Haacker, E.M.K., Kendall, A.D. & Hyndman, D.W. 2016. Water level declines in the high plains
aquifer: Predevelopment to resource senescence. Groundwater, 54(2): 231–242.

Hallstein, E. & Iseman, T. 2021. Nature-based solutions in agriculture: Project design for securing
investment. Rome, FAO and Virginia, The Nature Conservancy.
https://doi.org/10.4060/cb3144en

Herrero, M., Thornton, P.K., Mason-D’Croz, D., Palmer, J., Benton, T.G., Bodirsky, B.L., Bog-
ard, J.R. et al. 2020. Innovation can accelerate the transition towards a sustainable food sys-
tem. Nature Food, 1(5): 266–272.

HLPE (High Level Panel of Experts on Food Security and Nutrition). 2014. Food losses and waste
in the context of sustainable food systems. A report by The High Level Panel of Experts on Food Se-
curity and Nutrition. HLPE Report 8. Rome. www.fao.org/3/i3901e/i3901e.pdf

HLPE. 2015. Water for food security and nutrition. A report by The High Level Panel of Experts on
Food Security and Nutrition. HLPE Report 9. Rome. www.fao.org/fileadmin/user_upload/
hlpe/hlpe_documents/HLPE_Reports/HLPE-Report-9_EN.pdf

Hoogeveen, J., Faurès, J.M., Peiser, L., Burke, J. & Van De Giesen, N. 2015. GlobWat – A global
water balance model to assess water use in irrigated agriculture. Hydrology and Earth System
Sciences, 19(9): 3829–3844.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 333
Horst, L. 1998. The dilemmas of water division: Considerations and criteria for irrigation system de-
sign. Colombo, IWMI. https://publications.iwmi.org/pdf/H023588.pdf

IIASA (International Institute for Applied Systems Analysis). 2019. The digital revolution and
sustainable development: Opportunities and challenges. Report prepared by The World in 2050
Initiative. Laxenburg. http://pure.iiasa.ac.at/id/eprint/15913/1/TWI2050-for-web.pdf

IPCC (Intergovernmental Panel on Climate Change). 2019. 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. Geneva. www.ipcc.ch/srccl

IUCN (International Union for Conservation of Nature). 2020. IUCN global standard for na-
ture-based solutions: A user-friendly framework for the verification, design and scaling up of NbS.
1st edition. Gland. https://doi.org/10.2305/IUCN.CH.2020.08.en

IUCN & FAO. 2020. 008 - Developing agroecological practices as nature-based solutions.
In: IUCN. Cited 9 September 2021. www.iucncongress2020.org/motion/008

IWMI. 2021. IWMI and ReWater MENA support online course on water reuse for agriculture.
In: ReWater MENA. https://rewater-mena.iwmi.org/2021/06/08/iwmi-and-rewater-me-
na-supports-online-course-on-water-reuse-for-agriculture

Izzi, G., Denison, J. & Veldwisch, G. 2021. The farmer-led irrigation development guide: A what,
why and how-to for intervention design. Washington, DC, International Bank for Reconstruc-
tion and Development/The World Bank.
https://pubdocs.worldbank.org/en/751751616427201865/FLID-Guide-March-2021-Final.pdf

James, C. 2014. Brief 49: Global area of commercialized biotech/GM crops. In: International Ser-
vice for the Acquisition of Agri-biotech Applications (ISAAA). Cited 3 March 2022.
www.isaaa.org/resources/publications/briefs/49

Jurgilevich, A., Birge, T., Kentala-Lehtonen, J., Korhonen-Kurki, K., Pietikäinen, J., Saikku,
L. & Schösler, H. 2016. Transition towards circular economy in the food system. Sustain-
ability, 8(1): 69.

Karimi, P., Bastiaanssen, W.G.M. & Molden, D. 2013. Water accounting plus (WA+) – A wa-
ter accounting procedure for complex river basins based on satellite measurements.
Hydrology and Earth System Sciences, 17(7): 2459–2472.

Kassam, A., Friedrich, T. & Derpsch, R. 2018. Global spread of conservation agriculture. Interna-
tional Journal of Environmental Studies, 76(1): 29–51.

Kaune, A., Droogers, P., Van Opstal, J., Steduto, P. & Perry, C. 2020. REWAS: REal WAter Savings
tool: Technical document. FutureWater Report 200. Rome.
www.futurewater.nl/wp-content/uploads/2020/06/FAO_REWAS_v08.pdf

Kay, M. 2020. Improving irrigation water use efficiency: A synthesis of options to support capacity de-
velopment. Istanbul, Turkish Water Institute. www.suen.gov.tr/Suen/en/catdty.aspx?val=456

Kay, M.G. & Renault, D., eds. 2004. Capacity development in irrigation and drainage: Issues, chal-
lenges and the way ahead. FAO Water Report 26. Rome, FAO.
www.fao.org/3/y5524e/y5524e00.htm

334 5. RESPONSES AND ACTION AREAS


Keene, S., Troell, J. & Ginsburg, C. 2020. Whose water? A comparative analysis of national laws and
regulations recognising peoples indigenous peoples’, afro-descendants’, and local communities’
water tenure. Washington, DC, Rights and Resources Initiative and Environmental Law Insti-
tute. https://rightsandresources.org/wp-content/uploads/2020/02/WhoseWater.pdf

Kibblewhite, M.G.G. 2018. Contamination of agricultural soil by urban and peri-urban high-
ways: An overlooked priority? Environmental Pollution, 242(B): 1331–1336.

Kristensen, P. 2004. The DPSIR framework. Presented at Workshop on a comprehensive/de-


tailed assessment of the vulnerability of water resources to environmental change in Africa
using river basin approach. Nairobi, UNEP.
https://wwz.ifremer.fr/dce/content/download/69291/913220/.../DPSIR.pdf

Kummu, M., De Moel, H., Porkka, M., Siebert, S., Varis, O. & Ward, P.J. 2012. Lost food, wasted
resources: Global food supply chain losses and their impacts on freshwater, cropland, and
fertiliser use. Science of the Total Environment, 438: 477–489.

Lausche, B. 2019. Integrated planning: Policy and law tools for biodiversity conservation and climate
change. IUCN Environmental Policy and Law Paper No. 88. Gland, IUCN.
https://doi.org/10.2305/IUCN.CH.2019.EPLP.88.en

Lehner, B., Liermann, C.R., Revenga, C., Vörömsmarty, C., Fekete, B., Crouzet, P., Döll, P.
et al. 2011. High-resolution mapping of the world’s reservoirs and dams for sustainable riv-
er-flow management. Frontiers in Ecology and the Environment, 9(9): 494–502.

Lobos, V. 2021. Gobernanza del agua en territorios agrícolas: Estudio de caso en Chile: Subcuenca del
río Tinguiririca. Santiago, FAO. www.fao.org/3/cb3617es/cb3617es.pdf

Löw, F., Biradar, C., Dubovyk, O., Fliemann, E., Akramkhanov, A., Vallejo, A.N. & Waldner, F.
2017. Regional-scale monitoring of cropland intensity and productivity with multi-source
satellite image time series. GIScience & Remote Sensing, 55(4): 539–567.

Lowenberg-DeBoer, J. & Erickson, B. 2019. Setting the record straight on precision agriculture
adoption. Agronomy Journal, 111(4): 1552–1569.

Lubell, M., Blomquist, W. & Beutler, L. 2020. Sustainable groundwater management in Califor-
nia: A grand experiment in environmental governance. Society & Natural Resources, 33 (12):
1447–1467.

Lundqvist, J. & Unver, O. 2018. Alternative pathways to food security and nutrition – Water pre-
dicaments and human behavior. Water Policy, 20(5): 871–884.

Lundqvist, J., Malmquist, L., Dias, P., Barron, J. & Wakeyo, M.B. 2021. Water productivity,
the yield gap, and nutrition: The case of Ethiopia. Land and Water Discussion Paper 17.
Rome, FAO. http://www.fao.org/3/cb3866en/cb3866en.pdf

Mansourian, S., Dudley, N. & Vallauri, D. 2017. Forest landscape restoration: Progress in the
last decade and remaining challenges. Ecological Restoration, 35(4): 281–88.

Munroe, R. & Mant, R. 2014. REDD+ and adaptation: Identifying complementary responses to cli-
mate change. www.unep-wcmc.org/resources-and-data/un-redd-programme-launch-
es-info-brief-on-redd-and-adaptation

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 335
McCartney, M., Foudi, S., Muthuwatta, L., Sood, A., Simons, G., Hunink, J., Vercruysse, K.
& Omuombo, C. 2019. Quantifying the services of natural and built infrastructure in the con-
text of climate change: The case of the Tana River Basin, Kenya. IWMI Research Report 174.
Colombo, IWMI. www.iwmi.cgiar.org/publications/iwmi-research-reports/iwmi-re-
search-report-174

McElwee, P., Huber, B. & Nguyễn, T.H.V. 2020. Hybrid outcomes of payments for ecosystem
services policies in Vietnam: Between theory and practice. Development and Change, 51(1):
253–280.

Naranjo, L. & Willaarts, B.A. 2020. Guía metodológica: Diseño de acciones con enfoque del Nexo
entre agua, energía y alimentación para países de América Latina y el Caribe. Serie Recursos Na-
turales y Desarrollo, no. 197 LC/TS.2020/117. Santiago, Comisión Económica para América
Latina y el Caribe. www.cepal.org/es/publicaciones/46078-guia-metodologica-diseno-ac-
ciones-enfoque-nexo-agua-energia-alimentacion-paises

Nizzetto, L., Futter, M. & Langaas, S. 2016. Are agricultural soils dumps for microplastics of ur-
ban origin? Environmental Science and Technology, 50(20): 10777–10779.

OECD (Organisation for Economic Co-operation and Development). 2015. OECD principles on
water governance. Paris.
www.oecd.org/cfe/regionaldevelopment/OECD-Principles-on-Water-Governance.pdf

OECD. 2017. Guiding principles of effective environmental permitting systems. Paris.


www.oecd.org/env/outreach/37311624.pdf

OECD. 2018. Implementing the OECD principles on water governance: Indicator framework and
evolving practices. OECD Studies on Water. Paris. https://doi.org/10.1787/9789264292659-en

OECD. 2021. Financing a water secure future. Paris.


www.oecd.org/water/brochure-financing-a-water-secure-future.pdf

OECD. 2022. The OECD principles on water governance. Paris.


www.oecd.org/governance/oecd-principles-on-water-governance.htm

Ostrom, E. 2005. Understanding institutional diversity. Princeton, Princeton University Press.

Ostrom, E. 2010. Beyond markets and states: Polycentric governance of complex economic sys-
tems. American Economic Review, 100(3): 641–672.

Otoo, M. & Drechsel, P. 2005. Resource recovery from waste: Business models for energy, nutrient
and water reuse. London, Earthscan.

Oweis, T. 2016. Effective mechanised rainwater harvesting: Coping with climate change implications in
the Jordanian badia. Amman, International Center for Agricultural Research in the Dry Areas.
www.icarda.org/publications/6061/effective-mechanized-rainwater-harvesting-cop-
ing-climate-change-implications

Oxfam & ILC (International Land Coalition). 2016. Common ground: Securing land rights and
safeguarding the earth. Briefing Paper. Oxford, Oxfam. https://policy-practice.oxfam.org/re-
sources/common-ground-securing-land-rights-and-safeguarding-the-earth-600459

336 5. RESPONSES AND ACTION AREAS


Pahl-Wostl, C. 2015. Water governance in the face of global change: From understanding to transfor-
mation. Springer. https://link.springer.com/content/pdf/10.1007%2F978-3-319-21855-7.pdf

Pahl-Wostl, C., Mostert, E. & Tàbara, D. 2008. The growing importance of social learning in
water resources management and sustainability science. Ecology and Society, 13(1): 24.

Pahl-Wostl, C., Sendzimir, J., Jeffrey, P., Aerts, J., Berkamp, G. & Cross, K. 2007. Managing
change toward adaptive water management through social learning. Ecology and Society,
12(2): 30.

Pahl-Wostl, C., Arthington, A., Bogardi, J., Bunn, S.E., Hoff, H., Lebel, L., Nikitina, E. et al.
2013. Environmental flows and water governance: Managing sustainable water uses.
Current Opinion in Environmental Sustainability, 5(3–4): 341–351.

Pfitzer, M. & Krishnaswamy, R. 2007. The role of the food & beverage sector in expanding economic
opportunity. Corporate Social Responsibility Report No. 20. Cambridge, MA, Kennedy School
of Government, Harvard University. www.issuelab.org/resources/1708/1708.pdf

Pham, T.T., Thi, M.N.B., Thi, L.C.Ð., Hoàng, T.L., Pham, H.L. & Nguyen, V.D. 2018. The role
of payment for forest environmental services (PFES) in financing the forestry sector in Vietnam.
Center for International Forestry Research.

Plummer, R., Crona, B., Armitage, D.R., Olsson, P., Tengö, M. & Yudina, O. 2012. Adaptive co-
management: A systematic review and analysis. Ecology and Society, 17(3): 11.

Plusquellec, H. 2014. Improving performance of canal irrigation systems is long overdue.


http://fliphtml5.com/pjhk/uwyq/basic

Reed, M. & Massie, M. 2013. Embracing ecological learning and social learning: UNESCO bio-
sphere reserves as exemplars of changing conservation practices. Conservation and Society,
11(4): 391–405.

Renault, D. & Wallender, W.W. 2000. Nutritional water productivity and diets. Agricultural Water
Management, 45(3): 275–296.

Roobroek, D., Van Asten, P.J.A., Jama, B., Harawa, R. & Vanlauwe, B. 2015. Integrated soil fer-
tility management: Contributions of framework and practices to climate-smart agriculture.
Climate-smart Agriculture Practice Brief. Copenhagen, CGIAR Research Program on Climate
Change, Agriculture and Food Security. https://ccafs.cgiar.org/resources/publications/inte-
grated-soil-fertility-management-contributions-framework-and

Rosa, L., Chiarelli, D.D., Rulli, M.C., Dell’Angelo, J. & D’Odorico, P. 2020. Global agricultural
economic water scarcity. Science Advances, 6(18).

Ros-Tonen, M.A.F., Reed, J. & Sunderland, T. 2018. From synergy to complexity: The trend to-
ward integrated value chain and landscape governance. Environmental Management, 62(1):
1–14.

RRI (Rights and Resources Initiative). 2017. Securing community land rights: Priorities and oppor-
tunities to advance climate and Sustainable Development Goals. https://rightsandresources.org/
wp-content/uploads/2017/09/Stockholm-Prorities-and-Opportunities-Brief.pdf

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 337
RRI & ELI (Environmental Law Institute). 2020. Whose water? A comparative analysis of na-
tional laws and regulations recognising indigenous peoples’ and local communities’ rights to
freshwater. Washington, DC.
https://rightsandresources.org/wp-content/uploads/2020/03/Whose-Water.pdf

Searchinger, T., Waite, R., Hanson, C. & Ranganathan, J. 2019. Creating a sustainable food fu-
ture: A menu of solutions to feed nearly 10 billion people by 2050. World Resources Report.
Washington, DC, World Resources Institute.
https://files.wri.org/d8/s3fs-public/wrr-food-full-report.pdf

Seckler, D. 1996. The new era of water resources management: From “dry” to “wet” water savings.
Research Report No. 1. Colombo, International Irrigation Management Institute.
www.iwmi.cgiar.org/Publications/IWMI_Research_Reports/PDF/pub001/REPORT01.PDF

Settle, W. & Garba, M.H. 2011. Sustainable crop production intensification in the Senegal and
Niger River basins of francophone West Africa. International Journal of Agricultural Sustain-
ability, 9(1): 171–185.

Shafiee-Jood, M. & Cai, X. 2016. Reducing food loss and waste to enhance food security and en-
vironmental sustainability. Environmental Science and Technology, 50(16): 8432–8443.

Shah, T. 2009. Taming the anarchy: Groundwater governance in South Asia. Washington, DC, Re-
sources for the Future and Colombo, IWMI.

Shah, T. 2022. Water-energy-food nexus in action: Global review of policy and practice.
Thematic Background Report for SOLAW 2021. Rome, FAO.
www.fao.org/land-water/solaw2021/en

Sjoerd, E., Stofberg, S.F., Yang, X., Liu, Y., Islam, M.N. & Hu, Y.F. 2017. Irrigation and drainage
in agriculture: A salinity and environmental perspective. In: Current Perspective on Irrigation
and Drainage. IntechOpen.

Springmann, M., Clark, M., Mason-D’Croz, D., Wiebe, K., Bodirsky, B.L., Lassaletta, L., de
Vries, W. et al. 2018. Options for keeping the food system within environmental limits.
Nature, 562(7728): 519–525.

Tripathi, V., Edrisi, S.A., Chaurasia, R., Pandey, K.K., Dinesh, D., Srivastava, R., Srivastava, P.
& Abhilash, P.C. 2019. Restoring HCHs polluted land as one of the priority activities during
the UN-International Decade on Ecosystem Restoration (2021–2030): A call for global ac-
tion. Science of the Total Environment, 689: 1304–1315.

UNECE (United Nations Economic Commission for Europe). 2021. UN Watercourses Conven-
tion. In: UNECE. Cited 3 March 2022.
https://unece.org/environment-policy/water/un-watercourses-convention

UNEP. 2015. Building inclusive green economies in Africa: Experience and lessons learned 2010-2015.
Nairobi. www.unep.org/resources/report/building-inclusive-green-economies-africa-ex-
perience-and-lessons-learned-2010-2015

UNEP. 2019. Environmental rule of law: First global report. Nairobi.


https://wedocs.unep.org/handle/20.500.11822/27279

338 5. RESPONSES AND ACTION AREAS


UNEP. 2021. First draft of the Post-2020 Global Biodiversity Framework. CBD/WG2020/3/3.
www.cbd.int/doc/c/abb5/591f/2e46096d3f0330b08ce87a45/wg2020-03-03-en.pdf

United Nations. 2015. Transforming our world: The 2030 agenda for sustainable development.
New York. A/RES/70/1. www.un.org/en/development/desa/population/migration/generalas-
sembly/docs/globalcompact/A_RES_70_1_E.pdf

United Nations. 2018. Sustainable Development Goal 6: Synthesis report 2018 on water and
sanitation. New York.
www.unwater.org/publications/sdg-6-synthesis-report-2018-on-water-and-sanitation

USAID (United States Agency for International Development). 2016. Fact sheet: Land tenure
and women’s empowerment. In: LandLinks. Cited 3 March 2022.
www.land-links.org/issue-brief/fact-sheet-land-tenure-womens-empowerment

Varshney, R.K., Ribaut, J.M., Buckler, E.S., Tuberosa, R., Rafalski, J.A. & Langridge, P. 2012.
Can genomics boost productivity of orphan crops? Nature Biotechnology, 30(12): 1172–1176.

Venkatesh, M. & Hurrell, R., eds. 2018. Food fortification in a globalized world. Academic Press.

Viña, C.S.-L. 2020. Beyond title: How to secure land tenure for women. In: World Resources Insti-
tute. Washington, DC. Cited 3 March 2022.
www.wri.org/insights/beyond-title-how-secure-land-tenure-women

Whitehead, A.L., Kujala, H. & Wintle, B.A. 2017. Dealing with cumulative biodiversity im-
pacts in strategic environmental assessment: A new frontier for conservation planning.
Conservation Letters, 10(2): 195–204.

WMO (World Meteorological Organization) & GWP. 2017. Selecting measures and designing
strategies for integrated flood management: A guidance document. Geneva, WMO.
https://www.floodmanagement.info/publications/guidance%20-%20selecting%20mea-
sures%20and%20designing%20strategies_e_web.pdf

WOCAT (World Overview of Conservation Approaches and Technologies). 2013. Water harvest-
ing: Guidelines to good practices. https://www.wocat.net/library/media/25/

World Bank Group. 2020. Mobilizing private finance for nature. Washington, DC. http://pubdocs.
worldbank.org/en/916781601304630850/Finance-for-Nature-28-Sep-web-version.pdf

WWF (World Wide Fund for Nature). 2017. Natural and nature-based flood management: A green
guide. https://files.worldwildlife.org/wwfcmsprod/files/Publication/file/538k358t40_
WWF_Flood_Green_Guide_FINAL.pdf

Zhang, Y., Massel, K., Godwin, I.D. & Gao, C. 2018. Applications and potential of genome editing
in crop improvement. Genome Biology, 19: 210.

Zhang, H., Zhou, Y., Huang, H., Wu, L., Liu, X. & Luo, Y. 2016. Residues and risks of veteri-
nary antibiotics in protected vegetable soils following application of different manures.
Chemosphere, 152: 229–237.

Ziadat, F., Bunning, S. & De Pauw, E. 2017. Land resource planning for sustainable land manage-
ment: Current and emerging needs in land resource planning for food security, sustainable liveli-
hoods, integrated landscape management and restoration. Land and Water Division Working
Paper 14. Rome, FAO. www.fao.org/3/a-i5937e.pdf

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE 2021 339
6
Conclusions and
recommendations
The land and water resources behind global food supply systems are rarely recognized as critical pathways
to the transformation advocated by the 2021 United Nations Food Systems Summit. Yet more than
95 percent of food is directly dependant on land and freshwater, and 68 percent of aquaculture production
is derived from inland waters. The FAO outlook for 2050 projects that agricultural production will need
to add 50 percent more food, fibre and biofuel to satisfy human demand than in 2012. The implications
for land and water are profound. Over 33 percent of agricultural land is degraded, and any expansion will
necessarily involve further deforestation or recovery of degraded land. With climate change factored in,
global agricultural freshwater withdrawals would need to increase by as much as 30 percent above the level
in 2012. This would take total global withdrawals to within 10 percent of the annually renewable freshwater
resources generated on land.
Land suitability is expected to shift poleward under climate change. In addition, the impacts of climate
change on agriculture are felt through water. Changing patterns of rainfall and relative humidity
determine all aspects of crop growth. The management of soil moisture and its deficits lies at the centre
of agricultural adaptation.

It is for these reasons that SOLAW 2021 makes the case that immediate threats to land and water can
be addressed by transforming approaches to land and water governance and adaptive management to
keep productive land in play while contributing towards zero net GHG emissions. The functions of these
systems have to be safeguarded.
freshwater pollution on food production is
©FAO/Mohammad Rakibul Hasan
spreading and accelerating in low-income
countries. In addition, without well-designed
land policy and enforcement, rapid urbaniza-
tion in developing countries often takes over
some of the most productive arable land.

Current patterns of agricultural intensi-


fication are not sustainable. Agricultural

6.1 The state of production remains far below sustainabil-


ity levels. Changing land-use patterns and

play offers no room concentrating inputs are producing unac-

for complacency
ceptably high levels of pollution and GHG
emissions. These patterns of production
stretch the productive capacity of agricultural
The current state of land, soil and water systems to the limit and severely degrade
resources and the trends in their use reflect their associated environmental services.
the pressures and drivers imposed by grow-
ing populations and expanding economies. The combination of land degradation and

World food demand is predicted to increase water scarcity threatens food security.

by 50 percent over the next 30 years, with Agricultural intensification degrades soil

the greatest needs in developing countries. structure and water for other uses, and

While it is expected that production will depletes nutrients. These are reversible by

respond to rising demand, this will not be the reducing agriculture pollution, restoring

only measure of success. The environmen- the land, improving water quality, and

tal sustainability of the productive land and remediating soils to maintain productivity

water systems and their capacity to satisfy and reduce GHG emissions. Groundwater is in

the livelihood requirements of urban and crisis due to overexploitation from irrigation

rural populations will be essential criteria. and pollution derived from agricultural inputs
and untreated urban waste. Groundwater
Since SOLAW 2011, most of the growth in depletion and degradation is the first sign
global agricultural production has been of water scarcity, affecting vulnerable rural
derived from input intensification, particu- populations that depend on access to land
larly on prime agricultural land equipped for and water for subsistence and then spreading
irrigation. By contrast, rainfed systems in the at scales that affect national food security.
tropics and mountain regions have exhibited
slower increases in productivity. Many uses
of land and water systems are continuing
to impose negative impacts on ecosystem
services.
©FAO/Giulio Napolitano

There is little room for expanding the


productive land area. Deforestation and land
drainage continue to deplete protected areas,
despite attempts to limit encroachment. The
local impact of physical water scarcity and

342 6. CONCLUSIONS
5. RESPONSES
AND RECOMMENDATIONS
AND ACTION AREAS
6.2 Socioeconomic
Farming systems are becoming polarized.
The social and economic structure of most

development populations is finely tuned to natural resource

pathways are
access, even as populations concentrate in
urban areas. Large-scale commercial hold-

diverging ings dominate agricultural land use to supply


global food. Land tenure patterns restrict and
concentrate up to 500 million smallholdings
Demographic and economic growth increase
(less than 2 ha) in subsistence farming on
food demand, placing unprecedented pres-
lands susceptible to climate change, degra-
sures on ecosystems and limited renewable
dation and water scarcity.
land, soil and water resources. Socioeconomic
trends, including urbanization, migration Globally, 80 percent of the extreme poor live
and technological change, will continue to in rural areas. Most live in the developing
drive the distribution of these pressures world; their livelihoods are disproportionally
on available natural resources. Higher
dependent on agriculture, which is highly
incomes and urban lifestyles change food
exposed to current and future climate risks.
demand towards more resource-intensive
Ensuring equitable access to land and water
consumption of animal proteins, fruits and
resources is key for promoting inclusive rural
vegetables. At the same time, malnutrition
development. The lack of adequate access
persists among the urban and rural poor
and user rights and increasing disparities in
who are disconnected from markets or have
capacities to take advantage of land and water
limited access to productive land due to
endowments are exacerbating rural poverty.
poverty or geography.

Policy and legal frameworks governing land


Increasing population means there are
and water resources are fragmented, lack
reduced natural resources available
implementation or have proven ineffective.
per capita. Underlying the patterns of
economic growth, competition for land and Many are not adapted to cope with the range

water resources is intensifying. Increasing and depth of environmental shocks that

population is reducing the amount of natural are anticipated in the future under climate
resources available per capita. In 2018, more change. They also risk perpetuating the
than 733 million people lived in countries with current trends that concentrate land under
high (70 percent) and critical (100 percent) large commercial concerns and fragment
water stress areas, accounting for almost tenure among smallholder communities.
10 percent of the global population. Over
the past two decades, sub-Saharan Africa
experienced a 40 percent reduction in
water availability per capita between 2000
and 2017. West Africa, Northern Africa and
Southern Africa now have less than 1 700 m3
of agricultural land per capita, which is
considered to be a level that compromises
©FAO/Noah Seelam

a nation’s ability to meet water demand


for food. Similarly, there was a 22 percent
decline in arable land per capita between
2000 and 2019.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE
AGRICULTURE: 2021
SOLAW 2021 343
6.3 The risks to a
The depth of the socioeconomic trade-offs
between agricultural production and envi-
ronmental services depends on land-use
food-secure future
are proliferating
management. Such trade-off decisions will
require sound data and information to fully
understand the consequences of socioeco-
nomic outcomes and environmental impacts. Land and water systems are under pressure
Decisions taken will need to include ways – and some are at breaking point. There
of reducing the risks and their impacts to are mounting pressures on productive land,
avoid further degrading natural resources soil and water resources that are creating
while maintaining food security and poverty comprehensive land degradation and water
targets. There may be important syner- scarcity. Unprecedented heat and shifting
gies and trade-offs that cannot be addressed rainfall patterns already affect agricultural
by single sector strategies and investments production. Long-term adaptation to climate
alone. Initiatives in the water–food– change is necessary. Limits on the global food
energy nexus approach can help to opti- system need to be recognized, and alternative
mize resource-use efficiency, but ultimately land- and water-management approaches
land-use planning and the process of water taken to avoid, mitigate or manage risks.
allocation will need to become truly inclusive.
Pressures on land and water systems are
Overall recommendation: Inclusive and compromising agricultural productivity,
effective land and water governance will precisely at times and in places where growth
need to be applied, to underpin the required is most needed to meet global food security
productivity gains to meet global food targets. Land degradation and water scarcity
demand. Governance over land and water raise risk levels for agricultural production
resources can perform if there is an enabling and ecosystem services. A converging range
environment in which land- and water- of economic drivers and climate variability
management actions take place at multiple are affecting the long-term viability of global
levels of decision-making. Social, agricul- food systems.
tural and environmental policies need to be
mutually reinforcing if they are to recon-
cile competition over land and water. There
is progress in land tenure initiatives, but
land and water allocation adjustments will
be possible only when explicit instruments
are joined up and resources management
decision-making becomes inclusive. Inte-
grated land and water planning is urgently
needed to guide land and water use and
not just to promote sustainable resources
management. To establish a realistic scope
for reducing GHG emissions, land-use plan-
© Frank Nyakairu

ning will need to become much more inclu-


sive and focus on alternative strategies for
crop fertilization and soil management.

344 6. CONCLUSIONS
5. RESPONSES
AND RECOMMENDATIONS
AND ACTION AREAS
The risk to agricultural production from
land degradation is significant. However, it
is rarely factored in until cropland soils and
pastures are significantly depleted or lost
©FAO/Giulio Napolitano

because of human-induced erosion, saliniza-


tion and pollution. Climate change is expected
to further hamper growing conditions for
crops and natural ecosystems in subtropical
developing countries, whereas warming in
Climate change adds uncertainty to the agro- temperate latitudes could extend growing
climatic risks facing producers, particularly seasons for some cereals. Sustainable land
those who are least able to buffer shocks and and water management across all agrocli-
who are food insecure. Climate volatility and matic zones will become a priority to reduce
extreme hydrological and temperature events GHG emissions and increase food production.
will affect all producers, but risks are greater
in areas with minimal resource endowments, Water scarcity increases agricultural

a growing population and limited economic production risks as water supply, storage and

power to adapt local food systems or find conveyance systems reach their design limits.

substitutes. Specifically, climate change is In many areas with high water stress, farmers

expected to increase evapotranspiration manage their production risks by abstracting

from cropped land and alter the quantity and shallow groundwater for irrigation; in some

distribution of rainfall, leading to changes in cases, they use non-renewable groundwater.

land/crop suitability. Greater variations in However, competition for diminishing

river runoff and groundwater recharge are quantities of high-quality groundwater

expected, and will affect rainfed and irrigated is intensifying as aquifers suffer from

agriculture. At the same time, changes in overabstraction and saline intrusion. Many

overall temperature and rainfall regimes are aquifers also suffer from agricultural and

expected to result in poleward shifts in land industrial pollution. Droughts are slow to

suitability. develop and not easily recognized at first,


but they can quickly become a crisis when
Looking to the future. FAO estimates that by severe and damaging impacts emerge that are
2050, agriculture will need to produce almost widespread and have underestimated impacts
50 percent more food, fibre and biofuel than in on societies, ecosystems and economies.
2012. Agricultural production in sub-Saharan
Africa and South Asia will need to more than
double (increase of 112 percent) to meet esti-
mated calorific requirements. The rest of the
world will need to produce at least 30 percent
more. Achieving this will mean increasing
crop yields and cropping intensities, as there
are limited options for expanding the culti-
© Giulio Napolitano

vated area. The current annual cereal yield


growth rate remains below 1 percent and is a
warning that staple food production can fall
behind growing demand.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE
AGRICULTURE: 2021
SOLAW 2021 345
Water pollution from agriculture is prolif- Acquiring the data to support planning is
erating, as is pollution from domestic and vital. The tools for sustainable land, soil and
industrial processes. New and emerging water planning and management are avail-
pollutants are adding to clean-up costs and able to assess the potential impacts of climate
challenging technological solutions on land change on crop production and to tackle the
and in lacustrine and nearshore marine growing pressures on freshwater ecosystems
environments. and degrading land, soil and water quality.
Data collection needs to be comprehensive
Overall recommendation: The operational and smarter if it is to provide the basis for
decisions for agricultural production should transparent water accounting for agricul-
be better informed of economic and envi- tural water use and the quality of return
ronmental consequences. The risk to food flows. Monitoring the accumulated impact of
production can be mitigated by changing climate change in relation to agroecological
agricultural and land-management practices suitability will prove essential for planning
to reduce impacts on livelihoods, human resource use along the entire food value and
health and the delivery of ecosystem services. supply chains.
Using LRP tools together with climate models
provides invaluable insights into how these Integrated multisectoral approaches need not
changes will redistribute land available for be complex, they can be intuitive. Solutions
production for different crops and live- require close collaboration across sectoral
stock, and identifies potential impacts on boundaries where interests align. Planning
productivity and yield gaps. However, none and implementing measures that sustain
of this can be done without the land planning productivity, reduce pollution, seques-
process engaging with urban development ter carbon and mitigate emissions can be
and poverty reduction strategies that affect straightforward. Tested technologies in SLM
spatial planning including water governance. can be combined with inclusive planning
approaches at scale when good land and

6.4 Responses water governance is in place.

should be better
planned – the
tools are in place
Taking stock of land and water assets is
necessary to address the range of adaptive
management processes and attain national
emission targets for agricultural land. It is
therefore important to consider realistic
©FAO/Simon Maina

forms of spatial planning for agricultural


land use for which economic trade-offs can
be evaluated, and policies in public subsidies
can be directed.

346 6. CONCLUSIONS
5. RESPONSES
AND RECOMMENDATIONS
AND ACTION AREAS
The tools for sustainable land, soil and water
planning and management are available.
Planning tools can define critical thresholds
in natural resource systems, leading to the
reversal of land degradation when wrapped
up as packages or programmes of techni-
©FAO/Giulio Napolitano

cal, institutional, governance and financial


support. In this respect, LDN under the
UNCCD Global Mechanism can help govern-
ments set targets and plan interventions
based on the principle of Avoid > Reduce >
Reverse land degradation.
Resource planners and policymakers can
respond to the challenge using a mix of Preventing degradation is far less costly
policy tools (e.g. incentives, impact invest- than restoration. The LDN approach is advo-
ments and regulations), remote sensing, cated to find pathways out of human-induced
big data and innovative analytical meth- cycles of land degradation and water scarcity.
ods that are revolutionizing approaches to
resource assessment and planning. A wide Overall recommendation: Land resources plan-
range of resource assessment and planning ning is needed at all levels of decision-making
tools and approaches support decision- to address challenges set by changing human
makers, planners and practitioners, working demands. When matched with appropriate
at global, national and local levels to plan, financing mechanisms and combined with
take action, and scale out SLM and sustain- environmental regulation, LRP can provide
able water-management options. the essential impetus to reverse trends in
land degradation.
Models are now essential tools for land-use

6.5 Channel
planning and land and water resources
planning. They are increasingly being used
together with participatory approaches to
actions in four
response areas
develop better-adapted food and agricul-
tural systems. Combining land and water
resource assessment and planning tools,
including GAEZ methods, with the latest The governance of land and water resources
climate models provides invaluable insights underpins productive food systems. Demand-
into how changes will redistribute land avail- ing that these systems are, at the same time,
able for agricultural production and affect efficient, resilient and inclusive of those
water availability. These include shifts in who produce them and those who depend
areas suitable for different crop and livestock upon them is a tall order. Actions in four key
species and farming systems, and identify- response areas can enable and facilitate a
ing potential impacts on productivity and transition by all actors to sustainable land and
yield gaps. In particular, shifting to a risk water management. Taken together, this set
management approach can significantly of responses can transform current patterns
lessen drought risks and impacts. of land and water management in agriculture
and reduce GHG emissions from land.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE
AGRICULTURE: 2021
SOLAW 2021 347
The world has largely met the increasing multiple objectives related to natural
demand for food, feed and fibre over the past resources management, trade-offs, and
century by expanding the cultivated area and related ecosystems and services. Coherence
intensifying the use of land, soil and water is needed across all levels of government
resources. However, increasing resource and policy areas, as decisions outside the
scarcity and inequalities of access are chang- water and land domain can significantly
ing the global dynamics concerning food, affect natural resources. Understanding and
climate, energy and the allocation of financial recognizing the relationship of customary
resources to solve social and environmental and statutory land and water rights and the
problems. The ensuing economic tension has role of hybrid legal systems for inclusive
exposed the reality of shared dependency. water and land tenure regimes can form the
The ongoing COVID-19 pandemic is another basis for achieving a diverse array of policy
complicating factor, together with the and development goals.
suppression of agricultural production due to
armed conflict. Ensuring land, soil and water Collaborative decision-making and learning

resources are used in a sustainable manner requires deliberate linkages across insti-

requires careful negotiation of competing tutions, scales and sectors to capitalize on

interests in economic growth, equity and a stakeholders’ diverse knowledge, experiences

healthy environment. Negotiations will inev- and values to ensure negotiated trade-offs are

itably involve trade-offs among interests, but realistic, innovative and equitable. Actions

more importantly, they should open the way will also need to be inclusive across physical

for improved forms of agricultural practice and economic landscapes.

on land. Land and water management can


Land and water governance needs to shift
respond to meet the challenges of climate
up a gear. Reducing pressures on produc-
change through environmental, social and
tive land and water systems and adjusting
governance approaches to investment in
their allocation will be possible only when
actions that are better planned, informed
explicit land and water policy instruments
and implemented, but above all, better
are collaborative and resources management
governed. Without doubt, governance is the
decision-making is inclusive.
most important element in putting technol-
ogies into practice at territorial scale with
the consent of all interests. Improved prac-
tice and technology will have little impact
without it.

6.5.1 Action area I:


Adopting inclusive land
and water governance
Effective and inclusive governance is
essential for building capable and informed
© FAO/Simon Maina

institutions and organizations. However,


advances in land and water governance
require coherent and integrated policies
in the various sectors to deliver on the

348 6. CONCLUSIONS
5. RESPONSES
AND RECOMMENDATIONS
AND ACTION AREAS
management practices within a systems
approach for tackling mitigation and adapta-
tion in agriculture. This emphasizes improved
management of soil carbon, soil health and
soil fertility under grassland, cropland and
©FAO/Aris Mihich

integrated systems, including water quality


and watershed management.

6.5.3 Action area III:


Embracing innovative
6.5.2 Action area II: technologies and
Implementing integrated management
solutions at scale
Technical responses are now better targeted
Agriculture’s “solution space” has expanded. across agriculture to significantly improve
Advances in agricultural research have the sustainable management of land, soil
broadened the technical palette for land and and water. Digital agriculture needs to be
water management. Nature-based solutions accessible to all, and combined with advances
can address pest control, water quality and in farming hardware to target production
quantity attainment, crop phenology and and improve the environmental footprint
biodiversity. Applied at scale, they can reduce of agriculture. Mobile technologies are
the build-up of environmental pressures and spreading rapidly across the agricultural
obtain LDN. sector, together with innovative on-farm
mechanization. Combining these with
There is no “one size fits all” solution
remote-sensing services, cloud-based
– a “package” of solutions is envisaged.
computing, and open access to data and
Increasing land and water productivity is
information (“big data”) on crops, natural
crucial for achieving food security, sustain-
resources, climatic conditions, inputs and
able production and SDG targets. A “pack-
markets already benefits smallholder farmers
age” of workable solutions is now available to
by integrating them into digitally innovative
enhance food production and tackle the main
agrifood systems. However, care is needed
threats from land degradation, increasing
to avoid a “digital divide” among those with
water scarcity and declining water quality.
different levels of access to new technologies.
But these will succeed only when there is
a conducive enabling environment, strong
political will, sound policies and inclusive
governance, and full participatory planning
processes across all sectors and landscapes.

Measures to adapt to and mitigate the


impacts of climate change in agriculture are
©FAO/Jim Morgan

part of a continuum ranging from address-


ing the drivers of vulnerability to explic-
itly targeting climate change impacts. The
landmark KJWA places soil- and water-

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE
AGRICULTURE: 2021
SOLAW 2021 349
to tackle misuse, underuse and overuse of
fertilizers. Management of animal waste
through controlled spreading on agricultural
lands can be incentivized through an appro-
priate mix of governmental incentive and
regulatory policies, including the treatment
©FAO/Marco Longari

of biosolids and climate-smart biodigesters.

Responses to drought can shift to a risk


management approach and reduce impacts
on rainfed and irrigated production. This
Sustainable land-management techniques will require investment in early warning
can address the ten main soil degradation and monitoring systems to be combined
threats while conserving water resources, with outreach of climate-smart agriculture
vegetation and biodiversity. These include techniques, especially in areas where rural
controlling soil erosion through rainfall and communities are particularly vulnerable.
runoff management, and replenishing SOM,
SOC and nutrient balance in soil systems. Water scarcity is driving renewed interest in
The WOCAT database offers over 1 500 tried irrigation. The modernization of irrigation
and tested technologies to support SLM and to focus attention on farmer demand for
contribute to LDN targets in avoiding and more flexible and reliable water supplies will
reducing degradation and restoring degraded raise productivity. New planning, design and
lands. The network enables practitioners evaluation technologies, water accounting
and technical experts worldwide to select, and auditing, ICT and automation have to be
document and share their best practices and part of a package of modernization if they
experiences from specific socioeconomic and are to result in real water savings per unit of
biophysical contexts that support SLM and production. Traditional methods of improv-
restoration. ing water-use efficiency in crop production to
“save” water are being challenged. Research
Reducing chemical inputs and the concen- shows that so-called “efficient technolo-
tration of animal waste on land and water gies”, such as pressurized, sprinkler and
should be a global priority. The mobility and drip irrigation and canal lining, can increase
persistence of many agricultural nutrients evaporative consumption, reduce recharge
and chemicals in soils, surface water and and return flows to others.
groundwater are affecting human health and
ecosystem function. Input control and use
of alternative conservation methods need to
be mainstreamed in all farming systems and
targeted in priority river basins and catch-
ments. Integrated pest management and the
Fertilizer Code are some of the instruments
© Believe Nyakudjara

designed to counter the trend towards unsus-


tainable agricultural intensification and the
potential for increased use and harmful
effects of fertilizers, pesticides and herbi-
cides. The Fertilizer Code offers guidance

350 6. CONCLUSIONS
5. RESPONSES
AND RECOMMENDATIONS
AND ACTION AREAS
At river basin level, freshwater storage
6.5.4 Action area IV:
provides an essential buffer for managing
climate uncertainty and variability while
Investing in long-
juggling differences in supply and demand term sustainability
and building overall societal resilience. As
Future investments will need to move away
ageing infrastructure is being repaired or
from pure hardware solutions to sustain-
replaced, a new agenda is changing the
ing rainfed and irrigated production systems
current thinking about storage to address
through improved governance, integrated
all forms of freshwater storage, natural and
interventions at scale, and innovation in
built, in an integrated system that provides
management and technology. Investment is
multiple benefits.
needed in data and information collection
and management, to improve connectivity
Overall, the agricultural sector needs to
among all producers, markets and regulators.
better manage environmental risks – and
On-farm decision-making will then improve
at scale. The circular economy concept is
dramatically, and agronomic innovation can
just as applicable to agricultural land and
be combined with early warning systems and
water management as it is to the broader
performance monitoring offered by advances
global food system. It offers opportunities
in near-real-time dissemination of environ-
to use non-conventional waters that might
mental data.
otherwise go to waste, such as saline and
brackish water, agricultural drainage water
Private investment in land management will
and reclaimed water. Other aspects of reuse
prove decisive. Farmers can be incentivized
within the farming system include nutri-
to become investors in sustainable land and
ent recycling, regenerating soil health, and
water management when supported through
reducing non-renewable energy and mate-
innovative financing and instruments that
rials and inputs used in rainfed and irrigated
reconcile production and environmental
systems. In this sense, agriculture NbSs can
management.
provide a low-impact green development
strategy for transforming the agricultural Public investment will be essential to develop
sector into a beneficiary and a custodian of capacities across producer associations
ecosystems. and applied research institutions. A well-
regulated land and water governance
framework that can promote the adoption
of innovative management and technology
with targeted financing for impact is a real-
izable goal.

Land and water management can respond


to meet the challenges of climate change
based on investment that is compliant
© Rosetta Messori

with environmental, social and governance


approaches through actions that are better
planned, informed and implemented, but
above all, better governed.

THE STATE OF THE WORLD’S LAND AND WATER RESOURCES FOR FOOD AND AGRICULTURE
AGRICULTURE: 2021
SOLAW 2021 351
This report has taken a global view of the available land and water data derived from
national statistical sources where measures of land and water resource use have been
reported up to 2019. It is not prescriptive about specific regions or entry points. Rather,
the report recommends a set of mutually reinforcing responses to address the critical
issues of human-induced land degradation and water scarcity. Observations on the state
of land and water systems give rise to policy recommendations generally applicable in
agriculture programmes. These recommendations reflect the key messages developed for
each chapter.

352 6. CONCLUSIONS
5. RESPONSES
AND RECOMMENDATIONS
AND ACTION AREAS
Annex:
Country groupings

Continent Subregion Countries


Regions

Africa Algeria, Angola, Benin, Botswana, Burkina Faso, Burundi, Cameroon,


Cabo Verde, Central African Republic, Chad, Comoros, Congo,
Côte d’Ivoire, Democratic Republic of the Congo, Djibouti, Egypt,
Equatorial Guinea, Eritrea, Eswatini, Ethiopia, Gabon, Gambia, Ghana,
Guinea, Guinea-Bissau, Kenya, Lesotho, Liberia, Libya, Madagascar,
Malawi, Mali, Mauritania, Mauritius, Morocco, Mozambique, Namibia,
Niger, Nigeria, Rwanda, Sao Tome and Principe, Senegal, Seychelles,
Sierra Leone, Somalia, South Africa, Sudan, Togo, Tunisia, Uganda,
United Republic of Tanzania, Zambia, Zimbabwe

Northern Algeria, Egypt, Libya, Morocco, Tunisia


Africa

Sub-Saharan Angola, Benin, Botswana, Burkina Faso, Burundi, Cameroon, Cabo


Africa Verde, Central African Republic, Chad, Comoros, Congo, Côte
d’Ivoire, Democratic Republic of the Congo, Djibouti, Equatorial
Guinea, Eritrea, Eswatini, Ethiopia, Gabon, Gambia, Ghana, Guinea,
Guinea-Bissau, Kenya, Lesotho, Liberia, Madagascar, Malawi, Mali,
Mauritania, Mauritius, Mozambique, Namibia, Niger, Nigeria, Rwanda,
Sao Tome and Principe, Senegal, Seychelles, Sierra Leone, Somalia,
South Africa, Sudan, Togo, Uganda, United Republic of Tanzania,
Zambia, Zimbabwe

Sudano-Sahelian Burkina Faso, Cabo Verde, Chad, Djibouti, Eritrea, Gambia, Mali,
Mauritania, Niger, Senegal, Somalia, Sudan

Gulf of Guinea Benin, Côte d’Ivoire, Ghana, Guinea, Guinea-Bissau, Liberia, Nigeria,
Sierra Leone, Togo

Central Africa Angola, Cameroon, Central African Republic, Congo, Democratic


Republic of the Congo, Equatorial Guinea, Gabon, Sao Tome and
Principe

Eastern Africa Burundi, Ethiopia, Kenya, Rwanda, Uganda, United Republic of


Tanzania

Southern Africa Botswana, Eswatini, Lesotho, Malawi, Mozambique, Namibia, South


Africa, Zambia, Zimbabwe

Indian Ocean Comoros, Madagascar, Mauritius, Seychelles


Islands

353
Continent Subregion Countries
Regions

Americas Antigua and Barbuda, Argentina, Bahamas, Barbados, Belize, Bolivia


(Plurinational State of), Brazil, Canada, Chile, Colombia, Costa Rica,
Cuba, Dominica, Dominican Republic, Ecuador, El Salvador, French
Guiana, Grenada, Guatemala, Guyana, Haiti, Honduras, Jamaica,
Mexico, Nicaragua, Panama, Paraguay, Peru, Puerto Rico, Saint Kitts
and Nevis, Saint Lucia, Saint Vincent and the Grenadines, Suriname,
Trinidad and Tobago, United States of America, Uruguay, Venezuela
(Bolivarian Republic of)
Northern Canada, Mexico, United States of America
America
Northern Canada, United States of America
America
Mexico Mexico
Central Antigua and Barbuda, Bahamas, Barbados, Belize, Costa Rica, Cuba,
America and Dominica, Dominican Republic, El Salvador, Grenada, Guatemala,
Caribbean Haiti, Honduras, Jamaica, Nicaragua, Panama, Puerto Rico, Saint
Kitts and Nevis, Saint Lucia, Saint Vincent and the Grenadines,
Trinidad and Tobago
Central America Belize, Costa Rica, El Salvador, Guatemala, Honduras, Nicaragua,
Panama
Greater Antilles Cuba, Dominican Republic, Haiti, Jamaica, Puerto Rico
Lesser Antilles Antigua and Barbuda, Bahamas, Barbados, Dominica, Grenada,
and Bahamas Saint Kitts and Nevis, Saint Lucia, Saint Vincent and the Grenadines,
Trinidad and Tobago
Southern Argentina, Bolivia (Plurinational State of), Brazil, Chile, Colombia,
America Ecuador, French Guiana, Guyana, Paraguay, Peru, Suriname,
Uruguay, Venezuela (Bolivarian Republic of)
Guyana French Guiana, Guyana, Suriname
Andean Bolivia (Plurinational State of), Colombia, Ecuador, Peru, Venezuela
(Bolivarian Republic of)
Brazil Brazil
Southern Argentina, Chile, Paraguay, Uruguay
America
Asia Afghanistan, Armenia, Azerbaijan, Bahrain, Bangladesh, Bhutan,
Brunei Darussalam, Cambodia, China, Democratic People’s Republic
of Korea, Georgia, India, Indonesia, Iran (Islamic Republic of), Iraq,
Israel, Japan, Jordan, Kazakhstan, Kuwait, Kyrgyzstan, Lao People’s
Democratic Republic, Lebanon, Malaysia, Maldives, Mongolia,
Myanmar, Nepal, Occupied Palestinian Territory, Oman, Pakistan,
Papua New Guinea, Philippines, Qatar, Republic of Korea, Saudi
Arabia, Singapore, Sri Lanka, Syrian Arab Republic, Tajikistan,
Thailand, Timor-Leste, Turkey, Turkmenistan, United Arab Emirates,
Uzbekistan, Viet Nam, Yemen
Middle East– Armenia, Azerbaijan, Bahrain, Georgia, Iran (Islamic Republic of),
Western Asia Iraq, Israel, Jordan, Kuwait, Lebanon, Occupied Palestinian Territory,
Oman, Qatar, Saudi Arabia, Syrian Arab Republic, Turkey, United Arab
Emirates, Yemen
Arabian Bahrain, Kuwait, Oman, Qatar, Saudi Arabia, United Arab Emirates,
Peninsula Yemen
Caucasus Armenia, Azerbaijan, Georgia
Islamic Republic Iran (Islamic Republic of)
of Iran
Near East Iraq, Israel, Jordan, Lebanon, Occupied Palestinian Territory, Syrian
Arab Republic, Turkey

354
Continent Subregion Countries
Regions

Central Asia Afghanistan, Kazakhstan, Kyrgyzstan, Tajikistan, Turkmenistan,


Uzbekistan

Southern and Bangladesh, Bhutan, Brunei Darussalam, Cambodia, China,


Eastern Asia Democratic People’s Republic of Korea, India, Indonesia, Japan,
Lao People’s Democratic Republic, Malaysia, Maldives, Mongolia,
Myanmar, Nepal, Pakistan, Papua New Guinea, Philippines, Republic
of Korea, Singapore, Sri Lanka, Thailand, Timor-Leste, Viet Nam
South Asia Bangladesh, Bhutan, India, Maldives, Nepal, Pakistan, Sri Lanka
East Asia China, Democratic People’s Republic of Korea, Japan, Mongolia,
Republic of Korea
Southeast Asia Brunei Darussalam, Cambodia, Indonesia, Lao People’s Democratic
Republic, Malaysia, Myanmar, Papua New Guinea, Philippines,
Singapore, Thailand, Timor-Leste, Viet Nam

Europe Albania, Andorra, Austria, Belarus, Belgium, Bosnia and Herzegovina,


Bulgaria, Croatia, Cyprus, Czechia, Denmark, Estonia, Faroe
Islands, Finland, France, Germany, Greece, Holy See, Hungary,
Iceland, Ireland, Italy, Latvia, Liechtenstein, Lithuania, Luxembourg,
Malta, Monaco, Montenegro, Netherlands, North Macedonia,
Norway, Poland, Portugal, Republic of Moldova, Romania, Russian
Federation, San Marino, Serbia, Slovakia, Slovenia, Spain, Sweden,
Switzerland, Ukraine, United Kingdom of Great Britain and Northern
Ireland

Western and Albania, Andorra, Austria, Belgium, Bosnia and Herzegovina,


Central Europe Bulgaria, Croatia, Cyprus, Czechia, Denmark, Faroe Islands, Finland,
France, Germany, Greece, Holy See, Hungary, Iceland, Ireland,
Italy, Liechtenstein, Luxembourg, Malta, Monaco, Montenegro,
Netherlands, North Macedonia, Norway, Poland, Portugal, Romania,
San Marino, Serbia, Slovakia, Slovenia, Spain, Sweden, Switzerland,
United Kingdom
Northern Europe Denmark, Faroe Islands, Finland, Iceland, Norway, Sweden
Western Europe Andorra, Austria, Belgium, France, Germany, Ireland, Liechtenstein,
Luxembourg, Netherlands, Switzerland, United Kingdom
Central Europe Bosnia and Herzegovina, Bulgaria, Croatia, Czechia, Hungary,
Montenegro, Poland, Romania, Serbia, Slovakia, Slovenia
Mediterranean Albania, Cyprus, Greece, Holy See, Italy, Malta, Monaco, North
Europe Macedonia, Portugal, San Marino, Spain

Eastern Europe Belarus, Estonia, Latvia, Lithuania, Republic of Moldova, Russian


Federation, Ukraine

Eastern Europe Belarus, Estonia, Latvia, Lithuania, Republic of Moldova, Ukraine

Russian Russian Federation


Federation

Oceania Australia, Cook Islands, Fiji, Kiribati, Micronesia (Federated States of),
Nauru, New Zealand, Niue, Palau, Samoa, Solomon Islands, Tonga,
Tuvalu, Vanuatu

Australia and Australia, New Zealand


New Zealand

Pacific Islands Cook Islands, Fiji, Kiribati, Micronesia (Federated States of), Nauru,
Niue, Palau, Samoa, Solomon Islands, Tonga, Tuvalu, Vanuatu

355
Continent Subregion Countries
Regions

World Afghanistan, Albania, Algeria, Andorra, Angola, Antigua and Barbuda,


Argentina, Armenia, Australia, Austria, Azerbaijan, Bahamas,
Bahrain, Bangladesh, Barbados, Belarus, Belgium, Belize, Benin,
Bhutan, Bolivia (Plurinational State of), Bosnia and Herzegovina,
Botswana, Brazil, Brunei Darussalam, Bulgaria, Burkina Faso,
Burundi, Cambodia, Cameroon, Canada, Cabo Verde, Central African
Republic, Chad, Chile, China, Colombia, Comoros, Congo, Cook
Islands, Costa Rica, Côte d’Ivoire, Croatia, Cuba, Cyprus, Czechia,
Democratic People’s Republic of Korea, Democratic Republic of
the Congo, Denmark, Djibouti, Dominica, Dominican Republic,
Ecuador, Egypt, El Salvador, Equatorial Guinea, Eritrea, Estonia,
Eswatini, Ethiopia, Faroe Islands, Fiji, Finland, France, French Guiana,
Gabon, Gambia, Georgia, Germany, Ghana, Greece, Grenada,
Guatemala, Guinea, Guinea-Bissau, Guyana, Haiti, Holy See,
Honduras, Hungary, Iceland, India, Indonesia, Iran (Islamic Republic
of), Iraq, Ireland, Israel, Italy, Jamaica, Japan, Jordan, Kazakhstan,
Kenya, Kiribati, Kuwait, Kyrgyzstan, Lao People’s Democratic
Republic, Latvia, Lebanon, Lesotho, Liberia, Libya, Liechtenstein,
Lithuania, Luxembourg, Madagascar, Malawi, Malaysia, Maldives,
Mali, Malta, Mauritania, Mauritius, Mexico, Micronesia (Federated
States of), Monaco, Mongolia, Montenegro, Morocco, Mozambique,
Myanmar, Namibia, Nauru, Nepal, Netherlands, New Zealand,
Nicaragua, Niger, Nigeria, Niue, North Macedonia, Norway, Occupied
Palestinian Territory, Oman, Pakistan, Palau, Panama, Papua New
Guinea, Paraguay, Peru, Philippines, Poland, Portugal, Puerto Rico,
Qatar, Republic of Korea, Republic of Moldova, Romania, Russian
Federation, Rwanda, Saint Kitts and Nevis, Saint Lucia, Saint Vincent
and the Grenadines, Samoa, San Marino, Sao Tome and Principe,
Saudi Arabia, Senegal, Serbia, Seychelles, Sierra Leone, Singapore,
Slovakia, Slovenia, Solomon Islands, Somalia, South Africa, Spain,
Sri Lanka, Sudan, Suriname, Sweden, Switzerland, Syrian Arab
Republic, Tajikistan, Thailand, Timor-Leste, Togo, Tonga, Trinidad
and Tobago, Tunisia, Turkey, Turkmenistan, Tuvalu, Uganda, Ukraine,
United Arab Emirates, United Kingdom, United Republic of Tanzania,
United States of America, Uruguay, Uzbekistan, Vanuatu, Venezuela
(Bolivarian Republic of), Viet Nam, Yemen, Zambia, Zimbabwe

356
Glossary

Agricultural land: Agricultural area as the sum of areas under: (a) arable land, (b) permanent
crops (land cultivated with long-term crops that do not have to be replanted for several years)
and (c) permanent meadows and pastures.

Agroforestry: Land-use systems or practices in which trees are deliberately integrated with
crops and/or animals on the same land management unit.

Arable land: Land under temporary agricultural crops, temporary meadows for mowing or pas-
ture, market and kitchen gardens, and land temporarily fallow (less than five years). The aban-
doned land resulting from shifting cultivation is not included in this category. Data for “arable
land” are not meant to indicate the amount of land that is potentially cultivable.

Biodiversity: The variability among living organisms from all sources including, inter alia,
terrestrial, marine and other aquatic ecosystems and the ecological complexes of which they are
part; this includes diversity within species, between species and of ecosystems.

Carbon sequestration: The process of removing carbon from the atmosphere and depositing it
in reservoirs such as oceans, forests or soils through physical or biological processes.

Conjunctive use (of surface water and groundwater): The integrated management and use of
surface water and groundwater supplies.

Conservation agriculture: An approach to managing agroecosystems for improved and sus-


tained productivity, increased profits and food security, while preserving and enhancing the
resource base and the environment. Conservation agriculture is characterized by three prin-
ciples: continuous minimum mechanical soil disturbance, permanent organic soil cover and
diversification of crop species grown in sequences or associations.

Conservation tillage: An approach to soil management that excludes conventional tillage op-
erations that invert the soil and bury crop residues. There are five types of conservation tillage
systems: no-tillage (slot planting), mulch tillage, strip or zonal tillage, ridge tillage (including
no-till on ridges) and reduced or minimum tillage.

Consumptive use of water: The part of water withdrawn from its source for use in agriculture,
industry or domestic purposes that has evaporated, transpired or been incorporated into prod-
ucts. The part of water withdrawn that is not consumed is called return flow.

357
Contaminant: Any substance not intentionally added to food, which is present in such food as
a result of production (including operations carried out in crop and animal husbandry), manu-
facture, processing, preparation, treatment, packing, packaging, transport or holding of such
food or as a result of environmental contamination. The term includes chemical and biological
substances not desirable in food but does not include insect fragments, rodent hairs and other
extraneous matter.

Cropland (or cultivated land): The land that is under agricultural crops. In statistical terms,
cropland is the sum of arable land (see definition above) and permanent crops.

Desertification: The degradation of land in arid, semi-arid and dry subhumid areas resulting
from various factors, including climatic variations and human activities.

Dry lands: Arid, semi-arid and dry subhumid areas (other than polar and subpolar regions)
in which the ratio of mean annual precipitation to mean annual reference evapotranspiration
ranges from 0.05 to 0.65.

Ecosystem: A dynamic complex of plant, animal and microorganism communities, and the
non-living physical components of the environment (e.g. air, soil, water and sunlight), inter-
acting as a functional unit.

Ecosystem services (or environmental services): The benefits people obtain from ecosystems.
These include provisioning services (e.g. food and water), regulating services, supporting ser-
vices (e.g. soil formation and nutrient cycling) and cultural services (e.g. recreational, spiritual,
religious and other non-material benefits).

Erosion: The wearing away of the land by running water, rainfall, wind, ice or other geological
agents, including such processes as detachment, entrainment, suspension, transportation and
mass movement.

Eutrophication: The enrichment of freshwater bodies by inorganic nutrients (e.g. nitrates and
phosphates), typically leading to excessive growth of algae.

Evapotranspiration: The combination of evaporation from the soil surface and transpiration
from the plants.

Fertilizer: A substance that is used to provide nutrients to plants, usually via application to the
soil, but also to foliage or through water in rice systems, fertigation, hydroponics or aquaculture
operations.

Freshwater: Naturally occurring water on the Earth’s surface in glaciers, lakes and rivers, and
underground in aquifers. Its key feature is a low concentration of dissolved salts. It excludes
rainwater, water stored in the soil, untreated wastewater, seawater and brackish water. In this
report, when not otherwise specified, the term water is used as a synonym of freshwater.

Groundwater: All water which is below the surface of the ground in the saturation zone and in
direct contact with the subsoils.

358
Institution: The laws and regulations governing the management, development, protection
from pollution and use of water resources; the governmental bodies at all levels, in charge of
the administration and enforcement of the laws and regulations; the judiciary; and the formal
or informal water user-level organizations.

Integrated pest management: An ecosystem approach to crop production and protection that
combines different management strategies and practices to grow healthy crops while mini-
mizing the use of pesticides.

Integrated water resources management: A process that promotes the coordinated devel-
opment and management of water, land and related resources to maximize the resultant
economic and social welfare in an equitable manner without compromising the sustainability
of vital ecosystems. Sustainable Development Goal target 6.5 measures the degree and imple-
mentation of integrated water resources management.

Internal renewable water resources: The conventional measure of freshwater available to


a nation (surface water and groundwater), comprising resources deriving from the rainfall
within a nation’s boundaries. It excludes transboundary and fossil water resources.

Land degradation: The reduction in the capacity of the land to provide ecosystem goods and
services over a period of time for its beneficiaries.

Land degradation neutrality: A state whereby the amount and quality of land resources nec-
essary to support ecosystem functions and services to enhance food security remain stable, or
increase, within specified temporal and spatial scales and ecosystems.

Modernization: In irrigation, modernization is a process of technical and managerial upgrad-


ing (as opposed to mere rehabilitation) of irrigation schemes combined with institutional
reforms, if required, with the objective to improve resource utilization (e.g. labour, water
economics and environment) and water delivery service to farms.

Nutrient imbalance: An excess or lack of nutrients (mainly nitrogen, phosphorus and potas-
sium) in the soil as a consequence of bad land use and management. It may result in soil
contamination when nutrients are in excess and in loss of inherent fertility when nutrients
are mined.

Payment for ecosystem services: A voluntary transaction whereby a service provider is paid by
(or on behalf of) beneficiaries for land-use practices that are expected to result in continued
or improved environmental service provision beyond what would have been provided without
the payment.

Rangeland: Land on which the indigenous vegetation (climax or subclimax) is predominantly


grasses, grass-like plants, forbs or shrubs that are grazed or have the potential to be grazed,
and which is used as a natural ecosystem for the production of grazing livestock and wildlife.

359
Return flow: The part of the water withdrawn from its source that is not consumed and returns
to its source or to another body of groundwater or surface water. Return flow can be divided
into non-recoverable flow (flow to salt sinks, uneconomic groundwater or flow of insufficient
quality) and recoverable flow (flow to rivers or infiltration into groundwater aquifers).

Riparian: Relating to land adjoining a stream or river.

Runoff: Part of the water from precipitation, melted snow or irrigation that flows over the land
surface in stream flow and is not absorbed into the ground.

Salinization: The process by which salt accumulates in or on the soil. Human-induced saliniza-
tion is mostly associated with poor irrigation practices.

Sodic soil: A soil that contains sufficient sodium to adversely affect the growth of most crop
plants (sodic soils are defined as those soils which have an exchangeable sodium percentage of
more than 15).

Soil acidification: The lowering of the soil pH of the build-up of hydrogen and aluminium ions
in the soil and the leaching of base cations such as calcium, magnesium, potassium and sodium.
Soil acidification negatively affects soil fertility and compromises the production capacity of
most agricultural soils.

Soil biodiversity loss: The decline in the diversity of (micro- and macro-) organisms present in
a soil. It prejudices the ability of soil to provide critical ecosystem services.

Soil compaction: The increase in density and a decline of macro-porosity in a soil that impairs
the functions of both the top- and subsoil, and impedes roots penetration and water gaseous
exchanges.

Soil degradation: The decline in soil quality caused by its improper use by humans, usually for
agricultural, pastoral, industrial or urban purposes. Soil degradation may be exacerbated by
climate change and encompasses physical, chemical and biological deterioration.

Soil health: The capacity of soil to function as a living system. Healthy soils maintain a diverse
community of soil organisms that help to control plant disease, insect and weed pests, form
beneficial symbiotic associations with plant roots, recycle essential plant nutrients, improve
soil structure with positive repercussions for soil water and nutrient-holding capacity, and
ultimately improve crop production.

Soil organic carbon loss: The decline of organic carbon stock in the soil affecting its fertility
status and climate change regulation capacity.

Soil pollution: The presence of a chemical or substance out of place and/or present at higher
than normal concentration that has adverse effects on non-target organisms.

Soil salinization: The increase in water-soluble salts in soil which is responsible for increasing
the osmotic pressure of the soil. This negatively affects plant growth because less water is made
available to plants.

360
Sustainable development: The development that meets the needs of the present without com-
promising the ability of future generations to meet their own needs. Sustainability is a para-
digm for thinking about the future in which environmental, societal and economic consider-
ations are balanced in the pursuit of an improved quality of life.

Wadi: The ravine or valley of a seasonal stream in arid or semi-arid areas that is usually dry
except for a short time after spate flow events (a few hours to a few days).

Water accounting: A systematic acquisition, analysis and communication of information relat-


ing to stocks, flows and fluxes of water (from sources to sinks) in natural, disturbed or heavily
engineered environments.

Water auditing: A systematic study of the current status and future trends in water supply and
demand, with a particular focus on issues relating to accessibility, uncertainty and governance
in each spatial domain.

Water governance: The processes, actors and institutions involved in decision-making for
the development and management of water resources and for the delivery of water services,
encompassing the political, administrative, social and economic domains along with the formal
and informal systems and mechanisms involved.

Water harvesting: The process of collecting and concentrating runoff water from a runoff area
into a run-on area, where the collected water is either directly applied to the cropping area
and stored in the soil profile for immediate use by the crop (i.e. runoff farming) or stored in an
on-farm water reservoir for future productive uses (i.e. domestic use, livestock watering and
aquaculture irrigation).

Waterlogging: The state of land in which the water table is located at or near the soil surface,
affecting crop yields.

Water productivity: The ratio between the amount or value of output (including services)
provided by water, in relation to the volume of water used to produce the output. Crop water
productivity refers to the ratio between crop yield and water supply. Economic water productiv-
ity is expressed as the ratio between added value of a product and water supply.

Water scarcity: An imbalance between supply and demand of freshwater in a specified domain
(e.g. country, region, catchment or river basin) as a result of a high rate of demand compared
with available supply, under prevailing institutional arrangements (including price) and infra-
structural conditions. Symptoms are unsatisfied demand, tensions between users, competition
for water, overextraction of groundwater and insufficient flows to the natural environment.
Artificial or constructed water scarcity refers to the situation resulting from overdevelopment
of hydraulic infrastructure relative to available supply, leading to a situation of increasing
water shortage.

361
Water scarcity (absolute): An insufficiency of supply to satisfy total demand after all feasible
options to enhance supply and manage demand have been implemented. It is measured as the
level at which all freshwater resources available is less than 500 m3/capita annually.

Water scarcity (chronic): The level at which all freshwater resources available are being used.
Beyond this level, water supply can be made available only through the use of non-conventional
water resources such as agricultural drainage water, treated wastewater or desalinated water, or
by managing demand. The level at which all freshwater resources available ranges between 500
and 1 000 m3/capita annually.

Water security: The reliable availability of an acceptable quantity and quality of water for
health, livelihoods and production, coupled with an acceptable level of water-related risks,
while ensuring that the environment is protected and enhanced.

Water shortage: A shortage of water supply of an acceptable quality and/or low levels of water
supply at a given place and a given time, relative to design supply levels. This may arise from
climatic factors, or other causes of insufficient water resources, such as a lack of, or poorly
maintained, infrastructure, or a range of other hydrological or hydrogeological factors.

Water stress: The symptoms of water scarcity or shortage, such as widespread, frequent
and serious restrictions on use, growing conflict between users and competition for water,
declining standards of reliability and service, harvest failures and food insecurity. Sustainable
Development Goal indicator 6.4.1 measures water stress as the ratio of freshwater withdrawals
by all major sectors to the available freshwater resources after taking into account environmental
water requirements.

Water tenure: The relationship, whether legally or customarily defined, between people, as
individuals or groups, with respect to water resources.

Water-use efficiency: The ratio of the amount of water used for a specific purpose to the
amount of water withdrawn or diverted from its source to serve that use. In irrigation,
water-use efficiency presents the ratio between estimated irrigation water requirements
(through evapotranspiration) and actual water withdrawal. It is dimensionless and can be
applied at any scale (plant, field, irrigation schemes, basin, country). Sustainable Development
Goal indicator 6.4.2 measures water-use efficiency as the ratio of the gross value added per unit
of water used (in USD/m3).

Water-use right: In its legal sense, a legal right to abstract or divert and use water from a given
natural source; to impound or store a specified quantity of water in a natural source behind a
dam or other hydraulic structure; or to use or maintain water in a natural state (ecological flow
in a river, and water for recreation, religious/spiritual practices, drinking, washing, bathing or
animal watering).

Water withdrawal: Water abstracted from streams, aquifers or lakes for any purpose (e.g.
irrigation, industrial, domestic or commercial). It includes conveyance losses, consumptive
use and return flow. It can include water from renewable freshwater resources as well as water
from overabstraction of renewable groundwater or withdrawal from fossil groundwater, direct
use of agricultural drainage water, direct use of (treated) wastewater and desalinated water.

362
THE STATE OF THE
WORLD’S LAND AND
WATER RESOURCES FOR FOOD
AND AGRICULTURE 2021
Systems at breaking point

Satisfying the increased demand for food is placing pressure on


the world’s land, soil and water resources. Agriculture has its part
to play in alleviating these pressures and contributing positively
to climate and development goals. Sustainable agricultural
practices can lead to direct improvements in the state of land,
soil and water, and generate ecosystem benefits as well as
reduce emissions from land. Accomplishing all these requires
accurate information and a major change in how we manage the
resources. It also requires complementing efforts from outside
the natural resources management domain to maximize
synergies and manage trade-offs.

The objective of The state of the world’s land and water resources
for food and agriculture 2021 (SOLAW 2021) report is to build
awareness of the status of land and water resources, highlighting
the risks, and informing on related opportunities and challenges. It also
aims to underline the essential contribution of appropriate policies,
institutions and investments. Recent assessments, projections and
scenarios point to the accelerated depletion of land and water resources
and associated loss of biodiversity. The SOLAW 2021 report highlights the
major risks and trends related to land, soil and water resources, and
presents the means for resolving competition among users and
generating the desirable benefits. The report provides an update of the
knowledge base and presents a suite of responses and actions to enable
decision-makers to make an informed transformation from degradation
and vulnerability towards sustainability and resilience.

#SOLAW2021 ISBN 978-92-5-136127-6

9 789251 361276
CB9910EN/1/05.22

You might also like