Вы находитесь на странице: 1из 239

Water Security in a New World

Kamel Amer
Zafar Adeel
Benno Böer
Walid Saleh Editors

The Water, Energy,


and Food Security
Nexus in the Arab
Region
Water Security in a New World

Series editor
Zafar Adeel, Pacific Water Research Centre, Simon Fraser University, Burnaby,
BC, Canada
More information about this series at http://www.springer.com/series/13406
Kamel Amer Zafar Adeel

Benno Böer Walid Saleh


Editors

The Water, Energy,


and Food Security Nexus
in the Arab Region

123
Editors
Kamel Amer Benno Böer
Private Consultant UNESCO Liaison Office in Addis Abeba
Burlington, ON with the African Union and UNECA
Canada UNESCO
Addis Abeba
Zafar Adeel Ethiopia
Pacific Water Research Centre,
Faculty of Environment Walid Saleh
Simon Fraser University Food and Agriculture Organization of the
Burnaby, British Columbia United Nations
Canada Sana’a
Yemen

ISSN 2367-4008 ISSN 2367-4016 (electronic)


Water Security in a New World
ISBN 978-3-319-48407-5 ISBN 978-3-319-48408-2 (eBook)
DOI 10.1007/978-3-319-48408-2
Library of Congress Control Number: 2016955528

© Springer International Publishing AG 2017


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part
of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations,
recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission
or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar
methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specific statement, that such names are exempt from
the relevant protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publisher nor the
authors or the editors give a warranty, express or implied, with respect to the material contained herein or
for any errors or omissions that may have been made.

Printed on acid-free paper

This Springer imprint is published by Springer Nature


The registered company is Springer International Publishing AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Foreword

Water and energy are essential for building regional security and for securing a
sustainable economy that is focused on human well-being. However, the water–
energy nexus is complex and dynamic, and is subjected to rapid and unanticipated
demographic and natural changes. Our overexploited natural resources are highly
affected by external factors such as climate change, population growth, and eco-
nomic and trade fluctuations. Hence, identifying the nature of these changes is a
crucial step towards engineering appropriate tools to help us develop sustainable
resource management strategies based on a comprehensive understanding of this
water–energy nexus.
What is needed now is to attract voices from the region which are able to seek
more integrated approaches that will address the challenges and opportunities of
this nexus. These approaches should aim to foster integrated decision-making on
water and energy infrastructure systems and to create multiple scenarios and
solutions to tackle our water and energy needs on a larger scale. The need to
enhance the reliability and resilience of energy and water systems is crucial to the
promotion of responsible energy operations with respect to water quality, ecosys-
tems, and human dignity and well-being.
In addition to developing a regional water use efficiency programme through
technology and research, our region needs a shared strategy both within countries
and across borders that focuses on human–social–environmental concerns. We need
a strategy whose cooperation and contribution towards a safe and a secure region is
not limited by geographic borders. We need treaties that take into consideration the
demographic changes of the region and put “human dignity and equality” ahead of
“water storage capacity.”
The challenges and opportunities presented by climate change and water scarcity
in conflict zones in our region must be jointly addressed. As Tom Friedman has
noted, 45 million people are threatened in the Nile Delta within the next twenty
years by the rise in water levels of the Mediterranean. Recently, Iran warned that by
continuing to exploit 97 % of Iran’s surface water, approximately 55 million
people, 70 % of Iranians, will have no choice but to leave the country.

v
vi Foreword

We must also understand and respect our region’s carrying capacity—its capa-
bility to efficiently and equitably accommodate the needs and aspirations of its
population to the prerequisite standard of human dignity and security. Our under-
standing of carrying capacity must integrate the human, natural, and social
resources of the region without prejudice to the fundamental importance of balance,
vis-à-vis the natural ecosystems within which any notion of carrying capacity
necessarily operates. The region needs to formulate resilient strategies to maximize
the efficiency of our energy resources and to develop and invest in new, innovative
forms of renewable energy, such as smart and solar grids.
There is a growing need to understand the carrying capacity and capability of the
region since it denotes the capability of a state or region to efficiently and equitably
accommodate the needs and aspirations of its population to the prerequisite stan-
dard of human security and dignity.
This book provides a much-needed regional-level quantitative assessment of this
water–energy nexus in the West Asia North Africa (WANA) region. It emphasizes
the high dependence of energy systems on water resources and concludes that this
dependence on water must be addressed by the implementation and utilization of
new technologies that will lead us towards water and food security.
This book also highlights the necessity of employing integrated solutions to
address these challenges through the strengthening of research and development
efforts in order to lay the foundation for a more secure future for the region. I am
confident that it will be an indispensable reference for many.
Let us work together as a region towards securing water and energy resilience for
human prosperity.

H.R.H. Prince El Hassan bin Talal


Amman, the Hashemite Kingdom of Jordan
Preface

The water crisis is one of the most critical challenges facing the world today, and it
relates directly to food and energy security. This nexus of security challenges is
most prominent for the Arab Region and will continue to remain as such unless
significant rethinking is brought to bear on policies and actions. In this respect,
scientific inquiry, engineering innovations, and targeted education are important
ingredients to enhance the water, food, and energy security on a regional basis.
The necessity for better management of water resources increased agricultural
productivity and water-use efficiency, reduced dependency of food imports, and a
stronger outlook for diversified energy resources is critical. The management
of these interlinked resources also has to tie in with effective management of natural
and man-made ecosystems in the diverse Arab Region, which is characterized by
adverse environmental and climatical conditions.
The situation may worsen if one considers the trends in climate change,
increasing population pressure, and ongoing geopolitical conflicts. For example,
many have cited the origins of the “Arab Spring” in Tunisia in 2011 as tied to the
non-availability of basic needs, such as jobs and income, and families’ access to
clean drinking water, sufficient food supply, and energy.
Even though some of the Arab countries are among the world’s largest producers
of fossil fuels, factors such as population pressure, increasing internal per capita
consumption of oil and gas for power generation, transportation, desalination of sea
water, and fluctuating global fuel prices have made them reconsider their policies.
As a result, some of the countries are opting to turn to the production of nuclear
fuel, which comes with a host of related risks. There is also a growing recognition
that long-term sustainability and economic growth requires greater attention to
further development of environmentally friendly and renewable energy resources in
the Arab Region—including energy generation from solar, wind, thermal, tidal
energy, and biofuel options.

vii
viii Preface

The United Nations system, together with its constituent organizations and
agencies, has to play a pivotal role in facilitating change and enhancements around
the water, food, and energy security nexus in the Arab Region. For example, the
Food and Agricultural Organization (FAO), the World Food Programme (WFP),
the International Fund for Agricultural Development (IFAD), and other agencies
affiliated with the UN system are directly involved in the management of food
production and supply. UN-Water and UN-Energy are the two key coordination
mechanisms that enhance cooperative efforts of a multitude of players within the
UN system.
UNESCO has been involved in a range of water issues for 50 years, through its
International Hydrological Programme, devoted to water research, water resources
management, education, and capacity building. More recently, it has led the World
Water Assessment Programme (WWAP), which coordinates the publication of the
UN World Water Development Report. UNESCO also manages a global network
of accredited sites for nature conservation and sustainable development:
UNESCO’s World Network of Biosphere Reserves (WNBR). It promotes nature
conservation in reconciliation with sustainable development. Overall, UNESCO
plays an essential role to assist the quest for sustainable living based on the supply
of water, food, and energy.
UNU-INWEH—as the United Nations think tank on water—serves as the bridge
between the scientific community and the policy audiences to address major global
water challenges. It has focused on water security as the primary entry point to food
and energy security. A number of its recent publications have discussed the nexus
of the challenges faced in the Arab Region; the most notable of these are the
following books: “Water, Energy, and the Arab Awakening (2014)” and “Policy
Perspectives for Ecosystem and Water Management in the Arabian Peninsula
(2006).”
This volume, jointly published by UNU-INWEH and UNESCO, aims to provide
political decision-makers, regional and national authorities, food security, water
security, energy, science, education, and environmental stakeholders with
cutting-edge knowledge about the current situation as well as innovative ideas to
build the needed defences against water, food, and energy insecurities. By devel-
oping principles for efficient and ethical water security, food security, and energy
management, we move closer to sustainable development.
This volume draws heavily on interdisciplinary expertise to formulate guidance
that could help balance the hydrological, ecological, energy, and socio-economic
needs of the region, which comprises Algeria, Bahrain, Egypt, Iraq, Jordan, Kuwait,
Lebanon, Libya, Mauretania, Morocco, Oman, Palestine, Qatar, Saudi Arabia,
Sudan, Tunisia, Syria, the United Arab Emirates, and Yemen.
Preface ix

We commend the contributors to this volume, who have enabled a compre-


hensive contribution towards understanding the issues that underlie food security,
water security, and energy nexus in the Arab Region. We anticipate that the find-
ings presented in this volume will inform enhancement of management and policy
responses in the region.

Paris, France
Hamilton, Canada
June 2016

Flavia Schlegel
UNESCO Assistant Director-General
for Natural Sciences

Zafar Adeel
United Nations University
Institute for Water, Environment and Health
Director
Contents

1 Status of Water in the Arab Region . . . . . . . . . . . . . . . . . . . . . . . . . 1


Waleed Khalil Al-Zubari
2 State of Energy in the Arab Region . . . . . . . . . . . . . . . . . . . . . . . . . . 25
Kishan Khoday and Stephen Gitonga
3 Status of Food Security in the Arab Region . . . . . . . . . . . . . . . . . . . 43
Tariq Moosa Alzadjali
4 Desalinated Water for Food Production in the Arab Region . . . . . . 59
Hani Sewilam and Peter Nasr
5 Water, Energy, and Food in the Arab Region: Challenges
and Opportunities, with Special Emphasis on Renewable
Energy in Food Production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
John M. Bryden
6 Water, Energy, and Food Security in the Arab Region:
Regional Cooperation and Capacity Building . . . . . . . . . . . . . . . . . . 105
Atef Hamdy
7 Research and Development to Bridge the Knowledge Gap . . . . . . . 123
Khaled AbuZeid
8 Water-Energy-Food Security Nexus in the Arab Region:
Thoughts and Policy Options . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
Paul Sullivan
9 Managing Water, Energy, and Food for Long-Term Regional
Security . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
Zafar Adeel
10 Current Water for Food Situational Analysis in the Arab
Region and Expected Changes Due to Dynamic Externalities . . . . . 193
Rabi H. Mohtar, Amjad T. Assi and Bassel T. Daher

xi
xii Contents

11 Case Study: Masdar Renewable Energy Water Desalination


Program . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
Mohammad El Ramahi
12 Summarizing the Story . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
Benno Böer and Zafar Adeel
Contributors

Khaled AbuZeid Centre for Environment and Development for the Arab Region
and Europe, Heliopolis, Cairo, Egypt
Zafar Adeel Pacific Water Research Centre, Faculty of Environment, Simon
Fraser University, Burnaby, British Columbia, Canada
Waleed Khalil Al-Zubari Water Resources Management Program, College of
Graduate Studies, Arabian Gulf University, Manama, Bahrain
Tariq Moosa Alzadjali Arab Organization for Agricultural Development,
Khartoum, Sudan
Amjad T. Assi Department of Biological and Agricultural Engineering, Texas
A&M University, College Station, TX, USA
John M. Bryden Grounded and Inclusive Innovation Group, Norwegian Institute
for Bioeconomy Research, Oslo, Norway; University of Aberdeen, Aberdeen,
Scotland, UK
Benno Böer UNESCO Liaison Office in Addis Abeba with the African Union and
UNECA, UNESCO, Addis Abeba, Ethiopia
Bassel T. Daher Department of Biological and Agricultural Engineering, Texas
A&M University, College Station, TX, USA
Mohammad El Ramahi Asset Management and Technical Services, Masdar
Clean Energy, Masdar, UAE
Stephen Gitonga Regional Sustainable Energy Specialist, United Nations
Development Programme (UNDP), Regional Hub for Arab States, Amman, Jordan
Atef Hamdy Water Resources Management, CIHEAM-IAM Bari, Valenzano,
Bari, Italy
Kishan Khoday Climate Change, DRR and Resilience, United Nations
Development Programme (UNDP), Regional Hub for Arab States, Cairo, Egypt

xiii
xiv Contributors

Rabi H. Mohtar Department of Biological and Agricultural Engineering, Texas


A&M University, College Station, TX, USA; Zachry Department of Civil
Engineering, Texas A&M University, College Station, TX, USA
Peter Nasr The American University in Cairo, New Cairo, Egypt
Hani Sewilam RWTH Aachen University, Aachen, Germany
Paul Sullivan Georgetown University, Washington, USA; National Defense
University, Washington, USA
Disclaimer

The designations employed and presentations of material throughout this publica-


tion do not imply the expression of any opinion whatsoever on the part of the
United Nations University (UNU) and the United Nations Educational, Scientific
and Cultural Organization (UNESCO) concerning legal status of any country,
territory, city or area, or of its authorities, or concerning the delimitation of its
frontiers or boundaries. The views expressed in this publication are those of the
respective authors and do not necessarily reflect the views of the UNESCO or
UNU. Mention of the names of firms or commercial products does not imply
endorsement by UNESCO or UNU.

xv
Editorial Statement

We extend our greetings to the respected member states of UNESCO and UNU in
the Arab Region and to all of the people living in this vast and diverse region
extending from Ras al Hadd (Oman) over a geographical stretch of almost 8000 km
to Nuankchot (Mauretania). We invite the government authorities and the reader to
please take our science-based advices and warnings offered in this volume very
seriously. This is important and in the best interest of the people.
Since we have been invited, accepted the invitation, and decided to work on this
highly important subject of “The water, energy, and food security nexus in the Arab
Region,” we wish to highlight two highly important facts right in the beginning of
this science-based multidisciplinary book, that is being published at a time, when
the world, and in particular the Arab Region, faces very serious challenges that
relate to water, energy, and food availability, which, in turn, relates to good
governance.
1. Some places in the Arab Region, which are poor by natural resources, but rich
by per capita income, have developed rapidly from places of low human pop-
ulation densities with the best environmental footprints globally to places of the
fastest increase in human population and the highest rate of per capita water and
food and energy consumption and wastage, even though the water and food are
naturally not available, and energy availability has technical, geographical, and
volume limitations. Water has to be desalinized and/or imported and trans-
ported, with huge energy costs, over vast distances to large and major cities.
Food needs to be imported from places of high financial and energy costs.
Energy is mainly based on fossil fuel, naturally limited, and exposed to
unpredictable market fluctuations. Moreover, the attitude of the young genera-
tion has already lost the good virtues of “Bedouin” or “Samakiya,” who saved
every drop of water, has changed to be highly wasteful and careless, so, as if this
current richness will continue forever. This wastage of water, energy, and food
is unacceptable from the environmental point of view, the societal point of view,
and from the religious point of view.

xvii
xviii Editorial Statement

2. Even though water in the water-poor Arab Region is one of the most limiting
factors for sustainable human living and development, it remains available for
free in numerous compartments. This is a fact, even though, it has been known
for many years that good water governance urgently requires incentives for all,
for everybody, to contribute to good water management. We do not suggest or
recommend what needs to be done—but we urgently call for incentives—those
who are willing to listen, to understand, and to apply good water, food, and
energy management practices should one way or the other be rewarded by the
governments and authorities, because they do the right thing in the best interest
of the entire human community.
Water is and remains the most critical factor to provide the bare necessities. The
traditional Arabs, wherever they had water, they used it wisely, and they most
certainly avoided any wastage. This is no longer the case. Wastage is real and needs
to be readdressed. With those two major points being at our heart, we seriously
encourage the authorities and people to improve things—the wastage must no
longer go on as usual—things have to change—things have to improve—urgently.
In October 2002, “The 1st Gulf Regional Expert Meeting on Ecosystems and
Water Constraints,” hosted by the International Centre for Biosaline Agriculture
(ICBA) in Dubai, brought together a number of multidisciplinary experts, and
topics that required urgent attention on policy level were identified and discussed.
In Muscat, September 2003, during “The 2nd Gulf Regional Expert Meeting on
Ecosystems and Water Constraints,” which was hosted by the Ministry of Regional
Municipalities, Environment, and Water Resources of the Sultanate of Oman, these
issues were further discussed.
Based on these meetings, UNESCO and the United Nations University produced
a book in 2006 on “Policy perspectives for ecosystem and water management in the
Arabian Peninsula,” highlighting the great importance of the connectivity between
water resources management and ecosystem management, including marine,
coastal, and terrestrial man-made and natural ecosystems, in the Arabian Peninsula.
Six years later, in 2012, the situation most certainly did not really improve,
except for the willingness and knowledge capacity, but dependency on
carbon-intensive sea water desalinization, food import, and strategic
water-contingency planning remained. The Qatar National Food Security Program
(QNFSP), under the umbrella of His Highness the Emir of Qatar, Sheikh Tamim bin
Hamad al Thani, and spearheaded by His Excellency, Mr. Fahad Al Attiyah,
Chairperson of the QNFSP, invited UNESCO to assist organizing the International
Conference on Food Security in the Dry Lands (FSDL). UNESCO agreed sup-
porting the conference, which was supported by UN agencies, such as UNESCO,
the FAO, the WFP, and IFAD.
There were three main sessions at the conference, one on food security another
one on responsible investment, and one on the food security/water
security/ecosystem management—connectivity. This session was supported and
headed by UNESCO, and supported by the UNESCO Office in Doha, the Division
of Hydrological Sciences, and the Division of Ecological and Earth Sciences.
Editorial Statement xix

Later, the QNSFP requested UNESCO to produce this book, partly based on the
conference presentations, and partly on additional knowledge from the Arab
Region. UNESCO, in partnership with the United Nations University, agreed to
produce this volume on “The water, energy, and food security nexus in the Arab
Region.”
Jobs and income are related to social despair and frustration, illegal migration,
and full-fledged conflicts.
UNESCO and the United Nations University financially and technically sup-
ported the publication. We are very thankful to the Government of Qatar for hosting
the International Conference, and in particular His Highness Sheikh Tamim, and
His Excellency Mr. Al Attiyah. We also congratulate the numerous authors of the
chapters for their highly professional contributions, which led to the production of
this important volume.
The reader, concerned authorities, and the decision-makers who decide about
funding activities related to the theme of this book are encouraged to use this
publication as a supporting tool to select practical and replicable activities, and turn
from an era of too much talk and too much planning to real action that will bring
positive changes, including more water security, more food security, better energy
management, with related jobs and income in the best interest of the diverse Arab
Region, and its people.
They have to decide how best to utilize science, education, and how to enhance
food security, water security, and energy management towards sustainable human
living. Tunisia 2011 must not be forgotten, and the fire sparks that flew over its
borders to Egypt, Yemen, Syria, and other compartments in the Arab Region and
the world. The root causes of these conflicts are of course multidimensional.
However, they most certainly include the availability of jobs and income,
socio-economic realities, and the availability of the basic necessities for sustainable
human living. Sustainable human living, in turn, undoubtedly includes safe
drinking water, food, and energy.

Burlington, Canada Kamel Amer


Burnaby, Canada Zafar Adeel
Addis Abeba, Ethiopia Benno Böer
Sana’a, Yemen Walid Saleh
June 2016
Chapter 1
Status of Water in the Arab Region

Waleed Khalil Al-Zubari

Abstract The Arab Region has an extremely poor supply of water resources with
many areas experiencing unpredictable rainfall. Taking population size and growth
into consideration, the Arab Region is considered one of the world’s most
water-stressed regions, with continuously decreasing per capita freshwater avail-
ability. Despite the strenuous efforts made by the Arab countries in augmenting
their water supplies to meet increasing demands, the emphasis on the supply-driven
approach for water management has not only reached its physical and financial
limits for many countries and led to the over-exploitation of the region’s natural
water resources, but has also demonstrated its inability to deliver a substantial
degree of water sustainability or security. Currently, Arab countries are experi-
encing an alarming future of increasing water scarcity and increasing water
demands and supply costs, which might not only threaten their future development
and hamper human and socio-economic development efforts, but also the preser-
vation and sustainability of their past socio-economic achievements. A major
review of and shift in water policies in Arab countries, emphasizing demand
management and conservation is urgently needed, with the overall objective of
securing long-term water supplies while meeting strict criteria for socio-economic,
financial and environmental sustainability, and public health requirements. The
successful implementation of these policies would be the cornerstone for coping
with water scarcity in the region. Moreover, shared water resources should be given
high priority in order to reach agreements and form treaties regarding water allo-
cation, including quality considerations, according to international water law.
Finally, unless there is a major shift in the population policies of the region, the
water problem will continue to be a major constraint to its future development.


Keywords Resources and utilization Shared water resources  Desalination 
 
Treated wastewater Municipal Agriculture

W.K. Al-Zubari (&)


Water Resources Management Program, College of Graduate Studies, Arabian Gulf
University, Manama, Bahrain
e-mail: waleed@agu.edu.bh

© Springer International Publishing AG 2017 1


K. Amer et al. (eds.), The Water, Energy, and Food Security Nexus
in the Arab Region, Water Security in a New World,
DOI 10.1007/978-3-319-48408-2_1
2 W.K. Al-Zubari

1.1 Introduction

Water is one of the most valuable resources on Earth. It is an important vector in the
socio-economic development and for supporting the ecosystem. In Arab countries,
the importance and value of water is even more pronounced, for most of these
countries are located in an area considered as one of the world’s most water-stressed
regions. Located within an arid to extremely arid region, rainfall scarcity and
variability, coupled with high evaporation rates have characterized this part of the
world as having a limited availability of renewable freshwater. However, the
increasing scarcity of renewable water resources is not the only distinctive char-
acteristic of the region; weak governance and inadequate levels of management,
increasing water deficits, and the continuous deterioration of the quality of natural
water resources have become during the past decades equally distinguishing fea-
tures as well. Now, water scarcity is among the most pressing challenges con-
fronting the Arab countries, presenting a tangible threat that could negatively
impact socio-economic development efforts.
In the last three decades, the Arab Region has experienced rapid population
growth and accelerated socio-economic development. The population of the region
has doubled from about 170 million inhabitants in 1980 to more than 350 million in
2010 (UNDESA 2012). This growth was accompanied by a substantial increase in
water demands driven mainly by rapid urbanization and the implementation of
agricultural policies aimed at achieving national food security in many Arab
countries. To meet escalating demands, most Arab countries have focused their
efforts mainly on supply management and augmentation, where governments have
invested heavily in major infrastructures to secure supplies and to provide water
supply and irrigation services. Demands are being satisfied by increasing the
storage capacity for surface water, development of groundwater, extensive instal-
lation of desalination plants, and expansion in wastewater treatment and reuse.
However, the emphasis on “securing supplies” approach to meet increasing
demands has not only reached its physical and financial limits for many countries,
but has also led to the over-exploitation and quality deterioration of its natural water
resources.
In fact, implementing a water management system that is based on securing
supplies, without adequate attention to improving water efficiency and demand
management, has created “excess demand”, and has led to the emergence of a
number of unsustainable water uses in the region, such as low water use efficiency,
growing per capita water use, increasing cost of water production and distribution,
deterioration of water quality and land productivity, and increasing volumes of
effluent discharges polluting the limited surface and groundwater resources in the
region. The latter factor, in addition to its impact on human and ecological systems
health, has contributed to increasing the water scarcity in the region.
These conditions were further aggravated in many countries by the lack of
comprehensive long-term water policies and strategies that are based on
supply-demand considerations, and were further compounded by the institutional
1 Status of Water in the Arab Region 3

weaknesses, manifested by the multiplication and overlap of water agencies,


inadequate institutional and individual capacities, low levels of public participation,
and absence of enabled societies.
Furthermore, one of the major challenges facing the Arab Region is the high
overall dependency ratio of the region on shared water resources. As more than 60 %
of surface water resources originate from outside the Arab Region, this issue remains
a major concern threatening the region’s stability, food security, and complicates
national water resources management and planning. Conventions and agreements on
equitable sharing and management of water resources have not been signed by
riparian countries. In addition, some Arab countries are deprived of their water
resources by occupying powers (i.e., Occupied Palestinian Territories, Golan
Heights, and Southern Lebanon), which is another major issue in the region and is
constraining the development of the population of these countries. Moreover, the
critical nature of the current water situation in the Arab Region is expected to be
further aggravated by the impacts of climate change, such that it is anticipated that
water scarcity and quality deterioration in the region will increase due to precipitation
reduction, increase in domestic and agricultural water demands due to temperature
increase, and seawater intrusion of groundwater resources due to sea level rise.

1.2 Water Availability and Uses

Driven by severe aridity, water availability in the Arab Region has played a
dominant role in determining human activities, settlements, socio-economic inter-
actions, and growth more than any other part of the world. The Nile, Euphrates, and
Tigris Rivers hosted one of the greatest early civilizations on earth. The ancient
Yemen civilization was closely tied to the availability of water resources, and its
decline is historically related to the destruction of the ancient Ma’areb dam (AWC
2009).
Basic statistics clearly illustrate the magnitude of the scarcity challenge. Despite
covering 10.2 % of the world’s area, the Arab Region receives only 2.1 % of the
world’s average annual precipitation and contains as little as 0.3 % of its annual
renewable water resources (ACSAD 1997). More than 85 % of the Arab Region is
classified as arid and hyper arid, receiving an average annual rainfall of less than
250 mm. Areas that have relatively high rainfall rates are limited to locations
between the mountainous highlands on the coasts of the Mediterranean and those
on the coasts of the Atlantic Ocean, the southwestern parts of the Arabian
Peninsula, in addition to the southern parts of Sudan (Fig. 1.1). Furthermore,
rainfall rates are generally characterized by high spatial and temporal variability
with frequent drought cycles experienced in the region, thus negatively impacting
the reliability and availability of water resources.
Under these climatic conditions and poor endowment of water resources, and as
a consequence of the rapid population growth experienced by the region since the
mid 1970s, overall per capita freshwater availability decreased dramatically in all
4 W.K. Al-Zubari

Fig. 1.1 Rainfall distribution in the Arab Region (source Droubi et al. 2006)

Arab countries (Fig. 1.2). The majority of these countries are currently below the
water poverty line of 1000 m3/capita/year. In contrast, the world average for per
capita water availability is about 7240 m3/capita/year. The current (2011) overall
per capita freshwater availability in the Arab Region is at about 800 m3/capita/year.

Fig. 1.2 Trends in total renewable freshwater resources per capita in Arab countries (m3/
capita/year), 1992, 2002, and 2011 (source UNDP 2013; based on FAO 2013)
1 Status of Water in the Arab Region 5

Based on the projected population increase, it is expected that this indicator will
continue to decrease to reach about 500 m3/year by 2030 when the Arab Region
population will reach more than 500 million. This means that the whole region will
experience absolute water poverty, whereby water will become a major constraint
for development impacting the standard of living, health, and the environment
(Falkenmark 1989). In addition, by 2030, climate change is expected to have led to
a 20 % reduction in renewable water resources and more droughts in the region
(Doumani 2008), which would further exacerbate the current scarcity situation.
To meet escalating sectoral water demands, estimated at more than 255 billion
cubic meters (BCM) in 2010 (UNDP 2013), Arab countries rely on both traditional
water resources (including surface water and groundwater) and non-traditional
water resources (including desalinated water, treated wastewater, and irrigation
drainage water) to varying degrees (Fig. 1.3a). Most of the Mashreq, Nile Valley,
and Maghreb countries rely mainly on surface water resources, while the Arabian
Peninsula countries rely mainly on renewable and non-renewable groundwater
resources. All Arab countries are increasingly using treated wastewater, while
desalinated water progressively represents a major component in the water budget
of the Gulf Cooperation Council (GCC) countries. Reuse of agricultural drainage
water is practiced mainly in Egypt and Syria. The majority of water resources in the
region are being used for agriculture (85 %), while municipal and industrial sectors
consume about 8 and 7 % of the total water use, respectively (Fig. 1.3b).
The following is a description and analysis of the status and issues of water
resources and their uses faced by Arab countries.

Fig. 1.3 a Water resources


(SW surface water, GW
groundwater, AGDRW reused
agriculture drainage water,
DES desalinated water, TWW
treated wastewater); and
b water uses in the Arab
Region (AGR agricultural
sector, MUN municipal
sector, IND industrial sector).
Data source UNDP (2013)
6 W.K. Al-Zubari

1.2.1 Water Resources Availability

1.2.1.1 Conventional Water Resources

(a) Surface water resources


In the Arab Region, there are 23 major drainage basins of varying sizes that
contain either perennial rivers or ephemeral streams (wadis). A few medium size
rivers exist, mainly in Morocco, Algeria, Tunisia, Sudan, Lebanon, and Syria that
originate and flow within the national boundaries. Major rivers, such as the Nile,
Tigris, Euphrates, and Senegal flow from outside the region, while there are others
that are shared between the Arab countries of Syria, Jordan, and Lebanon. A few
wadis are also shared by some Arab countries of the Arabian Peninsula (Shahin
1989). Figure 1.4 displays the major drainage basins in the Arab Region, while the
region’s major rivers are listed in Table 1.1.
It is worth noting that more than 60 % of the surface water resources in the Arab
Region (about 174 billion cubic meters per year (BCM/year) of a total of 224
BCM/year) originates from outside the region. These include major rivers like the
Tigris, Euphrates, Nile, and Senegal. There are also many rivers that are shared
among many Arab countries that are comparatively smaller, which include the
Yarmouk in Syria and Jordan, and the Orontes and Nahr El-Kebir in Lebanon and
Syria. Agreements on equitable sharing and the management of these water
resources have not been signed by riparian countries. The issue of management of
shared water resources among countries outside the region, as well as between the
countries within the region, represents one of the major challenges facing the
region. Managing shared water resources is also a major determinant of regional

Dams
Lakes and
Swamps
Rivers
Major Permanent Stream Land Use
Permanent Streams Irrigated Crops
Major Ephemeral Streams Rainfed Crops Rocks and Lava Sheets
Ephemeral Streams (Wadis) Natural Pasture Swamps

Political Boundaries Forests Deserts

Fig. 1.4 Major drainage basins in the Arab countries (and land use) (source CEDARE and AWC
2004)
1 Status of Water in the Arab Region 7

Table 1.1 Major drainage basins in the Arab Region (sources Nile Basin Initiative n.d.; CIA
2003; Kibaro’lu 2004; Scheumann et al. 2011; Al-Mooji 2004; OMVS 2003)
Basin Tributaries Basin size River Average Riparian countries
(1000 km2) length discharge
(km) (BCM/year)
Nile Victoria Nile/Albert 3173 6693 109.5 Egypt, Sudan, South
Nile, Bahr El Jabel, Sudan, Burundi, DR
Bahr El Ghazal, Congo, Eritrea,
White Nile, Baro Ethiopia, Kenya,
Pibor-Sobat, Atbara, Rwanda, Tanzania,
Blue Nile and Uganda
Euphrates Sajour, 647,075 2330 32.0 Iraq, Syria, Turkey,
Jallab/Balikh, Jordan, and Saudi
Khabour Arabia
Tigris Batman, Khabour, 146,239 1718 52.0 Iraq, Syria, Turkey,
Greater Zab, Lesser and the Islamic
Zab, Adhaim, Republic of Iran
Diyala, Cizre, Wadi
Tharthar
Jordan Upper Jordan (Dan, 19,839 251 1.34 Lebanon, Syria,
River Hasbani, Banias, Jordan, Palestine,
Huleh valley, Lake and Israel
Taberias), Yarmouk,
Lower Jordan
Orontes Afrin and Karasu 37,900 448 2.8 Lebanon, Syria, and
(Al-Assi) Turkey
Nahr Al Noura el 0.991 90 0.330 Lebanon and Syria
Kebir Tahta-Aroussa and
Safa-Raweel
Senegal Falémé, Bafing, and 300 1800 22.0 Senegal, Mauritania,
Bakoye Rivers Mali, and Guinea

stability, and can threaten water food, and energy security, as well as the overall
national water resources planning.
To secure water supply, several Arab countries have made significant invest-
ments in water storage infrastructures. These are represented by major networks of
water storage and conveyance systems, which helps smooth supply between sea-
sons, as well as help reduce flood risks. Some countries, particularly those with high
variability and transboundary waters, have tried to minimize supply risks by
investing in water storage, while those located in hyper-arid areas have constructed
dams with the aim of recharging groundwater. The total dam capacity in the Arab
Region is estimated at about 262 BCM. More than 86 % of this total capacity is
owned by four countries having large populations, with agriculture representing a
main activity. These countries are Egypt (64.6 %), Iraq (19.2 %), Morocco (6.2 %),
and Syria (6.1 %) (UNDP 2013).
8 W.K. Al-Zubari

(b) Groundwater resources


The second major water resource in the Arab Region is the shallow and deep
groundwater supplies within the national boundaries or shared among Arab coun-
tries. In some countries, such as the Arabian Peninsula countries, Jordan, Occupied
Palestinian Territories (OPT), and Libya, groundwater dependency is high, and
reaches more than 80 % in some of these countries. Even in the relatively surface
water-rich countries, groundwater utilization and reliance is emerging and is on the
rise due to the steady increase in water demands, and groundwater dependency is
expected to increase with time. Groundwater resources are being utilized in these
countries mainly by the agricultural sector, followed by the domestic and industrial
sectors.
Groundwater occurrence, movement, exploitation, and water quality in both
shallow and deep aquifers is controlled by the major geological structures, strati-
graphic, and sedimentation processes. Extensive deep sedimentary formations in
northern Africa and the Arabian Peninsula countries contain major aquifers that are
non-renewable, or fossil (very limited recent recharge that took place during wet
periods 15,000–25,000 years ago), mostly confined, and exhibit large variation in
water quality (salinity range 200–20,000 mg/L). With some requiring treatment,
and with temperatures of 40–65 °C, their suitability for domestic consumption in
some locations is limited. Many of the deep aquifers are shared among Arab
countries of North Africa, the Arabian Peninsula, Jordan, and Iraq (UN-ESCWA
and BGR 2013). Table 1.2 displays a summary of the major aquifers in the region.
The unconsolidated deposits, mainly sand and gravel, belonging to the Neogene
and Quaternary periods form the shallow renewable aquifers that exist under river
beds and their flood plains, deltas, wadi beds, major depressions, and the interior
coastal plains. The aquifers of the Nile and its delta, the Tigris-Euphrates, various
intermountain valleys in North Africa, and wadis in the Arabian Peninsula allow
countries to store adequate reserves of good quality water, and receive frequent
recharge amounts from river flow, and during flooding events. These aquifers are
used extensively for domestic and irrigation purposes.
Currently, both renewable and non-renewable groundwater resources in most
Arab countries, and specifically in the Arabian Peninsula countries, are being
heavily utilized to meet increasing water demands, particularly irrigation water
demands, and are in critical condition. The excessive utilization of these resources
beyond their recharge rates has resulted in continuous water level decline and
quality deterioration due to seawater intrusion and connate water encroachment. In
addition to their over-exploitation and quality deterioration, groundwater resources
in most Arab countries are being threatened and polluted by numerous point and
non-point sources of pollution generated from anthropogenic sources (agricultural,
industrial, and domestic activities).
Table 1.2 Major groundwater systems in the Arab Region (source adapted from Ksia 2009; Sokona and Diallo 2008; Khater 2010)
Groundwater system Localities Countries sharing the Area Remarks
system (1000 km2)
Great desert Nubian sandstone aquifer system Libya, Egypt, Sudan and 2200 Sandstone
sandstone aquifer Chad
systems in North Continental intercalary Tunisia, Libya, Algeria 600 Sandstone
Africa
Terminal complex Tunisia, Libya, Algeria 430 Sandstone
Bechar western Algeria 240
Fazzan South-western Libya 450
1 Status of Water in the Arab Region

Eastern Cenomanian-Turonian limestones in Syria, Syria, Lebanon, Jordan, 48 Contributes to the flow of the
limestone/carbonate Lebanon, and Palestine; and Wadi As Sir Palestine Orontes, Litani, and other
Mediterranean limestone in Jordan Lebanese rivers, as well as the
aquifer system Jordan River
Hauran and Arab Syria, Jordan, and Saudi 15 Contributes to the flow of the
mountain basaltic Arabia Yarmouk and Azraq basins
aquifer system through El-Hamma, Azraq, and
Mazreeb springs
Eastern Arabia Umm Er Radhuma dolomite and limestone Bahrain, Kuwait, Oman, 1600 Primarily a limestone and
tertiary carbonate aquifer in the Arabian Peninsula and Iraq; Saudi Arabia, United Arab dolomite aquifer; hydraulically
aquifer system Dammam limestone and dolomite aquifer in Emirates, Yemen, Iraq, interconnected; a
the Arabian Peninsula (except for Yemen); Syria, and Jordan recharging-discharging aquifer
and Neogene aquifer in Bahrain, Kuwait, system
Oman, Qatar, and the UAE
Note The continental intercalary and the terminal complex together form the North Western Sahara Aquifer System (NWSAS) which has an area of
1,030,000 km2 extending over Tunisia, Libya, and Algeria
9
10 W.K. Al-Zubari

1.2.1.2 Non-conventional Water Resources

In the face of rising demands and the limited supply of conventional water
resources, non-conventional water supplies have been widely adopted in Arab
countries. These non-conventional supplies include desalinated water processed
through desalination plants, treated wastewater produced through reuse programs,
and irrigation drainage water. The volumes and contribution to meeting demands
have been increasing for all of these supplies. In terms of non-conventional irri-
gation drainage water, Egypt and Syria consume the most among Arab countries, as
their total volume of reused agricultural drainage in irrigation, generated from
mixing these waters with freshwater, is about 7500 and 2300 million cubic meter
per year (MCM/year), respectively (UN-ESCWA 2007).
(a) Desalination
The Arab Region leads the world in the use of desalination technology as it
possesses more than 50 % of the world’s desalination capacity. While desalinated
water represents a very small percentage of the water supply in Arab countries
(about 1 % of total water supply sources; Fig. 1.3a), it represents about 100 % of
the water supply to many Arab cities. The role of desalination as a major source for
domestic water supply is expected to increase in the future due to population and
urbanization growth, industrialization, and the depletion and/or quality deterioration
of conventional water resources, especially non-renewable resources.
Currently, the total cumulative capacity of all desalination plants contracted for
Arab countries is about 12 BCM/year (32.5 MCM/day; GWI 2010; Fig. 1.5). The
largest owners of desalination capacity in the region are the GCC member countries
(81 %), Algeria (8.3 %), Libya (4 %) and Egypt (1.8 %). It is expected that the
current high rate of annual increase will be maintained in the next decade

Fig. 1.5 Accumulated current and near term total contracted capacity in Arab countries (source
GWI 2010)
1 Status of Water in the Arab Region 11

(GWI 2010) to meet the escalating domestic water demands, with the total
desalination capacity in the region expected to reach about 25 billion m3/year (69
MCM/day) by the year 2016 (GWI 2010), and possibly 31.4 BCM/year by 2025
(Balaban 2008). However, the majority of the anticipated increase in desalination
capacity will be concentrated almost exclusively in the high-income, energy
exporting wealthier countries of the region, primarily the GCC countries (Al-Jamal
and Schiffler 2009). The increase in desalination capacity would be used for
municipalities and industries in these countries. Currently, the average share of
desalination in the municipal sector in the GCC countries is about 75 %, which is
either used directly or blended with groundwater, a ratio expected to increase in the
future with the continuous deterioration of groundwater resources.
In the Arab Region, particularly in the GCC countries, the thermal Multi-stage
Flash (MSF) technology still dominates the market, although installed capacity for
Reverse Osmosis (RO) has increased recently. The RO desalination technology
requires less energy and is easily scalable due to its high modularity (Al-Jamal and
Schiffler 2009). Electric energy consumption is almost the same or lower than
distillation, while it does not need any thermal energy (World Bank and BNWP
2004). Recently, RO-MSF hybrid systems are now being used as cogeneration
systems (Bushnak 2010). As experience and technology have developed, produc-
tion costs for desalination have fallen. New technologies, such as RO, electro-
dialysis, and hybrids, can deal with different types of input water, are more energy
efficient, or both. Furthermore, unit sizes have increased, bringing economies of
scale (World Bank 2007). These advances have driven prices down from an average
of US$1.0/m3 in 1999 to between US$0.5/m3 and US$0.8/m3 in 2004 (World Bank
and BNWP 2004). These trends in desalinated water cost reduction could mean that
desalination technology is becoming increasingly viable for lower income
countries.
While the financial cost of desalinated seawater from recently completed large
plants has been decreasing to around US$0.70 per cubic meter (without distribution
costs and depending mainly on plant size, duration of amortization, and energy
costs) in the US and other places, the average water production costs in GCC
countries remain somewhere between US$ 1 and 2 per cubic meter (World Bank
and AGFUND 2005). This can be attributed to many factors, including the dom-
inance of the public sector in the desalination industry. More importantly, the
investment needs for constructing new desalination plants to supply water for
rapidly growing populations and high per capita consumption in these countries are
enormous, especially under the government policies of heavy subsidies of the water
sector, which will not only exacerbate rapidly rising water demands, but will also
place an intolerable burden on national budgets. It is estimated that the annual
investment to produce, manage, and operate seawater desalination plants in Arab
countries will reach US$ 15–20 billion in the next decade (AFED 2010).
Moreover, desalination is an energy-intensive process. For example, in Saudi
Arabia, home to 35 % of the Arab Region desalination capacity, 25 % of Saudi oil
and gas production is used locally to generate electricity and produce water (i.e.,
Cogeneration Power Desalination Plants [CPDPs]), and if the current growth rates
12 W.K. Al-Zubari

of water demands continue, this fraction will be 50 % by 2030 (Al-Hussayen 2009).


Similarly, in Kuwait, CPDPs represent more than 50 % of the total energy con-
sumption in the country and if this trend continues, the energy required to meet
desalination plants demands will be equal to the country’s current fuel oil pro-
duction by the year 2035 (Darwish et al. 2010). Therefore, energy efficiency should
be a key criterion in commissioning new plants and upgrading old plants (Bushnak
2010).
Despite the importance of desalination in producing fresh water for Arab pop-
ulations and sometimes its positive impacts in reducing pressures on conventional
water resources, desalination plants have negative environmental impacts to con-
sider as well. In general, the environmental footprint of desalination has been
reduced through technological progress, however, some significant environmental
impacts remain. These impacts include air pollution from oxides emitted during
production, seawater and marine life pollution from rejected brines, which have
elevated temperatures and increased salt concentrations, the emissions of residual
treatment chemicals (e.g., anti-foaming and anti-scaling agents), and trace elements
picked up within desalination plants (Abderrahman and Hussain 2006).
Energy requirements for desalination can be met through renewable energy
resources that do not produce CO2 directly, with wind and solar energies most
commonly used, and wave power as a possibility in the future. Until recently
however, renewable energy was used for desalination only for small plants in
remote areas without access to electricity from the grid. In recent years, research
and development in this field has intensified, and several pilot desalination plants
using solar, wind, or geothermal energy have been installed, with the majority
operating successfully (Al-Jamal and Schiffler 2009). This new technological trend
to utilize renewable energy, particularly solar energy, is specially suited to the Arab
Region, which due to its geographic location has vast solar energy potential. In this
regard, several GCC countries are investing in renewable energy particularly for
desalination applications. Masdar City in Abu Dhabi, King Abdullah University for
Science and Technology (KAUST), and the recently established King Abdullah
City for Atomic and Renewable Energy (KA-CARE), have ambitious research and
development programs in solar energy and desalination.
(b) Treated Wastewater
Reuse of treated municipal wastewater in Arab countries constitutes another
increasing source of water driven by escalating water consumption in urban areas.
However, Arab countries vary in the scale and nature of their wastewater treatment
capacity, level of treatment, and reuse. Where it is not already being used, treated
wastewater is now considered as a potential non-conventional source of water in
most of these countries. Currently, treated wastewater represents about 1 % of the
total water used in the Arab Region, but this is expected to increase with time.
Generally, water scarcity, financial capabilities, and the importance of the agri-
cultural sector play a role in shaping wastewater treatment and reuse (Varis and
Tortajada 2009). While most of the region’s countries have programs for reusing
1 Status of Water in the Arab Region 13

Fig. 1.6 Total wastewater produced and wastewater treatment rate in some Arab countries, 2009–
2010, in BCM (source World Bank et al. 2011)

treated wastewater in irrigation (fodder crops, cereals, alfalfa, olive, and fruit trees
are most widely irrigated with treated water), few countries have institutional
guidelines for regulating the reuse of treated wastewater (MED WWR WG 2007).
Yearly data on generated, treated, and reused wastewater volumes in the Arab
Region are outdated and span over a long time period (1991–2006), which makes
proper analysis and comparison between countries difficult. Available data indicate
that a total of more than 13 BCM of wastewater is produced annually in Arab
countries, with only about 40 % being treated primarily or secondarily (UNDP
2013). The rest is discharged untreated to open water channels, the sea, or ground
reservoirs, which can cause several health concerns for both humans and the
environment. Another estimate (Qadir et al. 2007) indicates that the ratio of the total
volume of treated wastewater to that generated in the Arab Region is about 54 %,
which is higher than some other regions in the world (Asia [35 %], Latin
America/Caribbean [14 %], and Africa [1 %]). However, it should be noted that
this ratio varies greatly among Arab countries, from almost nil to almost 100 %.
Figure 1.6 displays the overall collection ratios and the treatment rate as a percent
of that collected for some Arab countries. If these waters were properly treated and
reused, they could support the water demands of some sectors, such as the agri-
cultural and industrial sectors, and potentially prevent health and environmental
concerns.
There are many advantages to using treated wastewater in the arid Arab coun-
tries, the most important is its potential to proportionally meet the increased need
for water with population and urbanization growth, and in alleviating the water
scarcity conditions faced by these countries. However, there are many handicaps to
the expansion of water reuse capacity. These include social barriers, technical
handicaps, and institutional and political constraints.
14 W.K. Al-Zubari

1.2.2 Water Uses

(a) Municipal Water Sector


During the period of 2005–2015, urbanization increased from 67 % of the Arab
population to about 70 %, and is expected to continue to increase by the same rate
over the next 10 years to reach 73 % by the year 2025 (UN Urbanization Prospects
2014). Along with this relatively rapid urbanization rate, domestic water con-
sumption has increased from about 14 BCM in the early 2000s to about 20.4 BCM
in 2011 (UNDP 2013), and is expected to increase to about 30 BCM by 2025
(Hamoda 2004), i.e., it is anticipated to increase by more than 50 % in the next
10 years.
Average per capita domestic water consumption in the Arab Region is about
200 L/day, but varies significantly among the countries of the region. For example,
domestic water consumption in the Gulf Cooperation Council (GCC) countries
ranges from 300 to 750 L/capita/day (World Bank and AGFUND 2005), ranking
among the highest in the world. The rise in per capita water consumption in these
countries is attributed to many factors including the absence of proper demand
management and of a price-signaling mechanism. Government policies have pri-
marily focused on the supply side of water production, coming from aquifers or
desalination plants. Water tariffs in the region are generally quite low, representing
no more than 10 % of the total cost on average, with no incentives for consumers to
save water.
Moreover, municipal water requirements are further exaggerated by the high
percentages of non-revenue water (NRW) in the municipal distribution network,
particularly its physical leakage component. The physical leakage can reach more
than 60 % in poorly maintained distribution networks in some Arab cities, and is
generally high in many countries, including the financially incapable and capable.
(b) Agricultural Sector
Although urban demand for water has been steadily increasing, the agricultural
sector continues to consume the largest amount of water in the Arab Region,
accounting for about 85 % of total water used. Agricultural water use rose from
about 165 BCM in 1995 (ACSAD 1997) to more than 218 BCM in 2012/2013
(UNDP 2013); a 32 % increase in a period of about 15 years.
During the last three decades, the Arab Region has witnessed an accelerated
developmental boom. To meet increasing food demand resulting from rapid pop-
ulation growth rates, economic policies favoring food self-sufficiency and
socio-economic development were adopted in many Arab countries. These policies
concentrated mainly on horizontal agriculture and played an effective role in pro-
moting the expansion of agricultural lands and the introduction of new arable areas,
and extension of irrigated cultivation. However, these policies were often made
without regard to the limited availability of water resources, were based on
supply-side management approaches, and without giving adequate attention to
1 Status of Water in the Arab Region 15

demand management and conservation, nor to the consequences of the widespread


use of water for irrigation on water resources and subsequently on agriculture itself.
Moreover, agricultural water demands are exaggerated by low irrigation effi-
ciencies that result from the use of traditional irrigation methods. Currently, surface
irrigation is by far the most widely used irrigation method in the Arab Region,
practiced on 80.3 % of the total full or partial control irrigation area, sprinkler
irrigation is practiced on 22.8 %, while micro-irrigation on only 2.8 % of the total
area (FAOSTAT 2008; LAS and UNEP 2010). It is estimated that approximately
50 % of water used for irrigation is wasted due to the inefficiency of the irrigation
methods used (Abu-Zeid and Hamdy 2004). Some studies even estimate irrigation
efficiencies as low as 30–40 % in the Arab Region (AFED 2010). These inefficient
methods, including deep percolation, evaporation, and surface runoff represent
foregone opportunities for water use. The outcome is generally poor agricultural
performance, and more dangerously, the unsustainable use of water resources for
agriculture as demonstrated by the continuous decline of water levels and salin-
ization. In many countries, after reaching a peak in agricultural production in the
past decades, percentage of food self-sufficiency has begun to decline as land and
water resources are depleted beyond their sustainable limits and populations con-
tinue to grow rapidly; water scarcity has become a critical constraint to agriculture.
Moreover, agricultural practices are also blamed for increased soil and water
salinity, toxic pollution from use of agro-chemicals, damming of rivers, and the loss
of biodiversity associated with wetlands destruction (AFED 2010).
As most of the Arab countries are heading towards a severe water scarcity, the
agricultural sector, with its lion share of water resources, is currently under various
types of pressures (IFAD 2009): First, it is under pressure to produce more food to
help reduce Arab countries’ enormous food imports bill (US$28 billion in 2006).
Second, as the largest employer of people in the rural and marginal areas, the
agricultural sector is under pressure to halt the decline in its ability to generate
employment opportunities, especially for young people. In 2006, 37 %, or
47.6 million people, out of an economically active population of 126 million, were
engaged in agriculture, down from 47.8 % in the 1990s. More employment in the
rural and marginal areas is expected to help reduce the influx of rural-urban
migration, respond to increasing market demand, and eventually reverse the decline
of the sector’s contribution to the gross domestic product (GDP) of Arab countries
(in 2005, the average agriculture’s contribution to GDP was quite low at 12.5 %
[ranging between 0.3 % in Kuwait and Qatar, to 34 % in the Sudan]). Third, the
agricultural sector is under mounting pressure to redirect progressively more sizable
amounts of its share of clean water to satisfy the growing domestic water needs of
cities and urban centers, and industries, or governments have to resort to expensive
or, for some countries, unaffordable desalination. In exchange, the agriculture sector
has to turn towards water reuse and towards using water of marginal quality to
cover its requirements for production. Fourth, the sector is facing growing pressure
to begin to adapt to climate change, as more severe droughts and, in some areas,
flash floods and crop-threatening weather anomalies are expected. The skilful
16 W.K. Al-Zubari

integration of weather and climate forecasts in agriculture and rural development


planning remains an unavoidable challenge.
In the GCC countries, agricultural subsidies for wells, fuel, and other inputs,
price support programs and trade protection in some, and a lack of controls on
groundwater extraction, have all drastically increased the area of irrigated land and
are contributing to the depletion of aquifers. Over the last two decades, the net
irrigated areas increased in all GCC countries by 100–300 %. Irrigation water is
often used inefficiently without considering the economic opportunity cost for
potable and urban/industrial purposes. Agriculture contributes less than 2 % to the
GDP in GCC countries but it over-exploits groundwater resources, most of which
are non-renewable fossil groundwater, resulting in their depletion and quality
deterioration due to seawater intrusion and the up-flow of saltwater. In most cases,
no clear “exit strategy” exists to address the question of “What comes after?” when
these resources are exhausted and what the replacement would be or alternative
water resources.
It is worth noting that in Saudi Arabia, the rapid expansion of irrigation areas,
due to generous subsidies, led to nearly a tripling of the volume of water used in the
country, from around 7.4 BCM in 1980 to 20.2 BCM in 1994, falling to 18.3 BCM
in 2000. In 2012, the volume of water used in Saudi Arabia is 17.5 BCM. It is
estimated that about 35 % of non-renewable groundwater resources in Saudi Arabia
were depleted by 1995 (Al-Turbak 2002). This unsustainable use of groundwater
for irrigation is shown by the continuous decline of water table levels, which in
some aquifers dropped more than 200 m during the last two decades. Recognizing
the consequences, the Saudi Government has taken steps since 2000 to reduce
groundwater depletion, encourage efficient irrigation water use, and reduce fiscal
burdens, such as stopping land distribution and reducing input subsidies. It has also
provided incentives for the use of water-saving technologies such as drip irrigation
and soil moisture sensing equipment. As a result, wheat production fell from
4 million tons in 1992 to 1.8 million tons in 2000 (World Bank and AGFUND
2005) and is to be completely phased out by 2016.

1.3 Major Water Issues

1.3.1 Trends and Competition Among Users

It is important to mention that potable water demand in the Arab Region will
increase due to population growth and development of urban areas. This trend will
certainly reduce quantities of water available for agriculture. For the agricultural
sector, developing new sources of water supply to meet its rising demand is
becoming enormously difficult. As indicated above, there are increasingly mounting
demands for the agricultural sector to reduce the pressure on water use and even
utilize marginal water quality for crop irrigation. That leaves the sector with a
1 Status of Water in the Arab Region 17

critical strategic policy thrust: quickening the pace of adopting water demand
management in the sector and facilitating this shift through innovative new policies,
capacity building, research, and training. Agricultural water demand management
must become a priority for policy-makers and water managers, especially since it is
the largest water consumer, and where major and effective savings can be achieved.

1.3.2 Shared Water Resources

More than 65 % of surface water resources originate outside of the Arab Region
(Table 1.1). Furthermore, the region has considerable regional groundwater sys-
tems, both renewable and non-renewable, that extend between neighboring Arab
countries and across the border of the region as a whole (Table 1.2). The majority
of these aquifer systems are non-renewable (or fossil), and are present in relatively
vast areas in the Arab Region, namely in the Sahara Desert and the Arabian
Peninsula, and are shared by many Arab and non-Arab countries. These are located
in a relatively deep geological formation and store significant amounts of water, but
this water has a finite life and quality limitations. In addition to perennial rivers and
large aquifer systems, smaller scale wadis (ephemeral streams) and alluvial aquifers
that cross political boundaries also exist, mainly in the Arabian Peninsula.
The water dependency ratio of some Arab countries, which indicates the per-
centage of total renewable water resources originating outside the country, is
extremely high (Table 1.3). Moreover, relatively low dependency ratio in some
Arab countries does not necessarily mean full control over their water resources, as
in the case of Palestine. Egypt, Syria, and Iraq are obliged to rely almost exclusively
on transboundary water resources emanating from outside its own borders. Shared
water resources thus play a significant role in the stability and development of the
region, creating hydrological, social, and economic relations and interdependencies
between riparian countries, both Arab and non-Arab. The fact that many Arab
countries depend on shared water resources in meeting the majority of their water
supply makes water a political issue, which may strain relations with neighbors and

Table 1.3 Water Country Water dependency ratio (%)


dependency ratio in the Arab
Region (source UN-ESCWA Kuwait 100
2011) Egypt 96.86
Bahrain 96.55
Sudan 76.92
Syria 72.36
Iraq 53.45
Jordan 27.21
Qatar 3.45
Palestine 2.99
Lebanon 0.79
18 W.K. Al-Zubari

could lead to armed conflict. Therefore, cooperation and coordination across


national borders and even across the region is essential if management of shared
water resources in a sustainable manner is to be attained (UN-ESCWA 2011).
Despite the high dependencies on shared surface water resources, a close look at
the legal systems governing shared surface water resources by Arab and non-Arab
countries clearly reveals the fact that most of these systems are partial and
incomplete, and that almost all the shared River Basins, if not all of them, suffer
from the lack of comprehensive international agreements, which constitute a major
challenge facing the region. As more than two-thirds of surface water resources
originate from outside the Arab Region and are managed unilaterally by the riparian
countries, this issue remains a major problem threatening the region’s stability, its
food security, and imposes high levels of uncertainty in water resources planning in
the down-gradient countries (LAS and UNEP 2010). Furthermore, those Arab
countries under occupation deprived of their water resources and the exploitation of
those resources by the occupying countries, represents another major issue in the
region and restricts their population development.
It is vital that the region attempts to further improve cooperation regarding
shared water development and management. A cooperation mechanism must be
developed to help member states reach equitable development and management of
their shared waters. Lack of appropriate cooperation and coordination mechanisms
for shared water resources at regional and inter-regional levels is a source of
concern. This issue is highly influenced by the prevailing relationships and power
asymmetry among states in the region and adjacent countries. Mutual cooperation
and coordination regarding the management of shared surface and groundwater
basins will help to achieve sustainable development within the region by con-
tributing to the rational development, utilization, and conservation of these crucial
resources.

1.3.3 Over-Exploitation of Groundwater Resources

Currently, both renewable and non-renewable groundwater resources in the Arab


Region in general, and in the Arabian Peninsula in particular, are facing critical
conditions. Renewable groundwater in the region is generally present in the form of
shallow alluvial aquifers recharged by the main rivers of the region or directly by
precipitation. This groundwater is being used extensively for domestic and irriga-
tion purposes in most of Arab countries, and at rates that far exceed natural
replenishment rates. This is leading to continuous and sharp declines in ground-
water levels and severe groundwater salinization due to seawater intrusion and
saltwater up-coning from lower strata (UN-ESCWA 1999; FAO 1997). Examples
of groundwater over-exploitation leading to quantity and quality deterioration are
the Saiss Basin in Morocco (ABHS et al. 2007), groundwater in the Gaza Strip
(PWA 2000), and the Amman-Azraq basin in Jordan (Hadidi 2005).
1 Status of Water in the Arab Region 19

In the Arabian Peninsula, where groundwater dependency is the highest and


provides more than 80 % of the total water supply in countries like Saudi Arabia
and Oman, almost all coastal aquifers are experiencing water level decline and
salinization due to seawater intrusion and salt water up-flow due to
over-exploitation. These include Sana’a Basin in Yemen (ACSAD and BGR 2005),
Dammam aquifer in Bahrain (Al-Zubari 2001) and Kuwait (Sayid and Al-Ruwaih
1995), Umm Er Radhuma aquifer in Saudi Arabia (Al-Mahmoud 1987), Al-Dhaid,
Hatta, Al-Ain, and Liwa areas in UAE (Rizk et al. 1997), and Al-Batinah coastal
plain aquifer and Al-Khawd fan in Oman (Macumber et al. 1997).
Furthermore, over-exploitation of groundwater and its depletion in the Arab
Region has had significant environmental impacts. Water levels have declined and
in some cases water salinization had caused the drying of natural springs and the
degradation or destruction of the surrounding habitats and ecosystems, as well as
the loss of the historical and cultural value of those areas. For example, most of the
historical springs in the Palmyra oasis in Syria have dried up, including Afka, where
the historical Kingdom of Zanobia once existed (ACSAD and BGR 2005). Natural
springs in Bahrain and Al Ahsa oasis in Saudi Arabia, most of the oases of the
Egyptian Western Desert, Al Kufrah oasis in Libya, the natural springs used to
irrigate Tozeur city in southern Tunisia, and the South Algerian oases have also
been lost due to excessive pumping and the lowering of groundwater levels.
In arid regions, salinization is very difficult to reverse, as it will require large
amounts of freshwater to push the freshwater-saline interface back. Furthermore,
over-pumping of groundwater is depleting national assets. While the economic
activities based on the extracted water increase GDP in the short term, the
over-exploitation undermines the country’s natural capital or wealth. Calculations
based on available data for four Arab countries show that the value of national
wealth consumed by over-exploiting groundwater is equivalent to as much as 2 %
of GDP (Fig. 1.7; Ruta 2005). Therefore, it is vital that groundwater resources in
Arab countries are carefully planned and managed by observing their natural
recharge rates, so that they can continue to contribute in the sustainability of human
socio-economic development and to support their dependent ecosystems.
Finally, groundwater resources, either renewable or non-renewable, must be
treated as a public-property (or alternatively common-property) resource, with the
state as the guardian or trustee of these resources, and be able to introduce measures

Fig. 1.7 Value of


groundwater depletion in
selected Arab Countries
(source Ruta 2005)
20 W.K. Al-Zubari

to prevent aquifer depletion and pollution. In other words, this means turning well
owners into well users that must apply to the state for water abstraction and use
right. Therefore, there is a high priority to put in place a system of groundwater
abstraction rights that is consistent with the hydrogeological realities. In the case of
non-renewable water resources, it is also important to agree to the level of gov-
ernment to which the decision on mining of aquifer reserves must be referred,
which should be given to the highest possible authority in the country. Moreover,
the value of detailed monitoring of groundwater abstraction and use, and the aquifer
state variables (groundwater levels and quality) response to such abstraction, cannot
be overemphasized. Monitoring of water quality, water levels, and water extraction
in an aquifer is the foundation on which groundwater resource management is
based. This should be carried out by the water resource administration, stakeholder
associations, and individual users. The existence of time-limited permits subject to
initial review will normally stimulate permit holders to provide regular data on
wells. It will be incumbent upon the water resources administration to make
appropriate institutional arrangements, through some form of aquifer database
(databank or datacenter) for the archiving, processing, interpretation, and dissem-
ination of this information.

1.3.4 Water Resources Pollution by Anthropogenic


Activities

In addition to their over-exploitation, groundwater resources and surface water


resources in the Arab Region are being threatened and polluted by numerous point
and non-point sources of pollution generated from anthropogenic, agricultural,
industrial, and domestic activities. As the quality of water resources deteriorates,
either from over-exploitation or direct pollution, its uses diminish, thereby reducing
water supplies, increasing water shortages, and intensifying the problem of water
scarcity in the region. In addition to rendering valuable water resources unusable,
water quality deterioration also increases health risks and damages the environment,
including fragile ecosystem regimes. Examples of such pollution are numerous in
the Arab Region, including the Gaza Strip (PWA 2000), Ras Al-Jabal region in
Tunisia (ACSAD and BGR 2005), Nile River in Egypt and Sudan (CEDARE and
AWC 2004), and upper Littani basin in Lebanon (Assaf and Saadeh 2008).
Currently, several countries in the Arab Region have acknowledged the prob-
lems associated with polluted groundwater and have undertaken initiatives to
protect these valuable resources from further degradation. In many Arab countries,
water quality has received increasingly considerable attention, and efforts to protect
their water resources have been made. Discharge of untreated wastewater is still an
important source of water pollution, however there has been substantial improve-
ment in wastewater treatment in many countries of the region, particularly Jordan
and Egypt. For example, the countries of Egypt, Tunisia, Algeria, and Morocco
1 Status of Water in the Arab Region 21

began to regularly monitor groundwater quality as of the 1990s. Furthermore, in


Egypt, since 1999 the direct disposal of untreated industrial effluents into the Nile
has been legally banned (CEDARE and AWC 2004). In 2001 the Sultanate of
Oman issued the Law on Protection of Sources of Drinking Water from Pollution,
and has since developed drinking water wellfield protection zones for all of its
groundwater basins (Sultanate of Oman 2001).

References

Abderrahman W, Hussain T (2006) Pollution impacts of desalination on ecosystems in the Arabian


Peninsula. In: Amer KM (ed) Policy perspectives for ecosystem and water management in the
Arabian Peninsula. United Nations Educational, Scientific and Cultural Organization/United
Nations University International Network on Water, Environment and Health, Hamilton,
Canada
ABHS (Hydraulic Basin Agency of Sebu), WWF (World Wildlife Fund), MENBO (Mediterranean
Network of Basin Organizations), and ACTeon (2007) Description of the river basin (Sebou,
Morocco): state of the art in the frame of pilot establishment of WFD tools. Report for the EC
contract 044357
Abu-Zeid M, Hamdy A (2004) Water crisis and food security in the Arab world: where we are and
where do we go?. In: Proceedings of the second regional conference on Arab Water 2004:
action plans for integrated development, Cairo, Egypt, pp 12–15
ACSAD (Arab Center for the Study of Arid Zones and Dry Lands) (1997) Water resources and
their uses in the Arab world. In: Proceedings of the first Arab symposium on water resources
and their uses in the Arab World. Organized by ACSAD, Arab Fund for Economic and Social
Development (AFESD), and Kuwait Fund for Arab Economic Development (KFAED), Kuwait
City, Kuwait, 8–10 March, 1997, pp 25–121 (in Arabic)
ACSAD and BGR (2005) Management, protection and sustainable use of groundwater and soil
resources in the Arab Region. Phase II Draft final report. Arab Center for the Studies of Arid
Zones and Dry Lands and the German Federal Institute for Geosciences and Natural Resources,
unpublished
AFED (2010) Arab environment: water, sustainable management of a scarce resource. In:
El-Ashry M, Saab N, Zeitoon B (eds) The Arab forum for environment and development.
Beirut
Al-Hussayen A (2009) Inaugural speech by the minster of water and electricity. Saudi Arabia,
Water and Power Forum, Jiddah, Saudi Arabia
Al-Jamal K, Schiffler M (2009) Desalination opportunities and challenges in the Middle East and
North Africa region. In: Jagannathan NV, Mohamed AS and Kremer A (eds) Water in the Arab
World: Management Perspectives and Innovations. The International Bank of Reconstruction
and Development/The World Bank, Middle East and North Africa Region, Washington, DC,
Retrieved from: http://siteresources.worldbank.org/INTMENA/Resources/Water_Arab_
World_full.pdf
Al-Mahmoud MJ (1987) Hydrogeology of Al-Hassa Oasis. MSc Thesis, King Fahd University of
Petroleum and Minerals, Dhahran
Al-Mooji Y (2004) Nahr el-Kabir Basin (Lebanon, Syria). consensus building for integrated
management: main characteristics of the basin. Retrieved from: http://webworld.unesco.org/
water/wwap/pccp/zaragoza/basins/nahr_el_kabir/nahr_el_kabir.pdf
Al-Turbak A (2002) Water in the Kingdom of Saudi Arabia: policies and challenges. In:
Proceedings of symposium on the future vision of the Saudi economy. 19–23 October, Riyadh,
Saudi Arabia
22 W.K. Al-Zubari

Al-Zubari WK (2001) Impacts of groundwater over-exploitation on desertification of soils in


Bahrain—A case study (1956–1992). In: Regional aquifer systems in arid zones –managing
non-renewable resources proceedings of international conference, Tripoli, Libya, 20–24
November 1999, pp. 311–22. IHP-V Technical Documents in Hydrology No. 42. United
Nations Educational, Scientific and Cultural Organization, Paris, Retrieved from: http://
unesdoc.unesco.org/images/0012/001270/127080e.pdf
Assaf H, Saadeh M (2008) Assessing water quality management options in the upper Litani Basin,
Lebanon, using an integrated GIS-based decision support system. Environ Model Softw
23:1327–1337, Retrieved from: www.weap21.org/downloads/WQLitani.pdf
AWC (Arab Water Council) (2009) MENA/Arab countries regional document. 5th World Water
Forum, 20–22 March, Istanbul (www.arabwatercouncil.org/administrator/Modules/CMS/
Technical%20Report%205_Arab-MENA_Regional_Document_WWF5.pdf)
Balaban M (2008) Desalination in Maghreb. In: EUROMED conference on desalination strategies
in South Mediterranean Countries
Bushnak A (2010) Desalination. In: El-Ashry M, Saab N, Zeitoon B (eds) Arab environment:
water, sustainable management of a scarce resource. The Arab Forum for Environment and
Development, Beirut
CEDARE and AWC (Centre for Environment and Development for the Arab Region; Arab Water
Council) (2004) State of the water report in the Arab region. CEDARE, Cairo
CIA (Central Intelligence Agency) (2003) CIA world factbook. New York, Retrieved from: https://
www.cia.gov/library/publications/the-world-factbook/fields/print_2093.html
Darwish MA, Al-Najem NM, Lior N (2010) Towards sustainable seawater desalination in the Gulf
area. Desalination 235:58–87, Retrieved from: http://www.seas.upenn.edu/*lior/lior%
20papers/Towards%20sustainable%20seawater%20desalting%20in%20the%20Gulf%20area%
20-published.pdf
Doumani FM (2008) Climate change adaptation in the water sector in the Middle East and North
Africa region: a review of main issues. PAP/RAC Workshop, Sardinia, May 19–21
Droubi A, Jnad I, Al Sibaii M (2006) ACSAD activity in the field of water resources management
and rainwater harvesting. Arab Center for the Study of Arid Zones and Dry Lands (ACSAD),
Damascus, Retrieved from: http://gwadi.org/sites/gwadi.org/files/RegionalDroubi.pdf
Falkenmark M (1989) The massive water scarcity now threatening Africa-Why isn’t it being
addressed. Ambio 18:112–118
FAO (2013) AQUASTAT database. Food and Agriculture Organization, Rome, Retrieved from:
www.fao.org/nr/water/aquastat/data/query/index.html
FAO Food and Agriculture Organization (1997) Irrigation in the near East region in figures. Food
and Agriculture Organization. Water Report 9. Rome, Retrieved from: www.fao.org/docrep/
w4356e/w4356e00.HTM
FAOSTAT (2008) FAOSTAT—FAO statistical databases. Food and Agriculture Organization of
the United Nations, Rome, Retrieved from: http://faostat.fao.org/site/357/default.aspx
GWI (Global Water Intelligence) (2010) Water market Middle East 2010. Global Water
Intelligence, 2010
Hadidi K (2005) Groundwater management in the Azraq basin. In: Proceedings of the Arab center
for the studies of arid zones and dry lands and the German Federal Institute for Geosciences
and Natural Resources workshop on groundwater and soil protection in the Arab region.
Amman, Jordan, 27–30 June
Hamoda MF (2004) Water strategies and potential of water reuse in the South Mediterranean
countries. Desalination 165:31–41
IFAD (2009) Fighting water scarcity in the Arab countries. International Fund for Agricultural
Development (IFAD), Rome
Khater AR (2010) Regional technical report on the impacts of climate change on groundwater in the
Arab region. In: Technical Document. UNESCO Cairo Office, Cairo. Retrieved from: www.unesco.
org/new/fileadmin/MULTIMEDIA/FIELD/Cairo/pdf/SC/Impacts%20of%20Climate%20Change
%20on%20Groundwater%20in%20the%20Arab%20Region-2%20Aug%202010-a.pdf
1 Status of Water in the Arab Region 23

Kibaro’lu A (2004) Water for sustainable development in the Euphrates-Tigris River Basin. In:
Proceedings of the 2nd Asia Pacific Association of Hydrology and Water Resources, 5–8
July, Singapore, Retrieved from: www.gap.metu.edu.tr/html/yayinlar/waterforsustainable
AKibaroglu.pdf
Ksia C (2009) Institutional and legal issues in managing shared water resources: the Arab region’s
experience. Department of Environment, Housing and Sustainable Development, the League of
Arab States, Cairo
LAS and UNEP (2010) Environmental outlook for the Arab Region. League of the Arab States
(Cairo) and UNEP (Nairobi), Retrieved from: www.unep.org/dewa/westasia/eoar
Macumber PG, Niwas JM, Al-Abadi A, Seneviratne R (1997) A new isotopic water line for
Northern Oman. In: The WSTA third Gulf water conference: towards efficient utilization of
water resources in the Gulf. Muscat, Oman, 8–13 March
MED WWR WG (2007) Mediterranean wastewater reuse report. In: Mediterranean wastewater
reuse working group—joint Mediterranean EUWI/WFD process, Retrieved from: http://www.
semide.net/media_server/files/c/e/Final%20report.1.pdf
Nile Basin Initiative (n.d.) Key facts about the Nile Basin. Entebbe, Uganda, Retrieved from:
www.nilebasin.org/newsite/index.php?option=com_content&view=article&id=135%3Akey-
facts-about-the-nile-basin&catid=75%3Astats&Itemid=68&lang=en
OMVS (Organization for the Development of the Senegal River) (2003) Senegal river basin,
Guinea, Mali, Mauritania, Senegal. In: UNESCO, water for people, water for life: the united
nations world water development report. UNESCO World Water Assessment Programme:
Paris and New York, Retrieved from: http://webworld.unesco.org/water/wwap/case_studies/
senegal_river/senegal_river.pdf
PWA (Palestinian Water Authority) (2000) Gaza coastal aquifer management program. Integrated
Aquifer Management Plan Task 3, Vol 1. Ramallah, Palestine
Qadir M, Sharma BR, Bruggeman A, Choukr-Allah R, Karajeh F (2007) Non-conventional water
resources and opportunities for water augmentation to achieve food security in water scarce
countries. Agric Water Manag 87:2–22
Rizk ZS, Alsharhan AS, Shindu S (1997) Evaluation of groundwater resources of United Arab
Emirates. In: The WSTA third Gulf water conference: towards efficient utilization of water
resources in the Gulf. Muscat, Oman, 8–13 March
Ruta G (2005) Deep wells and shallow savings: the economic aspect of groundwater depletion in
MENA countries. In: Background paper to making the most of scarcity: accountability for
better water management results in the Middle East and North Africa. World Bank,
Washington
Sayid SAS, Al-Ruwaih F (1995) Relationship among hydraulic characteristics of the Dammam
aquifer and wells in Kuwait. Hydrogeol J 3(1):57–70
Scheumann W, Sagsen I, Tereci E (2011) Orontes river basin: downstream challenges and
prospects for cooperation. In: Kibaroglu A, Scheumann A, Kramer W (eds) Turkey’s water
policy. Springer, Berlin, pp 301–312, Retrieved from: www.caee.utexas.edu/prof/mckinney/
ce397/Topics/Orontes/20-Orontes%20river%20basin.pdf
Shahin M (1989) Review and assessment of water resources in the Arab region. Water Int 14
(4):206–219
Sokona Y, Diallo OS (eds) (2008) The North-Western aquifer system: joint management of a
trans-border water basin. Synthesis Collection 1. Sahara and Sahel Observatory, Tunis,
Retrieved from: www.oss-online.org/pdf/synth-sass_En.pdf
Sultanate of Oman (2001) Royal decree no. 115/2001: issuing law on protection of sources of
potable water from pollution. Government of the Sultanate of Oman, Retrieved from: www.
pdo.co.om/hseforcontractors/blocks/documentation/docs/laws/decree_115_2001.pdf
UN Urbanization Prospects (2014) Population division: world urbanization prospects, the 2014
revision. United Nations Department of Economic and Social Affairs, Retrieved from: http://
esa.un.org/unpd/wup/
24 W.K. Al-Zubari

UNDESA (United Nations Department of Economic and Social Affairs) (2012) World population
prospects, the 2012 revision. New York, Retrieved from: http://esa.un.org/unpd/wpp/Excel-
Data/population.htm
UNDP (2013) Water governance in the Arab region: managing scarcity and securing the future.
UNDP Regional Bureau of Arab States, Retrieved from: http://www.arabstates.undp.org/
content/rbas/en/home/library/huma_development/water-governance-in-the-arab-region.html
UN-ESCWA (1999) Groundwater quality control and conservation in the ESCWA region.
E/ESCWA/ENR/1999/1. United Nations Economic and Social Commission for West Asia,
Beirut
UN-ESCWA (2007) ESCWA Water development report 2: state of water resources in the ESCWA
region. E/ESCWA/SDPD/2007/6. United Nations Economic and Social Commission for
Western Asia, New York, Retrieved from: http://www.escwa.un.org/information/publications/
edit/upload/sdpd-07-6-e.pdf
UN-ESCWA (2011) ESCWA Water development report 4: national capacities for the management
of shared water resources in ESCWA member countries. UNESCWA document
E/ESCWA/SDPD/2011/4. United Nations Economic and Social Commission for West Asia,
Beirut
UN-ESCWA and BGR (United Nations Economic and Social Commission for Western Asia;
Bundesanstalt fur Geowissenschaften und Rohstoffe) (2013) Inventory of shared water
resources in Western Asia. UN-ESCWA document number E/ESCWA/SDPD/2013/Inventory,
ESCWA, Beirut
Varis O, Tortajada C (2009) Water governance in the MENA region: policies and institutions,
extended report. In: An International conference on water governance in the MENA region.
Dead Sea, 7–11 June, Jordan
World Bank and AGFUND (2005) A water sector assessment report on the countries of the
cooperation council of the Arab States of the Gulf. In: Report no. 32539-MNA, water,
environment, social and rural development department, Middle East and North Africa. The
World Bank, Washington
World Bank (2007) Making the most of scarcity: accountability for better water management in the
Middle East and North Africa. MENA Development Report. Washington, DC, Retrieved from:
http://siteresources.worldbank.org/INTMNAREGTOPWATRES/Resources/Making_the_Most_
of_Scarcity.pdf
World Bank, AWC (Arab Water Council), IDB (Islamic Development Bank), and ICBA
(International Center for Biosaline Agriculture) (2011) Water reuse in the Arab World: from
principle to practice: voices from the field. In: A summary of proceedings: expert consultation
on wastewater management in the Arab World, 22–24 May, Dubai, Retrieved from: http://
water.worldbank.org/sites/water.worldbank.org/files/publication/Water-Reuse-Arab-World-
From-Principle%20-Practice.pdf
World Bank and BNWP (Bank–Netherlands Water Partnership) (2004) Seawater and brackish
water desalination in the Middle East, North Africa, and Central Asia: a review of key issues
and experience in six countries. In: Working paper no. 33515. World Bank, Washington
Chapter 2
State of Energy in the Arab Region

Kishan Khoday and Stephen Gitonga

Abstract The Arab Region is going through one of the most intense periods of
transformation in its history, with a convergence of political, security, economic,
social, and environmental drivers of change. More than any other region, energy
has defined the nature of development in the Arab world for decades, and it has
again arisen as a key factor in the region’s emerging agenda for change. Of par-
ticular importance in today’s development debates is the role of energy in crafting a
more sustainable and equitable model of development, with the nexus to poverty
reduction, food and water security, and social vulnerability gaining strong attention.


Keywords Energy and development Arab Region Middle east   Sustainable
  
energy Energy access Renewable energy Energy efficiency

2.1 Energy and Development in the Arab Region

The Arab Region is going through one of the most intense periods of transformation
in its history, with a convergence of political, security, economic, social, and
environmental drivers of change. More than any other region, energy has defined
the nature of development in the Arab world for decades, and it has again arisen as a
key factor in the region’s emerging agenda for change (UNDP 2012). Of particular
importance in today’s development debates on the future of the region is the role of
energy in crafting a more equitable model of development, with links to food and

K. Khoday (&)
Climate Change, DRR and Resilience, United Nations Development Programme (UNDP),
Regional Hub for Arab States, Cairo, Egypt
e-mail: kishan.khoday@undp.org
S. Gitonga
Regional Sustainable Energy Specialist, United Nations Development Programme (UNDP),
Regional Hub for Arab States, Amman, Jordan
e-mail: stephen.gitonga@undp.org

© Springer International Publishing AG 2017 25


K. Amer et al. (eds.), The Water, Energy, and Food Security Nexus
in the Arab Region, Water Security in a New World,
DOI 10.1007/978-3-319-48408-2_2
26 K. Khoday and S. Gitonga

water security gaining strong attention, in particular with regards the state of the
poor and vulnerable in society (Fattouh and El Katiri 2012).
While the energy nexus to human security is a global priority, as expressed in the
emerging post-2015 framework and the Sustainable Development Goals (UN
2014), it takes on even greater importance in the Arab world, given its status as the
world’s most food-import dependent, water scarce region. Furthermore, while big
picture governance reforms come into focus in many of the region’s transitional
countries, an important part of this must be on enhancing systems of energy gov-
ernance in particular (Khoday 2012).
The energy sector is central to the nature of development in the Arab Region,
making up approximately 40 % of the region’s GDP, upwards of 60 % of the
world’s oil reserves, and approximately 30 % of the world’s gas reserves (Fattouh
and El Katiri 2012). The region’s fossil fuel engines of growth are driven by
the group of high-income, energy-rich economies of the Arab Gulf, and with record
oil prices in recent years; this has translated into a surge of public revenue streams
and related development initiatives. With rapid growth in energy-intensive
urban-industrial growth, these countries are now seeing some of the world’s fast-
est rates of per capita energy consumption growth. Energy-intensive water desali-
nation facilities in particular are a source of rising energy demand.
The resulting drain on oil and gas reserves is now a top concern in most such
countries, given their criticality for future export revenues and fiscal stability.
Hundreds of billions of dollars in export revenue could be foregone in coming
decades if current rates of local energy consumption growth continue unmitigated.
As such, a strategic focus has arisen in energy-rich Arab economies to expand
energy efficiency and renewable energy to conserve oil resources for future export
revenues, and secure future social development dividends for the next generation.
Meanwhile for the region’s energy-import dependent countries, the nexus of
energy to human security is an even greater challenge. They make up the majority
of Arab countries and host 40 % of the region’s poor who still lack adequate access
to modern energy services (Khoday 2012). Further, while unequal access to energy
services is a problem in the region’s low- and middle-income countries, the nexus
of energy and human insecurity is even more challenged when considering the
plight of refugees and forced migrants, now among the most vulnerable in Arab
society (UNDG 2013; Bailey and Barbelet 2014). Sustainable access to energy will
be vital to address the plight of refugees and the strains placed on host communities.
For energy-import dependent countries, another challenge for addressing
inequality and social development goals has been to remain resilient to volatile
global energy prices. For instance, high global energy prices in recent years have
been a drain on public revenues, and converge with some of the world’s highest
levels of domestic energy subsidies creating serious challenges to fiscal sustain-
ability (Sdralevich et al. 2014). The recent dramatic drop in global energy prices is
welcome news for import-dependent countries in this regard, but measures are
needed to increase levels of energy security given a possible return to higher prices
in the future.
2 State of Energy in the Arab Region 27

The growing costs of remaining dependent on energy imports and high energy
subsidies bring risks to public welfare and social coherence, at a time when the
region is already going through an intense period of systemic change. For the poor
and excluded in society, higher expectations now exist for more rationale and
accountable use of energy as a public good, preventing the squandering of this vital
asset, and preserving its future value as a source of development dividends and
inter-generational equity. From a macro perspective, the Arab Region’s overall
energy challenges can be defined by the convergence of three key challenges:
(i) unprecedented demand growth with some of the world’s fastest rates of per
capita energy consumption growth, (ii) declining energy reserves, fragility in
security of energy supply and escalating costs from energy-import dependency, and
(iii) rising inequality between those who benefit from energy use, and the poor and
excluded in society, who rely on energy for various development goals including
food and water security.
The region now stands at a cross-roads. In confronting the downside risks that
energy insecurity brings to development, the region can engage opportunities to
shift to a more sustainable energy future. New strategies towards this end have been
endorsed in recent years, both at the regional level through the League of Arab
States (LAS), and at the more local level through a wave of new national energy
strategies and targets in many countries of the region. As noted further below, three
key opportunities now top this agenda:
(i) expand the share of renewable energy in the region’s supply mix,
(ii) scale-up energy efficiency and reduce energy intensity of growth,
(iii) expand access to energy to combat poverty and inequality.
Engaging these opportunities can help overcome development challenges
including the area of food and water insecurity, and set the stage for a transition to a
more sustainable energy future in the region (CEM and IRENA 2014). These three
opportunities also contribute to Sustainable Development Goal (SDG) number 7 on
Energy as part of the emerging post-2015 sustainable development agenda.

2.2 Renewable Energy Solutions

Among the most strategic priorities for a more sustainable energy future in the
region revolves around the potentials from solar and wind energy development. The
Arab Region leads the world with the planet’s highest levels of solar radiation,
alongside strong wind resources (El Katiri 2014a). However, most Arab countries
have yet to develop these potentials in a way that would lead to real shifts in energy
security. Relative to global trends, the region under-performs in finance and
investment levels, especially in the field of private investment in renewable energy
(Frankfurt School-UNEP Centre/BNEF 2014). The capacity of operational
renewable energy supply remains low, at only about 7 % of the overall energy mix,
28 K. Khoday and S. Gitonga

Fig. 2.1 Renewable energy targets and implementation gaps (source RCREEE 2015)

and even this is largely from traditional hydropower (RCREEE 2015). This despite
ambitious policy targets as seen in Fig. 2.1.
However, some signs of change are starting to emerge. A key feature in recent
years has been rapid cost reductions of solar technology components in the global
market, along with recognition within the region of renewable energy’s potential as
part of a high-tech knowledge economy (Ernst and Young 2013, p. 5). Globally,
such factors have increased the competitiveness of solar power in recent years, and
with the types of energy insecurity pressures noted above, this is also starting to
translate in the Arab Region. Today there are more than 2500 MW of renewable
energy projects under construction which will more than double the total installed
capacity which currently stands at 1974 MW (RCREEE 2015).
To ensure such initiatives have longer-term traction and can be scaled up
through greater levels of investment, a need exists to enhance two elements of the
current renewable energy system: policy frameworks, and institutional capacities to
implement and deliver results. In regards to the former element, a Regional Strategy
for Renewable Energy (2013–2030) has been endorsed for the first time by the
LAS, recognizing energy insecurity as a common concern across Arab society and
expressing the desire for member States to engage in common solutions. While
holding importance in itself, equally or perhaps more importantly, have been a
series of national policies in recent years meant to set the stage for scaled-up
investments and results.
2 State of Energy in the Arab Region 29

As shown by the Arab Future Energy Index (AFEX) (RCREEE 2015, pp. 5–7),
Morocco has an ambitious vision for renewable energy, with the National Energy
Strategy (2012–2020) seeking to install 2000 MW each of wind, photovoltaic
(PV) and concentrated solar power (CSP), with the goal of supplying 42 % of
energy mix by 2020. A comparably ambitious goal is in neighboring Algeria, where
a National Program for Renewable Energy and Energy Efficiency (2011–2030) has
set a 2030 goal of 7200 MW of CSP, 2800 MW of PV, and 2000 MW of wind
capacity. If achieved this would increase the share of renewable energy in the
national mix from about 6 % today to 40 % by 2030. In Tunisia, large investments
are planned in wind and PV capacities meant to take the country to 30 % of
installed capacity by 2030 while in Egypt, a National Renewable Energy Strategy
2020 sets a goal of 7200 MW of wind alongside 1320 MW each of PV and CSP,
with an overall target of 20 % of energy mix from renewables by 2020. Other
notable targets include a goal of 15 % renewables in the energy mix in Yemen by
2025, 12 % in Lebanon by 2015, 11 % in Sudan by 2031, 10 % in Jordan, and
Palestine by 2020. Palestine has the highest electricity prices in the region which
combines with issues of conflict and poverty to provide strong motivation for
exploring new renewable energy potentials.
While setting national targets is an important baseline factor, achieving policy
goals is a much broader task, especially in an environment like the Arab Region,
defined by a strong legacy of support to and reliance on fossil fuels. Overcoming
this legacy and achieving ambitious policy goals for renewable expansion requires a
set of complementary factors to achieve these new targets and overcome market
barriers to change. Policy measures go beyond targets to include changes to the
nature of energy subsidies, measures for attracting investments, project develop-
ment, feed-in-tariffs, systems for metering, etc.
While the emergence of ambitious policy frameworks is a welcome develop-
ment, it is also clear that rates of improvement will need to expand rapidly if such
goals are to be achieved in coming years. An implementation gap exists in many
countries, often related to the nature of institutions. Achieving renewable energy
goals and targets will require institutional capacities to deliver on new policy
frameworks, and address transparency, accountability, and participation in
decision-making processes.
Independent energy regulators for example can play an important role in
ensuring predictability to investors, setting tariffs, facilitating network connections,
and establishing competitive frameworks for private participation, etc. But across
the Arab Region, only five countries (Algeria, Egypt, Jordan, Palestine, and Sudan)
have such institutions, with varying degrees of effectiveness and capacity
(RCREEE 2015, pp. 11–14). They are important for many functions including
establishing baseline assessments on which to build renewable energy investments
and gauge progress, and for allocating land and related resources needed for
renewable energy ventures.
A related innovation in some countries around the world has been the emergence
of dedicated renewable energy agencies taking a lead in government for design of
specific regulations, procedures for project development, and initiation, etc. But
30 K. Khoday and S. Gitonga

whether dedicated agencies exist or not, the key issue is whether the broad insti-
tutional system is able to produce an enabling environment that attracts and facil-
itates greater investments, coordinates stakeholders and local partners, leads
research efforts, and solar/wind resource assessments, etc. Algeria, Jordan, and
Morocco for example have all produced a Solar Atlas, while Egypt, Jordan,
Lebanon, Morocco, and Tunisia have completed Wind Resource Assessments
(RCREEE 2015, pp. 11–14).
In summary, taking into consideration trends in policy frameworks, institutional
capacities and other factors like private investment, the region is seeing the grounds
for a future expansion of renewable energy. Figure 2.2 summarizes overall findings
from the 2015 Arab Future Energy Index (AFEX) report, the first native Arab index
dedicated to monitoring and analyzing sustainable energy competitiveness in the
Arab Region. North Africa leads the way in terms of aggregate improvement across
various renewable energy indicators with Algeria, Egypt, Morocco, and Tunisia all
ranking high on a number of policy, institutional, and other factors. Countries have
succeeded in pursuing a renewable energy agenda with a set of credible policies and
targets. Looking forward, cooperation is now needed in coming years to establish
the necessary institutional capacities and regulatory frameworks that can help
countries achieve the scaling up of investments needed to achieve these ambitious
goals.

Fig. 2.2 2015 Arab Future Energy Index (AFEX) renewable energy rankings (source RCREEE
2015)
2 State of Energy in the Arab Region 31

2.3 Energy Efficiency

While expansion of renewable energy is a top priority in terms of energy supply, an


even greater source of gains is on the demand side of the ledger, with energy
efficiency holding major potentials to reduce incremental energy needs from urban
expansion and growth. With the Arab Region now experiencing some of the
world’s fastest rates of energy intensity growth (Elasha 2010), there is massive
untapped potential for enhancing conservation measures across various sectors and
ensure that the critical energy resources that do exist are used in the most effective
and productive way.
Expanding renewable energy in the overall mix as elaborated above, while
simultaneously reducing the energy intensity footprint in key growth sectors can
together help bring about a sustainable energy economy in the region, and help set
the region on a stronger footing. As a highly urbanized society and with a good part
of the region’s oil export revenues now being re-invested into construction and
urban infrastructure expansion, energy efficiency improvements can now be found
in urban related sectors.
Key sectors now in focus for pursuing energy efficiency solutions in the region
include, but are not limited to: power generation, seawater desalination, buildings
and air conditioning, heavy industry, and transport, all of which are now major
sources of energy demand growth and are placing serious strains on energy secu-
rity. Given the economic challenges facing many countries in the region, energy
efficiency solutions also hold the benefit of being cost-effective and in some cases,
new technology solutions are in effect no-cost exercises with relatively short
periods of return on initial investments.
Trends in recent years show a growing gap between the region’s average elec-
tricity consumption growth, registered at 7.9 % per year in 2010, and the region’s
GDP growth rate, at about 3.9 % GDP growth the same year (RCREEE 2015, p. 3)
(Fig. 2.3).
The energy intensity gap is an important obstacle to energy security. A goal of
energy efficiency measures would be to reduce this gap and to start decoupling
economic growth from energy intensity growth. To bring energy efficiency solu-
tions to bear and at the scale needed to shift the trajectory of strong energy intensity
growth, a need exists as in the case of renewables to enhance two key elements of
the energy system—policy frameworks for energy efficiency, and institutional
capacities to implement and achieve results. At the regional level, Arab Guidelines
on Energy Efficiency have been enacted by the LAS. At the national level, a
number of countries have enacted energy strategies with energy efficiency com-
ponents—Algeria, Jordan, Iraq, Morocco, Palestine, Saudi Arabia, Tunisia, UAE,
and Yemen—while Lebanon, Jordan, Tunisia, Egypt, Palestine, and Sudan have
launched official National Energy Efficiency Action Plans (NEEAPs) meant to align
locally with the LAS Energy Efficiency Guidelines and set the stage for local
actions (RCREEE 2015, p. 5). In 2011, Lebanon was the first to adopt a NEEAP,
while Tunisia has been recognized as having the most comprehensive energy
32 K. Khoday and S. Gitonga

Fig. 2.3 Rising energy intensity of growth (source RCREEE 2015)

efficiency policy framework. Examples of 2020 targets include 5 % reduced energy


intensity in residential and industrial sectors in Lebanon, 6 % in residential and 3 %
in utilities in Palestine, and 18 % reduced energy intensity in the residential sector
in Sudan by 2020 (RCREEE 2015, p. 7).
In addition to target setting measures in national strategies, targeted regulatory
frameworks are needed to set the right incentives for shifting end-use consumer
behavior, enabling the role of market-based approaches and public-private part-
nerships, and enabling public finance for research and innovation. Many countries
now have in place framework laws and/or specific regulatory framework in this
regard. This includes a draft Energy Conservation Law in Lebanon, and existing
Energy Conservation Laws in Algeria, Morocco, and Tunisia, and specific regu-
lations for key sectors. Figure 2.4 summarizes overall findings from the 2015 Arab
Future Energy Index (AFEX) report on Energy Efficiency.
With regards to the important focus on industry standards, Tunisia is regarded as
a best practice example in the region owing in part to its comprehensive system of
regulations for industry including mandatory energy audits, energy management
systems, and reporting requirements. It also has an ambitious system for voluntary
agreements (VA’s) between Government and industry setting specific targets for
industrial actors, with close to 500 VAs now in place (RCREEE 2015). The gov-
ernment supports certain costs, such as for baseline energy audits and assessments
of potentials, as well as dedicated credit lines to support retrofitting of equipment,
etc. Likewise, Algeria has put in place mandatory measures for industry and a
Top-Industry initiative to support a shift to green solutions.
In the key buildings sector for example, regulations in many countries now target
standards for insulation and windows, and the best way to integrate efficient
electrical systems into overall design and layout of buildings. In addition,
2 State of Energy in the Arab Region 33

Fig. 2.4 2015 Arab Future Energy Index (AFEX) energy efficiency rankings (source RCREEE
2015)

opportunities exist for efficiency gains related to standards and labeling systems for
consumer appliances like air conditioners and refrigerators, which together account
for the vast majority of household energy consumption. In some parts of the Arab
Region, air conditioning for example in commercial, residential, and other facilities
accounts for the majority of overall national energy consumption.
Minimum energy standards for air conditioners and refrigerators now exist in
Algeria, Egypt, Jordan, and Tunisia for example, while Lebanon has voluntary
energy standards. Shifting to energy efficiency lighting is another example of a
low-cost and effective solution, with annual energy saving potentials between
11.4 % in Yemen and 8.3 % in Morocco to 7.7 % each year in Sudan and 7.8 % in
Jordan. Potential cost savings exist of over $1.4 billion each year for the region as a
whole (RCREEE 2015).
In addition to setting the overall policy environment for energy conservation,
such laws also establish financial incentives and dedicated funds to support tran-
sition to energy conserving technology; including the National Fund for Energy
Management in Algeria supported by taxes on natural gas, the Energy Development
Fund in Morocco, the Energy Efficiency Fund in Jordan, and the National Energy
Fund in Tunisia funded by taxes from vehicle and air conditioner sales.
Another critical element in the fight for energy efficiency across the region,
complementing policies, laws, and regulatory frameworks, is the rise of dedicated
institutions mandated to oversee the expansion of such measures and ensure
34 K. Khoday and S. Gitonga

compliance with mandatory standards. A dedicated energy efficiency agency has


served as indispensable in many country contexts, not only for leading formulation
and design of policies and regulations, but in leading coordination of stakeholders
and sector entities needed for effective implementation.
Almost half of Arab countries have such agencies, with varying capacities. One
example is the National Agency for Promotion and Rationalization of Energy Use
in Algeria (APRUE), which leads compliance and implementation of flagship ini-
tiatives related to energy efficiency in key sectors such as lighting and industry, and
manages the National Energy Management Fund. The Lebanese Center for Energy
Conservation (LCEC) led preparation of its NEEAP and various pilot projects over
the past several years. Capacities to implement national policy and regulatory
frameworks vary, with different degrees of progress on energy efficiency in
buildings, lighting, heavy industry, etc. In neighboring Morocco, efforts are led by
the National Agency for Development of Renewable Energy and Energy Efficiency
(ADEREE) while Tunisia has its National Agency for Energy Management
(ANME) with a number of affiliates at a research center for specific sectoral efforts
(RCREEE 2015, pp. 11–13). These and other examples stand as positive experi-
ences from which others in the region can draw lessons and models for scaling up
actions to achieve regional and national targets.

2.4 Energy Access for the Poor

As expressed above, positive trends do exist in some countries of the region for
expanding sustainable energy solutions. But expanding capacity does not neces-
sarily translate into social equity. Without proactive policy approaches, the benefits
of sustainable energy expansion could bypass the poor in society, who stand to
benefit greatly from access to modern, low-cost energy sources. Challenges of
inequality and social exclusion are now at the top of policy debates in the Arab
Region, and greater attention is now being placed on the role of energy in com-
bating poverty.
Despite significant hydrocarbon wealth in the Arab Region, there are a number
of countries like Egypt, Iraq, the Occupied Palestinian Territories, Sudan, Somalia,
Syria, and Yemen with low rates of energy access (El-Katiri 2014b, p. 15). A clear
priority is to expand access for poor and displaced communities that lack sus-
tainable access to energy services, often owing to lack of access to power grid
systems. For poor and displaced populations, lack of access to energy, such as
electricity and household cooking gas, results in various development constraints
such as impacts on income generation opportunities, issues of food and water
security, and family health. This is a particular concern for rural livelihoods where
energy for irrigation and drinking water pumps is often a limiting factor in pro-
ductivity, income, and various factors of well-being. “By means of water, we give
life to everything” reads a passage in the Koran. Water is recognized as a basic
necessity for life and is indispensable for a range of human development factors.
2 State of Energy in the Arab Region 35

2.4.1 Decentralized Energy Solutions

A key challenge in this regard is a lack of access to the national grid systems for
remote communities. In Sudan and Yemen for example, only 40 % of the popu-
lation has regular access to the grid, while rural communities in Egypt face similar
challenges (RCREEE 2015, p. 15). Given the ambitious targets noted above for
energy transitions, a trend across the region has been to scale-up investments into
utility-scale capacity. While universal grid access is a positive aspiration (Glemarec
et al. 2012), a sole focus on utility-scale capacity has resulted in less attention to
off-grid, decentralized renewable energy solutions that can serve as important
solutions to local needs, or as a ‘bridge’ for communities to expand energy access
while awaiting broader grid connectivity efforts. Examples include waste-to-energy
units to convert agriculture farm waste into energy, solar water pumps for drinking
water and irrigation, roof-top solar for schools and clinics, solar powered computers
for education, solar street lighting in remote communities, etc.
Off-grid, decentralized solutions can support a nexus approach that links energy
to other development sectors such as health, education, agriculture, and water
including the creation of new income opportunities. Benefits to local communities
include participatory approaches to achieving national energy targets, including
local business development potentials, with women entrepreneurship potentials in
particular. Another benefit is capture of local resource energy sources that may not
otherwise be utilized in large-scale national schemes, while also serving as a base
for local innovation of standard technologies to meet specific ecological and cli-
matic conditions and social needs.
But the ability to develop and deploy bottom-up, decentralized renewable energy
solutions faces many challenges (Waissbein et al. 2013). There is a general lack of
awareness on availability and types of solutions that can be engaged at the local
level, lack of efforts to pilot and scale-up solutions through adapted and tailored
business models and policy incentives, lack of financing options to achieve
affordability criteria, and lack of capacity among local governments on ways to
support sustainable energy goals in poor and excluded communities.
Even where communities are connected to grids, many suffer supply interrup-
tions and uncertain power outages, owing to problems in the grid system and supply
side uncertainties. A key trend in the Arab Region in this regard has been for
communities and businesses to rely on inefficient diesel generators as a back-up
during power cuts. Potentials exist to expand use of small-scale hybrid solar-diesel
units as a means of achieving the power back-up goal while reducing costs and the
health impacts of diesel use. This has potential particularly in key sectors such as
agriculture and tourism which are of benefit to rural communities. Market assess-
ments are needed to identify economically feasible business models for diesel to
solar retrofitting across the Arab Region, as well as new public-private partnerships
and pilot projects to explore and replicate good business models.
36 K. Khoday and S. Gitonga

2.4.2 Displaced Communities

An important segment of the poor needing special attention are the unprecedented
numbers of internally-displaced persons, forced migrants, and refugees seen across
the Arab Region, now among the most vulnerable communities in the region and
the world (UNDG 2013). In crisis contexts such as Iraq, Libya, Palestine, Somalia,
Syria, and Yemen, destruction or disruption of energy capacities is now a critical
issue facing the general population, with dire situations for both the displaced
communities themselves, and host communities like Lebanon and Jordan, where the
influx of displaced populations places extra pressures on already vulnerable states
of energy insecurity.
Recent years have seen one of the most dramatic rises in the Arab Region’s
history of forced migrants, internally displaced persons, and cross-border refugees.
Driven by conflicts, millions now face dire situations and are in need of critical
supplies of food, water, and energy. Furthermore, with many such situations
evolving from short-term to protracted states of displacement, response measures
need to factor in long-term sustainable solutions for host communities (Bailey and
Barbelet 2014). Factoring energy into response options has become an important
priority in contexts like Iraq, Libya, Somalia, Syria, Yemen, and in the protracted
situation in the Occupied Palestinian Territories. It is also critical in neighboring
Arab countries who currently host the vast majority of Syrian refugees and forced
migrants.
Lebanon is an important example. It now hosts over 36 % of all refugees that
have left Syria over the past four years, a major global contribution by a relatively
small country. Some estimates put the share of Syrian refugees as reaching 25 % of
the overall population in Lebanon, the highest relative share of any country hosting
refugees during the current crisis. This brings various risks to Lebanon’s devel-
opment pathway, including energy.
With the vast majority of refugees living in cities and towns, rather than camps,
pre-existing challenges of national and local energy insecurity are being exacer-
bated by the surge of extra energy demand from the large population influx. This
includes rising prices for basic fuel supply and impact of regular power cuts. These
new pressures create risks for the ability of both Lebanese host communities and
Syrian refugees alike. As part of Lebanon’s Response Plan to the Syria Crisis, a
series of activities have been launched by the government, with support of the
UNDP and Germany, to expand the use of renewable energy solutions for basic
energy needs such as household cooking, heating, and community lighting needs in
communities in North and Bekaa regions. These are among the poorest areas of the
country, while also the area where a majority of refugees reside.
In neighboring Jordan, the country relies on energy imports for about 97 % of
needs, and with 80 % of Syrian refugees and forced migrants residing in standard
cities and towns across the country rather than in refugee camps, overall electricity
consumption in the country has risen significantly since the onset of the Syrian
crisis. Estimates put increased energy demand from 4926 GWh in 2009 to
2 State of Energy in the Arab Region 37

6560 GWh in 2014, partly owing to the influx of refugees and forced migrants
(Jordan Response Platform 2015). In a country already energy poor and facing an
energy crisis, the extra demand has created an added element to the Kingdom’s
energy goals. To address this challenge, the Jordan Response Plan to the Syria
Crisis identifies a series of renewable energy and energy efficiency measures to help
expand energy access in refugee host communities, relieve socio-economic pres-
sure, and build resilience of both communities and the nation (Jordan Response
Platform 2015). This stands as an important example of the use of sustainable
energy for crisis recovery within the broader Regional Refugee and Resilience Plan
(3RP) (2015–2016).
Sustainable energy solutions can be a key factor to ensure an effective bridge
between humanitarian and development responses spanning short- and long-term
solutions. Expanding use of sustainable energy solutions like solar power to meet
the basic needs of refugees and displaced communities is important and urgent not
only from a humanitarian perspective, but also from a development perspective, as
it helps reduce pressures on host countries’ already strained energy systems and
helps achieve goals of inclusion and sustainability under SDG 7.

2.4.3 Energy Subsidy Reforms

In addition to technology solutions, a key issue in focus in the Arab Region has
been on provision of energy subsidies to reduce the net cost of electricity and fuel
access to consumers. The Arab Region is host to some of the world’s most
extensive systems of energy subsidies, an important factor in achieving relatively
good levels of energy access (Fattouh and El Katiri 2012). Energy subsidies have
been a core component of the social contract for development across the region for
many years, and a basic part of the social safety net for the poor in particular
(Sdralevich et al. 2014). But concerns over fiscal sustainability have now emerged.
Among the largest element of overall public subsidy systems in most countries in
the region, are at the core of widening deficits and public debts. For many countries,
achieving a more effective and efficient system of energy access is an existential
matter with risks to the very solvency of the country.
The risks posed by expensive energy subsidies in the region comes at a time
when greater attention has been placed globally in recent years on ways that energy
subsidies are actually an inefficient means to reach energy access goals. Often the
benefits of low energy prices accrue disproportionately to those who can already
afford access such as high-income households and industries. Globally, in devel-
oping countries the bottom 40 % of the population in terms of income distribution
receives only 20 % of fuel subsidies. This disproportionate impact holds true in
many Arab Region countries as well. Energy subsidy reform would thus benefit not
only expansion of sustainable energy options, but also better targeting of energy
subsidies to the poor.
38 K. Khoday and S. Gitonga

Meanwhile overall financial outlays for subsidy initiatives can be potentially


reallocated in ways that hold greater impact and benefit for the poor through
alternate schemes such as indirect or direct cash transfers to poor households, better
targeting those in need with proportionately greater benefit to overall goals to
combat energy poverty. Meanwhile, subsidies that lower prices can lead to over-
consumption, especially by those who can afford to do so, creating disincentives for
more rational and efficient use of increasingly scarce resources. Energy subsidy
reforms can increase the efficiency and effectiveness of fiscal strategies to achieve
economic, social and environmental goals (IEA et al. 2010, p. 3).
As a growing drain on public budgets and as a barrier to more efficient use of
energy, many countries in the region have commenced energy subsidy reform
efforts in recent times. Examples include Jordan, Egypt, Palestine, and Morocco,
with varying degrees of success and very specific contextual and political-economy
factors determining the causes for reforms in these countries and the possible
outcomes of reform policies. In many cases the outcomes of the energy subsidy
reform exercise are still far from clear. Thus, while evidence builds on the value of
energy subsidy reforms in the Arab Region, any such reforms also hold great
uncertainties and risks in terms of expected results and unintended consequences
for overall national development goals (Fattouh and El Katiri 2012). In the current
climate of change and uncertainty across the region, great care should therefore be
given when considering or designing energy subsidy reforms.

2.4.4 Global Partnerships for Energy Access

Last but not least, while various challenges exist for expanding energy access to the
poor within the region, Arab countries also stand as important global partners to
expand access to energy for the poor globally. Countries in the Arab Gulf like Saudi
Arabia, Kuwait, Qatar, and the UAE for example, are now some of the largest new
sources of official development assistance (ODA) and outward direct investment
(ODI) globally, with their expanding financial cooperation in less-developed
countries of Africa, Asia, and the Americas also now starting to connect with goals
of sustainable energy. Far from being a new partner in global development, the
Arab Gulf has a strong history of south-south cooperation (UN 2012). From 1973 to
2008, Arab ODA stood at about 1.5 % of GNI on average, double the target of
0.7 % set by the United Nations for developed countries of the Organization for
Economic Co-operation and Development (OECD) (Denney and Wild 2011). Saudi
Arabia, Kuwait, and the UAE all started dedicated bilateral development agencies
in the 1970s, all of which have channelled ODA to countries in Arab, African, and
Asian regions over the years (OECD 2011, pp. 211–218; Al-Yahya and Fustier
2010).
They have also played an increasingly important role in policy dialogues on
sustainable energy, water, and other development issues in LDCs. Recent years
have seen Qatar hosting the 2012 UN Climate Change Convention Conference of
2 State of Energy in the Arab Region 39

the Parties and the 2014 UN South-South Expo with a core focus on the
energy-food-water nexus, and UAE hosting of the 2014 Abu Dhabi Ascent on
Climate Change, annual gatherings of the World Future Energy Summit, and
hosting the International Renewable Energy Agency (IRENA). As the first global
multilateral agency to be hosted in the Arab Region, the presence of IRENA also
marks an important milestone as to the role the Arab Region can play in supporting
global capacities and institutions for energy access goals. But the Gulf is also
starting to move from dialogues to action, by also investing in global solutions.
Sustainable energy is now being integrated into the Arab Gulf’s ODA and ODI
strategies. The Saudi-based Islamic Development Bank (IsDB), the major multi-
lateral channel of Arab Gulf ODA, is for the first time launching an Energy Strategy
to guide its ODA, with an important focus on ways to achieve sustainable energy
and low-carbon goals in recipient countries in Africa, Asia, and the Arab Region
(IsDB 2013). For rural energy access programs, upfront long-term public invest-
ment is essential to developing the various functional capacities needed to scale-up
energy access to the step where market transformation can occur. At the national
level, capacities need to be strengthened in clean energy policy formulation, in the
proper targeting and reform of energy subsidies, and in providing a more enabling
environment for investment.
With the energy sector among the largest components of IsDB investments
globally, sustainable energy access for the poor has been receiving growing
attention with a new wave of grant and concessional lending for rural electrification,
renewable energy, energy efficiency, and other measures towards the goal of
poverty reduction in LDCs. Touching on the above, this has also seen an important
emphasis on cooperation to expand bottom-up decentralized solutions. IsDB also
took an important step at the 20th Anniversary of UN Conference on Sustainable
Development (Rio+20) by joining a new large-scale global endeavour to scale-up
low-carbon, sustainable transport options together with other international financial
institutions (AfDB et al. 2013). Another example is the OPEC Fund for
International Development (OFID) which has scaled-up assistance to developing
countries in recent years as part of a $1 billion commitment to support sustainable
energy access for the poor. This is meant to achieve the goals of the Rio+20
summit, help implementation the 2008 Riyadh Declaration on Energy for
Development, and its commitments to achieve the UN Sustainable Energy for All
process.
In parallel to Arab multilateral channels, progress is also seen at the level of
bilateral ODA. The UAE for example has formally integrated renewable energy
goals into its outward ODA strategies. This now includes more than $400 million
of assistance from the Abu Dhabi Fund for Development to help expand renewable
energy in developing countries, an initiative undertaken in concert with IRENA
(Ministry of International Cooperation 2013). These and other examples can be a
powerful force for the common goal of supporting new sustainable energy solutions
in LDCs around the world, with Arab ODA providers expected to scale up support
40 K. Khoday and S. Gitonga

in the future. New partnerships among development partners of Asia, Africa, and
Arab Regions could well reconnect age-old routes of trade, cooperation and
innovation, and to forge a ‘new energy silk road’ of development solutions.

References

AfDB, ADB, CAF, EBRD, EIB, IADB, ISDB, WB (2013) Progress report (2012–2013) of the
Multilateral Development Bank (MDB) working group of sustainable transport. African
Development Bank, Development Bank of Latin America, European Bank for reconstruction
and Development, European Investment Bank, Inter-American Development Bank, Islamic
Development Bank and the World Bank, Washington, DC
Al-Yahya K, Fustier N (2010) Saudi Arabia as a humanitarian donor: high potential, little
institutionalization. Global Public Policy Institute (GPPi) Berlin, Germany
Bailey B, Barbelet V (2014) Towards a resilience based response to the Syrian refugee crisis: a
critical review of vulnerability criteria and frameworks. UNDP, May 2014 Amman, Jordan
CEM and IRENA (2014) The socio-economic benefit of solar and wind energy. IRENA, Abu
Dhabi, UAE
Denney L, Wild L (2011) Arab donors: implications for future development cooperation.
European Development Co-operation to 2020, Policy Brief—ODI, London
Elasha BO (2010) Mapping of climate change threats and human development impacts in the Arab
region. UNDP Arab Human Development Report Research Paper Series, UNDP Regional
Bureau for Arab States, Arab Human Development Report, Research Paper Series, New York,
NY
El Katiri L (2014a) A roadmap for renewable energy in the Middle East and North Africa. Oxford
Institute for Energy Studies, Oxford, UK
El Katiri L (2014b) The energy poverty nexus in the Middle East and North Africa, OPEC Energy
Review
Ernst & Young (2013) MENA cleantech survey. Dubai, UAE
Fattouh B, El Katiri L (2012) Energy and subsidies in the Arab world. UNDP Arab Human
Development Report Research Paper Series, Regional Bureau for Arab States, New York
Frankfurt School-UNEP Centre/BNEF (2014) Global trends in renewable energy investment 2014.
Global Trends Report, Frankfurt, Germany
Glemarec Y, Rickerson W, Waissbein O (2012) Transforming on-grid renewable energy markets:
a review of UNDP-GEF support for feed-in tarrifs and related price and market-access
instruments. UNDP New York, NY
IEA, OPEC, OECD, WORLD BANK (2010) Analysis of the scope of energy subsidies and
suggestions for the G20 initiative. Prepared for Submission to the G-20 Summit Meeting,
Toronto 3
IsDB (2013) Approach to Improved Access to Electricity. Islamic Development Bank, Jeddah,
Saudi Arabia
Jordan Response Platform (2015) Comprehensive vulnerability assessment: energy sub-chapters.
Ministry of Planning and International Cooperation, Amman, Jordan
Khoday K (2012) Sustainable development as freedom: energy, environment and the Arab
transformation. Poverty in Focus Volume 23, at 11–13, International Policy Center for
Inclusive Growth, Brasilia
Ministry of International Cooperation (2013) Strategy for UAE international cooperation. MIC,
Abu Dhabi, UAE
OECD (2011) Development cooperation report 2011: 50th anniversary edition. OECD Publishing.
doi:10.1787/dcr-2011-en
2 State of Energy in the Arab Region 41

RCREEE (2015) Arab Future Energy Index (AFEX). Regional Centre for Renewable Energy and
Energy Efficiency (RCREEE), Cairo, Egypt
Sdralevich C, Sab R, Zouhar Y, Albertin G (2014) Subsidy reform in the Middle East and North
Africa. IMF, Washington, DC
UN (2012) South-south cooperation for development: review of progress made in implementing
the Buenos Aires Plan of Action, the new directions strategy for South-South cooperation, the
Nairobi outcome document and the decisions of the high-level committee. GA Document
SSC/17/1 (April 2012), New York, NY
UN (2014) Open working group proposal for sustainable development goals. Open Working
Group of the General Assembly on Sustainable Development Goals. UN Document A/68/970,
available at http://undocs.org/A/68/970
UNDG (2013) Position paper: a resilience-based development response to the Syria crisis.
Regional UN Development Group, UNDP, New York, NY
UNDP (2012) Arab development challenges report 2011: towards the developmental states in the
Arab region (2012). UNDP Regional Centre in Cairo, Cairo, Egypt
Waissbein O, Glemarec Y, Bayraktar H, Schmidt TS (2013) Derisking renewable energy
investment: a framework to support policymakers in selecting public instruments to promote
renewable energy investment in developing countries. UNDP, New York, NY
Chapter 3
Status of Food Security in the Arab Region

Tariq Moosa Alzadjali

Abstract Most Arab countries have made progress in enhancing their food secu-
rity situation during the last twelve years. Domestic production of food com-
modities increased considerably and is expected to continue to rise. The production
of grains in the Arab Region was estimated at about 55.5 million tons in 2013. In
the same year, the total number of livestock was estimated at about 345 million
head, comprised mostly of poor producing animals. The Arab Region produced
around 4.0 million tons of poultry meat by the end of 2013 and this is expected to
increase in the future. The poultry industry in the Arab Region relies heavily on
imported inputs and is consequently unsustainable. The production of the fisheries
sector in the region was estimated at 4.3 million tons of fish in 2013. The contri-
bution of aquaculture (about 25.7 %) to the total fish production in the Arab Region
is modest compared to its global contribution. Some Arab countries have over 30 %
of the population living in conditions classified as “poor”. About 308.9 million tons
of major food commodities were available for consumers in Arab countries in 2013,
an increase of 12.7 million tons over 2012. Thirteen Arab countries were classified
as “low” on the Global Hunger Index for 2014, and one country was classified
“moderate”. Three countries were classified “serious”, while another three were
classified “alarming”. Arab countries are unlikely to achieve high ratios of food self
sufficiency; however, they can maintain and improve the current ratios of food self
sufficiency.

  
Keywords Arab countries Food security Food availability Food utilization 
 
Subsidized food production Global hunger index Strategic food reserve

T.M. Alzadjali (&)


Arab Organization for Agricultural Development, Khartoum, Sudan
e-mail: dr.tariqalzadjali@gmail.com

© Springer International Publishing AG 2017 43


K. Amer et al. (eds.), The Water, Energy, and Food Security Nexus
in the Arab Region, Water Security in a New World,
DOI 10.1007/978-3-319-48408-2_3
44 T.M. Alzadjali

3.1 Introduction

According to the Food and Agriculture Organization of the United Nations (FAO),
“food security exists when all people, at all times, have physical and economic
access to sufficient, safe and nutritious food to meet their dietary needs and food
preferences for an active and healthy life.” In this context, food security in the Arab
Region is considered to be one of the biggest challenges facing the region, which
requires targeted food security interventions by governments both indirectly
through subsidized food production, or directly through certain subsidized food
commodities especially wheat and sugar. These interventions which vary in extent
and magnitude from one country to another are the primary method for reducing
food insecurity.
Food security in the Arab Region is affected by several factors such as unstable
political and social environments which suppress economic growth, war and civil
strife, trade deficit, natural resource constraints, inadequate education, limited rural
development, poor human resource development, poor health services, a limited
role of women; especially the rural women; flood and droughts, locust infestation,
and the absence of good governance.

3.2 State of Arab Food Security

Most Arab countries have made progress in enhancing their food security during the
last twelve years (2002–2013) excluding countries in which there was war or civil
conflict, or those that experienced serious economic and social challenges.
According to the Arab Organization for Agricultural Development (AOAD), the
region’s population of about 370 million in 2013 is predicted to reach about
550 million in 2030. This is a source of serious on-going threat to the status of food
security.

3.2.1 Food Availability

Despite the fact that Arab countries are heavily dependent on imported food
“especially cereal” domestic production of food commodities such as cereals, oil-
seeds, sugar crops, dates, and animal products increased considerably and is
expected to continue to rise as predicted by the Emergency Program of Arab Food
Security (EPAFS).
3 Status of Food Security in the Arab Region 45

3.2.1.1 Domestic Food Production

The production of grain in the Arab Region was estimated at about 55.5 mil-
lion tons in 2013, of which wheat constituted about 49.5 % (Table 3.1). The
average productivity of grain crops grown in the rain-fed areas is low compared to
the global average, probably due to the fluctuation of rainfall and the traditional
system of production characterized by the limited use of agricultural production
technologies.
Arab production of refined sugar is estimated at about 3.37 million tons
extracted from 24.11 million tons of sugar cane, and 13.67 million tons of sugar
beets. The productivity of sugar cane and sugar beets produced from irrigated
agriculture in the Arab Region is estimated at 108.4 and 51.0 tons/ha, respectively.
The productivity of sugar cane in the Arab Region is higher than the global average,
while the productivity of sugar beets is about (92.6 %) of the global average.
Production of vegetable oil amounted to (2.2) million tones extracted from about
(8.4) million tons of oil crops (olives, sunflower, cotton seeds, sesame, and peanut).
There are many challenges and constraints facing the vegetable oil sector such as
the limited use of production technologies, the absence of supplementary irrigation
in rain-fed agriculture, and inefficient implementation of the value chain concept.
Annually, the region produces about 109 million tons of vegetables, fruit, dates, and
potatoes (Table 3.2), which meets domestic consumption with some quantities
exported (AOAD 2013a).
The Arab Region produces significant quantities of animal and fisheries products
(Table 3.3). The total number of livestock was estimated at about 345 million head
in 2013, of which cattle and buffalo, sheep, goats, and camels constitute about
17.16, 52.11, 26.03 and 4.7 % respectively. The contribution of this large number
of livestock is estimated at about 5.0 million tons of red meat and 26.6 million tons

Table 3.1 Grain production Commodity Average 2002–2011 2012 2013


in the Arab Region (million
tons) (source AOAD 2013b) Grain 52.7 52.7 55.5
Wheat 25.2 26.1 27.5
Barley 5.8 5.4 6.6
Corn 7.5 8.9 8.8
Rice 6.5 6.6 6.7
Sorghum 6.6 3.9 5.6

Table 3.2 Production of Commodity Average 2002–2011 2012 2013


food in the Arab Region
(million tons) (source AOAD Refined sugar 3.06 3.24 3.37
2013b) Vegetable oil 1.73 2.02 2.18
Vegetables 47.90 53.43 53.10
Fruits 29.10 33.10 34.26
Dates 5.30 5.95 6.37
Potatoes 10.40 14.17 15.00
46 T.M. Alzadjali

Table 3.3 Animal and Commodity Average 2002– 2012 2013


fishery production in the Arab 2011
Region (million tons) (source
AOAD 2013b, c) All meats 7.4 8.6 9.0
Red meat 4.4 4.9 5.0
Poultry meat 3.0 3.7 4.0
Fish 3.7 4.2 4.3
Eggs 1.4 1.8 1.9
Milk and dairy 24.4 26.1 26.6
products

of milk and milk products. This production is not commensurate with the large
number of livestock in the Arab Region. The reason there is low productivity is that
most livestock in the region are kept under traditional pastoral systems with little or
no health care, and poor management.
The Arab Region produced around 4.0 million tons of poultry meat by the end of
2013 and this is expected to increase significantly in the coming few years. Five
countries produce about 72.4 % of the total Arab production of poultry meat, led by
Egypt (23.2 %), then Morocco (16.7 %), Saudi Arabia (15.9 %), Algeria (9.9 %),
and Jordan (6.7 %). The poultry industry in the Arab Region relies on the import of
most inputs, especially maize in which Arab countries suffer a huge deficit.
Therefore, the high self-sufficiency ratios of poultry meat and eggs achieved in the
Arab Region are misleading and unsustainable, and require joint Arab efforts to
develop maize and sorghum production vertically and horizontally.
The production of the fisheries sector in the region was estimated at 4.3 million
tons of fish in 2013 (Table 3.3).
Aquaculture contributed about 25.7 % of this total. The contribution of aqua-
culture to the total fish production in the Arab Region is considered to be modest
compared to the global contribution of 43.8 %. The progress witnessed in the
fisheries sector is not commensurate with available fisheries resources, and the
potential of aquaculture development in Arab countries.
The main constraints facing the development of the fisheries sector in the Arab
Region include:
– poor investments directed to the development of the fisheries sector;
– depletion of fish stocks and over-exploitation in many areas;
– lack of data and information about fish stocks in the Arab Region;
– prevalence of traditional fishing methods among small-scale fishermen and
small size of inland fisheries;
– inefficient value chain.
Considerable investment in aquaculture is planned in some Arab counties mainly
Egypt, Oman, and Saudi Arabia targeting an increase in production by more than
one million tons in the next ten years. The Arab Organization for Agricultural
3 Status of Food Security in the Arab Region 47

Development (AOAD) is preparing a strategy for sustainable development of Arab


Aquaculture in close consultation with and contribution from all Arab countries.

3.2.1.2 Food Export and Import

Since domestic production of major food commodities is not sufficient to make food
available to Arab people at all times and meet their dietary needs and food pref-
erences, food imports become essential for the existence of food security in the
region.
Fish, rice, wheat and flour, dates, fruit, vegetables, and potatoes are the main
export food commodities of the Arab Region. Re-exports of food commodities such
as refined sugar, wheat flour, and vegetable oils are a significant activity in some
Arab countries. The amount of grain exported by Arab countries has declined in the
past few years, especially wheat and flour and rice. Rice exports declined as a result
of the decrease in cultivated areas and export restrictions in Egypt, the main pro-
ducer of rice in the Arab Region (Table 3.4). Arab countries export about
9.3 million tons of fruit, vegetables, and potatoes which are expected to increase in
the near future as a consequence of the Ukraine crisis which might result in an
increase of Arab exports to Russia. The main exporting Arab countries for potatoes
are Egypt, Lebanon, Syria, Morocco, and Saudi Arabia. The main exporting Arab
countries of vegetables are Syria, Morocco, Jordan, Egypt, and Saudi Arabia,

Table 3.4 Exports of food commodities from Arab countries (source AOAD 2013b)
Commodity Average 2002–2011 2012 2013
Quantity Value Quantity Value Quantity Value
(ton) ($million) (ton) ($ million) (ton) ($ million)
Grain and 2,490,000 775.5 1,870,040 1038.11 1,866,430 1080.62
flour
Wheat and 1,231,900 274.8 670,020 304.62 691,640 320.31
flour
Corn 65,700 17.7 107,240 44.31 123,080 53.12
Rice 852,700 373.1 583,310 498.30 624,780 510.96
Barley 114,400 18.3 100,907 51.80 96,450 74.74
Sorghum 44,800 11.0 34,270 9.33 50,050 13.29
Potatoes 704,100 193.7 774,350 276.20 850,470 339.12
Legumes 258,900 162.0 200 360 193.86 237,340 197.18
Vegetables 2,985,700 1686.1 3,897,680 2780.98 3,915,500 2,988.85
Fruit 2,466,900 1,693.3 4,212,000 3291.0 4,524,630 3694.30
Refined 1,195,200 582.1 1,561,890 1334.06 1,582,350 1660.48
sugar
Vegetable 806,400 1203.9 965,930 1533.88 992,250 1553.00
oil
48 T.M. Alzadjali

Table 3.5 Exports of animal and fish products in the Arab Region (value: $million) (source
AOAD 2013b)
Commodity Unit Average 2002– 2012 2013
2011
Quantity Value Quantity Value Quantity Value
Live cattle Head 1,538,780 33.7 128,040 35.31 129,600 33.6
Sheep and goats Head 5,335,500 463.2 5,790,380 724.85 6,503,400 788.2
Red meat Ton 46,500 102.2 173,410 306.31 219,130 397.2
White meat Ton 55,800 97.0 70,260 160.91 72,890 173.4
Fish Ton 731,400 1875.9 851,460 2662.58 912,400 2956.4
Table eggs Ton 41,600 100.8 84,910 205.86 99,620 228.8
Milk and dairy Ton 2,289,100 851.6 2,959,430 1481.69 3,177,140 1699.7
products

whereas the main exporters of fruits are Egypt, Morocco, Syria, United Arab
Emirates (UAE), Lebanon, and Tunisia (AOAD 2013a).
Exports of live sheep and goats, cattle, red meat, and poultry meat are expected
to increase (Table 3.5), if more attention is given to the development of animal
health and quarantine services. Greater emphasis is required on the preparation of a
strategy and action plan for the development of the livestock industry especially in
Sudan and Somalia.
At present, the inadequate animal health services and quarantine facilities, weak
or limited infrastructure required for live animal export, and high prices of live
animals above global prices especially in Sudan, are the main obstacles hindering
the growth of live animal exports which, at present, do not exceed 6.6 million head
of sheep, goats, and cattle.
Poultry meat exports of Arab countries is increasing significantly, amounting to
about 73 thousand tons in 2013, an increase of 17 thousand tons compared to
average exports of the period (2002–2011). This trend is expected to continue in the
coming years after the implementation of large projects in Saudi Arabia and other
Arab countries. More than 3 million tons of milk and dairy products of a value of
about $1.7 billion were exported from Arab countries in 2013.
The share of Arab countries in the total world imports of grain, sugar, and
vegetable oils is considered to be very significant both in quantity and value of
those food imports. The imports of major food commodities of crop origin are about
102 million tons per year with a value of about $45.2 billion (Table 3.6). The value
of grain imports is about 42.3 % of the total value of major food imports, followed
by milk (9.9 %), refined sugar (9.4 %), vegetable oil (9.1 %), red meat (7.2 %), and
white meat (5.8 %).
Arab countries import 12.6 million head of live cattle, sheep and goats,
3.46 million tons of red meat and poultry meat, 3.5 billion eggs, and 11.6 mil-
lion tons of milk and dairy products. In addition, they import about 950.2 thousand
tons of fish. The total value of these imports is estimated at $18.4 billion, of which
3 Status of Food Security in the Arab Region 49

Table 3.6 Arab countries imports of food commodities of crops origin (source AOAD 2013b)
Commodity Average 2002–2011 2012 2013
Quantity Value Quantity Value Quantity Value
(thousand (million (thousand (million (thousand (million
tons) US$) tons) US$) tons) US$)
Wheat and 27438.3 6728.3 38259.5 11824.8 40794.7 12517.6
flour
Corn 13365.8 2715.2 16216.7 5167.4 16554.7 5749.5
Rice 3710.8 2346.1 4951.2 4176.1 5071.6 4268.3
Barley 8715.3 1706.8 11851.1 3529.9 11985.4 3565.7
Sorghum 727.0 198.8 710.8 256.7 388.5 157.1
Grains and 55419.9 20994.8 72700.5 25249.3 75421.5 26144.3
flour
Potatoes 699.6 341.1 842.7 502.1 840.2 496.6
Legumes 1162.6 688.6 1371.2 1160.0 1245.6 1098.8
Vegetables 2456.5 1060.8 3862.6 2083.8 3912.6 2101.8
Fruits 3527.7 1914.1 5516.3 3870.5 5639.5 3852.3
Sugar (raw) 8046.1 3098.8 9707.1 5814.6 10498.7 5859.9
Vegetable 4053.5 3642.4 4292.3 5241.4 4393.2 5641.2
oil

Table 3.7 Imports of animal and fish products in the Arab Region ($million) (source AOAD
2013b)
Commodity Unit Average 2002–2011 2012 2013
group Value Quantity Value Quantity Value Quantity
Cattle Head 653,100 393.0 564,700 673.3 596,200 702.6
Sheep and goats Head 12,280,800 974.4 14,041,600 1236.6 12,036,700 1205.4
Red meat Ton 804,900 1958.2 1,293,100 4318.4 1,392,500 4445.3
Poultry meat Ton 1,170,000 1762.6 1,995,200 3754.5 2,065,800 3588.43
Fish Ton 623,300 999.8 918,100 2024.1 950,200 2093.8
Egg Ton 92,900 158.2 163,800 259.6 192,300 268.2
Milk and dairy Ton 11,631,200 4197.1 11,785,300 5703.4 64,911,000 6143.9
products

milk and dairy products make up 33.3 %, while red meat, poultry meat, fish, live
sheep and goats, and live cattle make 24.1, 19.5, 11.3, 6.5 and 3.8 %, respectively
(Table 3.7).
The value of foreign trade in major food commodities is $99.9 billion with
exports accounting for 19.44 % of this amount. The percentage coverage of exports
to imports is about (24.13 %). The value of inter-Arab agricultural trade rose to
about $26.2 billion in 2013, an increase of about (24 %) over 2012.
50 T.M. Alzadjali

3.2.2 Access to Food

Access to food is mainly a function of the factors which affect the demand side,
notably income, poverty, food prices, and production costs. There is wide income
disparity among Arab countries and accordingly the proportion of family income
directed to food consumption ranges between 35 and 71 %.
In spite of the relative high income levels in many Arab countries, there are some
Arab countries with a percentage of poor exceeding 30 % of the population (The
World Bank 2014).
The food price index in the Arab world rose to 192.4 % in 2013, increasing by
22.3 % compared to 2012, while at the global level it rose from 216.3 to 232.3 %,
as shown in Table 3.8 (Sudan Trade Point 2014). Food price spikes in Egypt and
Sudan since 2011 are due to unfavorable political and economic conditions.
The rise in food prices in Arab countries is a result of a combination of internal
and external factors. Internal factors include high costs of agricultural production;
low agricultural productivity; unsustainable and ineffective agricultural policies;
low stocks of cereals, sugar, oil seeds, and vegetable oils; in addition to poor or
weak infrastructures such as roads, ports, and storage facilities. External factors are
more important for most Arab countries since they are mostly import dependent for
essential food commodity groups such as grains, cereals, sugar, and vegetable oil.
Therefore, any international crisis related to food commodities production, uti-
lization, and marketing will have a direct impact on the prices of these commodities
in Arab countries. These international crises include drought in the key

Table 3.8 Food price index Country 2011 2012 2013


in some Arab countries. Base
years (2002–2004) = 100 Jordan 143.4 150.0 155.6
(source data from FAO Bahrain 135.5 136.0 136.5
website, Sudan Trade Point Tunisia 126.5 136.1 –
2014) Algeria 166.5 188.0 193.6
Saudi Arabia 153.0 131.0 –
Sudan 169.6 230.0 314.0
Iraq 138.4 147.4 –
Oman 161.4 164.9 168.3
Kuwait 177.7 187.6 185.0
Palestine 148.1 151.3 152.9
Qatar 125.2 126.5 136.9
Lebanon 132.4 138.0 142.0
Egypt 260.2 284.1 319.4
Mauritania 223.9 233.5 243.8
Yemen 133.8 147.5 161.3
Arab countriesa 159.7 170.1 192.4
The World 203.7 216.3 232.3
a
Average of Arab countries
3 Status of Food Security in the Arab Region 51

grain-producing regions, devaluation of the American dollar, low international


stocks of cereals and oil seeds, rapidly fluctuating oil prices, increased feed stocks
used in the production of bio-fuels, in addition to conflicts and wars.
In the face of high and volatile food prices, most Arab countries have taken
various measures including direct support for consumers and farmers, provision of
production inputs, implementation of income generating projects, special food
security programs, social safety nets, and the reduction of taxes and duties on
imported food.

3.2.3 Stability of Food Supplies

The recent food crises were manifested in uncertainty and volatile prices in agri-
cultural markets, uncertain commitment of several key agricultural exporters during
food crises, and failure or limited ability of the private sector in meeting public
needs. The lessons learnt from these crises forced most Arab countries to take
serious measures towards stabilizing food supply for their people in order to
maintain a minimum level of food security and deal with the increasing incidences
of food emergencies. These measures include the setting up or the development of
strategic food reserves for both food emergencies and price stabilization.
The storage capacity for major food commodities such as cereal, sugar, and
vegetable oil are sufficient for periods ranging from two to four months in countries
such as Algeria, Egypt, Lebanon, Tunisia, and Yemen, while they range between
five and six months in Jordan, Morocco, Oman, and Qatar, and range between seven
to twelve months in Bahrain, Saudi Arabia, and Syria. Countries such as Bahrain,
Jordan, Oman, Qatar, and Saudi Arabia are about to increase their storage capacity of
strategic food reserves to become sufficient for more than twelve months.
Table 3.9 indicates that wheat stocks in nine Arab countries (Jordan, Bahrain,
Tunisia, Sudan, Iraq, Oman, Lebanon, Egypt, and Yemen) increased by 7.6 % in
2012 reaching up to 19.7 million tons. Rice stocks increased significantly by
79.4 % mainly due to export restrictions in Arab producing countries. Vegetable oil
stocks showed a slight increase (0.26 %), while sorghum and refined sugar
decreased by 9.4 % and 7.4 %, respectively. Wheat stocks represent about 86.3 %
of the total stocks of main food commodities in these Arab countries.

Table 3.9 Food stocks in Commodity World Arab countries


some Arab countries 2011/2012 2010 2011 2012
(thousand tons) (source
AOAD 2012) Wheat 183,100 15400.3 18304.5 19692.8
Rice 161,700 1091.7 1010.8 1813.6
Sorghum – 430.3 468.1 423.9
Vegetable Oil 37,800 270.4 269.7 270.4
Refined Sugar 66,100 693.5 666.6 617.5
52 T.M. Alzadjali

3.2.4 Food Utilization

3.2.4.1 Food Consumption

About 308.9 million tons of major food commodities were available for consumers
in Arab countries in 2013, increasing by 12.7 million tons compared to 2012
(Table 3.10). The availability of fruit, sugar, potatoes, and vegetables in 2013
increased by about 4.2, 12.2, 6.7 and 4.7 %, respectively compared to 2012.
Arab per capita major food consumption has improved over the years and it
increased in 2013 for most commodities with different ratios except for corn,
legumes, and milk and dairy products which decreased by 0.2, 5.2 and 0.6 %,
respectively (Table 3.11).
According to FAO (2010), the average daily per capita calorie, protein and fat
consumption in Arab countries was estimated at 2894.5 kcal, 83 and 72.7 g,
respectively. World Health Organization (WHO 2014) forecasted that by 2025 the
average daily per capita calorie consumption would reach about (2940) kilocalories
in the world and about 3440 kcal in the industrialized countries. The daily per
capita consumption of protein and fats at the world level is estimated by FAO to be
79.3 and 81.8 g, respectively.
It is expected that the average daily per capita consumption of calories, protein,
and fat in Arab countries as a whole will improve with the improvement of political,
or economical and social conditions in some Arab countries such as Djibouti, Iraq,
Libya, Somalia, Sudan, Syria, and Yemen.

Table 3.10 Food available for consumption in the Arab Region (thousand tons) (source AOAD
2013b)
Commodity Average 2002–2011 2012 2013
Wheat and flour 51702.4 63754.0 67648.0
Corn 20786.5 25087.0 25254.0
Rice 9267.1 10997.0 11224.0
Barley 14406.1 18746.0 19844.0
Other grain 9717 4969 5112
Potatoes 10416.8 14244.0 15203.0
Legumes 2368.2 2522.8 2417.4
Vegetables 47469.4 52684.0 55144.0
Fruit 29658.5 34006.0 35372.0
Refined sugar 8967.5 10583.0 11875.0
Vegetable oil 4921.2 5830.1 6167.5
Red meat 5175.4 6020.2 6156.8
Poultry meat 4082.2 5630.3 6023.6
Fish 3604.0 4251.9 4321.1
Table eggs 1494.6 1882.9 2000.7
Milk and dairy products 33686.6 34973.0 35119.0
3 Status of Food Security in the Arab Region 53

Table 3.11 Per capita food Commodity Average 2012 2013


consumption in the Arab 2002–2011
Region (gram/day) (source
AOAD 2013b) Wheat and flour 429 471 496
Corn 172 186 185
Rice 77 81 82
Barley 119 127 135
Potatoes 86 105 111
Legumes 20 19 18
Vegetables 394 395 404
Fruit 246 234 259
Refined sugar 74 79 87
Oils and fat 41 39 41
Red meat 43 45 45
White meat 39 42 44
Fish 30 31 32
Eggs 12 14 15
Milk and dairy products 279 259 257

3.2.4.2 State of Hunger

Despite the improvement in the state of hunger in Arab countries over the last two
decades, thirteen countries are classified low on the Global Hunger Index (GHI)1
for 2014 as they scored less than 5, with one country scoring slightly above at 5.9
which was therefore classified moderate (Table 3.12). Three countries were clas-
sified serious, as they scored between 11.9 and 19.5, and another three countries
were classified alarming as they scored between 23.4 and 29.5 (IFPRI 2014).
The state of hunger of the countries classified serious and alarming requires
immediate review of policies and approval of action plans to ensure that the situ-
ation will not worsen, and that their state of hunger will improve in a defined and
definite time frame. An intervention in the context of Arab League might be
required to shift the hunger situation in these countries. With the exception of Iraq,
these countries are not financially capable of implementing an effective action plan
on their own without the support of other Arab countries and the international
community.

1
GHI is multidimensional statistical tool used to describe the state of hunger. The index was
adopted and developed by International Food Policy Research Institute (IFPRI) to rank countries
on 100 points scale with (0) being the best score and (100) being the worst. The GHI combines
three equally weighted indicators in one index number: undernourishment, child under weight, and
child mortality for children younger than five years.
54 T.M. Alzadjali

Table 3.12 State of hunger Country GHI Classification


in Arab countries in 2014
(source IFPRI 2014) Algeria Less than 5 Low
Bahraina Less than 5 Low
Comoros 29.5 Alarming
Djibouti 19.5 Serious
Egypt Less than 5 Low
Iraq 12.7 Serious
Jordan Less than 5 Low
Kuwait Less than 5 Low
Lebanon Less than 5 Low
Libya Less than 5 Low
Mauritania 11.9 Serious
Morocco Less than 5 Low
Omana less than 5 Low
Qatara Less than 5 Low
Saudi Arabia Less than 5 Low
Sudan 26 Alarming
Syria 5.9 Moderate
Tunisia Less than 5 Low
United Arab Emiratesa Less than 5 Low
Yemen 23.4 Alarming
a
These countries state of hunger was estimated by AOAD

3.2.4.3 Self-sufficiency Ratios and Food Gap

Arab countries are unlikely to achieve high ratios of food self sufficiency due to the
challenges and constraints in the agricultural sector, and the high rates of population
growth. However, they can improve and maintain the current ratios of food self
sufficiency by investing in limited horizontal agricultural expansion, and improve
the productivity of current agriculture. Food self-sufficiency programs and projects
which were promoted and adopted in the 1970s and 1980s did not achieve food
security; instead they contributed to unsustainable agricultural development.
The self-sufficiency ratios of sugar, vegetable oil, and grain are less than 50 %.
However, potatoes, vegetables, fish, fruit, and eggs have high self sufficiency ratios
as seen in Table 3.13.
The value of the Arab food gap was $35.6 billion in 2013 which did not change
significantly over the last three years (Table 3.14). Cereal forms about (53.6 %) of
this food gap, followed by milk and dairy products (9.5 %), refined sugar (9.0 %),
vegetable oil, red meat (8.7 % each), and poultry meat (7.8 %).
Six Arab countries contributed around 75.31 % of the total Arab food gap in
2013 (Egypt: 20.44 %, Saudi Arabia: 19.56 %, Algeria: 12.77 %, the UAE:
10.59 %, Morocco: 6.86 %, and Yemen 5.09 %).
3 Status of Food Security in the Arab Region 55

Table 3.13 Food self-sufficiency ratios in the Arab Region (%) (source AOAD 2013b)
Commodity 2011 2012 2013 Average 2011–2013
Refined sugar 30.35 30.63 31.46 30.82
Vegetable oil 39.32 34.39 39.09 37.60
Grain group 45.59 42.67 43.02 43.76
Legumes 50.86 53.59 58.34 54.26
White meat 66.50 65.81 66.91 66.41
Milk and dairy products 73.77 74.76 75.88 74.80
Red meat 80.93 81.40 80.94 81.09
Eggs 97.52 95.81 95.37 96.24
Fruit 96.33 96.16 96.85 96.45
Fish 98.56 98.43 99.12 98.70
Vegetables 100.93 100.07 100.01 100.34
Potatoes 102.77 99.52 100.07 100.79

Table 3.14 Surplus and deficit of food commodities in the Arab countries (source AOAD 2013b)
Commodity group 2011 2012 2013
$ millions
Cereal (total) 18211.1 19342.8 19950.6
Wheat and flour 8516.0 9203.7 9609.0
Corn 4188.2 4092.9 4434.3
Rice 3011.1 2938.3 2980.8
Barley 2195.8 2778.7 2768.8
Potatoes 66.7 180.5 125.4
Legumes 848.6 771.9 717.7
Vegetables 668.69 697.22 887.05
Fruits 347.3 363.6 125.7
Refined sugar 3799.5 3579.6 3342.7
Vegetable oils 2936.4 2962.0 3254.2
Meat (total) 6385.0 6216.5 6121.6
Red meat 3311.7 3205.3 3222.3
Poultry meat 3073.3 3011.2 2899.3
Fish 373.45 510.08 686.61
Table eggs 44.3 42.9 31.3
Milk and dairy products 3653.6 3371.5 3537.6
Deficit value 36292.5 36831.3 37206.9
Surplus value 1042.15 1207.30 1573.66
Food gap value 35250.3 35624.0 35633.3
56 T.M. Alzadjali

3.3 Strengthening Arab Food Security

As food security is high on the agenda of Arab leaders, they approved in March
2007 during their regular summit held in Saudi Arabia in March, 2007, the
“Strategy for Sustainable Arab Agricultural Development for Two Decades (2005
2025)”, which was prepared by Arab Organization of Agricultural Development
(AOAD), and was considered as part of the “Joint Arab Strategy for Economical
and Social Development” (AOAD 2007). The Arab leaders in January, 2009 during
their Economic, Development and Social Summit held in Kuwait, in response to the
2007–2008 world food crisis, launched the “Emergency Program for Arab Food
Security (EPAFS)” which was also prepared by AOAD. In 2011, the Arab
Economical and Social Development Summit (AESD) held in Egypt, approved the
AOAD prepared action plan for the implementation of EPAFS over twenty years.
The action plan was divided into three stages: first stage began in 2011 and ends in
2016, second stage starts in 2017 and ends in 2021, and the third and final stage
begins in 2022 and ends in 2031.
The objectives of EPAFS are to increase the contribution of domestic food
production to food availability, reduce food import bills; increase access to food by
reducing unemployment rates, increasing income, and lowering food prices;
enhancing food supply stability; promoting effective public-private partnerships;
and maintaining better political and social stability (AOAD 2009).
The commodity framework of the EPAFS encompasses grains (wheat, barley,
rice, maize, and sorghum), sugar crops (sugarcane and sugar beet), oil crops
(sesame, peanuts, sunflower, and olive), dates, and animal products.
EPAFS has three components which are:
1. Productivity improvement and rationalization of water use in existing
agriculture.
2. Horizontal expansion of agriculture with water saved as result of the imple-
mentation of the first component.
3. Integrated investment projects related to the above two components.
In accordance with the resolution of the first Arab Economic, Development and
Social Summit, AOAD has to prepare a periodical progress report for the Arab
Economic and Social Council about the achievements in the context of the EPAFS.
The overall achievements of the Arab countries in the context of the program
could be considered satisfactory, but it would definitely be far better if funding and
foreign direct investments were made available to countries with considerable
agricultural resources such as Sudan and Mauritania. The limited investment
capacity of public and private sectors, and a shortage of funds for agriculture and
food security projects in these countries constrain their contribution to the Arab
food security status and they cannot even improve their own state of hunger.
Sudan launched a food security initiative in the AESD summit held in Saudi
Arabia in 2013, in order to increase Arab direct investments and funding for
agriculture and food security projects, to improve the food security status of Sudan,
3 Status of Food Security in the Arab Region 57

and make significant contributions to other Arab countries. The details of the
initiative are under preparation and are expected to be submitted to the AESD
summit to be held in Tunisia in 2015.
In the absence of a Pan-Arab funding mechanism and limited Arab direct
investments in agricultural development and food security projects in the Arab
Region, it would be difficult to expect that the Emergency Program for Arab Food
Security (EPAFS) achieves its objectives in a twenty-year timeframe. Arab leaders
should be approached for the establishment of a funding mechanism for agricultural
development and food security in the region to ensure the implementation and
success of EPAFS which they launched and approved.

References

AOAD (2007) The strategy for sustainable arab agricultural development for two decades (2005–
2025). Arab Organization for Agricultural Development
AOAD (2009) The emergency program for arab food security. Arab Organization for Agricultural
Development
AOAD (2012) Arab food security report. Arab Organization for Agricultural Development
AOAD (2013a) Arab food security report. Arab Organization for Agricultural Development
AOAD (2013b) Arab agricultural statistics yearbook volume 33. Arab Organization for
Agricultural Development
AOAD (2013c) Arab fishery statistics yearbook, volume 7. Arab Organization for Agricultural
Development
FAO (2010) Food and agriculture organization of the United Nations. Retrieved from: http://www.
fao.org/home/en/
IFPRI (2014) Global hunger index. International Food Policy Research Institute, Retrieved from:
http://www.ifpri.org
Sudan Trade Point (2014) Monthly bulletins on international food commodity prices
The World Bank (2014) The world development report, 2014. The World Bank, Washington, DC
WHO (2014) World Health Organization. Retrieved from: http://www.who.int/
Chapter 4
Desalinated Water for Food Production
in the Arab Region

Hani Sewilam and Peter Nasr

Abstract The Arab Region is one of the most water scarce regions of the world.
85 % of the water in the Arab Region is used for irrigation. The region is the
world’s largest importer of grains. The direct link between food and water limits the
potential of water-stressed Arab countries to promote food production. However,
the prospects for using unconventional resources for irrigation, such as desalination,
constitute priority for consideration and action. The high cost of desalination for
irrigated agricultural crops is the main reason it is not used. It is necessary to
analyze each factor (e.g. parts, chemicals, labor, membranes, and energy)
influencing the costs of water desalination. Energy costs range between 50 and
76 % of the total cost of the desalination process; however, the downward trend in
the energy use per desalinated m3 of water indicates that desalination technology is
becoming more viable for irrigation use. The concept of sustainability with its three
pillars: economic, environmental, and social should be at the forefront of planning
any food production initiative using desalinated water. In addition to the economic
feasibility of using desalination for irrigation, the carbon footprint, brine disposal,
and lowering water levels in shallow seas are serious environmental aspects to be
considered. Arab countries should learn from previous good and bad practices. The
trial of Saudi Arabia to be self-sufficient in water-intensive crops, such as wheat,
from the 1970s onwards is a clear example for unsustainable development.
Developing local capacity to adopt state-of-the-art desalination technologies should
be on the agenda of Arab Governments. Serious support for research and devel-
opment is urgently needed to help develop and pilot test new desalination tech-
nologies, such as Forward Osmosis.


Keywords Desalination for agriculture Forward osmosis  Reverse osmosis 
  
Sustainability Water scarcity Irrigation Water quality

H. Sewilam (&)
RWTH Aachen University, Aachen, Germany
e-mail: sewilam@lfi.rwth-aachen.de
P. Nasr
The American University in Cairo, New Cairo, Egypt
e-mail: pnasr@aucegypt.edu

© Springer International Publishing AG 2017 59


K. Amer et al. (eds.), The Water, Energy, and Food Security Nexus
in the Arab Region, Water Security in a New World,
DOI 10.1007/978-3-319-48408-2_4
60 H. Sewilam and P. Nasr

4.1 Introduction

4.1.1 Water Scarcity in Arab Region

The age of water scarcity is upon us. The world is facing increasing demands on
supplies of fresh water due to increased population, domestic and agricultural
consumption, and extraction for power production and industrial uses (Mayer et al.
2010). The Arab Region is one of the most water scarce regions of the world. The
region either suffers from physical water scarcity (Fig. 4.1) or is expected to
approach this state in the near future (UN Water 2014). Although the region con-
tains 6.3 % of world’s population, it has access to only 1.4 % of the world’s
renewable fresh water (World Bank 2013). It is expected that by the year 2050,
two-thirds of Arab countries could have less than 200 m3/person/year of renewable
water resources (Zafar 2014, n.d.).
Almost 85 % of the water available in the Arab Region is used for irrigation.
Adopted irrigation methods are not sustainable and lead to overuse of scarce
renewable water resources, which in turn results in soil salinization. The region’s
average water use efficiency in irrigation is only 50–60 %, compared to
best-practice examples of above 80 % efficiency under similar climate conditions in
Australia and southwest US (Hoff and Gill 2014). Similarly, physical water losses
in municipal and industrial supplies in the region are well above world averages
(World Bank 2007).
World Bank (2013) argues that there are many reasons for water scarcity in the
Arab Region:

Fig. 4.1 Global water scarcity (Source UN Water 2014)


4 Desalinated Water for Food Production in the Arab Region 61

• inappropriate and imbalanced withdrawal from surface water and ground


aquifers;
• low rainfall over space and time and high evaporation rates;
• inefficient use of freshwater through poor irrigation practices, leakage in water
delivery systems and unnecessary consumption;
• pollution of fresh water resources;
• 60 % of the region’s water flows across international borders, further compli-
cating the resource management.
That being said, water scarcity has the following implications (UN Water 2014):
• food insecurity and social instability of a growing population;
• degradation of natural systems by using low quality water for irrigation
purposes;
• contamination of groundwater;
• reduction of crop yields and loss of arable land;
• long-term damage to soils and aquifers that may not be easily recoverable due to
salinization;
• potential risk of water conflicts.
Many Arab countries are engaged in an unsustainable development path to
economic growth which overuses water resources. That is why some of the demand
for water must be met from carefully selected, economically efficient development
of new water sources, and use of non-traditional sources, such as desalinized
saltwater (World Bank 2012). The agricultural sector is under pressure to improve
its water use efficiency (“more crop per drop”) to maintain required production
levels. Water use efficiency can be improved by making the public sector more
efficient, the private sector more involved, and farmer groups more responsible
(IFPRI 2012).

4.1.2 Irrigation Water Demand in Arab Region

Major irrigation areas in the Arab Region include the Nile delta in Egypt, areas
along the Euphrates and Tigris Rivers in Iraq, central Saudi Arabia, western
Republic of Yemen, Oman’s Batinah coast, and the Sebou and Oum el Rbia sys-
tems in Morocco (Fig. 4.2). The total irrigated area in the Middle East and North
Africa Region (MENA) is approximately 21 million ha and the corresponding
irrigation water demand is approximately 213 km3/year (Hoff and Gill 2014).
Seven countries have 90 % of the Arab Region’s irrigated area, and two countries
account for 50 % (World Bank 2012). Future irrigation demand was determined by
the irrigation potential, defined as the difference between the currently irrigated area
and the total irrigable land for which renewable water resources are available.
Generally, irrigation potential is related to renewable water resources. Yet, in many
arid countries, irrigation is sustained through mining fossil groundwater reserves, an
62 H. Sewilam and P. Nasr

activity common in Jordan, Libya, Saudi Arabia, the United Arab Emirates, and the
Republic of Yemen (Ouda et al. 2011). Through depleting the aquifers, the area
under irrigation can exceed the irrigation potential (Zafar 2014).
Due to climate change impacts, water balance modeling predicts a very small
increase in the average flow of the River Nile into MENA as a result of likely
precipitation increases projected for the Upper Nile basin. However, this increase
will be more than offset by decreasing precipitation and increasing evapotranspi-
ration (ET) within the MENA region. By the year 2015, if global warming induces
a wetter and warmer climate, irrigation water demand will increase by 15 % over
current demand (Table 4.1). On the other hand, if the future climate is warmer and
drier, irrigation demand is expected to increase by 33 % (World Bank 2012). Under
the most probable trend, demand will increase by approximately 25 %. While
climate change will mildly affect irrigation water demand, it will significantly
shrink water resources. If the climate turns out to be drier than present, renewable
water resources could be reduced by more than 40 % (UN Water 2014).

4.2 Impact of Water Scarcity on Food Production

The most severe impact of water scarcity is in food production. As mentioned


previously, 85 % of the water in the Arab Region is used for irrigation. While there
is a potential for adopting more efficient irrigation practices, avoiding the need to
restrict the amounts of water used for agriculture in many parts of the region is
inevitable, which is institutionally and politically very challenging (Al-Sheikh,
n.d.). The region is going to be much more dependent on trade for food products
given less groundwater available for agriculture. Thus, as the region’s dependence
on trade increases, economic risks and volatility become significant issues. The
Arab Region currently imports almost 50 % of its grain. Recently, the price of grain
has experienced sudden increases, which is having a significant impact on

Fig. 4.2 Distribution of Arab Region equipped for irrigation in year 2000 (Source World Bank
2012)
4 Desalinated Water for Food Production in the Arab Region 63

Table 4.1 Irrigation water demand in the Arab Region (km3/year and % increase over current
demand)
Climate scenario Average Dry Wet
Current 2000–09 213 – –
2020–30 237 (+11 %) 254 (+19 %) 222 (+4 %)
2040–50 265 (+24 %) 283 (+33 %) 246 (+15 %)
Source World Bank (2012)

consumers in the region (World Bank 2013). Clearly, the volatility in food prices is
going to be experienced in the region resulting in political concerns.
Arab countries remain largely net importers of food. Vulnerability of food supply
chains and instability of food prices is mainly due to high dependence on food
imports, which was clearly demonstrated by events of the global food crisis in 2007
(Sadik 2014). Although the direct link between food and water limits the potential
of water-stressed Arab countries to promote food production, the prospects for
using unconventional resources for irrigation, such as desalination, constitute pri-
ority for consideration and action (World Bank 2013).

4.3 Desalination in the Arab Region

In the past, due to the difficulty and expense of removing various dissolved salts
from water, saline water was an impractical source of potable water. However,
starting in the 1950s, desalination became economically viable for ordinary use.
Desalination solutions were introduced in the Arab Region thousands of years ago
in places such as Alexandria and Palestine. In modern history, desalination took
place in the Red Sea during the late 19th century in cities like Sawakin, Abu-Qair,
and Aden (Bushnak 2010). Jeddah was the first city to rely on seawater desalination
for drinking for more than a decade. Desalination units were deployed in Bahrain
and soon after multi-effect distillers were installed in Kuwait, Dhahran, Ras
Tanurah, and Alkhobar. During the 1950s, Multi Stage Flash (MSF) was introduced
commercially first in Kuwait. The first seawater desalination plant to employ
reverse osmosis (RO) for municipal water supply outside the USA was commis-
sioned in Jeddah in 1978 (Bushnak 2010).
Current commercial desalination technologies have been developed through
large-scale applications in a number of Arab countries (World Bank 2012). By 2007
approximately 54 % of the world’s desalination potential was installed in the Arab
Region (Fig. 4.3). Today, member countries of the Gulf Cooperation Countries
(GCC), as well as Algeria, Libya, and Egypt are the largest users in the region, as
indicated by their total cumulative contracted capacity of desalination plants
(Fig. 4.4). Worldwide production of desalinated water by 2007 was approximately
44 km3 a year: 58 % from seawater, 22 % from brackish water, and 5 % from
wastewater (World Bank 2012). The high rate of annual increase in contracted
64 H. Sewilam and P. Nasr

capacity is expected to continue over the next decade. By 2016 the Arab Region’s
share of global demand is projected to account for approximately 70 % of the
increased global capacity for desalination (UNDP 2013). Of the 15 countries with
the largest conventional desalination installations, nine are in the Arab Region. Yet,
this large expansion requires a review of present policies and practices including
how to increase local capacity and knowledge.
Commercial technologies used today in desalination can be grouped into two
categories, namely, thermal and membrane (Fig. 4.5). Thermal technology sepa-
rates water from minerals through evaporation-distillation using multi-stage flash
technology; a very energy-intensive process (Mizuno et al. 2013). Multi-stage Flash
Distillation (MSF) desalinates by evaporating and condensing seawater in various
stages, each time functioning on lower pressure than the last (Buros 1990). The heat
required for the thermal part of the process is usually obtained from the steam from
the water stream cycle of a power plant. MSF is a proven technology, even with
high levels of salinity, and can be built to a very large scale (Krishna 2004). As
thermal desalination technologies are most common in these countries, being the
older technology, GCC countries tend to co-generate electricity and water in large
plants in order to increase fuel efficiency (Bushnak 2010).
On the other hand, membrane processes use reverse osmosis to force pressurized
saline water through membranes that exclude most minerals (Buros 1990). Usually,
membrane technologies are used when electric power is accessible or when feed
water is brackish water (World Bank 2012). In the Arab Region, the MSF tech-
nology still dominates, particularly in the GCC countries, although installed
capacity for reverse osmosis is growing (Fig. 4.6). The reverse osmosis technology,
easily scalable due to its high modularity, requires no thermal energy and less or
equivalent amounts of electric energy than distillation. Most Gulf countries still
prefer the thermal technology however, because they use the disposed heat in
cogeneration systems. More recently, hybrid reverse osmosis and multi-stage flash
systems are being used in cogeneration systems (UNDP 2013).
The choice of technology used for desalinating brackish water is dependent on
the level of salinity (ESCWA 2009). Reverse osmosis is used mostly for higher
salinity brackish water, while electro-dialysis is more efficient for lower salinity

Fig. 4.3 Distribution of


worldwide desalination
capacity in 2007 (Source
World Bank 2012)
4 Desalinated Water for Food Production in the Arab Region 65

Fig. 4.4 Contracted capacity of desalination plants since 1944 in m3/day (Source Bushnak 2010)

Fig. 4.5 Categories of current commercial desalination technologies in Arab Region (Source
Bushnak 2010)
66 H. Sewilam and P. Nasr

Fig. 4.6 Contracted desalination technologies in the MENA Region since 1944 (Source Bushnak
2010)

brackish water (Krishna 2004). Figure 4.6 provides a breakdown of the cumulative
contracted capacity by technology in the Arab Region since 1944. MSF process still
dominates, although installed capacity for reverse osmosis has increased recently.
RO is increasingly used because of its lower cost and improved membranes
(Lenntech 2014). Hybrid technologies, such as MSF/RO or MED/RO, can be used
in the future to increase efficiency when power generation is required (AMTA
2007). Future large co-generation plants may combine NF/MSF/and MED/RO if
present research and technical solutions prove to be commercially competitive (Yip
and Elimelech 2011).
With respect to volume, desalination plants in the Arab countries have a
cumulative capacity of about 24 million cubic meters per day. The highest
desalination capacity (Fig. 4.7) is in the Gulf countries (81 %), Algeria (8.3 %),
Libya (4 %) and Egypt (1.8 %) (UNDP 2013). Growth is expected to remain high
for the next decade to meet escalating domestic water demand. Desalinated water
will expand from 1.8 % of the region’s total water supply to an estimated 8.5 % by
2025 (World Bank 2012). Most of the anticipated increase in capacity will be
concentrated in the Gulf countries, where it will be used to supply water to cities
and industry. More than 55 % of the water supplied to cities in the Gulf countries
comes from desalinated water; used directly, or blended with groundwater. This
share is expected to rise as groundwater resources continue to deteriorate.
There are a number of new desalination technologies under development. These
new technologies include membrane distillation, carbon nanotubes membranes,
aquaporin (biomimetics) membranes, thin film nano-composite membranes, for-
ward osmosis, and electro-dialysis/deionization (Elimelech 2007; Kim et al. 2010;
Mayer et al. 2010; Zhao et al. 2012). However, such technologies need further
research and development so that one can claim that they hold great promise for
desalination of seawater. In addition, the use of renewable energies, mainly solar
and wind, are still underutilized and need more attention from Arab countries.
4 Desalinated Water for Food Production in the Arab Region 67

Fig. 4.7 Accumulated desalinated water in selected Arab countries in the years 2010 and 2016
(Source UNDP 2013)

4.4 Feasibility of Water Desalination for Agriculture

Because of high costs, desalination technologies are not simply used for agricultural
purposes. A thorough cost analysis is essential in order to determine whether water
desalination may be feasible to produce a water resource that could be used to
complement or substitute natural water resources in areas with water shortages. Yet
the current situation is quite different from that of decades ago, when water
desalination started its development. However, more experience is still needed in
order to determine whether water desalination is a solution to water scarcity and
especially whether desalinated water should be used in agriculture.
That being said, it is necessary to analyze the factors influencing the water
desalination costs of the different desalination technologies. Desalination tech-
nologies have evolved in the last few years, from being little used in the world,
limited to some oil rich countries where energy costs are low, to now being used
globally. At the beginning, desalination was only used to provide domestic and
industrial supplies. However, once this technology had been improved and its costs
decreased, its application was extended to other sectors, especially to agriculture.
To obtain an average cost of desalinated water, it is necessary to consider three
factors:
• desalination technology and energy requirement
• feed water quality
• product water quality.
68 H. Sewilam and P. Nasr

4.4.1 Desalination Technology and Energy Requirement

Water desalination has evolved from the traditional systems of water distillation,
with high energy consumption, to the most modern membrane technologies,
especially reverse osmosis, which is more energy efficient and requires lower
investment costs. Although distillation technologies were predominant in the past,
the appearance of RO membranes in the 1970s has completely changed the
desalination scene in the world, and especially the application of desalinated water
for agriculture (Wang et al. 2008).
Water desalination relies on energy consumption, which is the main cost of
desalinating water. Distillation technologies consume considerable energy regard-
less of the level of water salinity (AMTA 2007). However, energy consumption
with membrane technologies depends on the salt content of the feed water and of
the product water. RO can be adapted to different water salinity contents. This
flexibility has enabled the extension of the use of RO to new applications.
Electro-dialysis reversal (EDR) is less flexible than RO, and should only be used for
special brackish water applications in agriculture (Buros 1990).
A novel membrane technology, Forward Osmosis (FO), can be used to produce
water for irrigation. This type of FO application is Fertilizer Drawn Forward
Osmosis (FDFO), as demonstrated in Fig. 4.8. As Phuntsho (2012) clarifies, two
different solutions are used in the FDFO process: saline water (as the feed water) on
one side of the membrane, and highly concentrated fertilizer solution (as the draw
solution) on the other side of the membrane. The two solutions are always kept in
contact with the membrane through a countercurrent flow system, where fresh water
flows from the saline feed solution towards the highly concentrated fertilizer draw
solution. After extracting the water by the FO process, the fertilizer draw solution
becomes diluted, and can be used directly for fertigation, provided it meets the
water quality standards for irrigation in terms of salinity and nutrient concentration
avoiding the need for separation and recovery of the draw solution (Phuntsho et al.
2011). Although the potential for such an idea is very promising, research on this
model has not received much consideration until recently due to the lack of suitable
membranes.

Fig. 4.8 Typical FDFO setup (Source Phuntsho et al. 2012b)


4 Desalinated Water for Food Production in the Arab Region 69

Fig. 4.9 Comparison of average energy requirements for different desalination technologies
(Source Phuntsho 2012)

FDFO is a remarkably low energy desalination process. The only energy


required in the FDFO process is for sustaining the cross-flow of the feed and draw
solutions in contact with the membrane surface and providing sufficient shear force
to minimize the Concentration Polarization (CP) effects (Phuntsho et al. 2011,
2012a). Figure 4.9 shows the relative energy requirements for different desalination
technologies. The total energy saved, when compared to other current desalination
technologies on an equivalent work basis, can be between 72 and 85 % (McGinnis
and Elimelech 2007). The performance of NH3–CO2 as a draw solution (DS) could
vary from the fertilizer draw solutions (Phuntsho et al. 2011). Yet, given the fact
that the recovery of draw solutes from the diluted draw solution is not necessary,
the estimates in Fig. 4.9 signal that the energy required for FDFO will be signifi-
cantly lower.
Agricultural productivity is mostly affected by fertilizers and water availability
(FAO 2005). As agriculture is by far the largest consumer of fresh water in the Arab
Region, small savings in agricultural water use through improved techniques will
provide immense quantities of water available for the community and the envi-
ronment. Besides making irrigation water available using lower energy from saline
water sources, FDFO desalination provides nutrient-rich water for fertigation
(Phuntsho et al. 2012b). According to Kafkafi and Tarchitzky (2011) and Phuntsho
(2012), fertigation has several advantages in comparison to the application of water
and fertilizers separately:
• minimum loss of irrigation water due to leaching
• optimum nutrient balance by supplying the nutrients directly to the root zone
• control nutrient concentration in the soil solution
• saving on labor and energy costs
• offering flexibility in fertilizer application timing
• suitable for application in mixtures with other micronutrients such as pesticides
• accommodating and flexible technology as it can be easily integrated in any
already-existing fertigation scheme.
70 H. Sewilam and P. Nasr

4.4.2 Feed Water Quality

For highly profitable out-of-season crops, desalinated seawater could be considered


as an alternative source of irrigation water (Beltrán and Koo-Oshima 2004).
Generally, seawater is seen as the most promising resource for desalination in the
future. This is because of the enormous volume that this natural water resource
represents and its availability. However, brackish water desalination is also applied
in many areas.
The economic feasibility of brackish and seawater desalination for agricultural
applications needs to be assessed. The World Bank (2007) forecasts that more than
200 million people will encounter the problem of water scarcity by the year 2025
and that most of this population will be living within 50 km of the sea coast. In
addition, technologies to desalinate seawater and brackish water are available and
their efficiency is continuously improving, permitting desalinated water to cover the
agricultural demand in these areas. Distillation technologies can only be used for
desalination of seawater at a very high cost, while EDR is only used for certain
brackish waters with a medium to low salt content (Beltrán and Koo-Oshima 2004).
The flexibility of RO to the salt content of the product water makes it possible to
reduce costs, an advantage that is not feasible with the other technologies (Altaee
et al. 2014).

4.4.3 Product Water Quality

The required salinity of the irrigation water used to achieve sustainable agriculture
depends on a number of factors, such as climate, crops, soils, and water manage-
ment (Beltrán and Koo-Oshima 2004). Therefore, the design of desalination plants
has to carefully consider the agricultural needs, so that production costs can be
optimized.
In order to reduce the Leaching Requirement and the quantity of water applied,
desalinated water could be used for specific and profitable crops, such as lettuce,
orange, and pepper. In this way, the cost of desalination would be less than that of
typical irrigation water. Figure 4.10 shows irrigation water costs in relation to the
total costs for some crops. This information could be used when deciding whether
or not to use desalinated water in agriculture.

4.5 Challenges of Using Desalination for Irrigation


in the Arab Region

For Arab countries to make desalination a suitable source of irrigation water, the
following challenges must be appropriately addressed.
4 Desalinated Water for Food Production in the Arab Region 71

Fig. 4.10 Irrigation water costs as a percentage of total costs as a function of the water price
(Source Beltrán and Koo-Oshima 2004)

4.5.1 High Financial and Energy Cost

As currently practiced, desalination is capital and energy intensive. Costs per


delivered cubic meter of desalinated water are as high as $1.50 and even $4 in some
cases (UNDP 2013). The water is subsidized, however, and sold for as little as
4 cents/m3 in some Arab countries. With improvements in desalination technolo-
gies, production costs are continuously dropping. Technologies such as RO, elec-
trodialysis and hybrids are more energy efficient and better suited to different types
of water. As shown in Fig. 4.11, the price of multistage flash over 1985–2004

Fig. 4.11 Reduction in the unit cost of multi-stage flash desalination plants, 1955–2003 (Source
UNDP 2013)
72 H. Sewilam and P. Nasr

Fig. 4.12 Operating costs of desalination processes in cogeneration plants (Source Bushnak
2010)

dropped from $4.0–$2.0/m3 to $0.50–$0.80 (UNDP 2013). Similarly, the current


price of RO is estimated to be $0.99/m3 for seawater and $0.20–$0.70 for brackish
water (UNDP 2013). The World Bank (2012) confirms that energy requirements
vary from 3.5–5.0 kWh/m3 for RO seawater to 4–8 kWh/m3 for multi-stage flash
technology. This downward trend in the cost of desalinated water indicates that
desalination technology is becoming more viable.
While the unit capital cost in 2010 for seawater desalination plants ranges
between $1000 and $2000/m3/day of installed capacity (Bushnak 2010), the unit
capital cost for brackish water plants is estimated to be 25–45 % of the above unit
cost for seawater plants (Bushnak 2010). The relative operating costs (parts,
chemicals, labor, membranes, thermal energy, and electrical energy) of the three
main desalination processes (RO, MSF, and MED) for cogeneration plants are
illustrated in Fig. 4.12. The operating cost of thermal desalination plants is much
higher than those illustrated if waste heat or steam is not available on site. Thus,
desalination, as practiced, is the most expensive water source option among other
local options, especially if large volumes of water are needed for irrigation.
Arab countries, especially gulf countries, Algeria, and Libya, plan to increase
desalination capacity from 36 million m3/day in 2011 to about 86 million m3/day
in 2025 (Bushnak 2010). By the year 2025, needed investments are estimated at
$38 billion, 70 % of which are in the Gulf area. UNDP (2013) claims that although
costs will vary with interest rates and energy prices, the energy costs of the expected
expansion in desalination capacity by the year 2025 can be projected using the cost
breakdown of a typical RO desalination plant (Table 4.2). Assuming a 10 %
interest rate, the cost of a unit cubic meter of desalinated water would be $0.62.
Arab countries are expected to desalinate around 19 billion m3 in 2016 and about
31.4 billion m3 in 2025, at an average cost of $0.525/m3. The predicted annual
desalination costs are estimated at $10 billion in 2016 and $15.8 billion in 2025, of
4 Desalinated Water for Food Production in the Arab Region 73

Table 4.2 Cost breakdown Parameter Cost ($/m3)


for typical reverse osmosis 3
desalination plant in the Arab Capital cost, 800 m per day capacity
Region Energy consumption, 3.5 kWh/m3
Annualized capital cost (at 5 % interest rate) 0.180
Energy cost (at $0.06 a kWh) 0.210
Membrane replacement cost 0.035
Labour and chemicals 0.100
Total cost 0.525
Sources UNDP (2013)

which energy costs will be almost $4 billion in 2016 and $6.4 billion in 2025
(UNDP 2013).
Desalination is energy-intensive, so energy efficiency is important in developing
new plants, as well as upgrading old ones (Semiat 2008). Saudi Arabia uses a
quarter of its oil and gas production to generate electricity and produce water in
cogeneration power–desalination plants (Al-Sheikh, n.d.). Assuming water demand
continues to grow at the current rate, this share will increase by at least 50 % by the
year 2030 (UNDP 2013). Likewise, in Kuwait, cogeneration power desalination
plants consume more than 50 % of total energy generated. The energy required to
meet desalination plant demand is expected to be equivalent to the country’s current
fuel oil production by the year 2035 (UNDP 2013). This means that the Arab
countries cannot keep relying on fossil fuels to cover their energy demand in the
future. Serious plans and investments need to be considered to integrate desalina-
tion with renewable energy sources (solar, wind, tidal, thermal, and waste bio-fuel).

4.5.2 Sustainability

To accomplish financial and environmental sustainability of desalination for irri-


gation purposes, renewable energy abundantly available in the region should be
used. With proper R&D incentives, solar and wind power can contribute greatly to
achieving sustainability and reducing the carbon footprint. Arab countries should
cooperate regionally to maximize the use of their tremendous solar power, partic-
ularly for water supply.
Currently, cheaper fossil fuel will make renewable energy uncompetitive unless
governments are prepared to support their adoption due to its potential contribution
to energy security, reduction of the carbon footprint, and “green” energy trading
opportunities (World Bank 2012). Since in the longer term, fossil fuels are highly
unlikely to continue to be cheap into the Arab Region, oil and gas will become
more expensive due to the high demand from South Asia and China dominating
world markets. Moreover, use of fossil fuels could become even more expensive if
the international agreements to minimize greenhouse gas (GHG) emissions take
effect, obliging countries to pay premium prices to support this call (UN Water
74 H. Sewilam and P. Nasr

2014). In this context, renewable energy may become highly competitive with fossil
fuels. In addition, energy efficiency will not be enabled if available fossil fuel is
accounted for at well below the market price, as is the case in most GCC countries
like Saudi Arabia (Bushnak 2010). It is claimed that if Arab countries used only
5 % of their deserts to build concentrated solar power plants, they could satisfy the
world’s energy needs (UNDP 2013). The move to nuclear power cannot be a
sustainable solution with all associated risks, especially in a region that is politically
unstable like the MENA region including Israel. This makes it clear that renewable
energies are the future for the region to survive.
In addition to moving to the use of renewable sources of energies, the water
demand needs to be controlled and the efficiency of water usage must be increased.
For example, surface irrigation has to be based on modern technologies such as drip
or sprinkler irrigation systems. New technologies such as sub-surface irrigation
should be seriously considered, tested, adopted, and promoted in the whole region.

4.5.3 Environmental Impacts

In addition to the favorable effect of producing water of good quality and pre-
venting the soil salinization hazard of irrigation with brackish water, water
desalination technologies have certain environmentally unfavorable impacts. The
main environmental impact of most desalination plants is related to the source of its
energy and carbon footprint. Figure 4.13 shows the carbon footprint of common
processes used today in cogeneration plants, such as MSF, Thermal Vapor
Compression (TVC), MED and SWRO. The carbon footprint of MSF plants, which
ranges from 10 to 20 kg CO2/m3, is very close to that for MED/TVC, that ranges
between 11.2 and 19.6 kg CO2/m3 (Bushnak 2010). If waste heat is not available,

Fig. 4.13 Carbon emissions in kg CO2 per unit volume (m3) of water produced in cogeneration
plants (Source Bushnak 2010)
4 Desalinated Water for Food Production in the Arab Region 75

the carbon footprint is much higher than the illustrated figures. Also, to save capital
costs, almost all thermal plants built in the GCC countries do not use a low heat
cycle, resulting in a higher carbon footprint. It is also useful to note that the carbon
footprint of power generation plants ranges from 0.5 to 0.8 kg CO2/kWh depending
on the type of fuel used and plant efficiency (World Bank 2012).
There are concerns raised about the effects of brine discharge on the marine
environment. According to the World Bank (2007), discharge of brine, residual
chlorine, trace metals, volatile hydrocarbons, anti foaming and anti-scaling agents
are having an impact on the near-shore marine environment. The increasing number
of plants on the Gulf and the rising temperature of its water need thorough research
to determine future environmental challenges and find possible solutions (World
Bank 2013).
More studies are needed to make sure that irrigation using desalinated water does
not have a negative impact on the crop. Such studies should establish water quality
limits of irrigation water including the total concentration of soluble salts, the
relative proportion of sodium to the other cations, the bicarbonate concentration as
related to the concentration of calcium and magnesium, and the concentrations of
specific elements and compounds (FAO 1985). The amounts and combinations of
these substances define the suitability of water for irrigation and the potential for
plant toxicity.

4.5.4 Capacity Development

Developing local capacity and implementing state-of-the-art desalination tech-


nologies through providing financial and logistical support are required. Arab
Governments should provide generous support to help develop and pilot test new
desalination technologies, such as Forward Osmosis powered by solar and wind
energies. This can be done by awarding local and regional universities with funds
and generous scholarships to research and test the applicability of the new tech-
nologies. Some countries (e.g., Saudi Arabia) have large allocations for science and
technology initiatives. It is still a challenge to see how local universities will be able
to convert their intellectual research ideas to high-value economic assets.
Arab Governments should offer financial support to allow the establishment of
desalination training centers. The governments in partnership with local companies
should build and equip such centers. The Saline Water Conversion Corporation
(SWCC) in Saudi Arabia has the only desalination focused training center in the
region. In addition, the Arab Water Council is spearheading capacity building by
establishing the Arab Water Academy (AWA) and the Arab Desalination
Technology Network to facilitate networking, capacity building, and cooperation
among desalination experts in Arab countries (Bushnak 2010).
76 H. Sewilam and P. Nasr

4.6 The Opportunities of Using Desalination for Food


Production in the Arab Region

4.6.1 Saudi Arabia and Wheat Production

Dry climate, poor soil quality, limited water supply, and limited arable land sig-
nificantly restrict agriculture in Saudi Arabia. However, it became self-sufficient in
water-intensive crops, such as wheat, by investing in and implementing policies
supporting the agriculture sector from the 1970s onwards (World Bank 2012). The
Kingdom’s agricultural and economic development was supported by vast oil
resources inhibiting food price inflation. By 1984, Saudi Arabia became
self-sufficient in wheat and even became the sixth largest exporter of wheat in the
early 1990s. However, such an agricultural strategy was environmentally unsus-
tainable and depended mainly on subsidies from the government (Al-Sheikh, n.d.).
In 2006, wheat production reached over 2.6 million tons, though this came at the
high cost of depleting non-renewable underground water resources (Ahmed et al.
2013). Groundwater depletion led the country to back off from its self-sufficiency
strategy and encouraged the development of extensive water desalination projects to
meet growing water demand (FAO 2013).
Being one of the world’s driest regions, with rainfall averaging less than
130 mm/year, groundwater accounts for 84 % of water supply and desalination
accounts for about 8 % of water (Al-Sheikh, n.d.) Agricultural water demand is
almost 86 % of the country’s overall water consumption (World Bank 2012).
Government policies, such as artificially low water tariffs and the prioritization of
fresh water for agriculture, resulted in inefficient use of water (Ahmed et al. 2013).
As wheat production placed large demands on non-renewable aquifers, an imbal-
ance took place between water recharge and water discharge, and underground
water tables have fallen in high production regions (Hoff and Gill 2014). The
Government is now terminating subsidies to the wheat sector and is completely
prohibiting wheat production in the country in an effort to preserve their limited
water resources (Ahmed et al. 2013).
Saudi Arabia is planning to be 100 % dependent on wheat imports by the year
2016 (Ahmed et al. 2013). In 2011, the Government allocated US$12.3 billion for
the development of agriculture infrastructure, including electricity, transportation,
storage, and ports (Al-Sheikh, n.d.).

4.6.2 Qatar’s Desalination for Irrigation Projected Scheme

Qatar consumes about 1.2 million cubic meters per day of desalinated water, almost
30 % more than what Israel uses for a population nearly quadruple its size.
Agriculture, feeding Qatar’s 1.8 million residents, would consume up an additional
3.5 million cubic meters per day of desalinated water and that is possible only if
4 Desalinated Water for Food Production in the Arab Region 77

farmers use the most efficient water conservation techniques, such as greenhouses,
drip irrigation, and hydroponics (Baker 2012).
Each cubic meter of desalinated Gulf water produces 45 kg of salt (UNDP
2013). The byproduct of Qatar’s projected desalination scheme will produce around
157 million kg per day of salt, enough to fill 4620 shipping containers. Currently,
Qatar pumps the resulting brine back into the Gulf, as does Saudi Arabia, the
Emirates, and Kuwait (Al-Zubari 2013). This has enormous negative impacts on the
marine life.
Producing the water necessary for Qatar’s agricultural vision will require
1.8 GW of power generation capacity. That would translate to nearly 4000 ha
worth of solar panel area (Baker 2012). Only 1 % of the country’s terrain is arable,
the rest is sand dune, urban settlements, gravel, and vast stretches of salt-crusted
soil called “sabkha” (FAO 2014). The Ministry of Agriculture has identified a total
of 68,716 ha, or 6 % of the land, that can be farmed, which is much less than what
Qatar would need to be truly self sufficient in food. Even if Qatar decides to lower
expectations to produce only 1.7 million tons of food per year (64 % of its total
food needs) an additional 30,000 ha of soil would need to be reclaimed (Baker
2012).

4.6.3 Abu Dhabi’s Tight Water Budget

Abu Dhabi is food-secure, but not food self-sufficient, as food security necessitates
that all citizens and residents have access to enough food, both physically and
economically, to meet their needs (IFPRI 2012). As its population continues to
grow, and with it the demand for food and fresh water, how can Abu Dhabi
continue to be food secure, and how can the domestic water and agricultural
strategies contribute to food security?
In Abu Dhabi, there are three sources of water: groundwater which makes up
65 % of current supply; desalinated water which is the primary source of potable
water making up around 30 % of supply, and finally recycled water which counts
for around 5 %. Although the Emirate desalinates water, groundwater continues to
be strategically important for agriculture and natural ecosystems, and is the only
form of long term water storage. Almost 80 % of this groundwater, is currently
used for agriculture (Sadik 2014). However, groundwater in Abu Dhabi and much
of the region is essentially a non-renewable resource. As this water is used, and
because of the dry environment, there are now significant signs of depletion of
aquifers. In most intensive agricultural areas, the groundwater levels are falling up
to 5 m/year, and as fresh water is used, aquifers are also becoming more saline
(Dakkak 2013).
It is estimated that Abu Dhabi has no more than 50 years of usable (fresh and
brackish) groundwater left if the current rate is continued to be used, but in some
intensively irrigated areas the timeframe could be much shorter. Between 2060 and
2070, groundwater aquifers will be exhausted, provided that the current
78 H. Sewilam and P. Nasr

consumption rate continues (Zafar 2014). The increase in population and changing
diets are estimated to lead to a 70 % increase in demand for food by 2050 (UN
Water 2014). As a result, the global demand for food is likely to peak at the same
time that availability of fresh water is declining.
Abu Dhabi’s water budget is around 1460 million cubic meters per year. The
current water consumption is 3500 million cubic meters per year, meaning that
there is 60 % water over use (IFPRI 2012). In order to save the remaining fresh and
brackish groundwater, there is a need to explore the option for increasing the
utilization of saline groundwater. Saving groundwater will be beneficial not only
from the water standpoint, but also from a food standpoint. Currently, Abu Dhabi’s
domestic food production contributes around 10 % of total food requirements, so
90 % of its food products are imported, and it will continue to be heavily reliant on
imported food in the future (Sadik 2014). It is currently cheaper to import food than
it is to produce it domestically and so it makes sense to preserve water resources to
increase agricultural production in the future, and in case food imports become
more expensive than domestic production. To enable this to happen, Abu Dhabi
does not only need water to be available, but also needs to design a water-efficient
agricultural system that can be scaled up at relatively short notice (FAO 2013).
Desalinating more water to increase available water budget is an inevitable
option. Yet, one must consider the significant environmental and financial cost
implications, as desalination will consume energy in the form of oil in the domestic
market, rather than being available for export. A preferred option would be to
explore how to increase the water budget using sustainable novel desalination
technologies, making use of saline groundwater.
Another option is to explore the possibility of using saline groundwater and
seawater through techniques such as bio-saline agriculture (Sadik 2014). Saline
groundwater is also a non-renewable resource, but it is less valuable than fresh
groundwater, and more limited in its potential uses. It is worth noting that Abu
Dhabi has four times as much saline groundwater as it has fresh and brackish
groundwater.

4.7 Recommendations

The following actions are suggested to make seawater desalination a more sus-
tainable source of irrigation water:
• Oblige all new desalination plants to reduce energy consumption and reduce
carbon footprint per unit water produced. Arab Governments should set a
maximum limit on water carbon emissions.
• Implement newly developed solar powered desalination technologies for small
and large systems. Arab based technical solutions and products for solar
desalination and cogeneration can provide a strong economic base for many
4 Desalinated Water for Food Production in the Arab Region 79

countries in the region. Arab countries need to plan for exporting solar power for
their future prosperity as much as they rely on oil and gas exports today.
• Governments must provide generous support to private investments in R&D,
training, high technology venture capital, and knowledge based local industries.
Such support should be integrated to achieve desired national local economic
outcomes and meet export targets in strategic industries like desalination and
solar power.
• Arab countries should develop joint R&D programs in desalination and
renewable energy such as wind, solar, and possibly wave and tidal power. Such
programs would maximize the value of new ideas and research findings
emerging from new institutional knowledge centers such as King Abdullah
University for Science and Technology (KAUST) and Qatar Foundation.

References

Ahmed G, Hamrick D, Guinn A, Abdulsamad A, Gereffi G (2013) Wheat value chains and food
security in the Middle East and North Africa region. Duke University, Center on Globalization,
Governance and Competitiveness (Duke CGGC). Retrieved from: http://sites.duke.edu/
minerva/files/2014/04/2013-08-28_CGGC_Report_Wheat_GVC_and_food_security_in_
MENA.pdf
Al-Sheikh HMH (n.d.) Country case study: water policy reform in Saudi Arabia. Proceedings of
the Second Expert Consultation on National Water Policy Reform No. Appendix 8), FAO,
Retrieved from: http://www.fao.org/docrep/006/ad456e/ad456e0e.htm
Altaee A, Zaragoza G, van Tonningen HR (2014) Comparison between forward osmosis-reverse
osmosis and reverse osmosis processes for seawater desalination. Desalination 336:50–57.
http://doi.org/10.1016/j.desal.2014.01.002
Al-Zubari W (2013) Oil and water: how can Arabs turn energy into food? Arab Forum Environ
Dev 39(8), 12 March 2013. Retrieved from: http://www.afedmag.com/english/ArticlesDetails.
aspx?id=92
AMTA (2007) Water desalination processes. American Membrane Technology Association
Baker A (2012) Desert dreams: can the Middle Eastern country of Qatar learn to Feed itself? Time.
Retrieved from: http://science.time.com/2012/11/19/desert-dreams-can-the-middle-eastern-
country-of-qatar-learn-to-feed-itself/
Beltrán JM, Koo-Oshima S (2004) Water desalination for agricultural applications. FAO
Buros OK (1990) The ABCs of desalting, 2nd edn. International Desalination Association,
Topsfield
Bushnak AA (2010) Desalination. Arab environment: Water No. 8, Arab Forum for Environment
and Development, Beirut, Lebanon
Dakkak A (2013) Water scarcity. In: Egypt|EcoMENA. http://www.ecomena.org/tag/water-
scarcity-in-egypt/. Accessed 17 March 2014
Elimelech M (2007) Yale constructs forward osmosis desalination pilot plant. Membr Technol
2007(1):7–8
ESCWA (2009) Role of desalination in addressing water scarcity (Water Development No. 3).
United Nations Economic and Social Commission for Western Asia, New York
FAO (1985) Water quality for agriculture. http://www.fao.org/docrep/003/t0234e/t0234e00.HTM.
Accessed 6 March 2014
FAO (2005) Fertilizer use by crop in Egypt. Food and Agriculture Organization of the United
Nations, Rome
80 H. Sewilam and P. Nasr

FAO (2013) Water uses. http://www.fao.org/nr/water/aquastat/water_use/index.stm. Accessed 25


March 2013
FAO (2014) FAO country profiles. http://www.fao.org/countryprofiles/index/en/?iso3=QAT.
Accessed 22 Dec 2014
Hoff H, Gill T (2014) Beating water and land shortages in the Middle East and north Africa. http://
www.theguardian.com/global-development-professionals-network/2014/nov/04/water-and-
land-shortages-middle-east-north-africa-renewable-energy-desalination. Accessed 9 Dec 2014
IFPRI (2012) Water and agriculture—Middle East and North Africa: dimensions of food security.
Retrieved from: http://www.ifpri.org/book-6959/node/8227
Kafkafi U, Tarchitzky J (2011) Fertigation: a tool for efficient fertilizer and water management, 1st
edn. International Fertilizer Industry Association and International Potash Institute, Paris,
France
Kim SJ, Ko SH, Kang KH, Han J (2010) Direct seawater desalination by ion concentration
polarization. Nat Nanotechnol 5(4):297–301. http://doi.org/10.1038/nnano.2010.34
Krishna H (2004) Introduction to desalination technologies. Texas Water Development Board
Lenntech (2014) Reverse osmosis desalination: brine disposal. http://www.lenntech.com/
processes/desalination/brine/general/brine-disposal.htm. Accessed 11 April 2014
Mayer TM, Brady PV, Cygan RT (2010) Nanotechnology applications to desalination: a report for
the Joint Water Reuse & Desalination Task Force. Sandia National Laboratories
McGinnis RL, Elimelech M (2007) Energy requirements of ammonia–carbon dioxide forward
osmosis desalination. Desalination 207(1–3):370–382. http://doi.org/10.1016/j.desal.2006.08.
012
Mizuno H, Kansha Y, Kishimoto A, Tsutsumi A (2013) Thermal seawater desalination based on
self-heat recuperation. Clean Technol Environ Policy 15(5):765–769. http://doi.org/10.1007/
s10098-012-0539-5
Ouda S, Khalil F, Gamal A, Sayed AH (2011) Prediction of total water requirements for
agriculture in the Arab world under climate change. In: Fifteenth international water
technology conference, IWTC-15, Citeseer
Phuntsho S (2012) A novel fertiliser drawn forward osmosis desalination for fertigation. Doctoral
of Philosophy Thesis, University of Technology, Sydney (UTS), New South Wales, Australia,
Jan 2012. Retrieved from: http://epress.lib.uts.edu.au/research/handle/10453/21808
Phuntsho S, Shon HK, Hong S, Lee S, Vigneswaran S (2011) A novel low energy fertilizer driven
forward osmosis desalination for direct fertigation: evaluating the performance of fertilizer
draw solutions. J Membrane Sci 375(1–2):172–181. http://doi.org/10.1016/j.memsci.2011.03.
038
Phuntsho S, Shon HK, Hong S, Lee S, Vigneswaran S, Kandasamy J (2012a) Fertiliser drawn
forward osmosis desalination: the concept, performance and limitations for fertigation. Rev
Environ Sci Bio/Technol 11:147–168. http://doi.org/10.1007/s11157-011-9259-2
Phuntsho S, Shon HK, Majeed T, El Saliby I, Vigneswaran S, Kandasamy J, Hong, S, Lee S
(2012b) Blended fertilizers as draw solutions for fertilizer-drawn forward osmosis desalination.
Environ Sci Technol 46(8):4567–4575. http://doi.org/10.1021/es300002w
Sadik A-K (2014) The state of food security and agricultural resources. In: Arab environment: food
security, Arab Forum for Environment and Development, Beirut, Lebanon
Semiat R (2008) Energy issues in desalination processes. Environ Sci Technol 42(22):8193–8201.
http://doi.org/10.1021/es801330u
UNDP (2013) Water governance in the Arab region: managing scarcity and securing the future,
New York, USA
UN Water (2014) The United Nations world water development report 2014—Water and energy,
vol 1. United Nations, Paris
Wang LK, Chen JP, Hung YT, Shammas NK (eds) (2008) Membrane and desalination
technologies. In: Handbook of environmental and desalination technologies, vol. 13. Springer,
New York
4 Desalinated Water for Food Production in the Arab Region 81

World Bank (2007) Making the most of scarcity: accountability for better water management
results in the Middle East and North Africa. MENA Development Report, World Bank,
Washington DC
World Bank (2012) Renewable energy desalination: an emerging solution to close the water gap in
the Middle East and North Africa. World Bank, Washington DC. Retrieved from: http://
elibrary.worldbank.org/doi/book/10.1596/978-0-8213-8838-9
World Bank (2013) Dealing with water scarcity in MENA. From http://go.worldbank.org/
EAENEPWXA0 Accessed 9 Dec 2014
Yip NY, Elimelech M (2011) Supporting information: performance limiting effects in power
generation from salinity gradients by pressure retarded osmosis. Department of Chemical and
Environmental Engineering, Yale University, New Haven
Zafar S (2014) Water scarcity in MENA. http://www.ecomena.org/water-scarcity-in-mena/.
Accessed 9 Dec 2014
Zafar S (n.d.) Water scarcity in MENA. Retrieved from: http://www.ecomena.org/water-scarcity-
in-mena/
Zhao S, Zou L, Tang CY, Mulcahy D (2012) Recent developments in forward osmosis:
opportunities and challenges. J Membrane Sci 396:1–21. http://doi.org/10.1016/j.memsci.
2011.12.023
Chapter 5
Water, Energy, and Food in the Arab
Region: Challenges and Opportunities,
with Special Emphasis on Renewable
Energy in Food Production

John M. Bryden

Abstract The Arab Region has the world’s greatest deficit of water and arable
land. These shortages are being exacerbated by climate change, land degradation,
growth in population, urbanization, and incomes. Although having apparently
abundant fossil based energy resources, the demands of desalination and air con-
ditioning have increased with income and population growth, and some Arab oil
producers are forecasting a potential future decline in net oil exports. Renewable
Energy (RE) is therefore seen as an opportunity in the region. In particular, the
richer Arab countries plan RE for desalinization, providing domestic drinking
water, as well as water for irrigated farming systems, while nearly all have now set
ambitious targets for RE. However, RE has yet to provide cheap energy; desalinated
water is expensive, and water is in increasing demand for direct human use in the
region’s growing cities; desalination itself has polluting residues; the choice of
technologies for RE production is limited by the context; and the technologies are
usually expensive to implement. Social, economic, and political obstacles around
access to water, energy, and food add to the environmental and technological
challenges. Yet the region has significant clusters of wealth and human talent, and
there are increased efforts to understand and tackle the various challenges. There are
opportunities for innovation, manufacturing, and export in RE. There are also
important opportunities to develop better understanding of the institutions and
governance around policies and innovation systems that help to determine who gets
access to water, land, and energy, and under what terms.

Keywords Renewable energy  Water  Food security  Human rights 



Institutions Governance

J.M. Bryden (&)


Grounded and Inclusive Innovation Group, Norwegian Institute
for Bioeconomy Research, Oslo, Norway
e-mail: john.bryden@nibio.no
J.M. Bryden
University of Aberdeen, Aberdeen, Scotland, UK

© Springer International Publishing AG 2017 83


K. Amer et al. (eds.), The Water, Energy, and Food Security Nexus
in the Arab Region, Water Security in a New World,
DOI 10.1007/978-3-319-48408-2_5
84 J.M. Bryden

5.1 Introduction

The 360 million inhabitants of the Arab Region represent about 5 % of the world’s
population. As Table 5.1 shows, however, the countries of the region are highly
diverse, both between each other, and within each country. GDP per capita varies
between $100 (Somalia) and just under $100,000 (Qatar) per annum, indicating
great extremes of wealth and poverty. They have less than 1 % of global water
resources, but 57.6 % of the world’s proven oil reserves (2013). Apart from fossil
fuel-based energy, they have enormous and largely untapped reserves of solar
energy. They are among the least self-sufficient countries in food production in the
world, importing more than half of the calories consumed. They are therefore
vulnerable to high global food prices as well as global political conflicts in which
food may be used as a weapon.

Table 5.1 Population, rural population, and per capita GDP in the Arab Countries 2011
Population Rural Rural GDP per
population Pop. capita
(’000) 2011 (’000) 2011 (%) US$ current, 2011
Net oil exporting Algeria 36414 13960 38.3 5459
countries (NOEC) Bahrain 1234 218 17.7 19465
Egypt 80410 45895 57.1 2928
Emirates 8400 107 1.3 40320
Iraq 33226 10234 30.8 4907
Kuwait 4010 45 1.1 40139
Libya 6423 1407 21.9 5712
Oman 3295 760 23.1 21236
Qatar 1733 77 4.4 98964
Saudi Arabia 28376 4973 17.5 23594
Sudan N&S 33976 22591 66.5 2059
Syria 25084 9586 38.2 2400
Yemen 23830 16900 70.9 1343
Total, NOEC 286411 126753 44.3 7703
Net oil importing Djibouti 906 214 23.6 1366
countries (NOIC) Jordan 6249 1087 17.4 4622
Lebanon 4942 538 10.9 8112
Mauritania 3297 1558 47.3 1233
Morocco 32245 13443 41.7 2902
Palestine 4231 703 16.6 1768
Somalia 13040 7204 55.2 100
Tunisia 10674 3425 32.1 4350
Total, NOIC 75584 28172 37.3 3336
All 361994 154937 43 6791
Source AOAD (2012)
5 Water, Energy, and Food in the Arab Region … 85

“Water development and management programmes, if planned properly, can serve


multiple functions, from contributing to energy and food production to helping
communities adapt to climate change. A nexus approach to sectoral management,
through enhanced dialogue, collaboration and coordination, is needed to ensure
that co-benefits and trade-offs are considered and that appropriate safeguards are
put in place”

UN World Water Development Report 2014 Vol.1.

Fig. 5.1 Why a nexus approach?

Water, food, energy, and human existence are inextricably linked, and to a much
greater extent in the Arab countries than in any other region of the world. However,
while access to both food and water are human rights, so far, access to energy is
not. However, it is evident that many of the socio-economic goals contained in the
UN Covenant on Economic Social and Cultural Rights cannot be achieved without
access to energy, and so it can be argued that the right to physical and economic
access to modern energy services is already implied in existing human rights
obligations (Bradbrook 2005).
Recently many have argued for a ‘nexus approach’ to the linked problems of
water, food, and energy, as exemplified by Fig. 5.1. The nexus approach would
integrate management and governance across water, food, and energy sectors,
leading to ‘joined up’ governance.

5.2 The Challenges for Water Supplies and Consumption

Water is the essence of life. The body cannot function without it, we cannot grow
food without it, and indeed the whole ‘bioeconomy’ depends on it. As Table 5.2
shows, the Arab Region lives in a perpetual—and worsening—water deficit, having
about 5 % of the world’s population and but 1 % of its fresh water resources. Over
millennia, it has developed sophisticated, sustainable and effective means of
matching demand to supply, such as the Falaj system in Oman, parts of which are
said to have been functioning for up to 2500 years. Yet these systems are unable to
cope with declining water resources and rapidly increasing demand. Water also lies
at the heart of key regional conflicts, including the Israeli-Palestine-Jordan conflicts
(Jordan River), the Turkey-Iraq-Syria conflict around Tigris and Euphrates rivers,
the N-S Sudan conflicts, Nile river conflicts,1 conflicts between Jordan and Saudi

1
However, at the time of final editing the Presidents of Egypt, Sudan and Ethiopia had met to sign
a fair use agreement regarding the waters of the Nile. Al-Jazeera News, 24 March 2015.
86 J.M. Bryden

Table 5.2 Water withdrawals and resources in Northern Africa and the Middle East
World Northern Middle
Africa East
Total withdrawal Municipal km3/year 469 9 25
by sector % 12 10 9
Industrial km3/year 731 6 20
% 19 6 7
Agricultural km3/year 2702 79 231
% 69 84 84
Water Total km3/year 3902 94 276
withdrawala Per inhabitant m3/year 593 607 986
Freshwater Total km3/year 3753 82 268
withdrawal % of IRWRb % 9 176 55
Actual renewable water resources per m3/year 6148 284 1588
capita (2006)
Sources WWAP, with data from FAO (2013) and UN (DESA) (2011) [for population]; FAO
(2014)
Notes aIncludes use of desalinated water, direct use of treated municipal waste water and direct use
of agricultural drainage water
b
IRWR internal renewable water resources

Arabia over groundwater extraction from the Disi/Saq aquifer, and the
Ethiopia-Somalia conflicts (Jubba and Shabele Rivers).
Table 5.2 shows that the agricultural sector consumes about 85 % of the water in
the region, and yet Arab countries are increasingly unable to feed their growing
populations. According to a recent UNDP report, “More than half of the food
calories consumed in the region are now imported; this share is expected to rise to
64 % over the next two decades” (UNDP 2013). The Arab Region’s food imports
contained the equivalent of 83 billion cubic meters (BCM) of virtual water in 1994,
or about 12 % of the region’s annual renewable water resources. This proportion
was, even at that time, several hundred percent in Saudi Arabia, Libya, and Jordan
(World Bank 2009). Recent estimates are that the amount of virtual water imported
in the region doubled from 148 to 310 BCM between 2000 and 2010 (UNDP
2013). The agricultural sector will need to change rapidly in the region due to
anticipated increased water shortages and related uncertainty, as well as rising costs
of water production and increasing demand from industry and water utilities serving
people, not to mention the likely assertion of human rights.
Significant parts of the Arab Region have experienced simultaneous population
growth, economic growth, and urbanization in the past 40 or so years. Given the
scarcity of water in the region, there is growing competition between agriculture,
industry, tourism, and people for the increasingly limited water resources, and many
in the region argue that water should be diverted from agriculture to industrial and
urban water supplies. However, as Table 5.3 shows, access to improved drinking
5 Water, Energy, and Food in the Arab Region … 87

Table 5.3 Access to improved drinking water in rural and urban areas. Middle East and North
Africa
Urban improved Rural improved
Country (selected) Year Total Total Total Total
improved improved improved improved
(1000) (%) (1000) (%)
Algeria 2003 19167.0 90.2 9662.1 82.2
2012 24287.8 85.5 8001.0 79.5
Bahrain 2003 682.4 100.0 89.7 100.0
2012 1169.4 100.0 148.4 100.0
Djibouti 2003 532.0 91.9 111.1 63.0
2012 663.1 100.0 128.7 65.5
Egypt 2003 29448.1 98.8 37955.7 95.8
2012 35229.4 100.0 44940.9 98.8
Jordan 2003 3934.9 97.9 875.4 90.7
2012 5656.6 97.3 1080.0 90.5
Kuwait 2003 2056.4 99.0 38.8 99.0
2012 3162.3 99.0 55.7 99.0
Lebanon 2003 3186.2 100.0 504.0 100.0
2012 4059.7 100.0 587.4 100.0
Morocco 2003 15504.1 96.5 8024.6 59.4
2012 18377.9 98.5 8812.4 63.6
Oman 2003 1528.0 89.4 531.3 78.1
2012 2332.0 95.5 751.0 86.1
Qatar 2003 639.6 100.0 20.7 100.0
2012 2028.9 100.0 21.6 100.0
Saudi Arabia 2003 17671.6 96.0 4272.5 96.0
2012 22647.2 97.0 4801.6 97.0
Sudan 2003 8131.6 73.2 14437.2 54.3
2012 8182.1 66.0 12445.2 50.2
Syrian Arab 2003 8675.3 94.5 6588.8 81.1
Republic 2012 11407.2 92.3 8312.1 87.2
Tunisia 2003 6240.1 98.1 2767.8 79.4
2012 7233.2 100.0 3295.1 90.5
United Arab 2003 2733.5 99.6 624.2 100.0
Emirates 2012 7760.6 99.6 1412.2 100.0
West Bank and Gaza 2003 2225.7 89.4 798.5 85.1
2012 2567.6 81.6 883.4 82.3
Yemen 2003 4174.8 78.6 6864.5 49.9
2012 5647.3 72.0 7447.1 46.5
Middle East and 2003 191481.4 94.5 120107.4 75.9
North Africa (ALL) 2012 239022.4 93.6 133318.0 78.5
Source customised table derived from WHO/UNICEF (2014)
88 J.M. Bryden

water supplies has been improving in most of the region in recent years, especially
in urban areas of the richer countries. Access to improved water remains at unac-
ceptably low levels in the poorer countries, and in the rural areas the access in many
Arab countries is very poor. Of the 53.4 million people in the region without access
to improved water supply, 70 % or 37.2 million, are in the rural regions.
Desalination plays an increasingly important role in the supply of fresh water to
cities in the region, especially in the wealthy Gulf States which have 81 % of
desalination capacity in the region. However, the rising use of fossil fuels in
desalination has caused declining exports of oil. For example, in Kuwait, the energy
required to meet desalination plant demand is expected to equal the country’s
current fuel oil production by 2035 at current rates of growth. Equally, the region’s
oil exporters, and especially the Gulf Cooperation Council (GCC) countries, were
alarmed by a 2011 analysis by Chatham House (2011), which warned that if Saudi
Arabia continued its domestic consumption growth rates it would lose the ability to
export oil by 2020 and become a net importer by 2038 (UNEP/REN21 2013).
Desalination is also expensive in terms of investment and energy costs, and the cost
of water produced, although falling, remains at over $1 per cubic meter.2
Desalinated water is therefore unlikely to be utilized for agriculture.
There are several options for tackling the complex threats to the
water-food-energy problems of the region, but before discussing these, it is
important to emphasize that the Arab Region is extremely diverse. A few countries,
especially Lebanon, do not yet have an endemic water shortage, and their agri-
cultural sector remains robust. Some countries—especially the Gulf states—are
among the richest in the world in terms of GDP per capita, even if that GDP
remains highly unequally distributed. Others, like Yemen, Djibouti, Comoros, and
Mauritania are among the poorest in the world. Yet others are experiencing
on-going conflicts, including Syria, Israel-Palestine, Northern and Southern Sudan,
and Somalia. The nature of the physical, human, financial, and institutional
resources available to tackle the nexus of water, food and energy challenges is
therefore highly particular, and potential solutions need to be carefully matched to
context; one size does not—and will not—fit all.
The technologies to solve many of the problems are largely known, even if they
require adaptation to specific contexts. We know that source separation toilets can
reduce the water footprint by up to 90 % while also facilitating the return of
valuable nutrients to the land; we know how the Arab Region’s vast solar resources
can be used to make renewable energy; we know a lot about techniques of reducing
water consumption in agriculture, and dry-land farming. The problems are largely
those of adoption, which in turn concern human behaviour and its conditioning by
the social institutions that establish the ‘rules of the game’. More attention therefore

2
Although subsidized for consumers. See UNDP (2013) and Sowers (2011).
5 Water, Energy, and Food in the Arab Region … 89

needs to be paid to the role and functioning of human institutions, where innovative
new solutions can steer human decisions in the desired directions.
In this context, efforts to improve water governance have apparently stepped up,
but there is still a long way to go in securing human rights to adequate, clean, water
in the region, and in establishing integrated water management laws, regulations,
and other institutions that both provide for human rights and for other water uses
that go beyond basic needs. Thus, the Integrated Water Resources Management
(IWRM) approach outlined below has recently been strengthened in some cases by
a (human) rights based approach, which gives a higher priority to the fundamental
human right to adequate clean water. However, such principles of good water
governance based on a human rights approach frequently clash with other gover-
nance models promoted by some international agencies, such as New Public
Management, which lack attention to human rights, and stress efficiency and the
generation of financial surpluses from what are termed ‘services’ rather than basic
needs (Fig. 5.2).

Excerpts from international policy documents on the concept of Integrated Water


Resources Management

(a) “Integrated water resources management is based on the perception of


water as an integral part of the ecosystem, a natural resource, and a
social and economic good, whose quantity and quality determine the
nature of its utilization. To this end, water resources have to be
protected, taking into account the functioning of aquatic ecosystems and
the perenniality of the resource, in order to satisfy and reconcile needs
for water in human activities. In developing and using water resources,
priority has to be given to the satisfaction of basic needs and the
safeguarding of ecosystems. Beyond these requirements, however, water
users should be charged appropriately.” Source: Agenda 21. Chapter
18. Protection of the quality and supply of freshwater resources:
application of integrated approaches to the development, management
and use of water resources.
(b) “Halve, by 2015, the proportion of the population without sustainable
access to safe drinking water and basic sanitation”. Source: Millennium
Development Goals, Goal 7, Target 3.
(c) “The provision of clean drinking water and adequate sanitation is
necessary to protect human health and the environment. In this respect,
we agree to halve, by the year 2015, the proportion of people who are
unable to reach or to afford safe drinking water (as outlined in the
Millennium Declaration) and the proportion of people who do not have
access to basic sanitation).” Source: Johannesburg Plan of
Implementation of the World Summit on Sustainable Development,
Johannesburg, South Africa, 2002.

Fig. 5.2 Integrated water management and international policy (source Iza and Stein 2009)
90 J.M. Bryden

5.3 The Challenges for Energy Production

Most of the energy for electrical power (heavily utilized for cooling, water pumps,
industry, and urban domestic purposes) and for transportation, as well as desali-
nation come from fossil fuel sources. Domestic use competes with exports for oil in
the oil exporting countries, and adds to import burdens of the—often poorer—net
oil importing countries. Oil produces a major part of export earnings and govern-
ment revenues on the oil exporting countries, and there are concerns in several Arab
countries such as Kuwait, Saudi Arabia, and the Emirates about the diversion of oil
to meet growing domestic energy demands (UNEP/REN21 2013; UNDP 2013). As
can be seen in Fig. 5.3, total energy demand in the region has been increasing
rapidly due to population growth and economic growth, also linked to increasing
demand for water and air conditioning. Total Primary Energy Supply (TPES) in the
MENA region reached about 800 million tonnes of oil equivalent (Mtoe) in 2010,
an increase of 14.9 % compared to 2007, an average annual growth of 4.7 % over
the period. Electrical power generation reached about 1200 terawatt-hours
(TWh) in 2011 in the MENA region, a 20 % increase compared to 2008
(UNEP/REN21 2013). The growth was higher in the net oil exporting countries.
In the face of this increased demand, growing recognition of the link between
fossil fuel consumption and global warming, and high oil prices until late 2014,
there has been an increasing interest in renewable energy investment and

40 NOIC
NOEC
TOTAL MENA
35

30

25

20

15

10

0
2005 2006 2007 2008 2009 2010
Notes : NOIC= Net Oil Importing Countries
NOEC=Net Oil Exporting Countries

Fig. 5.3 Total primary energy consumption in MENA region 2005–2010 (quadrillion BTU per
annum) (source figure generated from IEA data)
5 Water, Energy, and Food in the Arab Region … 91

production in the region. While the RE share in the TPES of the MENA region
remained about 1 % from 2007 to 2010, the share increased from about 5 % to
almost 6 % in 2010 in the net oil importing countries of the region, with Tunisia
(14 % of total energy from renewables) and Morocco (approximately 5 %) leading
the way (IEA/OECD 2009; UNEP/REN21 2013). Interest in RE is particularly
strong in the net oil importing countries such as Jordan, Morocco, and Lebanon, but
there are also significant longer term RE targets in the oil exporting countries of the
Emirates and Saudi Arabia. These targets indicate strong future growth of RE in the
period to 2020, with a principal focus on wind turbines and solar technologies,
especially concentrated solar power. Hydro electricity is currently important as a
renewable source, but new opportunities in the region are limited, although
neighboring Ethiopia has considerable unrealized potential, as discussed later.
According to Bloomberg New Energy Finance (2013), new investment in renew-
ables in the MENA region was US$ 2.9 billion in 2012, an increase of almost 40 %
over 2011 and a 6.5-fold increase compared to 2004, and this is mostly in wind and
solar energy.
The production of energy from renewable sources other than hydro will not,
however, be cheap, and significant adaptation of the technologies will be needed to
meet the climatic conditions in the region; for example, extreme temperatures and
sandstorms are a problem for solar panels, as well as for wind turbines. Water is
also required for such technologies (for steam turbines in Concentrated Solar
Power, for washing of solar panels, etc.). It is also necessary to improve efficiency,
for example, by combining desalination with electricity generation and reducing
grid losses and supply fluctuations. Although regional research and innovation
capacity exists, for example in the Gulf States (MASDAR, IRENA) and in Egypt,
Jordan, Morocco, and Tunisia, this needs strengthening and re-orientation to more
participatory ‘co-learning’ approaches that engage users, producers, and authorities,
as well as researchers if indigenous innovation systems are to emerge that can also
underpin new manufacturing. It is important to add that the region is also investing
in nuclear energy, and coal fired electricity generation, both of which have demands
on water resources (Sowers 2011).
The cost of energy is thus one challenge that the region faces, and will face in the
future. In some countries, especially the net oil exporting countries, electrical
energy is subsidized in order to make it affordable. However, this policy has fre-
quently been challenged by organizations such as the IEA which argue that rather
than benefiting poor people, the subsidies benefit the growing middle classes and
wealthier people who consume the most energy. Therefore, there are clear political
challenges in reducing energy subsidies, and such a policy needs to be balanced by
increased spending on education, health, and social welfare to compensate.
The energy systems of the different Arab countries are not integrated, and the
grid infrastructure has been criticized as inadequate to improve the security and
quality of electricity supply. Recent regional initiatives include the multi-billion
dollar Gulf States Cooperation Council power grid project, creating an integrated
electricity network between Saudi Arabia, Qatar, Bahrain, Kuwait, Oman, and the
UAE (Al-Asaad and Ebrahim 2008). In addition, the Seven Countries
92 J.M. Bryden

Interconnection Project (SCIP), launched in the early 1990s, aims to interconnect


the grids of Libya, Egypt, Jordan, Syria, Iraq, Turkey, and Lebanon. An HVDC line
is also planned between Saudi Arabia and Egypt which aims to improve security of
supplies and energy trading opportunities. The training programme REGRID also
aims to empower professionals such as grid operators in Algeria, Egypt, Jordan,
Lebanon, Morocco, Syria, and Tunisia and equip them to cope with high volumes
of renewable power in their electricity systems (RENAC 2011).
Despite these and other efforts, the figures in Table 5.4 show that many com-
munities and individuals in the region remain unserved by the electricity grid, and
these are mainly in the rural regions of the poorer countries. The countries of the
Middle East and North Africa generally have high electrification rates—99 % or
more in 13 of the 21 countries. However, only nine have 99 % or greater access in

Table 5.4 Electrification rates and rural electrification, MENA countries, 2005 and 2010
MENA country/group Electrification Rural electrification rate %, 2010
rate %
2005 2010
Net oil exporters
Algeria 98.1 99.3 98
Bahrain 99 99.4 95
Egypt 98 99.6 99
Iran, Islamic Republic of 97.3 98.4 95
Iraq 15 86 94
Kuwait 100 100 100
Libya 99.8 99
Oman 95.5 98 93
Quatar 70.5 98.7 69
Saudi Arabia 96.7 99 94
Sudan 36
Syrian Arab Republic 90 92.7 84
United Arab Emirates 91.9 100 100
Yemen 36.2 39.6 23
Net Oil Importers
Djibouti n/a 50 20E
Jordan 99.9 99.9 99
Lebanon 99.9 99.9 99
Malta n/a 99.9 100
Morocco 85.1 97 97
Palestine n/a 99.7 99E
Tunisia 98.9 99.5 99
Sources IEA (2006), (2012); UNEP/REN21 (2013)
Notes NOEC net oil exporting countries
NOIC net oil importing countries
5 Water, Energy, and Food in the Arab Region … 93

the rural areas, and three have less than 50 % rural access. Electricity access
remains seriously inadequate in rural areas of Yemen (23 %), Sudan, and Djibouti
(estimated at around 20 %). Based on IEA data for 2010, 23.3 million people in the
MENA region lacked electricity access, and most of these live in rural areas.
Although data on rural electrification rates are not available for 2005, the IEA gives
rural electrification rates of 79.9 % for North Africa and 76.6 % for the Middle East
in 2000. The overall trend for electricity access has therefore been positive, and
notable cases are Iraq, Qatar, and Morocco, where substantial improvements in
electricity access have been achieved (IEA 2002; UNEP/REN21 2013).
Nevertheless, some 23 million people in the MENA region lack electricity and more
than that still use traditional biomass for cooking, and these are overwhelmingly in
rural regions. In addition, 64 % or some 27 million people in former Sudan had no
access to electricity and less than 1 % of people in the new South Sudan have
access. Somalia and Eritrea also have low rates of rural electrification although
recent data is not available.
Solar photovoltaic (PV) and solar water heaters are common rural off-grid
solutions for light, battery charging, and heat at the household level, as well as for
schools and small clinics, etc. Several village-based solar PV systems are in
operation in the MENA region, for example in Djibouti. Tunisia’s 2008 Renewable
Energy Plan included a major effort to develop renewable energy applications as a
means for rural electrification, and for use in the agricultural sector. The plan’s
objectives included installation of 63 wells with PV pumping stations and water
desalination units; 200 water pumping stations for irrigation, and two industrial
units connected to the network for combined heat and power from biogas; in
addition to the equipment of 200 farms with biogas units for domestic use; elec-
trification of 1000 rural households by hybrid systems; electrification of 1700 rural
households by PV systems; and electrification of 100 farms and tourist centres by
hybrid systems.
It is important that rural communities are encouraged to develop their own
renewable energy systems. Experience in Europe, India, North America, and Africa
has shown that this is feasible (OECD 2012). This is because renewable energy uses
rural resources of land and water, wind and sun, and provides economic opportu-
nities to rural people, especially if they have an ownership stake. This is true of both
rural and urban communities that currently lack energy supplies, but also more
generally where there are large future opportunities for solar and wind power in the
region.
The lack of adequate electrification in the rural regions of poorer African
countries is a challenge for the adaptation of agrarian systems, including those using
flash and gravity irrigation systems. More efficient and effective trickle irrigation
and the use of various forms of recycled waste water, do require electrical pumps.
Electricity is also needed for light to read, for schools and clinics, and to avoid
depletion of scarce biomass currently used for cooking purposes.
94 J.M. Bryden

5.4 The Challenges for Food Production

In 2011, 43 % of the population of Arab countries lived in rural areas, and agri-
culture remains an important source of rural livelihoods and employment in the
region, accounting for 37 % of the economically active population in 2006 (IFAD
2009). It is a major water user, accounting for some 85 % of the region’s annual
water use. Yet the region remains highly dependent on imported food, and vul-
nerable to both international food price fluctuations and global political conflicts.
Most of the agriculture in the region is either based on extensive dry-land
farming systems, on rain-fed farming in some regions of some countries, or on
irrigation. Most of the irrigation systems in use are considered inefficient, and have
severely depleted the region’s underground aquifers over the past 50 years—the
period of industrialization of the irrigated agricultural sector (Sowers 2011). The
situation in poor countries like Yemen is particularly severe, as Fig. 5.4 illustrates.
Given the water shortages in the region, the use of urban wastewater has
received some, but not enough, attention. So far at least, urban wastewater is little
used for agriculture in most Arab countries (ACWUA 2010). Low prices for water
used by irrigated agriculture, and subsidized energy for pumping systems, have not
been helpful for efforts to improve irrigation efficiency or encourage conservation
either. In addition, until recently policies in the region have encouraged greater
self-sufficiency, due to the high cost of imported food.
The use of various forms of wastewater could be greatly expanded, especially in
and near to the urban areas, but generally everywhere. There are cultural barriers—
public attitudes to use of wastewater for food crops can be adverse—as well as real

In Yemen, the situation is particularly problematic. Water per capita is less than 100
cubic meters per year, less than 2% of the world average. With no significant
perennial sources of surface water, Yemen relies on over-exploitation of
groundwater, extracting water at a rate three times its replenishment from the
limited rainfall. In the Sana’a basin, home to over 10% of the country’s population,
the water table is falling by 6 to 8 meters a year. The falling water table has
encouraged the use of oil digging rigs, further exacerbating the unsustainable
situation. By 2009, Yemeni Water and Environment Ministry officials estimated that
more than 800 private drill rigs were operating in the country, in contrast to only
three rigs in Jordan. In the meantime, the concentration of minerals in the water
has increased, leading to deterioration in water quality. The resultant water
shortages and increased cost of water have forced thousands of families to spend a
third of their incomes on purchasing water. Meanwhile, cultivation of qat, a water-
intensive crop that itself uses up to 40% of the water, has increased seven-fold since
the 1970s (UNDP, 2011:57)

Fig. 5.4 Water challenges in Yemen


5 Water, Energy, and Food in the Arab Region … 95

health risks to be tackled by adopting modern ecological recycling methods, and


sound regulation and regular checks are needed. Currently there are no risk man-
agement checks in place in most Arab countries, and wastewater is often discharged
into the sea. However, there is a large potential resource of wastewater for urban
and peri-urban agriculture that could be utilized, while at the same time allowing
scarce and expensive nutrients to be returned to the soil, given the correct invest-
ment decisions (the technologies are known, but often not used in the region), and
regulatory and supervisory structures. Morocco, for example, generates some 600
million cubic meters of wastewater per annum, which is largely unused, and re-use
is not even being considered in new wastewater treatment projects (ACWA 2010).
Given the growing urbanization and water deficits in the region, together with
reductions in ground water and its decreasing quality, the pressures on agriculture to
reduce water use are very large and growing. In the light of the climate change
prognosis for the region, this situation seems likely to remain for the foreseeable
future.
Lampietti et al. (2011) advance a number of suggestions for improving food
security in the region in the face of on-going food security risks and probably
continuing high food prices on world markets. The urban poor, landless rural
people, and small and marginal farmers are the worst affected by high food prices—
the poor spend between 35 and 65 % of their income on food, and the poor are
concentrated in rural areas.3 Many Arab countries have subsidies on food to ease
the financial burdens on the poor, but these subsidies are not sustainable in the
poorer non (net) oil-exporting countries such as Syria, Jordan, Egypt, Morocco, and
Yemen while even net oil exporters are challenged when oil prices fall, as they have
in 2014. Although writers such as Lampietti et al. suggest better targeting of food
subsidies through devices such as proxy means-testing, these methods stigmatize
the poor and are difficult to operate in rural areas where most of the poor are. Other
approaches including free nutritious school meals for all that ensure minimum daily
intakes for children or, where feasible, universal access to a minimum daily quantity
of subsidized food grains or bread would be preferable. However, the human rights
based approach recently adopted by India, and discussed below, perhaps offers the
most promise for the future.
Another approach hinges on more effective use of water in agriculture.
Desalinated water is too costly to use for food production, but water harvesting,
which captures run-off after rainfall, is another suggestion which is already in use,
and which can often be done at reasonable cost. However, ultimately such methods
can only provide a proportion of the water currently used by agriculture.
Better irrigation methods—drip and sprinkler—are also being introduced where
possible, but these require more sophisticated pumping and control systems.

3
Over 80 % of the poor in Yemen and Sudan were in rural areas, between 70 and 80 % in Egypt,
Mauritania and Tunisia, and between 60 and 70 % in West Bank and Gaza, Syria, and Morocco.
See Table 1 in Lampietti et al. (2011).
96 J.M. Bryden

It is also important to optimize the use of agricultural areas where rain-fed


agriculture is feasible, despite the generally dry climates and the adverse impacts of
climate change, for example, by using conservation agriculture techniques.
Table 5.5 shows that rain-fed agriculture is used on 62 % of cultivated land in the
region, and is especially important in Algeria, Iraq, Jordan, Lebanon, Libya,
Mauritania, Morocco, Somalia, Sudan, Syria, Tunisia, and Yemen. However, the
practice is mainly used by poorer and marginalized herders and farmers in the
region, who lack investment and other resources, and where extension services may
be weak or non-existent (Lampietti et al. 2011).
The International Center for Agricultural Research in the Dry Areas (ICARDA)
based in Syria—part of the CGIAR consortium—undertakes research on dryland

Table 5.5 Cultivated, irrigated and rain-fed land in the Arab Region in 2011
Cultivated Irrigated Rainfed % of cultivated
area area area area
(1000 ha) (1000 ha) (1000 ha) Irrigated Rainfed
Net oil exportingAlgeria 8445 987 4212 11.7 49.9
countries Bahrain 4 3 – 75.0 0.0
(NOEC)
Egypt 3620 3468 152 95.8 4.2
Emirates 234 234 – 100.0 0.0
Iraq 4480 2413 1097 53.9 24.5
Kuwait 10 10 – 100.0 0.0
Libya 2644 258 1536 9.8 58.1
Oman 75 75 – 100.0 0.0
Qatar 28 12 – 42.9 0.0
Saudi Arabia 4192 806 – 19.2 0.0
Sudan N&S 21,106 1603 18,797 7.6 89.1
Syria 5716 1400 3180 24.5 55.6
Yemen 1609 748 664 46.5 41.3
Total, NOEC 52,163 12,017 29,638 23.0 56.8
Net oil Djibouti – – – 0.0 0.0
importing Jordan 302 96 145 31.8 48.0
countries
Lebanon 245 126 106 51.4 43.3
(NOIC)
Mauritania 322 22 300 6.8 93.2
Morocco 10,078 2522 6060 25.0 60.1
Palestine 84 – 84 0.0 100.0
Somalia 1500 160 1012 10.7 67.5
Tunisia 5206 517 3667 9.9 70.4
Total, NOIC 17,737 3443 11,374 19.4 64.1
All 65,710 15,459 41,012 23.5 62.4
Source AOAD (2012)
5 Water, Energy, and Food in the Arab Region … 97

farming systems, and suggests two main strategies (Pedrick 2012). The first applies
to areas of high agricultural potential with good access to sustainable water supplies
(as in parts of Egypt) and involves ‘sustainable intensification’ on existing agri-
cultural land.
The most direct link between renewable energy and agriculture is of course the
link between the construction of the first and second Aswan Dam’s (mainly for
hydro-electricity) and the development of irrigated agriculture mainly in Egypt, but
also in Sudan and Ethiopia. Using over 30,300 km of channels and large canals (EI
Gamal 1999), Egypt has expanded cultivation of horticultural crops such as fruits,
nuts, vineyards, and vegetables on reclaimed desert land. Using drainage waters,
aquaculture has also expanded in the Nile Delta from some 20,000 tonnes in the
1980s to over 600,000 tonnes by 2009 (Conniff et al. 2013; Oczkowski et al. 2009).
Egypt also plans a large land reclamation project in the North Sinai Desert
involving the diversion of 4.45 billion m3/annum of Nile water to develop irrigated
agriculture west and east of the Suez Canal.
The use of the Nile waters is however a source of conflict within the region and
between neighbours. Egypt signed an agreement with Sudan in 1959, but Ethiopia
was not involved, and has considerable unrealized irrigation and hydro-power
potential on both the Blue Nile and Atbara rivers, even if the net benefits have been
disputed because of poor timing of water flows. Nevertheless, Ethiopia and seven
other Riparian countries in the Nile basin have interests that have so far have not
been taken into account nor made subject to common agreement. The agricultural
potential of southern Sudan, considered to be very considerable, has yet to be
realised due to bad land governance and conflicts rather than lack of agreements,
but the apparently considerable agricultural and hydro-electrical potential of
Ethiopia has also not been realised, if largely for different reasons. According to
reports, Ethiopia has a potential of some 3.7 million ha for irrigation, of which
about 50 % is in the Nile Basin. However, only 250,000 ha has so far been irri-
gated, of which less than 20,000 ha is in the Nile basin (Awulachew et al. 2007,
2009).
Dixon et al. (2001) have argued that there has been a strong urban bias for
decades in the region’s public goods investment policies as well as development
policies in general. Along with these policies have been the centralization of
decision making and public institutions, and the favouring of major cities for
provisions of energy, water, basic services, and communications. Centralization of
power, and rigid target systems, have stifled innovation and crop diversification.
Lack of forward thinking and poor planning regulations have at the same time
meant that high quality land around cities was lost for urban purposes. Illegal
transfers of communal land along with fragmentation of holdings and youth
migration has added to these institutional problems, for example in Sudan, and
exacerbated the situation of poorer rural people. A lack of regulation has “under-
mined the strength of older institutions and the system of communal range man-
agement which were designed to manage resources in a sustainable manner”. In
addition, “Extension systems have often been very top down in design and delivery,
98 J.M. Bryden

leaving little scope for farmer-driven initiatives and partnerships”. These issues are
part of the debate on reform in governance and institutional frameworks which is
vital to tackle the manifest problems in the food-energy-water nexus and especially
the human rights issues discussed below.

5.5 Institutional Issues in Food, Energy, and Water

As we have seen, there are serious institutional issues to be addressed in dealing


with the separate and combined challenges of food, energy, and water in the Arab
Region. First and foremost is how to make sure that every person in the region has
access to food, water, and energy, a goal which is enshrined in the Millennium
Development Goals, and reinforced by the codification of human rights to water
and food. The human right to energy may be added to these in the future because it
is increasingly seen as crucial for the realization of other human rights. While a few
Arab states have not signed all Human Rights conventions, or have made reser-
vations on religious grounds, these reservations do not generally apply to rights to
water and food.
Food, water, and energy are essential for the subsistence of human beings today
and in the future, and all three are, or are potentially, products of rural regions, as
well as being needed and used by rural people. Ownership to land and water use
rights is a crucial determinant of who benefits—and generally the distribution of
benefits—from provision of food, energy, and water. The UN (1966) International
Covenant on Economic, Social and Cultural Rights (ICESCR) outlines require-
ments that are a common international reference—also for the most part adopted by
the Arab countries—and that underlie much thinking about sustainable develop-
ment. Among these requirements are: that no people shall be deprived of its own
means of subsistence (Article 1); everyone’s right to an adequate standard of living
(Article 11); everyone’s right to work (Article 6); workers’ right to fair wages, equal
treatment, healthy working conditions, union activity, and the ability to make a
decent living for themselves and their families (Articles 7, 10 and 12); states’
obligation to prevent hunger by improved methods for producing and distributing
food (Article 11). Food, water, and by implication energy, are human rights pre-
cisely because they are important for subsistence. This focuses attention first on the
need to remove absolute poverty, but, second, also on counteracting social injustice
(Bryden and Gezelius 2016).
The Sussex ‘New Manifesto’, argues that to achieve poverty reduction and social
justice, the innovation process must address the so-called ‘3Ds’. First, it must
include affected people, especially the poor and underprivileged, so that these
people can shape the direction of changes needed. Second, such an inclusive
approach aims toward promoting fair distribution as a direct outcome. Third, the
context-sensitivity that follows from this bottom-up and needs-oriented innovation
process promotes diversity, meaning that functional local solutions are prioritized
5 Water, Energy, and Food in the Arab Region … 99

over standardized, technocratic top-down solutions. The New Manifesto outlined a


“vision for innovation” where science and technology, based on the 3Ds, would
work more directly for poverty alleviation, social justice, and the environment
(STEPS 2010).
Following the duty-ethics approach of the 3D’s, it can be argued that decisions
about allocations of water, productive land, or energy cannot be made through the
(Pareto) utilitarian ethics of neo-classical economics, where both the distributional
impacts of any policy or change and everything that cannot be reduced to monetary
values is ignored. This also underlies the earlier argument that the principles of New
Public Management are usually incompatible with the human rights based approach
to food, water, and energy. A different approach to governance and institutions
around the food-water-energy nexus is therefore needed in the Arab Region, as
elsewhere. Land, for example, is a critical resource for food production and water
access for that purpose, since groundwater accessed through wells and boreholes
frequently belongs to the well- or borehole-sinker. Land tenure rights remain
unclear in many Arab countries, and if clear, may be based on western practises of
titles and open land markets. In Sudan, for example, the 1971 land tenure reforms
effectively transferred all communal (unregistered) land rights to the State, which
subsequently reallocated them to traders, retired government officials, army and
police officers, and mafia and corporate land grabbers often under the banner of
agrarian development. In this way access to land and associated water rights “as a
source of wealth and power remains one of the main differentiating factors between
the central and peripheral regions of post-colonial Sudan” and, as a result, one of
the keys to the wider conflicts in Sudan (Elhadary 2014; Anseeuw et al. 2012). As
Polanyi (1944) recognised, land (and water) as the basis of livelihood for many
people, cannot be simply treated as other commodities and left to market forces
because this creates fundamental insecurity, depriving them of the guarantees
provided by traditional social institutions including communal tenure.
The Gulf States are themselves one of the major ‘land grabbers’ of the world,
including in other Arab states with water resources, such as South Sudan. This
strategy has replaced earlier strategies of developing irrigated agriculture based on
diminishing ground water resources, a strategy that has rightly been abandoned.
However, land acquisitions in poorer developing countries in order to feed an other
countries’ people does have its contradictions and problems, especially if it serves
to deprive already poor people of their own access to food.
India’s National Food Security Act (2013) is an encouraging model because it
moves from a welfare to a human rights approach to food security and deals with
both physical access and affordability and, remarkably, it also contains universal
elements, namely the Mid-Day Meal Scheme and the Integrated Child Development
Services Scheme. The Public Distribution Service will reach about three quarters of
the rural population and half of the urban population (two-thirds of the whole
population). However, the Act does not deal with inequalities in access to water and
land, or indeed access to energy.
100 J.M. Bryden

5.6 Opportunities for Renewable Energy and Food

The Arab Region has enormous reserves of solar energy, and considerable wind
energy resources. Some countries also have hydro-electric resources, and these are
in some cases still under development (as in Ethiopia, for example). The oppor-
tunities for renewable energy also give further possibilities for innovation, adap-
tation, manufacturing, and meeting the energy needs of rural and other poorer
sections of the population (UNEP/REN21 2013). Better access to electrical energy
can allow the adoption of more efficient and effective irrigation systems in farming,
both saving water and increasing production. It is essential for better education and
literacy in the rural population in particular. It can replace the use of scarce biomass
currently used for cooking, while solar heating can replace wood used for heating.
This in turn saves much time of rural people (especially women) who currently
forage, over sometimes long distances, for woody and other biomass resources.
Electricity is also important for cooling, for example of the milk from the Awasi
sheep, or for perishable fruits and vegetables. Although renewable energy does not
create more water, it helps to move it from surplus to deficit regions, as well as
introduce more effective irrigation systems, and provide opportunities for food
preservation. The role of public institutions and governance in making sure that
such benefits arise in practice will be crucial, and will be enhanced by the Human
Rights based approach discussed above.
There are also more opportunities for urban and peri-urban agriculture which can
make much better use of recycled waste-water, as well as the residues from better
management of sewage and other biological waste materials, as well as renewable
energy.
The recent (2015) agreement on the use of the Nile waters between Egypt,
Sudan, and Ethiopia opens up a considerable hydro-power resource, as well as
irrigation potential. As in so many other parts of the region, the resolution of
conflicts, so often linked to water issues will be important for ensuring that the
benefits of such developments will be realized as well as fairly distributed.

5.7 Conclusions

A new approach needs to be taken to the integrated management of water, pro-


ductive land, and energy in Arab countries if the human rights to water and food are
to become a reality, and if all people are to have access to energy, as desired by the
MDG. This approach prioritizes human rights to water, food, and energy and
considers both physical and economic access issues. In so doing, it will give higher
priority to necessary changes in the rural communities, where decentralized par-
ticipatory community-based solutions to the governance of water, food, and energy
5 Water, Energy, and Food in the Arab Region … 101

are likely to become more important for the future. In this respect, communal land
rights, ignored and even subverted in many countries, and not only Arab countries,
must be recognized and restored. However, participatory approaches are no ‘magic
bullet’, and where imposed by external donors and agencies as ‘conditionalities’
can do more damage to reform processes than good. In particular, they may simply
return power to traditional local elites (Shortall 2002; Guthrie 2008).
As Abers (2003) said, “The success of participatory institutions depends on a
dual process of commitment-building. Unless both state actors (ranging from
politicians to bureaucrats) and ordinary people are motivated to support, take part
in, and respect (Empowered Participatory Governance) experiments, those policies
are unlikely to become either empowered or participatory.”
In addition, the creative and effective use of waste water, including the devel-
opment of ecological sanitation systems that can not only provide water purification
but also return nutrients to the soil, should also be applied in the growing cities and
towns of the region, and will offer opportunities for expanded urban and peri-urban
agriculture (Refsgaard et al. 2005). Dryland farming techniques, for example the
systems being studied at ICARDA in the region, will also become increasingly
important, especially in those areas with rain-fed agriculture. Within this limited
area, selective crop and livestock breeding and no-till farming, as well as small
investments in dams and storage for irrigation at critical moments, will be impor-
tant, as will improved extension services and peripatetic training for the often small
and resource-poor farmers involved.
We have argued that the problems of water, food, and energy need to be dealt
with at both the consumption and production ends. Water saving, recycling, cap-
ture, and storage will all play an important part at the consumption end of the water
problems. Reforming land and water rights, including the restoration and codifi-
cation of communal rights and communal governance systems and the prevention
of ‘land and water grabbing’, as well as adequate pricing of water used for large
scale irrigation, will be important features of governance reform at the supply end.
In developing decentralized renewable energy, developing systems of communally
owned and operated supplies in so far unserved rural communities will be impor-
tant, as well as stimulation of learning and innovation systems for the adaption of
renewable energy technologies to local conditions. In this way, renewable energy
can also contribute to rural livelihoods, and so to new investments in agriculture.
However, reformed governance must also consider the human rights based
approach to guaranteeing basic levels of affordable nutritious food, clean water, and
energy to people everywhere, exemplified by India’s recent National Food Security
Act. For both food production and renewable energy in the form of hydro-power,
resolution of the currently unsatisfactory lack of agreement between riparian
countries of the Nile basin represents a major handicap to progress, as it has been
for some decades at least.
102 J.M. Bryden

References

Abers RN (2003) Reflections on what makes empowered participatory governance happen. In:
Fung A, Wright EO (eds) Deepening Democracy. Verson, London, pp 200–207
ACWUA (2010) Wastewater reuse in Arab countries. Arab Countries Water Utility Association,
Amman, Jordon
Al-Asaad HK, Ebrahim AA (2008) GCC interconnection authority. The GCC Power Grid:
Benefits and Beyond, Dammam, Saudi Arabia
Anseeuw W, Alden Wily L, Cotula L, Taylor M (2012) Land rights and the rush for land: findings
of the global commercial pressures on land research project. ILC, Rome
AOAD (2012) Arab agricultural statistical yearbook. Khartoum, AOAD. Vol 32, http://www.aoad.
org/Agricultural_%20Statistical_Book_Vol32.pdf Accessed 2 January 2014
Awulachew SB, Yilma AD, Loulseged M, Loiskandl W, Ayana M, Alamirew, T (2007) Water
resources and irrigation development in Ethiopia. In: Working paper 123. International Water
Management Institute, Colombo, Sri Lanka
Awulachew SB, Erkossa T, Smakhtin V, Fernando A (2009) Improved water and land
management in the Ethiopian Highlands: its impact on downstream stakeholders dependent on
the Blue Nile. Dissemination Workshop, Addis Ababa, Ethiopia
Bradbrook A (2005) Access to energy services in a human rights framework. http://www.un.org/
esa/sustdev/sdissues/energy/op/parliamentarian_forum/bradbrook_hr.pdf accessed 5 Jan 2015
Bryden JM, Gezelius S (2016) Innovation as if people mattered: The ethics of innovation for
sustainable development. In Bryden JM, Gezelius S, Refsgaard K, & Sutz J (eds) Inclusive
Innovation in the Bioeconomy. Innovation & Development. Special Issue, Forthcoming 2016
Chatham House (2011) (Authored by Glada Lahn and Paul Stevens) Burning oil to keep cool: The
hidden energy crisis in Saudi Arabia. Chatham House, London p 40
Conniff K, Molden D, Peden D, Awulachew SB (2013) Nile water and agriculture past, present
and future. GCIAR. http://www.iwmi.cgiar.org/Publications/Books/PDF/H045309.pdf
Accessed 5 Jan 2015
Dixon J, Gulliver A, Gibbon D (2001) Farming systems and poverty: improving farmers’
livelihoods. In: A changing world, 3, Middle East and North Africa. Rome, Italy
El Gamal F (1999) Irrigation in Egypt and role of National Research Center. In: Proceedings of the
annual meeting of the mediterranean network on collective irrigation systems, Malta, 3–6 Nov
1999, Options méditerranéennes Séries B, 31 http://www.iamb.it/parIactivities/research/
option_B31.pdf Accessed 24 Apr 2012
Elhadary YAE (2014) Examining drivers and indicators of the recent changes among pastoral
communities of Butana Locality, Gedarif State Sudan. Am J Sociol Res 4(3):88–101
FAO (2013) AQUASTAT http://www.fao.org/nr/water/aquastat/main/index.stm Accessed
December 2013
FAO (2014) AQUASTAT http://www.fao.org/nr/water/aquastat/main/index.stm Accessed 5
January 2015
Guthrie DM (2008) Strengthening the principle of participation in practice for the achievement of
the millennium development goals. In: UN (DESA) 2008
IEA (2002) World energy outlook 2002. IEA, Paris
IEA (2006) World energy outlook 2006. IEA, Paris
IEA (2012) World energy outlook 2012. IEA, Paris
IEA/OECD (2009) World energy outlook. IEA, Paris
IFAD (2009) Fighting water scarcity in the Arab countries. International Fund for Agricultural
Development, Rome
Iza A, Stein R (eds) (2009) RULE: Reforming water governance. IUCN, Gland, Switzerland
Lampietti JA, Michaels S, Magnan N, McCalla AF, Saade M, Khouri N (2011) A strategic
framework for improving food security in Arab countries. Food Secur 3(Suppl 1):S7–S22
Oczkowski A, Nixon SW, Granger SL, El-Sayed A-FM, McKinney RA (2009) Anthropogenic
enhancement of Egypt’s Mediterranean fisheries. Proc Natl Acad Sci 106(5):1364–1367
5 Water, Energy, and Food in the Arab Region … 103

OECD (2012) Linking renewable energy to rural development. OECD, Paris


Pedrick C (2012) Strategies for combating climate change in drylands agriculture: synthesis of
dialogues and evidence presented. In: The international conference of food security in dry
lands, Doha, Qatar, November 2012. Aleppo, Syria and Copenhagen, Denmark. The
International Center for Agricultural Research in the Dry Areas (ICARDA) and CGIAR
Research Program on Climate Change, Agriculture and Food Security (CCFAS)
Polanyi K (1944) The great transformation: the political and economic origins of our time. Beacon
Press, Boston. English edition: Originally published by Farrar and Reinhart (2001). New York
Refsgaard K, Jenssen PD, Magid J (2005) Possibilities for closing the urban-rural nutrient cycles.
In: Halberg N, Alrøe HF, Knudsen MT, Kristensen ES (eds) Global development of organic
agriculture: challenges and promises. CABI ecowiki.org/GlobalPerspective/ReportOutline
RENAC Renewables Academy (2011) RENAC: Grid integration of large scale renewables in the
MENA region. Press Release Berlin 18 May 2011
Shortall S (2002) Social exclusion in rural areas: are area based partnerships the way forward?
Occasional paper presented at Queensland University of Technology. Queens University
Belfast: School of Sociology, Brisbane, Australia
Sowers J (2011) Water, energy and human insecurity in the middle east. Middle East Research and
Information Project, MER 271 http://www.merip.org/mer/mer271/water-energy-human-
insecurity-middle-east
STEPS (2010) A new manifesto. ESRC STEPS Centre, University of Sussex, Brighton p 24
UN (2014) World Water Report, vols 1 and 2
UN (DESA) (2011) Population division http://www.un.org/en/development/desa/population/
UNDP (2011) Arab development challenges report 2011: towards the developmental state in the
Arab region. Cairo, Egypt
UNDP (2013) Water governance in the Arab region: managing scarcity and securing the future.
United Nations Publications, New York
UNEP/REN21 (2013) MENA renewables status report. Paris, UNEP/REN 21 (Authors Bryden J,
Riahi L and Zissler R)
WHO/UNICEF (2014) Joint monitoring programme data on access to improved water and
sanitation. http://www.wssinfo.org/data-estimates/tables/ Accessed 1 Feb 2015
World Bank (2009) Improving food security in Arab Countries. Washington, DC http://
siteresources.worldbank.org/INTMENA/Resources/FoodSecfinal.pdf
WWAP (n.d.) http://www.unesco.org/new/en/natural-sciences/environment/water/wwap/
Chapter 6
Water, Energy, and Food Security
in the Arab Region: Regional Cooperation
and Capacity Building

Atef Hamdy

Abstract In the Arab Region, water and energy are recognized as indispensable
inputs to modern economies and are driven by the three imperatives of security of
supply, sustainability, and economic efficiency. In most Arab Countries, despite the
links and the urgency for security of supply in both sectors, in existing policy
frameworks, energy and water policies are developed largely in isolation from one
another—a degree of policy fragmentation that is seeing erroneous development in
both sectors. For the Arab Region, the absence of a comprehensive understanding
of the links between energy, water, and food security is quite evident. Integrated
policy and management strategies and solutions need to be better identified, as well
as understanding where barriers exist to achieve that integration. Understanding and
managing the energy-water-food security nexus means different things to different
people. Nevertheless, as a challenge deeply embedded in our quest for sustainable
development, a better understanding of the links between energy, water, and food
security is essential in any attempt to formulate policies (Hoff 2011). This requires
urgent attention to building relationships and linkages between policy making
institutions of the three sectors. As far as water, energy, and food nexus security in
the region is concerned, the critical new institutional challenge should be directed at
developing policies, rules, and organization and management skills to enhance the
ability to evaluate and address the crucial questions related to policy choices and
modes of implementation among development options. This will result in having
informed decisions that integrate the energy, water, and food production sectors
based on the understanding of environmental potentials and limits, and on needs of
the people in the country concerned. Capacity building is, and will continue to be an
integral part and a fundamental supporting tool for water, energy, and food security
nexus. To successfully meet the nexus challenge, effective capacity building is
needed to find sustainable solutions to the increasing problems related to water
scarcity, energy shortages, and the food production gap. The road map towards
effective capacity building requires the active participation of a wide range of

A. Hamdy (&)
Water Resources Management, CIHEAM-IAM Bari,
Via Ceglie 9, 70010 Valenzano, Bari, Italy
e-mail: hamdy@iamb.it

© Springer International Publishing AG 2017 105


K. Amer et al. (eds.), The Water, Energy, and Food Security Nexus
in the Arab Region, Water Security in a New World,
DOI 10.1007/978-3-319-48408-2_6
106 A. Hamdy

stakeholders and individuals with vastly different perceptions of the issues at stake.
It is equally important for capacity building to enlarge its scope and go beyond
education and training to encompass the wider issues of institutions and both formal
and informal organizations. Those issues, in addition to others required to establish
effective and successful capacity building development strategies, are the focus of
this chapter.

Keywords Water  Energy  Food  Security  Capacity building

6.1 Introduction

The Arab Region is challenged by population growth, urbanization, economic


growth, availability and consumption of natural resources, water scarcity related
problems, and impacts resulting from climate change.
The region is also facing deforestation and desertification challenges. Although
these challenges are common to several countries, cohesion and collaboration
across national boundaries in tackling them seems to be limited.
All of the Arab Region is water deficient, whereby withdrawal of water far
exceeds renewable water resources. In addition, a potentially changing climate will
simply accelerate freshwater security challenges (Hafez 2010).
Management of future water needs in Arab countries facing water resources
scarcity will require a policy rethink. Deeper and more integrated understanding of
exactly where and how to manage the intersection of water with various other
factors within the economy, such as energy options, industrialization choices, and
agricultural strategies is needed. However, this requires improved data collection,
fresh integrated economic building, and new cross-cutting collaboration capacity
building programmes and institutional platforms (World Bank 2009a; AWMC
2011).
Regarding the energy resources in the region, it is quite evident that despite its
significant availability in some Arab countries, it still remains a financial and
economic burden on the economies of some other countries of the region. The
apparent disparities in the distribution of natural resources, including land, water,
and energy in the Arab Region are major driving forces in enhancing knowledge on
how to approach security in those sectors (LAS 2013).
For most Arab countries, increasing access to those vital resources is now the
major challenge. Challenges in securing enough water for energy and energy for
water will increase with population and economic growth, in addition, competition
for water resources will intensify and climate change will compromise solutions
(Hamdy et al. 2014).
6 Water, Energy, and Food Security in the Arab Region … 107

This nexus is evident in most arid countries of the Arab Region, those where
water, energy, and food resources face considerable stress.
The water food linkage represents another important and vital nexus in Arab
countries. The ability for the Arab countries to feed their growing population is
severely challenged by competition over increasingly scarce water resources. The
idea of food self sufficiency is now abound, and it is no longer rational or
sustainable.
Nowadays, assuring available water to fulfil the enormous gap in food pro-
duction in the region is the main challenge facing the farming sector. This shows
the strong relationship between Arab food security and water security, which are
both highly dependent on available energy (OECD 2011).
Currently, these links in and between water, food, and energy are not adequately
reflected in decision-making. Water, food, and energy policy are made in isolation,
without fully considering the implications for the other policy areas.
Therefore, to become more relevant, the water-food-energy nexus needs to get
beyond its superficial level and articulate what these links mean in practice.
The nexus initiative in the Arab Region should have the following three key
roles; to advance scientific knowledge of cross-sectoral interactions, globally and in
specific areas, to provide decision-making support, and tools to build awareness and
understanding of the water-energy-food nexus. These roles correspond to the three
pillars; research, policy engagement, and communications, and draw on the
expertise of a large number of scientists involved in nexus activities.
In the Arab Region a roadmap needs to be established on:
• how to implement nexus policies to increase efficiency of natural resources
management
• how to bridge science with policy business
• how can we learn from each other and how can we collaborate to address the
challenges ahead
At a regional level, there is a need for considering trade-offs among water,
energy, and agriculture policies in river basins, in particular through better coop-
eration between basin organizations and regional energy institutions.
Benefit-sharing mechanisms are a promising approach in this regard (Ringler et al.
2013).
Creating a political shift for movement towards resource efficient growth can be
achieved through the setting of new partnerships and collaboration involving
international agencies, international and domestic private sector actors, major
investors, NGOs, and academics. Equally, this action cannot be assured without
being backed with efficient capacity building programmes provided to the stake-
holders involved in water, energy, and food security (ADB 2013).
108 A. Hamdy

6.2 Nexus and South-South Cooperation (SSC)

Regional cooperation started the year 2008 by addressing the “Riyadh Declaration”
(LAS and AOAD 2008), that calls for sound trade and investment schemes for
enhanced food security in the short and long terms through inter-Arab agriculture
trade and public-private production ships (LAS and AOAD 2008). Equally, con-
secutive AFED reports (2010, 2011, 2012) reached similar conclusions empha-
sizing the important role of regional cooperation in enhancing food security in Arab
countries.
A regional strategic approach to address knowledge, communication and
resource gaps for nexus development in the Arab Region could be characterized by
(UNDP 2014):
1. Cooperation for nexus development in the Arab Region is premised on the
following assumptions.
– South-South Cooperation (SSC) as a driver for development, including its
dimensions that are directly or indirectly related to the Arab States is an
evolving process.
– A harmonized and coordinated effort could help to further catalyze and speed
up this process for the Arab Region by optimizing its intensity and focus.
This could also lead to an increase in the bargaining capacity of the region as
a whole, leading in turn to a stronger Arab voice in the global arena.
2. SSC clearly constitutes a shift from a single-sector to a multi-sector type of
relationship between countries in the region, however it must be kept in mind
that the countries are different in terms of the actual and anticipated structure and
proportion of inward and outward cooperation flows. In addition, the wealthier
countries in the region are poised to not only benefit from their role as partic-
ipants in SSC, but also supporting exchanges between other countries by filling
the financial resource gaps.
3. Building on existing institutional arrangements, capacities developed, and other
resources available (avoiding duplication of arrangements, capacities, and
resource facilities), while implementing the proposed approach can reduce
set-up costs and serve as a driver towards a more harmonized and inclusive
approach.
4. National-level coordination is crucial to ensuring scaled-up SSC within and
beyond the region, however, the approach should provide for harmonious and
non-bureaucratic horizontal linkages, along with local leadership and
ownership. In addition, it should be inclusive and systematic in the sense that
supportive activities should benefit as many existing and potential South-South
exchanges as possible involving various government civil-society and private
sector players.
6 Water, Energy, and Food Security in the Arab Region … 109

5. Finally, an understanding should be developed that standardization of individual


national-level approaches towards documenting and presenting expertise and
solutions, and defining needs may be crucial for scaling up SSC across the Arab
Region and beyond, while at the same time, being flexible and recognizing that
each country in the region is different.

6.3 Water, Energy, and Food Nexus:


The Capacity Building

Capacity building is and will continue to be the central part and a fundamental
supporting tool for water and energy resources development and management, in
general, and in the agriculture sector in particular. In the past few years, many Arab
countries received notable financial and technical support provided by international
organizations and donors, and this has broadened the range of capacity building
programs to be implemented by national institutions and regional organizations.
However, in spite of the achieved progress in both water and energy sectors, we are
still far away from our expectations. The irrigation sector is still experiencing poor
performance, low water use efficiency, bad water governance, and relatively high
water losses. In arid and semiarid developing countries in the Arab Region where
water crises are emerging and rapidly growing, hitting any possible development,
effective capacity building in water and energy resources, irrigation, and drainage
sectors, is needed to find sustainable solutions to the complex challenges those
countries are now facing. The roadmap towards effective capacity building requires
the active participation of a wider range of stakeholders and individuals with vastly
different perceptions of the issues at stake (Hamdy and Lacrigniola 1999).
Meeting challenges of water, energy, and food security nexus will depend on
these stakeholders having access to science-based information and on the capability
to evaluate the consequences of their decisions and actions. Equally important, is
for capacity building to enlarge its scope and go beyond education and training to
encompass the wider issues of institutions and involved formal and non-formal
organizations.
In the Arab Region, meeting such challenges requires additional efforts and
tedious work for a greater improvement in the water and energy sectors capacity
building. Equally, we still have to learn more on how capacity building should be
implemented and how we can avoid erosion of existing capacity. Capacity building
concepts, tools and instruments, benchmarks, and economic significance need to be
better addressed (Hamdy et al. 2014).
This calls for applying capacity building instruments, notably technical training
and education, to provide the water and energy sector institutions with human
resources having a greater capability to utilize, operate, and maintain the infras-
tructure for a longer time, as well as correctly planning, managing, and allocating
110 A. Hamdy

the available water resources and energy among the competing sectorial uses. This
should be done in coordination with updating and strengthening of the institutions.

6.4 Water and Energy Sector Capacity Building:


Experiences Gained and Lessons Learned

Over the past decades, in the Arab Region both sectors have benefited from large
investments in infrastructures. However, the question is: Are we in need of new key
characteristics of water and energy sector capacity building? If the answer is yes,
what are the reasons, and what are the key issues that should be introduced for
strengthening both sectors?
Nonetheless, many of these projects have partially failed to deliver the expected
benefits. The reasons for these shortcomings could be outlined in (World Bank
1996):
• strengthening the water, energy, and agricultural sectors by concentrating only
on investments and construction. The error was that emphasis was not directed
to institutional improvement projects and projects became structural parts of
longer-term programs
In addition, many failures can be attributed to systematic deficiencies in the
institutions that determine policy, project designing, and operational management
due to:
• absence of long-term integrated vision, and failure to account for constraining
factors such as limited water availability and steadily increasing demand;
• weak management of the organizations, such as utilities that are supposed to use
the new infrastructure;
• absence of a consistent economic rationale for the selected water and energy
projects, as well as poor prioritization and technology selection; and
• uncertainty on how to measure capacity or progress in building it.
Nowadays in most Arab countries, technically and politically, the importance of
building the capacities of the individuals, organizations, and institutions for an
appropriate water-energy-food security nexus approach is well recognized.
However, the capacity building programs are mostly geared to the managers and the
technical staff, and fully exclude the stakeholders, i.e. the users. The participation of
the individuals in the process has been ignored.
The top-down approach of many Arab governments and aid agencies failed to
produce the capacity needed. This could be one of the major causes behind the
failure of many of the projects undertaken in both water and energy sectors. Policy
reform, sector strengthening and reform, and effective regulations are impossible
without adequate participation of all stakeholders in the process of their develop-
ment at national, as well as local levels (Sidiqi and Anadon 2011).
6 Water, Energy, and Food Security in the Arab Region … 111

Indeed, the more effective the participation is, the lower the need will be for
top-down regulations, as participation will facilitate the exchange of information
and its dissemination.
We have to learn from the past to avoid repeating the mistakes in our new
capacity building supporting programs.

6.5 Capacity Building Improvement Is a Primary Need


in the Agricultural Sector to Achieve Food Security

Among the different sectorial water uses, agriculture is the highest in its water
consumption, receiving the lion’s share of the available water. Globally this stands
at around 60 %, while in the developing Arab countries it could reach values
exceeding 80 %. On the other hand, on-farm water use efficiency is very low, not
exceeding 45 % (Hamdy 2010).
The literature presenting several case studies regarding water use efficiency in
the irrigation sector and the associated water losses in most of Arab countries (FAO
1997, 2006, 2007, 2008; World Bank 2007, 2009b, c; Blue Plan—UNED 2002;
Louati 2008; AFED 2014a, b; ESCAW 2007; ICARDA 2011), all came to the
conclusion that in practice, too much of the irrigation water does not reach the
crops, and irrigation systems fall into disrepair almost as fast as new irrigation
schemes can be developed. The costs of this partial failure in most projects are high,
food production suffers a continuous increase in the food gap, leading to difficulties
in achieving both food and water security.
This requires changes in the existing mode of thinking but, for this to happen,
there is a clear need to improve the irrigation water sector performance as to the
following issues: efficiency, water allocation, effectiveness, and sustainability.
Nowadays, in the Arab Region there is a growing recognition that a new way is
needed for pursuing sustainability in the irrigation sector (Fig. 6.1).
Such a new way should be based on realistic planning and an integrated action
approach identifying the prerequisites of the stakeholders and assuring their full
participation in the process, building the needed capacities at all levels to guarantee
the presence of effective and efficient enabling conditions.

Fig. 6.1 A new approach to


sustainability quest in the Enabling
Capacities
irrigation sector system

Sustainability

Stakeholders
Prerequisites
participation
112 A. Hamdy

6.6 Nexus Approach Implementation:


The Required Capacities

Capacity development supporting programmes differ in the knowledge domain they


are addressing: some emphasize the technical domain, while others link technolo-
gies to the economic feasibility of their use. Integrating business skills, management
techniques, and organizational strengthening is less common, while programmes
with a focus on interactive policy-making, participation, and empowerment are
quite scarce.
For the region, capacity building programmes are needed to promote knowledge
and data sharing as well as the development of appropriate solutions for the
obstacles facing the implementations of the nexus approach.
To characterize the type of capacity development supported by the various
countries several variables have to be taken into account, such as: WWAP (2014).
1. The level of intervention (individuals, organizations, institutions);
2. The capacity domain to be strengthened (technical, economical and/or social
subject matter and/or methodology, managerial issues and organizational
capacity, or interactive policy-making and strategy design);
3. Categories of beneficiaries aimed at (scientists, producers, research managers,
education staff, trainers, extension agents, staff from support services);
4. Methodology applied (coaching, mentoring, short courses, formal education,
face to face interaction, distance/e-learning); and
5. Institutional delivery arrangements (individual scholarships, fellowships as part
of a research programme, or projects for organizational strengthening, institu-
tional partnerships, basket funding marked for capacity development in general,
competitive grant schemes, public tendering, etc.).
Capacity building, whether focused on individuals (skills-based training and
professional development) or institutions (management systems, coordination and
considered institutions), it should be a major element in the planning process, as
well as a significant aspect of the water, energy, and food security nexus.

6.7 Building Stakeholders Capacity

When considering capacity building in the water, energy, and food nexus, the key
requirement is to establish the capability in the different stakeholder’s communities
(including users, service providers and agencies, politicians, donors, and broader
society) to make critical assessments of the needs of the local situation. The
knowledge and capacity for analysis required by these different groups will vary
considerably.
In the Arab countries a great deal of attention should be given to enhance
meaningful stakeholders’ participation; and to bring diverse stakeholders to a
6 Water, Energy, and Food Security in the Arab Region … 113

common understanding of the objectives of the development intervention and the


planning processes for water management.
Any institutional change, whether strengthening or reform, implies that
numerous stakeholders be consulted for having their agreement and support.
Unfortunately, in the water sector in general and the irrigation one in particular,
important but often unrecognized stakeholder groups are women. Indeed, for many
Arab countries, imbalances between female and male users in access to water and
other services continue to be a problem, despite gender balance being a priority on
the agenda of policy makers, interventions agencies, water managers, and
researchers (Van Koppen 2002).

6.8 Building the Institutions Capacities:


Major Considerations

Creating appropriate conditions to build up the capacity of water, energy, and food
production institutions in Arab countries within a reasonable time frame, consid-
eration should be given to the following (Hamdy and Lacirignola 1999):
• Major efforts are urgently needed to assess national capacity building require-
ments that are not available at present for many Arab countries: it is essential to
plan, construct, and elaborate the programs required to properly build up the
institutional capacities.
• Natural resources and energy institutions should be dynamic and must contin-
uously change if they are to best match the evolving conditions. Problems
should be updated and evaluated periodically to adopt existing plans and to
formulate new programs.
• The experts sent by the External Support Agencies (ESAs), or invited by
national governments should have a wide knowledge of the physical, social,
institutional, legal, and cultural conditions and practices of the country
concerned.
• For effective capacity building, the most essential requirement is having a good
cadre of capable senior managers. My sense is that even with the best policies,
laws and institutions, and adequate availability of funds, if the right people are
not there, progress at best can only be slow and marginal.
• Universities and research institutions have to be fully involved in educating and
training people who can successfully plan, operate, and manage the nexus
approach, under the rapidly changing social, economic, environmental, and
political conditions.
• Opportunities should be taken for the networking of institutions, which is a
powerful tool for capacity building, and in particular for the exchange of
experiences and capacities between developing countries in the region.
114 A. Hamdy

6.9 Nexus and Capacity Development: The Need


for Comprehensive Education Programmes

When we talk about nexus capacity building in the water, energy, and food pro-
duction sectors we mean supporting the capacity of all categories at the different
levels involved, as well as creating the enabling conditions and the necessary tools
required for appropriate and efficient water and energy use and management, and
improving sectors’ performance (Hamdy 1997; Hamdy et al. 2000).
To develop capacities to implement the nexus approach, priority areas are rec-
ognized to be human resources development (HRD) and sharing of knowledge and
networking. In general, most governments and water organizations in the Arab
Region are well aware of the need of education and training. However, typically,
coherent and sustained strategies are absent. Commonly, scarce resources for
education, training, and data collection are spread too thinly over an array of
universities, institutes, and organizations. As a rule, most organizations involved in
nexus activities still view HRD as a way to consolidate the existing “unchangeable”
organization rather than looking at HRD as a means to change the staff, and use the
staff to change the organization into a better performing one (Hamdy 2002).
In addition, there is a need for a comprehensive and collaborative education
programme, aimed at advancing the goal of integrating planning, policy, and
management of water, agriculture, and energy to achieve security of all three. Such
a coordinated, comprehensive, and collaborative educational program can effec-
tively reach a significant number of people.
An educated public would be more receptive to consumer-level changes required
to achieve water, energy, and food security, which is essential to a sustainable
future for all nations.
One helpful way to illustrate the nexus is through “water foot-print.” Water is
embedded in the food we eat, the energy we use and the products and materials we
need. Making these complex interconnections understood on a personal level is an
essential step in gathering public support for the difficult policy initiatives.

6.10 Capacity Building Strategy:


The Holistic Joint Learning Projects

Nowadays with the acute shortage in available water resources, the notable increase
in energy demand, and the increasing gap in food production, following a “doing as
usual” approach is not valid. Programs for HRD should be updated and tackle the
heart of the issues water professionals are currently facing. In this context, holistic
joint learning projects are recommended as a promising capacity building strategy.
Indeed, it matches HRD with institutional development and creates a virtual space
for experimenting with different approaches to solve the sectors’ problems in a
holistic way. Such projects rely on involvement at the national and local levels, of
6 Water, Energy, and Food Security in the Arab Region … 115

agencies, universities, and key actors for decision making thereby, bringing toge-
ther the wide linkage gap already existing among the concerned stakeholders:
political, institutional, professional, water users, and the community.

6.11 Educational Training Programme: The New


Direction for a Successful Nexus Approach
Implementation

For most countries of the region there is an urgent need for a new direction in
educational training programs to mainly concentrate on (Fukuda-Parr et al. 2002):
• increasing the ability of professionals to deal with water sustainability in a
holistic fashion;
• addressing the need of the professionals to acquire multiple skills while
increasing recruitment and adaptive training;
• including HRD as a fundamental, mainstream action rather than a side issue;
• better disseminating information about water and energy to generate an
understanding of/and broader public support of needs to safeguard the limited
natural resources;
• building the capacity of local capacity builders: “train the trainers”;
• retaining trained personnel. More effort should be made by the water, energy,
and food production organizations to stimulate the trained staff to put new
knowledge and insight into practice;
• step-up conceptual work on capacity building tools and instruments including
criteria and benchmarks to assess progress.

6.12 A Tailor-Made Package for Education,


Training, and HDR

For the water, energy, and food security nexus, this package should apply tech-
niques other than the common classroom teaching, and should increasingly draw
from a larger variety of tools such as (Abu-Zeid and Hamdy 2010):
• distance learning;
• real-life simulation workshops and training sessions to build up:
– strategy development
– problem solving skills
– decision-making
– skills to work in multi-disciplinary teams
116 A. Hamdy

– continued professional education and training in-service, on site and


in-project training;
• networking on both national and regional levels.

6.13 Water, Energy, and Food Nexus:


Research Programme

For the Arab Region, research policies should be focusing on enhancing the uptake
of inter- and trans-disciplinary research approaches and strengthening the scientific
collaboration between natural and social scientists, by encouraging joint research
projects. The programme should also aim to decrease the existing knowledge and
communication gaps between scientists and users of research, ensuring the inte-
gration of decision makers, the private sector, and civil society more broadly in
running programmes.
Concerning the nexus approach, it is to be asked, how should research be shaped
in the Arab Region? And what are the major options?
In the region, these options can be outlined in the following (FAO 2014):
• enhance collaboration among countries in the region on global environmental
change and sustainability research. A first step might focus on mapping existing
regional networks and ongoing projects. Databases of relevant experts and
research institutions should be established;
• leverage national and regional research funding on water, energy, and food
production;
• promote scientific networking and provide a platform for engaging stakeholders
and reaching out to policy makers;
• support and facilitate capacity building and training activities in the region for a
better understanding of the nexus approach;
• ensure the continuity of these activities over time, and recognize the importance
of ensuring policy makers’ support.
Further enhancement of the existing stakeholder engagement in research activ-
ities could be realized by fostering dialogues and encouraging collaborative
research projects, exchanging best practices among countries, and carrying out
annual meetings between representatives of the government and the scientific
community.
It is much easier to identify cross-sectoral interactions in research than to address
them in practice. To operationalize the nexus, we need to understand conditions on
the ground, including the perspectives of different institutions and stakeholder
groups, at all relevant scales. Participatory processes are thus crucial, and can
produce more viable solutions to resource challenges (Bonn 2011).
6 Water, Energy, and Food Security in the Arab Region … 117

6.14 Building Research Partnerships

In the Arab Region there are several potential partners. The question is not only in
their presence, but, on how to engage them to work together. The partnerships
among research institutions such as: Masdar Institute of Science and Technology,
Kuwait Institute for Scientific Research, King Abdullah University of Science and
Technology (KAUST, Saudi Arabia), Biotechnology Research Centre, Libya and
Gulf Research Centre, among others, should be built on the strength of ongoing
work being conducted at such national and regional institutions and organizations.
In addition to such scientific and research institutions, the Arab Region is also
rich in other regional and NGOs organizations (the Arab network of UNESCO Man
and Biosphere Programme, The Arab Water Council, TWAS Arab Regional Office,
The Arab Academy For Science, Regional Organization for the Protection of the
Marine Environment, United Nations Economics and Social Commission for
Western Asia [ESCWA], Union for the Mediterranean, UNEP’s regional office), all
of whom can be relevant regional actors helping to form a backbone of strong
partnerships.
The Arab Region is rich in research institutions and a great number of scientists
in the water and energy sectors, as well as food production systems, however there
is no coordination or cooperation among them. The dissemination of research
findings among the national, as well as the regional institution are poor. Equally,
very little is known about initiatives and policies, lessons learned, best practices and
tools related to interdisciplinary studies, and co-designed research on water, energy,
and food security nexus for most Arab countries (Granit and Löfgren 2010).

6.15 Water Research and Public Water Policies

Water research organizations in the Arab Region are hampered by a lack of human
and financial assets and the absence of national assets and inadequate national
science and technology policies. Research agendas sometimes reflect the require-
ments of international funding organizations, rather than local community needs
and national goals. Underfunded, understaffed, poorly performing research orga-
nizations, in addition to poor quality and unavailability of data, continue to dom-
inate regional water research.
There is a lack of capacity for research and development in the Arab Region,
lack of integrated resource management, and little interaction between policy
makers and scientists. The political process is critical to implementing the nexus.
Politics is crucial to understand the roles of state institutional and organizational
structural aspects (rules, norms, and institutions) in shaping, confining, and other-
wise conditioning nexus related decision making.
Research and innovation are critical to setting the stage for effective water
policies that ensure sustainability, efficiency, and equity in access and use of the
118 A. Hamdy

Arab Region’s scarce water resources (Hamdy 2009). Therefore, water research
into institutional aspects should involve identifying arrangements that may con-
tribute to a functional, coordinated approach to important water-linked sectors such
as energy, agriculture, and the environment. It should include how compatible
policies in the interlinked sectors can be developed and implemented (Hellegers
et al. 2008).

6.16 Concluding Remarks and Recommendation

• In the Arab Region addressing nexus challenges will need to do more than
quantify biophysical interdependencies. This might sound obvious, but up to
now the nexus debate is surprisingly devoid of actors, i.e. people who are
actually affected by nexus challenges or those trying to address them. The lack
of people or actors in nexus thinking and analysis has the effect that much of the
nexus discourse happens in relatively abstract terms. Resources should be used
more effectively, sectors should be integrated, and policies should be harmo-
nized, but it is not clear who should do this. Who might be willing or able to do
this and how? Who will win and who will lose from the proposed changes?
• The strategic nexus action perspectives for the Arab Region should explore a set
of limited, albeit important questions to address nexus challenges. First, how to
identify nexus challenges, i.e. what is the problem/solution? Second, how to
disentangle interconnections, i.e. what can and should be integrated? And third,
how to respond to nexus challenges, i.e. what are conceivable ways forward?
For the region, there are three complementary perspectives to explore these
questions. An actor perspective to inductively investigate what the nexus
challenges are: network perspectives to conceptually and methodologically
disentangle connections; and a governance perspective to reveal new realities
and opportunities of collective action. As any conceptual, theoretical, or
methodological framing, the strategic action perspective is limited. To advance
nexus thinking a plurality of approaches will be needed.
• In most of the Arab countries there is an obvious scarcity of scientific research
and studies in the field of water-energy nexus and the interdependencies
between these two resources and their mutual values, which is leading to a
knowledge gap on the nexus in the region. Moreover, with climate change
deeply embedded within the water energy nexus issue, scientific research on the
nexus needs to be associated with the future impacts of climate change.
• In the Arab Region each country has to develop its own strategy for shifting to a
capacity and capability building thrust in the area of water and energy man-
agement, that recognize the need for operational integration of environment and
development with long-term commitment, through a participatory process. As a
consequence, a more holistic approach to strengthen the capacity of both indi-
viduals and organizations is required.
6 Water, Energy, and Food Security in the Arab Region … 119

• Adequate human resources, with skills training and experience in the scientific,
technical, managerial, and administrative functions is essential for the devel-
opment, conservation, and management of water and energy resources. There is
a need for training in planning, project monitoring, and evaluation.
A cost-effective way to train technicians on a continuous basis may be to set up
training schools that can train technicians for the various sub-sectors.
• As we apply nexus tools to policy challenges, we need to keep building sci-
entific knowledge. We also need to keep working to ensure we have clear
definitions, well-tested methodologies, and transparent, well-designed integrated
models. And we need to expand our perspective, focusing more on integrating
socio-economic and biophysical perspectives, and looking across scales, from
the local to the global. We need better tools to quantify cross-sectoral interac-
tions, as well as insights into how institutions and policies account for and
manage these interlinkages.
• Research institutes and universities need to be encouraged to direct their aca-
demic and research programs towards understanding the nexus and their inter-
dependencies and inter-linkages. Without the availability of such researches and
studies, the nexus challenges cannot be faced and solved effectively, nor can
these challenges be converted into opportunities in issues such as increasing
efficiency of water and energy use, informing technology choices, increasing
water and energy policy coherence, and examining the water-energy security
nexus.
• For an appropriate implementation of the nexus approach, the Arab Region
(Organization) needs to be in cooperation with governments, universities, and
research centres, as well as public and private stakeholders to support national
and regional efforts to deepen the knowledge of the relationship between, water,
food, and energy and to meet the requirements of socio-economic and envi-
ronmental development in the Arab Region.
• The prospects of integrating the use of water, agriculture, and energy resources
at the national level are wide, and can be studied through a scientific method-
ology that could lead to the identification of best practices with a positive impact
on the economic and social development. In this context, it is imperative for the
countries of the Arab Region to formulate multi-sectoral policies and adopt
development practices capable of achieving integrated and comprehensive
planning and management in the areas of energy, water, and food security for
current and future generations. A regional cooperation that involves
Governments, universities, and public sector stakeholders can support the
achievement of this goal.
• Wide-spread reforms in the energy and water sectors require serious institutional
and policy measures. Overconsumption cannot be checked, efficiency measures
cannot be adequately implemented, and renewable energy cannot spread out if
current subsidy regimes are not phased out. Private sector participation in the
energy and water sectors requires that policy makers establish the appropriate
enablers, including well-defined policies and sound regulatory frameworks.
120 A. Hamdy

• There is general recognition of the need for a sustainable increase in levels of


investment in nexus capacity development programmes if water energy and food
security are to be met. However, to utilize the increased resources in an effective
and sustainable way requires the enhancement of the organizational and insti-
tutional capacity of the major stakeholders in the innovation process to improve
the relevance, quality, and scale. Equally there should be re-balancing of pro-
viding support at the individual level on the one hand, and at the organizational
and institutional level on the other. More resources should be directed towards
supporting vocational training for young people in natural resources and food
production, and such capacity development should include support to learn
business management and entrepreneurial skills, as well as guidance on the use
of new information and communication technologies.
• Capacity strengthening initiatives should focus more on organizational and
institutional strengthening as opposed to individual training. Multi-stakeholder
initiatives should be pursued, as the available evidence suggests these are likely
to lead to larger impacts than those exclusively targeted towards individual staff
at research organizations and institutes of higher education.
• Monitoring, evaluation, and impact assessment procedures should become a tool
for institutional learning for the stakeholders involved in, and affected by the
capacity development supporting programmes on water, energy, and food
security. As well, elaborated systems for participatory Monitoring and
Evaluation (M&E) is an excellent mechanism to support learning and knowl-
edge exchange for all parties involved. An increase the number of the pro-
grammes functioning in this way is needed in order to strengthen the sharing of
good practice and experiences among the European countries and developing
Arab countries.
• Arab countries, both oil exporters and importers, are well endowed with
renewable sources of energy, primarily solar. For now, these are underutilized.
Together with enhanced energy efficiency and cleaner technologies, these
renewable sources can help diversify and power a more sustainable future. Key
prerequisites are aggressively advancing science and technology, and above all
pursuing regional cooperation.

References

Abu-Zeid M, Hamdy A (2010) Encyclopedia on water resources development and anagement in


Arid and semi-Arid Regions of the Arab world. In: Abu-Zeid E, Hamdy A (eds) Capacity
building for water resources development and management, vol 12, p 214
ADB (2013) Thinking about water differently: managing the water-food-energy nexus. Asian
Development Bank (ADB), Mandaluyong City, Philippines
AFED (2010) Arab environment. Water: sustainable management of a scarce resource. Retrieved
from: http://www.afedonline.org/Report2010/main.asp
6 Water, Energy, and Food Security in the Arab Region … 121

AFED (2011) Arab environment 4. green economy: sustainable transition in a changing Arab
World. Beruit, Lebanon. Retrieved from: http://www.afedonline.org/Report2011/PDF/En/Full-
eng.pdf
AFED (2012) Arab Environment 5. Survival options: ecological footprint of Arab countries.
Beruit, Lebanon. Reetrieved from: http://www.footprintnetwork.org/images/article_uploads/
Survival_Options_Eng.pdf
AFED (2014a) Institutional challenges for water—energy nexus. Arab perspective. In: Saab N
(ed) Thematic debate on the UN general assembly on the role of water, sanitation and
sustainable energy in the post 2015 development agenda. New York, pp 18–19, Feb 2014
AFED (2014b) Water efficiency handbook. Lebanon, Beruit, p 94
AWMC (Arab Ministerial Council for Water) (2011) AWMC strategy for water security in the
Arab region. League of Arab States, Cairo, Egypt
Blue Plan (Plan Bleu)—UNED (2002) Analysis of strategies and prospects for water in Tunisia,
Sophia—Anatoplis/ Marseilel, Retrieved from: www.planblue.org/publication
Bonn (2011) Messages from the Bonn 2011 conference: the water energy and food security nexus-
solutions for a green economy. In: The water energy and food security nexus—solutions for a
green economy Bonn. Bonn, Germany
ESCWA (2007) ESCWA water development report 2: state of water resources in ESCWA region.
Report E/ ESCWA/ SDPD/ 2007/ 6, ESCWA, Beirut, Lebanon
FAO (1997) Irrigation in the near east in figures. Water report No. 9, Rome
FAO (2006) AQUASTAT database Retrieved from: www.fao.org/ag/agl/aglw/aquastat/dbase/
indxstem.rom
FAO (2007) Improvement of irrigation water management in the kingdom of Saudi Arabia.
MTFN/SAM/011/SAM
FAO (2008) AQUASTAT database, Lebanon water report, 34. Food and Agriculture
Organization, Rome
FAO (2014) The water-energy-food nexus at FAO. Concept Note, Rome
Fukuda-Parr S, Lopes C, Malik K (2002) Institutional innovations for capacity development. In:
Fukuda-Parr S, Lopes C, Malik K (eds) Capacity for development—new solutions to old
problems. Earthscan Publications, New York, pp 1–21
Granit J, Löfgren R (eds) (2010) Water and energy linkages in the middle east—regional
collaboration opportunities. SIWI Paper 16, Stockholm International Water Institute,
Stockholm, Sweden, Apr 2010
Hafez M (2010) Water in the Arab world. Al- Ahram Center for Strategic and Political Studies
Hamdy A (1997) Capacity building for environmentally sound water management. In:
International conference on water management, salinity and pollution control towards
sustainable irrigation in the mediterranean region, vol 5, pp 11–25
Hamdy A (2002) Capacity building in water resources sector: human resources development. In:
Advances in soil salinity and drainage management to save water and protect the environment,
15–27 Oct 2002, Alger, Algeria, pp 443–473
Hamdy A (2009) Water crisis and food security in the Arab World and the future challenges.
CHIEAM, The Mediterranean Agronomic Institute of Bari, Bari
Hamdy A (2010) Improving the irrigation water sector performance and the need of effective
capacity building programs. In: Abu-Zeid M, Hamdy A (eds) Encyclopedia on water resources
development and management in Arid and semi-Arid Regions of the Arab world, vol 12,
pp 19–63 (Capacity building for water resources development and management)
Hamdy A, Lacirignola C (eds) (1999) Mediterranean water resources: major challenges towards
the 21st century. CIHEAM-MAI-Bari, Italy, p 570
Hamdy A, Lacirignola C, Trisorio-Liuzzi G (2000) Education, training and distance learning in
water sector capacity building. In: International conference on wadi hydrology. Sharm El
Sheikh, Egypt, 21–23 Nov 2000
Hamdy A, Driouech N, Hmid A (2014) The water-energy-food security nexus in the
mediterranean: challenges and opportunities. In: 5th International scientific agricultural
symposium, agrosym 2014, 23–26 Oct 2014
122 A. Hamdy

Hellegers PJGJ, Zilberman D, Steduto P, McCornick P (2008) Interactions among water, energy,
food and environment: evolving perspectives and policy issues. Water Policy 10(Supplement
1):1–10
Hoff H (2011) Understanding the nexus. In: Background paper for the Bonn 2011 conference. The
Water, Energy and Food Security Nexus, Bonn, Stockholm Environment Institute, Stockholm
Sweden
ICARDA (2011) Water and agriculture in Egypt. Presented at the Egypt-Australia-ICARDA
workshop On-Farm Water-use efficiency. International Center for Agricultural Research in the
dry areas, Cairo, Egypt, Jul 2011, p 84
LAS (2013) Arab strategy for the development of renewable energies (2010–2030). League of
Arab States
LAS and AOAD (2008) Riyadh declaration to enhance Arab cooperation to face world food crises.
League of Arab States and Arab Organization for Agriculture Development (AOAD),
Khartoum, Sudan. Retrieved from: http://www.aoad.org/strategy/RiadhDeceng.pdf
Louati MH (2008) Blueplan, efficiency of water utilization. National studies: the case of Tunisia
OECD (2011) Water, energy and agriculture: meeting the nexus challenge. OECD, Paris
Ringler C, Bhaduri A, Lawford R (2013) The nexus across water, energy, land and food (WELF):
potential for improved resource use efficiency? Curr Opin Environ Sustain 5(6):617–624
Siddiqi A, Anadon LD (2011) The water-energy nexus in Middle East and North Africa. Energy
policy 39(8):4529–4540
UNDP (2014) Mapping south-south cooperation: mechanisms and solutions in the Arab states.
Global South-South Development Academy Series, Retrieved from: http://www.undp.org/
content/dam/rbas/doc/South-South/UNDP_Mapping_South_South_Cooperation_Expo_Booklet_
Doha_Qatar_En_Jan_14.pdf
Van Koppen B (2002) A gender performance indicator for irrigation: concepts, tools and
applications. IWMI Research Report 59. Colombo, IWMI
World Bank (1996) Private sector participation in water supply and sanitation. Conference on
Institutional Options in the Water Supply and Sanitation Sector: Private sector Participation, 8–
11 Jul 1996, Johannesburg, South Africa
World Bank (2007) Making the most of scarcity: accountability for better water management in the
middle East and North Africa. The World Bank, Washington, DC p 235
World Bank (2009a) Water resources: managing a scarce, shared resource. Retrieved from: http://
siteresources.worldbank.org/IDA/Resources/IDA-Water_Resources.pdf
World Bank (2009b) b) Water in the Arab World: management perspectives and innovation.
World Bank, Washington, DC
World Bank (2009c) Improving food security in the Arab-countries. Washington, DC Retrieved
from: http://siteresources.worldbank.org/INTMENA/Resources/FoodSecifinal/pdf
WWAP (United Nations World Water Assessment Programme) (2014) The United Nations world
water development report 2014. Water and Energy, Paris, UNESCO
Chapter 7
Research and Development to Bridge
the Knowledge Gap

Khaled AbuZeid

Abstract The Arab Region, among the driest regions in the world with less than
1.5 % of the world’s renewable water resources, hosts more than 5 % of the world’s
population. It controls about 70 % of the world’s fossil oil reserves. Water security,
energy security, and food security can no longer be dealt with in isolation in the
Arab Region. It may be the most relevant region to a Water-Energy-Food Nexus
approach for development. Research and development on the Water-Energy-Food
(WEF) Nexus, and the extensive analysis of the water, energy, and food security
linkages, is needed to bridge the knowledge gap on their inter-dependencies.
Bridging this knowledge gap is essential to assist in policy making related to water,
food, and energy planning. This chapter presents some of the research and devel-
opment results on key policy elements related to the WEF Nexus in the Arab
Region. It presents some data and information on some key WEF indicators that
could help further research and development on the topic. This chapter considers
some of the linkages between water, energy, and food in the Arab Region,
addressing research needs for better policy decisions in selecting the appropriate
water resource, the appropriate water use, the appropriate water and energy savings
measures, and the appropriate food export-import strategies. The WEF nexus is
addressed through filling the knowledge gap on the externalities directly or indi-
rectly affecting the WEF security, such as non-conventional water resources,
transboundary water resources, blue and green water comparative advantage, as
well as food wastage and losses.

Keywords Water  Energy  Food  WEF  Security  Nexus  Wastewater 



Reuse Efficiency  Optimal

K. AbuZeid (&)
Centre for Environment and Development for the Arab Region and Europe,
2 ElHegaz Street, Heliopolis, Cairo, Egypt
e-mail: kabuzeid@cedare.int

© Springer International Publishing AG 2017 123


K. Amer et al. (eds.), The Water, Energy, and Food Security Nexus
in the Arab Region, Water Security in a New World,
DOI 10.1007/978-3-319-48408-2_7
124 K. AbuZeid

7.1 Introduction

Of all resources, water, food, and energy are most needed to sustain life on earth.
These three resources are tightly interconnected, forming a resource and policy
nexus. Their insecurity is an impediment to social stability and economic growth.
A nexus approach to water, energy, and food security can support the achievement
of sustainable development. The nexus approach would highlight the linkages
between Water, Energy, and Food into the development planning process to ensure
their sustainability. The range of potential risks and uncertainties relating to a single
resource is magnified when the links between different resources are taken into
account. Water, food, and energy insecurities are impediments to social stability
and economic growth (Abu-Zeid 2011).
Water scarcity in the Arab Region is considered one of the major and most
critical challenges facing Arab countries. This challenge is expected to grow with
time due to many pressing driving forces, including population growth, food
demand, unsettled and politicized shared water resources, and climate change,
forcing more countries into more expensive water resources, such as desalination,
to augment their limited fresh water supplies. The large financial, economic,
environmental, and social burdens to be borne cannot be overemphasized. This
challenge is exacerbated by its multiple nexuses with the various development
sectors, such as water and food, water and energy, and their interdependencies with
other resources such as land and wastewater, which carry within them many
cross-cutting security issues, mainly water security, energy security, and food
security. Water, food, and energy resources are tightly interconnected, forming a
policy nexus. Food production is the largest user of water globally. It is responsible
for 80–90 % of consumptive water use from surface water and groundwater. Water,
however, is also used to generate electricity, and about 8 % of global water
withdrawal is used for this purpose. Energy, in turn, is needed to transport and
fertilize crops. Food production and supply chains are responsible for around 30 %
of total global energy demand. Crops can themselves be used to produce biofuels,
but using water for crop production exclusively for biofuels may not be a preferable
choice in a water-scarce and food-short region such as the Arab Region.
A secondary level byproduct of biofuel from agriculture wastes may be a more
acceptable choice.
It is therefore important to address much more explicitly the various linkages of
the water sector with other sectors like energy, and food. Moreover, professionals in
all sectors, encouraged by governments’ strategies for sustainable development,
should think and act beyond the boundaries of their own sector, to achieve effective
and integrated resources planning and management.
According to the current trends in population growth and their associated water,
food, and energy demands in the Arab Region, water security, energy security, and
food security are closely linked, perhaps more than any other region in the world,
and actions in one area have strong impacts on the others. Hence, a nexus approach
that integrates management and governance across these three sectors can improve
7 Research and Development to Bridge the Knowledge Gap 125

security issues. This can also support the transition to a green economy, which
aims, among other goals, at efficient use of resources, and policy coherence.
A proper understanding of the nexus will allow decision makers to develop
appropriate policies, strategies, and investments to explore synergies, and to
identify trade-offs among the development goals related to water, energy, and food
security. Moreover, a nexus perspective increases the understanding of the inter-
dependencies across these three sectors and influences policies in other areas of
concern, such as climate change and environment.
In the water value chain, energy is required in all segments; energy is used in
almost every stage of the water cycle: extracting groundwater, feeding desalination
plants with its raw sea/brackish waters, and producing freshwater, pumping, con-
veying, and distributing freshwater, collecting wastewater, along with treatment and
reuse. In other words, without energy, mainly in the form of electricity, water
availability and delivery will not be possible.
It is estimated that in most of the Arab countries, the water cycle demands at
least 15 % of national electricity consumption and it is continuously on the rise
(Al-Zubari 2013). On the other hand, though less in intensity, water is also needed
for energy production through hydroelectric schemes (hydropower) and through
desalination (Co-generation Power Desalting Plants [CPDP]), for electricity gen-
eration and for cooling purposes, and for energy exploration, production, refining,
and oil recovery processes.

7.2 The Role of Research and Development

The scarcity of fresh water in the Arab Region promoted and intensified Research
and Development (R&D) in the technology of desalination and combined
co-production of electricity and water, especially in the Gulf Cooperation Council
(GCC) countries. Desalination, particularly CPDPs is an energy-intensive process.
Given the large market size and the strategic role of desalination in the Arab
Region, the installation of new capacities will increase the overall energy con-
sumption. As energy production is mainly based on fossil-fuels, a finite source, it is
clear that development of renewable energies to power desalination plants is nee-
ded. Meanwhile, to address concerns about carbon emissions, Arab governments
should link any future expansion in desalination capacity to investments in abun-
dantly available renewable sources of energy.
There is an urgent need for cooperation among the Arab countries to enhance
coordination and investment in R&D in desalination and treatment technologies.
Acquiring and localizing these technologies will help in reducing their cost,
increasing their reliability as a water source, increasing their economic added value,
as well as reducing their environmental impacts. Special attention should be paid to
renewable and environmentally safe energy sources, of which the most important is
solar, which is abundant in the Arab Region.
126 K. AbuZeid

Despite the strong relation, the water-energy-food nexus and their interrelation
has not been fully addressed or considered in the planning and management in
many Arab countries. However, with the increasing water scarcity, many Arab
countries have started to realize the growing importance of this nexus and the need
for R&D to bridge the knowledge gap in finding trans-disciplinary and
cross-sectoral solutions.

7.3 The State of Knowledge on the WEF Nexus


in the Arab Region: Bridging the Gap

In order to bridge the knowledge gap on the water, energy, and food nexus (WEF
Nexus) in the Arab Region, this chapter presents some key information required to
initiate research and further analysis. It provides in the following sections, several
linkages between different aspects of the WEF Nexus looking at water, energy, and
food from different angles. This type of information should support a nexus
approach to policy decisions related to implementing a fit-for-purpose water allo-
cation strategy. A strategy where the appropriate water resources are used for the
appropriate purpose at the appropriate location, and one that optimizes energy use,
water use, and food production.

7.3.1 The Water-for-Food Scarcity Impact on Energy

The per capita share of renewable blue (surface and groundwater) water resources is
below the water scarcity limit of 1000 m3/capita/year, in 17 out of the 22 Arab
countries (Fig. 7.1). To maintain an adequate level of food and water security, the
region has tapped into its reserves of fossil groundwater for food production, and
tapped into its reserves of fossil fuel to pump this groundwater, and desalinate
seawater for freshwater supply. The impossible sustainability of these fossil
reserves calls for the dire need to appropriately address the 3 securities in nexus.
The Arab countries are using about 24 BCM/year of fossil non-renewable
groundwater which put pressure on fossil energy as well to pump this groundwater.
This exploitation of non-renewable groundwater due to the unavailability of ade-
quate renewable freshwater resources is reflected in 10 out of 22 countries in the
region exceeding the 100 % threshold of their freshwater sustainability index,
indicating high potential for using more energy in these countries to make water
available for food production (Fig. 7.2).
With the severe water scarcity situation in the Arab Region, non-conventional
water resources including desalination of sea water, as well as treated wastewater,
are becoming important. Several countries have started to direct their treated
wastewater to food production. Energy requirements for wastewater treatment
7 Research and Development to Bridge the Knowledge Gap 127

Total Renewable Blue Water Resources Per Capita (CM/capita)


3,394

3,104

1,625 Total Renewable Blue Water Resources Per Capita


1,458

1,051

781
908

630

426

387

322

259

249

185

96

90

75

48

42

36

22

8
Mauritania

Iraq

Sudan (North)

Somalia

Syria

Morocco

Lebanon

Egypt

Tunisia

Djibouti

Algeria

Comoros

Jordan

Palestine

Bahrain

Yemen

Saudi Arabia

Libya

Oman

Qatar

UAE

Kuwait
Fig. 7.1 Total renewable blue water resources per capita (CM/capita) in the Arab Region (Source
AbuZeid et al. 2014)

Water Sustainability Index (%)


1,724.10

Water Sustainability Index


673.07

556.03

431.85

348.14

184.67

133.78

127.93

125.08

120.97

98.27

94.06

68.50

60.38

26.54

0.94

0.54

0.46

0.36

0.23

0.21

0.06

Fig. 7.2 Water sustainability index (%) in the Arab Region (Source AbuZeid et al. 2014)

versus desalination for agriculture purposes are going to play an important energy
security role in policy making for water and food security. The possibility of
generating energy through the wastewater treatment process adds another variable
to the Nexus equation. Municipal and industrial wastewater reached
23 billion cubic meters/year (BCM/year) in the region divided among the Arab
countries (Fig. 7.3). A small portion of this amount is treated and a smaller amount
of about 1.6 BCM/year is reused due to the low level of treatment. Although the
128 K. AbuZeid

Produced Wastewater (MCM/Year)

6,500

4,800

4,722
Produced Municipal and Industrial Wastewater (PMW)

1,364

1,200

1,023

730

650

546

520

310

293

270

182

76

74

62
60

42

15

13

4
Fig. 7.3 Produced municipal and industrial wastewater in the Arab Region (Source AbuZeid et al.
2014)

Reused Treated Industrial and Municipal Wastewater (MCM/Year)

550

Reused Treated and Industrial Municipal Wastewater

300 290

166
83 78
43 40 37 21 16 6 2 1 0 0 0 0 0 0 0 0
UAE
Syria

Libya
Egypt

Yemen
Jordan

Tunisia

Somalia
Iraq
Oman
Qatar

Lebanon

Algeria
Bahrain

Comoros

Palestine
Saudi Arabia

Kuwait

Morocco
Djibouti

Sudan (North)
Mauritania

Fig. 7.4 Reused treated industrial and municipal wastewater in the Arab Region (Source AbuZeid
et al. 2014)

countries that are efficiently reusing wastewater are mostly water-scarce and
energy-rich or economically better off, there are a few countries that are not con-
sidered energy-rich but may be considered developed in the state of knowledge on
water scarcity implications and the urgency of reuse (Fig. 7.4). It is obvious though
that the low level of wastewater reuse reflects the high energy costs required to
treat, pump, and transport wastewater for recycling. It also reflects the level of
urgent need for water, as well as the level of knowledge, research and development
in the area of wastewater treatment and reuse.
Agriculture drainage is another element in the WEF Nexus and it is estimated at
about 34 BCM/year in the Arab Region (Fig. 7.5). A small proportion of this
amount is recycled. Expansion in agriculture drainage reuse for agriculture
7 Research and Development to Bridge the Knowledge Gap 129

Produced Agricultural Drainage (MCM/Year)


7,860

7,500

5,191

4,699
Produced Agricultural Drainage (PAD)

2,952

1,717

1,031

967

500

375

282

211

173

140

134

89

66

40

34

-
Sudan (North)

Morocco
Iraq

Egypt

Syria

UAE

Yemen

Saudi Arabia

Mauritania

Oman

Lebanon

Jordan

Kuwait

Libya

Qatar

Bahrain

Palestine

Comoros

Djibouti

Algeria
Somalia

Tunisia
Fig. 7.5 Produced agricultural drainage by country in the Arab Region (Source AbuZeid et al.
2014)

purposes is constrained in many cases due to pollution levels from untreated


wastewater that reaches agriculture drains. Egypt is an example of a country that is
facing a water, energy, and food shortage. It reuses the largest amount of agriculture
drainage, for food production in the Arab Region, reaching about 5.5 BCM/year.
Historically depending on gravity flow surface irrigation systems in its Nile Valley
traditional agriculture lands, and generating substantial amounts of agriculture
drainage, Egypt is often faced with a WEF Nexus dilemma when it comes to reuse
of this agriculture drainage. On the one hand, it needs the energy to pump part of
this drainage water from the agriculture drains into the irrigation canals, to com-
pensate for shortage in water needed for food production. It also needs the energy to
pump the seepage groundwater in the Nile Valley to compensate for water shortage
in the downstream reaches of its irrigation canals. Farmers downstream also use
individual pumps and bear high energy costs to pump agriculture drainage water
into their farms to compensate for water shortages. On the other hand, switching
from gravity surface irrigation systems in traditional agriculture lands to pressurized
modern irrigation systems for more efficient on-farm water use, will not only
require a substantial amount of energy, but it will also reduce groundwater seepage
and agriculture drainage water which are important sources of water for food
production in other agriculture locations.
The water scarcity situation in the Arab Region has forced countries to depend
on an energy-intensive resource such as seawater desalination reaching a total
production of over 4 BCM/year in 2012 (Fig. 7.6). The growth rate in energy
consumption at 8 % in the region, and the low energy use efficiency reaching less
than 50 % in some cases within the region, which consumes more than 50 % of its
total energy for desalination in some countries, requires an optimal linkage between
130 K. AbuZeid

Produced Desalinated Water (MCM/Year)


1,033

950

Produced Desalinated Water


540

420

200

197

180

102

47

30

25

25

10

0
Saudi Arabia

UAE

Kuwait

Egypt

Oman

Qatar

Bahrain

Lebanon

Tunisia

Yemen

Mauritania

Jordan

Iraq

Libya

Morocco

Palestine
Algeria

Sudan (North)

Djibouti

Comoros

Somalia

Syria
Fig. 7.6 Produced desalinated water across the Arab Region (Source AbuZeid et al. 2014)

Withdrawals From Non -Conventional Resources (MCM/Year


6,000

Withdrawals From Non-Conventional Resources


1,240
1,199
550
540
498
234
223
118
93
51
49
47
31
26
7
7
5
0
0
0
0
Egypt

UAE

Saudi Arabia

Syria

Algeria

Kuwait

Oman

Qatar

Bahrain

Jordan

Tunisia

Lebanon

Libya

Yemen

Mauritania

Iraq

Morocco

Palestine

Sudan (North)

Djibouti

Comoros

Somalia

Fig. 7.7 Withdrawals from non-conventional resources within the Arab Region (Source AbuZeid
et al. 2014)

energy consumption and water production. Use of non-conventional water


resources (Fig. 7.7), such as treated wastewater, desalination, and agriculture
drainage water (currently reaching 11 BCM/year in the Arab Region), and the use
of deep non-renewable groundwater requires policy analysis that should consider
the “energy” costs involved in treatment, desalting, and pumping, which requires a
Nexus approach in allocating the appropriate “water” resource to “food” produc-
tion. There is still a knowledge gap in the optimization of water allocation to select
the most appropriate water resource for food production, by considering water and
energy sustainability, conserving freshwater resources, and utilizing the least
amount of energy.
7 Research and Development to Bridge the Knowledge Gap 131

7.3.2 Energy for Water

The World Water Development Report (2014) presents average energy require-
ments for developing different water resources (Abdel-Dayem 2014). These energy
requirements vary greatly: 0.37 kWh/m3 for water from lakes and rivers;
0.48 kWh/m3 for groundwater; 0.62–0.87 kWh/m3 for wastewater treatment; 1.0–
2.5 kWh/m3 for wastewater reuse; 1.5–3.5 kWh/m3 for drinking water sophisti-
cated treatment; and 2.58–8.85 kWh/m3 for salt water desalination.
A quick comparison of these energy requirements of developing different water
resources for the purpose of food production, demonstrates that pumping water
from open water bodies (being fresh surface water or recycled agriculture drainage
water) at 0.37 kWh/m3 would be more economical than pumping groundwater at
0.48 kWh/m3. When renewable freshwater is exploited, reusing treated wastewater
for food production at 1.5–3.5 kWh/m3 would be more economical than using
desalinated water at 2.58–8.85 kWh/m3.
The policy decisions that are made today will affect the policy decisions that
need to be taken in the future, and this is an important factor when it comes to the
exploitation of fossil non-renewable groundwater - a water resource that is often of
ultimate quality, and in most cases does not require sophisticated treatment for
drinking purposes. What is the best use for that resource? Should it be allocated to
agriculture (a water intensive user)? Or is it better for domestic and drinking water?
The nexus approach provides some insights to the policy decisions that need to be
made. Exploiting non-renewable groundwater in agriculture at 0.48 kWh/m3 will
quickly result in the need to make drinking water available at energy requirements
of 1.5–3.5 kWh/m3 or maybe even at 2.58–8.85 kWh/m3 if desalination is required.
Whereas, strategically using non-renewable groundwater for domestic and drinking
purposes at 0.48 kWh/m3 will sustain it for a longer period of time and will gen-
erate wastewater that could be treated and reused for agriculture at energy needs of
1.0–2.5 kWh/m3. This is an example of a water-for-food policy decision that, when
taken within a Nexus approach, could contribute to energy security by providing a
long term energy-efficient alternative.

7.3.3 Water for Energy

Mielke et al. (2010) indicates that water use for secondary oil extraction could reach
62.1 gal/MMBtu, and 10.3 gal/MMBtu for oil refining, as well as 315 gal/MWh for
steam turbine, and 4500 gal/MWh for hydropower (as evaporation losses).
It is obvious that moving into renewable solar and wind energy would be less
water-consumptive than fossil fuel extraction and refining, and even hydropower.
However, it is important to harness the little remaining hydropower potential of the
available water resources in the Arab Region (Table 7.1). Making use of in-stream
turbines for energy production is also important, as well as making use of all water
132

Table 7.1 Hydropower potential and utilization in the Arab Region


Country Gross Technically Economically Installed Hydro % production Hydro Planned
theoretical feasible feasible hydro generation in by hydro in capacity hydro
hydropower hydropower hydropower capacity 2008 or 2008 or most under capacity
potential potential potential (MW) average/most recent construction (MW)
(GWh/year) (GWh/year) (GWh/year) recent (average) (MW)
(GWh/year)
Algeria 12,000 4000 278 560 2.7 0 n/a
Comoros 1 2
Egypt >50,000 *50,000 2842 15,510 12.6 0 48
Iraq 90,000 67,000 2273 n/a 13 *30 800–
5000
Jordan 500 MW 12 62 0.45 0 n/a
Lebanon 280 750 7 76 n/a
Mauritania 30 120 n/a 0 n/a
Morocco 5203 4000 *1265 1318 6.6 *40 84–384
Somalia 600
Sudan n/a 19,000 575 4333 55 >1200 2000–
3600
Syria 1505 *8000 15 n/a n/a
Tunisia 1000 250 160 *70 160 3 n/a >20
Source AbuZeid et al. (2014)
K. AbuZeid
7 Research and Development to Bridge the Knowledge Gap 133

outfalls, be it wastewater outfalls, cooling water outfalls, or irrigation and drainage


gravity channels using mini-hydropower installations. There are still some knowl-
edge gaps in studying the potential of wastewater outfalls and mini-hydropower
installation that can maximize the optimal linkages between water and energy.

7.3.4 The Transboundary Dimension


of the Water-Energy-Food Nexus

The transboundary nature of the Arab Region’s water resources also puts another
dimension to the WEF Nexus, whereby 65 % of the region’s renewable surface
water resources originate outside the Arab countries. Adding the transboundary
dimension to the WEF Nexus complicates the issue even more, where water
security is closely linked to issues of energy and food security upstream and
downstream with different levels of effect depending on the number of riparian
countries affected. An upstream infrastructure that may ensure energy security for
an upstream country may, not only affect energy security for a downstream country,
but could also put its food security at risk. Therefore, the hydropower potential at
the transboundary river basin level needs to be assessed in an integrated manner and
should be planned jointly so that upstream hydropower projects do not affect
downstream hydropower installations or other food production agriculture activi-
ties. Likewise, upstream diversions for agriculture purposes upstream should be
planned jointly with downstream riparian countries to avoid negative impacts on
hydropower energy installations, or food production and agriculture activities
downstream.
Tables 7.2, 7.3 and 7.4 provide the state of hydropower potential and installed
capacity in the riparian countries of three of the major transboundary river basins
intersecting with the Arab Region, namely, the Senegal River Basin, the Euphrates
River Basin, and the Nile River Basin, respectively.

7.3.5 The Water-Energy Savings Nexus for Food

The low “on-farm” irrigation water use efficiency, reaching less than 60 % in some
cases, coupled with the need for increased energy use to switch to pressurized drip
or sprinkle irrigation systems saving on irrigation water use, requires an optimal
linkage between energy consumption and water consumption for food production.
Although low on-farm irrigation water use efficiency may result in increased
agriculture drainage, which some consider as wasted water, yet in some countries in
the region, this drainage water is reused several times for food production.
Improvements in any one sector may involve tradeoffs in others, and while modern
irrigation systems may reduce on-farm irrigation water used for food production, it
134

Table 7.2 Senegal River Basin hydropower potential


Country Gross Technically Economically Installed Hydro % production Hydro Planned
theoretical feasible feasible hydrocapacity generation in by hydro in capacity hydro
hydropower hydropower hydropower (MW) 2008 or 2008 or most under capacity
potential potential potential average/most recent construction (MW)
(GWh/year) (GWh/year) (GWh/year) recent (average) (MW)
(GWh/year)
Guinea 26,000 19,300 18,200 123 519 37.8 n/a >240
Mali *5000 155 >500 60 140 >100
Mauritania 30 120 n/a 0 n/a
Senegal 4250 2050 66 293 16 0 123
Source CEDARE-AWC (2012)
K. AbuZeid
Table 7.3 Euphrates River Basin hydropower potential
Country Gross Technically Economically Installed Hydro % production Hydro Planned
theoretical feasible feasible hydrocapacity generation in by hydro in capacity hydro
hydropower hydropower hydropower (MW) 2008 or 2008 or most under capacity
potential potential potential average/most recent construction (MW)
(GWh/year) (GWh/year) (GWh/year) recent (average) (MW)
(GWh/year)
Iraq 90,000 67,000 2273 n/a 13 *30 800–
5000
Syria 1505 *8000 15 n/a n/a
7 Research and Development to Bridge the Knowledge Gap

Turkey 433,000 216,000 140,000 13,700 48,000 25.4 8600 22,700


Source AbuZeid et al. (2014)
135
136

Table 7.4 Nile River Basin hydropower potential


Country Gross Technically Economically Installed Hydro % production Hydro Planned
theoretical feasible feasible hydrocapacity generation in by hydro in capacity hydro
hydropower hydropower hydropower (MW) 2008 or 2008 or most under capacity
potential potential potential average/most recent construction (MW)
(GWh/year) (GWh/year) (GWh/year) recent (average) (MW)
(GWh/year)
Burundi 6000 1500 600 50 208 100 1 177
D.R. 1,397,000 100,000 MW 145,000 2410 7303 100 >162 3690–
Congo 43,000
Egypt >50,000 *50,000 2842 15,510 12.6 0 48
Eritrea 0 0 0 0 0
Ethiopia *650,000 >260,000 162,000 669 2700 >95 1277 4170–
10,000
Kenya >24,300 1422 MW 747 3000 63 41 >160
Rwanda 400 MW 55 130 59 0 120–209
Sudan n/a 19,000 575 4333 55 >1200 2000–
(North 3600
&
South)
Tanzania 39,450 20,000 561 2098 61 0 1868
Uganda n/a >12500 *395.5 1391 67.9 337 *1000
Source AbuZeid et al. (2014)
K. AbuZeid
7 Research and Development to Bridge the Knowledge Gap 137

Overall Water Use Efficiency (%)


Overall Water Use Efficiency
95
86 86 84 84 81 80 76 76 75 73 73 72 72 70 67 63 63
58
50

29

Fig. 7.8 Overall water use efficiency (%) (Source AbuZeid et al. 2014)

could consume more energy and reduce return flows to groundwater and drainage
water that may be used for food production in other locations. It is important to
consider an objective of overall water use efficiency coupled with an energy use
efficiency objective rather than looking at on-farm irrigation efficiency in isolation.
Figure 7.8 shows the overall water use efficiency for Arab countries considering the
water reuse factor in these countries. It is obvious that a higher energy use is needed
to achieve higher water savings in the agriculture sector.

7.3.6 Food Security, Virtual Water Trade,


and the WEF Nexus

Each country in the Arab Region is far from having enough water to grow its basic
food needs. Thus, the idea of food self-sufficiency is now switched to the concept of
food security. Recent studies by CEDARE and the Arab Water Council shows that
although the Arab Region uses about 212 BCM/year of virtual water in food
production within the region, the Arab countries are importing 274 BCM/year of
virtual water in food products while exporting about 55 BCM/year of virtual water
in food products (AbuZeid et al. 2014). Figures 7.9 and 7.10 show the amount of
virtual water import and export respectively for the Arab countries for 2012.
It would be very expensive and energy intensive, and for some countries prac-
tically impossible, if the Arab Region would produce all its food needs locally to
achieve food self-sufficiency. Theoretically speaking, locally producing the
amounts of imported food products of year 2012, the Arab Region would require
about 1700 GWh/year of energy to desalinate the required amount of sea water, and
additional energy to convey the water for agriculture, and even more for pressurized
irrigation and fertilization.
The knowledge gap on the required amount of water and energy needed to
achieve food self-sufficiency, and the knowledge gap on sustainability issues of
138 K. AbuZeid

Virtual Water Import (BCM)

Import
45.00
40.00
35.00
30.00
25.00
20.00
15.00
10.00
5.00
0.00

Fig. 7.9 Virtual water import in the Arab Region (BCM) (Source AbuZeid et al. 2014)

Virtual Water Export (BCM)

20.00 Virtual Water Exports related to Trade in the Agricultural Sector

15.00

10.00

5.00

0.00

Fig. 7.10 Virtual water export in the Arab Region (BCM) (Source AbuZeid et al. 2014)

fossil groundwater and fossil fuel, are influencing strategies and policies that may
not be appropriate for a water scarce region, such as the Arab Region.

7.3.7 The Blue and Green Water Effect on the WEF Nexus

Using green water through rain fed agriculture for food production uses less energy
than using blue water through irrigated agriculture. Due to the scarcity of rainfall in
the Arab Region, agriculture is mostly irrigated in the Arab Region depending on
groundwater and surface water abstractions mostly from trans-boundary rivers
originating outside the region. Although green water (from direct rainfall for rain
fed agriculture) use for food production in the Arab Region reached 51 BCM/year,
7 Research and Development to Bridge the Knowledge Gap 139

blue water (abstracted from rivers and groundwater for irrigated agriculture) use for
food production reached 231 BCM/year, putting more pressure on energy. It is
recommended to make use of green water for food production whenever it exists to
relieve the pressure on energy resources.

7.3.8 Food Wastage: Adding to the Water-Energy-Food


Nexus

It is also worth mentioning, that food wastage is closely linked to virtual water
wastage, and could also be considered as an indirect contributor to energy wastage.
Figure 7.11 shows food losses in the Arab Region and Central Asia as a percentage
of food production indicating 20 % to more than 50 % wastage among several
categories of food including fruits and vegetables, root and tuber crops, cereals, fish
and seafood, oil seeds and pulses, meat, and milk and dairy products.
Figure 7.12 shows food wastage in tonnage during different stages including
food production, post harvesting and storage, processing and packaging, distribu-
tion and consumption. Translating this food wastage of about 80 million tons into
virtual water, it could reach an amount of about 50 BCM/year. Noting that the
region indicated here is larger than the Arab Region, this food wastage could also
translate into a substantial amount of energy wastage affecting not only food and
water security but also energy security.

Food Losses as a Percent of Food Production in North Africa, West and Central Asia

60%

50%

40%

30%

20%

10%

0%
Fruits & Veg. Root & tuber Cereals Fish & seafood Oilseeds & Meat Milk & Diary
crops pulses products

Fig. 7.11 Food losses as a percent of food production in North Africa, West and Central Asia
(Source AbuZeid 2013)
140 K. AbuZeid

Food Losses and Waste (1,000,000 tonnes) in North Africa, Central and Western Asia

40

30

20

10

0
Cereals Root & Oilseeds & Fruits & Meat Fish & Milk &
tuber crops pulses Veg. seafood Diary
products

Agricultural Production Post-harvest Handling & Storage Processing & Packaging

Distribution Consumption

Fig. 7.12 Losses and waste (1,000,000 tons/year) in North Africa, Central and Western Asia
(modified from FAO 2011)

7.4 Conclusion

Achieving water security, energy security, and food security in the Arab Region
requires advancing the state of knowledge, enhancing research and development in
the technologies used for efficient water use, efficient energy use, renewable energy
development, and increased crop productivity. Applied research on decision mak-
ing and multi criteria analysis is important in the field of water, energy, and food
planning. Policy decisions to be made in any of the water security, energy security,
or food security fields should not be done in isolation but rather within a nexus
approach. Research on the WEF Nexus should consider the fit-for-purpose water
use, the long term implications of using fossil reserves, the transboundary dimen-
sion of the WEF Nexus, the green water comparative advantage for food produc-
tion, the optimization of virtual water trade, the measures needed to reduce food
wastage, water wastage, and energy wastage, and the trade-offs between increasing
water efficiency, increasing water recycling, and the corresponding energy needs.

References

Abdel-Dayem S (2014) The state of water-energy-food nexus in the MENA/Arab region.


Presentation at the 3rd Arab Water Forum, Cairo, Egypt
AbuZeid M (2011) Water energy and food nexus. In: Water, nuclear, and renewable energy:
challenges versus opportunities. Alexandria TWAS-ARO 7th meeting, Egypt
AbuZeid K (2013) Water-land-waste nexus in the Arab Region. Workshop on advancing a nexus
approach to the sustainable management of water, soil, and waste. United Nations University,
Dresden, Germany
7 Research and Development to Bridge the Knowledge Gap 141

AbuZeid K, Elrawady M, CEDARE and Arab Water Council (2014) 2nd Arab State of the water
report-2012, 2nd Arab Water Forum, Water Resources Management Program-CEDARE and
Arab Water Council, ISSN 2357-0318, ISBN 978 977 90 3806
Al-Zubari W (2013) The water-energy nexus in the GCC countries: Evolution and related policies.
Sixth “Zayed Seminar” on green economy: success stories from the GCC, 8–9 May 2013, The
Arabian Gulf University, Bahrain
FAO (2011) Global food losses and food waste: Extent, causes and prevention. Food and
Agriculture Organization, Rome
Mielke E, Diaz Anadon L, Narayanamurti V (2010) Water consumption of energy resource
extraction, processing, and conversion. Energy Technology Innovation Policy Research Group,
Harvard Kennedy School, Belfer Center for Science and International Affairs, Oct 2010,
Cambridge, USA
Chapter 8
Water-Energy-Food Security Nexus
in the Arab Region: Thoughts and Policy
Options

Paul Sullivan

Abstract The Arab Region is under considerable water stress. It is also heavily
reliant on imported food. The region has countries with considerable energy exports
mixed in with those who are of differing degrees of reliance on imported energy.
The Arab Region is one that has significant problems with energy, water, and food
waste. It is also a place where energy, water, and food subsidies have added much
to the resource and economic stresses of many countries. Energy, water, and food
show not only stress today, but potentially much greater stresses in the future.
Energy, water, and food can be considered, and policies developed about their use
and development, separately. However, because of the connections between these
three via the water-energy-food nexus and via the water footprints of food and
energy systems and the energy footprints of water and food systems, as well as
other indirect and direct connections among them, policies need to be integrated
and coordinated across these three. There is much discussion of integrated water
policies in the region, but the region needs to go further and consider integrated
water-energy-energy-food nexus policies with nexus thinking. To develop these
changes educational, outreach, training, investment, and other programs need to be
developed in the region to help lead the people and their leaders to a better future.
Direct cash payments and other softening tools need to be implemented to mitigate
the effects of the phasing out of the massive subsidies and other distortions within
the water-energy-food nexus that would otherwise lead the region to resource,
economic, and political disasters. Reason and moderation will need to be defining
characteristics of these policy changes. Otherwise, the shocks of change could
backfire on an already unstable region. If the water-energy-food nexus policies are
not developed properly, in the long run, there could be extreme resource stresses in

All opinions are Professor Sullivan’s alone.

P. Sullivan (&)
Georgetown University, Washington, USA
e-mail: Drsullivenergy@gmail.com
P. Sullivan
National Defense University, Washington, USA

© Springer International Publishing AG 2017 143


K. Amer et al. (eds.), The Water, Energy, and Food Security Nexus
in the Arab Region, Water Security in a New World,
DOI 10.1007/978-3-319-48408-2_8
144 P. Sullivan

the region beyond what have been seen to date. These resource stresses could have
global implications in increased energy insecurity, increased terrorism, and
increased migration from the region to more stable, more economically developed
and more resource rich areas.


Keywords Water-energy-food nexus Arab world  Subsidies  Sustainability 
 
Energy security Water security Food security

8.1 Introduction to the Problems

A simple example of the water-energy-food nexus can be found in the use of pumps
to move water from an underground aquifer to irrigate crops. Energy is used to run
the pumps. The water is used to grow the crops. Hence, we have a quite simple
picture of the water-energy-food nexus. Expanding from this simple example to the
water-energy-food nexus of a region of an Arab country, such as the delta area of
northern Egypt, to the country level and then water regions that cut across countries,
such as the Tigris-Euphrates or the Nile basins, or to the entire Arab Region,
considerably increases the networked and systems complexity of the
water-energy-food nexus.
Most of the food for the Arab Region is imported. Some Arab countries are also
food exporters. A large percentage of the energy produced in the region is exported.
Yet some Arab countries are energy importers. Water is imported into the Arab
Region from various places connected to it. Small percentages of overall water
needs for the region are directly imported by ship. Water is exported from Syria to
Iraq via the Euphrates and some water moves across from Saudi Arabia to other
Gulf Cooperation Council (GCC) states in slowly moving underground aquifers.
There are many complex international trade and resource diplomacy aspects of the
water-energy-food nexus in the Arab Region.
A water-energy-food nexus will work more efficiently with non-distorted market
systems and proper public policies attached to them. In the Arab world, energy,
water, and food are often subsidized and sometimes heavily subsidized causing
over demand of the resources associated with them. This means that aquifers and
river basins are drained faster than they would otherwise be. This also means that
energy is overused, which brings traffic congestion, health issues, and sometimes
massive budget stress. When food is subsidized, then it also has demand beyond
what would normally be. More food is lost and wasted than would be without the
subsidies. The less value put on a resource the less efficiently it will be used. When
these subsidies within each system of water, energy, and food are connected across
the nexus then the distortions, losses, and waste can be multiplied. These distortions
can cause stresses and strains on an economy and a society. They can also stress
resources beyond energy, water, and food, such as land, minerals, and human
health. These compounded stresses have led to significant problems in the region. If
nothing is done, then these stresses could get much worse (Abruzezze 2014; Abu
8 Water-Energy-Food Security Nexus in the Arab Region … 145

Zeid 2013; Allan 2014; Al-Zubari 2013; Andrews-Speed et al. 2012, 2015;
Axworthy and Adeel 2014; FAO 2006).

8.2 Energy: Subsidies, Water Footprints of Energy,


Energy Waste and Losses

To say that the water-energy-food nexus in the Arab Region is unsustainable is


understating the problem. It is a region under enormous water stress, yet water is
subsidized. It is a region that imports a very large proportion of its food, for some in
the GCC this is about 90 %, yet food continues to be subsidized. Energy supplies
are more problematic than one might think even in the GCC. The only country in
the GCC that has enough gas to supply its needs is Qatar (Booz and Company
2010). Tunisia, Egypt, Morocco, Lebanon, and Jordan are net energy importers.
Syria is a net energy exporter, but via smuggling refined oil products mostly. Iraq is
a major oil exporter and also has problems with smuggling of oil and oil products.
Algerian refined oil products are smuggled into Morocco. Egyptian refined oil
products used to be smuggled into Gaza.
Oil and oil refined products are heavily subsidized in the GCC and this is
reducing the ability of these countries to export their oil for hard currency.
Subsidized electricity made in subsidized electricity generating plants is used to
produce subsidized desalinated water to grow subsidized crops in the GCC. Energy
subsidies are also creating entire industries, such as aluminum, that are based on
cheap, subsidized energy. Many of the manufacturing companies in the Arab
Region would be hard pressed to be competitive globally if their energy and water
subsidies were taken off (Charles et al. 2014; Espinoza 2012). About 50 % of the
world’s energy subsidies can be found in the Arab Region and these have caused
great distortions (Droogers et al. 2012; EIA 2014; IMF 2013, 2014; Sdralevich et al.
2014). Some of the cheapest gasoline in the world can be found in some countries
in the Middle East and North Africa (Global Petrol Prices n.d.).
The massive subsidy systems of the Arab Region have driven many countries in
it to budgetary stress that would otherwise not be needed. The pre-tax subsidies in
the Middle East and North Africa are on average about 9 % of GDP and 22 % of
government revenues. Post-tax subsidies are on average about 14 % of GDP and
34 % of government revenues (Devarajan 2014; IMF 2013). Interestingly, the
countries with the most subsidies per capita, such as Egypt, Syria, and Libya, were
the countries that had violent revolutions in recent years. Subsidies were supposed
to keep the peace in the streets. They did not.
Subsidies are often originally meant to help the poor and the near poor, but often
they end up helping the richer people in the Arab countries far more than the poor
(IMF 2014, p. 2). Subsidies encourage waste. The make energy, water, and food
security unsustainable. Subsidies also encourage smuggling, which finances illegal
activities and in some cases terrorism and rebellion (Sullivan 2014). They
146 P. Sullivan

discourage investments from the private sector. They drain budgets. They strangle
private sector development and innovation (Abdelaziz 2013; AFP 2013;
Al-Khatteeb 2014; Alexander and Krause-Jackson 2013; Al-Makhfi 2013;
Blanchard 2009; Del Grando et al. 2010; Erdil 2014; Fattouh and El-Katiri 2012;
Faucon and Albayrak 2014; Hashim 2006; Herbert 2014; IEA 2011; IMF 2014;
Lund 2014; Neuhauser 2014; ‘Turkey seizes smuggled…’ 2014; ‘Fuel crisis in
northern Iraq…’ 2014; UNDP 2012; “US official: ‘Islamic State’ a petrostate”
2014). They also increase pollution and greenhouse gases. In many countries in the
region, more is spent on maintaining energy subsidies than on education (IMF
2014, p. 2).
Fuel subsidies are the highest for the region in terms of percentage of GDP in
Saudi Arabia, which has some of the cheapest fuel prices in the world. However,
Egypt, Syria, Iraq, and Libya also have had large fuel subsidies. And these have not
been the most peaceful of states.
Such fuel subsidies have also added to traffic congestion and traffic injuries and
deaths in the region (World Bank 2014, p. 14). They have also added to motor
vehicle air pollution deaths in the region (World Bank 2014, p. 15). Energy sub-
sidies, getting back to our original simple example of the diesel pump for irrigation,
have increased water use for irrigation in the region (World Bank 2014, pp. 17–18).
A lot of water is used in energy extraction, production, transport, and other
aspects of the energy industry (Sullivan 2013). When water is subsidized then there
will tend to be an overuse of that water in the energy industry. This would not only
increase the actual resource costs of producing energy, but would also cause a faster
draining of water reserves and a higher percentage of use of renewable freshwater.
Water is used in extraction techniques, such as fracking, but also in enhanced oil
recovery with water injection into older hydrocarbon fields or such fields with
pressure problems to get the gas and oil to flow faster. Water is also used to refine
gasoline and diesel and to produce biofuels. Water is used to produce electricity in
concentrate solar power plants. Little water is used in solar photovoltaic facilities
compared to all thermal methods of producing electricity. Wind farms use a lot less
water per unit of energy produced than thermal or even concentrated solar power
plants (Water Footprint Network 2014). Conventional gasoline production uses a
lot less water per btu than the creation of biofuels or other refined liquid fuels
(Meilke et al. 2010). A huge amount of water could be saved by converting the
energy systems of the region to less water-using technologies. A lot of water could
be saved if the cooling systems for electricity plants, which are sometimes also
desalination plants in the region were changed to dry, hybrid, or other less
water-using cooling methods than those that mostly exist in the region today (IEA
2012; Andrews-Speed et al. 2015; Bauer 2014; Clean Energy Business Council
2014; DOE 2014).
What could be done with the massive fuel subsidies in the region if they were
redirected into renewable energy diffusion and especially for renewables for
desalination? However, as these subsidies stand now they are tending to thwart the
development of renewable energy systems in the region (Bridle et al. 2014).
8 Water-Energy-Food Security Nexus in the Arab Region … 147

For the typical thermal power plant, about 65 % of the energy produced is lost in
the cooling system. Then along the way to its final use, considerable energy is lost
in sending it along transmission lines, stepping it up and down, and in the final end
use. This is especially so for incandescent light bulbs, pumps, and pipes to use
water and in most air-conditioning systems. When electricity prices are low the
people at the end use stage are going to be far less efficient in the behavioral use of
the energy and also in their choice of what machines and other systems they will
use to be run on that electricity.
When the fuels going into the electricity generation system are subsidized, then
the incentives to use those fuels efficiently declines. Sankey diagrams for energy
flows, which can be found for most countries in the world on the International
Energy Agency web site, can give some insights into energy use in the region and
its countries (IEA 2013). A large amount of the energy that goes into electricity
production is lost and wasted energy. Again, this is not uncommon globally, but the
incentives to correct this via combined cycle systems or other systems that can
capture the wasted energy in its heat form to do things like produce more electricity
or heating and cooling are less when energy prices are so low.
Low gasoline and diesel prices also reduce the incentives to use more efficient
auto and truck engines, lighter weight trucks and cars, and the other changes that
will be needed to move toward more energy-efficient transportation systems. Large,
gas-guzzling cars are commonplace in the region.
The biggest sources of CO2 in the world are from transportation and electricity
production (EPA 2014). Subsidizing fuels for transport and electricity production and
subsidizing the consumption of electricity makes energy waste much worse than it
otherwise would be. The Arab Region’s countries are not amongst the largest
greenhouse gas producers in the world, excepting Saudi Arabia. However, on a per
capita basis, some Arab countries, such as Qatar, UAE, and Kuwait are right at the top
(Begum 2014; Joint Research Center 2013, 2014; Meltzer et al. 2014; Union of
Concerned Scientists 2014; United Nations Millennium Goals Indicators 2013).
Qatar’s per capita CO2 production is about 10 times the average for the countries of the
world. So there may be some future research potential to look at the carbon footprint
for the region, how it relates to the subsidy programs, and how the water-energy-food
nexus is connected in the region. In the GCC region in particular, there is a significant
carbon footprint to water (Griffiths-Sattenspiel and Wilson 2009; Waterfootprint.org
2014) given how much natural gas and oil go into desalination.

8.3 Food: Food Security, Food Losses and Waste, Water


and Energy Footprints of Food and Food Subsidies

Energy is used to produce, process, ship, freeze, cool, clean, cook, and more to the
food that is used in the home. So when food is wasted, energy is also wasted. If that
food was imported from far away, let us say an apple from New Zealand, then one
148 P. Sullivan

would really need to add up the water and energy used to produce, process, pack,
ship, sell, and more that apple in the country in the Arab world. “Every time food is
lost or wasted, all the natural resources used to grow, process, package, transport,
and market it also go wasted. For example, when an apple is lost or wasted, 70 L of
water used for its production go wasted as well” (FAO 2014b, p. 1).
Over 50 % of the region’s food is imported. The Arab Region is the largest
regional importer of wheat in the world (Economic Research Service USDA 2013,
p. 45). Its imports of wheat have grown from under 10 million tons up to the
mid-1970s to over 70 million tons more recently (UNDP 2013, p. 5). Many of the
GCC countries import between 70 and 90 % of their food, with Qatar being the
most food import dependent (Economist Intelligence Unit 2010, 2014; ‘Food
security in the Gulf…’ 2014). The region imports about 50 % of its cereals, 70 %
of its vegetables, and about 60 % of its sugar (FAO 2014b, p. 1, c).
The region takes in over 20 % of the entire world’s imports of wheat, coarse
grains, rice, butter, cheese, skim milk powder, whole milk powder, poultry, sheep,
and sugar. These shares of world imports are expected to continue to increase. Its
shares in world consumption of many products are quite considerably larger than its
shares in the production of these foods. The large import/production ratios for
foodstuffs in the region prove this (FAO 2014d, p. 6).
The region has a low percentage of lands that are arable. Per capita arable land is
miniscule in some countries in the region (FAO 2014d, p. 10). It has one of the
lowest per capita renewable water resources in the world (FAO 2014d, p. 13, 21,
112–113). Its cereal yields are about 56 % of the world average per hectare (FAO
2014b, p. 1, e).
The region is urbanizing quickly. Its population rates have been in decline in
most countries, but still remain high compared to world averages (FAO 2014b, p. 1,
e). The poorest countries in the region have some of the highest population growths
in the world. Millions in Yemen, Syria, Iraq, Gaza, and more are starving or
undernourished.
Food inflation has been significant in many countries in the region in recent
years, and has occurred at times in the past when there have been food price shocks
globally. This makes sense, of course, given how much reliance the region has on
food imports, and most particularly the GCC and the poorer countries, or those who
are in the midst of considerable political instability, such as Iraq, Syria, and Yemen
(Economist Intelligence Unit 2014; Fan 2014; Hajjar 2014; Hajjar and Sweeney
2014; World Food Program 2014a, b, c; Naja 2014).
Food security situations vary greatly across the countries of the region. The most
severe food security issues are found in Syria, Yemen, Sudan, Djibouti, Comoros,
Somalia, and the West Bank and Gaza. Algeria, Egypt, Jordan, Lebanon, Morocco,
and Tunisia also have significant food security issues. We could easily add the area
now under control by ISIS, but the data coming out of there are not clear enough to
solidify the problems. Iraq has from very severe to moderately severe food security
issues depending on the area of Iraq, the rains, the flows of water from Turkey, and
political security. Areas closest to the Tigris and Euphrates are the most likely to
have controllable food security stress. Those in the desert areas, especially those
8 Water-Energy-Food Security Nexus in the Arab Region … 149

with limited groundwater reserves will be the least food secure, especially when
they are cut off from food trade with more food secure areas within Iraq or from
foods imported into the country (Gereffi and Ahmed 2014; Bailey and Willoughby
2013; Daher 2014; Economist Intelligence Unit 2014; FAO 2013a, b, 2014b, d, e;
Laamrani 2014; Naja 2014; Spiess 2011). Some of the areas controlled by ISIS may
be the least food secure areas in Iraq and Syria, which may lead one to seriously
wonder about the long term sustainability of ISIS when faced with water and food
shortages, which seem inevitable for some time in the future. Such food insecurity
for ISIS will get even worse if its income sources from energy, people, guns, and
drug smuggling are stopped or curtailed.
The GCC and the richer countries of the region, have fairly good food security
when considering the wealth they have to buy the food that is needed for their
people. However, they need to rely on the security of major sea lane choke points,
such as the Bab el Mandab and the Straits of Hormuz, for the import of the food.
About 40 % of the GCC’s grain comes via the Bab el Mandab. About 80 % of its
rice comes via the Straits of Hormuz. The GCC is also greatly dependent on the
Suez Canal for about 81 % of its food imports from North America and Europe
(Bailey and Willoughby 2013; Duncan 2013).
Close to 80 million people in the region are undernourished. Over 30 % of the
children are stunted. Yet 25 % of the people are obese (FAO 2014d, p. 1). This is a
region of poverty and hunger, as well as a region of extreme wealth and obesity. It
is also a region where the poor can also be obese. It is a region of food security
contradictions and distortions.
Importing food is also importing water. The GCC countries can afford, so far, to
import that food and keep their people fed. In some GCC countries obesity is on the
rise, as is diabetes caused by obesity and other factors. So, in a way, subsidizing
imported food is a cause for expensive health problems. Subsidizing the production
and consumption of domestically produced food adds in other problems. This not
only causes over-consumption of the domestically produced food, as it does for the
imported food, but also causes overuse of the energy and water used to produce that
domestic food. A lot of the water and energy used goes to producing feed for
livestock as well.
A big aspect of the water-energy-food nexus in the GCC, and the rest of the Arab
world is the water footprint of the crops and livestock that are being developed in
the region. The most water-using sources of food for people are livestock, such as
cattle and sheep. Globally over one-quarter of all of the water used by humanity
goes into livestock production (Hoekstra 2012). It would be interesting to see a
water footprint accounting of the development of livestock in the Arab world.
The least water-using sources of food are vegetables. However, there are many
gradations on these water uses. The water used to grow wheat in a hot desert
climate is quite a bit more than the water normally used to grow it in climate more
suited to such a crop. Growing rice with flood irrigation in areas near to a desert,
which literally sops up the humidity from the evaporation of the flood irrigation
makes little sense, yet this is often seen in places like Egypt. Making sugars with
sugar beets makes more sense than with sugar cane in water-stressed areas.
150 P. Sullivan

Developing and using less-water-using varieties of many crops in Arab countries


could go a very long way towards reducing water use. Tightening up or changing
irrigation systems could also make a big difference.
Changing consumption and cropping patterns can go a long way towards
reducing the water footprints of a region that has both water and food stress, as well
as the underlying land stress. For examples, the production of beef requires about
15,500 L/kg, sheep is about 6100 L/kg, and chicken is about 3900 L/kg. Rice is
about 3400 L/kg, wheat about 1300 L/kg, maize about 900 L/kg, tomatoes about
180 L/kg and lettuce about 130 L/kg (Ercin and Hoekstra 2012, 2014;
Gerbens-Lennes et al. 2013; Hoekstra 2010, 2011, 2012, 2014; Hoekstra and
Mekonnen 2012; Hoekstra and Wiedmann 2014; Mekonnen and Hoekstra 2010a,
2010b, Mekonnen and Hoekstra 2011a, 2011c, 2012).1 These water footprints are
for the entire supply chain for these foods from farm to plate and other connected
activities to the supply chain for the product. Some foods use a lot more water per
kg of production than others. Livestock usually uses much more water than grains
because the water is used to produce the feed for the livestock, mostly grains first
and then the conversion from grain feed to meat, which is quite inefficient ener-
getically, in calories.
These numbers are for world averages. Given that some areas of the region are
quite hot and dry one would expect these numbers to be quite bit higher. It is a good
thing for regional water use for food production that many of these products are
imported into the region. It is simply quite odd that about 80 % of the water use in
the region goes to agriculture in an area with such water stress. This is especially so
for the GCC where there is not only extreme water stress, but also higher per-
centages of water use for agriculture than in places like Egypt, where renewable
fresh water is more plentiful per capita.
Add energy, water, and food subsidies together and the result is
over-consumption of water in water-stressed areas, over-consumption of food, and
over-consumption of energy to produce the livestock feed and food for the people
of the region. So far, the GCC seems to be able to handle the energy costs of
over-consumption. However, there is a growing concern not only about the drain
such complex nexus subsidy systems are making on energy supplies and potential
energy exports, but also on the budgets of the countries. All of the countries of the
GCC region are concerned with the over-consumption of water in desert climates.
Over-consumption of food has not been at the top of the planning and policy
committees in the region, but it could be a greater concern as their populations,
budgetary, and health stresses continue to grow—and if instability in the region
grows further, and those important sea lanes and ports for food imports are under
increasing risk.

1
See Waterfootprint.org (http://www.waterfootprint.org/?page=files/productgallery) for a truly
eye-opening set of figures for how much water goes into making various products from start to
finish in their product supply chains.
8 Water-Energy-Food Security Nexus in the Arab Region … 151

Yemen is a special case in the region. It is not part of the GCC, but a close and
very poor neighbor of these relatively wealthy countries. It is energy, water, and
food short. About 60 % of its water goes to growing qat. A large proportion of its
energy, maybe 30 % or more goes to getting water out of the ground and moving it
from one place to another, including by truck or pipelines. Yemen is wearing down
its energy resources. It is also importing about 80 % of its food given how little land
and water goes to producing food crops, instead, growing the narcotic qat. Yemen
may either have to move its capital, Sana’a, to another place closer to the sea and
desalination plants that could be constructed near the sea, or it will have to spend
massive amounts of money not only on desalination plants near the sea, but also in
moving that water up to where most Yemenis live in the highlands far away from
the sea (Heffez 2013; Kramer 2014; Sullivan 2012).
Saudi Arabia has developed massive dairy farms, cattle farms, and a significant
export industry of food for the region and beyond based on subsidized water, often
fossil water of desalinated sea water, subsidized land, subsidized labor, and in a tax
free environment. Much of this is based on the power of its oil export revenues.
However, if the world energy markets change, especially the energy used for
transportation, which is mostly based on oil, and if Saudi Arabia cannot move
towards non-oil desalination, then Saudi Arabia is in real trouble economically. The
Saudis could be walking into a water-energy-food trap that they would not be able
to get out of soon enough if they do not plan now, invest now, and develop policies
now to make their water-energy-food nexus more sustainable in the long run,
instead of just tossing out oil revenues to create desalinated water, which could be
used for other economic development and economic diversification efforts. If Saudi
trends in the use of oil in desalination continue, by 2040 they could be burning 8
million barrels a day to produce fresh water. They are already burning about 1.5
million barrels a day to desalinate. Imagine what the diversion of 8 million barrels a
day to desalination could do to world oil markets (World Bank 2012, p. 12). Similar
trends could be found for Kuwait and other large oil exporters in the GCC.
The region imports tens of millions of tons of wheat each year and wastes about
45 % of it. It wastes about 20 % of its cereals, 50 % of its fruits and vegetables,
16 % of its meat, and 27 % of its fish and seafood (FAO 2014f, p. 1). All food
wastage represents a missed opportunity. If one apple is wasted, then the numbers
might seem minimal to most. However, if millions of parts of apples, sheep, lamb,
cheese, and other heavy water-using agricultural products are wasted, then the water
and energy waste connected to the food waste can be just mind boggling. In some
countries in the region, the amount of food lost and wasted is greater the food deficit
in the country (FAO 2013c, p. 18). A good breakdown of the food lost and wasted
in the region can be found in some superior FAO documents on the subject (FAO
2011, 2014a, c). There is loss and waste at all stages of the farm-to-plate supply
chains in the region, and also along the import-to-plate supply chains for the region.
For the Arab Region about 36 % of the potential food consumed is lost and wasted.
About 10.8 % at harvest, 7.8 % post-harvest, 6.3 % at processing and packing,
5.6 % at distribution, and 5.5 % at consumption (FAO 2014a, p. 27). These
numbers may seem exacting, but as with other aspects of the water-energy-food
152 P. Sullivan

nexus in the region, the quality and quantity of the data are often more lacking than
many may want to admit. However, the overall trends and indicators on food losses
and waste in the region are stunning (FAO 2014f, p. 1).
There are many poor people in the Arab Region and this multiplicative waste
makes them and the region all the poorer, hungrier, energy poor, and water stressed.
Now consider that most people in the Arab Region are Muslims and that one of the
five pillars of Islam is charity, this waste seems all the more outrageous. Also, there
is massive waste, which has been complained about by even the religious author-
ities in Mecca, during and after Ramadan feasts. Ramadan fasting is another pillar
of Islam. This is supposed to, amongst other things, develop a sensitivity in the
better off Muslims for the plight of the hungry and poor. With all of the waste
during Ramadan this sensitivity seems a bit blunted. The phrase “Allah loveth not
the wasters” can be found many times in the Koran. So all of this energy, water, and
food waste is not only an economic issue, it is also a social, ethical, moral, and
religious issue.
In the MENA region about 250 kg of food is lost or wasted each year per capita
(FAO 2014c). That is about 43 cubic kilometers of water lost and wasted for lost
and wasted food FAO 2014f). This is about 90 % of Egypt’s allocation of the Nile
waters under the 1959 treaty. It makes one wonder why it is not taken more
seriously at the highest levels in government. Add in the fact that for many Arab
countries their oil and other export revenues go in good part to import food and
other agricultural goods, and the waste seems bordering on the absurd. This is
another side of the energy-food nexus. A lot of energy is exported to pay for food
imports, and hence virtual water imports, and then much of that imported food and
virtual water is lost and wasted.
Part of the drivers for food loss waste are the food subsidies in the region. Food
subsidies can cause threats to the health of the people in an Arab states, such as
diabetes, which is a huge problem in the region (International Diabetes Foundation
2014; Jones 2014; Kerr 2014; Mehrtash 2014). More is spent on energy subsidies
than on health in the region. Here is another connection in the water-energy-food
nexus that many may not consider (Devarajan 2014).

8.4 Water: Water Security, the Water-Energy


Nexus and Desalination in the GCC, More
on the Water-Food Nexus in the Region,
Water Losses and Waste

If the region was lush with rivers, sopped with connected and constantly refilled
aquifers, and had a lot of predictable rains, then many of the food, and even many
of the energy problems of the region could be resolved. This is a region of many
countries with often quite different water supplies, water demands, and water
security problems (FAO 2009a, b, c, d, e, f g, h, i, j, k, l, m, n; 2013a). However, the
8 Water-Energy-Food Security Nexus in the Arab Region … 153

region is one of the most water short in the word. It is a region of mostly no rain or
unpredictable rain patterns (UNDP 2013, p. 28). It is also a region of historically
declining per capita fresh water supplies (Devlin 2014). Total unmet demand in the
region, about 42 cubic kilometers, is about 80 % of the water allocation to Egypt in
its 1959 agreement on the Nile (or, interestingly about the same amount of water
that is lost and wasted in lost and wasted food in the region). The expected unmet
demand for water by 2030 could be almost three times that. The total unmet
demand by 2050 could be four times that (World Bank 2012, p. 4). These are truly
stunning numbers. The decline in water availability per capita that could happen in
the next few decades should be a cause for alarm in the region and in the world,
given the importance of this region for the economic and political stability of the
world (World Bank 2012, p. 27 and 31). If one were to look at some projected
demands and supplies of water for the region under various climate change sce-
narios, one could see some real problems brewing, especially in Egypt, Iraq, Jordan,
Morocco, Saudi Arabia, Syria, the West Bank, Gaza, and Yemen (World Bank
2012, pp. 34–35). Yemen’s capital Sana’a may have to move due to the declining
water table in its aquifers (Heffez 2013).
There are many ways of solving the looming problem of water gaps for the
region (World Bank 2012, pp. 7–9). One is desalination, and this is more prevalent
in the GCC than other areas of the region, but eventually more desalination may be
needed in the other areas. About 50 % of all desalination capacity in the world can
be found in the region, and almost all of that is in the GCC (World Bank 2012,
p. 9). All along the cost of the Arabian Gulf, desalination plants have been built to
fill in the water supply gaps that have developed as Saudi Arabia, Kuwait, Qatar,
Bahrain, and the UAE have developed economically. Bahrain gets about 79 % of
its fresh water from desalination. Qatar gets about 75 %, Kuwait about 70 %, the
UAE about 67 %, Oman about 44 %, and Saudi Arabia about 15 % (World Bank
2012, p. 64; Booz and Company 2014). Massive amounts of oil and money have
gone into running and paying for this desalination. About half of all the desalination
capacity of the world is found in the GCC. If the trends of oil use to desalinate water
in Saudi Arabia continue, for example, then the country will be using most of its oil
to create its fresh water in just a few years. Kuwait, the Bahrain, and the UAE are
also under the threat of considerable oil supply stress if the trends in their use of oil
for desalination continue. Qatar uses mostly natural gas for desalination given its
massive reserves. However, the UAE uses mostly natural gas for desalination, and
much of that natural gas is imported. In the GCC between 10 and 25 % of the
electric power produced goes to desalination of water (Fath et al. 2013).
Again, a large looming issue is what may happen if Saudi Arabia is so
oil-stressed by its desalination efforts that it can no longer export as much oil in the
world oil markets to be a big player. What would that mean for Saudi Arabia’s
economy? What would that mean for OPEC? What would that mean for the world
oil markets? These are questions that few are asking. Some thought and effort is
being put into moving towards renewable energy systems to desalinate water in the
region. Concentrated Solar Power production of electricity and desalination seems
to be one of the better options. However, the movement towards this seems quite
154 P. Sullivan

slow. The region has massive renewable energy potential and it should be used to
help solve this water crisis (World Bank 2012).
Little real effort seems to have been put forward in the GCC region towards
demand management. Most of the investments and policy efforts have been towards
supply development. Energy, water, and food are all subsidized in the region. This
is a region that is mostly desert and one of the most water-stressed in the world.
About 80 % of the water used goes to the agricultural sector in desert states. The
water is at a cost well below what would be needed to have its use sustainable. In
some countries in the GCC, the gap between costs of producing the water and the
revenues from selling the water is as high as 92 % (Booz and Company 2014, p. 7).
Some cities in this increasingly urbanizing Arab Region are doing better than
others. Non-revenue water can be over 40 % of the water use in places like the
West Bank, Cairo, Alexandria, Algeria, Jordan, and Djibouti (UNDP 2013, p. 45).
Water losses in the GCC are quite high (Booz and Company 2014, p. 8). Prices for
water in the GCC are unusually low for such wealthy countries (Booz and
Company 2014, p. 10). Recovery of operating costs for water utilities is rare, in
places like Rabat and Casablanca. The UAE and Kuwait only pay off about 10 % of
their overall costs. Alexandria, Cairo, and Bahrain may pay off about 25 % (Booz
and Company 2014 p. 51). Water subsidies and weak water governance and
supervision add up to water losses and waste. None of the Arab states is paying
back the costs of water in real terms when considering the losses and waste of this
vital natural resource. On average about 20 plus percent of the water used is either
wasted or lost in the region.
The percentage of water use in agriculture completely outweighs the importance
of agriculture to the economies of the Arab Region. In the GCC region agriculture
may be about 2 % of the overall GDP (Booz and Company 2014, p. 42), yet over
80 % of this precious water is going to agriculture (Booz and Company 2014,
p. 40). For the overall region agriculture has its highest percentage of GDP in the
poorer countries, such as Sudan, Syria, Morocco, Tunisia, and Egypt (UNDP 2013,
p. 42; Booz and Company 2014), but these are also countries where water waste
and losses may be at their worst. It is truly astonishing how small a percentage of
water goes to industry in the total Arab Region, and particularly in the GCC. It is
also clear more water is going to agriculture than would be optimal given all of the
other resource stresses involved (Booz and Company 2014, p. 12). For the GCC it
is clear they are using too much of their very scarce water, and clearly too much for
agriculture (Booz and Company 2014, p. 5).
Another way to cut back on the use of water and to make it more efficiently used
is to use less for agriculture, but using more efficient irrigation technologies and
methods and, possibly far more important, to focus on the water footprints
described earlier in this chapter, and change cropping and livestock patterns. Water
can be reused in agriculture. However, that would mean a lot of energy used to
clean up the water that already was poured onto the crops along with pesticides,
herbicide and fertilizers. Possibly using more efficient and sustainable pest control
8 Water-Energy-Food Security Nexus in the Arab Region … 155

and fertilizer techniques could cut the costs of reusing this water. Most of the
fertilizers used today are flushing into the local aquifers, rivers, lakes, and the sea
because the crops simply cannot absorb more than a small percentage of it. Putting
too much fertilizer on crops is not only water wasting, it is also energy wasting
because often these fertilizers are based on fossil fuels and they use a lot of energy
to make them.
Waste water treatment and reuse is another option for supply augmentation
(World Bank 2011). However, this is a lot rarer in the region than one might hope
(UNDP 2013, p. 24). Egypt has made the most effort on this. Syria was making
progress prior to its present nightmare. Saudi Arabia, Tunisia, the UAE, and a few
others are putting more effort and finance into this. One would hope that this could
be a good source for future water if it is done right, but, still, the biggest returns can
likely be found in the largest user of water in the region, agriculture.
It is very rare when the region considers the externalities connected with the
over-consumption and pollution of water. GDP losses due to water pollution and
degradation are not insignificant (UNDP 2013, p. 27). The gap between supply and
demand is growing. Water reuse is a tiny proportion of the solution to the problem
so far. Also, the evaporation of the agricultural water, in particular in the desert
climates of the region, pretty much assures that specific water, or a good part of it,
will not return to the place where it was taken out. The humidity that the evapo-
ration caused will be blown in the wind to somewhere else.
The costs of ground water depletion have been estimated by The World Bank to
be about 1–2 % of GDP each year for Tunisia, Yemen, Egypt, and Jordan, with
Jordan being the highest percentage loss (World Bank 2012, p. 37). Losses to Saudi
Arabia and others from the losses of ground water over the many years of trying to
grow wheat in the desert and other such projects may prove to be much higher in
the long run.
Another way of creating longer term supplies of fresh water is to set up more
dam systems, but these seem to have reached their limits in many areas. Also, there
are considerable international tensions to be considered in the river areas where
such dam systems might be set up, or even in the areas where underground aquifers
are pumped out to create surface lakes with dam systems, such as in Libya (UNDP
2013, p. 24). Water dependency ratios are quite high for many Arab states (UNDP
2013, p. 28). This dependency in an increasingly water-stressed environment could
either lead to more water cooperation or more conflict.
One of the biggest water, energy, and food changing policies would be the phase
out the subsidies for all three that are a growing budgetary burden, especially for the
poorer Arab states. Phasing out these subsidies could also help these countries with
their growing resource burdens, such as where they are going to get all of the
energy, water, and food they will need for their future populations and other needs.
Also, cutting back on the loss and waste of food could be a huge source of
recovered water.
156 P. Sullivan

8.5 Some Determinants of the Nature


of the Water-Energy-Food Security
Nexus of a Country

Natural resource endowments are one of the main determinants of the


water-energy-food security nexus in the Arab Region. Some countries, such as
Egypt, Sudan, Iraq, and Syria, have significant river systems and river basin
endowments. Others have very few rivers or none, yet have large underground
aquifers of fossil water, such as Libya, Saudi Arabia, and Jordan. Some have
somewhat significant rainfall in parts of the country, yet are mostly dry areas, such
as Algeria, Oman, and Yemen. Some countries rely on outside sources of water
more than others. Egypt is almost totally reliant on water from other countries in the
Nile Basin, especially Ethiopia. Iraq and Syria rely greatly on water emanating from
Turkey. Underground water flows from Saudi Arabia feed into aquifers in Kuwait
and Bahrain. Then there are countries, such as Qatar and the UAE, that are under
such massive water stress that their leadership is nearly apoplectic about it. Some
countries extract well over 100 % of their renewable fresh water resources. Many
Arab countries are living well beyond their means when it comes to fresh water so
artificial water, desalinated water, needs to be produced.
Some countries have newer, better maintained water systems. Others have older,
not so well-maintained water systems. Some have vast areas of their countries that
have no piped water and delivery of water is often by small truck, such as in Yemen
and parts of Jordan.
Some Arab countries have large endowments of oil and gas, such as Saudi
Arabia, Kuwait, Iraq, the UAE, Qatar, Libya, and Algeria. Others have to import
their oil and gas, such as Morocco, Tunisia, Egypt, and Jordan. Yet others are
wearing down their oil and gas endowments very quickly, such as Yemen. Then
there are the moderately endowed countries, such as Oman. Some Arab countries
have great potential resources in geothermal electric energy, such as Saudi Arabia,
Djibouti, and Jordan. Geothermal heating and cooling could be used in almost all of
the region given that at certain depths underground all across the region the tem-
peratures would remain at about 50°F. Solar radiation exists throughout the region,
but it is surely more intensive in the Sahara, and in many parts in Saudi Arabia,
Qatar, and the UAE. Wind energy endowments are also significant for many areas
in the region. Solar, wind, and geothermal are options for many areas in the region
and this needs to be looked into. Energy resources are not just oil and gas.
Some countries in the region have very well developed electricity and refining
infrastructures. Others do not. Some countries keep up with the maintenance of
these important infrastructures. Others do not or cannot. Libya, Yemen, and Syria,
countries in the midst of nightmarish civil wars or near civil wars, cannot keep up
with their energy needs and energy investments. Egypt was set back in its energy
investment and maintenance plans by its multiple revolutions and continued
instability. The GCC states have huge financial resources to expand and maintain
8 Water-Energy-Food Security Nexus in the Arab Region … 157

electricity and refining networks. Egypt, Jordan, Morocco, Yemen, and other poorer
countries have to stretch their budgets to keep up.
Some countries of the region have significant arable land resources and have put
considerable effort and investment into reclaiming land, such as Egypt. Others have
little arable land compared to their entire land mass, such as the UAE, Qatar,
Bahrain, and Saudi Arabia. Some countries, such as Egypt, Syria and Iraq, have
been blessed with being near river basins, which is a great thing for farming. Others
have vast deserts dotted with oases and little immediate possibilities for productive
food and other agricultural lands, such as Libya, and much of Algeria. Morocco has
the luck of location and topography in its arable land resources. Yemen does as well
in parts of the country, but that has been wasted as a potential food resource by
using much of that land and its underground water resources for growing the
narcotic qat.
The water-energy-food complex can be partly determined by events. For
example, the droughts that ravaged northern Syria, Iraq, and part of Iran in recent
years changed the nature of the water-energy-food nexus in those areas. Aquifers
were drained. The lack of rain prompted greater use of not only aquifers, but also of
the fuel to pump the water out of those aquifers. The drought prompted a massive
outflow of people from northern Syria, which some believe helped spark the Syrian
revolution. The completion of the Great Ethiopian Renaissance Dam could sig-
nificantly impact the water flows to Sudan and Egypt not only as the reservoir
behind it is filled, and especially if the water in the reservoir is used for irrigation.
The water-energy-food complex can be partly determined by personalities of the
leadership involved in the policy developments related to energy, water, and food.
Saddam Hussein redirected water to develop dam systems and run-off lakes. He
also drained the areas of the south where some of his opposition was found. The
leadership of Saudi Arabia for many years allowed its aquifers to be drained in
order to grow wheat in the desert. Muammar Ghaddafi started the Great Man-Made
River Project in his attempts to green a country that is about 90 % desert and make
Libya a large food exporter. Hosni Mubarak tried to develop the massive Toshka
project to create a new branch of the Nile. During his time energy, water, and food
subsidies continued to grow as a burden on the Egyptian budget, even with attempts
to reduce such subsidies. President El-Sisi phased out some of the gasoline sub-
sidies, yet did not face unrest. If such a change happened during the last years of
Mubarak, then considerable unrest may have happened. Jordan reduced its energy
subsidies and some dissent and demonstrations happened.
Good leadership and much better policy development and resource uses across
the water-energy-food nexus could make a huge difference in the sustainability of
not only the nexus in the region, but of many countries in the region.
158 P. Sullivan

8.6 Part of the Solutions to the Energy-Water-Food Nexus


Stress and Its Potential Results

Part of the solution in the medium to long run, is for a country to take a hard look at
what it is using its water for, how it gets its water, the market for its water, and the
technologies it is using for water production, treatment, reuse, and efficiency. The
country also needs to take a hard look at what it is using its energy for, how it gets it
energy, the market for its energy (both domestically and internationally), and the
technologies it is using for energy production, processing, distribution, and effi-
ciency. A country also needs to look hard at its food, feed, and other agricultural
markets. What food is it producing? How is it producing the food? How is the food
distributed and sold? How is it packaged and transported? What are the markets like
for food, feed, and other agricultural products produced in the country and imported
from the outside? How can water use be changed via changed cropping, irrigation,
and other parts along the supply chain for agricultural goods? Does the country
really know how much water goes into its food along the entire life-cycle of each
and every agricultural product? How can water use be changed along the supply
chain for the country’s energy? How can energy use change along the supply chain
for the country’s water? How can both energy use and water use be changed along
the supply chains for both at the same time? How can food consumption and
production be changed along with the water consumer and produced to create the
food along the entire supply chain of the food-water nexus? How can energy use be
changed along the supply chain of food in the country?
Other important questions are: How much food is wasted in the country? How
much water is wasted in the country? How much energy is wasted in the country?
How much of this waste in energy, food, and water can be reduced by having
different supply chains for each, and how can each supply chain within and across
these nexus products and services be developed in different ways to reduce this
waste?
A big part of the problem of finding solutions to the waste and abuse of resources
in the Arab Region is the lack of data to figure out what to do. Water use data are
often lacking in many parts of the region, especially in places like Yemen. The use
of the Nile water energy use data also has significant problems due to the lack of
proper metering and the problem of non-revenue electricity in some parts of the
region. There is considerable smuggling of gasoline and diesel between Syria and
Turkey, and between Algeria and Morocco.
Getting the water-energy-food nexus right could add considerably to the growth,
development and sustainability prospects for the region. The region is facing water
scarcity now and will face worsening water scarcity in the future under most
scenarios. Developing more diverse and economical water production methods can
help. Developing much better demand management can make a very big difference.
Developing a fully functioning integrated water policy within and across countries
can make a major dent in the overuse and abuse of water in the Arab Region.
Phasing out the subsidies to water and setting proper pricing for water could cause
8 Water-Energy-Food Security Nexus in the Arab Region … 159

some political and social upset if it is not done properly. However, by not phasing
in proper pricing, the region could face increasingly severe water stress and
increasing severe social stress and conflict.
Water needs to be more carefully included in the agricultural and energy policies
of the region. Energy extraction, processing, transport, and end uses have some
water component to them. A mostly dry area such as the Arab Region needs to
reconsider how it will produce electricity in the future. There are many good
arguments for developing replacement electricity methods that use much less water,
such as solar, wind, and other renewable energy sources that can use much less
water, than the typical thermal electricity plant based on gas and oil that are in the
region. Concentrated solar power can use a lot of water, but clearly not as much as
thermal generation. Concentrating Solar Power (CSP) can also be used to desalinate
water. This would not only save water, but also create fresh water. Refineries, crude
oil facilities, and other energy uses for water can also start to focus on water-saving
technologies and methods.
In a place that is so dry, it makes little sense to have 80 % of its water going to
agriculture. It makes far more sense to import the water-using foods and other
agricultural goods rather than draining aquifers or using up massive amounts of
energy to create desalinated water to pour onto hot, dry soils to grow crops. There
may still be a significant place for agriculture in the region, especially in areas
where there is good arable land and good sources of blue, green, and grey water,
such as in the delta of Egypt or some of the best farm lands of Morocco and
Lebanon. However, subsidizing water and energy to produce dairy goods in the
deserts is unsustainable by definition. Technological and scientific experiments,
also heavily subsidized, can help postpone, if only psychologically, the final
decisions to move away from a distorted water-energy-food nexus, but those final
decisions will surely have to be made.
A main question is: how long can countries like Qatar, the UAE, Saudi Arabia,
and others fool themselves into thinking that their water deficits will not require
massive changes in the water-energy-food nexus and the energy, water, and food
systems of their countries and the region? Oil money can be converted into inef-
ficient agricultural systems and desalination for just so long before the costs of
doing so will eat up the budgets of many countries that rely on oil export revenues.
The water deficits of the future in the region will happen in a very bad way unless
new energy and food systems are developed correspondingly with water reuse,
treatment of polluted water, water demand management, and more.
Relying on just desalinated water to fill in the expected water gaps will be
absurdly expensive when changes in prices and governance can make the transi-
tions cheaper and more politically bearable. Given that agriculture uses the grand
majority of water in the region, then the biggest bang for the buck could come from
tightening up the use of water in agriculture by improving irrigation methods and
by changing crop types to those that make more technical, economic, and resource
sense in a dry, and possibly increasingly drier environment in the region. It is also
odd that agriculture is such a small proportion of the GDPs of the GCC countries,
160 P. Sullivan

yet a massive drain on their water reserves and mostly quite limited renewable
water.

8.7 Some Potential Global Implications


(and There Are a Lot More)

If the water-energy-food nexus policies are not developed properly in the long run,
there could be extreme resource stresses in the region beyond what have been seen
to date. These resource stresses could have global implications in increased energy
insecurity, increased terrorism, and increased migration from the region to more
stable, more economically developed, and more resource rich areas.
Continuing to waste water, energy, and food in the region, given the deep and
complex relations amongst these resources could lead to, for example, much less oil
being exported from the region. How might this work? Let us take the example of
Saudi Arabia, which uses massive amounts of oil to desalinate water for industrial,
residential, and also agricultural use. As desalination needs increase with increased
population, then more oil would be needed to produce those increased water needs.
More water would be needed for the increased food needs in the country. Saudi
Arabia has a considerable agricultural and food processing sector that will likely
grow. As it grows it will need more subsidized water produced with internally
subsidized energy to help develop the subsidized food production to be purchased
at markets in the Kingdom at subsidized prices.
These nexus linkages will only get more stressful unless the subsidies are lifted,
in a phased manner most likely, and if demand for water, energy, and food are
incentivized in more sustainable ways than they are now.
If we look at other countries in the Arab Region, such as Egypt, Jordan, Algeria,
and Yemen, the effects could be quite a bit different than mere energy stress from
less exports of energy. The water-energy-food stresses could quite easily, if not
handled properly and in an integrated, nexus and long term manner, turn into
drivers for greater political instability, insurrections, revolutions, terrorism, and
more. The populations most at risk for pushing some of these countries over the
tipping points toward these political, police, and military problems would be the
youth, who would begin to lose even more hope as they see the resources of their
countries get worn down further and their economic opportunities get more limited.
Also, those declining resources and economic opportunities could most likely
drive more people, and especially young people, to immigrate in much larger
numbers towards Europe and beyond. This would create even more stress between
the Arab states and the richer states near to them—or even some very far from them.
Hungry and thirsty people with little resource and economic hope may also turn to
terror groups for support and sustenance. These resource stresses could act as threat
multipliers, not only for insurrections and revolutions, but also for terrorism.
8 Water-Energy-Food Security Nexus in the Arab Region … 161

8.8 This Is All Mostly Solvable via Integrated


and Sustainable Nexus Policies and Investments

The problem with oil export stress due to increased oil use to desalinate water for
many uses, including in the production of food, could be partially solved by
changing the technologies for desalination to solar, such as concentrated solar
systems, solar panels, solar power towers, and more, and also towards geothermal,
wind, and other renewable technologies. Nuclear desalination is also an option,
although there are political and other effects that would need to be considered with
the development of nuclear desalination in the region. Saudi Arabia could also
benefit by getting rid of the distorting and stress-inducing subsidies that cut across
the water-energy-food resource nexus. Water, energy, and food waste is prevalent
in the country and this cannot continue in the long run without serious implications
for the political stability and sustainability of the country.
The same could be said for Egypt, Jordan, Algeria, Yemen, and many other
countries in the region. Food waste is massive. Water waste is gigantic. Energy
waste is stunning. And all of these add up to a nexus of waste, which if not handled
with politics that cut across the nexus, could lead to insurrections, revolutions, civil
wars, and more. This has already started in Yemen. Its conflicts are as much, if not
more, resource based as ethnically based. Differing ethnicities and ethnic stress can
tend to magnify resource stresses. This can apply to Iraq, Saudi Arabia, Bahrain,
and many other countries in the region.
Subsidy, pricing, regulatory, land, technology, environment, energy, food, and
water policies all have to be developed with the nexus in mind—and with the nexus
of waste and stress in mind. The time to think through what is needed for the future
is now. A new set of leaders needs to be educated in the nexus of resources, waste,
stress, and more. There need to be serious education and training programs to get
the leadership and the people up to speed about what could be facing them in the
future if the nexus connections in many levels are not considered and incorporated
into policies and practices.
When we see major Arab leaders speaking about the problems and prospects of
the nexus in their countries, and what needs to be done, then we might be on the
right path to a better future for the region. If countries still waste, subsidize, and
continue on the path of unsustainable policies and practices, then the hope for a
better Arab world could be stunted considerably—and that is the last thing this
region needs.

8.9 Policy Recommendations List

1. There needs to be considerably better and more comprehensive data collection,


analysis, and sharing on the water-energy-food nexus in the region.
162 P. Sullivan

2. It needs to be understood that each Arab country in the region has a different
water-energy-food nexus than the others, but they are connected in many
complex ways. One-size-fits all policy prescriptions and technology solutions
may not work across the region, yet may work well in one country rather than
another.
3. The variety of cultures, politics, economics, and histories of the peoples of the
region need to be considered when local, country, and regional policy and other
options are being developed.
4. Comprehensive education and outreach programs need to be developed in the
region directed at all social, economic, and educational strata of the region on
the meaning, economics, technologies, and policies related to energy, water,
and food and the water-energy-food nexus.
5. Comprehensive education and outreach programs need to be developed in the
region to explain what energy and water footprints mean and how they are
related to the everyday lives of the people of the region.
6. Comprehensive education and outreach programs need to be developed in the
region to explain the importance of energy, water, and food efficiency.
7. Comprehensive education and outreach programs need to be developed in the
region to explain the real and total costs of energy, water, and food waste in the
region.
8. Arab universities, think tanks, and other related institutions need to obtain the
funding and expertise to work on the solutions to the looming
water-energy-food nexus crises.
9. Arab universities, schools, think tanks, and other related institutions need to
obtain the funding and expertise to develop economic, resource, financial, and
technological literacy for its people, managers, leaders, militaries, and others to
help understand the connections between economics, politics, technology,
resource thinking, finance, and the water-energy-food nexus.
10. The water-energy-food nexus should be part of the curriculum of every public
educational institution and private educational institutions with public funding
from pre-school to Ph.D. programs. There are many brilliant Arabs. Give them
the ideas and the background and they will help solve the problems.
11. Governments of the region and multilateral institutions should invest or begin
to invest in developing the jobs of the future related to energy, water, and food
security and the water-energy-food nexus. The militaries of the region could
also play their part in this.
12. The governments of the region need to develop the capacity to handle at the
policy, technological, and economic levels the problems and crises that will
come their way in energy, water, and food security as well as the stresses within
the water-energy-food nexus.
13. If any government policy changes are to happen, then there needs to be a
serious and forthcoming outreach program explaining these policy changes to
the people, businesses, and other institutions and organizations of the region.
This could prove to be vital on the issues of the inevitable phasing out of
subsidies in energy, water, and food.
8 Water-Energy-Food Security Nexus in the Arab Region … 163

14. When subsidies are going to be phased out in energy, water, and food they need
to be done in a coordinated way and with a sensitivity to the political and
economic environments of each country. It would be best to phase these out
while considering the water-energy-food nexus as the basis of decision making.
15. Taxation and investment policies in the countries need to be more focused on
dealing with the resource, economic, social, and political stresses related to the
water-energy-food nexus.
16. Pilot programs need to be started not only at the leadership levels, but also at
the village, town, and other levels to work as communities and in private-public
partnerships on developing a better way of handling the energy, water, land,
and other resources of the region.
17. A serious set of competitions should be established in the region with signif-
icant cash prize awards and other benefits to people, private companies, and
government and military leaders to develop solutions to the water-energy-food
nexus crises of the future and in new ways of thinking about and developing
ideas for the future of these issues.
18. The best of the militaries in the region could act as vanguards for technological
and other changes directed at improving energy, water, and food security.
19. Nexus thinking should become the norm in leadership right down to the
average person in the Arab world. Why not have the Arab world as the van-
guard for change in thinking about these issues, given how important the nexus
is for its prosperity, peace, survival, and more in its future?
20. The Arab Region should become the melting pot of ideas from all over the
world to help resolve the issues related to the water-energy-food nexus.
21. Significant scholarships should be developed by the richer Arab states not only
for their people, but for persons from the less well-off Arab states to have their
best and brightest go abroad for training in the issues related to the
water-energy-food nexus. Corruption and nepotism in the scholarship programs
should be minimized.
22. Guaranteed higher level jobs should be waiting for the best and brightest who
got these scholarships in order to groom them for the future resource leadership
positions of the region. They should be required, with no exceptions, to work
on water-energy-food nexus issues for at least for the number of years that they
were away for training.
23. Some thought and public and private funding could be put towards the
development of water-energy-food nexus extension services in the region, as
have been developed for agricultural extensions services globally.
24. Considerable internal and external diplomatic efforts need to happen to begin to
change the social contracts in the region regarding energy, water, and food and
the water-energy-food nexus.
25. The domestic and international tensions and politics regarding water need to be
handled in more effective and practical ways.
26. International laws with regard to shared energy and water resources need to
have some enforcement mechanism to make sure they are followed. This is
particularly important in the Arab Region given how many underground
164 P. Sullivan

aquifers, rivers, lakes, and more are shared by countries within the Arab Region
and with others. The building tensions related to the Great Ethiopian
Renaissance Dam between Egypt and Ethiopia is a clear example to have better
mechanisms for conflict resolution before the conflict gets out of hand.
27. There should be an open and transparent network of energy, water, food, and
other libraries and digital resources for the region, which connect the people,
scholars, inventors, and others in the region to the information and opportu-
nities for making better decisions, better resource trade-offs, and more for the
region.
28. Many of these policy prescriptions could be labor-using and
education-developing, which could surely help the youth of the region.
29. Desalination will likely be a big part of the water future of the region. Given
how much oil and gas may be used in desalination if the trends continue, there
should be great efforts and investment put forward to look into more energy
efficient desalination techniques and more desalination from renewable energy,
such as solar, wind, geothermal, and others. Some efforts and investments may
also be needed in the further development of nuclear-powered desalination.
30. Desalinated water being used to irrigate water-intensive crops and to run
water-intensive businesses, such as dairies needs to be reconsidered.
31. There needs to be some way to value and charge for the use of the underground
aquifers of the region that are being drained, in order to either help replenish
them to some degree or to at least postpone the inevitable water disasters that
may happen, especially in places like Yemen.
32. There should be strict and transparent policies developed for the reuse and
treatment of water. Reuse of water in the Arab Region will need to be far more
common in the future as it heads to much greater water supply gaps in the
coming decades.
33. There should be incentives developed to make more efficient uses of water at
the household and other levels be worthwhile for those who will invest in them.
34. More efforts and investments need to be put toward reducing food waste in the
region. There is a shocking amount of food waste for a place that has a serious
problem with food import dependence as well as stunting and obesity, and with
the health issues that go with that.
35. There should be incentives developed at the household to all other levels to
reduce food waste and losses. These investments need to be developed along
the entire food and other agricultural products chains, not just at the consumer
levels.
36. More efforts and investments need to be put towards either importing or
developing less energy intensive ways of treating, moving, and extracting
water.
37. More efforts and investments need to be put towards either importing or
developing less water-intensive irrigation and other agricultural methods all
along the food chain in the region.
38. There needs to be a greater understanding and application of the water foot-
prints of all of the products produced in the region and activities engaged in
8 Water-Energy-Food Security Nexus in the Arab Region … 165

within the region. With that greater knowledge, better policies and water
pricing can be developed.
39. More effort could be put into food and other trade to reduce the water footprint
of agriculture in the region. The biggest use of water in the region is in agri-
culture. Often the crops grown and the foods produced do not correspond to the
water and other resource stresses in the region because of distorted pricing of
water and those other resources.
40. The water systems of the region do not pay for themselves. This has to change
not only to make them more financially viable, but also to make sure that the
pricing of water pays back the investments made in the water systems.
41. Energy needs to be used more efficiently and effectively from household to
industry, government, and other levels. This will not happen until energy prices
are brought in line with their natural resource and other costs.
42. Low fuel prices for gasoline and diesel prompt overconsumption of those fuels.
This leads to greater traffic congestion, more fatal and non-fatal traffic acci-
dents, great pollution, and related health effects.
43. Low fuel prices, especially for diesel, prompt farmers that use diesel pumps for
irrigation to over-irrigate the fields. This leads to over-abstraction of water,
salinity of fields, over-use of fuels, and a degradation of the soil, water, and
more. This may also lead to lower crop yields over the long run, hence making
the food security situation worse.
44. There need to be policies developed to make the choices of more energy
efficient technologies for households, factories, farms, hotels, military bases,
and more be worthwhile for the changes to occur. Without the proper incentives
little will likely happen.
45. Tax, investment, and other incentives should be coordinated across the
water-energy-food nexus. Having separate policy changes is not enough.
46. There is a lot of talk and writing about having integrated water policies in the
region. There should be integrated energy, water, and food policies in the
region that take into account the water and energy footprints of products and
processes involved, and in the overall energy, water, and food security of the
target country and of the region.
47. Reasoned and moderated policy changes should start the changes off. Too strict
or too shocking policy changes could backfire with the public and others.
48. Some form of mitigating compensation should be given, especially to the poor
and the near poor, when the subsidy and other changes that may put them at
economic risk are applied. These changes must be phased in an integrated
manner and in consideration of the on-the-ground realities of each area or
country.
49. It is not just separate energy policies or water policies or food policies that will
save the region from potential resource disasters and the political and other
fallout from them. These policies have to work together in a nexus manner.
50. Good leaders need to be developed in the region that will focus on the
water-energy-food nexus. Without good leaders and change champions, many
of the best of ideas could end up to be the causes of more harm than good.
166 P. Sullivan

51. Some countries are up to the challenges of quicker changes in policies, tech-
nologies, and social contracts. Others are not. Changes need to take this into
consideration.
52. Changes need to be set up into something like a portfolio diversification, rather
than wholesale tossing out of the old to be replaced with the new.

References

Abdelaziz K (2013) At least 27 dead in Sudan fuel subsidy protests: medical source. Reuters, 26
Sep 2013. Retrieved from: http://www.reuters.com/article/2013/09/26/us-sudan-protest-
idUSBRE98P09120130926
Abruzezze L (2014) Global food security. IMF, 24 Feb 2014. Retrieved from: http://www.imf.org/
external/np/seminars/eng/2014/food/pdf/Abruzzese.pdf
Abu Zeid K (2013) Water-land-waste nexus in the Arab region. water-soil-waste nexus workshop
Dresden, Germany 11–12 Nov 2013. Retrieved from: https://flores.unu.edu/wp-content/
uploads/2013/11/09-Abu-Zeid-Water-Land-Waste-Nexus-Issues-in-the-Arab-Region.pdf
AFP (2013) Smuggled Libyan oil fuels Tunisia economy—and instability. Global Post, 2 Apr
2014. Retrieved from: http://www.globalpost.com/dispatch/news/afp/140402/smuggled-libya-
oil-fuels-tunisia-economy-and-instability
Al-Khatteeb L (2014) How Iraq’s black market in oil funds ISIS. CNN, 22 Aug 2014. Retrieved
from: http://edition.cnn.com/2014/08/18/business/al-khatteeb-isis-oil-iraq/
Al-Makhfi J (2013) Algeria smuggling crackdown cuts fuel line to Morocco. Arab News, 28 Sep
2013, Retrieved from: http://www.arabnews.com/news/466102
Al-Zubari K (2013) Water-energy-food nexus in the arab region. Arab regional implementation
meeting for the twentieth session of the commission for sustainable development (CSD-20),
29–30 May 2013, Dubai, UAE. Retrieved from: http://css.escwa.org.lb/sdpd/2044/s22.pdf
Alexander C, Krause-Jackson F (2013) Syrians turn to backyard refining as war reaches oil.
Bloomberg Business, 25 Apr 2013. Retrieved from: http://www.bloomberg.com/news/2013-
04-24/syria-open-backyard-refineries-as-war-reaches-oil-field.html
Allan T (2014) Water, food, trade—energy and climate change. OCP policy center conference
series, King’s College London, 11–13 June 2014. Retrieved from: http://www.ocppc.ma/
ckfinder/userfiles/files/1%20Allan-Nexus%2020140611%281%29.pdf
Andrews-Speed P, Bleischwitz R, Boersma T, Johnson C, Kemp G, VanDeveer SD (2012) The
global resource nexus: the struggles for land, energy, food, water, and minerals. Trans-Atlantic
Academy. Retrieved from: http://www.transatlanticacademy.org/sites/default/files/
publications/TA%202012%20report_web_version.pdf
Andrews-Speed P, Bleischwitz R, Boersma T, Johnson C, Kemp G, VanDeveer SD (2015) Want,
waste or war?: the global resource nexus and the struggle for land, energy, food, water, and
minerals. Routledge, NY p 240
Axworthy T, Adeel Z (eds) (2014) Global agenda 2013: water, energy and the Arab awakening.
United Nations University UNU-INWEH. Retrieved from http://inweh.unu.edu/wp-content/
uploads/2014/10/Water-Energy-and-the-Arab-Awakening_Web.pdf
Bailey R, Willoughby R (2013) Edible oil: food security in the Gulf. Energy, environment and
resources, Chatham House, Nov 2013. Retrieved from: http://www.chathamhouse.org/sites/
files/chathamhouse/public/Research/Energy,%20Environment%20and%20Development/bp1113
edibleoil.pdf
Bauer D (2014) Water-energy nexus: challenges and opportunities. US Department of Energy, 5
Sep 2014. Retrieved from: http://energy.gov/sites/prod/files/2014/09/f18/2014-09-05_Energy
%20Water%20Nexus_SEAB%20Presentation.pdf
8 Water-Energy-Food Security Nexus in the Arab Region … 167

Begum K (2014) CO2 Emissions in GCC Countries. BQ Magazine, 14 Aug 2014 http://www.
bqdoha.com/2014/08/co2-emissions-gcc-countries
Blanchard C (2009) Iraq: oil and gas legislation, revenue sharing US policy. In: CRS report for
congress, 3 Nov 2009, Washington, DC
Booz and Company (2010) Gas shortage in the GCC: how to bridge the gap, 13 June 2010.
Retrieved from: http://www.strategyand.pwc.com/media/file/Gas_Shortage_in_the_GCC.pdf
Booz and Company (2014) Achieving a sustainable water sector in the GCC: managing supply and
demand, building institutions, 8 May 2014. http://www.strategyand.pwc.com/global/home/
what-we-think/reports-white-papers/article-display/achieving-sustainable-water-sector-gcc
Bridle R, Kiston L, Wooders P (2014) Fossil-fuel subsidies: a barrier to renewable energy in five
Middle East and North African countries. Global subsidies initiative research report, Sep 2014
IISD. Retrieved from: http://www.iisd.org/gsi/sites/default/files/fossil-fuel-subsidies-
renewable-energy-middle-east-north-african-countri%20%20%20.pdf
Charles C, Moerenhout T, Bridle R (2014) The context of fossil-fuel subsidies in the GCC region
and their impact on renewable energy development. Global subsidies initiative research report,
July 2014 IISD. Retrieved from: http://www.iisd.org/gsi/sites/default/files/ffs_gcc_context.pdf
Clean Energy Business Council (2014) Water and energy in MENA: challenges, opportunities and
potential. Clean Energy Business Council, Jan 2014. Retrieved from: http://www.
cleanenergybusinesscouncil.com/site/resources/files/Energy-and-Water-in-MENA.pdf
Daher B (2014) Food security in the desert: what does it take? OCP policy center conference
series, King’s College London, 11–13 June 2014. Retrieved from: http://www.ocppc.ma/
ckfinder/userfiles/files/1%20-%20Daher_The%20Water%20Food%20Energy%20Nexus%20in
%20Drylands_Bridging%20Science%20and%20Policy_OCP%20Policy%20Center_Rabat_June
%202014%281%29.pdf
Del Grando J, Coady D, Gillingham R (2010) The unequal benefits of fuel subsidies: a review of
evidence for developing countries. IMF Working Paper (WP/10/202). Retrieved from: http://
www.imf.org/external/pubs/ft/wp/2010/wp10202.pdf
Devarajan S (2014) Corrosive subsidies in MENA. Future development, The World Bank, 12 Nov
2014. Retrieved from: http://blogs.worldbank.org/futuredevelopment/corrosive-subsidies-mena
Devlin J (2014) Is water scarcity dampening growth prospects in the Middle East and North
Africa? Brookings Institution, 24 June 2014. Retrieved from: http://www.brookings.edu/
research/opinions/2014/06/24-water-scarcity-growth-prospects-middle-east-north-africa-devlin
DOE (2014) The water-energy nexus: challenges and opportunities. US Department of Energy,
July 2014. Retrieved from: http://energy.gov/sites/prod/files/2014/06/f16/Water%20Energy%
20Nexus%20Report%20June%202014.pdf
Droogers P Immerzeel WW, Terink W, Hoogeveen J, Bierkens MFP, van Beek LPH and Debele B
(2012) Water resource trends in the Middle East and North Africa towards 2050. Hydrol Earth
Syst Sci 16:1–14. http://www.futurewater.nl/wp-content/uploads/2012/09/hess-2012-137-
typeset_manuscript-version5.pdf
Duncan G (2013) Middle East instability poses greatest threat to GCC’s food security, report says.
The National Business, 11 Nov 2013. Retrieved from: http://www.thenational.ae/business/
industry-insights/economics/middle-east-instability-poses-greatest-threat-to-gccs-food-security-
report-says
Economic Research Service USDA (2013) US wheat trade, 24 Jan 2013. Retrieved from: http://
www.ers.usda.gov/topics/crops/wheat/trade.aspx#world
Economist Intelligence Unit (2010) The GCC in 2020: resources for the future. A report from the
economist intelligence unit. Retrieved from: http://graphics.eiu.com/upload/eb/GCC_in_2020_
Resources_WEB.pdf
Economist Intelligence Unit (2014) Food security in focus: Middle East and North Africa 2014.
A report from the economist intelligence unit. Retrieved from: http://foodsecurityindex.eiu.
com/Home/DownloadResource?fileName=EIU%20GFSI%202014_Middle%20East%20and%
20North%20Africa%20regional%20report.pdf
168 P. Sullivan

EIA (2014) Total primary energy consumption per capita. US Energy Information Administration.
Retrieved from: http://www.eia.gov/cfapps/ipdbproject/iedindex3.cfm?tid=44&pid=45&aid=
2&cid=regions&syid=2005&eyid=2009&unit=QBTU
EPA (2014) Global greenhouse gas emissions data. US Environmental Protection Agency.
Retrieved from: http://www.epa.gov/climatechange/ghgemissions/global.html
Ercin AE, Hoekstra AY (2012) Carbon and water footprints: concepts, methodologies and policy
responses. UNESCO, Paris http://unesdoc.unesco.org/images/0021/002171/217181E.pdf
Ercin AE, Hoekstra AY (2014) Water footprint scenarios for 2050: a global analysis. Environ Int
64(2014):71–82 http://waterfootprint.org/media/downloads/Ercin-Hoekstra-2014-WaterFoot
printScenarios2050.pdf
Erdil M (2014) Smuggled fuel hits Turkish Markets. Hurriyet Daily News, 9 July 2014. Retrieved
from: http://www.hurriyetdailynews.com/smuggled-syria-fuel-hit-turkish-market.aspx?page
ID=238&nID=68858&NewsCatID=348
Espinoza R (2012) Government spending, subsidies and economic efficiency in the GCC. Oxford
centre for the analysis of resource rich economies. University of Oxford, July 2012. Retrieved
from: http://www.oxcarre.ox.ac.uk/files/OxCarreRP201295.pdf
Fan S (2014) Overcoming the triple burden of malnutrition in the MENA region. International
Food Policy Research Institute, 16 Oct 2014. Retrieved from: http://www.ifpri.org/blog/
overcoming-triple-burden-malnutrition-mena-region
FAO (2006) Livestock’s long shadow: environmental issues and options. Rome. Retrieved from:
http://www.fao.org/docrep/010/a0701e/a0701e00.HTM
FAO (2009a) AQUASTAT database, Bahrain. Food and Agriculture Organization, Rome.
Retrieved from: http://www.fao.org/nr/water/aquastat/countries_regions/BHR/index.stm
Accessed 10 May 2015
FAO (2009b) AQUASTAT database, Egypt. Food and Agriculture Organization, Rome. Retrieved
from: http://www.fao.org/nr/water/aquastat/countries_regions/EGY/index.stm Accessed 10
May 2015
FAO (2009c) AQUASTAT database, Iraq. Food and Agriculture Organization, Rome. Retrieved
from: http://www.fao.org/nr/water/aquastat/countries_regions/IRQ/index.stm Accessed 10
May 2015
FAO (2009d) AQUASTAT database, Jordan. Food and Agriculture Organization, Rome.
Retrieved from: http://www.fao.org/nr/water/aquastat/countries_regions/JOR/index.stm,
Accessed 10 May 2015
FAO (2009e) AQUASTAT database, Kuwait. Food and Agriculture Organization, Rome.
Retrieved from: http://www.fao.org/nr/water/aquastat/countries_regions/KWT/index.stm
Accessed 10 May 2015
FAO (2009f) AQUASTAT database, Lebanon. Food and Agriculture Organization, Rome.
Retrieved from: http://www.fao.org/nr/water/aquastat/countries_regions/LBN/index.stm
Accessed 10 May 2015
FAO (2009g) AQUASTAT database, Libya. Food and Agriculture Organization, Rome. Retrieved
from: http://www.fao.org/nr/water/aquastat/countries_regions/LBY/index.stm Accessed 10
May 2015
FAO (2009h) AQUASTAT database, Oman. Food and Agriculture Organization, Rome. Retrieved
from: http://www.fao.org/nr/water/aquastat/countries_regions/OMN/index.stm Accessed 10
May 2015
FAO (2009i) AQUASTAT database, Qatar. Food and Agriculture Organization, Rome. Retrieved
from: http://www.fao.org/nr/water/aquastat/countries_regions/QAT/index.stm Accessed 10
May 2015
FAO (2009j) AQUASTAT database, Saudi Arabia. Food and Agriculture Organization, Rome.
Retrieved from: http://www.fao.org/nr/water/aquastat/countries_regions/SAU/index.stm
Accessed 10 May 2015
FAO (2009k) AQUASTAT database, Sudan. Food and Agriculture Organization, Rome. Retrieved
from: http://www.fao.org/nr/water/aquastat/countries_regions/SDN/index.stm Accessed 10
May 2015
8 Water-Energy-Food Security Nexus in the Arab Region … 169

FAO (2009l) AQUASTAT database, Syria. Food and Agriculture Organization, Rome. Retrieved
from: http://www.fao.org/nr/water/aquastat/countries_regions/SYR/index.stm Accessed 10
May 2015
FAO (2009m) AQUASTAT database, UAE. Food and Agriculture Organization, Rome. Retrieved
from: http://www.fao.org/nr/water/aquastat/countries_regions/ARE/index.stm Accessed 10
May 2015
FAO (2009n) AQUASTAT database, Yemen. Food and Agriculture Organization, Rome.
Retrieved from: http://www.fao.org/nr/water/aquastat/countries_regions/YEM/index.stm
Accessed 10 May 2015
FAO (2011) Global food losses and food waste: extent, causes and prevention. Rome. Retrieved
from: http://www.fao.org/docrep/014/mb060e/mb060e.pdf
FAO (2013a) Coping with water scarcity in the Near East and North Africa. In: Regional
conference for the near east fact sheet (NERC-32). Retrieved from: http://www.fao.org/3/a-
as215e.pdf
FAO (2013b) Food security and nutrition in the Southern and Eastern Rim of the Mediterranean
Basin. FAO Regional Office for the Near East, Cairo, 2013. Retrieved from: http://www.fao.
org/3/a-i3206e.pdf
FAO (2013c) Report of the expert consultation meeting on food losses and waste reduction in the
Near East Region: towards a regional comprehensive strategy. FOA Regional Office for the
Near East. p 67. Retrieved from: http://www.fao.org/documents/card/en/c/c4e9e0a0-52a8-
4e28-bf9b-cf2986808ecc/
FAO (2014a) Food losses and waste in the context of sustainable food systems. A report by the
high level panel of experts on food security and nutrition of the committee on world food
security, Rome 2014. Retrieved from: http://www.iufost.org/iufostftp/FLW-%20FAO.pdf
FAO (2014b) Food security and nutrition in the Near East and North Africa. In: Regional
conference for the near east fact sheet (NERC-32). Retrieved from: http://www.fao.org/3/a-
as214e.pdf
FAO (2014c) Reducing food loss and waste in the Near East and North Africa. In: Fact sheet,
regional conference for the Near East (NERC-32) http://www.fao.org/3/a-as212e.pdf
FAO (2014d) State of food and agriculture in the Near East and North Africa Region. In: FAO
regional conference for the Near East (NERC 14/4), 24–28 Feb 2014, Rome, Italy. Retrieved
from: http://www.fao.org/docrep/meeting/030/mj390e.pdf
FAO (2014e) FAO statistical yearbook 2014: Near East and North Africa food and agriculture. In:
Food and agriculture organization of the United Nations regional office for the Near East and
North Africa, Cairo, 2014. Retrieved from: http://www.fao.org/docrep/019/i3591e/i3591e.pdf
FAO (2014f) Tackling food loss and waste in the Near East and North Africa. Food and
Agriculture Organization of the United Nations Regional Office for the Near East and North
Africa, Cairo, 2014. Retrieved from: http://www.fao.org/neareast/perspectives/food-waste/en/
Fath H, Sadik A, Mezher T (2013) Present and future trend in the production and energy
consumption of desalinated water in GCC countries. Int J Therm Env Eng 5(2):155–162. http://
www.iasks.org/sites/default/files/ijtee201305020155165.pdf
Fattouh B, El-Katiri L (2012) Energy subsidies in the Arab World. Arab human development
report research paper series UNDP, Oxford, UK. Retrieved from: http://www.arab-hdr.org/
publications/other/ahdrps/Energy%20Subsidies-Bassam%20Fattouh-Final.pdf
Faucon B, Albayrak A (2014) Islamic State funds push into Syria and Iraq with labyrinthine
oil-smuggling operation: west pushes for crackdown as Jihadists move fuel with pipeline, truck
mule. Wall Street J, 16 Sep 2014. Retrieved from: http://online.wsj.com/articles/islamic-state-
funds-push-into-syria-and-iraq-with-labyrinthine-oil-smuggling-operation-1410826325
Food security in the Gulf: how to keep stomachs full (2014) The Economist, 22 Feb 2014.
Retrieved from: http://www.economist.com/news/middle-east-and-africa/21596978-gulf-arabs-
are-debating-how-best-feed-themselves-how-keep-stomachs-full
Fuel crisis in northern Iraq revives market in smuggles gasoline. (2014) Middle East Monitor, 11
July 2014. https://www.middleeastmonitor.com/news/middle-east/12708-fuel-crisis-in-northern-
iraq-revives-market-in-smuggling-gasoline
170 P. Sullivan

Gerbens-Leenes PW, Mekonnen MM, Hoekstra AY (2013) The water footprint of poultry, pork
and beef: a comparative study in different countries and production systems. Water Resour Ind
1–2(2013):25–36 http://waterfootprint.org/media/downloads/Gerbens-et-al-2013-waterfoot
print-poultry-pork-beef.pdf
Gereffi G, Ahmed G (2014) Wheat global value chains, global corporations and food security in
MENA. Center on globalization, Governance & competitiveness, Duke University,
Harvard IGLP Workshop, Doha, Qatar, 7 Jan 2014. Retreived from: http://sites.duke.edu/
minerva/files/2013/08/2014-01-06_IGLP_Wheat-GVC-in-MENA_Gereffi-Ahmed_WEBSITE.
pdf
Global Petrol Prices (n.d.) Gasoline prices, liter. http://www.globalpetrolprices.com/gasoline_
prices/. Accessed 10 May 2015
Griffiths-Sattenspiel B, Wilson W (2009) The carbon footprint of water. River Network, Portland,
Oregon. Retrieved from: http://www.rivernetwork.org/resource-library/carbon-footprint-water
Hajjar M (2014) Addressing the root causes of food inflation in the MENA. Farrelly and Mitchell
Food & Agri-Business Specialists, Sep 2014. Retrieved from: http://farrellymitchell.com/wp-
content/uploads/2014/08/Insights-September-2014.pdf
Hajjar M, Sweeney M (2014) Opportunities for supply chain consolidation in GCC food sector.
Farrelly and Mitchell food and agri-business specialists, Nov 2014. Retrieved from: http://
farrellymitchell.com/wp-content/uploads/2014/11/GCC-Food-Report-Farrelly-Mitchell-Nov-
2014.pdf
Hashim AS (2006) Insurgency and counter-insurgency in Iraq. Cornell University Press, Ithaca
512
Heffez A (2013) How Yemen chewed itself dry: farming qat, wasting water. Foreign Affairs
Snapshot, 23 July 2013. http://www.foreignaffairs.com/articles/139596/adam-heffez/how-
yemen-chewed-itself-dry
Herbert M (2014) Partisans, profiteers and criminals: Syria’s illicit economy. Fletcher Forum Int
Aff 38(1):69–86. Retrieved from: http://www.fletcherforum.org/wp-content/uploads/2014/04/
38-1_Herbert1.pdf
Hoekstra A (2010) The water footprint: water in the supply chain. Environmentalist 1(93):12.
Retrieved from: http://waterfootprint.org/media/downloads/Hoekstra-2010-The
Environmentalist_01March_Issue93.pdf
Hoekstra AY (2011) Understanding the water footprint of factory farming. Retrieved from: http://
waterfootprint.org/media/downloads/Hoekstra-2011-FarmAnimalVoice_1.pdf
Hoekstra A (2012) The hidden water resource use behind meat and dairy. Anim Front 2(2):3–8.
Retrieved from: http://waterfootprint.org/media/downloads/Hoekstra-2012-Water-Meat-Dairy.
pdf
Hoekstra AY (2014) Water for animal products: a blind spot in water policy. Environ Res Lett 9
(9):091003 Retrieved from: http://waterfootprint.org/media/downloads/Hoekstra-2014-Water-
for-animal-products.pdf
Hoekstra AY, Mekonnen MM (2012) The water footprint of humanity. PNAS 109(9): 3232–3237.
Retrieved from: http://waterfootprint.org/media/downloads/Hoekstra-Mekonnen-2012-
WaterFootprint-of-Humanity.pdf
Hoekstra A, Wiedmann TO (2014) Humanity’s unsustainable water footprint. Science 344
(6188):1114–1117. Retrieved from: http://www.waterfootprint.org/Reports/Hoekstra-
Wiedmann-2014-EnvironmentalFootprint.pdf
IEA (2011) IEA analysis of fossil-fuel subsidies. International Energy Agency, World Energy
Outlook Paris, 4 Oct 2011. Retrieved from: http://www.iea.org/media/weowebsite/energy
subsidies/ff_subsidies_slides.pdf
IEA (2012) Water for energy: is energy becoming a thirstier resource? Excerpt from the World
Energy Outlook 2012. Retrieved from: http://www.worldenergyoutlook.org/media/
weowebsite/2012/WEO_2012_Water_Excerpt.pdf
IEA (2013) Sankey diagrams. Retrieved from: http://www.iea.org/Sankey/
IMF (2013) Energy subsidy reform: lessons and implications. IMF, 28 Jan 2013. Retrieved from:
http://www.imf.org/external/np/pp/eng/2013/012813.pdf
8 Water-Energy-Food Security Nexus in the Arab Region … 171

IMF (2014) Energy subsidies in the Middle East and North Africa: lessons for reform. IMF, Mar
2014. Retrieved from: https://www.imf.org/external/np/fad/subsidies/pdf/menanote.pdf
International Diabetes Foundation (2014) IDF diabetes atlas 2014. Retrieved from: http://www.idf.
org/sites/default/files/DA-regional-factsheets-2014_FINAL.pdf
Joint Research Center (2013) Trends in global CO2 emissions: 2013 report. PBL Netherlands
Environmental Assessment Agency, The Hague. Retrieved from: http://edgar.jrc.ec.europa.eu/
news_docs/pbl-2013-trends-in-global-co2-emissions-2013-report-1148.pdf
Joint Research Center (2014) GHG (CO2, CH4, N2O, F-gases) emission time series 1990–2012 per
region/country. Emission Database for Global Atmospheric Research. Retrieved from: http://
edgar.jrc.ec.europa.eu/overview.php?v=GHGts1990-2012
Jones R (2014) Diabetes epidemic hits Persian Gulf region. Wall Street J, 10 Feb 2014. Retrieved
from: http://www.wsj.com/news/articles/SB10001424052702304773104579268223173652920
Kerr S (2014) Diabetes burdens Middle East health systems. Financial Times, 14 Nov 2014.
Retrieved from: http://www.ft.com/cms/s/0/7685b8f0-6be5-11e4-b1e6-00144feabdc0.
html#axzz3Nft2Xkqa
Kramer M (2014) Water scarcity and human security in Yemen. Middle East Institute, National
University of Singapore, 12 Mar 2014. Retrieved from: https://mei.nus.edu.sg/index.php/web/
publications_TMPL/Perspectives-2-Water-Scarcity-and-Human-Security-in-Yemen
Laamrani H (2014) Nexus water energy food security in Arab Countries: a policy perspective and
the way forward. OCP Policy Center Conference Series, 11–13 June 2014. Retrieved from:
http://www.ocppc.ma/ckfinder/userfiles/files/4%20-%20Rabat%20Conf%20Nexus%20June%
2013%20%202014%281%29.pdf
Lund A (2014) Cold winter coming: Syria’s fuel crisis. Carnegie Endowment for International
Peace, 13 Monday 2014. Retrieved from: http://carnegieendowment.org/syriaincrisis/?fa=
56917
Meilke E, Anadon LD, Narayanamurti V (2010) Water consumption of energy resource extraction,
processing and conversion. Belfer Center for Science and International Affairs, Harvard
Kennedy School. Retrieved from: http://belfercenter.ksg.harvard.edu/files/ETIP-DP-2010-15-
final-4.pdf
Mekonnen MM, Hoekstra AY (2010a) The green, blue and grey water footprint of crops and
derived crop products volume 1: main report. Value of Water Research Report Series No 47,
UNESCO-IHE, Delft, The Netherlands http://waterfootprint.org/media/downloads/Report47-
WaterFootprintCrops-Vol1.pdf
Mekonnen MM, Hoekstra AY (2010b) The green, blue and grey water footprint of crops and
derived crop products volume 2: appendices. Value of Water Research Report Series No 47,
UNESCO-IHE, Delft, The Netherlands http://waterfootprint.org/media/downloads/Report47-
WaterFootprintCrops-Vol2.pdf
Mekonnen MM, Hoekstra AY (2011a) National water footprint accounts: the green, blue and grey
water footprint of production and consumption volume 1: main report. Value of Water
Research Report Series No. 50, UNESCO-IHE, Delft, The Netherlands http://waterfootprint.
org/media/downloads/Report50-NationalWaterFootprints-Vol1.pdf
Mekonnen MM, Hoekstra AY (2011b) National water footprint accounts: the green, blue and grey
water footprint of production and consumption volume 2: appendices. Value of Water
Research Report Series No. 50, UNESCO-IHE, Delft, The Netherlands. http://waterfootprint.
org/media/downloads/Report50-NationalWaterFootprints-Vol2.pdf
Mekonnen MM, Hoekstra AY (2011c) The water footprint of electricity from hydropower. Value
of Water Research Report Series No. 51, UNESCO-IHE, Delft, The Netherlands http://
waterfootprint.org/media/downloads/Report51-WaterFootprintHydropower.pdf
Mekonnen MM, Hoekstra AY (2012) A global assessment of the water footprint of farm animal
products. Ecosystems 15:401–415
Meltzer J, Hultman N, Langley C (2014) Low carbon energy transitions in Qatar and Gulf
cooperation council region. Global economy and development program, Feb 2014, Brookings.
Retrieved from: http://www.brookings.edu/*/media/Research/Files/Reports/2014/03/low%
172 P. Sullivan

20carbon%20energy%20transitions%20qatar%20meltzer%20hultman/00%20low%20carbon%
20energy%20transitions%20qatar%20meltzer%20hultman%20executive%20summary.pdf
Mehrtash H (2014) Obesity in the Middle East. Global Health Middle East, 2 Apr 2014. Retrieved
from: http://www.globalhealthmiddleeast.com/obesity-in-the-middle-east/
Naja F (2014) Towards a food secure Arab World, assessment as a start? Retrieved from: https://
kim.uni-hohenheim.de/fileadmin/einrichtungen/fsc/FSC_in_Dialog/previous_FSC_in_dialog/
2014/23-1-2014_Naja.pdf
Neuhauser A (2014) Funding the Islamic State. US News and World Report, 11 Sep 2014.
Retrieved from: http://www.usnews.com/news/articles/2014/09/11/funding-the-islamic-state
Sdralevich C, Sab R, Zouhar Y, Albertin G (2014) Subsidy reform in the Middle East and North
Africa: recent progress and challenges ahead. IMF. Retrieved from: http://www.imf.org/
external/pubs/ft/dp/2014/1403mcd.pdf
Spiess A (2011) Food security in the Gulf cooperation council (GCC) economies (Working Paper).
Hamburg: NDRD. Retrieved from: http://www.ndrd.org/Spiess_-_Working_Paper_on_Food_
Security_in_the_GCC.pdf
Sullivan P (2012) Water, food, energy, qat and conflict: Yemen. Future Directions International,
23 Apr 2012. Retrieved from: http://futuredirections.org.au/wp-content/uploads/2012/04/FDI_
Associate_Paper_-_23_April_2012.pdf
Sullivan P (2013) A mutually dependent relationship. About Oil Newsletter, 23 Sep 2013, pp 24–
28. Retrieved from: http://www.abo.net/en_IT/flip-tabloid/oil_23_en/index.html#/24/
Sullivan P (2014) The energy-insurgency revolution nexus: an introduction to issues and policy
options. J Int Aff 68(1):35–63
Turkey seizes smuggled Syrian fuel worth 6.7 mln liras in 45 days (2014) Hurriyet Daily News, 21
July 2014. http://www.hurriyetdailynews.com/turkey-seizes-smuggled-syrian-fuel-worth-67-
mln-liras-in-45-days.aspx?pageID=238&nID=69142&NewsCatID=341
United Nations Millennium Goals Indicators (2013) Carbon dioxide missions (CO2), metric tons
of CO2 per capita (CDIAC). Accessed July 2013. http://mdgs.un.org/unsd/mdg/SeriesDetail.
aspx?srid=751
UNDP (2012) Arab human development report. Research Paper Series Energy Subsidies in the
Arab World. http://www.arab-hdr.org/publications/other/ahdrps/Energy%20Subsidies-Bassam
%20Fattouh-Final.pdf
UNDP (2013) Water governance in the Arab Region: managing scarcity and securing the future.
UNDP, New York. Retrieved from: http://www.arabstates.undp.org/content/dam/rbas/doc/
Energy%20and%20Environment/Arab_Water_Gov_Report/Arab_Water_Gov_Report_Full_
Final_Nov_27.pdf
Union of Concerned Scientists (2014) Each country’s share of CO2 emissions, 18 Nov 2014.
Retrieved from: http://www.ucsusa.org/global_warming/science_and_impacts/science/each-
countrys-share-of-co2.html#.VKa93SvF-So
US official: ‘Islamic State’ a petrostate (2014) Deutsche Welle, 23 Oct 2014. Retrieved from:
http://www.dw.de/us-official-islamic-state-a-petrostate/a-18018010
Waterfootprint.org, (2014) “Links: Carbon Footprint”, http://www.waterfootprint.org/?page=files/
LinksCarbonFootprints
Water Footprint Network (2014) Product water footprint: what is the water footprint of a product?
Water Footprint Network. Retrieved from: http://www.waterfootprint.org/?page=files/Water-
energy
World Bank (2011) Water reuse in the Arab World: from principle to practice. A summary of
proceedings expert consultation wastewater management in the Arab world, 22–24 May 2011,
Dubai, UAE http://water.worldbank.org/sites/water.worldbank.org/files/publication/Water-
Reuse-Arab-World-From-Principle%20-Practice.pdf
World Bank (2012) Renewable energy desalination: an emerging solution to close the water gap in
the Middle East and North Africa. World Bank, Washington, DC. http://water.worldbank.org/
sites/water.worldbank.org/files/publication/water-wpp-Sun-Powered-Desal-Gateway-Meeting-
MENAs-Water-Needs_2.pdf
8 Water-Energy-Food Security Nexus in the Arab Region … 173

World Bank (2014) Corrosive subsidies. World Bank Middle East and North Africa region MENA
economic monitor, Washington, DC. Retrieved from: http://www-wds.worldbank.org/external/
default/WDSContentServer/WDSP/IB/2014/10/08/000350881_20141008084801/Rendered/
PDF/912100WP0Box380RSION0OCTOBER0402014.pdf
World Food Program (2014a) Sudan—food security monitoring. Retrieved from: https://www.
wfp.org/content/sudan-food-security-monitoring
World Food Program (2014b) Syria emergency. Retrieved from: http://www.wfp.org/emergencies/
syria
World Food Program (2014c) Comprehensive food security survey: Yemen. World Food
Programme, Nov 2014. Retrieved from: https://www.wfp.org/content/yemen-comprehensive-
food-security-survey-november-2014
Chapter 9
Managing Water, Energy, and Food
for Long-Term Regional Security

Zafar Adeel

Abstract This chapter presents a contemporary and robust definition of regional


security that encompasses flow of resources, sustainable economic development,
poverty reduction, and peaceful co-existence. Regional integration and political
stability are key ingredients for achieving regional security. A great level of
diversity in human security, quantified by using the Human Development Index as
a surrogate, persists in the region. The chapter focuses on the role played by water,
food, and energy in regional security; it presents some inter-related drivers of
change that impinge on regional security: the burgeoning population with a sig-
nificant ‘youth bulge’ and accompanying widespread youth unemployment; the
economic impacts as a result of globalization, particularly in food and energy
sectors; the rise in extremist ideologies and their intersection with efforts to enhance
democratic processes; and, geopolitical tussles that are often aimed at greater
control of the region’s various resources. A major factor in the regional insecurity is
the lack of adequate environmental management, resulting from poor environ-
mental governance; the environmental management gaps are also tied to capacity
gaps in human, technological, and institutional resources. The pros and cons are
discussed for a number of approaches for sustaining regional water, energy, and
food security; these include agricultural land acquisition in Africa, increasing focus
on smallholder rain-fed agriculture, mastering the water-energy-food nexus
including renewable energy sources, and enabling a favorable policy environment.
It is concluded that in order to convince policymakers and governments in the Arab
Region that the Water-Energy-Food Nexus (WEF Nexus) is central to regional
security, supporting arguments must be presented in quantifiable economic and
social terms.

 
Keywords Regional security Drivers of insecurity Water-Energy-Food nexus 

Land acquisition Smallholder agriculture

Z. Adeel (&)
Pacific Water Research Centre, Simon Fraser University, Burnaby, British Columbia, Canada
e-mail: zadeel@sfu.ca

© Springer International Publishing AG 2017 175


K. Amer et al. (eds.), The Water, Energy, and Food Security Nexus
in the Arab Region, Water Security in a New World,
DOI 10.1007/978-3-319-48408-2_9
176 Z. Adeel

9.1 An Overview of Regional Security Concepts

Conventional definitions of regional security focus on manifest and measurable


interactions between states, and often focus on threats, whether perceived or real,
between states (Lake 1997). In this approach, security is often viewed through the
lens of military apparatus and capabilities, which presumably allows states to
achieve a strong negotiating position on matters of regional interest and national
security (Adeel 2012). In the case of overt armed conflict, military capabilities are
typically combined with international diplomacy to achieve strategic objectives.
However, one may argue that investments into militarization of security often
compete with and divert vital resources away from development activities, which
could play a greater role in achieving regional security. Thus a different and more
inclusive concept of regional security has to be sought, one that incorporates
economy and livelihoods, politics and policies, resources and people, and human
security (Werz and Hoffman 2013).
The notion of human security, quite central in this context, is characterized using
the UNDP definition: “Human security can be said to have two main aspects. It
means, first, safety from such chronic threats as hunger, disease and repressions.
And second, it means protection from sudden and hurtful disruptions in the patterns
of daily life—whether in homes, in jobs, or in communities. Such threats can exist
at all levels of national income and development” (UNDP 1994). This concept of
human security is central to human survival and quality of life, and thus, invariably
linked to economic growth and poverty eradication, freedom and human rights, and
equity and justice (Brauch 2005).
It is, therefore, safe to conclude that regional security is a broad concept that
surpasses but includes the conventional military- and conflict-based securitization.
Put differently, regional security cannot be solely tied to military expenditure or
investments into maintaining law and order (Zou’bi 2014). A contemporary and
robust definition of regional security would encompass flow of resources, sus-
tainable economic development, political stability, poverty reduction, and peaceful
co-existence.

9.1.1 Determinants of Regional Security

Regional Integration: With the increased attention on linkage of economies and


related development processes across national boundaries, the notion of regional
cooperation has emerged and gained strength as a nested construct within the global
political and development agenda. This increased interest in regionalization of
issues is based on the premise that distribution of resources, volumes of trade and
commerce, and political stability all play a role in ensuring regional security.
9 Managing Water, Energy, and Food for Long-Term Regional Security 177

The rate (and ease) of flow of resources—including manpower, capital, and natural
resources—within a region determines the rate of economic growth, and eventually,
creation of jobs and reduction of unemployment rates (AMF 2014). Increasingly,
regional trade has been a significant factor in determining economic growth rates;
the European Union is often cited as a success story, although the 2008–2009
economic crisis has put the strength of this regional arrangement under a severe test.
Political Stability: Similarly, political stability is also a key determinant of
regional security—this construct of stability surpasses national boundaries, and the
knock-on impacts from political turmoil can hamper development at regional scale.
A key example is South Asia, in which political instability in Afghanistan, growth
of terrorism in Pakistan, internal freedom movements in India, and the Maoist
rebellion in Nepal have chronically impaired the region’s ability to achieve its full
economic and development potential. Foreign Direct Investment (FDI), a key
ingredient for economic growth and stability, tends to shy away from regions and
countries in which poor governance and a state of insecurity increases risks for
investors (Burnside and Dollar 2000). The somewhat stagnated economic growth
and rampant poverty in South Asia is often contrasted with its neighboring
Southeast Asian region, in which Viet Nam, Singapore, Republic of Korea, and
Malaysia are examples of robust and rapid economic development coupled with
political stability (Cho et al. 2014).

9.1.2 The Economic Terrain in the Arab Region

In terms of economic and social development indicators, there is a very broad range
of values observable throughout the region, in contrast to the common perception of
it as a cohesive and monolithic region. As shown in Table 9.1, the Gross Domestic
Product (GDP) per capita varies by about two orders of magnitude between the least
and most prosperous countries (World Bank 2015). This variability is also observed
in the Human Development Index (HDI), in which Bahrain, Kuwait, Qatar, Saudi
Arabia, and United Arab Emirates demonstrate a ‘very high human development’
(UNDP 2015). In contrast, Djibouti, Mauritania, Sudan, and Yemen have a ‘low
human development,’ with other countries falling in between. This broad range
inter alia also underpins fairly different natural resource endowment in the Arab
Region, as discussed in Sect. 9.2.1. Using HDI as a surrogate for human security,
we can conclude that the region’s level of human security is not uniform either. The
challenge, therefore, is to improve the average level of security in the region, which
would require improvements in human security in the countries where HDI levels
are quite low. Learning from the European Union, there has to be a concerted effort
to reduce intra-regional disparities in the Arab Region.
178 Z. Adeel

Table 9.1 A decade-based comparison of Gross Domestic Product (GDP) per capita in the Arab
Region, 2005 and 2014; all figures in current US$
Country 2005 GDP/cap 2014 GDP/cap HDI 2014 (rank)
Qatar 53,207 96,732 0.850 (32)
United Arab Emirates 40,299 43,963 0.835 (41)
Kuwait 35,694 43,594 0.816 (48)
Bahrain 18,418 24,855 0.824 (45)
Saudi Arabia 13,274 24,161 0.837 (39)
Oman 12,399 19,310 0.793 (52)
Lebanon 5339 10,058 0.769 (67)
Libya 8159 6573 0.724 (94)
Iraq 1849 6420 0.654 (121)
Algeria 3102 5484 0.736 (83)
Jordan 2326 5423 0.748 (80)
Tunisia 3218 4421 0.721 (96)
Egypt 1197 3199 0.690 (108)
Morocco 1932 3190 0.628 (126)
Sudan 662 1876 0.479 (167)
Djibouti 910 1814 0.470 (168)
Mauritania 693 1275 0.506 (156)
Syria 1592 N/A 0.594 (134)
Yemen 817 1408a 0.498 (160)
Arab World 3763 7386 0.686
Source World Bank (2015)
Human Development Index (HDI) data from the Human Development Report (UNDP 2015)
a
2013 data

9.2 The Context for Security in the Arab Region

The history of Arab civilization is rich and spans many centuries and millennia;
many comprehensive treatises explore and investigate this rich history. Awareness
of this historical context is central to fully understanding the present-day turmoil
and conflicts prevalent throughout the region. This chapter does not intend to
replicate that historical analysis, and instead guides readers to published works by
other established authors and historians (Lewis 1996, 2002; Mousalli 2014; Zou’bi
2014).
However, it is useful to understand economic, social, and political developments
within the past few decades—particularly the developments and transitions that
have a direct bearing on real or perceived regional security issues. In the context of
this chapter’s focus on water, food, and energy in regional security, some
inter-related elements are of great interest: the burgeoning population bulge with
widespread youth unemployment; the economic impacts as a result of globalization,
particularly in food and energy sectors; the rise in extremist ideologies and its
9 Managing Water, Energy, and Food for Long-Term Regional Security 179

intersection with efforts to enhance democratic processes; and, geopolitical tussles


that are often aimed at greater control of the region’s various resources. The picture
becomes complex when one starts to consider geopolitical interests of those
countries outside and immediately neighboring the Arab Region—countries like
Turkey, Iran, and Israel, which have all vested interests in supporting domination of
competing political and religious ideologies in the Arab Region. That picture
becomes immensely more complex with other, more peripheral players such as
Russia, United States, European Union, etc., also becoming stakeholders. There
also less obvious external players like OPEC group of countries, including China
and Canada, that have an indirect interest in ensuring that the geopolitical devel-
opments in the Arab Region do not undermine their own strategic or economic
interests.
Let us briefly analyze each of these elements.

9.2.1 Population Growth and Resource Distribution

The population boom in the Arab Region has been phenomenal in the last three
decades. As shown in Table 9.2, the total population has grown from about
200 million in 1990 to about 340 million in 2013. Even more remarkable is
long-term population growth trends in some of the smaller countries in the region:
Qatar grew from 400,000 to nearly 2 million people, and United Arab Emirates
grew from 1.7 to 8.5 million in the same time period. This population growth has
the most obvious consequences of reduced resources available per capita and
increased population density per square km. The former is discussed elsewhere in
this volume in terms of water availability per capita, including forecasts based on
population trends and climate change. The result of the latter is some of the densest
populations in the world in Bahrain and Lebanon. The annual average population
growth rate for the region is between 2.3 and 2.4 %, based on 2013 and
decade-long figures, respectively. It is challenging to maintain a resource supply
with such population growth rates, resulting in perceivable impacts on regional
security. As discussed in Sect. 9.2.2, the average population growth rates also mask
some challenging demographics and related drivers of regional insecurity.

9.2.2 Demographics and Unemployment

Like many developing regions, the Arab Region has a prominent ‘youth bulge’—an
increasing population of young people between the ages of 15–25 (LaGraffe 2012).
This youth bulge is perhaps much more prominent here compared to almost any
other region in the world; data presented in Fig. 9.1 show an average of 60 % of the
180 Z. Adeel

Table 9.2 Population statistics for the Arab Region


Population Annual growth Land area Population HDI
(thousands) Rate % (km2) density 2014
(person/km2)
1990 2013 2013 2000– 1990 2013
2010
Qatar 418 1830 2.2 10.82 11,607 37 158 0.850
United Arab 1773 8533 1.07 10.68 83,600 21 102 0.835
Emirates
Kuwait 2142 3965 3.71 4.91 17,818 120 186 0.816
Bahrain 503 1255 0.76 6.77 707 711 1776 0.824
Saudi Arabia 15,187 29,994 2.73 3.02 2,000,000 7 15 0.837
Oman 1625 3855 6.4 1.45 309,500 5 12 0.793
Lebanon 2550 4096 0.64 0.65 10,452 245 392 0.769
Libya 4229 8,554 3.24 3.26 1,775,500 2 5 0.724
Iraq 17,890 35,087 2.61 3.04 435,052 41 81 0.654
Algeria 25,022 38,229 2.04 1.69 2,381,741 11 16 0.736
Jordan 3468 6530 2.22 2.33 89,342 39 73 0.748
Tunisia 8154 10,893 1.07 0.99 155,566 52 70 0.721
Egypt 51,911 84,629 2.52 2.2 1,009,450 52 84 0.690
Morocco 24,167 32,954 1.1 1.14 710,850 34 46 0.628
Sudana 23,436 36,164 3.16 2.98 1,882,000 9 19 0.479
Djibouti 520 1011 3.1 3.1 23,200 22 44 0.470
Mauritania 1980 3612 2.43 2.43 1,030,700 2 4 0.506
Syria 12,116 21,768 1.83 2.37 185,180 65 118 0.594
Yemen 12,860 25,244 2.92 2.86 555,000 23 45 0.498
Arab World 209,951 342,387 2.33 2.44 0.686
Source AMF (2014); UNDP (2015)
a
Land Area of Sudan after the South’s cession in July 2011

population under the age of 30, which is the second highest in the world after
sub-Saharan Africa (Hoffman and Jamal, 2012). Many leading researchers have
written extensively about the social and economic consequences of this demo-
graphic growth and that it acts as a “force multiplier” for root causes of conflict and
instability, such as unemployment, poverty, urbanization trends, and declining
economic growth (LaGraffe 2012). The governments of the region need to heed
calls for economic reforms that would accelerate economic growth and create more
jobs.
The unemployment trends vary throughout the Arab Region, ranging between
estimates of 13 % (Raphaeli 2006) and 17 % (AMF 2014). This represents the
highest unemployment rate in the world, even ahead of the sub-Saharan Africa
region which is otherwise the poorest in the world. Based on a comparison between
1995 and 2005, there is no significant downward trend in unemployment rates
9 Managing Water, Energy, and Food for Long-Term Regional Security 181

Fig. 9.1 Percentage of population under the age of 30 years in select Arab countries (adapted
from Hoffman and Jamal 2012)

suggesting that employment growth is not keeping pace with the population growth
(Raphaeli 2006). These unemployment rates have a relatively greater impact on
youth, which in turn can and has become the leading cause of volatility in the Arab
society.
The so-called Arab Spring that started in 2012 is a remarkable example of the
social, political, and economic unrest that can result from a sense of disenfran-
chisement amongst youth. It is understood that the level of frustration of the youth
with the political oppression in the Arab Region runs high (Hoffman and Jamal
2012). What started as a lone protest in Tunisia, spread swiftly through a number of
Arab countries: Syria, Egypt, Yemen, Jordan, Algeria, Iraq, and so on. Some
governments like the one in Jordan brought about swift transitions in governance
structures, as well as changes to the political leadership, which reduced tensions and
minimized adverse societal impacts. The other end of the extreme is Syria, where
political unrest triggered an all out civil war that rages on without any signs of
abatement. The Syrian situation not only has greatly impacted the population and
economy of the country, its spillover effects in terms of armed conflict and refugees
has also engulfed a significant part of Iraq. The Syrian refugees are faced with a
tragic humanitarian crisis and can inadvertently further exacerbate the
socio-political challenges in host countries like Lebanon and Jordan, where refugee
influx constitutes a 25 and 10 % increase in the host population, respectively
(Balsari et al. 2015).
It is clear that regional stability and security depends significantly on how well
these broader economic and social challenges are addressed. Because the unrest is
not always contained within national boarders, regional efforts to improve
182 Z. Adeel

economic growth and lower unemployment can have knock-on benefits for
everyone. It remains to be seen whether such narrative around employment and
economy goes beyond lip service and results in significant job growth.

9.2.3 Impacts of Globalization on Food and Energy Sectors

One may argue that a sharp rise in food prices, combined the long-term droughts
and crop failures, was the trigger for unrest in many countries in the Arab Region—
most notably in Tunisia and Syria. The food security equation very directly inter-
sects the water availability, as is discussed in detail elsewhere in this volume. While
the green revolution of the 1970s led to increased food production worldwide, the
Arab Region was no different, there has been a consistent decline in food pro-
duction, and countries in the region have increasingly relied on food imports; at this
time the agricultural value contribution to the GDP stands at 5 % for the Arab
Region (AMF 2014). This increases the reliance on world food market and results
in a heightened sense of insecurity; if the present situation with population growth
and stagnant agricultural production continues, the “food gap” will grow to be
about US$ 96 billion by 2030 (AMF 2014). Some Arab countries have responded
by increasing their focus on self-sufficiency in food production and agricultural
output. While this intensification of agriculture may sound superficially attractive
and politically expedient, it increases threats to the water resources of the region
and risks a total or near-total collapse of the water systems. Not only does food
security link in a fairly direct way to regional security, if not handled and managed
properly it also can become a major factor in destabilizing the water economy.
Similarly, fluctuating fortunes in the energy sector can also result in instability
and insecurity. Even with new reserves of oil and natural gas discovered in recent
years, the Arab Region’s shares of the world’s reserves have been falling; this can
be contrasted by about 5 % increase in the energy demand in the region (AMF
2014). To a great degree many of the oil-producing Arab countries are dependent
on traditional petro-based economic activities as a significant component of their
national economy. Ostensibly, wide fluctuations in the international crude oil pri-
ces, as shown in Fig. 9.2, can play havoc with national economies. This volatility,
coupled with the recent steep decline in oil prices has led the Arab member
countries of OPEC to dip into their national reserves, limiting the capital available
for development activities and lending (AMF 2014).

9.2.4 Extremism and Democracy

A number of research papers and dissertations explore in depth the roots of


extremism in the Middle East region, and tie those to lack of democracy. Many
have argued that the absence of democracy at national and subnational scales, when
9 Managing Water, Energy, and Food for Long-Term Regional Security 183

Fig. 9.2 Ten-year fluctuation in crude oil prices (Crude Oil WTI) (source Nasdaq 2015)

combined with economic problems and social injustice, become the rallying
argument for those who see extremism as a way of breaking out of this ostensibly
closed loop of neglect and misery (Pillar 2011). The objective in this chapter is not
to undertake an exhaustive analysis of these aspects. Nonetheless, it is obvious from
observing published research and reading newspaper headlines that the real and
perceived regional security is seen in the context of extremism and terrorism. The
rise in 2014 of the so-called Islamic State in Iraq and Syria (ISIS) has clearly shown
that extremist threats are not always bound by national and international borders.
The ‘experiment’ in Libya over the past five years also demonstrates that a
forceful (and armed) solution to modify the governance mechanism in the absence
of other mitigating factors would likely not yield positive outcomes. Conversely,
gradual and slow societal transitions may not keep pace with the fast-acting drivers
of change discussed in this chapter. It is not so far-fetched to argue that in the
absence of drastic changes to the governance paradigm and evolution of democratic
institutions, there is little that stands in the way of growing extremism (Zou’bi
2014). Achieving regional security requires positive solutions to economic, social,
and political challenges, which can likely overcome extremist ideologies in the long
run.
184 Z. Adeel

9.3 The Environmental Dimensions of Security


in the Arab Region

9.3.1 Climate Change as a Driver of Regional Security

Many researchers have linked environmental factors to human security, which as


we discussed earlier, ties directly into regional security. A simple causal structure is
often presented: changes to the water cycle as a result of climate change and other
anthropogenic factors result in water scarcity, which in turn has an impact on food
security and could trigger conflicts (Kloos et al. 2013). Other environmental haz-
ards and extreme climate events such as floods or prolonged droughts could also be
triggers for insecurity. However, the underlying causality is typically more com-
plicated and it is quite difficult to divide the causes into neat categories of envi-
ronmental and non-environmental factors. A fundamental mismatch between the
natural resource endowment and societal demands is the primary trigger. Syria is a
case in point: Gleick (2014) argues that challenges associated with climate change
and the impact of water scarcity on food production have contributed to the present
conflict. At the same time, he suggests a number of other factors also contributed as
triggers including the state of national economy, the wave of political reform
triggered by the so-called Arab Spring, and “a broad set of religious and
sociopolitical factors.”
Johnstone and Mazo (2013) argue that “the Arab Spring would likely have come
one way or another, but the context in which in it did is not inconsequential. Global
warming may not have caused the Arab Spring, but it may have made it come
earlier.”
The most significant and obvious resource challenge is that of water scarcity,
which is forecasted to get worse as the region becomes hotter and drier a result of
climate change impacts (IPCC 2007). As is discussed in other chapters in this
volume, freshwater availability is significantly constrained in almost all the coun-
tries in the Arab Region—leading to obvious challenges in drinking water supply,
water for irrigation and food production, and water for industrial activity. The
economic and social impacts of this water scarcity are far-reaching, and yet there
are very few examples of water management at the regional scale. The Jordan river
is a case in point in which Syria, Jordan, and Israel collectively divert about 95 %
of its flow for agricultural, industrial, and urban purposes (Zou’bi 2014). There is an
ongoing plan to divert water from the Red Sea to the Dead Sea via a 180-km long
canal, en route generating energy and providing freshwater through desalination
(GSI 2011); the throughput of saline water would help restore the Dead Sea levels
that have been declining because of water diversions in the Jordan river basin.
Although this ambitious project is still short of the funds required to fully imple-
ment it and a number environmental impact concerns are not fully addressed yet, it
presents an example of regional collaboration—between Israel, Jordan, and
Palestinian Authority—to collectively address larger water security issues.
9 Managing Water, Energy, and Food for Long-Term Regional Security 185

9.3.2 Gaps in Environmental Governance

A major factor in the regional insecurity is the lack of adequate environmental


management, resulting from poor environmental governance. For a number of
reasons, requisite environmental governance mechanisms—including legal and
regulatory frameworks—and related institutions have not evolved through most of
the region. The lack of awareness or a vision for environmental sustainability on
part of the politicians and policymakers signifies the inability to fully understand
the associated opportunities and risks, which further exacerbates the regional
insecurity (AFED 2009). Economic policies and their implementation for urban
growth and some industrial development has led to significantly adverse impacts on
a range of natural habitats in the region (van Lavieren et al. 2011).
The environmental management gaps are tied to capacity gaps in human,
technological, and institutional resources. In particular, environmental management
agencies often lack skilled researchers, laboratory facilities, and appropriate
financial resources (van Lavieren et al. 2011). Hamdy (this volume) discusses these
capacity gaps at length and offers options for addressing them strategically. Few
research organizations exist within the region that can undertake a broader view of
environmental management while also offering specific and targeted advice on
real-time and real-world problems. At the political level, the Council of Arab
Ministers Responsible for the Environment (CAMRE) is an institution designed as
a forum for making policy level decisions at the regional scale (van Lavieren et al.
2011). Similarly, an Environmental Affairs Committee also exists within the Gulf
Cooperation Council (GCC), which has facilitated the adoption of a unified code of
standards for environmental protection.

9.4 Approaches for Sustaining Regional Water, Energy,


and Food Security

The notion of Water-Energy-Food (WEF) Nexus is relatively new to the interna-


tional development community (Hoff 2011), and is even newer to the Arab Region.
It argues that the three elements are cross-linked as “services” essential to eco-
nomic, social, and environmental development. These three services or sectors have
a triangular relationship but obviously do not operate in isolation of other elements
in the economic domain, such as urbanization patterns, population growth and
dynamics, and climate change-related impacts. Translation of this nexus-based
thinking into active policy frameworks remains an uphill battle. Because the con-
cept is relatively new, there are very few practical examples of this available. Some
examples do exist of the sub-components of the WEF Nexus: examples include
hydro-power and irrigation schemes (water-energy nexus) and biofuel production
systems (energy-food nexus).
186 Z. Adeel

Some major hurdles must be overcome to achieve successful policy imple-


mentation of the WEF Nexus concepts. First, administrative and bureaucratic silos
in governments and institutions are difficult to integrate, particularly in the absence
of hard evidence that integration leads to efficiency in operation, and importantly,
cost-savings. Second, the lack of evidence is driven by very limited research and
empirical data available outside the realm of traditional hydropower generation
schemes. Significant new investments are needed for adequate research, monitor-
ing, and policy analysis. Despite these limitations, the countries in the Arab Region
have been testing and deploying approaches that are starting to address parts of
WEF Nexus; a few selected approached are discussed here in the context of
achieving regional security.

9.4.1 Land Acquisition in Africa to Achieve Water-Food


Security

Recognizing the limitations on availability of freshwater and arable lands, there is a


growing trend in the wealthy Arab nations to ensure water and food security by
acquiring agricultural land and associated water rights in Africa. Table 9.3 sum-
marizes the scale of this land acquisition; Egypt, Saudi Arabia, and United Arab
Emirates have all acquired large-scale agricultural land in Uganda, Tanzania, and
Sudan, respectively. Such acquisitions multiply the arable land available to a nation
—by about three times for Qatar and about six times for United Arab Emirates.
Even for a country like Egypt where agricultural productivity has been historically
high, land acquisition in Uganda amounts to an increase of about 30 % in its arable
land. The impact on the arable land of the recipient countries is variable, but in
some cases such as Egypt-in-Uganda (15 %) or Saudi Arabia-in-Tanzania (5 %),
this can be quite significant.

Table 9.3 Acquisition of lands in Africa by Arab countries


Arab country Recipient country in Land acquired % of Arable land in recipient
Africa (ha)a country
Egypt Ethiopia 20,000 0.1
Egypt Uganda 860,000 15.2
Jordan Sudan 30,000 0.1
Libya Mali 100,000 2.1
Libya Liberia 20,000 4.3
Qatar Kenya 40,000 0.8
Saudi Arabia Tanzania 500,000 5.2
Saudi Arabia Sudan 10,000 <0.1
United Arab Sudan 380,000 1.8
Emirates
Source Mo Ibrabim Foundation (2011)
a
Figures rounded to the nearest 10,000 ha
9 Managing Water, Energy, and Food for Long-Term Regional Security 187

The obvious objective of these large-scale land acquisition exercises is to secure


food supply for otherwise water-scarce Arab countries. This is typically accom-
panied by provisioning of new financial investments, in many cases FDI, in the
recipient country in Africa (Khouri et al. 2011). Those in favor of this approach
argue that it can help increase agricultural productivity in recipient African coun-
tries through introduction of newer agricultural technologies and practices, create
livelihood opportunities for rural populations, and contribute to development of
rural infrastructure. Conversely, such arrangements can impinge of the traditional
land and water tenure of communities and increase food insecurity for the local
population. There is also a question as to whether the recipient African countries
have truly evaluated the long-term consequences of these transactions on their own
economies and societies (Khouri et al. 2011); in most cases, the legislative and
regulatory frameworks for assessing such long-term impacts simply do not exist.
Similarly, investments by the Arab countries are premised on political and social
stability in the recipient African countries; history tells us that such presumptions
may not hold up in regions that are undergoing economic transitions and are
otherwise prone to political upheavals.

9.4.2 Increasing Focus on Small-Scale Agriculture

Haddad et al. (2011) argue that rain-fed agriculture, practiced extensively in the
Arab Region (up to two-thirds of the region’s cropland), can help ensure better food
security at the regional scale. They indicate that with enhanced investments in
research and extension, and by engaging the private sector through enabling poli-
cies could result in more sustainable practices by smallholder farmers. Through use
of adequate technologies, particularly aimed at increasing food productivity per
drop of water and using supplemental irrigation, smallholder farmers can become
part of the solution for the region’s food security challenges. A key obstacle in
achieving this is that smallholder farmers are typically ill-equipped to cope with the
impacts of climate change, as well as regional and international drivers of food
markets (Alwang and Norton 2011). “Safety Nets” for smallholder farmers could
help alleviate the situation by incorporating targeted cash transfers, food stamps,
food distribution, targeted sales in food markets, and food-for-work programs.

9.4.3 Mastering the Water-Energy Nexus

It has been demonstrated that water and energy are closely inter-dependent sectors,
and both underpin economic growth and social development (Schuster-Wallace
et al. 2014). Mutual optimization of both sectors can enhance energy efficiency,
reduce water pollution, minimize the costs of energy and water provisioning,
increase access to these services, and support climate change mitigation by reducing
188 Z. Adeel

greenhouse gas emissions. In the Arab Region, there has been a tendency to follow
silo-ed policy approaches when dealing with water and energy—seawater desali-
nation and hydropower generation are obvious and salient exceptions to this.
Hydropower generation is also linked in most cases to irrigation schemes and thus
to food production.
Subsidies in the energy sector have been a topic of discussion around economic
reforms in the Arab Region; the most recent estimate puts regional energy subsidies
at about US$ 287 billion, accounting for about 12 % of the region’s GDP (AMF
2014). As most of these subsidies are aimed at the petro-based industry, the
opportunities to invest in alternative energy sources remain limited. There are some
indications that there is a gradual trend for removing these subsidies (Arouri et al.
2012); however, a detailed discussion of this element of the energy sector is beyond
the scope of this chapter. Overall, there has been a rising awareness in the general
public of the climate change impacts of the energy choices made; this may translate
to more sustainable forms of energy generation (Arouri et al. 2012). One notable
example is that consumers in the Arab Region are switching to more energy effi-
cient cars.
Considerable re-thinking is still needed in the energy sector on the regional level,
particularly as electricity consumption continues to grow over 8 % per year in the
GCC countries (Alnaser and Alnaser 2011). New investments into renewable
energy, notably solar (e.g., local photo-voltaic cell production) and wind-generated
energy, are an indication that the vagaries of the petro-based energy sector (see
Sect. 9.2.3) have not gone completely unnoticed at the policy level. There is an
obvious cross-over to water management approaches, in which thermal desalination
processes being driven by solar energy are considered a promising prospect
(Doukas et al. 2011); achieving such integration, however, can only take place in a
policy environment that is conducive.

9.4.4 Creating an Enabling Policy Environment

Creation of enabling policies that address the challenges encountered by the Arab
Region at the nexus of water, food, and energy is critical. These enabling policies
have to cover the whole gamut of value chain that comprises interdependent
components (Zou’bi 2014). Consequently, a bottom-up approach of engaging all
stakeholders is essential. Creation of new partnerships that bring together a broad
range of stakeholders, including inter alia researchers, policymakers, private sector,
and nongovernmental organizations, at the regional scale has also been suggested
(Khouri et al. 2011). It can be argued that creating substantial and sustainable
on-the-ground impacts requires that such partnerships are mobilized, adequately
resourced, and supported through active policy measures—national governments
can play a key role in providing resources and policy support for these partnerships.
There is also a need for creating and enabling initiatives that go beyond the
national boundaries in order to capture larger-than-the-sum-of-parts benefits. It can
9 Managing Water, Energy, and Food for Long-Term Regional Security 189

be envisioned that regional forums such as the League of Arab States (aka. Arab
League) and the Organization for Islamic Cooperation (OIC) can play central role;
scarcity of resources has historically prevented the achievement of such regional
objectives (Zou’bi 2014). For example, a regional plan for food security was
announced by the Arab League in 2007, which has since received considerable
support but requires upwards of US$ 144 billion for its implementation (Khouri
et al. 2011). In an atmosphere of serious regional-scale geopolitical conflicts,
including the civil wars in Syria, Libya, and Yemen, mobilizing such large-scale
new investments remains a challenge.

9.5 Conclusion: Achieving Long-Term Regional Security

The Arab Region is diverse in its challenges, endowments, and the underlying
social, economic and political drivers. This fact makes generalizations around
water, food, and energy difficult and less meaningful. However, commonalities of
threats to regional security demand that mutual solutions to the water-energy-food
challenges be sought to optimize the scale of operation and to pool together
financial, manpower, and intellectual resources. Considering the political turmoil in
a significant number of Arab countries, it also makes sense that efforts to achieve
resource security are interlinked with regional security, and aim to minimize the
adverse drivers. A broader approach to regional security that centers around human
security—safety from threats and repressions, and protection against disruption to
the patterns of daily life—must be included in such a regional vision.
In order to convince policymakers and governments in the Arab Region that the
WEF Nexus is relevant and central to regional security, supporting arguments must
be presented in economic and social terms. The economic benefits (creation of jobs
and sustainability of commercial and industrial activity) and reduced risks (of cli-
mate change impacts, food insecurity, regional instability) of integrated policy-
making must be presented in a quantifiable format. It must also be clearly presented
that the enhancement of benefits and reduction of risks would only take place in a
policy environment that supports integrated planning and implementation. These
approaches also offer the possibility of looking beyond national boundaries to arrive
at truly regional perspectives on security.

References

Adeel Z (2012) Human development approach to water security. In: Bigas H (ed) The global water
crisis: addressing an urgent security issue. Papers for the Inter Action Council, UNU-INWEH,
Hamilton, Canada
AFED (2009) Arab environment: climate change, impact of climate change on Arab countries. In:
Tolba MK, Saab NW (eds) Report of the Arab forum for environment and development
(AFED). Beirut, Lebanon
190 Z. Adeel

Alnaser WE, Alnaser NW (2011) The status of renewable energy in the GCC countries. Renew
Sustain Energy Rev 15:3074–3098
Alwang J, Norton GW (2011) What types of safety nets would be most efficient and effective for
protecting small farmers and the poor against volatile food prices? Food Secur 3(Suppl 1):
S139–S148
AMF (2014) The joint Arab economic report 2014 (overview and statistical annexes). Arab
Monetary Fund, Abu Dhabi, United Arab Emirates. Accessed 31 Dec 2015 http://www.
arabmonetaryfund.org/content/joint-arab-economic-report
Arouri MEH, Youssef AB, M’henni H, Rault C (2012) Energy consumption, economic growth
and CO2 emissions in Middle East and North African countries. Energy Policy 45:342–349
Balsari S, Abisaab J, Hamill K, Leaning J (2015) Syrian refugee crisis: when aid is not enough.
The Lancet 385(9972):942–943
Brauch HG (2005) Threats, challenges, vulnerabilities and risks in environmental and human
security. SOURCE ‘Studies of the University: Research, Counsel, Education,’ publication
series of UNU-EHS No. 1/2005. United Nations University Institute for Environment and
Human Security, Bonn, Germany
Burnside C, Dollar D (2000) Aid, policies, and growth. Am Econ Rev 90:847–868
Cho C, Choi YC, Song YC (2014) Key development needs of South Asia and priority sectors of
Korean ODA. World Economic Update, Korea Institute for International Economic Policy,
Sejong-si, Korea
Doukas H, Patlitzianas KD, Kagiannas AG, Psarras J (2011) Renewable energy sources and
rationale use of energy development in the countries of GCC: Myth or reality? Renew Energy
31:755–770
Gleick PH (2014) Water, drought, climate change, and conflict in Syria. Weather Clim Soc 6
(3):331–340
GSI (2011) Red sea to dead sea water conveyance (RSDSC) study: dead sea research team.
Geological Survey of Israel (GSI) Report Number GSI/10/201
Haddad N, Duwayri M, Oweis T, Bishaw Z, Rischkowsky B, Hassan AA, Grando S (2011) The
potential of small-scale rainfed agriculture to strengthen food security in Arab countries. Food
Security 3(Suppl 1):S163–S173
Hoff H (2011) Understanding the nexus. In: Background paper for the Bonn 2011 conference: the
water, energy and food security nexus. Stockholm Environment Institute, Stockholm, Sweden
Hoffman M, Jamal A (2012) The youth and the Arab spring: cohort differences and similarities.
Middle East Law Governance 4:168–188
IPCC (2007) Climate change 2007: synthesis report. In: Pachuri RK, Reisinger A
(eds) Contribution of working groups I, II and III to the fourth assessment report of the
intergovernmental panel on climate change. IPCC, Geneva, Switzerland
Johnstone S, Mazo J (2013) Global warming and the Arab Spring. In: Werrell CE, Femia F
(eds) The Arab spring and climate change—a climate and security correlations series. Center
for American Progress, Washington DC, US
Khouri N, Shideed K, Kherallah M (2011) Food security: perspectives from the Arab. World Food
Security 3(Suppl 1):S1–S6
Kloos J, Gebert N, Rosenfeld T, Renaud F (2013) Climate change, water conflicts and human
security: regional assessment and policy guidelines for the Mediterranean, Middle East and
Sahel, Report No. 10 Aug 2013. United Nations University Institute for Environment and
Human Security, Bonn, Germany
LaGraffe D (2012) The youth bulge in Egypt: an intersection of demographics, security, and the
Arab spring. J Strateg Security 5(2):65–80
Lake DA (1997) Regional security complexes: a systems approach. In: Lake DA, Morgan PM
(eds) Regional orders: building security in a new world. Pennsylvania State University Press,
US
Lewis B (1996) The Middle East: a brief history of the last 2000 years. Simon & Schuster, New
York, US
9 Managing Water, Energy, and Food for Long-Term Regional Security 191

Lewis B (2002) What went wrong?: the clash between Islam and modernity in the Middle East.
Oxford University Press, New York, USA
Mo Ibrahim Foundation (2011) African agriculture: from meeting needs to creating wealth.
Ibrahim Forum 2011, Tunis, Tunisia. Accessed 31 Dec 2015 http://static.
moibrahimfoundation.org/downloads/publications/2011/2011-facts-&-figures-african-
agriculture-from-meeting-needs-to-creating-wealth.pdf
Moussalli A (2014) The dynamics of Arab uprisings and Middle Eastern geopolitics. In:
Axworthy TS, Adeel Z (eds) Water, Energy, and the Arab Awakening. UNU-INWEH,
Hamilton, Canada
Nasdaq (2015) http://www.nasdaq.com Accessed 31 Dec 2015
Pillar PR (2011) Alienation and rebellion in the Arab World. Mediterranean Quarterly 22(4):8–19
Raphaeli N (2006) Unemployment in the Middle East: causes and consequences. Inquiry and
Analysis Series Report No. 265, 10 Feb 2006, Middle East Media Research Institute,
Washington DC, US
Schuster-Wallace CJ, Qadir M, Adeel Z, Renaud F (2014) Putting water and energy at the heart of
sustainable development. UNU-INWEH, Hamilton, Canada
UNDP (1994) Human development report 1994. United Nations Development Programme
(UNDP), New York, US
UNDP (2015) Human development report 2015: work for human development. United Nations
Development Programme (UNDP), New York, US
van Lavieren H, Burt J, Feary DA, Cavalcante G, Marquis E, Benedetti L, Trick C, Kjerfve B,
Sale PF (2011) Managing the growing impacts of development on fragile coastal and marine
ecosystems: lessons from the Gulf. A policy report, UNU-INWEH, Hamilton, Canada
Werz M, Hoffman M (2013) Climate change, migration, and conflict. In: Werrel CE, Femia F
(eds) The Arab spring and climate change—a climate and security correlations series. Center
for American Progress, Washington DC, US
World Bank (2015) GDP per capita (current US$) updated 22 Dec 2015. Accessed 31 Dec 2015
from: http://data.worldbank.org/indicator/NY.GDP.PCAP.CD
Zou’bi MR (2014) The Arab pseudo-spring?: a snapshot of the underlying politics and economics,
and the challenge of water insecurity. In: Axworthy TS, Adeel Z (eds) Water, energy, and the
Arab awakening. UNU-INWEH, Hamilton, Canada
Chapter 10
Current Water for Food Situational
Analysis in the Arab Region and Expected
Changes Due to Dynamic Externalities

Rabi H. Mohtar, Amjad T. Assi and Bassel T. Daher

Abstract Uneven distribution of water, food, and energy (WEF) resources, toge-
ther with demographic, geographic, political, and other natural constraints, burden
WEF security plans in many regions of the world. Decision makers are under
pressure to bridge the WEF supply-demand gap, and often propose reactive, rather
than preventive, strategies that are associated with uncertainty, challenges of sus-
tainability, and various socio-economic, environmental, cultural, and political
drawbacks. Arab countries, like most regions of the world, face internal and
external challenges to managing, sustaining, and securing these scarce, unevenly
distributed natural resources. According to the World Bank (2007), current man-
agement plans for addressing the WEF security challenges faced by the Arab
countries fail to take into account the complicated internal and external dynamics of
the region. This chapter looks holistically at the region and provides a situational
analysis of the major stresses on WEF securities by specific location; it describes
how these are impacted by externalities such as climate change, population growth,
and economic development. The chapter explores the uncertainties associated with
predicting the future of these external stresses and identifies management risks
associated with the lack of understanding of the soil-water and associated spatial
variability. The chapter concludes with a vision for adaptive management
approaches that address the external stresses on regional WEF security, and offer a
vision for increased resilience of local communities. The paper concludes by dis-
cussing ways to expand the vision and adapt it to offer a more general, broader
security by localizing water and food securities.

R.H. Mohtar (&)  A.T. Assi  B.T. Daher


Department of Biological and Agricultural Engineering,
Texas A&M University, College Station, TX 77843-2117, USA
e-mail: mohtar@tamu.edu
R.H. Mohtar
Zachry Department of Civil Engineering, Texas A&M University,
College Station, TX 77843-2117, USA

© Springer International Publishing AG 2017 193


K. Amer et al. (eds.), The Water, Energy, and Food Security Nexus
in the Arab Region, Water Security in a New World,
DOI 10.1007/978-3-319-48408-2_10
194 R.H. Mohtar et al.

 
Keywords Water-food nexus Localizing resources Primary resources security 
Water security  Food security 
Arab Region 
Adaptive management 
Non-stationary stresses

10.1 Introduction

Arab countries face multiple internal and external challenges to manage, sustain,
and secure three scarce and unevenly distributed natural resources: water, energy,
and food (WEF). The uneven distribution of these resources is a general charac-
teristic of the region, and together demographic, geographic, political, and other
natural constraints, exert burdens on WEF security plans there. The decision makers
in these countries are under continuous pressure to seek solutions to bridge the
WEF supply-demand gap. Thus, proposed strategies, are most commonly reactive
rather than preventive, are usually associated with the uncertainties regarding
sustainability, and have various socio-economic, environmental, cultural, and
political drawbacks. According to the World Bank (2007), current management
plans for addressing the WEF security challenges faced by the Arab countries fail to
take into account the complicated internal and external dynamics of the region.
In Arab countries, water security is the most critical challenge. The renewable
water resources are the lowest in the world at 1500 m3/capita/year (WRI 2005). The
majority of the water running through major rivers in the region originates outside
of the region, exposing it to vulnerabilities that lead to political unrest, which in turn
shakes its fragile WEF security. The Southeast Anatolia Development project, the
Grand Ethiopian Renaissance Dam, and the Israeli diversion of the Jordan River are
examples of the vulnerability of WEF systems to externalities in Syria, Iraq, Egypt,
Sudan, Jordan, and Palestine, respectively. Another external pressure affecting
WEF security in the Arab Region is climate change, which is expected to hit this
region the hardest, though with huge variations among the different countries
(Droogers et al. 2012). Climate change is associated with other social, economic,
and political dynamics that will shape the future of water-energy-food security in
the region.
The general strategic plan for securing food supply in the Arab Region is
through food imports (virtual water trade), predicted to constitute more than 60 %
of the food basket by 2030 (World Bank 2009). Grains form the bulk of this import,
provide 30–40 % of total caloric intake of an average person in the region (Larson
et al. 2013; Wright and Cafiero 2011). The decision to rely on food imports was
made under pressure and led to unintended consequences such as the degradation of
the quality of the primary natural resource, arable soil. The decision also had
significant socio-economic impact on farmers, who comprise a large segment of
Arab society. It added vulnerability to global markets. Barnes (2013) warned
against the misuse of the “virtual water” concept and its application in agricultural
lands in Egypt, particularly if adopted without a clear plan and mitigation measures.
10 Current Water for Food Situational Analysis … 195

A key solution for facing the variability in challenges is localizing water and food
securities. In this paper, we will shed light on one significant resource, “green water”,
whose use can save water, food, and energy for many countries in the area. We also
want to encourage innovative thinking through optimized use of natural resources,
promote the benefits to be gained from technological advances, and stress the value
of continuous inclusion of the farmer, the primary steward of this resource.

10.2 Situational Analysis of the Internal and External


Stresses on Arab Countries’ Water-Food Securities

The water-food security of the Arab world is constrained by several internal and
external stresses. Key among them are high population and economic growth
(socio-economic), arid and semi-arid climate, dependency on external water
resources and food supply, climate change, and political unrest. There is a high
variability and uneven distribution of the two main resources for food production,
namely arable land and water. There are generally great disparities in wealth,
natural resources, and economic systems; thus, adaptive strategies for maintaining a
sustainable level of water-food security must vary. This section provides a situa-
tional analysis of these internal and external dynamic stresses and describes how
these stresses impact water-food security in the Arab world. To capture the vari-
ability of these stresses among the Arab countries, we categorize the Arab world
into four geographical units:
(1) Arabian Peninsula: Bahrain [BHR], Kuwait [KWT], Oman [OMN], Qatar
[QAT], Saudi Arabia [SAU], United Arab Emirates [UAE], and Yemen
[YEM];
(2) Middle East: Iraq [IRQ], Jordan [JOR], Lebanon [LBN], State of Palestine
[PSE], and Syria [SYR];
(3) North-eastern Africa: Comoros [COM], Djibouti [DJI], Egypt [EGY], Somalia
[SOM], Sudan [SDN];
(4) Maghreb: Algeria [DZA], Libya [LBY], Mauritania [MRT], Morocco [MAR],
and Tunisia [TUN].

10.2.1 Population Growth

In the past three decades, the population in the Arab Region has more than doubled.
Though fertility rates have consistently declined, the population rate continues to be
high when compared to the rest of the world (UNDP 2012). This rapid increase in
population makes the Arab Region one of the youngest regions (Brown and
Crawford 2009), a reality that presents a clear opportunity for the Arab countries if
196 R.H. Mohtar et al.

Population in the Arab Word [1980, 2014, 2050]

2,50,000 1,000

900

2,00,000 800

700

Percent Increase (%)


Population (Thousands)

1,50,000 600

500

1,00,000 400

300

50,000 200

100

0 0
Middle East
Arabian Peninsula

IRQ

Maghreb

MRT

MAR
BHR

KWT

OMN

QAT

SAU

UAE

YEM

JOR

LBN

PSE

SYR

COM

EGY

SOM
DJI

SDN

DZA

LBY

TUN
NE Africa 1980 2014

2050 %increase 1980-2014

Fig. 10.1 The population growth in the Arab world from 1980 to 2014 and the projected
population in 2050. Source ESA UN (2012)

the right investments were made in human capital and training for the labor market.
Youth represents a major economic asset (Madsen et al. 2007), but it also presents
serious challenges in the absence of adequate infrastructure for education and health
care, increased unemployment, malnutrition, and slow economic growth (UNFPA
2012). Further, it imposes an overburdening strain on already limited water
resources and arable soil with regard to meeting growing water-food demands. The
population in the Arabian Peninsula is two and a half times more than it was in
1980 (Fig. 10.1). Projections show that the population in the Arab world as a whole
will reach 692 million by 2050 (Zyadin 2013). It is projected that North East Africa
will maintain the highest population, while the highest percentage of population
growth is expected to take place in the Arabian Peninsula countries. Arab countries
across different regions continue to grow at highly variable rates due to different
patterns of fertility and immigration.

10.2.2 Economic Development

The landscape of economies in the Arab world is characterized by great variability


across different countries (Fig. 10.2). In 2013, Qatar recorded the highest GDP per
capita in the world (USD 102, 100) while Somalia bottomed the global list with
USD 600 (ESA UN DESA 2012). Mainly catalyzed by a combination of high oil
and gas returns and low local population, the Arabian Peninsula enjoys a higher
GDP per capita compared to the other regions. The sectors that contribute to the
10 Current Water for Food Situational Analysis … 197

Fig. 10.2 The gross domestic product (GDP) and gross domestic product per capita (GDPPP) in
year 2013 for countries in the Arab world. Source World Bank (2014)

GDP across countries in the Arab world also highly vary: some have significant
agricultural activity, while others solely rely on industry and services. In 2012, for
example, 14 % of Egypt’s GDP came from agriculture, while 40 and 46 % came
from industry and services respectively. The value added to GDP from agriculture,
industry, and services for the same year was 7, 19, and 73 % for Lebanon, 1, 60,
and 39 % for UAE (World Bank 2014). Economic growth continues to be slow in
the Arab world, for different reasons in different locations. This variability mainly
results from the lack of a stable macroeconomic environment, inadequate human
capital, and excessive reliance on public investment (Elbadawi 2004; Sala-i-Martin
and Artadi 2002). Youth unemployment is nearly double that of the global rate and
continues to grow due to the rigid labor market and its inefficiencies (World
Economic Forum 2013, 2014). A different set of interventions are needed to curb
existing trends across different countries.

10.2.3 Water Availability and External Water Resources


Dependency

The Arab world is considered to be one of the most water-scarce regions in the
world. This scarcity has several dimensions.
(1) The physical “hydrological” scarcity due to the limited internal renewable
water resources, which seems to become more severe with the climate change.
(2) The high dependence on the external renewable water resources from other
Arab countries and from non-Arab countries. Currently, the external renew-
able water resources of Egypt, Sudan, Iraq from non-Arab countries are 97,
89, and 61 % of their total renewable water resources, respectively (see
Fig. 10.3).
198 R.H. Mohtar et al.

Internal and External Renewable Water Resources Per Capita [2014]


100
3,000 97 97 96 100
88 89
2,500
80

Percentage (%)
2,000 61
57 59
m3/capita/yr

58 60
1,500
40
1,000 27
21
20
500 9
4
2 3 1 0
0 0 0 0 0 0 0
0 0
BHR

QAT

UAE

PSE
SYR

EGY

DZA

TUN
OMN
Arabian Peninsula

SAU

YEM

IRQ

LBN

DJI

SDN

LBY
KWT

MRT
MAR
JOR

COM

SOM

Maghreb
NE Africa
Middle East

ERWR [2014] IRWR [2014]


Absolute Water Scarcity Water Scarcity
Water Stress Water Vulnerable
Percent of External RWR from Total RWR

Fig. 10.3 The internal and external renewable water resources in the Arab world. Source FAO
(2014)

(3) The socio-economic dynamics represented by the sharp increase in population,


shift in life styles, and economic growth. For example, economic growth in
Qatar and United Arab Emirates caused a population growth of nine times
more than it was three decades ago, imposing substantial strains on already
scarce water resources. Such a socio-economic situation has never been
experienced in any region except in the Arabian Peninsula. Moreover, this
region is classified as an absolute water scarcity region (i.e. <500 m3/
capita/yr) according to the physical water stress classification of Falkenmark
et al. (1989) (Fig. 10.3).
(4) The depletion of the vulnerable water resources to bridge the growing
supply-demand gap.
(5) The deterioration of water and land resource quality, due to exposure to
low-quality domestic and industrial wastewater and agricultural return flow.
Arab countries can be categorized into four groups based on the sources and
availability of water to meet the current demand: (1) renewable surface water-based
(Egypt and Sudan); (2) renewable surface and groundwater-based (Iraq, Lebanon,
Syria, Morocco, Comoros, Mauritania); (3) renewable and non-renewable
groundwater-based (Algeria, Tunisia, Libya, Saudi Arabia, Oman, Jordan, and
Palestine); (4) mainly non-renewable groundwater-based (Kuwait, Qatar, United
Arab Emirates, and Yemen).
10 Current Water for Food Situational Analysis … 199

The internal renewable water resource from rainfall, physical water scarcity, is
the lowest in the Arab world. Each of the Arabian Peninsula countries receives less
than 500 m3/capita/yr, and, except for Oman, each has already exceeded their
renewable water resources. In the Middle East, only Iraq and Lebanon have a better
water quota (between 500 and 1000 m3/capita/yr), while Jordan, Palestine, and
Syria are still categorized as absolute physical water scarcity countries. Except
Comoros, all the North East Africa countries are also classified as absolute physical
water scarcity countries. The same classification applies for all Maghreb countries
except Morocco, still considered to be a water-scarce country (Fig. 10.3). One
should keep in mind that food security is not only a function of water availability,
but also of land fertility and climate conditions. The considerable variability of land
fertility and climate conditions among these countries is an important characteristic
of the entire region and must be considered in any water-food security plans.
The Middle East region of the Arab world is characterized by political insta-
bility. Water security and political instability are interconnected. All of the coun-
tries in the Middle East share trans-boundary water bodies among themselves and
with other, non-Arab countries. Most of the 60 % renewable water resources in Iraq
and Syria originate in Turkey and flow through the Euphrates-Tigris basin. Given
the Turkish water-food-energy security plans for the Southeast Anatolia
Development project, it is to be expected that the growing water demand in Iraq and
Syria will not be easily obtained through these rivers; similarily, the sitation is not
any better in other parts of this region. More than 66 % of the flowing water in the
Jordan River is extracted by Israel, while only 18, 12, <1, and 0 % is utilized by
Jordan, Syria, Lebanon, and Palestine, respectively. As for the Nile River basin,
around 97 % of the total renewable water resources of Egypt and 89 % of those in
Sudan, flow through the Nile River. Thus, the Grand Ethiopian Renaissance Dam is
of great concerns to both countries. In this part of the world, water policy plays a
pivotal role in shaping the future of water-food security.

10.2.4 Climate Change

According to many global and local climate change models, the Arab world will
receive less precipitation and an associated increase in temperature. The spatial and
temporal variations of these two climate characteristics are high, not only between
the countries, but also within a given country. The uncertainties in the predicted
values of various climate change models are very high. A study for the World Bank
by Immerzeel et al. (2011), assessed the future water supply-demand under various
climate change and socio-economic scenarios. It was found that climate change will
be responsible for 16 % of the water shortage in the region by 2050. Whereas,
Drooger et al. (2012) found that 22 % of the water shortage in the region will be
due to climate change, and 78 % due to the socio-economic dynamics by 2050.
In the study by Immerzeel et al. (2011), one can recognize the high variation in
projected precipitation and actual evapotranspiration based on the results of nine
200 R.H. Mohtar et al.

Median Precipitation and Reference Evapotranspiration Anomalies of the 9 GCMs as a


percentage:[2040-2050] with respect to [2000-2009]

Arabian Peninsula
BHR
KWT
OMN
QAT
SAU
UAE
YEM
Middle East
IRQ
JOR
LBN
PSE
SYR
NE Africa
DJI
EGY
Maghreb Median Precipitation Anomalies
DZA
LBY Median Reference Evapotranspiration
MAR Anomalies
TUN
-15.0 -10.0 -5.0 0.0 5.0 10.0 15.0
Percentage [%]

Fig. 10.4 The median precipitation and reference evapotranspiration anomalies of 9 GCMs as a
percentage comparing the (2040–2050) scenarios with respect to current situation (2000–2009).
Source Immerzeel et al. (2011)

different Global Climate Models (GCMs). With regard to precipitation, the pre-
diction intervals were [−21 %, 35 %] in the Arabian Peninsula with a median value
of 2.1 %; [−19 %, 7 %] in the Middle East with a median value of −6.4 %;
[−11 %, 22 %] in Egypt and Djibouti of the North-east Africa region with a median
value of 1.5 %; and finally [−19 %, 7 %] in Maghreb with a median value of
−7.3 %. Only five countries are expected to have more precipitation by 2050, three
of these are in the Arabian Peninsula: Oman, United Arab Emirates, and Yemen
(Fig. 10.4). Rainfall intensity brings another uncertainty in the future of water-food
security in the region. Changes in the climate are expected to accelerate the
hydrologic cycle, thus increasing the probability of flooding and desertification:
both of which impose substantial strain on the water-food scarcity in the region.
The increase in temperature also increases water demand, mainly from the
agricultural sector, by increasing evapotranspiration. Evapotranspiration in the Arab
World is expected to increase by 1–3.6 % by 2050. The greatest will be in the
Middle East, with a median value of 3.1 % (Fig. 10.4).
Climate change is considered to pose a major threat to water-food security in the
region and is expected to be responsible for around 20 % of the supply-demand gap
by 2050. The uncertainty in the predicted values are very high and the temporal
domain of the occurrence of such extreme events is not clear. Severe droughts can
hit at any time, and would aggravate socio-economic and political instability in the
region.
10 Current Water for Food Situational Analysis … 201

10.2.5 Global Food Prices and Their Implications on Arab


Countries

The Arab world is well-known for dependence on cereals in its diet. About 30–
40 % of total calories of an average person in the region comes from wheat. Thus, it
is not surprising that more than 29 % of the global wheat export between 2008 and
2010 was to the region (Larson et al. 2013). Several Arab countries attempted to
achieve self-sufficiency, at least in wheat production. For example, in 1984, Saudi
Arabia subsidized wheat production, becoming a net exporter of wheat in less than
10 years. However, this success was costly as it resulted in depleting valuable
groundwater resources, affecting the diversity of agricultural products in the
kingdom (Sowers et al. 2010). Currently, the plans are to cease wheat production in
Saudi Arabia and to either import cereals or secure access to arable land in other
countries (mainly in Africa and Asia), sufficient to meet the increasing demand for
cereals.
Grain import (virtual water trade) seems to offer an easy solution to bridging the
unmet water and food demand in these countries. Due to the socio-economic,
hydrologic, and climate change constraints, the tendency is toward boosting food
imports up to 64 % by 2030 (World Bank 2009). Such a plan is associated with
many risks: some large grain exporters are converting their product into ethanol and
it is expected that by 2050, 182 million tons of cereals will be grown for
bio-fuel production, compared to only 65 million tons in 2005 (Alexandratos and
Bruinsma 2012). This will impact both global food prices, and the local markets in
the Arab world, where the poor and vulnerable spend 30–70 % of their income on
food. On average, local food markets showed 5–10 % variation as a response to the
global market prices in the period 2006–2011. The highest variation was observed
in Egypt, Yemen, Iraq, and Palestine (Larson et al. 2013). Poor people in poor
countries are the least resilient to changes in food prices. These represent a quarter
of the population in the region. It is not only the vulnerable people, but also the
precious land that could suffer from the side effects of the increased food imports.
Barnes (2013) addressed the drawbacks of the Egyptian policy to decrease the area
of water-intensive rice cultivation and increase rice imports, noting that it was the
land and the farmer who received the shock.

10.3 Uncertainties Associated with the External Stresses

It is clear that socio-economics, politics, and climate change will play an influential
role in shaping the future of food and water security in the Arab world. However,
substantial uncertainties are associated with these projections. The uncertainties
increase the vulnerability of the regions’ inhabitants and reduce the resilience of
water-food security adaptive strategies. In the last decade, rapid population growth
in the Arab Peninsula resulted from the greatly increased demand in the labor
202 R.H. Mohtar et al.

market; three wars in Iraq and Syria caused significant increase in the population of
Jordan and Lebanon in a very short time. It is challenging for any adaptive plan to
accommodate socio-economic and political dynamics in such a short time frame.
Given that the socio-economic factor is anticipated to be responsible for about 80 %
of the unmet demand by 2050, it is critical that water strategies, in the face of such
dynamics, focus on local-scale water supply projects and technical solutions. These
would enforce the resilience of water-food security systems that require the
engagement of local societal factors. Moreover, global biofuel production is
expected to double by 2050 and result in a shortage of (or increased
competition/demand for) cereals. There is no guarantee about the future production
of bio-fuel, cereals production for the food market, and their prices. These uncer-
tainties are alarming for Arab world countries, which plan to increase food imports,
mainly of grains.
Several studies using different GCMs have been conducted to predict future
climate changes. Two common results can be concluded from these studies: (1) in
average climatic scenarios, climate change will be responsible for about 20 % of the
water supply-demand gap by 2050, and (2) there is high variation among the
different models. For example, in the Immerzeel et al. (2011) study, the precipi-
tation anomalies, as a percentage in the Arab world, are predicted to be within the
interval [−18 %, 18 %] with a median value of −2.1 %. The questions then
become: when can we expect the worst case scenario? Are we ready for it? Can our
large-scale projects and plans cope with it? All that we observe from recent drought
and flood events tell us we are not ready yet. One way for mitigating such risks is to
develop reliable local climate predictions. Under these uncertainties, water experts
and decision makers should work on creating innovative solutions by understanding
the local environmental system, and engaging the societal actors.

10.4 Toward Innovative Solutions for a Better Future


of Water-Food Security in the Arab Region

Agriculture is considered to be a main constituent of national security and social


identity in many countries. Agriculture not only depends heavily on water, but also
on land: both of which are threatened by climate change. Furthermore, considerable
amounts of energy are embedded within different stages of food production. The
agricultural sector embeds within it vast resources and is complex in terms of its
attributes and its value to society. Agriculture consumes the most water in a region
classified as the most water-scarce in the world. The agricultural sector is also the
main source of income for the most vulnerable parts of the society in the Arab
Region. An important step towards identifying ways to encourage agricultural
resilience within the Arab countries is to investigate the potential of local resources
and their ability to withstand the stresses facing these resource systems.
10 Current Water for Food Situational Analysis … 203

Good management of this sector demands a holistic, yet localized understanding


of the inter-linkages among different resources (water, energy, soil, and land),
externalities (economics, population and demographics, politics, and climate), and
involved stakeholders (governments, farmers, investors, and civil society). No
single water-food security strategy could be generalized over the entire region due
to the enormous variability in resource availability, external factors, and the range
of involved stakeholders. This variability is highly pronounced among different
countries, and even among different sub-regions in the same country. Therefore,
localized adaptation strategies need to be developed, based on these elements.
Developing local-specific plans will contribute to reducing the vulnerabilities of
local societies and the stresses that face resource systems. It will also help improve
levels of resilience of natural and socio-economic systems as they face external
stresses such as climate change.
A 199 m3/yr water shortage is expected by year 2050 (Droogers et al. 2012) in
the Middle-East and North African (MENA) region. With the increase in the water
shortage, supplying domestic households from fresh water resources will take
priority over agriculture, and will create a challenge for agriculture experts and
decision makers to meet market needs. This could lead to reduction in local agri-
cultural production and to bridging shortfalls in food demand through imports,
while accepting associated risks and vulnerabilities. Alternatively, localized adap-
tation scenarios could be developed to maintain a certain level of local food pro-
duction. Immerzeel et al. (2011) discusses nine adaption scenarios to manage the
unmet water demand. The conclusion of the study favors improved agriculture
practices, such as heat and salt tolerant crop varieties. It also favors desalination as
part of an efficient adaptation strategy for water resources. This conclusion is in
alignment with the need to localize food production, farming systems, and adaption
plans, in order to increase resilience of the agricultural system as it faces a wide
array of dynamic stresses. Moreover, Verner et al. (2013) concluded that farmers
and rural inhabitants are key players in increasing the resilience of the agricultural
sector’s response to climate change. They are at the front-line of efforts to protect
their source of income in a non-stationary world. The study highlighted the
importance of regular, on-farm agricultural experiments to identify, evaluate, and
manage the changes in local production systems rather than adopt ‘one-size-fits-all’
solutions.
Arab countries currently meet a significant portion of their food market needs
through import. Amounts and types of food import vary from one place to another,
depending on resource availabilities and enabling conditions. Agricultural trade
plays an important role in today’s global economy, and will continue to be a major
contributor in supplying markets worldwide. Nevertheless, the fact remains that
many potentials for improving levels of local food production in the area are
under-tapped, or yet to be developed. The Arabian Peninsula is rich with energy
resources (renewable “solar” as well as non-renewable “fossil”) yet, very poor in
fresh water and fertile soil. Technology is advancing rapidly, and it is only a matter
of time before efficiency and storage issues are overcome for solar energy. Such a
clean, cheap, and renewable source of energy can bring agricultural life to the area.
204 R.H. Mohtar et al.

Blue and Green Water in the Maghreb Countries (km3/yr)


200

150
km3/yr

100

50

0
Algeria Egypt Libya Morocco Tunisia Mauritania
Precipitation 198.6 36.3 76.6 113.6 44.7 89.9
Green Water flow 182.1 35.6 73.1 100.3 38.5 83.0
Geenwater storage 11.3 0.6 3.1 7.5 3.3 1.5
Blue Water 5.2 0.1 0.4 5.7 2.9 5.4

Fig. 10.5 The green and blue water portions from the total precipitation. Source Schuol et al.
(2008)

Nonetheless, unique localized farming systems need to be adopted where


soil-water-atmosphere can be monitored and controlled. Sahara Forest Project1
launched as a pilot project in Qatar, and now under construction in Jordan, provides
case studies for localized farming systems in which technologies can be combined
and used to alleviate the impact of severe environmental externalities such as
limited water resources, limited arable soil, and year round very high temperatures.
Another under-appreciated asset in the region is green water, defined by World
Soil Information (Ringersma et al. 2003) as ‘the fraction of rainfall that infiltrates
the soil, and includes soil water holding capacity’. This water needs to be better
quantified and managed. It maintains, determines, and even improves the produc-
tivity and functionality of the soil-water-atmosphere continuum. Once quantified
and managed, it can reduce soil erosion, maintain soil fertility, and most impor-
tantly, maintain the equilibrium between the soil-water-plant system and its sur-
rounding climate. Such a function will be essential to face the stresses associated
with climate change. Figure 10.5 shows the amounts of available blue and green
water in Maghreb countries (Schuol et al. 2008). The data shows that blue water
comprises only 4 % of the renewable water resources, while green water constitutes
the remaining 96 %. A great deal of research, human capacity, and management
must be channeled into developing technical solutions to better utilize the 4 % (blue
water), but the 96% remains under studied: 91 % of green water flow, evapotran-
spiration, and 5 % of green water is stored in the soil profile. Figure 10.6 shows the
portion of precipitation counted as ‘internal renewable water resources’ across

1
See (http://saharaforestproject.com/).
10 Current Water for Food Situational Analysis … 205

How Much Rainwater Accounted as Internal Renewable Water Resources


25,000 80
70
70
60
20,000 60
m3/capita/yr

Percentage (%)
50
15,000 37
32 40
30
30
10,000 19
15 20
12
6 7 7 6 8 5
5,000 2 4 2 2 2 2 4 3 10
0 1 1 0
0
0 -10
Arabian Peninsula

Middle East

NE Africa

Maghreb
OMN
KWT

COM

MAR
YEM

SOM

MRT
BHR

DJI
QAT

UAE

EGY

DZA
SAU

LBN

SDN

LBY
SYR
JOR

PSE
IRQ

IRWR [2015] Precipitation[Long term average] Percent of IRW from Precipitation


3,000

Vulner
Water
2,000

able
Stress
Water
1,000
0

Middle East
Arabian ...

KWT

NE Africa
BHR

UAE

SYR

Maghreb
JOR

COM
YEM

LBN

SOM

MAR
EGY

SDN

MRT
DZA

TUN
LBY
SAU

DJI
PSE
IRQ
QAT
OMN

IRWR [2015] ERWR [2015]

Fig. 10.6 The amount of rainwater accounted as internal renewable water resources. Source FAO
(2014)

countries in the Arab world. Only eight of the 22 countries show that precipitation
contributes to more than 10 % of blue water. Currently, the simplified represen-
tation of the soil-water medium limits proper accounting and management of this
precious but under-appreciated water. Braudeau and Mohtar (2014) present a new
paradigm that has the capability to quantify green water as it exists in the organized
soil-water system, with thermodynamic equilibrium and its surrounding climate
(temperature, humidity, and pressure). Once quantified, it can be better managed.
206 R.H. Mohtar et al.

Better accounting and utilization of green water offers a promising leap toward
better adaptation to climate change challenges facing food supply chains.

10.5 Conclusions and Recommendations

The Arab world faces water and food security challenges with a set of
non-stationary internal and external pressures. Arid and semi-arid climates, recur-
ring water stresses, high population growth rates, frequent political unrest, and
climate change are characteristic of the challenges that face the region. The high
level of variability in those challenges, whether among different geographic groups
(Arabian Peninsula, Middle East, North-east Africa, and Maghreb), different
countries within a group, or among different areas within a single country, is
another characteristic of the challenges in the region. This makes it difficult to adopt
one-size-fits-all adaptive strategies. Uncertainty of future projections for those
challenges adds to the complexity of prescribing proper strategies that could
respond to possible scenarios.
Under predictions of high uncertainty and variability of future stresses, large-scale
technical solutions cannot provide high levels of resilience, and would threaten the
livelihood of communities across the region. Alternatively, localizing adaptive
strategies, with pivotal participation of local communities, increases the resilience of
water-food security in the face of uncertain stresses. Moreover, lessons learned at the
farm level, would contribute to increasing the ability of farmers and farming systems
to maintain a sustainable level of food security. Innovative thinking is needed to
understand the potential of available and under tapped resources, and to better utilize
them at the local scale, in order to better contribute to bridging the water-food
supply-demand gap. This does not eliminate the role of food imports, rain water
harvesting systems, or existing hydraulic structures contributing to bridging the gap,
but would come at higher risks and lower levels of resilience. Localizing the farming
system, once adopted as a strategy moving forward, would increase the overall
resilience of the water-food systems. Such a strategy should be inclusive and must
account for the natural, social, economic, and technical environments after which
local solutions are prescribed and geared toward optimizing, rather than maximizing,
the use of water, soil, plant and atmosphere. The proposed approach is a
socio-techno-environmental approach as opposed to a solely technical one.

References

Alexandratos N, Bruinsma J (2012) World agriculture towards 2030/2050: the 2012 revision. ESA
Working Paper No. 12-03 June 2012, FAO, Rome
Barnes J (2013) Water, water everywhere but not a drop to drink: the false promise of virtual
water. Criti Anthropol 33(4):371–389. doi:10.1177/0308275X13499382
10 Current Water for Food Situational Analysis … 207

Braudeau E, Mohtar RH (2014) Integrative environmental modeling. Encyclopedia of agricultural,


food, and biological engineering, 2nd edn. doi:10.1081/E-EAFE2-120049111
Brown O, Crawford A (2009) Rising temperatures, rising tensions: climate change and the risk of
violent conflict in the Middle East. International Institute for Sustainable Development,
Retrieved from: https://www.iisd.org/pdf/2009/rising_temps_middle_east.pdf
Droogers P, Immerzeel WW, Terink W, Hoogeveen J, Bierkens MFP, Van Beek LPH, Debele B
(2012) Water resources trends in Middle East and North Africa towards 2050. Hydrol Earth
Syst Sci 16(9):3101–3114. doi:10.5194/hess-16-3101-2012
Elbadawi IA (2004) Reviving growth in the Arab World. World Bank, Washington, DC, July
2004. Retrieved from: http://siteresources.worldbank.org/DEC/Resources/Arab_growth_
revised_July_22_2004.pdf
ESA UN (2012) Population division of the department of economic and social Affairs of the
United Nations Secretariat, World population prospects: the 2012 revision. Retrieved from:
http://esa.un.org/unpd/wpp/index.htm
Falkenmark M, Lundqvist J, Widstrand C (1989) Macro-scale water scarcity requires micro-scale
approaches. Aspects of vulnerability in semi-arid development. Nat Res Forum 13:258–267
FAO (2014) AQUASTAT database. Food and agriculture organization of the United Nations
(FAO), Rome. Accessed 30 Oct 2014
Immerzeel WW, Droogers P, Ternik W, Hoogeveen J, Hellegers P, Bierkens M, van Beek R (2011)
Middle-East and North Africa water outlook. World Bank Study. Future Water Report 98
Larson DF, Lampietti J, Gouel C, Cafiero C, Roberts J (2013) Food security and storage in the
Middle East and North Africa. World Bank Econ Rev 28(1):48–73. doi:10.1093/wber/lht015
Madsen EL, Daumerie B, Hardee K (2007) The effects of age structure on development: policy
and issue brief. Population Action International, Washington, DC
Ringersma J, Batjes NH, Dent DL (2003) Green water: definitions and data for assessment. Report
2003/2 ISRIC—World Soil Information, Wageningen, The Netherlands
Sala-i-Martin X, Artadi EV (2002) Economic growth and investment in the Arab world.
Discussion Paper No. 0203-08. Department of Economics, Columbia University, NY.
Retrieved from: http://hdl.handle.net/10022/AC:P:523
Schuol J, Abbaspour KC, Yang H, Srinivasan R, Zehnder AJB (2008) Modeling blue and green
water availability in Africa. Water Resour Res 44:W07406. doi:10.1029/2007WR006609
Sowers J, Vengosh A, Weinthal E (2010) Climate change, water resources, and the politics of
adaptation in the Middle East and North Africa. Clim Change 104(3–4):599–627. doi:10.1007/
s10584-010-9835-4
UNDP (2012) United Nations Development Programme—Regional Bureau for Arab States
(RBAS). Retrieved from: http://www.arab-hdr.org/data/profiles/AS.aspx
UNFPA (The United Nations Population Fund) and The Government of Yemen (2012) Country
programme action plan. Retrieved from: https://data.unfpa.org/downloadDoc.unfpa?docId=238
Verner D, Lee D, Ashwill M, Wilby R (2013) Increasing resilience to climate change in the
agricultural sector of the Middle East: the cases of Jordan and Lebanon. A World Bank Study,
Washington, DC. doi:10.1596/978-0-8213-9844-9
World Bank (2007) Making the most of scarcity: accountability for better water management
results in the Middle East and North Africa. World Bank, Washington, DC
World Bank (2009) Improving food security in Arab countries. A joint report by FAO, IFAD and
the World Bank. World Bank, Washington, DC
World Bank (2014) Indicators database. World Bank IBRD/IDA. Retrieved from: http://data.
worldbank.org/indicator/
World Economic Forum (2013) World Economic Forum—The Arab World Competitiveness
Report 2013. Retrieved from: http://www3.weforum.org/docs/WEF_AWCR_Report_2013.pdf
World Economic Forum (2014) World Economic Forum—The Global Competitiveness Report
2014–2015. Retrieved from: http://www3.weforum.org/docs/WEF_GlobalCompetitiveness
Report_2014-15.pdf
208 R.H. Mohtar et al.

WRI (2005) World resources 2005—the wealth of the poor: managing ecosystems to fight poverty.
World Resources Institute (WRI) in collaboration with United Nations Development
Programme, United Nations Environment Programme, World Bank and World Resources,
September 2005, Washington, DC
Wright B, Cafiero C (2011) Grain reserves and food security in the Middle East and North Africa.
Food Security 3(S1):61–76. doi:10.1007/s12571-010-0094-z
Zyadin A (2013) Water shortage in MENA region: an interdisciplinary overview and a suite of
practical solutions. J Water Res Prot 2013(5):49–58
Chapter 11
Case Study: Masdar Renewable Energy
Water Desalination Program

Mohammad El Ramahi

Abstract In 2013, Masdar launched a renewable energy desalination pilot program


to research and develop energy-efficient, cost-competitive desalination technologies
that are suitable to be powered by renewable energy. The long-term goal is to
implement renewable energy-powered desalination plants in the United Arab
Emirates, as well as the wider MENA region and to have a commercial scale facility
operating by 2020. Through a competitive tender, four commercial partners—
Abengoa, Suez, Sidem/Veolia and Trevi Systems—were selected to support the
development of the program. Each of them was to develop and operate a
next-generation pilot seawater desalination plant. The four plants are to test a range
of innovative approaches in boosting operational efficiency. The program consists
of two stages:
• Pilot phase (2013–2016): Based in Ghantoot, Abu Dhabi, the commercial
partners constructed four small-scale desalination pilot plants. They are to
operate continuously for at least 18 months. The performance of the plants will
be assessed, rigorously monitored and tested by Masdar.
• Implementation and development (after 2017): Scaling-up of technologies that
meet predefined criteria as commercially viable, large-scale seawater desalina-
tion plants. The plants will be entirely powered by renewables.
The Masdar Institute of Science and Technology conducts four research projects
that operate alongside and, in addition to, the pilot program. The desalination
project is sponsored by the Abu Dhabi Government, with co-funding provided by
the industry partners. Masdar is leading the project management and coordinating
the program with the Abu Dhabi Water and Electricity Authority (ADWEA),
Regulation & Supervision Bureau (RSB), Environment Agency—Abu Dhabi
(EAD), and Abu Dhabi Sewerage Services Company (ADSSC).

 
Keywords Seawater desalination Energy efficiency Renewable energy Pilot 
 
program Innovative desalination technology United Arab Emirates Masdar 
M. El Ramahi (&)
Asset Management and Technical Services, Masdar Clean Energy, Masdar, UAE
e-mail: melramahi@masdar.ae

© Springer International Publishing AG 2017 209


K. Amer et al. (eds.), The Water, Energy, and Food Security Nexus
in the Arab Region, Water Security in a New World,
DOI 10.1007/978-3-319-48408-2_11
210 M. El Ramahi

11.1 Introduction

In view of the rising water scarcity in the world and particularly in the Gulf region,
it is crucial to identify a sustainable solution to meet future long-term water needs.
In Abu Dhabi, the Abu Dhabi Water and Electricity Company has predicted a
deficiency of 339 million imperial gallons per day (MIGD) (equivalent to 1.5
million m3/day) by the year 2020; this deficiency is expected to increase to 634
MIGD (2.9 million m3/day) by the year 2030.
For a long time, desalination has been the major source of potable water in arid
regions like the United Arab Emirates (UAE). As an example, desalination has
provided a reliable source of fresh water to the growing population and economies
in the Gulf region for nearly half a century. No real development in society or
industry in this region would have been possible without the parallel development
and implementation of desalination.
Clearly, desalination is a critical component of sustaining life and economy in
the Gulf region. Some countries in the area rely on desalination to produce 90 % or
more of their drinking water, and the overall capacity installed in this region
amounts to about 40 % of the world’s desalinated water capacity.
The desalination industry that we know today is founded on the premise that was
created in the Middle East back in the 1950s to support the region’s people and
industry in these largely expanding countries. One can even speculate that there
would not have been an oil and gas industry, had there not been a desalination
industry serving the community and supporting incredible development in the
Middle East.
In the past, the desalination market in the Middle East was largely driven by
robustness rather than efficiency, and water production was realized with costly and
energy-intensive processes.
Today, the situation is quite different. Desalination has become a solution to
water problems well beyond the traditional arid areas of the Middle East and North
Africa. With the emergence of desalination as part of mainstream water resource
management in many parts of the world, the desalination industry has also paid
increasing attention to reducing energy consumption and increasing environmental
responsibility in its practice. We have seen the success of these efforts, with energy
consumption in desalination overall declining by nearly 50 % over the past two
decades. However, much remains to be done, with enormous benefits still to be
realized.
A gradual but constant change in perspective and environmental understanding
is driving desalination technology towards more efficient and environmentally
friendly options and towards energy optimization.
Energy consumption in the desalination sector is a typical case where innovation
has played an important role, especially in the last decade. The energy consumption
of seawater reverse osmosis (RO) technology has dropped to levels in the range of
4–5 kWh/m3 thanks to both the adoption of more sophisticated energy recovery
devices and also to the development of membranes that can operate at higher
11 Case Study: Masdar Renewable Energy Water Desalination Program 211

recovery and better flux. However, the energy intensity required, even by
state-of-the-art desalination plants, is still far too high and represents a real barrier
to allow a medium- to large-scale desalination plant to be powered economically by
a renewable energy source.
While a lot has already been done to reduce the energy intensity of desalination
plants, it is essential that the industry continue to look ahead in order to make sure
that in the medium to long distant future, our choices are truly sustainable, and that
renewable energy solutions can be applied to desalination as well.
The solution to this problem lies in innovative concepts that can challenge
established technologies. In the past, the main issue was that these technologies
were rarely given a chance to be proven beyond the laboratory or pilot scale project.
Therefore, there is a risk that some of them will never be able to develop properly
into an industrial reality. Abu Dhabi and Masdar have made a commitment to
decrease desalination energy intensity and to power desalination technologies with
the region’s abundant renewable resources, such as solar, wind, and geothermal as a
solution to improve water security, reduce the environmental impact of desalina-
tion, and reduce the cost of the produced water.
Against this backdrop, Masdar launched the Renewable Energy Desalination
Program (the “Program”) in January 2013 after receiving a mandate from Abu
Dhabi’s leadership. This forward-thinking mandate directed Masdar to develop and
demonstrate advanced and innovative technologies in desalination for the purpose
of both ensuring water security and reducing energy consumption of the sector in
order to meet the UAE’s energy reduction targets. The successful application of
renewable energy for desalination in the region will depend on several factors.
These are primarily the capacity to create a strong platform of interest that involves
researchers, investors, and governments, as well as success in setting policies that
seriously promote this application on an ongoing basis.
In this respect, Masdar’s Program aims at transforming the desalination industry
into a more sustainable model by enabling the piloting of new innovations in the
field, including not only those that are based on existing technologies, but also those
that implement completely novel and potentially revolutionary concepts.

11.2 Strategic Goal

The goal of the Program is to identify industrial-scale and commercially viable


desalination technologies that will address sustainable access to water both in the
region, and globally. Working with four leading technology partners, the Program
will demonstrate the performance of four pilot seawater desalination plants operated
over 18 months.
The uniqueness of this initiative is that the pilot phase is will be designed to pave
the way to a further step of commercialization of larger scale industrial plants, thus
bridging the gap between Research and Development (R&D) and commercial
implementation of sustainable innovative desalination solutions.
212 M. El Ramahi

Abu Dhabi’s government through Masdar, established long-term strategic


partnerships with the selected technology providers to develop and implement
advanced and innovative renewable energy powered water desalination infras-
tructure in the UAE, leading to win-win situations.
Specifying sustainability and energy efficiency as the primary focus represents a
novel approach in the industry and is designed to create the right synergies between
the academic and research worlds, industry, and public institutions. From there,
gradually, renewable desalination can develop on a larger scale.
Notably, the task of achieving more sustainable and environmental water pro-
duction cannot be viewed in isolation without a parallel task of maintaining low
water tariffs, as well as availability and reliability of the technology.
This is an important aspect in the process of making the innovative desalination
solutions bankable so that they can be supported by investors in an industrial
setting.
For the UAE, the main benefits of the Program are:
• An improvement in the energy efficiency of future desalination plants;
• A diversification of energy supply and enhancement of energy security;
• A reduction in the cost of desalinated water;
• A reduction in the environmental impact of water desalination; and
• Support of the UAE’s endeavor to strengthen its position as a leader in sus-
tainable development.
For the selected technology providers, the main benefits are:
• An opportunity to develop advanced and innovative desalination technologies
and intellectual property with co-financing and support from Masdar;
• An opportunity to demonstrate the developed desalination technologies “on the
ground” under the auspices of Masdar;
• An opportunity to implement a reference project in a region that accounts for
around 40 % of global desalination capacity and establish a strong commercial
footprint; and
• An opportunity to become one of the developers of future large-scale renewable
energy driven desalination plants in the UAE.

11.3 Objectives of the Program

The main objectives of the Program are to develop and demonstrate advanced and
innovative technologies for seawater desalination that are less energy intensive (i.e.,
have a reduced energy consumption per unit of produced water) and are less costly
(i.e., have a reduced cost of the produced water considering the whole desalination
system on a lifetime basis) in comparison to the currently installed seawater
desalination plants in the UAE.
11 Case Study: Masdar Renewable Energy Water Desalination Program 213

It is expected that the Program will encourage the development and large-scale
installation of advanced and innovative seawater desalination technologies that are
energy efficient and suitable to be powered by renewable energy sources.
The Program is also designed to stimulate growth, promote investment, advance
the desalination sector, and champion the commercialization of sustainable and
energy-efficient desalination technologies for the region and the world.
Masdar is coordinating the Program with key Abu Dhabi stakeholders, such as
the Abu Dhabi Water & Electricity Authority, the Regulation & Supervision
Bureau, the Environment Agency of Abu Dhabi, and the Abu Dhabi Sewerage
Services Company.
The Program consists of two stages:
• Stage 1 (2013–2016): Piloting phase to develop and demonstrate
energy-efficient desalination systems at a small scale. The plan is to build and
operate four pilot plants, which will be located in Ghantoot, Abu Dhabi, UAE.
The pilot plants are expected to be operated on a continuous basis for at least
18 months to demonstrate the reliable performance of the developed
technologies.
• Stage 2 (after 2017): Implementation of the developed energy-efficient desali-
nation technologies in large-scale, fully renewable energy powered seawater
desalination plants.
Stage 1 comprises the design, engineering, construction, testing, operation, and
evaluation of the pilot plants to develop advanced and innovative energy-efficient
seawater desalination technologies suitable to be powered by renewable energy
sources. The pilot plants are required to be reliable and suitable for conditions found
in the UAE.
Stage 1 foresees the development and demonstration of two different categories
of seawater desalination technologies:
• Advanced seawater desalination technologies; and
• Innovative seawater desalination technologies.
Advanced desalination technologies are defined as technologies that are based
on commercially proven and used desalination systems (using thermal, membrane,
or other processes) and that are further optimized and improved as part of the scope
of the Program to reduce the specific energy consumption of the system and to
enhance its compatibility for coupling with renewable energy sources.
The developed advanced technologies are expected to have the potential to lower
the specific energy consumption to reach the following target values:
• Reverse osmosis plants: less than 3.6 kWh/m3 of electric energy;
• Thermal desalination plants: less than 1.0 kWh/m3 of electric energy and a
specific thermal energy consumption varying in accordance to the temperature
of the inlet energy stream as demonstrated in Fig. 11.1.
214 M. El Ramahi

Fig. 11.1 Target specific thermal energy consumption of thermal desalination plants using
advanced desalination technologies in dependence of the temperature of the energy supply

Fig. 11.2 Target specific thermal energy consumption of thermal desalination plants using
innovative desalination technologies in dependence of the temperature of the energy supply

Innovative desalination technologies are defined as technologies that are based


on novel concepts for seawater desalination that are not necessarily based on
commercially proven or used processes.
The developed innovative desalination technologies shall have the potential to
lower the specific energy consumption to reach the following target values:
• Membrane based processes: less than 3.1 kWh/m3 of electric energy;
• Thermal desalination processes: less than 1.0 kWh/m3 of electric energy and a
specific thermal energy consumption varying in accordance to the temperature
of the inlet energy stream as shown in Fig. 11.2.
Figures 11.1 and 11.2 show the target specific thermal energy consumption of
thermal desalination plants in dependence of the temperature of the energy supply.
11 Case Study: Masdar Renewable Energy Water Desalination Program 215

The targeted energy values have been constructed considering the exergetic value
of the heat, the limited capability of thermal desalination systems to achieve high
performance ratios at low temperatures, and the economic value of the heat supply.
The target values indicated in Figs. 11.1 and 11.2 are intended to encourage not
only reducing the specific energy intensity of the desalination process but also using
low grade thermal sources (below 70 °C) where heat is often dispersed into the
environment with harmful impact.
The target values refer to an annual average and are based on feed water with a
concentration of total dissolved solids of 42,000 mg/L and a temperature of 30 °C.
It is further noted that the energy consumption values refer to the complete
desalination process and include the energy consumption of the feed water intake,
pre-treatment, desalination, post-treatment, remineralization, brine discharge, and
all other auxiliary processes.
For the second stage of the Program the following solutions are envisaged:
• Use of one or more large-scale seawater desalination plants using the highly
energy-efficient desalination technologies that are developed and successfully
demonstrated in the first stage of the Program. The desalination plants shall be
powered by renewable energy, which may be:
– Geothermal;
– Solar ponds;
– Solar thermal collectors;
– Photovoltaic systems;
– Concentrated solar power;
– Wind turbines;
– Ocean energy; or
– A combination of the above listed renewable energy sources.
The electrical energy consumption of the desalination plant shall be provided by
renewable power plants. These power plants do not necessarily need to be
co-located or directly coupled with the desalination plant. Only renewable resources
(geothermal or solar) shall be used as the thermal energy input.

11.4 Selection Process

In February 2013, Masdar invited 180 different companies and organizations in the
water desalination industry to participate in the Program. Masdar received 24
submissions and assessed them in collaboration with ILF Consulting Engineers.
The evaluation process was conducted in two phases: qualification, followed by
down-selection of four technologies. The evaluation process was conducted over a
period of eight months. The most important selection criteria were the cost of
produced water (assuming 100 % renewable electric and thermal energy supply),
energy efficiency, environmental impact, quality, health, safety, and environmental
216 M. El Ramahi

(QHSE) compliance, innovativeness, engineering competency of the bidder, project


delivery, cost sharing, and openness for collaboration with the Masdar Institute of
Science and Technology on related R&D.
Given the innovative nature of the technologies offered and the dynamism of the
technological solutions in the medium to long terms, the evaluation criteria also
included the projected development outlook for the proposed desalination tech-
nology in terms of cost and energy intensity, as well as the potential for industrial
application in terms of bankability.
After thorough evaluation of all received proposals, Masdar decided to partner
with well-known players in the desalination industry (Abengoa Water, Suez, and
SIDEM/Veolia) and an innovative Californian start-up company, Trevi Systems.
The selected technologies have not yet been commercialized, and do not exist on
a utility scale anywhere in the world. They have however, passed research, lab
testing, modeling, and prototyping.

11.5 Site for the Pilot Plants

The site of a decommissioned desalination plant in Ghantoot was selected as project


site due to accessibility to deep seawater, availability of electricity, and to natural
gas. Ghantoot is a coastal place around 65 km northeast of the city of Abu Dhabi
and around 60 km southwest of Dubai in the Emirate of Abu Dhabi. A satellite
view of the site is depicted in Fig. 11.3.
The seawater quality in Ghantoot is of good quality with a concentration of total
dissolved solids of around 45,000 mg/L.

11.6 Pilot Plants

The four desalination pilot plants will be built at the same site, and auxiliary
facilities will be shared among the four plants as shown in Fig. 11.4.
The auxiliary facilities will include seawater abstraction, seawater break tank,
brine discharge, supply infrastructure for gas or heat, electricity, water and sewage,
potable water export line, water tank, mobile office, telecommunication, and related
infrastructure.
The electricity consumed by the pilot plants will be supplied by the national grid.
The heat for the desalination pilot plants will be provided by a central boiler unit,
which is fired by natural gas.
Each desalination pilot plant comprises the equipment to desalinate the feed
water and will include the required pre-treatment, desalination unit, post-treatment,
and all the related piping, valves, pressure vessels, pumps, etc.
Each desalination plant will deliver potable water in accordance with Abu
Dhabi’s potable water quality standards as released by the Abu Dhabi Regulatory &
11 Case Study: Masdar Renewable Energy Water Desalination Program 217

Fig. 11.3 Satellite view of Ghantoot

PILOT PLANT

AUXILIARY DESALINATION AUXILIARY


FACILITIES PLANT FACILITIES

PRE-
TREATMENT

SEAWATER
NETWORK

PUMP
SEAWATER
INTAKE
BREAK
TANK

DESALINATION
BRINE UNIT
DISCHARGE

POTABLE
WATER
REMI TANK
PLANT
POTABLE
WATER

ELECTRICITY HEAT
SUPPLY SUPPLY

AUXILIARY
FACILITIES

Fig. 11.4 Schematic layout of a pilot plant


218 M. El Ramahi

Fig. 11.5 Rendering of the pilot plant facility

Supervision Bureau. The water will be pumped to an intermediate tank provided by


Masdar. The intention is that the product water will then be used for interconnection
with the Abu Dhabi potable water network by means of potable water transfer
pumps.
Each desalination plant will be designed, engineered, constructed and operated
by the selected technology provider.
Figure 11.5 depicts the envisioned setup of the pilot plant facility in Ghantoot.

11.7 Selected Partners

Masdar selected the companies as technology providers for the Program


(Table 11.1).
The key features of the desalination technologies span improved pretreatment
processes to the use of novel components for energy recovery and brine manage-
ment, all geared towards lowering overall energy consumption of the desalination

Table 11.1 Companies selected as the technology providers for the program and their projected
capacities
Partner Technology Capacity [m3/d]
Abengoa Reverse osmosis + membrane distillation 1080
Suez Reverse osmosis + ion exchanger 100
SIDEM (Veolia) Reverse osmosis 300
Trevi systems Forward osmosis 50
11 Case Study: Masdar Renewable Energy Water Desalination Program 219

process. This shall allow desalination processes to be cost-effectively powered by


renewable energy sources, thereby reducing the dependence on natural gas for
water production. The selected technology mix includes both advanced and inno-
vative desalination technologies.

11.8 Project Time Schedule

The time schedule for the program is outlined in Table 11.2.

11.9 Operation Regime and Expected Results

The desalination pilot plants will be operated over 18 months.


The assessment of the performance of the pilot plants will follow a rigorous
testing regime that has been set up in advance based on a stage-gate process. The
mode of operation and time-line is outlined in Table 11.3.
The initial two months of operation are dedicated to the start-up and initial trials.
The operator of the pilot plant shall not have any obligation to dispatch the product
water to the water network during this period. However, it is expected that during

Table 11.2 Milestones as Milestone Date


projected by each company
for the project Installation of auxiliary facilities Q1/2015
Installation of Trevi System plant Q2/2016
Installation of Abengoa plant Q4/2015
Installation of Suez plant Q4/2015
Installation of SIDEM/Veolia plant Q3/2015
Completion of operation of desalination plants Q4/2017

Table 11.3 Duration of Pilot plant operation time Operating mode Duration
operating modes for
assessment of pilot plants Months 1–2 Initial set-up
Trial run 3 days
End of 2nd month Acceptance test 12 h
Month 3 Availability test 30 days
Months 4–12 Operation period
End of 6th month Performance test 12 h
End of 9th month Performance test 12 h
End of 12th month Completion test 7 days
Months 13–18 Optimization
period
220 M. El Ramahi

this period the average operational availability will be at least 30 % of the design
capacity.
At the completion of this initial setup period, the desalination pilot plant shall
achieve a reliable operational pattern at the design capacity. Upon completion of the
initial setup period, the operator of the pilot plant shall undertake a three-day trial
run where the plant shall run continuously at full capacity. Successful completion of
this trial run will provide eligibility to the operator of the pilot plant to proceed to
the acceptance test.
During the acceptance test, the desalination pilot plant shall be operated with at
least 100 % of its design capacity during a period of 12 h. If Masdar deems the
daily plant availability, all performance data, and water quality data satisfactory, the
operator may proceed to the next stage of the project; the availability test.
The purpose of the availability test is to demonstrate over a period of 30 days
that the desalination pilot plant performs in a reliable manner and achieves the
anticipated performance parameters as specified by the technology provider in the
tender phase of the Program. In this respect, the operator of the pilot plant shall
conduct performance monitoring, including water quality monitoring, throughout
the availability test period using the on-line instrumentation, along with spot
sampling for off-line laboratory analysis.
It is only after passing the availability test that the pilot plant will be allowed to
enter into the phase of operation and dispatch. During this period, which shall last
nine months, the desalination pilot plant shall be operated with a minimum avail-
ability of 80 %. The measured performance parameters of the pilot plant shall be
within the anticipated performance parameters. Upon completion of the operation
period, a six-month optimization period shall give the operator the possibility to
operate the plant in more aggressive conditions without performance and dis-
patching obligations in order to enable further optimization of the process and pilot
plant performance.

11.10 Related R&D Programs

Concurrently with the pilot projects in Ghantoot, each of the four selected tech-
nology providers will perform an R&D project in collaboration with the Masdar
Institute of Science and Technology to support the pilot project in Ghantoot.
One of the R&D projects will involve the study and evaluation of membrane
scaling and fouling in membrane distillation modules with the aim of determining
the best ways to mitigate these problems. It is expected that this study will develop
appropriate protocols for membrane cleaning, as well as an evaluation report that
can serve as a troubleshooting catalogue for commercial membrane distillation
plants.
In another R&D project, an optimized design of a full-scale solar energy pow-
ered seawater reverse osmosis (SWRO) plant using the most practical and eco-
nomical photovoltaic and solar thermal energy technologies will be developed.
11 Case Study: Masdar Renewable Energy Water Desalination Program 221

Capacitive de-ionization for treatment of permeates from the first-pass RO is the


focus of the third R&D project. This study will demonstrate the potential for cost
and energy savings in SWRO with the application of capacitive de-Ionization, by
replacing the second-pass RO process in SWRO plants.
The fourth R&D project will focus on the development and testing of
high-temperature forward osmosis membranes and manufacturing techniques.
Expected results from this study are a recipe for composition and structure of
advanced forward osmosis membranes able to withstand elevated temperatures.
Chapter 12
Summarizing the Story

Benno Böer and Zafar Adeel

Abstract This volume investigates the need for a more open and inter-disciplinary
dialogue on the nexus of food security, water security, and energy security in the
Arab Region. The water security challenges of the Arab Region are well known and
well documented, including by some of the authors in this volume. The water
security concerns link directly with the achievement of food and energy security but
more importantly, also to broader regional security and peace. It has been argued in
a number of chapters that achieving sustainable economic development is tied to the
security of the water-energy-food nexus, which is in turn essential for achieving
sustained peace. Achieving these lofty objectives is much more challenging and a
discussion of various approaches is provided throughout the volume; this chapter
intends to provide a broader overview.

 
Keywords Water-security Food-security Energy nexus in the Arab Region 

Regional security and peace Achieving sustainable economic and environmental
 
development UN Sustainable Development Goals Shift in water and population

policies Synchronization and inter-disciplinary approaches on government man-
agement level

B. Böer (&)
UNESCO Liaison Office in Addis Abeba with the African Union and UNECA,
UNESCO, Addis Abeba, Ethiopia
e-mail: b.boer@unesco.org
Z. Adeel
Pacific Water Research Centre, Faculty of Environment, Simon Fraser University,
Technology and Sciences Complex 2, Room 8800, 8888 University Drive,
Burnaby BC V5A 1S6, British Columbia, Canada

© Springer International Publishing AG 2017 223


K. Amer et al. (eds.), The Water, Energy, and Food Security Nexus
in the Arab Region, Water Security in a New World,
DOI 10.1007/978-3-319-48408-2_12
224 B. Böer and Z. Adeel

12.1 Emerging Context

The contributing authors of this volume highlight the well-known water challenges
in the Arab Region, but focus really on identifying innovative ways to achieve
sustainable economic and social development—well-being of people and regional
peace remain at the heart of these discussions.
In putting together the various chapters in the volume, great care was taken to
consider the recent major developments at the international arena; for example, the
launch of the new Sustainable Development Goals (SDGs) as part of the 2030
Agenda for Sustainable Development, which have superseded the former
Millenium Development Goals. The 2030 Agenda brings a set of 17 new goals, all
inter-connected, and each applicable universally to all UN member states. These 17
SDGs, combined with 169 underlying targets, provide a new roadmap for the next
15 years. This volume relates to most, if not all of the SDGs in one way or another,
and in particular to the three SDGs included in the title of this volume, that is SDGs
numbered 2, 6, and 7 (SDG 2: end hunger, achieve food security and improved
nutrition, and promote sustainable agriculture; SDG 6: ensure availability and
sustainable management of water resources and sanitation for all; and SDG 7:
ensure access to affordable, reliable, sustainable, and modern energy for all).

12.2 Origins of This Volume

In the United Nations system, there are a number of organizations, programmes,


and funds that work towards water, food, and energy security. UNESCO has a clear
mandate in the sciences, ecology, water, and education, all of which are essential
ingredients to achieving security around the water-energy-food nexus. The United
Nations University (UNU) is mandated to serve as a global community of scholars
that can understand, address, and respond to global challenges; in doing so, it
operates like a think tank that serves as a bridge connecting the research and
scientific community to those involved in policymaking in governments and the
broader UN system. Generally speaking, the broad mandates afford these two
partner organizations an excellent opportunity to focus on the regional challenges
facing the Arab Region.
UNESCO’s Science Report—towards 2030 has been published (UNESCO
2015)1 in support of achieving the 2030 Agenda for Sustainable Development. It
provides comprehensive global information and data on Member States, including
the Arab Region. It provides the scientific community with country profiles for
Algeria, Bahrain, Egypt, Iraq, Jordan, Kuwait, Lebanon, Libya, Mauretania,
Morocco, Oman, Palestine, Qatar, Saudi Arabia, Sudan, Tunisia, Syria, the United
Arab Emirates, and Yemen. Similarly, UNU’s Institute for Water, Environment and

1
http://unesdoc.unesco.org/images/0023/002354/235406e.pdf.
12 Summarizing the Story 225

Health (UNU-INWEH) has published two policy reports on challenges faced in


achieving the water-related SDG targets, and approaches to potentially overcome
these challenges.
This new volume provides comprehensive assessments of the current
water-security, food-security, energy-security nexus in the Arab Region. It also
provides a realistic outlook into likely development scenarios, and, most impor-
tantly, it provides science-based guidance on best practices that need to be urgently
implemented in order to achieve the SDGs. In other words, it shows what needs to
be done and it provides guidance to the UN member states in the Arab Region to
move from rhetoric to action. To state the obvious, the sooner the best practices are
being implemented, the better the outcome.

12.3 Highlights of Contributions

The individual chapters are written by some of the best known scientific scholars in
the region, supported by a highly experienced editorial team, including those from
UNESCO and UNU. We provide here a brief synopsis of each chapter to orient the
reader.
Walid Al Zubari compiled a chapter on Status of Water in the Arab Region. In
summarizing the overall status of water resources, he highlights the extremely poor
supply in the Arab Region—which will likely be exacerbated in terms of the impact
from climate change on the average regional water distribution. The chapter
explores the unpredictable rainfall, population size, and growth, and the dangerous
decline in per capita water availability, despite continuous efforts to improve the
situation. He draws a picture of an alarming future of increasing water scarcity and
increasing water demands and supply costs, and recommends that a major review
of, and shift in water policies and population policies is urgently needed. This
overview by Al Zubari sets the stage for subsequent analysis of the consequences
for energy and food security.
Kishan Khoday and Stephen Gitonga provide a parallel analysis through their
chapter on the State of Energy in the Arab Region. These experts argue that the
Arab Region is currently going through one of the most intense periods of trans-
formation in its history, with a convergence of political, security, economic, social,
and environmental drivers of change. They underscore the fact that energy remains a
key factor in the region’s development debates, with a view to poverty reduction,
food and water security, and social vulnerability. They provide some insights into
how the changing demands and market pressure on the energy sector serve as
feedback loops impacting the economic and social development in the Arab Region.
Tariq Alzadjali presents a chapter pertaining to the third element of the nexus
and reviews the Status of Food Security in the Arab Region. He provides
science-based data on how most Arab countries have made progress in enhancing
their food security situation in the last few years. However, he also provides data on
the Arab countries’ Global Hunger Index, with most of them being classified as
226 B. Böer and Z. Adeel

‘poor’, ‘low’, ‘serious’, and even ‘alarming’. The author predicts that these coun-
tries are unlikely to achieve high ratios of food self sufficiency, particularly because
of a scarce and diminishing water support, and yet there remains hope for
improvement.
Hani Sewilam and Peter Nasr provide a chapter on the highly important con-
temporary aspects of using Desalinated Water for Food Production in the Arab
Region. These experts point out that the prospects of desalination constitutes a
priority for consideration and action in terms of irrigation-water for
food-production. At the same time, they warn about the high associated costs, and
argue for the need for costs reduction of desalination technology in the future. They
also highlight that economic, environmental, and social issues should be at the
forefront of planning any food production initiative using desalinated water. In
particular, they highlight the need for the best environmental practices and the need
to learn from historic failures.
John Bryden elaborates on Food, Energy, and Water in the Arab Region:
Challenges and Opportunities, with Special Emphasis on Renewable Energy in
Food Production. He argues that the great deficit of water and arable land is likely
to be exacerbated by several factors, such as climate change, land degradation,
growth in population, urbanization, incomes, and increased demands driven by
desalination and air conditioning. He suggests that urgent attention needs to be
given to the development of renewable energy as an opportunity in the region.
Atef Hamdy explores the Water, Energy, and Food Security in the Arab
Region: Regional Cooperation and Capacity Building. He points out that in most
Arab Countries, in existing policy frameworks, energy and water policies are
developed in isolation from one another. A better understanding of the links
between energy, water, and food security is essential in any attempt to formulate
policies. This requires urgent attention to building relationships and linkages
between policy making institutions of the three sectors; this will result in having
informed decisions.
Khaled Abu Zeid shines some light on the Research and Development to Bridge
the Knowledge Gap. He highlights the need for extensive analysis of the water,
energy, and food security linkages, to assist in policy making related to water, food,
and energy planning. His chapter presents data and some of the research and
development results on key policy elements related to the WEF Nexus in the Arab
Region. The chapter addresses research needs for better policy decisions in selecting
the appropriate water resource, the appropriate water use, the appropriate water and
energy savings measures, and the appropriate food export-import strategies.
Paul Sullivan provides Thoughts and Policy Options on Water-Energy-Food
Security Nexus in the Arab Region. One of his most important statements is that
the energy, water, and food availability shows not only stress today, but potentially
much greater stresses in the future. He urges the need to integrate policies across
energy, food, and water. Education, outreach, training, investment, and other pro-
grams need to be developed to help lead the people and their leaders to a better
future, and to avoid resource, economic, and political disasters. Reason and mod-
eration will need to be defining characteristics of these policy changes. He warns
12 Summarizing the Story 227

about extreme resource stresses in the future beyond what has been observed to date
in the region. These, in turn, could cause increased energy insecurity, increased
terrorism, and increased migration.
Zafar Adeel provides long-term prospectives in his chapter on Managing Water,
Energy, and Food for Long-Term Regional Security. His chapter presents a
contemporary and robust definition of regional security that encompasses the flow
of resources, sustainable economic development, poverty reduction, and peaceful
co-existence. The chapter focuses on the role played by water, food, and energy in
regional security, highlighting the risky ‘youth bulge’ and widespread youth
unemployment. He warns about the associated rise in extremist ideologies, as well
as the lack of adequate environmental governance. Mastering the water-energy-
food nexus includes not only the development and application of renewable energy
sources, but it needs to convince policymakers and governments in the Arab Region
that the water-energy-food nexus is central to regional security.
Rabi Mohtar, Amjad Assi, and Bassel Daher present a chapter on the Current
Water for Food Situational Analysis in the Arab Region and Expected Changes
due to Dynamic Externalities. The authors propose that decision makers must
revert from reactive decision-making to a new pro-active approach, taking into
account the complicated internal and external dynamics of the region. The chapter
explores the uncertainties associated with predicting the future and it concludes
with a vision for adaptive management approaches. It offers a vision for increased
resilience of local communities by discussing ways to broaden security by local-
izing water and food securities.
Mohammad El Ramahi presents a Case Study: Masdar Renewable Energy
Water Desalination Program. Masdar launched a renewable energy desalination
pilot program to research and develop energy-efficient, cost-competitive desalina-
tion technologies that are suitable to be powered by renewable energy in the wider
MENA region. The commercial scale application of such technologies would also
need to take into consideration the involvement and importance of the private
sector; protection of public interests obviously constitutes an important dimension
of any future engagement of the private sector.

12.4 Key Take-Away Points

There are four broad take-away messages from the book.


First, the threats and risks are here and not just in the distant future. All the
authors highlight the obvious fact: the Arab Region is under considerable water
stress, and the situation will continue to get worse in tandem with a number of
global changes—most notably to climate and the related regional water distribution.
Interestingly, the authors no longer talk about an emerging water-crisis, but about a
real and present dangerous situation of water-scarcity, linked to insecurity of food
and energy provisioning, as well as threats to sustainable living conditions and
human well-being. The consequence of such insecurity are very obvious and, as has
228 B. Böer and Z. Adeel

been observed in the case of Syria, quite detrimental to the overall regional
geopolitical stability and peace. In other words, no country can afford to view their
water-energy-food challenges in isolation, but rather have to analyze and respond to
the challenges collectively.
Second, viable solutions are available in the Arab Region and can be imple-
mented through innovative policies, judicious use of new technologies, and ener-
gizing public opinion. A key ingredient to the success of these solutions is regional
cooperation and pooling together of resources, as argued by both Hamdy and
Sullivan. While there are a number of emerging collaborations, these tend to be
focused on a singular dimension—say either energy or water—and rarely ever
venture into a consolidated vision. Two other related challenges also need to be
overcome: (a) provisioning of sufficient financial resources for effective imple-
mentation, often not achieved due to inappropriate prioritization at the national and
regional level; and (b) producing enabling policies that afford the opportunity for
these innovative solutions to be implemented at large scale (read: national or
regional scale). In view of the latter, excellent solutions often remain stuck at the
pilot scale, and do not achieve effective implementation.
Third, integration across water, energy, and food sectors is obviously needed but
achieving it in practice will be very challenging. The Arab Region is not alone in
grappling with this problem. Equally, most of the developed nations elsewhere have
encountered bureaucratic inertia and deeply-rooted turf protection as major hurdles
towards achieving a common, cohesive vision of jointly managed water, energy,
and food. Budgeting and planning at the national scale continues to be silo-driven,
with little thought put into achieving synergies through a common vision and
implementation. In our opinion, there lies a great opportunity for the Arab Region
to lead by example. The 2030 Agenda for Sustainable Development offers a
common framework and template around which national development policies can
be formulated in an integrated manner. In the same way, the SDGs open the door to
jointly achieving the underlying targets at the regional scale.
Fourth, there are some gaps in scientific knowledge but at the same time there is
a wealth of data and synthesized information that can guide decision-making. There
is always the need of conducting more research and collecting more data. In the
Arab Region, there may be sufficient quantities of data available in the water,
energy, and food sectors but little effort has been put into collective analysis of
various future scenarios. Mohtar et al. discuss ways of addressing these gaps and
connecting the Arab Region to the rest of the world when analyzing development
pathways. Similarly, regional scale impact of climate change and its consequence
for the water-energy-food nexus is now being undertaken, and a lot more practical,
policy oriented information is still required; that can be a target for the research and
policy community. Research institutions like the Masdar Institute offer a lot of
promise for addressing these knowledge gaps. Having said all of this, however, the
available data and information are sufficient to formulate broad policy and strategic
visions at the regional level. A number of forums exist at the regional scale that can
be utilized to form a common vision, allocate resources, empower through enabling
policies, and providing incentives for success.
12 Summarizing the Story 229

We hope that this volume, rich with contributions from experts on the Arab
Region, offers timely advice and guidance. It is also meant to serve as a resource
that points to repositories of knowledge that the policy- and decision-makers can
tap. Such collective, regional-scale visioning is intended to contribute to a new era
of water, energy, and food security. What is very promising is that the dialogue has
become much more open, integrated, and inter-disciplinary. These are indeed the
first steps towards a sustainable and peaceful future for the Arab Region.

Вам также может понравиться