Sunteți pe pagina 1din 309

Edited by:

Nicola Lamaddalena, Muhammad Shat anawi ,


Mladen Todorovic, Claudio Bogliotti, Rossella Albrizi o
SERIES B: Studies and Research
Number 57
Water Use Efficiency
and Water Productivity
WASAMED Project
(EU contract ICA3-CT-2002-10013)
Proceedings of 4th WASAMED Workshop
W
a
t
e
r

U
s
e

E
f
f
i
c
i
e
n
c
y

a
n
d

W
a
t
e
r

P
r
o
d
u
c
t
i
v
i
t
y

W
A
S
A
M
E
D


P
r
o
j
e
c
t
2007
O
p
t
i
o
n
s

M

d
i
t
e
r
r
a
n

e
n
n
e
s
Number
B57
CIHEAM
CIHEAM / IAMB -EU DG Research

CIHEAM
Centre International de Hautes Etudes
Agronomiques Mditerranennes
Secrtariat Gnral / Secretary General: Bertrand HERVIEU
International Centre for Advanced
Mediterranean Agronomic Studies
Prsident / : Moun HAMZ Chairman
11, rue Newton
75116 PARIS
Tel. + 33 1 5323 9100 - Fax + 33 1 5323 9101
Email : secretariat@ciheam.org
IAM
Instituts Agronomiques Mditerranens
Mediterranean Agronomic Institutes
Bari - Chania - Montpellier - Zaragoza
IAM - Bari
IAM - Chania
IAM - Montpellier
IAM - Zaragoza
Directeur : Cosimo LACIRIGNOLA
Via Ceglie 9 - 70010 Valenzano, Bari, ITALIE
Tel. + 39 080 4606 111 - Fax + 39 080 4606 206
Email : iamdir@iamb.it
www.iamb.it
Directeur : Alkinoos NIKOLAIDIS
P.O. Box 85 - 73100 Chania, Crete, GREECE
Tel. 30 2821 03 50 00 - Fax 30 2821 03 50 01
Email : alkinoos@maich.gr
www.maich.gr
Directeur : Vincent DOLL
3191, route de Mende
34033 Montpellier Cedex 5, FRANCE
Tel. 33 04 6704 60 00 - Fax 33 04 6754 25 27
Email : thirion@iamm.fr
www.iamm.fr
Directeur : Luis ESTERUELAS
Apartado, 202 - 50080 Zaragoza, SPAIN
Tel. 34 976 71 60 00 - Fax 34 976 71 60 01
Email : iamz@iamz.ciheam.org
Www.iamz.ciheam.org



Options Mditerranennes, Sries B n. 57








Water Use Efficiency
and Water Productivity



Proceedings of 4
th
WASAMED
(WAter SAving in MEDiterranean agriculture)
Workshop
Amman (Jordan), 30 Sept. - 4 Oct. 2005
EU contract ICA- CT - 2002- 10013

2

















































Opinions, data and facts exposed in this number are under the responsibility of the
authors and do not engage either CIHEAM or the Member-countries.


Les opinions, les donnes et les faits exposs dans ce numro sont sous la responsabilit
des auteurs et nengagent ni le CIHEAM, ni les Pays membres.


CIHEAM
Centre International de Hautes Etudes Agronomiques
Mditerranennes
Options
Directeur de la publication: Bertrand Hervieu




SERIES B: Studies and Research
Number 57

Water Use Efficiency
and Water Productivity
Proceedings of 4
th
WASAMED
(WAter SAving in MEDiterranean agriculture)
Workshop
Amman (Jordan), 30 Sept. 4 Oct. 2005
EU contract ICA- CT - 2002- 10013

Edited by:
Nicola Lamaddalena, Muhammad Shatanawi,
Mladen Todorovic, Claudio Bogliotti, Rossella Albrizio






2007

mditerranenne
4
La maquette et la mise en page de ce volume de Options
Mditerranennes Sries B
ont t ralises lAtelier dEdition de lIAM Bari
This volume of Options Mditerranennes Series B has been
formatted and paged up by the IAM Bari Editing Board








Water Use Efficiency
and Water Productivity



Proceedings of 4
th
WASAMED
(WAter SAving in MEDiterranean agriculture)
Workshop
Amman (Jordan), 30 Sept. - 4 Oct. 2005
EU contract ICA- CT - 2002- 10013


Edited by:
Nicola Lamaddalena, Muhammad Shatanawi,
Mladen Todorovic, Claudio Bogliotti, Rossella Albrizio



Bari: CIHEAM
(Centre International de Hautes Etudes Agronomiques Mditerranennes)
p. 294, 2007
Options Mditerranennes, Sries B, N. 57



ISSN : 1016-1228
ISBN : 2-85352-355-1



CIHEAM, 2007
Reproduction in whole or in part
is not permitted without the consent
of Options Mditerranennes
Reproduction partielle ou totale
interdite sans lautorisation
d Options Mditerranennes

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
1
CONTENTS

FOREWORD........................................................................................................................... 3
INTRODUCTION .................................................................................................................... 5


KEYNOTE PAPERS

WATER USE EFFICIENCY IN IRRIGATED AGRICULTURE: AN ANALYTICAL REVIEW. 9
A. HAMDY
FUTURE OPTIONS AND RESEARCH NEEDS OF WATER USES FOR SUSTAINABLE
AGRICULTURE.................................................................................................................... 21
M.R. SHATANAWI
RELATING WATER PRODUCTIVITY AND CROP EVAPOTRANSPIRATION................... 31
L.S. PEREIRA
SYSTEMATIC APPROACH TO THE IMPROVEMENT OF AGRICULTURAL WATER USE
EFFICIENCY......................................................................................................................... 51
T.C. HSIAO
ON THE CONSERVATIVE BEHAVIOR OF BIOMASS WATER PRODUCTIVITY ............ 59
P. STEDUTO, T.C. HSIAO, E. FERERES
TECHNICAL INTERVENTIONS TO IMPROVE WATER USE EFFICIENCY IN IRRIGATED
AGRICULTURE.................................................................................................................... 63
A. HAMDY


COUNTRY REPORTS

TECHNIQUES FOR IMPROVING WATER USE EFFICIENCY IN GREENHOUSE
CULTIVATION IN CYPRUS ................................................................................................. 74
P. POLYCARPOU, D. CHIMONIDOU, I PAPADOPOULOS
REVIEW AND ANALYSIS OF WATER USE EFFICIENCY AND WATER PRODUCTIVITY
IN EGYPT ............................................................................................................................. 82
M. NASR ALLAM, R. ABDEL-AZIM
WATER USE EFFICIENCY AND WATER PRODUCTIVITY IN GREECE.......................... 92
A. KARAMANOS, S. AGGELIDES, P. LONDRA
IRRIGATED AGRICULTURE AND WATER USE EFFICIENCY IN ITALY....................... 102
M. TODOROVIC, A. CALIANDRO, R. ALBRIZIO
EFFECTS OF DEFICIT IRRIGATION ON YIELD AND WATER USE EFFICIENCY OF
SOME CROPS UNDER SEMI-ARID CONDITIONS OF THE BEKAA VALLEY OF
LEBANON.......................................................................................................................... 139
F. KARAM, K. KARAA, N. TARABEY

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
2
WATER USE EFFICIENCY AND WATER PRODUCTIVITY IN MALTA .......................... 156
G. ATTARD, J. MANGION, P. MICALLEF
IRRIGATION RESEARCH RESULTS IN THE SYRIAN ARAB REPUBLIC..................... 166
A. KAISI, Y. MOHAMMAD, Y. MAHROUSEH
WATER USE EFFICIENCY IN TURKEY............................................................................ 178
R. KANBER, M. UNLU, E.H. CAKMAK, M. TUZUN


OTHER CONTRIBUTIONS

WATER USE EFFICIENCY IN C
3
CEREALS UNDER MEDITERRANEAN CONDITIONS: A
REVIEW OF SOME PHYSIOLOGICAL ASPECTS............................................................ 192
E.A. TAMBUSSI, J. BORT, J.L. ARAUS
PRODUCTIVITY OF THE POTATO CROP UNDER IRRIGATION WITH LOW QUALITY
WATERS ............................................................................................................................ 208
N. BEN MECHLIA, K. NAGAZ, J. ABID-KARRAY, M.M. MASMOUDI
USE OF THE HEAT DISSIPATION TECHNIQUE FOR ESTIMATING THE
TRANSPIRATION OF OLIVE TREES................................................................................ 215
J. ABID KARRAY, M.M. MASMOUDI, J.P. LUC, N. BEN MECHLIA
WATER RESOURCES MANAGEMENT AT THE RIVER BASIN LEVEL: AN
INSTITUTIONAL ANALYSIS.............................................................................................. 221
A. BILLI, A. QUARTO, E. ZINI
THE ECONOMICS OF WATER EFFICIENCY: A REVIEW OF THEORIES,
MEASUREMENT ISSUES AND INTEGRATED MODELS ................................................ 231
A. BILLI, G. CANITANO, A. QUARTO
PROPOSAL FOR THE INTEGRATION OF IRRIGATION EFFICIENCY AND
AGRICULTURAL WATER PRODUCTIVITY...................................................................... 269
B. BLMLING*, H. YANG**, C. PAHL-WOSTL
EFFECTIVENESS OF INDICATORS FOR SUSTAINABLE WATER USE IN
AGRICULTURE.................................................................................................................. 287
G. ZEROL




Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
3

FOREWORD


The escalating water crisis in arid and semi-arid areas of the Mediterranean constitutes a major
threat to global progress towards sustainable development of the entire region. The region is one of
the largest food importers and the forecasts indicate an increase of food import due to strong
population growth and increased demand for limited freshwater resources.

Modern strategies for water saving in the Mediterranean region are focusing awareness on the
demand management of irrigated agriculture, while looking at improvement of water use efficiency,
throughout the whole water path from the source to the irrigation fields, and on a better productivity of
each m
3
of water used in agriculture with the primary objective of water saving and increasing crop
production and income of farmers. This requires considerable efforts in the implementation of the best
management practices in the region which are compatible with the technical, financial and socio-
economic capabilities of the irrigation environment.

In the Mediterranean, the success in water saving strategies depends on the level of
understanding and integration of cultural, economic, institutional and environmental contexts.
Nevertheless, water saving is still below expectations due to the lack of effective regional co-
ordination, communication and dialogue among all the relevant stakeholders, and the lack of a
common-shared knowledge to support formulation of adequate national and regional water saving
programmes and sustainable water policies. In this view, WASAMED (WAter SAving in
MEDiterranean agriculture) project has been envisaged and implemented at the beginning of 2003.

WASAMED is a 48-month Thematic Network (ICA3-CT2002-10013; http://wasamed.iamb.it),
granted by the International Scientific Cooperation of the Directorate General of Research, European
Commission, in the frame of the 5th Framework Programme. The Network is coordinated by
CIHEAM-Bari Institute and involves 42 partners from 16 countries including scientific institutions,
decision and policy makers, researchers, end-users, Water User Associations (WUAs) and NGOs.
One of the main targets of the WASAMED is seeking consensus on best options, goals and indicators
for water saving in the region mainly through the exchange of knowledge, dissemination of
information and realization of five Euro-Mediterranean Workshops and an International Conference by
the end of project.

The Workshop on Water Use Efficiency and Water Productivity, held in Amman (Jordan), in
October 2005, is the fourth of a series of five Thematic Workshops and presents the result of shared
efforts of all the WASAMED partners and, particularly, of the University of Jordan, the National Center
for Agricultural Research and Technology Transfer (NCARTT), and CIHEAM-IAMB which jointly
organized the event.

This volume of Option Mditerranennes includes twenty-one contributions presented in the
Workshop by both the partners of WASAMED and by external speakers, invited on the basis of their
reputation and experience in the subject. In this occasion, I would like to express my deepest thanks
to all the institutions and scientists and water experts contributing in the realization of the Workshop
and this Proceedings. I hope that the papers presented in this volume and the workshop
recommendations can respond to the need of building a regional knowledge on Water Use Efficiency
and Water Productivity and the implementation of best management practices on the ground in the
Mediterranean region.

Finally, on behalf of CIHEAM-Bari Institute and in the name of all institutions involved in the
WASAMED project, I would like to express my sincere appreciations and gratitude to the European
Commission INCO-MED programme for its financial support and valuable work in following the
progress of the project.


Cosimo Lacirignola

Director, CIHEAM-IAM Bari

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
4

























































Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
5

INTRODUCTION

The contemporary world is strongly linked to the economic parameterization and the water-related
efficiency terms are commonly used indicating the level of performance of a system when water is
transported, consumed and/or used in the process of production of a specific good. In each specific
process, the efficiency refers to the output/input ratio and it is usually expressed as a non-
dimensional value ranging between 0 and 1. Throughout the years, several water-related efficiency
terms have been applied to different fields (agriculture, engineering, economy, industry, etc).

From an engineering point of view, the concept of efficiency is mainly linked to the hydraulic
performance of water conveyance and distribution systems. Both indicates the ratio between the
water output at the outlet of the system and water input at the inlet of the hydraulic structure/system
under consideration. Consequently, in the engineering sector are commonly used the terms as: water
conveyance efficiency (the ratio between the volumes of water at the outlet and inlet of a water
conveyance network), water distribution efficiency (the ratio between the volumes of water at the
outlet and inlet of a water distribution network), water storage efficiency (the ratio between the volume
of water diverted for irrigation and the volume entering a storage reservoir for the same purpose),
irrigation system efficiency (the ratio between the volume of water effectively available for crop use
and the volume applied on the field), etc.

In the agricultural sector, the Water-Use Efficiency (WUE) term has been widely in use since the
middle of sixties when Viets has introduced it in his article on Fertilizers and the efficient use of
water (Viets, F.G., 1962, Adv. Agron., 14: 223-264) and on soil management practices (Viets, F.G.,
1966, Increasing water use efficiency by soil management. In Plant environment and efficient water
use, eds. W.H.Pierre et al. Madison, Wisconsin, USA: American Society of Agronomy). Since that
time, the WUE term has become a common tool to describe, at different scales, the relationship
between the crop growth development and the amount of water used. For example, at the leaf scale,
the plant physiologists use the Photosynthetic Water-Use Efficiency referring to the ratio of net
assimilation to transpiration; at the plant (canopy) scale, the agronomists employ both Biomass and
Yield Water-Use Efficiency indicating the ratio between the biomass and yield, respectively, and crop
(evapo)transpiration. Nevertheless, in all these cases, the WUE terms are not non-dimensional values
and they do not represent an output/input ratio of only one entity. In fact, they describe the processes
in which water is consumed and/or used to produce new entities (e.g. biomass, yield, etc.), indicating
the quantity produced per surface area from the unit amount of water. For this reason, several
alternatives have been proposed in the recent years to convert these WUE terms into some others,
more appropriate terms. Among such attempts, the Water Productivity (WP) term is going to assert
and to spread in the agricultural scientific community as it describes better the ratio between the
quantity of a product (biomass or yield) and the amount of water depleted or diverted. Certainly, the
WP term can be used from leaf to plant (canopy) and field scale whereas the choice of both, the
nominator and the denominator of the WP ratio, may vary with the objectives and domain of interest
of the study.

The Water Productivity term plays a crucial role in modern agriculture which aims to increase yield
production per hectare per unit of water used, both under rainfed and irrigated conditions. This
objective can be pursued either increasing the marketable yield of the crops for each unit of water
transpired, or reducing the outflows and the atmospheric water depletion, or enhancing the effective
use of rainfall, of the water stored in the soil, and of the marginal quality water. The first option
concerns to the need for improving crop yield; the second one intends to increase the beneficial
depletion (transpiration) of water supply against the non-beneficial portion (evaporation); the third
aims to utilize efficiently the water resources, further than the water diverted from reservoirs, streams
or groundwater sources. All these principles lead to the improvement of the on-field management
aspects of crop growth, emphasizing the importance of the application of the best crop management
practices which will permit to use less water for irrigation, decrease evaporation losses, optimize
fertilizer supply, pest control, energy consumption, soil conditions, etc. This is of particular importance
in arid and semi-arid regions with limited water supply, where the farmers are frequently constrained
to apply deficit irrigation strategies and to manage water supply in accordance with the sensitivity of
crops growing stages to water stress. In those situations, the economic aspects of WUE would get
additional importance due to farmers interest to improve economic return from the investments in
irrigation water supply.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
6

The 4th WASAMED Workshop on WUE and WP, held in Amman (Jordan) from Sept. 30 to Oct.
4, 2005, was dedicated to both Water-Use Efficiency and Water Productivity. The main objective of
this Workshop were to assess the actual situation of research and agricultural practices related to the
WUE and WP on the leaf, canopy and field scale in the arid and semi-arid areas of the Mediterranean
region and to address the opportunities for water saving in irrigated agriculture. Accordingly, the
Workshop highlighted and discussed the following main issues:
a) the hydrological aspects and agronomic management strategies of WUE and WP (e.g. evapo-
transpiration estimates, crop water requirements, irrigation scheduling and soil water balance),
b) eco-physiological aspects of WUE and WP (e.g. plant-water relationships, crop growth, capture
and/or use efficiency of resources and yield production) and
c) economic aspects of WUE and WP (e.g. water allocation cropping pattern optimization at
different scales and costs of management practices);
d) sustainability of WUE and WP options, accounting social, economic, environmental and policy
dimensions and inter-linkages among them.

This volume of Option Mditerranennes represents only a part of presentations carried out
during the Workshop and consists of twenty-one scientific contributions. In the first part are given six
keynote documents covering different aspects of water use efficiency and water productivity in
irrigated agriculture and indicating future options, research needs and possible technical interventions
to improve water use in agriculture sector. The second chapter is dedicated to the country reports
presented for eight Mediterranean countries (Cyprus, Egypt, Greece, Italy, Lebanon, Malta, Syria and
Turkey). The last part of this volume is related to the research results of several experiments carried
out in the Mediterranean region and the analysis of socio-economic and institutional and policy
aspects of water resources management.

We truly hope that this volume of Options Mditerranennes will contribute in searching for a
consensus on the framework of interventions (measures) to disseminate the best-performing WUE
and WP practices accounting for different socio-economic conditions, environmental scenarios and
scaling aspects characterizing the Mediterranean region.



Nicola Lamaddalena
Muhammad Shatanawi
Mladen Todorovic
Claudio Bogliotti
Rossella Albrizio






Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
7




























KEYNOTE PAPERS































Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
8

























































Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
9

WATER USE EFFICIENCY IN IRRIGATED AGRICULTURE:
AN ANALYTICAL REVIEW



A. Hamdy
CIHEAM - Mediterranean Agronomic Institute-Bari, Italy


SUMMARY The growing water scarcity and the misuse of available water resources are nowadays
major threats to sustainable development for most developing arid and semiarid countries of the
Mediterranean. For most countries of the region, the important role agriculture could play in not only
feeding and clothing burgeoning population is well recognized, but also in increasing the limited
available water supply by reducing water losses and by increasing the water use efficiency in the
irrigation sector. Avoiding water conflicts among the water user sectors, achieving water security and
food security is fundamentally a matter of water use efficient rate in the irrigation sector. The
importance of efficiency in water use clearly varies across regions and nations as well as through
time. The prevailing geographic, economic and social conditions of the nation play an important role
in examining the efficient use of water. In agriculture, water use efficiency may be defined quite
differently by a farmer, a manager of an irrigation project, or a river basin authority. Efficiencies in the
use of water for irrigation consists in various components and takes into account losses during
storage, conveyance and application to irrigation plots. Identifying the various components and
knowing what improvements can be made is essential in making most effective use of this vital and
scarce source in the Mediterranean agriculture areas. This what will be addressed in the presented
analytical review paper.

Key words: water use efficiency, definitions, irrigation.


INTRODUCTION

Water scarcity has been reflected in traditional social and economic systems in arid areas of the
Mediterranean region. During the past thirty to forty years, population growth, urbanization and
economic development have depleted the region's economically exploitable water resources. Water
scarcity, exacerbated by water quality deterioration and the lack of effective water management, has
become a major problem in several arid countries, and even in the humid ones.

Water shortage is not a new phenomenon in the Mediterranean countries. What is new, however,
is that it is occurring in an increasingly changed environment and this makes it more serious and long-
lasting. The most recent droughts in the summers of 1989 and 1990 marked a turning point. They
highlighted the vulnerability of water supplies even in the industrialized northern Mediterranean
countries which had always relied on adequate per capita of rainfall.

Water crisis is endemic or permanent in some southern Mediterranean areas, but it has now even
reached towns and villages in France, Spain, Italy and Greece, obliging them to impose temporary
restrictions. The shortfall in quantity has been compounded by a decrease in quality due to
contamination of surface or underground water.

Why scarcity emerges as a major problem in most of arid and semi-arid countries of the
Mediterranean?
The answer to this question implies several reasons that could be outlined in the following:
water withdrawals reach the physical limits of available natural water resources (Palestine, Israel,
Egypt, Libya and Jordan);
physical conditions make inter-basin transfers or development of deep aquifers to balance
supply and demand very costly, often requiring special provisions for maintenance and
operation combined with uncertainties regarding sustainability of the new sources (Morocco,
Tunisia);
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
10
low water production efficiencies due to low cost of water or weaknesses in the deployment of
recent advances in technology (Algeria); and
loss of affordable potable water because of pollution and environmental degradation.

Regardless of the specific causes, existing institutions are not amenable to cope with the spiralling
increases in aggregate demands for water by municipal, industrial, tourist, and agricultural uses while
preventing pollution.


WHAT IS EFFICIENT WATER USE?

The term water use efficiency originates in the economic concept of productivity. Productivity
measures the same amount of any given resource that must be expended to produce one unit of any
goods or service. Thus, water productivity might be measured by the volume of water taken into a
plant to produce a unit of the output. In general, the lower the resource input requirement per unit, the
higher the efficiency.

In any environmental resource context, however, the efficiency concept must be extended to
include considerations of quality. Any effort to improve water use efficiency should be consistent with
maintaining or improving water quality. Taking both quantity and quality into account, the following
definition applies.

Water use efficiency includes any measure that reduces the amount of water used per unit of any
given activity, consistent with the maintenance or enhancement of water quality.

The importance of efficiency in water use clearly varies across regions and nations, as well as
through time. Geographically, for instance, water availability will condition the manner in which use
patterns develop. Other things being equal, arid and semi-arid regions require a greater efficiency of
water use than humid ones.

Economic conditions will often lead to greater or lesser water use efficiency. Many regions of the
world have been assisted in their development through public financing of water development. While
the benefits or costs of such projects in efficiency terms are often debatable, the main point here is
that economic factors can influence water use efficiency.

Social conditions also play an important role in examining the efficient use of water resources. The
literature reveals many areas where public education has led to conservation and better use of
available water supplies.

Here it is of interest to report some of the definitions for efficient water use presented by several
authors during the International Seminar on "Efficient Water Use", Mexico City (1991).

J. Bau (1991) affirms that the efficient water use consists in optimizing water use.
Walker, Richardson and Seveback (1991) pointed that efficient water use means optimizing water
usage and ensuring efficiency in its use.

Arreguin (1991) stated that efficiency may be obtained by optimizing the use of water and
infrastructures through active participation by users with a sense of social responsibility.

Gloss (1991) indicated that efficient water use should be considered from different points of view.
There is absolute efficiency to use the least amount of water possible, economic efficiency, which
seeks to derive maximum economic benefits, social efficiency which strives to fulfill the needs of the
user community, ecological efficiency which guarantees natural resource conservation, and
institutional efficiency which qualifies the function of an institution regarding its water related tasks.
Depending on the conditions of each user system, these non-exclusive definitions can be achieved
simultaneously.

The abovementioned definitions indicate that an examination of water use efficiency requires a
multi-dimensional approach. In addition to the physical elements, social, economic and environmental
factors must be carefully considered.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
11


WATER USE EFFICIENCY IN AGRICULTURE

In agriculture, water use efficiency may be defined quite differently by a farmer, a manager of an
irrigation project or a river basin authority. For example, on-farm irrigation efficiencies and project
efficiencies may be low, but substantial water losses may infiltrate in the soil, recharge the acquifers
and may be pumped up again for re-use, either in the same project area or in another downstream.
Other losses, such as overland flow, may feed drainage systems or rivers, and may be pumped or
diverted for re-use. By recycling losses, river basin efficiencies could become very high. The water
saving gained from introducing new technologies would be restricted to saving in evaporation losses
from wetted land surfaces and water puddles, and evaporation losses from non-beneficial vegetation,
which may be substantially less than the savings experienced on the farm. Clearly, any water
conservation project should be carefully appraised by using adequate geo-hydrological information to
study the project's effect on the water balance in the river basin.

In water use a distinction should be made between technical efficiency and economic efficiency.
On one hand, technical efficiency may be low in a project area, but may be high in the river basin if
water is recycled. On the other hand, water losses in project area and recycling particularly when high
pump lifts are involved may reduce economic efficiency.

Initial water losses may lead to undesirable effects, such as waterlogging and salinity. A third way
to express water use efficiency is through production per cubic meter of water.


IRRIGATION EFFICIENCIES IN THE MEDITERRANEAN REGION

In the Mediterranean area irrigation represents 72% of the total water withdrawals. Irrigation is
extremely water intensive. It takes about 1000 tons of water to grow one ton of grain and 2000 tons to
grow one ton of rice.

Despite the high priority and massive resources invested in the water resources development, the
performance of large public irrigation systems has fallen short to expectation in developing and
developed countries of the Mediterranean area.

Competitive and inefficient use of limited regional water supplies by irrigated agriculture is a major
threat to sustainability of water supplies. Very often the conveyance losses of conduits (unlined
canals or leak pipes) are much too large, a 30% loss percentage of the available water is common in
irrigation systems.

Another cause of inefficient water use is the emphasis on meeting demand by constructing new
supply facilities rather than improving the efficiencies of the existing ones.

In most countries of the region, significantly more water is delivered per unit area than is required,
leading to low irrigation efficiencies. The area irrigated in many irrigation systems is much less than
the area commanded and annual cropping intensities are lower than anticipated. Water deliveries
rarely correspond in quantity and timing to the true requirements of the farmer's crops, leading to loss
in productivity. In many irrigation systems water is distributed inequitably between farmers near the
head-end reaches of the system, where water is short, and those less fortunate farmers located
downstream.

The quantities of water consumed by the crops in an irrigation project are considerable. But the
volumes of water handled by the project system have to take account of system efficiency, a product
of efficiency during:
(i) conveyance,
(ii) distribution,
(iii) field application.

Average losses in irrigation projects suggest that only about 45% of water diverted or extracted for
irrigation actually reaches the crops. But losses vary widely, those in the conveyance system taking
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
12
the water to the irrigation site may vary between 5 and 50 percent. (Table 1) shows the technical
irrigation efficiencies that could be expected for conveyance and distribution efficiency in the
Mediterranean region. Project efficiencies range from 30 to 65 percent, depending on the
sophistication of the irrigation system and the on-farm irrigation technology in use.

Table 1 - Typical project irrigation efficiencies (in percent)
Category Conveyance
and
Distribution
Field
Application

Project *
A. Large-scale irrigation
1. Traditional open canal system (manual control) (e.g.
Turkey)

60

50

30
2. Open canal systems with hydraulic control and
surface irrigation (e.g. Morocco)

70

60**

40
3. Open canal systems with manual control, on-farm
storage and sprinkler/drip (e.g.Jordan)

75

70

55
4. Open canal systems with hydraulic control, buffer or
on-farm storage and sprinkler/drip

85

70

60
5. Pipe conveyance systems with sprinkler/drip (e.g.
Cyprus)

95

70

65
B. Groundwater irrigation
6. Lined field channels and on-farm surface (gravity)
80

50

40
7. Pipe systems and on-farm sprinkler/drip 95 70 65
Notes:
* Gravity (surface) irrigation on the farm
** Project efficiencies are rounded to nearest 5 percent

Efficiency in the use of water for irrigation consists of various components and takes into account
losses during storage, conveyance and application to irrigation plots.

Identifying the various components and knowing what improvements can be made is essential to
making the most effective use of this vital but scarce source in Mediterranean agricultural areas.


IRRIGATION EFFICIENCY DEFINITIONS

An efficiency is generally defined as the ratio of output over input, and is expressed as a
percentage. In irrigation, efficiency was first defined by Israelsen (1932) as the ratio of irrigation
water transpired by the crops of an irrigation farm or project during their growth period, over the water
diverted from a river or other natural source into the farm or project channel or channels during the
same period of time.

This definition has not essentially changed in the years that have since passed although numerous
definitions of irrigation efficiency were developed (see Appendix). Basically, they can be divided into
three main groups:
Definitions based on measured volumes of water;
Definitions based on measured depths of application;
Definitions based on other criteria, mainly related to yield.


A Definitions based on volumes of water

Most of definitions presented in Appendix are based on ratios of water volumes. These definitions
have the advantage of being relatively easy to use: the volume of water delivered to the soil, a field, or
a distribution system can be measured, and the volume of water stored in the rootzone during the
irrigation or evapotranspired by the crop can be either measured or calculated.

Definitions based on total volumes delivered to or in a system have disadvantage that, for some
efficiencies, the uniformity of application or supply cannot be taken into account: the volume of water
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
13
delivered can match the volume required, but the division of water can be such that a part of the
system or field is under-irrigated and another part is over-irrigated.

B Definitions based on measured depths of application

The ASCE (1978) on-farm efficiency definitions use depths of application. If the application
depths are only average values found by dividing the total volume delivered to the system by the area
there is no difference between these efficiencies and those based on volumes of water. If, however,
the depths are actual values measured at specific representative locations in the field, these
definitions can, after the measured depths have been processed, also take into account the uniformity
of application.

The definitions based on measured depth of field application have the disadvantage that they
cannot be used for a whole system: it is impractical or impossible to measure the depth of application
in all the fields of irrigation system.


C Definitions based on other criteria, mainly related to yield

These definitions are seldom used due to the fact that crop yield are influenced by many factors
other than water supply alone (fertilizer application, treatment of plant diseases etc.). Crop yields are
measure if agricultural production and not of water supply, although managers of irrigation water
supply systems are sometimes judged by crop yields.

Many ratios are presented as efficiencies and are called efficiencies although they are not
efficiencies in the sense of a ratio of output to input, whereby the output (of some quantity) is a
conversion of an input (of the same quantity). Examples are: a deep percolation efficiency (Hart et
al, 1979), the ratio of the volume of water required to fill the available rootzone water storage minus
the deficit to the volume of water absorbed by the soil through infiltration (the ratio, moreover, is
high when the deep percolation is low); an operation efficiency (Schuurmans, 1989), as the ratio
effective volume of water supplied to actually supplied volume of water.


DEFINITIONS RELATED TO THE CONVEYANCE OF WATER

The efficiencies related to the conveyance of water have not changed much since Israelsen (1932)
defined the water conveyance and delivery efficiency. Jensen et al. (1967), Bos and Nugteren
(1974), and ICID (1978) use the same definitions, except where Isrealsen mentions water delivered
to farms (in the numerator), Jensen uses water delivered by the conveyance system, and ICID uses
water delivered to the distribution system.

The distribution system consists of tertiary units, each of which can be one farm only, or can
incorporate many farms. The distribution system is usually under the control of farmers or groups of
farmers. These efficiencies express how much water is delivered from the conveyance system to the
distribution system. The not-delivered water includes operational spills, seepage, and evaporation.
These definitions do not intend to cover the uniformity of the division of water over the various parts of
the distribution system.


DEFINITIONS RELATED TO WATER STORAGE IN THE ROOTZONE

These definitions have the problem that rootzone is not exactly defined. Its dept varies with the
growing stage of the crop, the type of crop, the prevailing groundwater depth, etc.

The term water stored in the rootzone is used by Hansen (1960) in the numerator of the water
storage efficiency: the ratio of water stored in the rootzone during irrigation, over water needed in the
rootzone prior to irrigation. This ration was introduced because of the inability of the Isrealsens water
application efficiency to reflect conditions of under-irrigation. However, rootzone water storage
efficiencies do not cover conditions of over-irrigation: if more water is applied than needed, the excess
is not accounted for.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
14

The three definitions of Hart et al. (1979) deal neither with the conveyance of water nor with crop
growth. These definitions seem to originate from irrigation at field level because the area to which
water is delivered is not explicitly defined. If that area is a field or a farm, the efficiencies do not, for
example, cover operation spills in the system or seepage from conveyance channels.


DEFINITIONS RELATED TO WATER USED FOR PLANT GROWTH

These definitions take into consideration:
a) Volume of water used for plant growth, and
b) Volume of water supplied.


Volume of water used for plant growth

The difficult point in irrigation efficiency is to determine the amount of water used for plant growth.
In the definitions, this amount is expressed in various ways:
irrigation water transpired by the crops (Isrealsen, 1932),
normal consumptive use of water (Hansen, 1960),
useful water applied (Hall, 1960),
volume of water transpired by plants, plus volume of water evaporated from soil (Jensen et al.,
1967),
consumptive use effective rainfall (Erie, 1968),
beneficially applied depth of water (ASCE, 1978), including applications for such purposes as
salt leaching, frost projection, crop cooling,
volume of water needed, and made available, for evapotranspiration by the crop to avoid
undesirable water stress in the plants throughout the growing cycle (Bos, 1980).

The actually used volume of water for plant growth will always be an estimate and it can only be
measured on an experimental scale. Crop transpiration and evaporation of water the soil are usually
combined in evapotranspiration. The reason for the formulation of Bos is that the link between crop
water use and water storage in the rootzone is not easily made. With this definitions, the problem of
how to account for crop water use is shifted from rootzone moisture to evapotranspiration.

A water application efficiency defined as water stored in soil, or in the rootzone, over water
delivered to that rootzone is not useful is not related to evapotranspiration. And it is exactly this
relationship that is made by Isrealsen consumptive use efficiency. The product of this consumptive
use and the water application efficiency is the ratio between irrigation water transpired by the crop
and irrigation water delivered to the farm.

The time period considered is important too. The efficiency of one application at field level,
measured with either volumes or depths, gives no information on an average application efficiency for
a growing season, irrigation season, or for an area other than the measured field. Hall (1960) defined
a season application efficiency to extend his application efficiency to the entire irrigation season.
Israelsen and ICID explicitly mention growth period or growing cycle in their definitions.


Volume of water supplied

When the numerator is concerned with water used for plant growth, the scope of the denominator
determines which efficiency is evaluated. For example Isrealsens water application efficiency
concerns irrigation water delivered to the farm, and Halls application efficiency concerns gross
volume of water delivered to the field. Typically, the broadest scope of volume of water supplied is
the scheme: Jensens irrigation efficiency deals with total volume of water diverted, stored, or
pumped for irrigation, and ICIDs project or overall efficiency deals with volume of water diverted or
pumped from the river.


RELATIVE WATER SUPPLY
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
15

Relative Water Supply (RWS), proposed by Levine (1982), is the ration of water supply to the
demand for water. Defined as input over output, the RWS resembles the reciprocal of an irrigation
efficiency, but there is a difference. Generally, the crop irrigation water requirements is arrived at by
subtracting the effective precipitation to be expected from the crop water requirement (potential
evapotranspiration), whereas in the definition of RWS, the effective precipitation is taken as a water
delivery. The inclusion of seepage and percolation on the demand side (denominator) of the definition
suggests that the RWS ratio has been especially formulated to evaluate rice cultivation. Levine (1982)
uses rainfall in the numerator of the definition, whereas Weller et al. (1988) use effective rainfall.


THE ICID DEFINITIONS AND WATER SHORTAGE

Water shortage usually implies that the supply of water is not enough for crop growth. The
terminology used includes under-irrigation or under-sipply.

In the ICIDs irrigation efficiencies, the crop irrigation water requirements is Vm the volume of
water needed, and made available for evapotranspiration by the crop to avoid undesirable water
stress in the plants throughout the growing cycle. The starting point for Vm is the theoretical crop
irrigation water requirement. However, the definition restricts itself to well-watered conditions. This
conditions are presumed, and if this presumption were to be strictly adhered to, the ICID (crop
related) definitions could be only used for planning and designing irrigation. When Vm is taken as
potential evapotranspiration (ETpot), and Vc is the volume diverted or pumped, the value of ETpot/Vc
can become greater than 100% under conditions of water shortage. Then, it is not longer an efficiency
in the sense of a ratio of output to input. If, however, Vm is taken as actual evapotranspiration (lower
than ETpot), values will be high, although never higher than 100%.

Beside conveyance efficiency, distribution efficiency, field application efficiency and overall or
project efficiency (see Appendix), the ICID also defined a tertiary unit efficiency and an irrigation
system efficiency. They are, when the non-irrigation supplies are neglected, combinations of
respectively the distribution and field application processes and the conveyance and distribution
processes.


EFFICIENCY AND UNIFORMITY

In irrigation, uniformity is used to express the variation in depths of application or supplied
volumes. Christiansen (1942) defined uniformity coefficient CU for the comparison of sprinkler
patterns. ICID (1978) extended the use of this coefficient to cover infiltration water too.

Efficiency and uniformity can be described at different levels in an irrigation system: for a field, for
a tertiary unit, and for a scheme. Within a river basin, the division of flow over successive irrigated
areas can be important, especially if the river flows through several countries or states.


Efficiency and uniformity at field level

Most efficiency definitions concerns volumes only, because these are relatively easy to measure,
but, for a field, the total volume delivered is not an absolute measure of how much water each part of
the field receives. For sprinkler-irrigated fields, uniformity has been frequently investigated with cans
on the field catching the water from the sprinkler. The uniformity of the actual field application is then
assumed to be equal to the application depths measured in the cans. For other methods of field
application (e.g. basin or furrow), uniformity is more difficult to measure and is then usually estimated.
The actual depth of water applied to the soil can be found by measuring the soil moisture content at
several representative locations in the field before and after irrigation. Because of the practical
limitations of these measurements, the depth of water is commonly taken as a function of the
opportunity time for water to infiltrate. For the latter assumption to give reliable results, the soil within
a field must be reasonable homogeneous.

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
16
Till and Bos (1985) assumed a normal distribution of applied water depths to parts of an irrigated
field. If the volume of irrigation water equals the volume of water needed for evapotranspiration (field
application efficiency is equal to 100%), 50% of the area is under-irrigated and the other 50% is over-
irrigated. And about 84% of the area receives less water than the mean plus the standard deviation. If
the uniformity of application is not changed and the volume supplied to the field is increased, the
efficiency will decrease, but the under-irrigated area of the field will also decrease. The assumption of
a normally distributed soil moisture over the field implies that, to give 75% of the area an average
supply, the supply should be raised to mean plus 0.67 times the standard deviation. By increasing the
volume of water delivered to the field, the field application efficiency decreases, but an increasing part
of the field receives more than the mean water supply.


Efficiency and uniformity at tertiary unit level

Within a tertiary unit, water is delivered to fields. The deliveries to these fields can be regarded as
observations in a population to which the same statistics can be applied as for fields, provided there
are enough data.


Efficiency and uniformity at scheme level

Within a scheme, the uniformity of flow division over the main and secondary channels and tertiary
units follows from measured flows. Efficiencies and uniformities can be calculated from the same
data.

Wolters et al. (1987) reported on the uniformity of the division of flow over the irrigated lands of the
Fayoum Governorate, in Egypt, with gross irrigated area of 151, 865 ha. The division of flow was
measured from the main intake down to where the area is subdivided into five parts which are served
by a secondary channel (4,155 ha). For each sub-area, the uniformity was expressed by the ratio of
actually delivered flow over intended flow. The intended flow for the Fayoum is based on proportional
flow division over the irrigable area, for which the system was designed. The results were that:
a) Areas with a disproportional supply are easily spotted;
b) The operation needs improvement throughout the system;
c) The most important improvement in the investigated secondary channel would be to
provide it with an adequate quantity of irrigation water.

Making any other improvements in the area would be of little use as long as it remains under-
irrigated.

The timeliness specifications of water deliveries might be usefull in schemes where the objective,
the design, and the operation rules are very strict as to the timing of the water delivery when water
delivered before or after pre-defined period of time is considered as lost.


INCREASES IN THE EFFICIENCY OF IRRIGATION WATER USE

When an increase in the efficiency of irrigation water use is being considered, the following
questions arise:
Is an increased efficiency needed?
Is an increased efficiency possible?
It is realistic to expect an increased efficiency from the proposed measures?


The need for increases in the efficiency of irrigation water use

The need for increases in the efficiency of water use depends on the balance between the
following positive and negative effects of increased efficiencies.

The positive effects of increased efficiency of irrigation water use are:
a larger area can be irrigated with the same volume of water;
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
17
the competition between water users can be reduced;
the effect of a water shortage will be less severe;
water can be kept in storage for the current (or another) season;
groundwater levels will be lower, which can lead to lower investments costs
for the control of waterlogging and salinity;
there will be less flooding;
better use will be made of fertilizers and pesticides and there will be less
contamination of groundwater and less leaching of minerals;
health hazards can be reduced;
energy can be saved;
there will be fewer irrecoverable losses;
instream flows, after withdrawals, will be larger, thereby benefitting aquatic
life, recreation, and water quality.

The negative effects of increased efficiency of irrigation water use are:
soil salinity can increase because of reduced leaching;
wetlands and other wildlife habitats may cease to exist;
groundwater levels will fall and aquifers will receive less recharge;
water retention in upstream river basin areas will be reduced;
there will be a need for more accurate operation and monitoring;
and a need for a more expensive infrastructure.

When considering measures that could lead to an increase in irrigation efficiencies, the possible
effects of the proposed measures on the factors in this list have to be investigated.

Many factors in the list of positive and negative effects of increased irrigation efficiencies have a
relationship with the water quality of the components of the water balance, which means that water
quality investigations should complement efficiency investigations.


The possibilities of increasing the efficiency of irrigation water use

The possibilities of increasing efficiencies are influenced by what is technically possible and,
moreover, by the general rule of the economic feasibility of improving the system. The benefits from
improvements should outweigh the costs.

Investigating the possibility of increasing the efficiency of irrigation water use implies establishing
the components of the water balance of the system. This will often reveal whether, how, and where
the efficiency can be increased. If such a comparison reveals that improvements are possible, targets
values for one or more efficiencies have to be set and the measures needed to reach these target
values have to be taken. Finally, the results of the actions have to be assessed.


Effect of proposed measures to increase irrigation efficiencies

Several issues were recommended as proper tools for the improvement of irrigation efficiencies.
Among those issues water charges, lining canals, improvement of irrigation structures and
modernization of irrigation systems, etc.

The analysis of the literature and comparing the results obtained on the irrigation efficiencies by
the implementation of the previous mentioned measures, the contraries in the data found by the
researchers demonstrates that it is not always realistic to expert increased efficiencies from measures
aimed at that goal.

This could be explained on the ground that irrigation efficiencies are basically ratios of volumes in
the water balance of an irrigation scheme. The studies carried concentrated mainly on the
relationships between the components of the waterbalance of irrigation systems and the factors that
might influence these relationships. However, there are other factors that are related to irrigation
efficiencies such as the management of crops, the socio-economic and legal environments of
irrigation systems, the capacity building in the irrigation sector and the quality of water.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
18

Those factors, together with those related to the components of the water balance of irrigation
systems should be fully considered when final decisions are being made on water use in the
agricultural sector.



CONCLUSIONS AND RECOMMENDATIONS

Irrigated agriculture is by far the greater user of water. The limits to the availability of water
and land for irrigated agriculture necessitate the careful use of these resources. Nevertheless,
an increased water can have negative as well as positive effects.
Low irrigation efficiencies are mainly attributed to the following common problems:
1) Lack of measurement devices
2) Lack of data on water flows and cropping patterns, and
3) Low irrigation efficiencies because of seepage or excessive water applications

Evaluation of development irrigation projects identify additional problems (e.g. inadequate project
formulation and discontinuity after donor involvement).
Lack of data is a serious constraint to improving irrigation. There is a continuous need for
data on water flows, crops, state of system repair, and agricultural practices. Monitoring
systems should be an integrated part of irrigation system management. Many irrigation
schemes, especially those in the formal modern sector, do not live up to expectations
because of over-optimistic assumptions on:
1) the time needed for the design, construction, and realization of benefits, leading to higher
than expected project costs, and benefits being realized later in time;
2) irrigation efficiencies, leading to, for instance, a conveyance system that delivers less
water than planned, and a lower than expected performance in economic terms.
There are considerable seasonal variations in efficiencies whereas, usually, the values are
only high, and only to be high, for a short period of one or two months in the season. That
period is the critical part of the season for design and operation. Further research is needed
into the implications of this phenomenon for the design of irrigation systems.
When considering an increased irrigation efficiency in a certain system, it is more useful to
regard that system as unique, and to use the list of positive and negative effects of increased
irrigation efficiencies on which to base a decision, rather than to rely on general relationships
between characteristics and efficiencies.
Many factors in the list of positive and negative effects of increased irrigation efficiencies bear
a relationship to the water quality of the components of the water balance, which means that
water quality investigations should complement efficiency investigations.


REFERENCES

American Society of Civil Engineers/ASCE. 1973. Consumptive Use of water and Irrigation Water
Requirements. The Technical Committee on Irrigation Water Requirements of the Irrigation and
Drainage Division of the ASCE.
American Society of Civil Engineers/ASCE. 1978. Describing Irrigation Efficiency and Uniformity. The
On-farm Irrigation Committee of the Irrigation and Drainage Division of the ASCE. Proceedings of
the ASCE 104, IR 1:35-41.
Bos, M.G. 1980. Irrigation Efficiencies at Crop Production Level. ICID Bulletin 29.2: 18-26. New Delhi.
Bos, M.G. and W. Wolters. 1990. Water Charges and Irrigation Efficiencies. Irrigation and Drainage
Systems 4: 267-278. Kluwer Academic Publishers. Dordrecht. The Netherlands.
Erie, L.J. 1968. Management : A Key to Irrigation Efficiency. Proceedings of the ASCE 94, IR3:285-
293.
Greenland, D.J.and S.I. Bhuiyan. 1980. Rice Research Strategies in Selected Areas: Environment
Management and Utilization. Special Internationl Symposium on Rice Research Strategies for the
Future, 21-25 April 1980. IRRI, Manila, Philippines. Cited in: Water Management Study at Kaudalla
Irrigation Scheme.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
19
Hall, W.A. 1960. Performance Parameters of Irrigation Systems. Transactions of the American
Society of Agricultural Engineers/ASAE 3(1): 75-76, 81.
Hamdy, A. and C. Lacirignola. 1999. Mediterranean Water Resources: Major Challenges towards
21st Century. Mediterranean Agronomic Institute of Bari, Italy.
Hansen, V.E. 1960. New Concepts in Irrigation Efficiency. Transactions of the American Society of
Agricultural Engineers/ASAE 1960: 55-64.
Hart, W.E., Peri, G. and Skogerboe. 1979. Irrigation Performance: An Evaluation. Journal of the
Irrigation and Drainage Division of the ASCE. 105, IR3: 275-288.
International Commission on Irrigation and Drainage/ICID. 1978. Standards for the Calculation of
Irrigation Efficiencie. ICID Bulletin 27. 1:91-101. New Delhi.
Israelsen, O.W. 1932. (1st Edition). Irrigation Principles and Practices. John Wiley, New York.
Jensen, M.E., Swarner, L.R. and J.T. Phelan. 1967. Improving Irrigation Efficiencies. In: Irrigation of
Agricultural Lands. Agronomy Series: 11, American Society of Agronomy, Wisconsin, USA.
Keller, J. 1986. Irrigation System Management. In: Irrigation Management in Developing Countries.
K.C. Nobe and Sampath, R.K. (Editors). Studies in Water Policy and Management, 8: 329-352.
Westview press.
Levine, G. 1982. Relative Water Supply: An Explanatory Variable. Technical Note 1. The
Determinants of Developing Country Irrigation Problems Project. Cornell and Rutgers University,
Ithaca, N.Y.
Schuurmans, W. 1989. Impact of Unsteady Flow on Irrigation Water Distribution. In, Irrigation: Theory
and Practice. J.R. Rydzewski and C.F. Ward (Editors), Pentech Press, London.
Weller, J.A.; Payawal, E.B. and Salandanan, S.. 1988. Performance Assessment of the Porac River
Irrigation System. Asian Regional Symposium on the Modernisation and Rehabilitation of Irrigation
and Drainage Schemes (held at the Development Academy of The Philippines, 13-15 Feb. 1989).
ODU/Hydraulics Research Ltd. Wallingford, UK.
Wolters, W. , Nadi Selim Ghobrial and Bos, M.G.. 1987a. Division of Irrigation Water in The Fayoum.
Egypt. Irrigation and Drainage Systems. 1:159-172. Kluwer Academic Publishers. Dordrecht. The
Netherlands.























Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
20








































Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
21

FUTURE OPTIONS AND RESEARCH NEEDS
OF WATER USES FOR SUSTAINABLE AGRICULTURE



Muhammad R. Shatanawi
Professor of Water Resources and Irrigation, University of Jordan,
Amman, Jordan, Email: shatanaw@ju.edu.jo


SUMMARY - The increased demand for other uses coupled with recurrent drought and climatic
changes in countries of limited water resources, is producing unprecedented pressure for reducing
the share of fresh water used in irrigation. Many countries in the Mediterranean region, give priority of
water allocation to the domestic sector followed by tourism and industry, and what is left is allocated
to agriculture. At the same time that agriculture is asked to give water to other uses, the increasing
population demand requires increase in food production. This creates a conflict that should be
resolved and should be alleviated by examining different options of water uses for sustainable
agriculture. Increasing the efficient use of water is a key non-structural approach to water resources
management. The agriculture water use efficiency and water productivity is very important as they are
largely inefficient in so many countries due to poor distribution systems and excess irrigation. The
over all global average agricultural water use efficiency in the region is in the order 40%. This paper
analyzes the water situation in the region and shows how agriculture will be affected by water
shortages and giving priority of water uses to other sectors. Water planners and decision makers as
well as researchers are faced with different challenges including; resources, economical,
environmental and institutional. There are some options that can be used to sustain agriculture in the
region considering the above constraints and challenges. They include: improving water use
efficiency, reducing crop consumption of water, irrigation with reclaimed water, practicing deficit
irrigation and irrigation with desalinated water. Lessons learned from Jordan and other countries will
be illustrated in this presentation. The research vision for the next 25 years could include the following
set of actions: have more efficient use and allocation for water use in irrigation; improved water
productivity by introduction of new management measures such as deficit irrigation; and the
introduction of high yielding low water demanding varieties. Desalination of brackish and sea water
can offer limitless fresh water that can be used for agriculture. Therefore, there is a need to find
cheap methods of desalination such as the use of solar and renewable energy. As reclaimed water of
urban wastewater is becoming a new source, it is necessary to maintain efficient and sustainable
agriculture production while using them. Therefore, researchers should consider these needs in
setting up their research priorities.

Key words: irrigation, water saving, research, education, sustainability.


INTRODUCTION

Many countries in the Mediterranean region are located in the arid to semi arid regions of the
world that are classified with limited water resources and increasing water scarcity. Managing these
resources for sustainability will become increasing complex and difficult in the future as climate
changes increases the frequency and intensity of drought and water shortages. The decline in the
available water supplies as a result of the above biophysical factors as well as other geopolitical
factors, rapid expansion in population, urbanization and economic development will result in depletion
of the exploitable water resources. These conditions will produce an unprecedented pressure on the
share of fresh water for irrigating the agricultural sector and will force many countries to reform their
water allocation policy by giving priorities to the domestic and industrial water demand. As such,
irrigated agriculture will be the most effect sector claiming what is left and using the reclaimed
wastewater. In many cases such as Jordan, agriculture is asked to give its share to other uses in
spite of increasing demand for food production. This situation creates a conflict that should be
resolved and could be alleviated by adopting different options of water uses for sustainable
agriculture. Increasing the efficient use of water in agriculture would be a prime option as it is
considered a key non-structural approach to water resources management.

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
22
The issues of agricultural water use efficiency and water productivity are very important as they
are largely inefficient in many countries due to many factors such as poor distribution system, excess
irrigation and lack of proper irrigation water management. Therefore, improving water uses efficiency
and productivity on sustainable basis is an enormous challenge for the water stress and scare
countries of the Mediterranean. They should set forth strategies and plan to design and carry out
programs that increase water use efficiency by farmers while increasing farm incomes. This will
require a combination of technology, education and extension services coupled with research
programs and effective policy framework in order to reflect the real opportunity cost of water. The
focus should be concentrated on crop selection for better water use efficiency and productivity with
improvement of marketing performance to add value and increase the return on agricultural
investment in the irrigated sector. The problem is not limited to countries of the south Mediterranean
but it has reached other humid high rainfall countries like France, Spain, Italy and Greece obliging
them to impose temporary restrictions. The agricultural productivity and competitiveness in the whole
region is adversely affected by water scarcity and inefficient use of water. The average water use
efficiency for the sector in general ranges from 40-45% (Hamdy 2005 and Osman 2006).

Efficiencies in the use of water for irrigation consists of various elements taking into account
losses during storage, conveyance and application to irrigated area while optimal water productivity
can take into consideration, proper water allocation and scheduling, selection of high value crop,
timing of irrigation and other field practices. These practices and elements have been discussed in
many articles but this paper concentrate on the options of research, education and information to
improve irrigation water efficiency and its productivity.


CHALLENGING FACING RESEARCH

Water scarcity is the single most important resource management challenge in the region. Inspite
of that, most countries do not treat their water as a scarce resource. Before exploring different options
in research, education and information, it is necessary to discuss the issues and challenges facing
them. Researchers and decision makers should be aware of the resources and management
challenges as well as socio-economic and environmental issues and institutional setup.


Resources Challenges

The water shortage of the region has been traditionally addressed by increasing supply of water.
The most common approach was to extend exploration and make massive investment in water
resources development. Over the years, most of water resources have been almost developed, so
the rate of investments is currently shrinking. Expanding the supply is unlikely to make dramatic
changes in the future because their development will technically be unfeasible and economically
expensive. Therefore, an essential part of any resources program must focus on managing water
demand. Most of the water saving will come from agriculture by improving water use efficiency.


Economic and Social Challenges

The impact of reduced share of water to the agricultural sector will directly affect the issues of food
self reliance. This may cause much economical and social reform that can create additional
challenges, thus encourage decision makers, researchers and farmers for the efficient use of water
and improved productivity.


Environmental Challenges

With improved water use efficiency, the extra amount of water applied for leaching can be reduced
thus increasing soil salinity. Also, the agricultural sector will relay on a great percentage of its water
supply on reclaimed wastewater and low quality water. These conditions will create environmental
problems that should be addressed in research and management.

Management Challenges
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
23

Supply and demand management, through the more economically efficient supply and use of
water and through changes in production practices as well as reduction of loses, is a vital issue in
resource management strategies and planning. Legislation an economic approach to water demand
management is properly the most important instrument that water resources managers has to
develop and use. They should work together with researcher in order to come up with an optimum
approach to water saving.


Institutional Challenges

For any water management program to be successful and effective, there should be a close link
between researchers and decision makers, between extension workers and researchers and between
farmers or farmers associations and extension workers. For these reasons, an institutional set up
must be created where the four groups are integrated; researchers focus on on-farm research and
extension agents focuses on adoption process. In developing countries, irrigation extension services
dose not exist thus establishing extension agencies for irrigation and water management imposes a
great challenge.


RESEARCH OPTIONS

Universities and research institution are asked to participate in the improved management of
irrigation water and they should demonstrate that they are capable of developing sustainable and
integrated research programs. On the other hand, public sectors (represented by ministries of
agriculture and water and irrigation) and the private sector should give the universities the leading role
in research to gain experience in improving agricultural water use efficiency and productivity.
Research results in many developing countries demonstrates that agricultural demand for water can
be reduced without decreasing the total irrigated areas or the value and the quality of the agriculture
production. Proper water usages along with crop selection might actually increase agricultures
contribution to the economy and at the same time decrease its water usage. However all research
efforts and results are considered worthless unless they can reach and be adopted by the end users
and farmers. Therefore, universities and research institutions should have extension units linked to
the public sector extension services that are capable of disseminating the results to the farmers. The
later are in need of information which are site specific and time dependant because many of them
have realized that optimal agricultural production requires good water management.

The research on supply management are limited to improving the quantity of supply by using
various management tools through modeling and enhancing water supply such as water harvesting
and recharge of groundwater. However, other minor options such as cloud seeding, fog harvesting
and water desalination should not be ignored. On the demand side, the research challenges and
options are attractive because water saving and increasing the value per cubic meter of water are
highly achievable. Any improvement in the irrigation efficiency means expanding irrigated area by the
same percentage. Below are some options that might be adopted by researchers in order to orient
their research program in water demand management aiming at increasing water use efficiency and
improving its productivity.


Crop Water Requirements

Previous calculations of crop evapotranspiration using different formulae and procedures have
always overestimated the actual consumptive use of different crops. Due to lack of data and research
results, these methods were applied to be the basis for the design of many existing projects. In these
projects, the system capacity and the delivery schedule allow for over irrigation most of the time. The
FAO paper 56 has suggested a new procedure in calculating ET based on the revision of previous
methods. Although, FAO procedure is general but the results of field trials have shown that about
20% of ET have been overestimated by previous methods. For example, the average peak ET for
citrus in the northern part of the Jordan Valley was calculated as 5.4 mm/day using FAO methods
compared to previous calculation of 6.5 mm/day according to Blaney-Criddle method. The results of
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
24
recent research (Shatanawi et al 2006) and the farmers practices in the same area have shown that
the average ET has drop to about 4.8 mm/day without any impact on the produce quality and quantity.

On the other hand, crop requirement for near maximum yield is determined by plant physiology,
there are associated management factors manipulating the microclimate environment that can
provide some advantages. For example, the irrigation requirement for open field vegetables may be
twice or triple that for crops grown in plastic houses. A complementary approach is to select planting
times and growing seasons that minimize the atmospheric demand for water consumption. Another
possible action aimed at reducing ET is to change cropping patterns in favor of high value crops
indented to the export that have relatively smaller water requirement.

With the availability of new technologies like real time automatic weather stations and modern
devices like the Eddy correlation, it would be possible to determine the exact amount of daily ET. The
use of remote sensing data using satellite images coupled with ground truthing, ET and crop
coefficient can be determined at district and regional level.

The EU has supported a research project (STRP) entitled Improved Management Tools for
Water-Limited Irrigation: Combining ground and satellite information through models, (IRRIMED) with
the participation of 6 Mediterranean countries under FP6. The aim of this project is the establishment
of tools to support efficient management for water used for irrigation as well as to test scenarios for
long term sustainable policies. Accurate knowledge of water demand and use by irrigated agriculture
is the key to an effective water management strategy. The general scientific objective is the
assessment of temporal and spatial variability of water consumption of irrigated agriculture under
limited water resources condition. Intensive measurement campaigns with eddy correlation equipment
will allow combining ground and satellite measurements into models, to ultimately produce simple
methods to assess evapotranspiration (ET) over large areas.

The accurate assessment of actual ET over selected crop during the growing season, will allow
validating models and to update the crop calendar and crop water requirements. Also, remote sensing
of crop extension and evolution during the growing season will help to measure the actual acreages
of the different crops. Refining existing methods for simple ET estimation will be used to deriving ET
maps from satellite data. This line of research will continuously update information that can be revised
annually based on agro-climatic conditions.


Precision Irrigation

The issue of irrigation scheduling (in when to irrigate and how much water to apply) is a matter of
delivery schedule and farmers decision. With the availability of soil moisture sensors and stem water
potential devices, it is possible to irrigate at the exact time when water is needed by the plant. These
devices can be installed in the soil at two depths or can measure the tension in the leaves or fresh
stems may be connected to electronic control panel that can tell the farmers the need to irrigate.
Research on who to integrate these modern sensors to the irrigation systems is an option that should
be exploited in the future.

Precision irrigation is not limited to irrigation scheduling but can be extended to incorporating them
into the design of various irrigation systems. In surface irrigation, laser land leveling can insure good
uniformity distribution and improved irrigation efficiency. In pressurized irrigation system, the systems
can be operated through automatic control panel. Also, leaks and uneven distribution of irrigation
water along the lateral and subunits can be detected easily. The introduction of such technology will
certainly improve irrigation efficiency and water productivity as well as reducing water losses.


Use of Reclaimed Water

Reclaimed wastewater has become a significant source of the water resources in many countries
of the Mediterranean like Jordan were it contribution to the irrigation sector has reached about 15% in
2005 and will reach 40% by the year 2020. Research in this area is scattered and is limited to treating
this source as low quality water. Research in this regards should be extended to include long-term
impact of using the reclaimed water on soil and the environment, changes in on-farm practices
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
25
especially those related water use efficiency, adopting more high value crops and social and
economical impact of the reuse of treated effluent. The research should focus on finding appropriate
tools to help farmers to overcome problems that will face them and the sector and to develop new
attitude and behavioral patterns.


Desalination of Brackish Water

The area of the Mediterranean, especially the south has a significant reserve of saline water that is
considered a potential resource in the future. It is possible to irrigate certain crops with this kind of
water provided that the soil exhibits good drainage conditions while applying extra water for leaching
purposes. Research results have shown that there are few success cases where the production is
economically feasible. The reduction in yield of up to 50% may not justify the investment provided for
the irrigation and drainage systems as well cost of pumping and delivery. There are few cases where
it is possible to irrigate fodder crops in sandy soils with good natural drainage system.

An alternative to that would be to desalinize this water in which the cost is justified. Experience
from the Jordan Valley has shown that the cost of desalination of saline water (2000 to 5000 ppm)
can reach as low as 0.2 $/m3 using medium size reverse osmosis plants with a capacity of 40 to 50
m3/hr. Irrigation with the blended water of 500 ppm has increased the yield of high consumptive crops
more than twice. Banana yield has increased from 20 ton/ha to 40 ton/ha with good quality produce
while the irrigation water requirements have been reduced from 2500 mm to 1800 mm.

The investment can be farther justified if this water is used to irrigate seedling nurseries and cash
crops like strawberry. Therefore, new research ideas should be explored on conducting comparative
studies, reducing the cost of desalination, and evaluating the environmental and economical impact.


Deficit Irrigation

Deficit irrigation means applying less water than cumulative ET, thereby allowing roots to utilize
stored soil water in the winter or pre-season irrigation. Therefore, the irrigation water requirements in
early irrigation in the spring season can be less than that indicated by ET calculation. Also, deficit
irrigation may be regulated for the rest of the season avoiding critical periods. Such management
practices results in water saving in irrigation without affecting or reducing yield. There are two types of
deficit irrigation; sustained and regulated. In sustained deficit irrigation, the irrigation is reduced during
the whole season while regulated deficit irrigation starts with normal irrigation and then gradually
irrigation is reduced. Regulated deficit irrigation is an irrigation strategy based on limiting non
beneficial water losses by reducing the amount of water for crop during non-critical phonological
stages. The deficit irrigation is controlled during times when the adverse effects on productivity are
minimized. There are a lot of research activities on DI that are going on field crops and vegetables.
Field demonstration conducted by Shatanawi and the French agriculture Mission in Jordan (1996)
showed that 40% reduction in water consumptive from the farmer's practices did not affect the yield.
Observation and communication with some farmers concluded that reducing water application by 30-
40% during drought years did not reduce yield economically. However, research on fruit trees is
limited and should be evaluated in estimating the actual ET under deficit irrigation in order to
maximize the water unit productivity. Such research should include applying different level of irrigation
while measuring the soil moisture content and leave water potential.

It is worth mentioning at this point that EU has supported a research project on deficit irrigation
entitled Deficit Irrigation for Mediterranean Agricultural Systems (DIMAS). The objective of this
project is to evaluate the concept of deficit irrigation (DI) as a means of reducing irrigation water use
while maintaining or increasing farmers profits. The DI concept will be the subject of multidisciplinary
research at different scales, geographic locations, and with different perennial and annual crops. The
objective will be to develop a workable, comprehensive set of irrigation (DI) strategies that can be
disseminated quickly among the various agricultural systems of the Mediterranean Region. The
project addresses directly the first topic of the FP6- INCO-2002-B1.2 specific measure, research on
sustainable irrigation, including deficit irrigation. Eleven partners from seven different countries
(Greece, Italy, Jordan, Morocco, Spain, Tunisia and Turkey), including research and water
association institutions will work for three years on the project. Their main activities will be: a) the
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
26
development of a general summary model of crop yield as a function of water supply, b) the validation
of the model for the main irrigated annual (wheat, sunflower, cotton,) and perennial crops (olive,
pistachio, citrus), using common research protocols, c) a survey on physical, socio-economic and
cultural conditions for each crop and irrigated area, and d) scaling up by combining the yield model
with economic optimization modules that will generate optimum DI strategies compatible with the
specific socio-economic characteristics of each area under study.

The results of the project will provide recommendations for reducing irrigation water use while
ensuring the sustainability of irrigated agricultural systems in the Mediterranean basin. Feedback with
project end-users will take place via participation of farmers associations and irrigation water
agencies who will contribute their expertise in managing water scarcity, thus ensuring that all relevant
issues are addressed.


Irrigation Techniques

The efficiency of the on-farm water use and the water productivity can be increased with improved
irrigation techniques. Innovation in this area should be pursued between researchers and the
irrigation industry. Although micro irrigation is known to be high efficient irrigation system but
experience shows that if the system is not well design and not operated probably, the efficiency can
be as low as 50%. In addition to that, research on irrigation accessories such as filters, pressure
regulators should be incorporated into the system design and management. In this area, the use of
sand filter with proper sand gradation automatic filter systems, emitters, acid and chlorine injection
should be tested and experimented on crops highly sensitive to water stress. For other irrigation
systems like surface irrigation and sprinkler irrigation, there is a high potential for innovation such as
molding of surface irrigation, irrigation cut back and surge flow irrigation. The design of all irrigation
system should provide flexibility and simplicity required for successful operation under different soil
variables and topographic variation. Research should be oriented toward proper and careful selection
of pumps, pipes and on-farm sprinkler equipment in order to sustain high uniformity at a specified
application rate. The research in irrigation system should also concentrate on the energy aspect by
introducing and testing low pressure micro and sprinkler irrigation in order to reduce the cost of
operation and maintenance. Technology so far, has produced sprinkler system of low energy
precision application (LEPA) and low pressure compensating emitters that can give high uniform
application rate and efficient irrigation. A probably designed and managed system incorporating all of
the above technology can have efficiency as high as 98%. Research should be further pursued to
explore along with the industry new technology that can save in water and produce uniform irrigation.


EDUCATION AND TRAINING OPTIONS

The overarching goal in promoting the efficient and the effective management of water is the
investment in human resources development. The venue in this regards has many options ranging
from University diploma and higher graduate research to in-job training and tailor training programs.
Most universities in the region offer B.Sc degrees and M.Sc Degrees in water and irrigation and few
Ph.D programs. However, capacity building is not only limited to new graduate and extension agents
but should be extended to the decision makers, legislators and stakeholders.

Training on the state of art on water management can help in establishing water resources
management agencies and creation of irrigation advisory units. Experience from Jordan has shown
that farmers receiving training from the University of Jordan in irrigation scheduling and management
have reported improvement in irrigation efficiency by 30% (Shatanawi, 2004). The link between
farmers and educators as well as researcher should be the responsibility of the extension services
who will convey the result of the research to the farmers. Therefore training of extension personal will
facilitate the flow information to the stakeholder groups such as water user associations. They should
be also trained to carry on activities such as managing demonstration sites, organizing working
session with farmers groups, hosting educational opportunities for professional, disseminating
information and encouraging individuals to participate in improving water use efficiency and
productivity. The results of the demonstration programs that have been undertaken on pilot basis,
aiming at improving on-farm water use efficiency can be disseminated by extension services.

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
27
The irrigation sector in many countries of the Mediterranean needs high quality extension services
with innovative ideas to translate policies into action plans. The extension services should not be
limited to the public sector but the various irrigation companies and private agricultural enterprises
who are using the state of art technology must all of them take action in training extension services as
well as farmers.


INFORMATION AND TECHNOLOGY TRANSFER

Information is considered a key element in the day to day management, decision making and
undertaken current and future water balance and flow. The flow and availability of information to
researchers, operators and farmers is also of prime important to determine irrigation scheduling by
farmers and water delivery schedule by farmers. There are different levels of information required. For
operators of the irrigation system at the project, it is necessary that information regarding flows of
surface water and reclaimed wastewater, and stocks of groundwater be available at the present and
future. This kind of information coupled the projected cropping pattern and agro-climatic data can help
in better water allocation and optimization, thus improving water use efficiency and productivity. The
role of the irrigation authority should not be limited to operation at the sub-unit level, which is
responsibility of the farmers group, but they should work at higher level of planning and management
at the project level. The Water Management and Information System (example from the Jordan
Valley) can provide such a model for irrigation districts.

Research and technology transfer institution can play an important role in utilizing the field and
climatic data, information of crop pattern and meteorological data into daily out put that can be used
by farmers. An example on that is the Irrigation Information System established by the National
Center for Research and Technology Transfer (NCARTT) where the whole country is covered by a
network of automatic weather stations. Daily ET is calculated for different location from the data
transmitted from these stations using the FAO method of paper 56.

On the other hand, expansion of pilot projects and demonstration farms will help in obtaining
accurate information on the impact of improved irrigation practices on water use and economic return
of unit volume of water. They will serve as a venue for comprehensive extension and technology
transfer programs. The demonstration sites of concentrated water practices could be divided to cover
the different climatic zone of any region. Demonstration activities can include the following:
1. Demonstration on micro irrigation operation and maintenance.
2. Demonstration of deficit irrigation strategies.
3. Demonstration of winter soil moisture conservation technique.
4. Demonstration on optimum irrigation practices.

Each country should have an entity responsible to coordinate national agriculture research on
water management and technology transfer activities. The entity can act as a focal point for
capitalizing and facilitating information and technology transfer including social and economic
marking. The focal points should work among a network of partners such as universities, irrigation
authorities, extension services and other environmental societies.

The experience of the Scientific Irrigation Scheduling Services (SISS) from California State and
Washington State can be technically adopted by some countries. SISS are enterprises that market full
service or self service SIS products such as a low cost tensiometers, water marks and portable soil
water sensors (Neutron Probe). Also, the technology of Aqua-Card have been widely used in Italy
could be adopted by other Mediterranean countries.


CONCLUSIONS

Many countries in the region are characterized by limited water supplies while providing additional
resources will be an expensive option. Therefore, efforts in the optimal management of water
resources should concentrate on the demand side management. As the agricultural sector is
consuming the bulk of water supply, good management of irrigation water can be translated into
significant amount of saving in the water resources. In addition, the agriculture sector will be most
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
28
affected by water shortage and would be asked to give water to other uses such as the domestic and
the industrial sector. Therefore, increasing the efficient use of water in the agricultural sector would be
an overarching goal in changing certain policies or adopting new ones with the objectives of
improving on-farm water management and maximize agricultural return per unit of water. Optimization
of water use at the farm level involves getting the maximum value output for minimum amount of
water. There are several activities that can be carried out to achieve this goal ranging from field
management practices by the farmers to water management approach by decision makers. However,
these activities can attain high degree of success if it is supported by good research programs and
extension services where both need the availability and flow of information.

Research program and activities should be oriented towards the adoption of improved
technologies, management practices and policies which contribute to the national development as
well as addressing the problems and needs of the irrigated agriculture. Research projects must be
planned and formulated in such a way that these activities are directed towards effectively conserving
and managing the national resources utilized in agricultural development so it would not contribute to
the loss of and/or deterioration of these resources.

In order to cope with future challenges facing irrigated agriculture, research and technology
transfer, system must be capable of responding to the needs of an increasingly more sophisticated
and more competitive agricultural sector. Research projects will be demand driven and should be
applied and/or adaptive in nature. In order to ensure the most efficient use of the limited resources
available for research and technology transfer activities, project should be multidisciplinary in their
approach.

The role of institutions and agencies engaged in agricultural research must be clearly defined,
along with the areas of research and technology transfer they are capable of implementing and
carrying out. Agricultural research should take into consideration population increase, changing
consumer habits, limited water resources and the need to increase water use efficiency and improve
production practices to reduce production costs and increase the competitiveness of the production.

The goal of research and technology transfer in the area of irrigated agricultural is to achieve an
efficient and economic production, diversification and marketing crops for domestic and export
market. In order to achieve the above goals, research strategy and technology transfer activities
should be directed towards the following objectives:
2. Improve water use efficiency
3. Utilize non-traditional water resources
4. Intensify cropping systems
5. Improve and test new management practices


REFERENCES

Al-Jayyousi, O. and M. Shatanawi. (1995).Analysis of Future Water Policies in Jordan. Water
Resources Development, Vol. 11, No. 3, pp. 315-330.
Allen, R., A. Clemmens, C. Burt, K. Solomon, and T. O'Halloran. (2005). Prediction Accuracy for
Project-Wide Evapotranspiration Using Crop Coefficients and Reference ET. J. Irrig. Drain. Eng.
1(24), 24- 36.
ESCWA, (2004). The Optimization of Water Resources Management in the West Asia Countries,
Economic and Social Commission for Western Asia, UN, New York.
Fereres, E., Goldhamer, D. A., and Parsons, L. R. (2003). Irrigation Water Management of
Horticultural Crops. HortScience. 38(5), 1036-1042.
Fereres, E. (2005). Managing Deficit Irrigation: from the Crop to the District. a paper presented at
the 4th WASAMED Workshop on Water Use Efficiency and Productivity, University of Jordan,
Amman, Jordan, 1-4 October, 2005.
Hamdy, A. (2001). Sectorial Water Use Conflict and Water Saving Challenges in the proceeding of
the workshop entitled Water Saving and Increasing Water Productivity; Challenges and Options.
March 10-23, 2001. Amman, Jordan.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
29
Hamdy, A. (2005). Water use Efficiency in Irrigated Agriculture: An Analytical Review a paper
presented at the 4th WASAMED Workshop on Water Use Efficiency and Productivity, University of
Jordan, Amman, Jordan, 1-4 October, 2005.
Osman, M.E. (2005). Reflection on the Status of on-Farm Water Use Efficiency in Selected Countries
of West Asia, a paper presented at the 4th WASAMED Workshop on Water Use Efficiency and
Productivity, University of Jordan, Amman, Jordan, 1-4 October, 2005.
Shatanawi, M. R. (1986). Efficiency of the Jordan Valley Irrigation System. DIRASAT (Agricultural
Sciences). 13 (5), 121-142.
Shatanawi, M. R. et al., (1994). "Irrigation Management and Water Quality in the Central Jordan
Valley", A Baseline Report Prepared for the USAID Mission to Jordan, by the Irrigation Support
Project for Asia and the Near East (ISPAN) and the Water and Environment Research and Study
Center, University of Jordan, Amman, Jordan.
Shatanawi, M. (2002). Policy Analysis of Water, Food Security and Agriculture Policies in Jordan,
Review paper submitted to the World Bank.
Shatanawi, M. (2004). Improved Management Tools in Irrigation for the Jordan Valley, an internal
report submitted to the Deanship of Scientific Research, University of Jordan, Amman, Jordan.
Shatanawi, M., M. Duqqah and S. Naber. (2006). Agriculture and Irrigation Water Policies for Water
Conservation in Jordan. A paper presented in the 5th WASAMED Workshop on Harmonization
and Integration of Water Saving Options: Convention and Promotion of Water Saving Policies and
Guidelines. Malta, 3-7 May, 2006.
World Bank. (2002). Water Sector Review Update, Main Report, The Hashemite Kingdom of Jordan,
World Bank, Washington D.C.



































Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
30
























































Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
31

RELATING WATER PRODUCTIVITY AND CROP EVAPOTRANSPIRATION



Luis S. Pereira*
*Center for Agricultural Engineering Research, Institute of Agronomy,
Technical University of Lisbon, Tapada da Ajuda, 1349-017 Lisbon, Portugal,
Fax: +351 21 362 1575; email: lspereira@isa.utl.pt



SUMMARY - Water productivity became an important issue in improving the performance of
irrigation, including when focusing water saving issues. However, various concepts may be
considered, which requires appropriate definitions and related analysis. Because it represents a ratio
between harvesting yield and water use, a main component of the latter is crop evapotranspiration.
This calls for a discussion relative to its concepts and computation, thus to identify both how crop
evapotranspiration may be managed to improve water productivity and existing gaps in its knowledge.
In addition, a discussion on the economic aspects relative to improved water productivity and saving
is presented, including a simple analysis of economic issues and respective gaps in knowledge.

Keywords: water productivity, land productivity, evapotranspiration, aerodynamic resistance, surface
resistance, crop coefficients, stress coefficients.


WATER PRODUCTIVITY VS. IRRIGATION EFFICIENCY

The term efficiency is commonly applied an irrigation systems or sub-system: water storage,
conveyance, distribution off- and on-farm, and application at the farm. It can be defined by the ratio
between the water depth delivered by the sub-system under consideration and the water depth
supplied to that sub-system, usually expressed as a percentage. Adopting an output/input non-
dimensional ratio, the term efficiency could be applied to evaluate the performance of any irrigation
and non-irrigation water system but the term is almost exclusive of irrigation (Pereira et al., 2002a).
Misleading interpretations are therefore common by water managers, which should be avoided.

For farm irrigation systems, the application efficiency Ea may be defined by the ratio between the
average water depth added to the root zone storage in the quarter of the field receiving less water to
the average water applied. However, this indicator should be used together with others, mainly those
relative to distribution uniformity (Burt et al., 1997; Pereira, 1999; Pereira and Trout, 1999). Moreover,
this indicator Ea should not be used for characterizing seasonal irrigation but only each irrigation
event because soil water availability and water depths applied vary from an irrigation to the next and
they highly influence this performance indicator as analysed by those authors. In addition, weather
conditions, mainly wind speed and temperature in case of sprinkler irrigation, may also vary from an
irrigation event to another and largely influence Ea.

Farmers do not see the improvement of farm application efficiencies as a must. Application
efficiencies become higher when farmers apply water timely and the distribution uniformity is higher.
Improved uniformities decrease differences in amounts of water made available for the crop in the
under-and over-irrigated parts of the field. As discussed by many authors, e.g. Keller and Bliesner
(1990) and Mantovani et al. (1995), this leads to more even crop development and higher yields.
When the farmer adopts an appropriate irrigation scheduling, then yields are positively impacted; in
addition, the application efficiency becomes higher as well as the economic results of irrigation
(Ortega et al., 2005). Thus, improving irrigation efficiency is not a farmers objective but to achieve
higher yields and economic profit.

Improving transport and distribution efficiencies may be an objective of farmers management of
irrigation systems when seepage, leaking or overflow would decrease the availability of water to tail-
end distributor canals and tail-end farmers, or when improvements aim at an easier control of
deliveries to branch canals, distributors and farms. In other words, the interest of farmers is to have
improved service performances. Thus, the former efficiency terms are being replaced by indicators of
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
32
canal and pipe systems that refer to service performance, such as reliability, dependability and equity
(Molden and Gates, 1990; Lamaddalena and Pereira, 1998; Lamaddalena and Sagardoy, 2000;
Pereira et al., 2003a; Bos et al., 2005).

Improving irrigation efficiencies is often said to be an objective associated with water savings.
However this is only true when the farmers and canal managers have the appropriate tools and farm
and off farm irrigation systems are designed and managed in such a way that delivery and irrigation
scheduling can be applied effectively, as discussed by Pereira et al. (2002a, b). Otherwise water
savings may not be achieved in that way but through under-irrigating the crop.

Nowadays, there is a trend to call for increasing water productivity as a must (FAO, 2002; Molden
et al., 2003). The attention formerly paid to irrigation efficiency issues is therefore being transferred to
water productivity. However, this term is used with different meanings (Fig. 1). Water productivity may
be generically defined as the ratio between the actual yield achieved (Ya) and the water use,
expressed in kg/m3, but the denominator may refer to the total water use (TWU), including rainfall,
which is referred herein with the symbol WP:


TWU
Ya
WP =
(1)


WATER DIVERSION
Agriculture
and landscape
Effective rain
Transpiration Soil evapor
Seepage +
runoff
Conveyance +
distribution
Application to
cropped field
Percolation +
runoff
Non-crop ET
YIELD
reuse
Total WP
Irrig WP
WUE
Farm
WP
WATER DIVERSION
Agriculture
and landscape
Effective rain
Transpiration Soil evapor
Seepage +
runoff
Conveyance +
distribution
Application to
cropped field
Percolation +
runoff
Non-crop ET
YIELD
reuse
WATER DIVERSION
Agriculture
and landscape
Effective rain
Transpiration Soil evapor
Seepage +
runoff
Conveyance +
distribution
Application to
cropped field
Percolation +
runoff
Non-crop ET
YIELD
reuse
Transpiration Soil evapor
Seepage +
runoff
Conveyance +
distribution
Application to
cropped field
Percolation +
runoff
Non-crop ET
YIELD
reuse
Total WP Total WP
Irrig WP Irrig WP
WUE WUE
Farm
WP
Farm
WP



Fig. 1. Different definitions of water productivity and water use efficiency.

However, it may refer only to the irrigation water used (IWU) mobilized for at system level (WP
I
),

IWU
Ya
I
WP =
(2)
or to the total water use at farm or field level (TWU
Farm
), thus including rainfall and irrigation (WP
Farm
),

Farm
TWU
Ya
Farm
WP =
(3)
or relate to irrigation water (IWU
Farm
) only, thus (WP
I-Farm
):
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
33


Farm IWU
Ya
Farm I
WP =

(4)

The meaning of these indicators is necessarily different and may cause contradictions when the
wording water productivity is used without specifying which target is being considered.

The term water use efficiency (WUE) is also commonly used in irrigation but often with different
meanings. Some authors refer to it as a synonymous of application efficiency, thus as a non-
dimensional output/input ratio; others adopt it to express the water productivity of the irrigation water,
as a yield to water ratio. To avoid misunderstandings, the term water use efficiency should be limited
to physiological and eco-physiological purposes (Steduto, 1996) or, as some do, may be replaced by
the term transpiration ratio or similar.

The idea that improving water productivity or the water use efficiency leads to water savings is
also not entirely true because it is also required to distinguish between consumptive and non-
consumptive uses. The same amount of grain yield depends not only on the amount of irrigation
water used but also on the amount of rainfall water that the crop could use, which relates to rainfall
distribution during the crop season. Moreover, the pathways to improve yields are often not related
with water management but with agronomic practices and the adaptation of the crop variety to the
cropping environment. However, a crop variety that has a higher WUE than another has the potential
for using less water than the second when achieving the same yield. But this is a characteristic
intrinsic to the crop and is not depending upon irrigation management.


WATER PRODUCTIVITY CONCEPTS

Considering Fig. 1, one may approach the different concepts relative to water productivity and
assume some definitions aimed at irrigation management. Then the following base definition is
adopted:


TWU
Ya
WP = (1a)
where Ya is the actual harvestable yield in kg, and TWU is the total seasonal water use by the crop in
m
3
or, referring to the unit surface, in mm.
Eq. 1a may take a different form


NBWU LF ETa
Ya
WP
+ +
= (5)
where Ya is the actual harvestable yield and the denominator refers to the water use components;
ETa is the actual season evapotranspiration in mm, LF is the water used for leaching in mm and
NBWU is the non-beneficial water use in mm. This concerns the percolation through the bottom of the
root zone, runoff out of the irrigated fields, and losses by evaporation and wind drift, The beneficial
water use (BWU) is then constituted of ETa and LF.

If the seasonal water use is considered through the respective and diverse water sources, then
Eq. 5 is replaced by the following equation:

I SW CR P
Ya
WP
+ + +
= (6)

where P is the season precipitation, CR is the amount of capillary rise, SW is the difference in soil
water content between planting and harvest and I is the seasonal irrigation depth, all expressed in
mm.

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
34
One can observe that maximizing water productivity is to find out its limit when the maximal yield
Ymax is attained, which means that ETa = ETc, where ETc is defined as the ET of a healthy crop,
well managed and not short of water, thus cultivated under pristine conditions (Allen et al., 1998), and
that NBWU is at its minimum value:

( )
( ) NBWU LF ETc
Y
WP
min
max
max
+ +
= (7)

An high WP may also be obtained when a crop is water stressed (up to acceptable limits); then the
yield is reduced as well as the denominator terms in Eq. 5. But such an high productivity is obtained
with ETa < ETc, and with LF below its target value. If this option is non-controlled, and control is
difficult to be achieved including when farm systems have appropriate distribution uniformity, yields
may decreased below an acceptable level and therefore induce appreciable income loss to the
farmer. Observing Fig. 2 it may be seen that if the objective would be to maximize WP and not the
land productivity often farmers would not have advantage in practicing supplemental irrigation. The
figure also shows that the highest WP are obtained in rainfed wheat production. This may be
understood if the irrigation systems they use are poorly performing. However, the figure also show
that yields under rainfed production may often be below the economic viability, which clearly justifies
the farmers option to adopt supplemental irrigation.

y = -0.4278x
2
+ 4.7328x - 0.543
R
2
= 0.7611
0
5
10
15
20
0 2 4 6 8 10
Grain yield (t ha-1)
W
P

(
k
g

h
a
-
1

m
m
-
1
)
Irrigated Rainfed


Fig. 2. Relationship between water and land productivity for durum wheat in northern Syria (source:
Zhang and Oweis, 1999).

It is therefore important to consider the economic issues relative to water productivity since the
objective of a farmer is to achieve high income and profit.

Replacing the numerator of equations above by the monetary value of the achieved yield Ya, the
economic water productivity (EWP) is expressed as /m
3
and defined by:


( )
TWU
Ya Value
EWP = (8)


However, the economics of production is less visible in this form than when both the numerator
and the denominator are expressed in monetary () terms, respectively the yield value and the TWU
cost, thus yielding the following definition:

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
35

( )
( ) TWU Cost
Ya Value
EWP = (9)

Alternatively, this definition may be expressed assuming that all water costs are due to the costs of
irrigation, thus:


( )
( ) I Cost SW CR P
Ya Value
EWP
+ + +
= (10)

This may not be true if water conservation measure are used and therefore there are costs
associated with water harvesting or soil management that create additional mobilization of rainfall,
mainly increasing the infiltrated fraction and/or reducing soil water evaporation losses. Alternatively,
considering Eq. 5, the following definition may be used:

( )
( ) NBWU LF ETa Costs
Ya Value
EWP
+ +
= (11)

Determining the costs associated with the water use components as in Eq. 11 may be difficult but
ideally this equation may support the economic evaluation of measures to control the NBWU.

Maximizing EWP, when all costs not referring to water use are kept constant, means to find the
limit to the ratio between the yield value and the yield costs associated with water use, which
corresponds to maximize the crop revenue in which concerns water use:

( )
( )
( ) Income
Y Costs
Y Value
EWP max
max) (
max
max max = = (12)

This maximal EWP or maximal revenue is often different from the maximal yield and depends
upon the structure of the production costs (Fig. 3).

Fig. 3. Schematic representation comparing how maximizing farm incomes for a commercial and a
family farm lead to different approaches to economic water productivity (costs relative to water
volumes used are not considered for simplification).

For a farm where labour is by workers or using somewhat sophisticated equipment associated
with energy costs, then the irrigation costs grow almost linearly with the amount of irrigation water
use. Contrarily, for a farm using surface irrigation without energy costs associated nor large capital
investment, and where labour is provided by the family, thus is remunerated by the final yield, the
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
36
effective costs relative to irrigation are not depending upon the amount of irrigation water use.
Therefore, the maximum net income for the first may be close to the maximal EWP, while the
maximum income for the family farm are close to the maximal land productivity since land, not water,
is the main limiting factor determining farm income. The related economic impacts are however less
well known, insufficient data are available and tools for the respective analysis are insufficiently
developed (Victoria et al., 2005).

It is therefore important to understand how WP could be improved. Knowing that yields depend
upon the seasonal evapotranspiration, this analysis focuses this component of the water use.


CROP EVAPOTRANSPIRATION AND RESISTANCES TO VAPOUR FLUXES

The Penman-Monteith equation (Monteith, 1965) is generally considered to be able to represent
the evapotranspiration from any vegetated surface (Jensen et al., 1990; Allen et al., 1994, 1998;
Pereira et al., 1999). It can be expressed by the following combination equation:

|
|
.
|

\
|
+ +

+
=
a
s
a
a s
p a n
r
r
1
r
) e (e
c G) (R
ET

(13)

where R
n
is the net radiation, G is the soil heat flux, (e
s
e
a
) represents the vapour pressure deficit of
the air,
a
is the mean air density at constant pressure, c
p
is the specific heat of the air, represents
the slope of the saturation vapour pressure -temperature relationship, is the psychrometric constant,
and r
s
and r
a
are the (bulk) surface and aerodynamic resistances (Fig. 4)

soil
stomatal
a
i
r

f
l
o
w
r
r
s
a
(bulk) surface
resistance
aerodynamic
resistance
reference
level
evaporating
surface
cuticular


Fig. 4. Schematic representation of the resistances to vapour fluxes (Allen et al., 1998)


The Penman-Monteith approach as formulated above includes all parameters that govern energy
exchange and corresponding latent heat flux (evapotranspiration) from uniform vegetation canopies.
Most of the parameters are measured or can be readily calculated from weather data. The equation
can be utilized for the direct calculation of any crop evapotranspiration as the surface and
aerodynamic resistances are crop specific.

Aerodynamic resistance (r
a
) determines the transfer of heat and water vapour from the
evaporating surface into the air above the canopy. It can be expressed as:
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
37


( ) | | ( ) | |
z
m h
a
u k
z d z z d z
r
2
om oH
/ ln / ln
=
(14)

where r
a
is the aerodynamic resistance [s m
-1
], z
m
is height of wind measurements [m], z
h
is the
height of air humidity measurements [m], d is the zero plane displacement height [m], z
om
is the
roughness length governing momentum transfer [m], z
oh
is the roughness length governing transfer of
heat and vapour [m], k is the von Karman's constant, 0.41 [-], and u
z
is the wind speed at height z [m
s
-1
].

As discussed by Alves et al. (1998), the assumption that heat and vapour escape from the canopy
from the level d+z
oH
, as it is implied in Eq. 14, can be questioned. In alternative, r
a
can be calculated
from the top of the canopy to the reference height, using (Perrier, 1975; Stockle and Kjelgaard, 1996):

( ) ( ) | | ( ) | |
z
m c h
a
u k
z d z d h d z
r
2
om
/ ln / ln
= (15)

where h
c
is the crop height [m].

These equations are restricted for neutral stability conditions, i.e., where temperature, atmospheric
pressure, and wind velocity distributions follow nearly adiabatic conditions (no heat exchange). The
application of the r
a
equations for short time periods (hourly or less) may require the inclusion of
corrections for stability. However, when predicting ET
o
in the well-watered reference surface, heat
exchanged is small, and therefore stability correction is normally not required.

For its practical application, the parameters d and z
o
, if not measured, can be estimated from the
crop height h
c
[m] and LAI [-] (e.g. Brutsaert, 1982; Perrier, 1982):
( ) ( )
|
.
|

\
|
= 2 1
2
1 / LAI exp
LAI
h d
c
(16)
( )| | ) / LAI exp( / LAI exp h z
c om
2 1 2 = (17)
Genetic improvements and crop management influence these parameters through acting on hc
[m] and LAI but impacts are relatively small. Aiming at increasing WP (Eqs. 5 and 7) changes should
focus on decreasing ETa and ETc, thus on increasing the aerodynamic resistance, thus decreasing
both d and zom (vd. Eqs. 14 and 15); however the main impact on ra depends on weather conditions
through wind speed. Eqs. 16 and 17 show that d and zom increase for high and fully cover crops and
are smaller for low crop heights and partial cover crops. Thus, plant breeding improvements may
favour higher ra when crops become lower in height and LAI.

Surface resistance is more complex. The bulk surface resistance describes the resistance of
vapour flow through the transpiring crop and evaporating soil surface. Where the vegetation does not
completely cover the soil, the resistance factor should indeed include the effects of the evaporation
from the soil surface. If the crop is not transpiring at a potential rate, the resistance depends also on
the water status of the vegetation.

Plant physiologists consider rs to be a purely physiological parameter that accounts for the
stomatal control of transpiration. Stomata have been carefully studied and the factors that determine
their functioning are well known. Some of them, like radiation, either solar radiation Rn or PAR,
temperature (T) and vapour pressure deficit (VPD) are those that govern the physical process of
evaporation. Others, like soil (or plant) water potential (

) represent the true physiological control by


stomata which takes place mainly in water stress conditions. Other factors, like the age of the leaf, the
previous history of water stress of the plant and the position of the leaf in the plant, are also important
but less quantifiable (Alves and Pereira, 2000).

Scaling resistances from leaf to canopy, which constitutes the "bottom up" approach to rs, is full of
controversy. The standard procedure is to average stomatal resistance r
st
at different levels in the
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
38
canopy, weighted by leaf area index (Monteith, 1973). However, the values of r
s
determined this way
even with measured stomatal resistances seem to give good results only in very rough surfaces, like
forests, and partial cover crops with a dry soil. On complete cover crops, especially when the soil is
wet, average stomatal resistance can greatly depart, being normally lower, from the values of r
s

obtained as a residual term of the Penman-Monteith equation using the "top down" approach (e.g.
Baldocchi et al., 1991; Rochette et al., 1991). The following equation establishes the essential
relations between r
s
and weather variables (Alves and Pereira, 2000):

( )
( ) G R
VPD c
r r
n
p a
a s

+ +
|
|
.
|

\
|

1 1 (18)

where is the slope of the vapour pressure curve (Pa/C), is the psychrometric constant (Pa/C),
a

is the atmospheric density (kg/m
3
), c
p
is the specific heat of moist air (J kg
-1
C
-1
), R
n
-G is the energy
available at the crop surface, and is the Bowen ratio. This discrepancy has been regarded as to
indicate that not all leaves actually contribute to the total evaporation fluxes from the canopy. The
concept of "effective" leaf area was therefore introduced and linked to radiation interception, the
upper, well illuminated leaves being those that most contribute to transpiration.

The surface resistance r
s
is crop specific and relates to the stomatal resistance r
st
and to the
"effective" leaf area; increasing resistance to water stress implies increased stomatal control, thus
higher r
st
and r
s
. plant breeding for increased resistance to water stress may lead to increased r
st
and
r
s
; however, acting on these crop characteristics is difficult and should avoid that increasing r
st
and r
s

would lead to lower photosynthetic efficiency, which would decrease WUE. Referring to Eq. 18, non
considering the role of climate that may be the essential factor determining r
s
, it becomes evident that
the main factor to act on is the Bowen ratio : since it represents the ratio between sensible and latent
heat fluxes, a low indicates high water availability to the crop and an high is representative of
water stress conditions. Therefore, for unchanged climate conditions, high rs values are obtained
when the aerodynamic resistance is high and the crop is water stressed. However, limiting the
availability of water to the crop may lead to lowering the photosynthesis and to reduced yields. As for
r
a
, acting on r
s
is difficult, could have contradictory results in terms of crop yields, and may be less
efficient in increasing WP.


CROP EVAPOTRANSPIRATION AND CROP COEFFICIENTS

Crop evapotranspiration can be derived from meteorological and crop data by means of the
Penman-Monteith equation (Eq. 13). By adjusting the albedo and the aerodynamic and canopy
surface resistances to the growing characteristics of the specific crop, the evapotranspiration rate can
be directly estimated. The albedo and resistances are, however, difficult to estimate accurately as
they may vary continually during the growing season as climatic conditions change, as the crop
develops, and with soil surface wetness and soil water availability. The canopy resistance will further
be influenced by the soil water availability, and it increases strongly if the crop is subjected to water
stress.

As there is still a considerable lack of consolidated information on the aerodynamic and canopy
resistances for the various cropped surfaces, the crop coefficient approach is generally used to
estimate the crop evapotranspiration, ET
c
, which is calculated by multiplying the reference crop
evapotranspiration, ET
o
, by a crop coefficient, K
c
:

o c c
ET K ET = (19)

where ET
c
is the crop evapotranspiration [mm d
-1
], K
c
is the crop coefficient [dimensionless], and ET
o

is reference crop evapotranspiration [mm d
-1
].

Considering the original Penman-Monteith equation (Eq. 13) and the equations of the
aerodynamic and surface resistance described above, the FAO-PM reference ET
o
equation for daily
time-step computations has the following form (Allen et al., 1994):

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
39
) 34 . 0 1 (
) (
273
900
) ( 408 . 0
2
2
u
e e u
T
G R
o
ET
a s n
+ +

+
+
=

(20)
where ET
o
is the reference evapotranspiration [mm day
-1
], R
n
is net radiation at the crop surface
[MJ m
-2
day
-1
], G is soil heat flux density [MJ m
-2
day
-1
], T is air temperature at 2 m height [C], u2 is
wind speed at 2 m height [m s
-1
], e
s
is saturation vapour pressure [kPa], ea is actual vapour pressure
[kPa], e
s
-e
a
is saturation vapour pressure deficit [kPa], is the slope of the vapour pressure curve
[kPa C
-1
], and is the psychrometric constant [kPa C
-1
].

The FAO-PM equation (Eq. 20) determines the evapotranspiration from the hypothetical grass
reference surface and provides a standard to which evapotranspiration in different periods of the year
and in other regions can be compared and to which the evapotranspiration from other crops can be
related.

Most of the effects of the various weather conditions are incorporated into the ET
o
estimate.
Therefore, as ET
o
represents an index of climatic demand, K
c
varies predominately with the specific
crop characteristics and only to a limited extent with climate. This enables the transfer of standard
values for K
c
between locations and between climates. This has been a primary reason for the global
acceptance and usefulness of the crop coefficient approach and the K
c
developed in past studies.

The crop coefficient, K
c
, is basically the ratio of the crop ET
c
to the reference ET
o
, and it represents
an integration of the effects of four primary characteristics that distinguish the crop from reference
grass. These characteristics are:
- Crop height. The crop height influences the aerodynamic resistance term, ra, of the FAO
Penman-Monteith equation and the turbulent transfer of vapour from the crop into the
atmosphere. The ra term appears twice in the full form of the FAO Penman-Monteith
equation.
- Albedo (reflectance) of the crop-soil surface. The albedo is affected by the fraction of ground
covered by vegetation and by the soil surface wetness. The albedo of the crop-soil surface
influences the net radiation of the surface, Rn, which is the primary source of the energy
exchange for the evaporation process.
- Canopy resistance. The resistance of the crop to vapour transfer is affected by leaf area
(number of stomata), leaf age and condition, and the degree of stomatal control. The canopy
resistance influences the surface resistance, rs.
- Evaporation from soil, especially exposed soil.

The soil surface wetness and the fraction of ground covered by vegetation influence the surface
resistance, r
s
. Following soil wetting, the vapour transfer rate from the soil is high, especially for crops
having incomplete ground cover. The combined surface resistance of the canopy and of the soil
determines the (bulk) surface resistance, r
s
.

The K
c
in Equation 19 predicts ET
c
under standard, pristine conditions. This represents the upper
envelope of crop evapotranspiration and represents conditions where no limitations are placed on
crop growth or evapotranspiration due to water shortage, crop density, or disease, weed, insect or
salinity pressures. The ET
c
predicted by K
c
is adjusted if necessary to non-standard conditions, ET
c act
or

ET
a
, where any environmental condition or characteristic is known to have an impact on or to limit
ET
c
. For this reason, ET
a
is used when defining WP (Eq. 5) and ET
c
is used in Eq. 7 referring to the
maximal WP. These aspects are represented in Fig. 5.

Actual ET
c
can be less than the potential ET
c
under non-potential growing conditions including
water stress or high soil salinity. The non-potential ET
c
is termed actual ET
c
and is represented as
ET
c act
. It is defined as:
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
40


o act c act c
ET K ET = (21)
where K
c act
is the actual crop coefficient that includes effects of environmental stresses.


x =
ET
ET K K
water & environmental
stress
s c
adjusted
c adj
o
x
Radiation
Temperature
Wind speed
Humidity
climate
+
=
x =
well watered
well watered
grass
reference
crop
ET
o
ET
o
ET
c
K
c
factor
grass
crop
conditions optimal agronomic
Yield = Ym
Yield = Ya
x =
ET
ET K K
water & environmental
stress
s c
adjusted
c adj
o
x
Radiation
Temperature
Wind speed
Humidity
climate
+
=
x =
well watered
well watered
grass
reference
crop
ET
o
ET
o
ET
c
K
c
factor
grass
crop
conditions optimal agronomic
Yield = Ym
Yield = Ya
Yield = Ym
Yield = Ya
act
x =
ET
ET K K
water & environmental
stress
s c
adjusted
c adj
o
x
Radiation
Temperature
Wind speed
Humidity
climate
+
=
x =
well watered
well watered
grass
reference
crop
ET
o
ET
o
ET
c
K
c
factor
grass
crop
conditions optimal agronomic
Yield = Ym
Yield = Ya
x =
ET
ET K K
water & environmental
stress
s c
adjusted
c adj
o
x
Radiation
Temperature
Wind speed
Humidity
climate
+
=
x =
well watered
well watered
grass
reference
crop
ET
o
ET
o
ET
c
K
c
factor
grass
crop
conditions optimal agronomic
Yield = Ym
Yield = Ya
Yield = Ym
Yield = Ya
x =
ET
ET K K
water & environmental
stress
s c
adjusted
c adj
o
x
Radiation
Temperature
Wind speed
Humidity
climate
+
=
x =
well watered
well watered
grass
reference
crop
ET
o
ET
o
ET
c
K
c
factor
grass
crop
conditions optimal agronomic
Yield = Ym
Yield = Ya
x =
ET
ET K K
water & environmental
stress
s c
adjusted
c adj
o
x
Radiation
Temperature
Wind speed
Humidity
climate
+
=
x =
well watered
well watered
grass
reference
crop
ET
o
ET
o
ET
c
K
c
factor
grass
crop
conditions optimal agronomic
Yield = Ym
Yield = Ya
Yield = Ym
Yield = Ya
act


Fig. 5. Schematic representation on the relationships between reference, potential and actual crop
evapotranspiration and crop yields (adapted from Allen et al., 1998)

In its dual form, K
c
= K
cb
+ K
e
(Allen et al., 1998). The basal crop coefficient K
cb
represents the ratio
of ET
c
to ET
o
under conditions when the soil surface layer is dry, but where the average soil water
content of the root zone is adequate to sustain full plant transpiration. Additional evaporation due to
wetting of the soil surface by precipitation or irrigation is represented in the evaporation coefficient K
e
.
The total, actual K
c act
is the sum of K
cb
and K
e
reduced by any occurrence of soil water stress:


e cb s act c
K K K K + =
(22)
where K
cb
is the basal crop coefficient [0 - 1.4], K
e
is a soil water evaporation coefficient [0 - 1.4] and
K
s
is the stress reduction coefficient [0 - 1], which reduces the value of K
cb
when the average soil
water content of the root zone is not adequate to sustain full plant transpiration. K
s
is equal to 1 when
no stress occur, thus then the first term of Eq. 22 becomes equal to K
c
. K
e
represents the evaporation
component from wet soil that occurs in addition to the ET represented in K
cb
. The sum of K
cb
and K
e

can not exceed some maximum value for a crop, based on energy limitations, generally referred as
the K
c
. the value. An update and extension on the use of Eq. 22 is given by Allen et al. (2005a).

The linearized form used for mean K
c
and basal K
cb
curves in FAO-56 was introduced in FAO-24
(Doorenbos and Pruitt, 1977). The FAO K
c
curve is comprised of four straight line segments
representing the initial period, the development period, the midseason period and the late season
period (Fig. 6). These segments are defined by three primary K
c
values: K
c
during the initial period (K
c
ini
), K
c
during the midseason (full cover) period (K
c mid
) and K
c
at harvest (or at the end of the late
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
41
season) (K
c end
). The K
c ini
defines the horizontal portion of the K
c
curve during the initial period until
approximately 10% of the ground is covered by vegetation. The K
c mid
defines the value for K
c
during
the peak period for the crop, which is normally when the crop is at "effective full cover". This period is
described by a horizontal line extending through K
c mid
. The development period is defined by a
sloping line that connects the initial and midseason periods. The late season has a sloping line that
connects the end of the midseason period with the harvest (end) date.

Crop
Dev. Initial
Period Period Period Period
Mid Season
Late
Season
Time of Season, days
K
c
0.2
0.4
0.6
0.8
1.0
1.2
0.0
K
c ini
Kc mid
K
c end


Fig. 6. Schematic of the generalized Kc curve with four crop growth stages and three Kc or Kcb
values (Allen et al., 1998)

The K
c
values vary with a large number of factors as represented in Fig. 7. First they depend on
the crop through its characteristics determining the aerodynamic and surface resistances as defined
above. Thus the K
c
values and the respective crop stage durations vary from crop to crop reflecting
the respective heights and LAI determining r
a
, stomatal control, degree of soil cover by vegetation,
and plant density determining the bulk surface resistance and, the latter, influencing the albedo and
the soil evaporation component.

Secondly, the K
c
varies with the crop growth stage (Fig. 6). For annual crops, it varies from
planting to about 10% soil cover during the initial phase, then until full crop cover and from then to the
start of senescence of leaves, and finally until harvest. All factors mentioned above height, LAI,
number and functioning of stomata, soil cover gradually change along the crop season, thus also
the r
a
and r
s
, as well as the albedo and soil evaporation. For deciduous trees and shrubs changes go
from leaf initiation to full cover and, later, from starting leaf senesce to the fall of leaves. In addition to
these changes there are those in plant density while the crop is aging until attaining the target
development. Additional changes have to be considered relative to the occurrence of ground cover by
vegetation, which also uses water including during the dormancy period of the crop. For the
evergreen trees and shrubs changes in crop characteristics determining r
a
, r
s
and albedo only occur
during the first years of the crop but the influences of ground cover are also non-negligible. Finally,
there are also the permanent pastures whose characteristics vary with cuts and between cuts.

Management influences refer to soil management, which may be such that evaporation is
reduced, planting densities determining soil cover and vapour fluxes inside the canopy, and
harvesting date, since a crop may be harvested fully mature or, as for table crops, before that leaf
senescence affect the quality of the product. This influences particularly the K
c
end.


Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
42

1.0
frequent
in-
frequent
sugar cane
cotton
maize
cabbage, onions
apples
h
a
r
v
e
s
t
e
d
f
r
e
s
h
d
r
i
e
d
wetting
events
0.8
0.6
0.4
0.2
1.2
K
c
(short)
(long)
mid-season late season initial
crop
ment
develop-
60 .. % .. 25
ground cover
. 40 .
c
soil ground cover
crop type
evapo-
ration
plant
development (wind speed)
(humidity)
main factors affecting K
crop type
harvesting date
in the 4 growth stages

Fig. 7. Schematic of the variation of the Kc curve with crop, environmental and management factors
(Allen et al., 1998)

Crop management is also influencing K
c
: the crop may be managed for achieving the potential
yield Y
max
, or some aspects be poorly practiced and affect crop height, LAI and the stomatal control.
These factors include seedbed preparation, seeding dates, plant density, fertilising, pest and diseases
control, weed control, and irrigation.

In addition, the K
c
are influenced by the climate although the main climatic influences are
incorporated in ET
o
. The climate largely determines the duration of the crop growth stages.

The K
c ini
for most crops excepting the evergreen ones is essentially representing soil water
evaporation since the crop is not covering but a small fraction of the soil. It is therefore determined by
the frequency and amount of wettings during this stage and by the potential evaporation rate from the
soil. It is also influenced by the amount of water available in the upper soil layer, of 10 to 15 cm and
by the soil water holding capacity and its capillary rise potential to bring water stored below into this
evaporative soil layer. This process and formulation is well described by Allen et al. (2005b). Factors
referred above may largely vary from one location to another and, for the same location, from one
crop season to the next. This explains why a large variation of values is shown in Fig. 7. In addition,
differences in management add to those factors as referred in the following.

The K
c ini
may be largely modified by management practices such as direct seeding (no tillage), soil
mulching by straw or plastic, tunneling with plastic as for horticultural crops, and other similar
practices. Two effects occur: on the one hand, there is a decrease of energy at the soil surface, thus
a decrease of the evaporation rates; on the other hand, there is an increased resistance to vapour
transfer from the soil surface into the atmosphere, mainly relative to an increase in the surface
resistance. Therefore, K
c ini
may be much lower than under conditions when the soil is fully exposed to
radiation. However, impacts referred in literature are variable reflecting differences in soil cover; in
case of vegetal mulching this highly depend on the amount and distribution of soil coverage; in case
of plastic mulch they depend on its transparency to radiation of short and long wave, and on the
amount of openings in the plastic through where vapour may escape to the atmosphere. Anyway, in
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
43
general, soil covering with mulch reduces K
c ini
, thus soil evaporation, not affecting transpiration of the
canopy. Effects are kept but progressively reduced during the crop development phase and mostly
disappear at the mid-season for full cover crops. Adopting the dual crop coefficient approach (Eq. 22)
these effects are better studied since it becomes possible to separate the soil evaporation and the
transpiration components of ET
c


In case of tree and shrub crops where the soil is covered by active vegetation, which is often
required as a measure to combat erosion, the impacts of soil cover are totally different since this
vegetation also uses water and the crop water requirements of such crops are then increased relative
to bare soil conditions.

The K
c mid
essentially varies with the crop but is also affected by the climate, namely when
advective conditions occur. This relation with climate is analysed by Pereira et al. (1999). Therefore,
an adjustment to climate is proposed by Allen et al. (1998) aimed at exporting to other climates the
tabled values for K
c mid
. In fact these tabled values refer to a standard climate where the wind speed
u
2
is 2 m s
-1
and the minimum relative humidity RH
min
is 45% during the mid season of the considered
crop. Then the adjustment to any other climate is performed through the following equation:
| |
3 0
2
3
45 004 0 2 04 0
.
min ) climate standard ( mid c mid c
h
) RH ( . ) u ( . K K
|
.
|

\
|
+ = (23)

where K
c mid
and K
c mid (standard climate)
refer to the location where the application is performed and to the
standard conditions, and h is the crop height [m] during mid-season.

This equation shows well the aspects referred above relative to ra since u
2
is a main variable
controlling ra and determining this equation. he dominant. K
c
increases with wind speed and
decreases inversely. Therefore, reducing ET at the mid-season may be obtained by avoiding high
winds as it is commonly done in arid lands, either through cropping in areas less exposed to wind or
using wind breaks. RH
min
, for a certain extend, represents well the conditions for diffusion of vapour
into the atmosphere, very strong in arid climates where RH
min
is low. Changing RH to control
evaporation is generally non practical, but RH tends to increase when wind is reduced and thus the
transport of vapour from the air layer close to the crop. Eq. 23 shows that these impacts are larger for
tall crops and smaller for short ones, which agrees with the above referred increase of r
a
when the
crop is of low height.

The K
c end
is largely affected by management which determines harvesting, thus the end of the
crop season (Fig. 7). Moreover, harvesting earlier increases K
c end
relative to a late harvesting of the
same crop but shortens the duration of the end-season period. Harvesting earlier is practiced for food
crops that should be eaten fresh, and harvesting later is adopted for crops when conservation or
preservation is easier when they are stored dry, as for cereals. Climate also impacts the K
c end
when
the crop is harvested fresh. Then Eq. 23 applies but variables refer to the the late season. Controlling
ET during this period refers to the same aspects as for the mid-season.

K
c
values are well known for the temperature climate crops, mainly the annual crops (Allen et al.,
1998); some deficiencies in knowledge occur for tree crops due to differences in plant density, ground
cover by vegetation or mulch, and architecture of the plantations. However, literature is producing
consistent knowledge that provides for adopting coherent values for K
c
in other regions. Main gaps
refer to tropical and sub-tropical crops, which research is less abundant and largely published in
native languages. A great effort in improving the corresponding base-knowledge is necessary.

Frequent gaps in practical knowledge refer to the lack of adoption of appropriate estimation of K
c ini

when using the rough indicative values in Tables, and the lack in adjusting the K
c mid
and K
c end
to
climate (Eq. 23). Moreover, it is the use of K
c
without referring to the fact that the crop is managed for
potential yield or not, thus that crop factor is not differentiate between K
c
and K
c act
(Eq. 19 and 21).
Then is often said that FAO 56 tabled values are not appropriate when the main problem is to use
only part of the information produced in the guidelines and omitting the use of all other adjusting tools.

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
44
It is necessary to underline herein that looking to develop water productivity assessment without
fully considering (and understanding) the concepts and calculation tools for crop evapotranspiration is
inadequate since the main component in computing irrigation depths is Etc or ETa and, without
knowing these, one can only roughly know how much is the non-beneficial water use term NBWU
(Eq. 7 and 11).


WATER STRESS AND IMPACTS ON YIELD

The basic relationship between crop ET and yields may be represented by the Stewart model
(Stewart et al., 1977, Doorenbos and Kassam, 1979) relating the relative yield decrease with the
relative evapotranspiration deficit (see Fig. 5)

|
|
.
|

\
|
=
|
|
.
|

\
|

c
act c
y
m
ET
ET
K
Y
Ya
1 1 (24)
where K
y
is the yield response factor [-], ET
cadj
is the adjusted (actual) crop evapotranspiration [mm d
-
1
], ET
c
is the crop evapotranspiration for standard conditions (no water stress) [mm d
-1
], Ya is the crop
yield when ET = ET
cadj
, and Ym is the maximal crop yield corresponding to ET
c
. A better description
of ET impacts on yields is obtained when the yield response factors refer to specific crop phases and
the history of the crop stresses is taken into consideration as it is largely reported in the literature.

Recombining the terms of Eq. 24, the stress coefficient Ks (Eq. ) may be expressed as

|
|
.
|

\
|
=
m
a
y
s
Y
Y
K
K 1
1
1
(25)
which expresses how the relative yield decrease impact the stress coefficient as a function of the
yield response factor. However, for operational purposes, it is better to express Ks as a function of the
soil water depletion:

|
|
.
|

\
|

=
RAW TAW
D TAW
K
r
s
(26)
where TAW and RAW are respectively the total and readily available soil water, and Dr is the
cumulated soil water depletion between two wetting events by rain or irrigation. Eq. 26 indicates that
Ks < 1 when soil water is depleted below the RAW threshold.

Yields may be affected by other environmental and management factors such as salinity (Hamdy
and Karajeh, 1999; Minhas, 1996; Rhoades et al., 1992). A simplified approach for salinity impacts for
conditions where ECe > ECe threshold is
( )
100
1
b
EC EC
Y
Y
threshold e e
m
a
= (27)
where Y
a
is the actual crop yield, Y
m
is the maximum expected crop yield when EC
e
< EC
e threshold,
EC
e
is the mean electrical conductivity of the saturation extract for the root zone [dS m
-1
], EC
e threshold
is the
electrical conductivity of the saturation extract at the threshold of EC
e
when crop yield first reduces
below Y
m
[dS m
-1
], and b is the reduction in yield per increase in EC
e
[%/(dS m
-1
)]. Values for
EC
e threshold
and b are tabled by Rhoades et al. (1992) and Allen et al. (1998).

The combined impacts of water and salinity stress are expressed through

( )
|
|
.
|

\
|

|
|
.
|

\
|
=
RAW TAW
D TAW
EC EC
K
b
K
r
threshold e e
y
s
100
1 (28)

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
45
indicating that Ks < 1 when either the ECe threshold or the soil water threshold
t
(corresponding to
RAW) are attained (Fig. 8).

The concepts described by Eq. 24 to 28 are the base for deficit irrigation together with the
knowledge of the crop development phases when water stress impacts are smaller. Using
those equations it is then possible to easily compute


o act c act c
ET K ET = (21 bis)
from appropriate estimation of Ks given as above

e cb s act c
K K K K + =
(22 bis)
or simply

c s act c
K K K =
(29)

0.00
0.20
0.40
0.60
0.80
1.00
: soil water content
FC t WP
D : depletion from root zone (mm)
r
K
s
0 RAW TAW
low
high
with soil salinity
n
o

s
o
i
l

s
a
l
i
n
i
t
y


Fig. 8. Schematic of soil water and salinity determining the stress coefficient Ks (Allen et al., 1998)

The great difficulty in adopting an irrigation management that allows for crop stress during
selected phases of the crop season is the insufficient knowledge about the respective economic
impacts. The literature is abundant on which crop phases are less or more sensitive to water and salt
stress; numerous field observation tools allow to assess the soil water status and, less often, the
salinity conditions; a variety of models may be used to simulate the soil water balance and therefore
provide appropriate information for irrigation scheduling. Although a few tentative exist (Victoria et al.,
2005), the great gap refers to combining physical assessment and simulation with economic
assessment of impacts.

In an example referring to Tunisia and Portugal (Rodrigues et al., 2003; Zairi et al., 2003) it is
shown that more than following the ET-yield relations, it is necessary to analyse the relationships
between ET or irrigation and the economic impacts of deficit irrigation. Reducing ET leads to reduced
yields (land productivity) but increased water productivity. This option may be feasible when
decreasing Gross Margins per unit land (GM/ha), i.e. the economic land productivity, it results in
increased GM per unit water, thus increased economic water productivity as exemplified in Fig. 9

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
46
0
200
400
600
800
1000
1200
1400
0 40 80 120 160 200 240 280
Season irrigation (mm)
G
M


(
U
S
D

/

h
a
)
(1) (2) (3) (4) (5) (6) (7)
Number of irrigations
0.0
0.5
1.0
1.5
2.0
0 40 80 120 160 200 240 280
Season irrigation (mm)
G
M


(
U
S
D

/

m
3
)
(1) (2) (3) (4) (5) (6) (7)
Number of irrigations


Fig. 9. Gross margins per unit surface (a) and per unit volume of water applied (b) for alternative
deficit irrigation strategies for the wheat crop under average ( ), and very high ( )
climatic demand conditions (Zairi et al., 2003)

The farmer incomes then reduces but, when water is lacking, that income is higher than reducing the
cropped area. However, for crops growing out of the rainy season, these relations are different as
shown for a tomato crop in the same region (Fig. 10)


0
1000
2000
3000
4000
5000
240 320 400 480 560 640 720 800 880
Season irrigation (mm)
G
M


(
U
S
D

/

h
a
)
(6) (8) (10) (12) (14) (16) (18) (20) (22)
Number of irrigations
0.0
0.2
0.4
0.6
0.8
240 320 400 480 560 640 720 800 880
Season irrigation (mm)
G
M


(
U
S
D

/

m
3
)
(6) (8) (10) (12) (14) (16) (18) (20) (22)
Number of irrigations


Fig. 10. Gross margins per unit surface (a) and per unit volume of water applied (b) for alternative
deficit irrigation strategies of tomato crop in Siliana for average ( ), high ( ) and very
high ( ) demand conditions (Zairi et al., 2003).

Thus, for crops having a high demand for water such as summer crops, including if they have a
favourable ratio between yield price and water cost, as it is the case for tomato in Tunisia, when
GM/ha decreases due to less water application the economic water productivity do not increase but
may decrease. It is then questionable to adopt deficit irrigation. This response is different with
different structure of production costs as exemplified for center-pivot irrigated maize in Portugal (Fig.
11).
(a) (b)
(a) (b)
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
47

Fig.11. Variation of the gross margins per hectare cropped (---) and per m3 of water applied (___)
relative to the maize crop when irrigation depths decrease from full irrigation to heavy deficits
for average ( ), high ( ) and very high ( ) climatic demand conditions (Rodrigues et
al., 2003)


CONCLUSION REMARKS

The analysis above shows that the recently adopted concept of water productivity may be
advantageous relative to the old concept of irrigation efficiency. However, it is required to well define
both the numerator and the denominator of the water productivity term in order to further perform
appropriate analysis of irrigation performances in relation to yield.

Crop evapotranspiration is an essential term when assessing water productivity. Thus, it is
required to well understand which factors lead to potential or non-potential (below the potential) crop
evapotranspiration and how these factors may be manipulated or managed to achieve a controlled
crop demand with minimal negative impacts on yields. Current knowledge provides the appropriate
tools for such assessment despite gaps in knowledge relative to tropical and sub-tropical crops in
particular. However the main gaps identified in the practice refer to a less good use of existing tools
and concepts. Improved modelling tools may help solving this but making better use of present know
how is critical.

Moreover, it shows to be essential the appropriate combination of physical assessment tools with
economic assessment. Economic water productivity seems to be relevant since farmers decisions are
based in income considerations. Therefore, an improved knowledge about the economic relations is
required and economic considerations must be integrated with engineering approaches and not left
as external to be just analysed by economic specialists.


REFERENCES

Allen, R.G., Willardson, L.S., Frederiksen, H.D., 1997. Water use definitions and their use for
assessing the impacts of water conservation. In: J. M. de Jager, L.P. Vermes, and R. Ragab (Eds.)
Sustainable Irrigation in Areas of Water Scarcity and Drought (Proc. ICID Workshop, Oxford),
British Nat. Com. ICID, Oxford, pp. 72-81.
Allen, R.G., Smith, M., Perrier, A., Pereira, L.S., 1994. An update for the definition of reference
evapotranspiration. ICID Bulletin, 43(2): 1-34.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
48
Allen, R.G., Pereira, L.S., Raes, D., Smith, M., 1998. Crop Evapotranspiration. Guidelines for
Computing Crop Water Requirements. FAO Irrig. Drain. Paper 56, FAO, Rome, 300 p.
Allen, R.G., Pereira, L.S., Smith, M., Raes, D., Wright, J.L., 2005a. FAO-56 Dual crop coefficient
method for estimating evaporation from soil and application extensions. J. Irrig. Drain. Engng.
131(1): 2-13.
Allen, R.G., Pruitt, W.O., Raes, D., Smith, M., Pereira, L.S., 2005b. Estimating evaporation from bare
soil and the crop coefficient for the initial period using common soils information. J. Irrig. Drain.
Engng. 131(1): 14-23.
Alves, I, Pereira, L.S., 2000. Modelling surface resistance from climatic variables? Agric. Water
Manag. 42: 371-385.
Alves, I., Perrier, A., Pereira, L.S., 1998. Aerodynamic and surface resistances of complete cover
crops: How good is the big leaf approach? Trans. ASAE 41(2): 345-351.
Baldocchi, D.D., Luxmoore, R.J., Hatfield, J.L., 1991. Discerning the forest from the trees: an essay
on scaling canopy stomatal conductance. Agric. For. Meteorol. 54: 197-226.
Bos, M.G., Burton, M.A., Molden, D.J., 2005. Irrigation and drainage Performance Assessment.
Practical Guidelines. CABI Publ. Wallingford.
Brutsaert, W., 1982. Evaporation into the Atmosphere. R. Deidel Publ. Co, Dordrecht,
Burt, C.M., Clemmens, A.J., Strelkoff, T.S., Solomon, K.H., Bliesner, R.D., Hardy, L.A., Howell, T.A.,
Eisenhauer, D.E., 1997. Irrigation performance measures: efficiency and uniformity. J. Irrig. Drain.
Engng. 123: 423-442.
FAO, 2002. Crops and drops: making the best use of water for agriculture. FAO, Land and Water
Development Division, Rome, Italy
Hamdy, A., Karajeh, F., (Eds.) 1999. Marginal Water Management for Sustainable Agriculture in Dry
Areas. (Proc. Advanced Short Course, Aleppo, Syria), ICARDA, Aleppo and CIHEAM/IAM-B,
Istituto Agronomico Mediterraneo, Bari.
Jensen, M.E., 1996. Irrigated agriculture at the crossroads. In: Pereira, L. S., Feddes, R. A., Gilley, J.
R., Lesaffre, B. (Eds.) Sustainability of Irrigated Agriculture. Kluwer Acad. Publ., Dordrecht, pp. 19-
33.
Jensen, M.E., Burman, R.D., Allen, R.G. (Eds.), 1990. Evapotranspiration and Irrigation Water
Requirements. Am. Soc. Civ. Eng. Manual No. 70, 332 pp.
Keller, J., Bliesner, R.D., 1990. Sprinkler and Trickle Irrigation. Van Nostrand Reinhold, New York,
652 pp.
Lamaddalena, N., Pereira, L.S., 1998. Performance analysis of on-demand pressurized irrigation
systems. In: Pereira, L.S., Gowing, J.W. (Eds.) Water and the Environment: Innovation Issues in
Irrigation and Drainage, E &FN Spon, London, pp. 247-255.
Lamaddalena, N., Sagardoy, J.A., 2000. Performance Analysis of On-Demand Pressurized Irrigation
Systems. FAO Irrigation and Drainage Paper 59, FAO, Rome, 132 pp.
Mantovani, E.C., Villalobos, F.J., Orgaz, F., Fereres, E., 1995. Modelling the effects of sprinkler
irrigation uniformity on crop yield. Agric. Water Manage. 27: 243-257.
Minhas, P.S., 1996. Saline water management for irrigation in India. Agric. Water Manage. 38: 1-24.
Molden, D.J. and Gates, T.K., 1990. Performance measures for evaluation of irrigation-water-delivery
systems. J. Irrigation and Drainage Engineering 116(6): 804-823.
Molden, D., Murray-Rust, H., Sakthivadivel, R., Makin, I., 2003. A water-productivity framework for
understanding and action. In: Kijne JW, Barker R, Molden D (eds.), Water Productivity in
Agriculture: Limits and Opportunities for Improvement, IWMI and CABI Publ., Wallingford, pp 1-18.
Monteith, J.L., 1965. Evaporation and the environment. XIXth Symposia of the Society for
Experimental Biology. In the State and Movement of Water in Living Organisms. University Press,
Swansea, Cambridge, 205234.
Monteith, J.L. 1973. Principles of Environmental Physics. Edward Arnold, London.
Ortega, J.F., de Juan, J.A., Tarjuelo, J.M., 2005. Improving Water Management: The Irrigation
Advisory Service of Castilla la Mancha (Spain). Agric. Water Manage. 77: 37-58.
Pereira, L.S., 1999. Higher performances through combined improvements in irrigation methods and
scheduling: a discussion. Agric. Water Manage. 40 (2): 153-169.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
49
Pereira, L.S., 2003. Performance issues and challenges for improving water use and productivity
(Keynote). In: T. Hata and A H Abdelhadi (Session Organizers) Participatory Management of
Irrigation Systems, Water Utilization Techniques and Hydrology (Proc. Int. Workshop, The 3rd
World Water Forum, Kyoto), Water Environment Lab., Kobe University, pp. 1-17.
Pereira, L.S., Trout, T.J., 1999. Irrigation methods. In: van Lier, H.N., Pereira, L.S., Steiner, F.R.
(Eds.) CIGR Handbook of Agricultural Engineering, vol. I: Land and Water Engineering, ASAE, St.
Joseph, MI, pp. 297-379.
Pereira, L.S., Perrier, A., Allen, R.G., Alves, I., 1999. Evapotranspiration: Review of concepts and
future trends. J. Irrig. Drain. Engng. 125(2): 45-51.
Pereira, L.S., Cordery, I., Iacovides, I., 2002a. Coping with Water Scarcity. UNESCO IHP VI,
Technical Documents in Hydrology No. 58, UNESCO, Paris, 267 p. (accessible through
http://unesdoc.unesco.org/images/0012/001278/127846e.pdf)
Pereira, L.S., Oweis, T., Zairi, A., 2002b. Irrigation management under water scarcity. Agric. Water
Manag. 57: 175-206.
Pereira, L.S., Calejo, M.J., Lamaddalena, N., Douieb, A., Bounoua, R., 2003. Design and
performance analysis of low pressure irrigation distribution systems. Irrigation and Drainage
Systems 17(4): 305-324.
Perrier, A., 1975. Etude physique de l'vapotranspiration dans les conditions naturelles. III -
vapotranspiration relle et potentielle des couverts vgtaux. Ann. Agronomiques 26: 229-243.
Perrier, A., 1982. Land surface processes: vegetation. In: Eagleson, P.S. (ed) - Land Surface
Processes in Atmospheric General Circulation Models, Cambridge Univ. Press, Cambridge,
Mass., pp 395 - 448.
Rhoades, J.D., Kandiah, A., Mashali, A.M., 1992. The Use of Saline Waters for Crop Production. FAO
Irrigation and Drainage Paper 48, FAO, Rome.
Rochette, P.; Pattey, E.; Desjardins, R.L.; Dwyer, L.M.; Stewart, D.W.; Dube, P.A., 1991. Estimation
of maize (Zea mays L.) canopy conductance by scaling up leaf stomatal conductance. Agric. For.
Meteorol. 54: 241-261.
Rodrigues, P.N., T. Machado, L.S. Pereira, J.L. Teixeira, H. El Amami, A. Zairi, 2003. Feasibility of
deficit irrigation with center-pivot to cope with limited water supplies in Alentejo, Portugal. In: G.
Rossi, A. Cancelliere, L. S. Pereira, T. Oweis, M. Shatanawi, A. Zairi (Eds.) Tools for Drought
Mitigation in Mediterranean Regions. Kluwer, Dordrecht, pp. 203-222.
Rossi, G., Cancellieri, A., Pereira, L.S., Oweis, T., Shatanawi, M., Zairi, A. (eds.), 2003. Tools for
Drought Mitigation in Mediterranean Regions. Kluwer, Dordrecht, 357 p.
Steduto, P., 1996. Water use efficiency. In: Pereira, L. S., Feddes, R. A., Gilley, J. R., Lesaffre, B.
(Eds.) Sustainability of Irrigated Agriculture. Kluwer, Dordrecht, pp. 193-209.
Stockle, C.O.; Kjelgaard, J., 1996. Parameterizing Penman-Monteith surface resistance for estimating
daily crop ET. In: Camp, C.R.; Sadler, E.J.; Yoder, R.E. (eds) Evapotranspiration and Irrigation
Scheduling (Proc. Int. Conf., San Antonio, Texas, 3-6 Nov), ASAE, St. Joseph, MI, pp. 697-703.
Victoria F.B., Viegas Filho J.S., Pereira L.S., Teixeira J.L., Lanna A.E., 2005. Multi-scale modeling for
water resources planning and management in rural basins. Agric. Water Manage. 77: 4-20.
Zairi, A., H. El Amami, A. Slatni, L.S. Pereira, P.N. Rodrigues, T. Machado, 2003. Coping with
drought: deficit irrigation strategies for cereals and field horticultural crops in Central Tunisia. In: G.
Rossi, A. Cancelliere, L. S. Pereira, T. Oweis, M. Shatanawi, A. Zairi (Eds.) Tools for Drought
Mitigation in Mediterranean Regions. Kluwer, Dordrecht, pp. 181-201.
Zhang, H., Oweis, T., 1999. Water-yield relations and optimal irrigation scheduling of wheat in the
Mediterranean region. Agric. Water Manag. 38: 195-211.









Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity
50



































Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

51
SYSTEMATIC APPROACH TO THE IMPROVEMENT
OF AGRICULTURAL WATER USE EFFICIENCY



T.C. Hsiao
Professor Emeritus, Dept. of Land, Air, and Water Resources, University of California, Davis.
1 Shields Ave., Davis, California, 95616, U. S. A. Email: tchsiao@ucdavis.edu


SUMMARY - Effective management of scarce water resources requires a systems approach. Starting
at the source of water, a cascade of events leads to the final production of crops or animal products at
the expense of water. These events are mostly sequential, with each process step in the sequence
having its own efficiency of output per unit of input. Using a simple sequence of three hypothetical
steps, it is shown that the overall efficiency of a process is the product of the efficiencies of each
sequential step. That is, efficiencies of individual process steps are multiplicative in determining the
overall efficiency. Thus, improvement in any one of the efficiency steps has equal effect in improving
the overall efficiency, and the overall improvement is more than the sum of the individual
improvements. This principle provides a simple and quantitative means to optimize the allocation of
limited resources in improving water use efficiency.
Crop production in relation to water use are considered in terms of the pertinent sequence of
efficiency steps for irrigated conditions. Rainfed conditions will be considered in another presentation
in the Rainfed and Drought session. Efficiency steps and the sequences are outlined and discussed
and the likely improvements assessed quantitatively for some scenarios. The universal applicability of
this approach to different cropping as well as animal production systems when water is limiting is
emphasized.

Key words: Irrigation, crop productivity, water saving, management, resource allocation, optimization.


INTRODUCTION

The relentless growth of human population, coupled with the intensifying desire for higher living
standard, including the continuous shifting to diets based more and more on meat and dairy products,
are straining the water resources all over the world, especially in the more arid regimes. Adding to the
problem is the increased awareness of the need for water in the preservation of the environment and
ecosystems. Since the fresh water resources are essentially finite on earth, making more efficient use
of the water must be a major focal point in coping with water shortage. Numerous ways have been
devised or advocated and major efforts have been made to improve the efficiency of water use in
agriculture. The production of crops and animals with water as a key input involves complicated
processes with myriad of facets that are subjected to the impact of management decisions and
environmental influence. A systematic and quantitative approach is needed to analyze where the
inefficiency lies, to assess the potential improvements, and most importantly, to determine how to
allocate limited available resource to maximize the improvement in water productivity. This paper
describes briefly a relatively simple and yet quantitative and comprehensive framework for these
purposes. A more complete treatment is given in a paper in the forthcoming special issue of Irrigation
Science (Hsiao et al., 2005).


THE CONCEPT OF CHAIN OF EFFICIENCY STEPS AND ITS SIGNIFICANCE

Generally and as commonly used in economics, efficiency of any production process may be
defined as the ratio of input to output for that process, both measured in quantitative units. The units
to use vary depending on the situation; if the same units define both, then the efficiency ratio is
unitless. For example, if the resource input as well as the production output is measured in monetary
units such as dollars or euros, the efficiency ratio (or simply efficiency) would be in fractions or
percentage. If the measure of input and output are in different units, then the units for the efficiency
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

52
must be given for the efficiency to be meaningful. For example, fuel efficiency of a car may be
expressed in km per liter, the ratio of the distance traveled to the volume of gasoline consumed.

When the production of a product is complicated and the starting resource input goes through
many processing steps sequentially ending in the product, a simple approach is available to quantify
the overall efficiency of the whole in terms of the efficiency of each of the component steps. Because
the processing steps are in sequence and comes one after another, the output of the first step is the
input of the second step, and the output of the second step is the input of the third step, etc. In
equation form:

1 1 i
Input Output
+
= , and

1
1
1
Input
Output
E = (1a)

1
2
2
2
2
Output
Output
Input
Output
E = = (1b)

2
3
3
3
3
Output
Output
Input
Output
E = = (1c)
where E designates efficiency of a step in the efficiency chain, and the subscripts i, a running number,
designates the steps; 1, 2, and 3 refer to the specific steps, 1 being the first step and 3 being the last
(third).

If there are only three steps in the whole efficiency chain, the overall efficiency (E
all
) would be the
ratio of the final output (output
3
) to the initial input (input
1
), i. e.,

1
3
all
Input
Output
E =
Because the steps are sequential, the output of the preceding step is the input of the following step, as
can be seen by a close examination of Eq. 1a, 1b, and 1c. This gives rise, inevitably, the following
relationship between the efficiency of individual steps and the overall efficiency:

2
3
1
2
1
1
all
Output
Output
Output
Output
Input
Output
E = (2)

It is easily seen from the right side of the equation that the numerator of the first fraction cancels
out the denominator of the second fraction, and the numerator of the second fraction cancels out the
denominator of the third fraction, leaving only the ratio of the last output (output
3
) to the first input,
(input
1
), which is E
all
. So the overall efficiency is the product of the individual efficiency steps as long
as the steps for the whole process are sequential. This simple mathematical outcome holds true
regardless of the number of individual steps in the whole process, although Eq 2 is written for an
efficiency chain consisting only of three steps.

When analyzing a production process, it is important not only to know the efficiencies of the
different component steps, but also to know how improvements in the efficiency of the steps affect the
overall efficiency. It turned out that by expressing the improvement as a fraction of the original
efficiency, a simple equation to calculate the new overall efficiency can be obtained. Denoting the
fractional improvement by , an expression for the improved efficiency of a step (E
new
) is:

original new
E ) 1 ( E + = (3)
Applying Eq. 3 to all the steps in an efficiency chain and designating each step by the running
number j (j = 1, 2, 3, etc. depending on the position of the step in the chain), a general expression of
the new overall efficiency (E
all
,new) in terms of and the original overall efficiency (E
all
, original) is as
follows:
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

53
( )
j
j
original , all new , all
1 E E + = (4)
where is the multiplication operator over items j. Expressed in words, one plus the fractional
improvement for each step, when multiplied together, and multiplied again by the original overall
efficiency, is the new overall efficiency. Eq. 4 is general, and can be applied to any efficiency chain. It
also applies to cases where there is a reduction in efficiency of some or all the steps, simply by
denoting the fractional change in efficiency () as negative.

There are some important features to note regarding Eq. 2 and 4: (1) The treatment is quantitative,
and by simple mathematics, demonstrates the fact that the overall efficiency is the products of the
efficiencies of individual steps (and not the average of the efficiencies). (2) Even though the efficiency
of each step may be high, the overall efficiency is considerably or much lower because of the
multiplicative effect of individual efficiencies. (3) By the same token, the same multiplicative effect
makes it possible to improve the overall efficiency substantially by making minor improvement in
several of the individual efficiencies. (4) The impact of a change in the efficiency of one step on the
overall efficiency is strictly according to the proportional change in the efficiency of that step,
regardless of where the step is located in the efficiency chain or how efficient the step is originally.
Some of these features may not be intuitively obvious until some examples are given. Befitting the
objectives of this conference and as an example, the chain of efficiency steps concept is applied in
the next section to irrigated crop production to quantify water productivity or water use efficiency.


EFFICIENCY OF IRRIGATED CROPPING AND POTENTIAL FOR MPROVEMENT

The chain of efficiency steps approach, though not so called, is sometimes used in the literature to
evaluate the delivery of water from a reservoir or other sources to the soil of the root zone of the crop.
This covers the water and irrigation engineering aspects but not the agronomic and crop aspects. In
this paper the concept is extended all the way to crop yield, starting from water diversion from the
reservoir. Beginning with the engineering aspects, one may divide up the processes into some
obvious sequential steps. Water, the input, is first conveyed from the reservoir outlet to the farm gate,
and this constitutes the first efficiency step in the whole process. The efficiency of this step may be
termed conveyance efficiency (E
conv
) and is calculated as the ratio of the quantity of water (W)
diverted out of the reservoir (W
vo
) for that farm, to the quantity of water received at the farm gate
(W
fg
). The water loss along the way is by leakage and also commonly by evaporation. The efficiency
of this step depends of course on the circumstances and engineering and management practices,
and can vary from very low to very high. In Table 1 the range of efficiency for this first and each
following step are given, one for poor situations when the efficiencies are low, and one for good
situations when the efficiencies are high. These ranges are based on literature and our general
understanding and do not include the more extreme values, especially those in the poor situation
category. Also given in Table 1 are the overall efficiency (E
all
) for the poor and good situations,
calculated according to Eq. 2 from the mid-value (average of the two limits of the range) of each step
efficiency. In addition, the numerator and denominator of the efficiency ratio for each step are also
given, as well as the efficiency units.

After the water arrives at the farm, it is stored or not stored depending on the farmer, and
distributed to the fields for irrigation. For simplicity, we will combine the storage and on farm
conveyance to the field into one step and call its efficiency farm efficiency (E
farm
). The output is water
at the field edge (W
fd
) and the input is water at the farm gate (W
fg
). The ranges of efficiency for the
poor and good situations are also given in Table 1. Once the water is at the field edge, it is applied as
irrigation to the crop in the field. The crop can only use the water retained in its root zone (W
rz
), water
that runs off the surface of the field or drains below the root zone represents losses. This step is well
known in irrigation engineering and its efficiency is designated as application efficiency (E
appl
). The
output is W
rz
, and the input, W
fd
. Applying Eq. 2 to link the three efficiency steps described, as well as
the subsequent five steps leading to crop yield to be described later, the whole efficiency chain and
the overall efficiency are:

all
vo
yld
bm
yld
as
bm
tr
as
et
tr
rz
et
fd
rz
fg
fd
vo
fg
E
W
m
m
m
m
m
W
m
W
W
W
W
W
W
W
W
W
W
= = (5)
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

54
Again, because the output of the preceding step is the input of the following step, all the terms on the
left side of Eq. 5 cancel out except for the denominator of the first and numerator of the last efficiency.
Note that the efficiency steps do not have to be all in the same units and can involve quantities of
different nature. In this case the first five steps are all concerned with quantity of water (W), and the
last two steps are concerned with mass of materials of different nature. Efficiency of the sixth step is
the mass (m) of carbon dioxide assimilated per unit of water transpired by the crop. Units of each
efficiency as used in this paper are given in Table 1.

Table 1. Range of efficiencies for the steps in the efficiency chain from water diverted out of the
reservoir to yield of annual grain (or fruit) crops. Two ranges are given, one for poor circumstances
and practices, and the other for good circumstances and practices. Also given are the overall
efficiency for the two situations, calculated from mid-values of the efficiency steps. The denominator
of the efficiency ratio is the input, and the numerator, the output, for each efficiency step.
Efficiency

Efficienc
y step


Efficiency
ratio


Units
Poor
circumstances
and practices
Good
circumstances
and practices
E
conv
W
fg
/W
vo
unitless 0.5 0.7 0.8 0.96
E
farm
W
fd
/W
fg
unitless 0.4 0.6 0.75 0.95
E
appl
W
rz
/W
fd
unitless 0.3 -0.5 0.7 0.95
E
et
W
et
/W
rz
unitless 0.85 0.92 0.97 0.99
E
tr
W
tr
/W
et
unitless 0.25 0.5 0.7 0.92
E
as
m
as
/W
tr
kg
CO2
m
water
-3
6.0 8.0 9 14
E
bm
m
bm
/m
as
kg
biomass
kg
CO2
-1
0.22 0.36 0.4 0.5
E
yld
m
yld
/m
bm
unitless 0.24 0.36 0.44 0.52
E
all
m
yld
/W
rz
kg m
-3
0.0242 1.22

With the chain of efficiency steps fully written out in Eq. 5, we now return to describe the remaining
steps (from the fourth step onward), which concern the plant and agronomic aspects. The fourth step
is consumptive efficiency (
rz et et
W W E = ), a measure of the proportion of water in the root zone
removed by evapotranspiration (W
et
). The loss of efficiency in this step is due to water left in the soil
at harvest time. The next step is transpiration efficiency (
et tr tr
W W E = ), a measure of the proportion
of water taken up by the crop and transpired (W
tr
), as distinguished from water evaporated from the
soil. The next step is assimilation efficiency (
tr as as
W m E = ), a measure of the mass of carbon dioxide
assimilated by photosynthesis (m
as
) relative to the volume of water transpired. The measurements
here now include the mass of assimilated carbon dioxide as well as the volume of water. The next
step is biomass conversion efficiency (E
bm
), a measure of the plant biomass produced (m
bm
) relative
to the mass of carbon dioxide assimilated. This efficiency is primarily determined by the chemical
composition of the crop and is not easily changed. The last step is yield efficiency (E
yld
), a measure of
the proportion of plant biomass that ends up in the harvested yield (m
yld
), and is equivalent to harvest
index (HI), a well known parameter in the crop and agronomic literature.

The most striking results (Table 1) of applying Eq. 2 or 5 to irrigated cropping is that the difference
in overall water use efficiency (last line, Table 1) between the poor situation and the good situation is
huge, in spite of the fact that for each efficiency step the difference between the two situations is not
that large or even minor. Nonetheless, E
all
for the poor situations is only 2% of E
all
for the good
situations. The reason for this huge difference lies in the multiplicative nature of the efficiency chain,
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

55
as already noted. This 50 fold difference in water use efficiency to produce yield (grain or fruits of
annual crops) indicate that there is much room for improvement in many situations. It should also be
noted that the comparison is not between the extremely poor and the extremely good situations, but
between the mid-values of the efficiency steps for the two situations.


DETERMINANTS OF EFFICIENCY OF THE STEPS AND IMPROVEMENTS OF EFFICIENCY

Some of the more important factors that impact the various efficiency steps are now discussed
briefly, along with potential improvements that can be made at relatively low costs, and a sample
improvement in the overall efficiency is calculated to illustrate the potentials. Starting with the first step
of the chain of efficiencies, a poor E
conv
implies leaky conduits or open conveyance over long distance
with much loss by evaporation. Improvement could be very costly (e.g., converting open channel to
closed conduits) or at least more than nominal (e.g., repairing cracks widely spread along the conduit
length). The next step efficiency, E
farm
, is more amenable to improvement. A common cause for low
E
farm
is water leakage from unlined or poorly lined storage pond and conveyance ditches. Lining with
plastic sheeting could be relatively inexpensive and could raise E
farm
from poor to the good level in
Table 1. The next step, E
appl
, may also be improved at nominal cost. One common cause of low E
appl

for surface irrigation is applying the water too fast or too slow relative to the infiltration rate of the soil
and slope of the land, resulting, respectively, either in too much deep drainage at the head, or too
much drainage at the tail end, of the field. Better control of the application rate to match the infiltration
rate and slope should entail only minimum cost. For sprinkler irrigation, E
appl
may be improved by
avoiding irrigating under strong wind, and by pressurizing the sprinkle line adequately to ensure even
water distribution. The next step, E
et
, is already relatively high for the poor situation; improvement is
more readily made in E
tr
. Low E
tr
is the result of too much soil evaporation relative to crop
transpiration. Since soil evaporation is high when coverage of the ground by foliage canopy of the
crop is low and when the soil surface is frequently wetted (Ritchie and Burnett, 1971), E
tr
is raised if
the crop is planted more densely and more uniformly distributed over the soil to provide better canopy
cover, and the soil is not irrigated frequently to minimizing wetting of the soil surface. The water
transpired by the crop is in exchange for the cabon dioxide assimilated photosynthetically by the crop.
E
as
is generally higher for C
4
species than C
3
species, and higher if mineral nutrients, especially
nitrogen (Steduto et al. 2005), are not deficient. E
as
is also affected by evaporative demand of the
atmosphere, being higher under cooler temperature and higher humidly (Hsiao, 1993b; Xu and Hsiao,
2004). If switching from a C
3
to a C
4
crop is not an option, It may be possible to change the planting
time so growth of the crop takes place under the lower evaporative demand of the cooler part of the
season. Better fertilization would improve E
as
as well and the extra cost of the fertilizer may pay for
itself by increasing yield in addition to enhancing E
as
. Next step is biomass efficiency, E
bm
. Because it
is largely a function of the chemical composition of the crop, it is not easily changed except for the
possibility of reducing respiratory loss of assimilates by growing the crop under a cooler temperature
regime. The last step is E
yld
, the ratio of harvested yield to the total crop biomass. E
yld
has been
improved considerably during the last century as the result of breeding for crops with higher yield. The
higher yields turned out to be largely the result of partitioning more biomass to fruit or grain and less
to vegetative parts (Evans, 1993). For a number of crops, the partitioning can be modulated by water
status of the plant, and hence by irrigation scheduling. Unusually high water status induces more
vegetative growth in many species and can reduce E
yld
. Mild water deficit after the crop canopy is fully
grown may improve E
yld
, but very severe water deficit at pollination time would reduce it markedly.
Moderate water stress also reduces E
yld
during grain filling because of accelerated leaf senescence,
especially if the crop is relatively low in nitrogen. These effects are more thoroughly discussed
elsewhere (Hsiao, 1993a; Hsiao et al., 2005). It suffices to say that strategically better timed irrigation
provides a means to improve yield efficiency (equivalent to harvest index) at a minimum or no
additional cost.
Just how much increase in E
all
of the poor situation in Table 1 can be expected if some of the
nominal or low cost improvements in the individual efficiency steps discussed above are carried out?
Eq. 4 shows that if the improvement is only in one step, say a 55% increase in the efficiency of that
step ( = 0.55), then the improvement in E
all
is also 55%. This holds regardless which step is being
improved. On the other hand, if improvements are made in a number of the steps and none of them
are major, there would be marked improvement in E
all
. To illustrate, the improved E
all
is evaluated by
applying Eq. 4 assuming the following: the original E
all
is that for the poor situation in Table 1; E
farm
is
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

56
increased 40% ( = 0.40) by lining the ditches but not the storage pond with plastic sheeting; E
appl
is
increased 37% by taking more care to regulate water application for the furrow irrigated field; E
tr
is
increased 25% by reducing irrigation frequency somewhat while increasing the water applied per
irrigation to ensure good water supply to the crop; and the other efficiency steps in the chain remain
unchanged. The (1 + ) values for the improvements in the order given are: 1.4, 1.37, and 1.25, and
their product is 2.4. That is, the new overall efficiency is now 2.4 times the original overall efficiency,
and calculates out to be 0.058 compared to the original 0.0242 kg of yield per m
3
of water. If some
additional but still not costly improvements are made in the steps, E
all
could be raised still much
higher. For example, if the storage pond is spread with clay to reduce the porosity of the soil bottom
and E
farm
is increased by 78% as the result instead of only 40%, E
as
is increased 19% by improved
nitrogen fertilization, and E
yld
is increased 24% by better control of irrigation to restrict leaf growth after
canopy closure. The overall efficiency would be increased 4.5 fold in this case, to 0.109 kg of yield
per m
3
of water. Note that the overall improvement is marked although still much lower than that for
the good situation in Table 1. The point is that a systematic and integrated approach must be taken to
produce more crop per unit of water, by examining all the individual steps for potential improvements
at nominal cost, and not just focus the attention on one or two of the step. That way limited resources
can go a long way in improving water use efficiency.


APPLICATION TO OTHER PRODUCTION SYSTEMS AND ON LARGER SCALES

Because the principle and equations are general, the chain of efficiency steps is application to any
production systems as long as the steps in the production process are largely sequential. Water is of
paramount concern in rainfed cropping systems in less humid areas. To apply this approach, the
engineering aspects, from conveyance from the reservoir to placing water in the root zone, are
replaced by a couple of efficiency steps involving infiltration of rain water into the soil and retention of
the water in the root zone. From that point onward the steps are the same as those starting on line 4
(E
et
) of Table 1. The concept is also valid for animal production. By adding animal production steps
following the biomass step (for forage fed animal) or yield step (for grain fed animal), the final
outcome is animal product instead of crops. These interesting applications are discussed elsewhere
(in a presentation in the Rainfed and Drought session of this conference, and in Hsiao et al, 2005).
The treatment here is confined implicitly to the local scale. In fact, the unit considered is a single
field. For practical use, it is necessary to account for more complex situations such as a farm with a
number of fields of different crops, or an irrigation district comprised of many farms and several
distribution canals. These situations certainly make the calculations more complicated, but the
principle and basic equations still apply. A way to integrate the basic equations for application at large
scale has been worked out and is discussed in Hsiao et al. (2005). Another complication is the need
to account for the use of recycled runoff and drainage water, also discussed in Hsiao et al.


USE IN ECONOMICAL ANALYSIS

The ability to quantify the contribution of improvement in any efficiency step to the improvement in
overall efficiency makes this approach extremely useful. Different steps have difference efficiencies
and the cost of their improvement also differ. Often the cost of raising a step efficiency to a top level is
very high, but raising it to a modest level is low or moderate. Eq. 4 indicates that generally it is better
to allocate resources to improve the steps with the lowest efficiencies, because the overall
improvement is proportional to the fractional improvement of a step. So a given percentage
improvement (e.g., 20%) in a low efficiency step (e.g., from 0.4 to 0.48) has exactly the same effect
on the overall efficiency as the same percentage improvement in a relatively high efficiency step (e.g.,
from 0.8 to 0.98). When many step efficiencies are less than the good situation, how to allocate the
limited resources for improvement among the steps is not simple and requires optimization. The
approach here provides the quantitative fundaments for that process.


Acknowledgement

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

57
The intellectual development of the approach described here owns much to discussions with and
suggestions from Dr. Pasquale Steduto. I also thank Prof. Elias Fereres for valuable input, criticism
and encouragement.


REFERENCES

Evans, L. T. (1993). Crop evolution, adaptation, and yield. Cambridge University Press, Cambridge.
Hsiao, T. C. (1993a). Growth and productivity of crops in relation to water status. Acta Hortuculturae
335:137-148.
Hsiao, T. C. (1993b). Effects of drought and elevated CO
2
on plant water use efficiency and
productivity. In: M. B. Jackson and C. R. Black (eds.) Interacting Stresses on Plants in a Changing
Climate, NATO ASI series. Vol. I 16. Springer-Verlag, Berlin. pp. 435-465.
Hsiao, T. C., Steduto, P, and Fereres, E. (2005) A Systematic approach to the improvement of water
use efficiency. Irrigation Sci., in preparation.
Ritchie, J. T. and Burnett, E. (1971). Dryland evaporative flux in a subhumid climate: II. Plant
Influences. Agronomy J. 63:56-62.
Steduto, P, Hsiao, T. C., and Fereres, E. (2005) On the conservative behavior of biomass water
productivity. Irrigation Sci., in preparation.
Xu, L.-K. and Hsiao, T. C. (2004). Predicted vs. measured photosynthetic water use efficiency of
crops stands under dynamically changing field environments. J. Expt. Bot. 55:2395-2411.






























Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

58













































Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

59
ON THE CONSERVATIVE BEHAVIOR OF
BIOMASS WATER PRODUCTIVITY
1




P. Steduto
*
, T.C. Hsiao
**
, E. Fereres
***

* Division of Land and Water, FAO, United Nations, via delle terme di Caracalla, Rome, Italy
(pasquale.steduto@fao.org)
** Department of Land, Air and Water Resources, University of California, Davis, CA, USA
(tchsiao@ucdavis.edu)
*** Instituto de Agricultura Sostenible, University of Cordoba, Spain (ag1fecae@uco.es)


INTRODUCTION

Food production and water use are two closely linked processes. As the competition for water
intensifies worldwide, water in food production must be used more efficiently. Of the different steps in
water use in the crop production process, the most fundamental is the exchange of water lost by
transpiration for the assimilation of carbon dioxide. The net gain of carbon and energy by the plant in
this process then leads to the production of biomass. It turned out that for biomass production, the
efficiency of water use is relatively constant after the variation in two key environmental factors,
evaporative demand of the atmosphere and air carbon dioxide concentration, are accounted for by
normalization.

This conservative behaviour has been analyzed and discussed in detail several times in the past
half century (e.g., de Wit, 1958; Tanner and Sinclair, 1983). In light of the urgent need to answer the
question of how much the efficiency of water use in agriculture can be improved, and to further
analyse the implications for agricultural systems sustainability (e.g., Fereres et al., 1993), we are
revisiting the issue here to see how conservative biomass water productivity is and the extend of
possible improvements.

The conceptual basis for the conservative behaviour is reviewed and the ways to normalize for
evaporative demand and carbon dioxide concentration illustrated. It is hoped that this discourse will
help to focus better the potential means to improve the efficiency of water use, and also lead to a
simple means of modelling crop productivity based on water use.


THEORETICAL FRAMEWORK AND EXPERIMENTAL EVIDENCE

The focus of this note is biomass water productivity (WP
b
), also referred to as biomass water use
efficiency (WUE
b
) in the literature. From an agronomic standpoint, it is the amount of crop biomass
output per unit of water consumed in transpiration by the crop and evaporation from the soil (together,
evapotranspiration). From a physiological standpoint, only the water transpired is considered because
evaporation from the soil is not in exchange for carbon assimilated. Here, WP
b
is defined as the
aboveground dry matter (kg m
-2
) produced per unit of water transpired (m
3
m
-2
, or mm). Therefore, the
units of WPb are g m
-2
mm
-1
or g biomass per m
3
of water transpired (g m
-3
). Only above-ground
biomass is considered in our discussion as for most crop species, except root crops, only a small
portion of the total biomass is in roots and because there is a general homeostatic growth response
towards a near constant root:shoot ratio.

In developing the theoretical background and the appropriate framework for analyzing the
constancy of WP
b
, we follow a stepwise scaling-up approach, from leaf to whole crop field, in the
analysis of the two basic processes involved, water transpiration (T) and net carbon assimilation (A),
and its conversion to biomass.

At the leaf level, we define photosynthetic water productivity (WP
p
) as the ratio of leaf net carbon
dioxide assimilation (A
l
) to leaf transpiration (T
l
), both expressed as flux rates on a leaf area basis

1
The present note is abstracted from a paper under revision that will be published in Irrigation
Science early in 2006.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

60
(mol m
-2
s
-1
for A
l
, and mol m
-2
s
-1
for T
l
) and directly proportional to the gas gradient (CO
2
for A
l
and
vapour for T
l
) and inversely proportional to the resistance encountered along the path (e.g., boundary
layer, stomatal, metabolic). Plants have apparently evolved physiological mechanisms to keep the
importation (from ambient air to leaf interior via stomata) and depletion (from leaf interior to the
cellular carboxilating sites) of CO
2
in balance most of the time so that the leaf-internal CO
2

concentration (c
i
) is conservative. This implies that photosynthetic capacity and stomatal opening are
coordinated and operate in concert in the leaf. This suggests that when one of the two opposing
processes, either the importation or depletion, is perturbed, the other adjusts with some lag to keep
the system in balance and c
i
nearly constant. There has been substantial experimental evidence
showing that for many species, c
i
tends to remain constant under a range of conditions including
temperature, radiation, water and salinity stresses, especially when the stress develops gradually, as
it generally occurs in the field. The ample evidence of the tendency of c
i
to remain constant at a
constant ambient CO
2
concentration (c
a
), i.e. a constant c
i
/c
a
ratio, is an indication that stomata
perform at the leaf scale in a manner that leads to a constant WPp.
At the canopy level, we define, canopy photosynthetic water productivity (
c
p
WP
) as the ratio of
canopy net carbon dioxide assimilation (A
c
) to canopy transpiration (T
c
). As we scale up from leaf to
canopy, there are additional features that must be taken into account because the consideration is
now on a land area basis instead of leaf area basis. The extent of radiation capture by a crop
depends on the amount of leaf area, on the geometric arrangement of the leaves within the canopy,
as well as on the angle and intensity of incident radiation. As is the case at leaf level, the process of
T
c
shares the same source of captured energy as A
c
. Of the total captured solar radiation, though,
only the fraction that is photosynthetically active (PAR) is effective in CO
2
assimilation, while the
whole spectrum is used for transpiration. PAR, however, is a fairly constant fraction of the incident
solar radiation as is the ratio of absorptance of PAR to non-PAR radiation for the leaves of many
species. Consequently, any change in the amount of radiation captured by the canopy would affect in
a similar way A
c
and T
c
so that also
c
p
WP
tends to remain constant.

At crop field level, the variable we want to focus on is the biomass water productivity (WP
b
).
Changing from A
c
to biomass requires an analysis of the respiratory costs in relation to A
c
and of the
chemical composition and carbon cost of the biomass. A constant WP
b
, then, would be expected only
if the relationship between assimilation and respiration is also linear. This seems the case, provided
that the composition of biomass does not change significantly. More and more evidence is appearing
indicating an approximate fixed ratio of assimilation to respiration for crops (e.g., Albrizio and Steduto,
2003) where the reproductive organ has no high protein and/or oil content, such as soybean and
sunflower. Constant WP
b
seems to be the case even under varying environmental conditions.
Although WP
b
addresses situations where only aboveground biomass is considered, constant WP
b

has also been described for root and tuber crops.


NORMALIZATION OF BIOMASS WATER PRODUCTIVITY FOR CLIMATE

To extrapolate water productivity values between climatic zones and between atmospheric CO
2

statuses, there is a need to normalize them for the climate, specifically, for the evaporative demand of
the atmosphere and for the atmospheric CO
2
concentration, respectively. Ways of normalizing WP
b

for the evaporative demand of the atmosphere (Steduto and Albrizio, 2005) and for the atmospheric
CO
2
concentrations (Hsiao, 1993) are expressed by Eq. (1) and Eq. (2), respectively, where
*
p
WP
is
the normalized value of WP
b
. In Eq. (1), the summation is over a total number of time intervals (n),
with i being the running number designating each interval and ti the length of the interval (in days);
Biomass denotes the gain in biomass from the beginning to the end of the summation period. In Eq.
(2), the subscript o indicates the reference situation; the summation is over a number of days (n);
w is the water vapour concentration difference between the leaf intercellular air space and the
atmosphere.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

61

E
T
t
Biomass
WP
i
n
1 i 0
c
i
*
b

=
|
|
.
|

\
|
= (1)

=
=
=
=

=
n
1 i
i
n
1 i
i o
n
1 i
i o , a
n
1 i
i a
o , b
*
b
) w (
) w (
) c (
) c (
WP WP (2)


CONCLUSION

The implications that the near constancy of WP
b
has in the improvement of water productivity in
agriculture cannot be overemphasized. The presented stepwise approach from leaf to the whole crop
has provided a conceptual and theoretical framework to explain the basis for the constancy of
biomass water productivity. An important implication of normalizing biomass water productivity is that
it allows the comparison of water productivity data across the globe on equal footing, after accounting
for differences due to variations in evaporative demand of the climate, and in atmospheric carbon
dioxide concentration when applicable. Such comparisons will reveal more definitively the intrinsic
properties of the crop or the management practices that alter such productivity. Most importantly,
normalized WP
b
will provide a head start in knowing the WP
b
values at a new location or new time
period when CO
2
concentration is different, whether in the future or in the past. This offers an
invaluable tool for modelling purposes, providing an effective way of extrapolating WP
b
values
between different locations and seasons. Crop modelling based on radiation use efficiency (RUE), in
fact, has a limited normalizing capability (Steduto and Albrizio, 2005).


REFERENCES

Albrizio R. and Steduto P., 2003. Photosynthesis, respiration and conservative carbon use efficiency
of four field grown crops. Agric. For. Meteorol., 116: 19-36.
de Wit C.T., 1958. Transpiration and crop yields. Versl. Landbouwk. Onderz. 64.6 Inst. Of Biol. and
Chem. Res. on Field Crops and Herbage, Wageningen, The Netherlands.
Fereres E., Orgaz F. and Villalobos F.J., 1993. Water use efficiency in sustainable agricultural
systems. International Crop Science, CSSA, I: 83-89.
Hsiao T.C. (1993) Effects of drought and elevated CO2 on plant water use efficiency and productivity.
In Jackson, M.D. and Black, C.R. (eds) Global Environmental Change. Interacting Stresses on
Plants in a Changing Climate. NATO ASI Series. Springer-Verlag, New York, pp. 435-465.
Steduto P. and Albrizio R., 2005. Resource-use efficiency of field-grown sunflower, sorghum, wheat
and chickpea. II Water Use Efficiency and comparison with Radiation Use Efficiency. Agric. For.
Meteorol., 130: 269-281.
Tanner C.B. and Sinclair T.R., 1983. Efficient water use in crop production: research or re-search? In:
Taylor, H.M., Jordan, W.A. and Sinclair, T.R. (eds). Limitations to Efficient Water Use in Crop
Production. ASA, Madison, Wisconsin, pp. 1-27.
















Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

62



























































Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

63
TECHNICAL INTERVENTIONS TO IMPROVE
WATER USE EFFICIENCY
IN IRRIGATED AGRICULTURE



A. Hamdy
Professor Emeritus, CIHEAM - Mediterranean Agronomic Institute of Bari, Italy


SUMMARY - The increasing scarcity of water in the dry areas of the Mediterranean is now a well
recognized problem. High rate of population growth and economical development require continuous
diversion of agricultural water to higher priority sectors. The need to produce more food with less
water poses enormous challenges to transfer existing supplies, encourage more efficient water use
and promote natural resources conservation. On-farm water use efficient-techniques if coupled with
improved irrigation management options, better crop selection, appropriate culture practices and
timely socio-economic interventions would help achieving this objective. In arid and semiarid
countries of the region, water is a more limiting factor to production than land; hence, maximizing
water productivity should have higher priority over maximizing yield in the strategies of water
management. This implies that planning water and land use should be based on the comparative
advantages of the dry areas, but, within the framework of maximizing the return from the limited
available water resources. Achieving the greatest water productivity needed to resolve water shortage
problems will not happen automatically. There is great need to find appropriate ways and proper tools
for water saving and to achieve greater efficiency and equity in the irrigation system. Equally,
procedures and practices for the assessment of the performance of irrigation must be improved with
better management systems for water conveyance, allocation and distribution. Such issues, beside
others, will be addressed in this paper, with major emphasis on the technical interventions that could
be adopted in order to increase water productivity through water saving and the improvement of the
rates in water use efficiency in irrigated agriculture.

Key words: water use efficiency, irrigation.


INTRODUCTION

The availability of freshwater is, today, one of the great issues facing the human kind, and in some
ways the greatest, because problems associated with its availability affect the lives of millions of
peoples, particularly those in the developing countries. In the Southern and Eastern countries of the
Mediterranean, the agriculture sector is, by far, the largest water user. On a consumptive-use basis,
80 to 90% of all the available water resources are consumed by the agriculture sector. However,
average losses in the irrigation projects suggest that only about 45% of the delivered or extracted
water for irrigation actually reaches the crops. Very often, the conveyance losses of conduits (unlined
canals or leak pipes) are much too large, a 30% losses of the available water is common in the
irrigation systems.

Another cause of inefficient water use is the emphasis on meeting the demand by constructing
new supply facilities rather than improving the inefficiencies of the existing ones. Furthermore, for
most countries of the Mediterranean, on-farm irrigation practices deliver significantly more water per
unit area that what is required, leading to such low irrigation efficiencies.

Up to now, water conservation and efficient use of water have not been given the attention they
deserve. As agriculture is, by far, the largest water user, efficient irrigation management will
undoubtedly be a major conservation option in the future. An increase in water use efficiency in the
agriculture sector will serve a double purpose: it will free sufficient amounts of water for alternative
industrial uses, while at the same time, it is quite possible to take away a certain amount without
scarifying food production. In deed, it is possible to produce more out of less, but, this requires
finding out the appropriate ways to achieve greater efficiency and equity in the irrigation systems.

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

64
Several issues were recommended as proper tools for the improvement of irrigation efficiencies;
among those issues: lining canals, improvement of irrigation structure, modernization of irrigation
systems, etc. However, as irrigation efficiencies are basically ratios of volumes in the water
balance of an irrigation system, concentrating mainly on the relationships between the components of
the water balance of the irrigation system and the factors that might influence these relationships, it is
not always realistic to expect increased efficiencies.

There are other factors related to irrigation efficiencies such as management of the crops, socio-
economic and legal environment of the irrigation systems, capacity building in the irrigation sector and
quality of the water. These factors, together with the ones related to the components of water
balance, should be fully considered in all the programs and projects dealing with water use and
efficiency improvement in the irrigation sector.

Indeed, efficient use of irrigation water more crop per drop and improving the productivity of
water in agriculture through an appropriate water management is not an easy process. Significant
challenges still remain in the areas of technological, managerial and policy innovation and adaptation,
human resources development, information transfer and social environment considerations. Our
success and/or failure is a matter of our capability in finding sustainable solutions to the challenges
we are facing.


WATER USE EFFICIENCY IN IRRIGATION

Efficiency in the use of water for irrigation consists of various components and takes into account
losses during storage, conveyance and application to irrigation plots. Identifying the various
components and knowing what improvements can be made is essential to making the most effective
use of this vital but scarce source in Mediterranean agricultural areas.

Generally speaking, efficient water use is defined as the ratio between the actual volume of water
used for a specific purpose and the volume extracted or derived from a supply source for that same
purpose.

Functionally expressed, we have:
Ve
Vu
Ef =


where: Ef - Efficiency, dimensionless
Vu - Volume utilized, m
3

Ve - Volume extracted from the supply source, m
3
.

In this case, water use efficiency refers exclusively to irrigation. In accordance with the definition
proposed by the International Irrigation and Drainage Committee (quoted by Burmau et al., 1981)
efficiency in the use of water for irrigation has several different components (Table 1).

It is not redundant to insist on the fact that one can properly judge the hydraulic performance of
irrigation only when both its efficiency and its effectiveness are assessed.

The product of the three types of efficiency (Table 1) is the total efficiency of water use of irrigation
Ei. Functionally expressed, we have:
Eu Ec Es Ei =

The term of irrigation efficiency, which is different from efficiency of water use in irrigation, is also
used. According to Israelsen (1963), such efficiency is expressed as:
Up
Vt
Ea =

where Vr is the volume of usable water stored in the exploration zone of the plants' root, Vp is the
volume received by the plot and Ea is the irrigation efficiency. Also if we consider the volume of runoff
loss and E and the seepage loss as D then this efficiency can also be expressed as follows:
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

65
Up
D E Up
Ea

=

Irrigation efficiency has several different components (Table 2). Losses in the distribution system
are due to leakage and evaporation; losses during application to the field are due to wind, evaporation
and runoff, and losses from the soil are due to excess water, applied beyond what the crop uses.

Table 1 - Components of efficiency of water use in irrigation
1. Storage efficiency Es.

Ve
Vd
Es = Ratio between
The volume diverted for irrigation (Vd) and the volume entering a storage reservoir (Ve) for the
same purpose.
2. Conveyance efficiency Ec

Vd
Vp
Ec = Ratio between
The volume of water delivered to irrigation plots (Vp) and the volume diverted from the supply
source (Vd).
3. Irrigation efficiency Eu.

( )
Vp
* Vu
Eu = Ratio between
The volume used by plants throughout evapotranspiration process, Vu, and the volume that
reaches the irrigation plot, Vp.
* (Vu) is equal to the volume of evapotranspiration by plants minus the volume of effective rainfall.

Table 2 - Irrigation efficiency factors
1 - Conveyance efficiency - ratio of water delivered to water diverted from source.
2 - Application efficiency - ratio of water reaching the soil to water diverted.
3 - Water use efficiency - ratio water available for the crop to water applied to soil.


IRRIGATION EFFECTIVENESS

As defined above, irrigation efficiency is an indicator of how much water is being lost in the
process of irrigation and nothing more. It is a usual and misleading habit to take irrigation efficiency as
an indicator of the overall hydraulic performance of a system. There is a tendency to think that as long
as efficiency is high, the system is performing well overall. However, one can easily get very high
efficiencies if he under-irrigates over the entire field, in which case there is in fact a bad performance.

Effectiveness is another criterion of performance in that it indicates the degree of achievement of
the desired irrigation objective, consisting of replenishing the root zone after depletion by root
extraction.

In general terms, effectiveness is usually defined as the ratio of output achieved to the target or
desired output. It is thus a measure of the degree of achievement of a fixed desired objective. In the
case of irrigation, effectiveness can thus be defined as:

( ) 100
irrigation before deficit moisture Soil
zone root the in stored water of Amount
% Ef =


E
f
is a parameter which may vary from 0 to 100% and is actually an indicator of the degree of
under-irrigation since it falls below 100% in such a case. Walker and Skogerboe (1987) referred to
this same criterion as water requirement efficiency and in other instances it is termed storage
efficiency.

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

66
Another criterion which has also been used for judging the extent of under-irrigation and which is
area-based rather than volume-based is the percentage of area adequately irrigated (Pa) which can
be defined as:

( ) 100
area irrigated Total
d replenishe zone root with Area
% Pa =

The drawback of this criterion is that while it indicates the spatial extent of under-irrigation, it does
not give any indication on the severeness of under-irrigation. One can thus theoretically have a root
zone that has been replenished to a level of 95% over the entire field, in which case Ef will be 95%
while Pa will be 0% since no point on the field would have had the root zone entirely replenished.
Effectiveness is thus a better indicator of the fulfillment of the irrigation objective than is the % of area
adequately irrigated.


GENERAL DISCUSSION AND CONCLUSIONS

The way to water saving and whenever possible to its re-use, is still open. From the purely
technical point of view, important water savings are possible, if one thinks that under realistic
conditions water efficiency can vary depending on the cases, the modes as well as equipment, for
instance from 75% to 25% about, one understands that moving from the former to the latter value
means to triple the irrigable surface at equal water use efficiency, with technologies and methods
available today, agriculture could cut its water demand by 10 to 15 percent.

If one considers not only crop requirements and the pedological environment but also the fact that
quite often the water saving techniques are labor-capital-energy-multiple-factors consuming
techniques, on one hand, and the new concerns about environment and some social problems
particularly related to the frequently low school levels of farmers, on the other hand, one understands
indicate that the solution through water saving is not simple but a complex one.

To achieve a sound use of water for irrigation with higher efficiency and better water saving
requires, on one hand:
deeper scientific and technical knowledge which is still far from being perfect although using
models which are hoped to be clearer and more and more widespread (Ait Kadi,1992);
a more systematic and permanent monitoring of supplying unbased data at reasonable cost, on
the other hand;
a closer participation and collaboration of the whole technical environment and the involvement of
the farmers in the implementation of the program for some tariffing criteria, or in the improvement
of the modes of use of water (Abu Zeid, 1990-1992). A great equilibrium is necessary in
evaluating the needs of the different user's groups and more flexibility and reconversion have to
be ensured. As for the technical aspects again, it should be noticed that, however, irrigation is a
tool used to maximize the general objectives of the whole agricultural management.

These are complex objectives, with a strong although not unique economic component, quite
different depending on the type of enterprise and scenarios, changing over time and space. It is then
very superficial and misleading to think of optimizing the use of water resource by maximizing only
irrigation water efficiency (in terms of money) as such, this could be a utopian ideal of a given
hypothetical owner of irrigation water only.

In practice some remarks are made: the first one is that the scenario where food demand is
greater (countries with a high population rate) and the natural rainfall regime is more limiting (arid and
semi-arid areas), the objectives of irrigation approach more the one of maximum water use efficiency;
it is less so and more related to income, in the case of more developed countries.

The second is that the irrigation efficiency is the result of all interactions with other production
factors and this may need adequate investments also in other factors (for instance, fertilizers, new
varieties, modern machinery, etc...).

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

67
The third remark concerns the need to have a team of well specialized technicians available for
irrigation, and then also the need to ensure training and update knowledge of these cadres.

It is a great effort and not only a technical scientific one. Cultural and social progress of the
agricultural world should be promoted as much as possible.

Operation and maintenance (O&M) is one of the most underestimated aspects of irrigation
projects in developing countries. As a result, the efficiency of the project continues to decline, and
during a crisis situation, generally the problem faced is more complex to resolve technically and more
funds have to be extended than had the maintenance works been carried out on regular basis.
Another issue worth noting is the fact that poor though O&M is for irrigation, it is generally even worse
in drainage.

The impending crisis of land and water scarcity by the turn of the century particularly in the arid
and the semi-arid region of the Mediterranean countries implies that research should be oriented to
cope with such alarming situation and to increase the productivity of available land and water
resources in a sustainable manner. In this regard, the following teams of research are suggested:

1) - Modern Irrigation methods made appropriate to the least developing technological context:
irrigation techniques have to be adapted and made appropriate to self-sustained technological
improvements, taking into account economic, technical, maintenance and cultural aspects. Adaptive
research is needed to modify the hardware of modern methods to permit local construction and
engineering use of local and low cost materials and skills. The process will also help in enabling users
to master the technologies and facilitate easier maintenance. Similarly the software modern methods
would also need adaptation to accommodate cases where data are inadequate or unreliable.

Another topic which should be carefully considered is the regulation of land networks; its
sequence regulation of water deliveries not only has the direct advantage of rising the crop yield but
also contributes to control waterlogging and salinity, beside other advantages such as the reduction of
seepage losses and sedimentation, control of disease vectors and the reduction of weed growth.

The introduction of pressurized irrigation systems as a tool for a better water saving and a better
economy in fertilizer use and less pollution of ground water, is among the points that should receive
the attention of researchers. Research is required to adapt such methods to specific soil, topographic
ecological and social conditions in order to ensure cost-effectiveness and sustainability. A part of the
research should be devoted to the rehabilitation and modernization of large systems, with the
objectives of greater decentralized operations, improved water conservation, and reliable and
equitable water delivery to self-managed groups. Systems now in operation will have to meet the
challenges of water scarcity and hence of conservation; moreover, rapid changes in all aspects of the
economy and in agriculture, call for systems that offer flexibility in crop production.

2 - Decision support at various levels: the potential decision support system (DSS) as planning,
policy-information and operational tools at a national, regional and farmer level is becoming evident.
At national and regional level, the combined use of models, remote sensing and geographical
information system can help predict the effects of different policy scenarios and allow better
management of existing land and water resources. At the farmer's level, DSS can help to improve
crop production and irrigation efficiency.

Existing models should be further developed, improved and tested in different situations; also new
models are required.

At the local level, research should focus on the operational use of models. Research is needed
with respect to parameter estimation for different crops and soils and model calibration and
verification in combination with field tests. Models should be adapted to the users by making them
user-friendly (generation of input data, understandable output, etc.). Furthermore, attention should be
given to the introduction of water quality aspects especially salinization into the soil-vegetation-
atmosphere (SVAT) model.

3 - Sustainable use of marginal quality water: there is increasing awareness of the need to augment
water supply and improve water quality through the use of the treated sewage and drainage water.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

68

For a sustainable use of drainage and brackish water for irrigation, research work is needed on the
physical parameters governing the flow of water and salt in the irrigated soil profile with an estimation
of the risk of salinization through modeling and testing improved methodologies to reduce salinity
through drainage. Research should be also directed towards the establishment of new strategies
including different crop rotations based on proper soil, water and crop management that favor the
reuse of drainage and brackish water in irrigation without deterioration either in crop production or soil
productivity. Adaptive research programs should be planned on the recycling of used water which
could lead to a great deal of progress, but which is still limited by the lack of research in this field. The
approach appears to be promising and will enable the impact of scarcity to be minimized in times to
come. A part of the program should include the waste water technologies and methodologies to
achieve the most appropriate ones to be used in irrigation to produce an effluent which meets the
recommended microbiological and chemical quality guidelines, both at low cost and with minimum
operational and maintenance requirements.

An important aspect of the use of marginal quality water for plant production is the possibility of
ground water contamination with undesiderable pollutants. Further basic and applied research is
needed to preserve and/or to improve the quality of ground water from serious contamination. Lack of
knowledge about long-term effects of treated waste water has to be searched; this will greatly
encourage and facilitate its use in irrigation on a wider scale.

4 - Performance assessment, monitoring and evaluation of irrigation and drainage systems: one main
reason for poorer economic performance of irrigation projects is deficiency in project planning and
management. The result is a restricted budget for operation and maintenance, crucial the integrity
and sustainability of an irrigation system. Research is needed to formulate a methodology for the
collection, storage and processing of administrative, technical and environmental data required for an
effective and efficient management of soil and water resources. The conveyance of information from
evaluators to planners and designers creates a positive feedback mechanism leading to a better
design which is another important area of research.

5 - Water optimization: crop water requirements and irrigation scheduling; efficient irrigation: in order
to optimize water application by irrigation to different crops under different soil types, irrigation
systems and climatic conditions, research should be continued to provide better knowledge of soil-
water-plant relationships, reviewing the concept of an optimal water supply. The management of
irrigated crops to cope with droughts should receive priorities in research. In addition the link between
crop water requirements and irrigation schemes reliability should also be considered.

6 - Environmentally sound practices for sustainable land and water use: new approaches and
methods are needed to solve some of the old problems in irrigation. These approaches need to be
cost-effective and bearable by farmers in order to be ecologically and economically sustainable. Many
fields of knowledge including engineering, biological science, and social science need to be drawn on
the development of these approaches and methods.

Research is needed in areas like: control of soil erosion and soil fertility conservation, water
conservation practices, economic criteria for irrigation and drainage of heavy saline and/or other
problem-affected soils; also alternative crops and cropping systems better suited to problem-affected
soils and other ecological constraints; integrated pest management and pest control; and
development of irrigation management practices to cope with drought and aridity, including water
conservation practices and the reuse of drainage and waste water for irrigation.

7) Socio-economic studies: management and design interaction in relation to community dynamic and
local managerial capability: The design of schemes has traditionally been the domain of irrigation
engineers whose concern has been to optimize technical efficiency of water use. The projected
returns to such schemes are too often unrealized because they rely on an unrealistic precision in
water allocation and water use across the design of irrigation area. Efforts need to be made to design
schemes whose layout is simpler to be operated and managed by farmers and their communities.
The challenge is for engineers to work with local farmers at whatever level of technology is
appropriate, and design more manageable schemes to economize, where possible, to compare real
social and economic impact in relation to the money spent as between simpler and more complex
designs under similar conditions.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

69





REFERENCES

Abu Zeid, M. (1993). Irrigation Cost Recovery In Developing Countries. Cahiers Options
Mditerranennes. Vol. I, n 1 Water Resources: Development and Management in Mediterranean
Countries. pp. 13.1-13.8.
Abu-Zeid, M. (1990). Some Technical and Economic Considerations on Irrigation Water Pricing.
Water Science Magazine, Issue n 7, Cairo, Egypt.
Ait Kadi, M. (1993). The Application of Optimization Techniques to Water Resources. Cahiers Options
Mditerranennes. Vol. I, n 1 Water Resources: Development and Management in Mediterranean
Countries. pp. 9.1 - 9.15.
Arreguin, F. (1991). Efficient Water Use in Cities and Industry. Proceedings of the International
Seminar on Efficient Water Use, Mexico, Oct. 1991, pp. 749-756.
ASCE (1974). Water Management through Irrigation and Drainage: Progress, Problems and
Opportunities. Committee on Research of the Irrigation and Drainage Division, ASCE, IR2, 1974.
Bau, J. (1991). Research on Water Conservation in Portugal. Proceedings of the International
Seminar on Efficient Water Use. Mexico, Oct. 1991. pp. 736-743.
Burman, R.D.; Nixon, P.R.; Wright, J.L. and Pruitt, W.O. (1981). Water Requirements. In: Design and
Operation Farm Irrigation Systems. Eds. M.E. Jensen, ASAE, St. Joseph, Michigan, 1981.
Caswell, M.F. (1989). The Adoption of Low-Volume Irrigation Technologies as a Water Conservation
Tool. Water International, Vol. 14, n 1, International Water Resources Association, 1989.
CEMAGREF, CEP, RNED-HA. (1990). Irrigation-Guide Pratique. Montpellier
Falkenmark, M. and Widstrand, C. (1992). Population and Water Resources: a Detailed Balance.
Population Bulletin, Population Reference Bureau.
FAO (1992). Wastewater Treatment and Use in Agriculture. (ed) Pescod, M.B. Irrigation and Drainage
Paper. 47 Rome.
Gleick, P.H. (1993). Water in a Crisis: Guide to the Worlds Fresh Water Resources. Oxford University
Press.
Gloss, S. (1991). The Legal and Institutional Conundrum of Efficient Water Use in the Western United
States. Proceedings of the International Seminar on Efficient Water Use, Mexico, Oct. 1991, pp.
523-530.
Horst, L. (1983). Irrigation System - Alternative Design Concepts, Irrigation Management Network
Paper 7c April. Overseas Development Institute.
International Commission on Irrigation and Drainage (ICID). (1989). Planning the Management,
Operation and Maintenance of Irrigation and Drainage Systems, World Bank Technical Paper
Number.
Lyle, W.M. and Dixon, D.R. (1977). Basin Tillage for Rainfall Retention. Trans.Am.Soc. Agri.Engr. 20,
pp. 1013-1017 & 1021.
Lyle, W.M. and Bordovsky, J.P. (1983). LEPA Irrigation System Evaluation. Transactions of American
Society of Agricultural Engineers 26, pp. 776-781.
Merriam, J.L. (1977). Efficient Irrigation. California Polytechnic State University. San Luis Obispo,
California.
Nayan, S.; Pandey, A.D. and Debajyoti, C. (1991). Feasibility for Canal Automation to an Existing
Canal -a Case Study. Proceedings of the International Seminar of Efficiency Water Use, Mexico,
Oct. 21-25, 1991.
Norum, E.M. (1991). Irrigation System Selection: Secondary Considerations. In: Irrigation and
Agricultural Development, FAO/Pergammon Press, 1980.
Phene, C.J.; Howell, T.A. and Sikorski (1985). A Travelling Trickle Irrigation System. In: Advances in
Irrigation. Vol. 2 (Ed) Daniel Hillel Academic Press, London, pp. 2-49.
Postel, S. (1986). Effective Water Use for Food Production. Water International. n 11, 1986.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

70
Saffaf, A.Y. (1980). Selection of Appropriate Irrigation Methods for Semi-Arid Regions. In: Irrigation
Agricultural Development, FAO/Pergammon Press, 1980.
Von Bernuth, R.D. and Gilley, J.R. (1985). Evaluation of Centre Pivot Application Packages
Considering Droplet Induced Infiltration Reduction. Trans. Am. Soc. Agr. Engr. 28(6), pp. 1940-
1946.
Walker, W.; Ruchardson, S. and Sevebeck, K. (1991). A Comprehensive Approach to Water
Conservation. Proceedings of the International Seminar on Efficient Water Use, Mexico, Oct.
1991. pp. 763-769.
World Bank (1992). World Bank Development Report. Oxford Press. New York.

















































Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

71

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

72



























COUNTRY REPORTS
































Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

73



























































Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

74
TECHNIQUES FOR IMPROVING WATER USE EFFICIENCY
IN GREENHOUSE CULTIVATION IN CYPRUS



P. Polycarpou, D. Chimonidou, I. Papadopoulos
Agricultural Research Institute
1516 Nicosia, Cyprus



SUMMARY - In Mediterranean region, there is an increasing concern about the effective and efficient
utilization of water for agriculture and water conservation in general. The promotion of effective water
use and on farm water management, were identified as an important contribution to the management
strategy needed to address problems of water scarcity and practicing intensive agriculture on
environmentally sound grounds. Recently particular emphasis was laid on protected cultivation and
more specific on cultivation of vegetables and flowers on substrates and soilless cultures (closed
systems and open with minimum drainage). New unites with soilless cultivation (mainly perlite,
coconut and rockwool) have been established applying modern greenhouse technology and fully
computerized irrigation-fertigation methods. The only way to increase productivity, improve quality
and control growth and production, is through the application of modern greenhouse technology, new
techniques of cultivation and integrated protection and production management. At the Agricultural
Research Institute the use of local materials i.e. perlite, mixtures of perlite with pomace, almond
shells, pine bark, gravel, etc. have been tried successfully. In this paper, results of the application of
modern techniques, hydroponic cultures, re-circulation of irrigation water and nutrient solution in
closed systems and control of the climatic conditions in the greenhouse (temperature, air humidity,
CO2, etc) will be discussed. The introduction of modern technology and soilless culture in
greenhouse cultivations (vegetables and flowers) resulted in higher production, better quality, efficient
and effective use of water and fertilizers and minimize the use of chemicals for pest and disease
control. The use of closed recirculation systems has reduced the water needs of the cultivations close
to the evapotranspiration levels of the crop. Ongoing research on using a green lagoon to de-
nitrifigate the reject water from the closed system, when the undesired elements in it reach toxic
levels, seems to be very promising. The grown Sudan-grass in the lagoon can be used as animal
feed or as an energy plant.


Key words: soilless culture, hydroponic systems.


INTRODUCTION

In Mediterranean region, there is an increasing concern about the effective and efficient utilization
of water for agriculture and water conservation in general. The promotion of effective water use and
on farm water management, were identified as an important contribution to the management strategy
needed to address problems of water scarcity and practicing intensive agriculture on environmentally
sound grounds (Papadopoulos and Chimonidou, 1997, 2004).

In Cyprus, the irrigated land amounts to 35000 ha (16.2% of the total agricultural land) with
provision to be expanded. The irrigated agriculture in semi arid countries like Cyprus demands large
amounts of water and faces the serious challenge to increase or at least sustain agricultural
production while coping with less and/or lower quality water (Papadopoulos and Chimonidou, 1997,
2004).

Recently particular emphasis was laid on protected cultivation and more specific on cultivation of
vegetables and flowers on substrates and soilless cultures (closed systems and open with minimum
drainage). New unites with soilless cultivation (mainly perlite, coconut and rockwool) have been
established applying modern greenhouse technology and fully computerized irrigation-fertigation
methods (Chimonidou 1999, 2003a, 2003b).

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

75
The demand for increased production and better yield quality, the lack of good quality water for
irrigation and the need for protection and conservation of the environment require the implementation
of new technologies in greenhouse cultivation. In this respect, the first measures to be taken are
improvement of greenhouse structures and automation of the systems. Moreover, modern techniques
have to be applied such as cultivation on artificial substrates, hydroponic cultures, re-circulation of
irrigation water and nutrient solution in closed systems and control of the climatic conditions in the
greenhouse (temperature, air humidity, CO2, etc). Apart from higher production of good quality (10 %-
25 % yield increase in vegetables and up to 30 % in flowers), such technology would result in efficient
and effective use of water and fertilizers and in minimizing the use of chemicals for pest and disease
control, especially after the prohibition in using Methyl Bromide for soil sterilization. Moreover some
cultural practices, like soil cultivation and weed control are avoided, and land not suitable for soil
cultivation can be used (Polycarpou and Hadjiantonis 2004).


SOILLESS CULTURE

Recently particular attention was given in soilless cultivation and the area under soilless culture is
rapidly expanding. The first soilless culture in Cyprus started with rose cultivation on rockwool in 1996
at the area of Monagrouli (Limassol) with 0.2 ha and expanded in 0.4 ha in 1997 with a fully
automated computerized open system. Since then, the cultivation of roses on substrates has rapidly
expanded. The application of soilless culture in vegetables (tomatoes, cucumber, lettuce,
strawberries) came later due to the low and unpredictable market prices of the products.

The total area cultivated today in Cyprus using soilless techniques is about 11.6 ha. A break down
of this area is given in Table 1. This area is expected to increase rapidly in the near future due to the
grants schemes announced by the Ministry of Agriculture as a measure to improve the quality of the
agricultural products as a result of Cyprus joining the European Community.

Soilless culture has some fundamental differences compared to growth on soil. Unlike soil culture,
soil1754less cultivation starts with media free of soil born diseases and gives the grower flexibility to
control growth by manipulating water and nutrient supply. The substrate can be chosen to have the
desirable characteristics so that with the appropriate water and nutrient management yield can be
maximized and quality be improved.


Table 1. Soilless Cultivation in Cyprus (Source: Department of Agriculture)
Area of Soilless Culture in Cyprus
(ha)
Substrate
Flowers
(Roses & Cerbera)
Vegetables Total
Perlite 0.3 0.2
0.5
Coco Peat 3.8 1.6 5.4
Rock wool 0.7 5.0 5.7
TOTAL 11.6


However, there are some difficulties in soilless cultivation like:
High initial installation cost
High skill requirements on growers
Sensitive system with no buffering capacity of nutrients No error tolerance.
High water quality requirements. Risk for environmental pollution if not properly
managed.That is the reason for the slow application of the method until now.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

76


EXPERIMENTAL WORK ON THE CULTIVATION ON SUBSTRATES

Roses on soilless cultivation

A lot of attention was given on the research on roses and experiments started since 1990 on the
physiology of roses, the critical levels of development in relation to water stress and the effect of
irrigation, shading and salinity on yield and quality (Chimonidou, 1996,1997,1998).

Experiments on soilless cultivation of roses started in 1997 using different substrates i.e. perlite,
perlite (70%) + pomace (30%), Coconut, 100% rockwool and pine bark (70%) + straw (30%) as inert
media. The aim was to compare different inert media (imported and local) and evaluate the cultivation
in bags and containers using the technique of shoot bending. Irrigation based on the moisture content
of the media, is kept constant at -5 to -8 kPa at the area of the root zone. The total amount of
irrigation water was the same but the frequency was different according to the holding capacity of
each substrate. The productivity and quality characteristics (stem length, fresh weight, flower bud
diameter and height) have been recorded. Results were very encouraging for the local substrates
tested.

Experimental work on the cultivation of Roses on soilless culture expanded in fully automatic
greenhouses at different locations aiming at the use of new substrates to face the problem of low
quality waters in Mediterranean climates using the minimum drainage and accumulate the salts at the
periphery of the container.

A joint programme between the Agricultural University of Athens - Greece and the Agricultural
Research Institute of Nicosia - Cyprus (2001-2004), aimed at studying the development and
photosynthetic activity of roses cultivated in four different substrates and two irrigation regimes.
Roses cv "Eurored" were cultivated on four different substrates in a heated greenhouse at the ARI
using local materials i.e. perlite 100%, mixtures of perlite 50% with pomace 50%, perlite 50% with
almond shells 50% and almond shells 50% with pine bark 50%. The two irrigation regimes applied,
were 800ml (6 times/day X 2 min) and 530ml (4 times/day X 2 min).

The photosynthetic rate, stomatal conductance, CO
2
concentration and transpiration rate of the
rose plants grown in the four substrates and under the two irrigation regimes were measured as well
as the total productivity and quality characteristics (stem length, fresh weight) of the roses produced.
Results showed that concerning the interaction between substrate and irrigation level, higher
production was recorded with the roses growing in the substrate pine bark 50% and almond shells
50% irrigated with the reduced irrigation level and the substrate of perlite 50% and pomace 50%
irrespective of irrigation level. The quality characteristics of the roses produced under all treatments
were marketable with mean stem length between 75-85cm.

No significant differences were observed concerning the photosynthetic rate, the stomatal
conductance, the CO
2
concentration and the transpiration rate under the four substrates and the two
level of irrigation. It seems that the lower irrigation level did not create conditions of water stress and
did not affect negatively the physiological activities of the rose plants. Concluding remarks showed
that the local substrates could be used successfully as substrates for the rose cultivation in the
region.


Cultivation of Lysianthus (Eustoma) on substrates

Two different experiments on Lysianthus were conducted during the years 2002-2003 (at the
Agricultural Research Institute and at Zygi experimental station), aiming at higher productivity and
year round production.

The first experiment in cooperation with the Aristotle University of Thessaloniki was aimed at
evaluating the productivity and quality characteristics of Eustoma grandiflorum on two substrates and
two irrigation regimes. The substrates used were perlite 70% with coco 30% and perlite 50% with
pomace 50%. The irrigation was performed using drippers of 4l/h and the irrigation intervals were: 6
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

77
times x 2 min (800 ml/ day) and 4 times x 2 min (530ml/ day = reduction 33%). The pH and the EC
were kept constant at the levels of 6,5 and 1,7-1,8 DS/m respectively. Drainage for both cases was
only 5%. Results showed that no significant differences were existed between the different substrates
or the stressed and not stressed plants with respect to the total productivity (number of stems) or the
quality characteristics (number of flower buds, stem length and fresh weight) of Lysianthus. Vase life
of the plants was not affected by the cultivation in different substrates. On the contrary, the plants
under the low level of irrigation lasted more days in vase with or without preservative (Chimonidou et
al. 2003).


Greenery Cultivated in Different Substrates

Experiments on the cultivation of flowers and greenery on substrates started in 1995 with the
Greenery Rumohra adiantiformis cultivated on four different substrates in comparison to the
cultivation on soil in an unheated greenhouse at Zyghi experimental station. The aim was to increase
productivity improve quality (particularly stem length with increasing shading levels) and avoid soil
born diseases by improving soil structure and aeration of the root zone using local material (i.e. pine
bark, mushroom compost, pomace etc.).

Results show that pine bark 30% with peatmoss 70% appeared to be the most promising
substrate among those tested during 1995-97, concerning their quality characteristics (Chimonidou,
1999).

The experiment was extended in 1998 with the cultivation of greenery and other cut flowers on
soilless cultivation. The species Limonium ottolepis, Rumohra adiantiformis, Gerbera jamesonii, and
Pteris vitata, were tested in a mixture of perlite 70% and pomace 30% at the Agricultural Research
Institute in Nicosia. The total productivity and the quality characteristics (stem length and fresh
weight) as well as their keeping quality after harvest were studied. Results appeared to be very
encouraging for all species tested and specially for Limonium ottolepis, concerning the total
productivity and quality characteristics in comparison to previous research work carried out on soil
(Chimonidou 1999).


HYDROPONIC SYSTEMS

The open system for soilless culture, Fig.1, is at present most favored commercially in Cyprus due
to its simplicity, mainly in managing the nutrient solution.

Pollution of the environment (underground water), waste of fertilizers and water are though only
some of the problems faced in open hydroponic systems. The leachate is usually collected in a
reservoir and is used for the fertigation of open cultures or greenhouse cultivations in the soil. This
results in approximately 30 % loss of fertilizers and water from the system.

For this reason ARI started a research program in order to develop a locally adopted closed
hydroponic system (Fig. 2), using locally available inert substrates, like crashed gravel produced in a
copper mine in Cyprus. The leach ate from the substrates is collected in a tank and is recirculated
after being sterilized passing through a UV lamp. The EC and pH of the water are regulated using an
automatic fertilizer-mixing unit as by the open system. The water consumption of a good managed
closed system is reduced to the evapotranspiration level of the plants. The system requires water of
very good quality that is difficult to find in Cyprus. At the coastal areas where greenhouse cultivation
has developed due to the favorable climatic conditions, the ground water salinity ranges from 1.5 to 4
dS/m, whilst the salinity of water coming from dams is around 1 dS/m. The fresh water supplied to the
closed system can be therefore rainwater collected from the greenhouses or water treated by a small
reverse osmosis unit. Thus the need for replacing the nutrient solution due to the increasing
concentration of chlorides and sodium is minimized. When the nutrient solution comes to the point
that it has to be replaced, the possibility is studied to pass it through a green lagoon planted with
sudan-grass is studied. By this method the water is denitrificated and can be safely disposed to the
environment. The sudan- grass harvested from the lagoon could be used for animal feed or for the
production of bio-mass for energy purposes. The experiments are carried out at the ARI research
station at Zygi on tomato cultivation (Polycarpou and Hadjiantonis 2004).
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

78



Fig. 1. Schematic diagram of an open hydroponics system applied commercially in Cyprus.

In addition, an open system using a mixture of locally available organic materials with perlite or
peat moss as substrate is being studied in floriculture. In this zero loss system the nutrient solution
is supplied to the plants, planted in big boxes (substrate volume 15 liters/plant), in such a quantity that
leaching just starts. In this way the water and fertilizer loss from the system is minimal. The salts are
pushed by the irrigation water away from the root zone and are accumulated in the outer volume of
the substrate not affecting the growth of the plants.
















Fig. 2. Schematic diagram of a closed recirculating hydroponics system applied at the ARI in Cyprus.

In designing and operating such a closed hydroponic system the following main parameters are to
be considered:
AC
TANK A
B
Leachate Tank
Fertilizer Mixer
Fresh Water
Reject water
UV
ACID
Leachate Tank
Fertilizer Mixer
Fresh Water B
TANK A
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

79
Crop related matters such as the life span of the crop, the water and nutrient requirements
(recipe), and the cultural practices needed.Method for fertilizer mixing and supply of irrigation
water (Using simple volumetric fertilizer injectors or automatic fertilizer mixing units).
Use of locally available inert substrates like perlite, coarse sand, crashed gravel vs. imported
inert materials like rock wool.
Climate Control in Greenhouses, like monitoring the aerial climate requirements
(temperature, relative humidity, light, CO
2
, etc), the root zone requirements (root temperature
and O
2
supply in the root zone) and improving the PAR transmission of covering materials
and lowering their NIR transmission.
Due to the advantages of the closed hydroponic system compared to the open one, ARI is
investing a lot of effort in optimizing its parameters, simplifying its operation and training the growers
in its effective management and utilization (Polycarpou and Hadjiantonis 2004).


CONCLUSIONS

In Mediterranean countries where water is limited and of high cost, diversion to intensive
irrigated agriculture, protected cultivation and soilless culture are promising alternative and
innovative approaches.

Soiless culture using locally available substrate materials (perlite, mixture of perlite + peat,
pomace, pine bark etc) could be the solution. Experimental results so far in terms of yield,
quality and water use efficiency are very encouraging.

It is important that a suitable closed system is developed that is based on low cost local
materials, which are both effective and easily disposable after use in order to avoid
environmental pollution. The system should be easily adaptable to the growers according to
their potential skills. The technology used should be locally supported in order to avoid long-
term maintenance problems.


REFERENCES

Chimonidou-Pavlidou Dora. 1996. Effects of water stress at different stages of rose development.
Acta Horticulturae 424:45-51.
Chimonidou-Pavlidou Dora. 1997. Use of saline waters for irrigation in Cyprus: New developments
and Management practices. In Proceedings of the International Conference on Water
management, salinity and pollution control towards sustainable irrigation in the Mediterranean
Region. Italy, 22-26 Sept.,1997.p. 21-33.
Chimonidou-Pavlidou Dora. 1998. Yield and quality of rose Madelon as affected by four irrigation and
three shading regimes. Acta Horticulturae 458:95-103.
Chimonidou-Pavlidou Dora. 1999. Protected Cultivation and Soilless Culture in Cyprus. First
meeting of FAO thematic Working group for Soilless Culture, August 31- Sept. 6, 1999, Halkidiki,
Macedonia, Greece.
Chimonidou Dora 2002a. Flowers under protected cultivation with special emphasis on roses and
new cut flowers. FAO/AUB First National Conference on Integrated Production and Prodection
Management of Greenhouse Crops, p.105-114.
Chimonidou Dora, 2002b. Country report of Cyprus. In Proceedings of the Regional Experts Meeting
on Flowers for the Future. Izmir-Turkey, 8-10 Oct,2002, p.15-30.
Chimonidou Dora, M. Stavrinides, I. Papadopoulos and P. Polycarpou, 2003. Intensive cultivation of
floricultural greenhouse crops. Proceedings of the Conference on Greenhouse Crop Production in
the Mediterranean Region. Nicosia, Cyprus 10th November 2003.
Papadopoulos,I. and Chimonidou Dora, 1997. Nutrient and Agro- chemical management for pollution
control under intensive irrigated agriculture. In Proceedings of the International Conference on
"Water management, salinity and pollution control towards sustainable irrigation in the
Mediterranean Region". Italy, 22-26 Sept, 1997. p. 45-65.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

80
Papadopoulos I. and Chimonidou Dora, 2004. Participatory Water Saving Management and Water
Cultural Heritage: Cyprus Country Report. Option Mediterraneennes Series B, No: 48, p. 97-111.
Polycarpou P. and Hadjiantonis Ch.,2004. Cyprus Country Report. FAO Regional Training Workshop
on Soilless Culture Technologies. Izmir-Turkey, 3-5 March 2004, p.12-15.












































Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

81



























































Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

82
REVIEW AND ANALYSIS OF WATER USE EFFICIENCY
AND WATER PRODUCTIVITY IN EGYPT



M. Nasr Allam
*
and R. Abdel-Azim
**

* Prof., Head of the Irrigation and Hydraulics Dept., Faculty of Eng., Cairo University, Egypt
** PhD, Ministry of Water Resources and Irrigation, Egypt


SUMMARY - This paper presents analysis of water resources availability and water uses in Egypt.
More focus will be given to agricultural water use and productivity since agriculture sector represents
the major water-consuming sector in Egypt. Nile River is the major source of water where rainfall is
rare and groundwater is limited. The nature of the irrigation network and system of the Nile river is
rather unique. Water lost from one point is usually used in the downstream, and hence the global
water use efficiency is relatively high. Moreover, there is a multiple system for water use in Egypt.
Water is used for agriculture, municipal and industry, navigation, hydropower generation, and
fisheries. Emphasis in this paper is given to agriculture water use and productivity. Agriculture sector
consumes about 85% of the total available water. There are several applied policies for improving
water use efficiency. Irrigation Improvement project is currently implemented to enhance water
distribution and minimize water losses through physical and institutional measures. Participatory
approach is also being implemented in water management followed by modification of laws and
institutional reform. Future water policies include introducing integrated water management approach
to increase water use efficiency and maximize water productivity. The paper presents the productivity
of the main crops in Egypt and water crop productivity compared to some international figures. The
results show that crops have a relatively high productivity particularly rice crop which has the highest
productivity over the world.

Key words: water resources, cropping pattern, crop production, Egypt.


INTRODUCTION

Water, always, plays an essential role for providing the basis of population stability and civilization.
The Nile River in Egypt has supported the longest civilization over the world which lasted more than
seven thousands years. Egyptians, throughout the history, were skillful enough to utilize the Nile
water. During this century, they installed an invaluable water structure; High Aswan Dam (HAD). This
dam helped in providing more controllable water releases pattern over the year, serving the Egyptian
population who are living on the small batch along the river. Egypts annual share of Nile water that
controlled by the dam is 55.5 billion cubic meters as stipulated and agreed upon between Egypt and
Sudan in 1959. After the construction of the HAD in 1986, the cultivated area has been expanded to
reach 8.0 million feddans, which is cultivated about twice a year. The government of Egypt continues
to invest heavily in expanding the cultivated area, and is planning to add another 3.4 million feddans
of the cultivated land by year 2017.

Water Supply augmentation has been practiced in Egypt since several decades through recycling
the drainage water and use of shallow groundwater. Drainage reuse started in 1970s and reached
now a level of 4.0 billion cubic meters annually and it is planned by the government to increase the
reuse up to 8.0 billion cubic meters annually. Simultaneously, the groundwater withdrawals are
planed to increase from 4.0 to 7.0 billion cubic meters per year in Nile valley and Delta. The
groundwater aquifers in the Nile Delta and Nile valley are replenished from the leakage water from
the river, irrigation canals and drains. Therefore water supply augmentation through recycling
(drainage water reuse) and capturing of the water losses (groundwater utilization) increases the
overall efficiency of the water resources system in Egypt.

Although the water requirements of municipal and industrial sectors have the first priority to be
met, the agricultural policies and innovations had its impact and pressure on the operation of the Nile
water system. During the last decade the government of Egypt has liberalized the cropping pattern
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

83
and farmers became free to select the crops they like to grow, except for rice which is restricted to the
permission of the government.

The Nile system is viewed as two sub-global systems; Nile valley and Delta. In Nile valley,
drainage water returns back to the river and drainage reuse practices are limited and occur only on a
very small scale. Delta region is the potential area for drainage reuse promotion. The linkage between
the different hydrological areas, i.e. between global or sub-global and the local levels, is the water
flows and change in salts concentration. The water and salt balance analyses are, therefore, viewed
essential for evaluating these levels.

At present, operation of the Nile system is successful in meeting the current water demands.
However, Egypt must do more with less water (Abu Zeid, 1997) to cope with future development
plans for the country and with projected future increase in population. Egypt has introduced different
innovations to the existing system in order to save water from old lands to be diverted to the new
lands. These innovations included Irrigation Improvement Project which started in 1980s, and
drainage water reuse program which started in 1970s. The irrigation improvement project activities
include improvement of secondary and tertiary irrigation delivery network and leveling of agricultural
fields. It is expected that this project will result in saving irrigation water and improving the agricultural
productivity. Drainage water is one of the valuable water resources in Egypt created by the intensive
and large irrigation/drainage system. It accumulates the excess of irrigation water with appropriate
quality that can be reused within the system.


WATER RESOURCES

Water resources in Egypt are represented with the annual quota from the Nile water; the limited
amount of rainfall; the shallow and renewable groundwater reservoirs in the Nile Valley, the Nile Delta
and the coastal strip; and the deep nonrenewable groundwater in the eastern desert, the western
desert and Sinai. The non-traditional water resources include reuse of drainage and municipal waste
water, and desalination of seawater and brackish groundwater. The rainfall is very limited, and mainly
at the north coast on the order of 200 mm/year. The more we go southward to Cairo the less this
amount is, then the decrease is rapid until the southern areas where there is almost no rainfall.
Rainstorms in Egypt take place in autumn, winter and spring and their frequency and intensity differ
from year to year. Table 1 shows the available water resources in Egypt and the water use of the
different sectors at year 2000.
Table 1. Water Uses and Available Resources in Year 2000

Water Uses (bcm/year) Water Resources (bcm/year)
Sector Amount Resource Amount
Municipalities 5.25 Nile river 55.5
Industry 3.5 Groundwater (Delta
and Valley)
5.5
River Transport 0.25 Deep Groundwater 0.8
Fisheries - Drainage Water
Reuse

Hydropower - - Canals in the Delta
Region
4.5
Agriculture 63.00 - Nile river and Bahr
Youssef
5.0
- Illegal Uses 3.0
Waste Water Reuse 0.2
Rainfall and Flash
Floods
0.5
Evaporation Losses (3)
Total 72.00 Total 72.00


Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

84


GOVERNMENT POLICIES AND FARMER PRACTICES ON CROPPING PATTERN

Crop Liberalization Policy

Selection of crops, mainly cotton and basic food grains, to be grown every year was controlled by
the government till year 1985. Farmers were responsible to deliver a preset quota of grains or cotton
to the government. The government was controlling all marketing channels and prices. This policy
resulted in an average of 30% net tax on agricultural outputs, which enabled the government to
continue subsidizing consumers and financing industrialization. The government controlled also the
crop rotation among growers. Figure (1) shows a typical crop rotation types (3-turn crop rotation).
During the period 1985-1990, partial free cropping pattern policy was applied. It included liberalizing
most of crops except for cotton and rice.

Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
Year 1
Year 2
Year 3
Rice
Wheat/Long Berseem
Cotton
Short Berseem
Short
Berseem
Beans/Berseem
Maize
Beans/Berseem
Nili
Long Berseem
/Wheat


Fig. 1 Three-Turn Crop Rotation

By year 1995, all crops were liberalized except for rice and sugar cane, which was restricted to the
availability of water resources and future expansion plans. In fact this policy resulted in a wide
variation in crop selection among growers. Cotton and rice areas were clearly affected, as most of
growers shifted from cultivating cotton to rice. Although the MPWWR restricted the rice area not to
exceed 1.0 million feddans, but the actual cultivated rice area reached 1.5 million feddans in year
1995, i.e. 50% more than the targeted area. On contrary, the cotton area declined to reach about 0.7
million feddans. Then, the government started to motivate cotton growers through creating new
mechanisms for cotton marketing.

Among different regions in Egypt, crop diversification varies according to the climatic and soil
conditions. In Upper Egypt sugar cane is the largest single crop in summer, particularly in Aswan and
Qena governorates. The second summer crop which competes with sugar cane in this region is
sorghum and maize. In winter, wheat and clover are the major crops. In middle Egypt region, maize is
major summer crop where it covers about 40% of the cultivated area. In winter, wheat and clover
cover, situation looks different as the temperature is rather lower and soil conditions are different. In
southern part of the Delta, rice is not allowed to be grown. Then, maize is the major summer crop and
wheat is the major crop in winter. In the northern part of Delta rice is allowed to be grown where the
water table is rather high. Figure (2) shows crop diversification for the different regions of Egypt.

Introducing short-duration crops

The government of Egypt has launched, in 1999, a program for introducing the short-term varieties
of rice, to replace the log-duration one. The short-duration rice can be harvested after 120 days
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

85
instead of 160 days of the long-term rice. It is assumed that the short-term varieties will consume less
water compared to the traditional varieties.

Transplanting and Salt-Tolerant Crops

Cultivating the drought tolerant varieties will result in maximizing the crop water use efficiency.
Moreover, using the transplanting method in cultivating some crops instead of direct seed sowing will
reduce the water requirements of those crops. In the case of rice, the reduction in the water
requirement is estimated by 10-15%.

Replacement sugarcane by sugar beet for sugar production, is another governmental plan. The
government already introduced with great success the sugar beet in the western delta. The problem
in upper Egypt that sugar cane is necessary for the sugar factories in the region. Sugar cane
agriculture in upper Egypt therefore is expected to continue along the life time horizon of these
factories. Introduction of modern irrigation in the sugar cane fields is being carried out to reduce the
water application and to increase water use efficiency of this high water requirement crop.

Cultivating the salt tolerant crop varieties will save the fresh water as they can be irrigated by
marginal low quality water. Agricultural drainage water can be used either directly or after mixing with
fresh water, based on the salinity of drainage water. This is practiced in Egypt on a large scale and
mainly in the north of the Nile Delta as well as in EI-Salam canal project in the eastern Delta and
Sinai.


Intensifying Cropping Pattern

The crop intensity indicates how land is cultivated. In many case the land is cultivated twice in one
year; winter and summer crops. The crop intensity ranges from 1.3 to 2.36. The higher intensity
indicates that the land is cultivated three times a year. Crop intensity is higher in Lower Egypt
compared to other regions. The overall crop intensity in Egypt is about 1.81.

Winter Crops

Long-duration clover occupies about 85% of the total cultivated area in Port Said. Damietta and
Menoufia Governorates showed also about 50% and 42% of its area, respectively, grown with long-
duration clover.

Wheat and Long-duration clover occupies about 38% for each, of the cultivated in Fayoum. While
in Beni Suef, wheat occupies about 40% of the cultivated area and long duration clover occupies less
than 20%. Long-duration clover is the dominant crop in Giza that constitutes more than 35% of the
cultivated area, but wheat represents less than 10%. n upper Egypt governorates, the dominant crop
is wheat which occupies 40%, 57%, 37%, 22% of the cultivated area in Assuit, Sohag, Qena, and
Aswan. Long-duration Clover is rather less than 20 % in these governorates.

Summer Crops

Maize, Rice, Cotton and Sorghum are the major summer crops in Egypt as illustrated in Figure 2.
Rice occupies the largest area in Lower Egypt (35% of the total area). Maize Occupies 37% of the
cultivated area in Middle Egypt, 24% in Lower Egypt and 23% in Upper Egypt. Cotton occupies 18%
of the cultivated area in Lower Egypt, 13% in Middle Egypt and 6% in Upper Egypt. Sorghum is a
dominant crop in Upper Egypt where it occupies 25% of the total cultivated area and 4% in Middle
Egypt.

Nili Crops

Nili Crops are planted during the period July September. These crops are minors and occupy
7%, 21% and 8% in Lower, Middle and Upper Egypt, respectively. Nili crops include mainly maize.


Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

86
Perennial Crops

Perennial crops include Sugar Cane and Orchards. Sugar Cane is the most dominant crop in
Upper Egypt where it occupies 24% of the total cultivated area. In Middle Egypt Sugar Cane occupies
only 8%. Orchard occupies 9%, 8% and 4% in Lower Egypt, Middle Egypt and Upper Egypt,
respectively. Orchards showed a high percentage in Desert Governorates; 47% of the

Diversity of Winter Crops among Regions in Egypt
0%
10%
20%
30%
40%
50%
Lower Egypt Middle Egypt Upper Egypt Desert Governorates
%

o
f

C
o
m
m
a
n
d

A
r
e
a
Long-term Clover Wheat Short-Term Clover
Diversity of Summer Crops among Regions in Egypt
0%
10%
20%
30%
40%
50%
Lower Egypt Middle Egypt Upper Egypt Desert Governorates
%

o
f

C
o
m
m
a
n
d

A
r
e
a
Maize Rice Cotton Sorghum
Diversity of Perennial Crops among Regions
0%
10%
20%
30%
40%
50%
Lower Egypt Middle Egypt Upper Egypt Desert Governorates
%

o
f

C
o
m
m
a
n
d

A
r
e
a
Sugar Cane Orchards


Fig. 2 Crop diversity in Different Regions

CROP PRODUCTION

Crop yield and production is affected by different factors such as availability of water, quality of
irrigation water, soil type and climatic conditions. As mentioned before, irrigation water deteriorates as
moving northwards downstream the system but still within the safe limits. However, the new
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

87
reclaimed areas at the end of the system such as in the north of Delta are depending, on a great
extent, on drainage water which has salinity of higher than 1500 ppm.

Winter Crop Yield

Table 2 presents the crop yield on governorate basis (MALR, 1996). It includes Wheat, Barely,
Beans, Clover and Tomato. It could be concluded from these data that there is no big difference of
wheat yield in Lower, Middle and Upper Egypt where the average yield are 17.03, 17.71 and 16.5
Ardeb/fed, respectively. But, there is significant difference in yield for tomato crop. The average yield
of tomato was 14.3, 12.24 and 20.29 in Lower, Middle and Upper Egypt regions, respectively. This
may indicate the impact of low water quality on tomato production.

Table 2. Winter Crop Yield at Governorate Level (Source: MALR, 1996)

Governorate
Old
Land
New
Land
Old
Land
New
Land
Old
Land
New
Land
Old
Land
New
Land
Old
Land
New
Land
Nobaria 11.03 8.87 9.44 19.40 13.77
Alxandria 15.54 13.00 8.49 8.18 8.90
Beheira 17.04 13.83 7.91 28.27 7.23
Gharbia 16.75 17.05 10.72 23.84 9.73
Kafr Elsheikh 17.58 13.63 11.97 8.23 10.06 25.77 20.00 9.17
Dakahlia 17.05 14.42 17.50 12.87 9.75 21.38 14.50 7.24
Damietta 14.41 9.22 8.59 21.96 6.55
Sharkia 17.1 13.36 16.52 9.64 7.00 25.35 14.00 13.76
Ismailia 13.87 13.01 13.52 8.24 6.36 4.00 14.11 29.18
Port Said 8 4.00 7.00 15.08 17.00 6.00
Suez 13.19 13.13 13.99 10.30 6.88 19.98 8.70
Menoufia 17.99 14.18 16.00 8.92 6.00 28.18 26.60 8.98
Qaliobia 16.79 17.74 8.91 33.34 14.16
Cairo 12.06 6.76 13.50 19.51 7.91
Total for Lower Egypt 17.03 14.16 9.34 9.37 25.40 14.28 13.77
Giza 20.67 14.00 18.55 8.50 32.73 14.24
Beni Suef 15.86 13.77 13.00 5.91 32.80 13.00 14.26
Fayoum 16.81 14.00 11.70 12.00 7.64 18.60 20.00 9.90
Menya 19.39 13.00 12.79 10.64 6.88 22.00 12.00 11.02
Total for Middle Egypt 17.71 12.08 6.92 24.41 12.24
Assuit 16.49 12.44 14.10 8.05 35.52 15.35
Sohag 16.96 14.00 12.63 8.84 6.96 6.98 34.26 39.50 25.07
Qena 15.79 11.45 10.95 11.22 6.16 5.64 23.58 23.34 25.64
Aswan 15.99 5.00 8.94 6.02 21.65 6.06
Luxor 18.13 10.56 0.00 7.50 6.50 26.11 13.00
Total for Upper Egypt 16.49 11.00 7.59 6.01 34.25 21.77
Total for Nile Valley & Delta 17.06 12.11 13.50 9.38 8.55 9.28 25.84 20.29 16.78
New Valley 12.83 10.24 12.31 7.49 23.65 13.15
Matrouh 1.17 1.73 8.50 22.00
North Sinai 1.56 1.81 4.00
South Sinai 4.14 0.00 10.53
Total for Desert Govern. 3.95 2.23 12.31 7.84 23.32 10.07
Tomato
(Tons/fed)
Wheat
(Ardeb/fed)
Barely
(Ardeb/fed)
Beans
(Ardeb/fed)
Clover
(Tons/fed)

Note: Ardeb = 150 kg; feddan = 0.42 hectares



Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

88
Summer Crop Yield

Table 3 shows the crop yield for Sorghum, maize, rice, sugar cane and cotton. There is no
significant difference in maize yield among different regions in Egypt. But, cotton yield showed
significant variation among regions. Cotton yield was found 6.97 Qentar/ fed1. In Lower Egypt, 8.2
Qentar/fed. In Middle Egypt and 11.42 Qentar/fed. In upper Egypt. This indicates the impact of water
quality on cotton yield among different regions.

Table 3. Summer Crop Yield at Governorate (Source: MALR, 1996)

Governorate
Old
Land
New
Land
Old
Land
New
Land
Old
Land
New
Land
Old
Land
New
Land
Old
Land
New
Land
Nobaria 8.25 21.22 21.76
Alxandria 18.10 3.21 40.00
Beheira 23.40 3.65 24.73 6.52
Gharbia 22.70 3.41 41.75 5.81
Kafr Elsheikh 20.63 14.47 3.39 31.94 5.39
Dakahlia 21.85 17.00 3.65 31.20 5.85
Damietta 19.98 3.05 31.25 6.21
Sharkia 20.28 3.48 39.37 7.95
Ismailia 18.72 3.04 3.78
Port Said 8.48 2.29 4.74
Suez 15.73 2.00 29.46
Menoufia 19.79 16.44 2.71 36.01 8.19
Qaliobia 20.33 20.00 3.41 40.30 9.06
Cairo 12.27 28.81
Total for Lower Egypt 8.25 21.25 3.51 36.07 21.76 6.97
Giza 15.82 22.74 33.51 4.14
Beni Suef 19 18.75 25.63 8.75
Fayoum 11.1 19.13 5.95 2.96 31.33 6.17
Menya 14.78 21.92 19.50 46.54 9.04
Total for Middle Egypt 11.74 21.07 2.96 44.33 8.20
Assuit 14.09 20.12 3.49 40.77 11.80
Sohag 12.89 12.45 20.60 20.00 47.66 10.64
Qena 12.74 11.65 18.72 13.61 46.88 42.49 4.18
Aswan 11.65 15.95 47.42
Luxor 10.37 15.63 8.00 48.95
Total for Upper Egypt 13.4 11.07 19.77 47.22 42.49 11.42
Total for Nile Valley & Delta 13.11 20.98 19.57 46.73 40.47
New Valley 11.49 8.81 2.76 20.00
Matrouh 14.00
North Sinai
South Sinai
Total for Desert Govern. 3.95 2.23 12.31 7.84 23.32 10.07
Cotton (hair)
(Qenta/fed)
Sorghum
(Ardeb/fed)
Maize
(Ardeb/fed)
Rice
(Tons/fed)
Sugar Cane
(Tons/fed)

Qnetar = 50 kg for cotton





Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

89

Crop Yield in New Lands

New lands have more deteriorated water quality in delta region. The current cultivated new lands
is still have safe limits for water quality, but the future reclamation areas may have more deteriorated
water quality due to using drainage water on a great extent particularly on Northern Delta. The current
production of crops in new lands showed nearly equal levels of production particularly for vegetables
and maize.

It could be concluded that the crop production is higher in the upstream reaches of the Nile (Upper
Egypt) and then decline as moving northwards to Delta region. But the difference is not very big. The
current production of new lands is not far less than the old lands.


Crop yield compared to international Figures

Table 4 presents a comparison for crop yields for different countries. Egypt showed a high level of
crop yield compared to other countries. Egypt is ranked on the top for sugar cane, rice, sesame and
sorghum yield.


Table 4 Egypt's Rank among World Countries According to Productivity of Main
Crops

Egypt's Rank Crop Yield (ton/fed)
1 Sugar cane 50.4
1 Rice 4.133
1 sesame .531
1 Sorghum 2.397
2 Peanut 1.329
2 Broad bean 1.372
3 Lentil .74
4 Wheat 2.755
7 Dry onion 14.064
9 Cotton 1.1
10 Maize 3.466
10 Barely 1.153
12 Potatoes 10.27
13 Sugar beet 20.29
Note: Number of countries used for comparison are 20
Source: Compiled by the General Department of Agricultural Studies (From FAO data)


Water Use Efficiency


The overall efficiency of the water system in year 2000, a shown in Figure 3, equals the
consumption as a percentage of the total inflow, is about 71%. This efficiency is relatively high taking
into consideration that the prevailing irrigation method is surface irrigation, which has a low efficiency.
This high system efficiency is probably attributed to the intensive efforts of government in O&M, and
to the current recycling practices, in addition to the considerable experience of Egyptian farmers.





Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

90













`






Note: all figures are in BCM/y




CONCLUSIONS

Although water resources are limited and scarce in Egypt, great efforts have been and are being
conducted to increase water use efficiency and water productivities. The current water use efficiency
exceeds 70% on the national level. However, plans are being prepared to increase this level of
efficiency in order to increase the cultivated area by about 40% by year 2017. These plans include
improvement of irrigation delivery systems, introducing low-water-consuming crops, introducing salt-
tolerant crops, and reuse of drainage water. Intensifying cropping pattern was one of the factors
contributed to increasing the water productivity.


REFERENCES

Ministry of Water Resources and Irrigation, (1998). National Policy for Drainage Water Reuse,
APRP-Water Policy Reform Project, Report No. 8, Cairo, Egypt.
Abdel Azim, R. A., (1999). Agricultural Drainage Water Reuse in Egypt: Current Practices and a
Vision for Future Development, Ph.D. thesis, Faculty of Engineering, Cairo University, Cairo,
Egypt.
Ministry of Agriculture and Land Reclamation (1999), The Economic Sector, Agro-economic Bulletins,
several issues, Cairo, Egypt.
Allam, M.N. (2000), Water Resources: Utilization and Management, Cairo, Egypt.







HAD
Rainfall
0.5
Municipalities
and Industry
Old
55.5
Evaporation
3.0
Consumption
1.75
Fig. 3 Water Status in Year 2000
Evapotranspiration

38

13.255
Out Flow
Agricultre
Sea
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

91




























































Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

92





WATER USE EFFICIENCY
AND WATER PRODUCTIVITY IN GREECE



A. Karamanos*, S. Aggelides**, P. Londra**
* Laboratory of Crop Production. Agricultural University of Athens, 75 Iera Odos, 11855 Botanicos,
Greece. E-mail: akaram@aua.gr
** Laboratory of Agricultural Hydraulics. Agricultural University of Athens, 75 Iera Odos, 11855
Botanicos, Greece. E-mails: ags@aua.gr and v.londra@aua.gr


SUMMARY This work presents a summary on water resource availability in Greece with an
emphasis on irrigation water use since about 85% of water withdrawal is for agricultural purposes.
Irrigation methods, crop water requirements, crop production and crop water productivity data are
elaborated are discussed. The state of art of research and agricultural activity as related to Water Use
Efficiency (WUE) and Water Productivity (WP) in Greece is analyzed by means of: the hydrological
aspects and agronomic management strategies, the eco-physiological aspects and sustainability
including social, economic, environmental and political measures. Furthermore, the initiatives and
activities to improve Water Use Efficiency and to increase food production by using less water are
reported.

Key words: irrigation, water use efficiency, crop production.


INTRODUCTION

Greece has a population of about 11,000,000 and occupies an area of 131,962 Km
2
. The types of
land use are:
a. Arable 30%
b. Forest 19%
c. Pastures 43,4%
d. Others 7,6%
The cultivated area is 34,638 Km
2
out of the total arable area of 38,986 Km
2
. The irrigated land
occupies 14,305 Km
2
with the following percentage of cultivations (National Statistical Service of
Greece, 2001):

65,06%
7,94%
24,19%
2,81%
crops vegetables trees vines
Total irrigated land

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

93

Fig. 1. Total irrigated land: percentage of cultivation

The climate is typically Mediterranean with all the subtypes of Mediterranean climate due to
topographical relief, from the warm and dry type (e.g. Cyclades) to the alpine type (mountainous
areas above 1500 m).

The annual precipitation ranges from 400 mm in Athens and Cyclades islands to more than 1500
mm in the high mountainous areas with values of 700 mm in Eastern Greece and 1000-1200 mm in
Western Greece and the Ionian islands.

The total annual precipitation is estimated to be 116,689 hm
3
with 50.9% (59,371 hm
3
) and 31.9%
(37,190 hm
3
) to be lost by evaporation and runoff respectively and only 17.2% (20,133 hm
3
) infiltrated
into soils. The total water consumption was estimated at 5,500 hm
3
/year increasing by more than 3%
annually (Ministry of Development, 2002).

The major water use in Greece is for agricultural purposes (85%) whereas the domestic and
industrial uses are 13% and 2% respectively. The increased demand for water cannot always be met
despite adequate precipitation. Water imbalance is often experienced, especially in the coastal and
south-eastern regions, due to the temporal and spatial variations of precipitation. This results from the
increased water demand during the summer months and the difficulty of transporting water due to the
mountainous terrain.

The irrigation methods used in Greece are surface, sprinkler and drip irrigation. The general trend
is to abandon gradually surface irrigation and move on to more efficient methods like the sprinkler and
drip irrigation (Fig. 2).

0
50,000
100,000
150,000
200,000
250,000
300,000
1
9
8
4
1
9
8
6
1
9
8
8
1
9
9
0
1
9
9
2
1
9
9
4
1
9
9
6
1
9
9
8
Year
I
r
r
i
g
a
t
e
d

a
r
e
a
(
h
a
)
Surface irrigation Sprinkler irrigation Drip irrigation


Fig.2. The trends of irrigation methods used in common reclamation works

It is accepted that the water application efficiencies for surface, sprinkler and drip irrigation are
75%, 85% and 90% respectively. The water conveyance and distribution efficiency is 70%, 85% and
95% for transportation of water with earthen, concrete channels and pipelines respectively (Ministry of
Agriculture, Land Reclamation Service, 1991).
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

94

The amounts of annual reference evapotranspiration (ETo) have been calculated in national level
and ranged from 1058 mm in Northern Greece up to 2015 mm in the arid and semiarid zone of South-
eastern Greece.

The irrigation requirements depend on the crop and the local soil-climatic conditions and they
reach 170 mm for vines in the temperate regions of Northern Greece (Thessaloniki) and 440 mm for
maize in the warm and wet regions of Southern Greece (Pyrgos) (Ministry of Agriculture, 1990).

To meet these requirements, irrigation has been extended (irrigation covers 40% of the total
cultivated area) under common reclamation works and private irrigation systems. The water sources
used differ between public and private networks. The public networks mainly use surface water, while
the private ones use underground water.

The maximum calculated crop water requirements reach a value of 4,089 Km
3
of water and the
actual water use is 6,833 Km
3
. This means that the Total Water Use Efficiency is 60%. This low value
is due to a poor land levelling, the aged water distribution systems, the high percentage of surface
irrigation systems and transportation and distribution of water (Tsanis et al., 1996).

In Greece, there are values of crop water productivity in terms of crop production (kg) in relation to
the volume (m
3
) of water used. These values come from irrigation experiments conducted in several
parts of the country on behalf of the Ministry of Agriculture. The values given below are a
representative selection and refer to good years, i.e. in seasons when crops gave the best
production for the region (Ministry of Agriculture, 1990).

Table 1. Crop production, water use and crop water productivity in Greece
(Source: Ministry of Agriculture, 1990)
Crop Production
(kg/ha)
Cubic meters of water used for
actual evapotranspiration-ETc
(m/ha)
Crop water
productivity
(kg/ m)
Maize 12000 4980 2.41
Cotton
(not ginned)
3550 4500 0.79
Industrial tomato 97000 3900 24.87
Watermelons 43000 3570 12.04
Vines 14800 1850 8.00
Potatoes 30000 4800 6.25
Lettuce 36000 4600 7.82


WATER USE EFFICIENCY AND WATER PRODUCTIVITY

The state of research and agricultural activity as related to Water Use Efficiency (WUE) and Water
Productivity (WP) in Greece can be focused in the following three aspects:
The hydrological aspects and agronomic management strategies of WUE and WP.
The eco-physiological aspects of WUE and WP.
Sustainability and economic aspects of WUE and WP.

The hydrological and agronomic management aspects of WUE and WP
Agricultural and irrigation engineers facing the problem of rational irrigation planning, rarely
have at their disposal extensive information on soil-plant-atmosphere system. Therefore, there
is demand for simpler approaches to estimate actual evapotranspiration. The attempts that
have been made in Greece related to the above issue are:

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

95
1. The crop coefficient (Kc) for sweet shorgum was estimated under the conditions of Biotia area
(Dercas et al., 1996).
2. Poulovassilis et al. (1998) compared the estimated heat flux densities obtained through the
various methods used with actual heat flux densities, as these are measured into the soil
profiles. The comparison was encouraging and therefore heat flux densities calculation
methods may be used for following heat flow into the soil profile.
Another attempt to help rational irrigation planning was made by Anadranistakis et al. (1999a),
who estimated actual and maximum evapotranspiration for neutral instead of actual
atmospheric stability, employing a model based on the equation of Shuttleworth and Wallace
(1985).
Anadranistakis et al. (2000) investigated two parametarizations of the aerodynamic resistance
control vapor transfer and they have applied to an evapotranspiration estimating model in order
to find out their effect on the estimated (separate) values of transpiration and evaporation
throughout the biological cycle of a crop.
A semi-empirical approach was proposed for estimating actual water losses from crops (cotton,
wheat and maize) assuming that the ratio of actual to maximum evapotranspiration (ET/ETm) is
an exponential function of the water content in the root zone (Poulovassilis et al., 2001).
Spanomitsios and Paraskevopoulou-Parousi (2001) showed that the comparison of strawberry
plant transpiration rates with a linear equation of transpiration rate and the simplified form of
Penman-Monteith equation agreed well with the last equation.

Furthermore several attempts have been made by studying the reference evapotranspiration
equations under the arid and semi-arid conditions in Greece.

The results of the comparative evaluation of reference evapotranspiration estimated according
to seven different versions of the Penman method were presented. The Penman 1963 VPD#3
was considered as the most reliable for the climatic conditions of Copais (Greece)
(Poulovassilis et al., 1996).
An attempt was made to estimate the wind function parameters, entering the aerodynamic term
of Penman equation for hourly evapotranspiration estimates. Useful results were obtained
concerning the possibility of using the Penman equation in estimating hourly or daily reference
evapotranspiration (Kerkides et al., 1997).
The establishment of a relationship between G/RN and L both for day and night was attempted
during the development of a wheat crop, under varying soil moisture regimes (Anadranistakis et
al., 1997).
An attempt was made to estimate the wind function parameters entering the aerodynamic term
of Penman equation for hourly evapotranspiration estimates. Useful results were obtained
concerning the possibility of using the Penman equation in estimating hourly or daily reference
evapotranspiration (Kerkides et al., 1999).
The Penman-Monteith method for estimating reference evapotranspiration (ETo) is considered
as the most reliable. Using measurements from the Copais (Greece) area, the expression of Rn
as a function of Rs and of Rs as a function of the daily sunshine duration (n) is attempted,
considering that G is a portion of Rn. Then ETo is calculated both with measured and estimated
Rn. The comparison of the corresponding ETo values was proved satisfactory (Alexandris,
2000).
A new empirical equation for estimating hourly reference evapotranspiration ETo is proposed.
For the calibration of the proposed model, data collected from Copais (Greece) and from CIMIS
equation (Davis Sacramento, CA) were used (Alexandris and Kerkides, 2003).

For an appropriate irrigation scheduling, the atmospheric demand for water vapor, the soil
characteristics and the plant specific features, i.e. the soil-plant-atmosphere continuum as a
dynamic system, should be considered. In semi arid regions, such as Greece, rainfall is
unevenly distributed over the year. More than 80% of the mean annual precipitation falls during
the months October-February. Meanwhile during the dry seasons of spring and summer tourist
mobility is at its peak, exercising a substantial pressure on good quality water reserves. The
fact that irrigation is also required mostly in spring and summer means that water allocation and
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

96
water use efficiency are very important. It is therefore necessary to develop models to
conveniently estimate actual Crop Water Requirements.
For the Mediterranean countries the future challenge will be the increase in food production using
less water. In Greece many attempts has been made to this direction. Some of them are:

Louizakis (1994) estimated the water requirements and water productivity for tobacco in
Greece.
Crop water requirements were determined using the Penman-Monteith method with climatic
data from meteorological stations in Greece and the USDA method. It was found that the total
irrigation requirements for Greece for the year 1991 were ranged from 3073 to 4069 Mm3/yr
(Tsanis et al., 1996).
A relationship between maize yield and water consumption has been developed in Thessaloniki
plain during a four year experiment. Maximum yield was attained when the depth of applied
water ranged from 700 to 900 mm (Panoras et al., 1997a).
The same relationship in the same place but for winter wheat (cv. Yecora) was established after
a five year experiment. It has been shown that maximum yield was attained when the amount of
applied water ranged from 400 to 500 mm and the best application time of water was the
second and the third critical stage of the winter wheat (Panoras et al., 1997b).
Water productivity and water requirements were studied for cotton (cvs Zeta-Z and Korina) in
Central Greece for the year 1997 (Polychronidis, 1998).
Anadranistakis et al. (1999b) represented a model for estimating crop water requirements
throughout crop development. The model has been validated with meteorological and crop data
collected from experimental fields of the Agricultural University of Athens. The results taken
from three crops (cotton, wheat and maize) against soil moisture profile changes were very
satisfactory. The agreement between observed and estimated evapotranspiration was within
8%.
Crop water requirements were studied for sorghum during a two-year experiment in Central
Greece (Dercas and Liakatas, 1999).
An estimation of the total water requirements in the Prefecture of Larissa irrigated both by
private and public boreholes and by surface waters during the year 1999 by using FAO
Penman-Monteith method was carried out (Sakellariou-Markantonaki and Vagenas, 2003).

Saving irrigation water in arid and semiarid zones where there is scarcity of water is a demand
of major priority. The method of irrigation ultimately chosen must be the best as far as the water
use efficiency is concerned. The water related efficiencies from water conveyance, storage,
distribution and application must be high. Some works related to this subject are:

Water use efficiency was studied under the soil-plant-atmosphere conditions of Greece by
Poulovassilis et al. (1991) and they proposed measures to increase it.
Papamichail and Papadimos (1996) studied the water use efficiency for graded border
irrigation.
The same authors (Papamichail and Papadimos, 1996) studied the water use efficiency for
furrow irrigation used for irrigation.
Ampas (1998) studied the irrigation efficiency in furrows used for row crops. He found that the
calculation of inflow rate at each time step maximizes the application efficiency.
Mayropoulos (2001) studied water use efficiency and he recommended several measures at
technical, economic and legislation level.


The eco-physiological aspect of WUE and WP
Here there are some aspects of the actual situation of research and agricultural practices
concerning crop water productivity in Greece.

The water productivity of potatoes was studied experimentally in the region of Attika and it was
found to be 6.25 kg/m3 (Aggelides et al., 1984).
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

97
The water productivity of vine (Cardinal) was studied experimentally in the region of Attika and it
was found to be 8.0 kg/m3 (Panagiotou and Aggelides, 1987).
Poulovassilis et al. (1993a) studied the crop water productivity reduction due to irrigation with
brackish groundwater.
The crop water productivity of sweet sorghum was experimentally studied in Copais area with a
high water table. Results showed a high crop water productivity due to the small amount of
irrigation water Dercas et al., 1994).
The crop water productivity of sweet sorghum was studied in relation to plant density (Dalianis
et al., 1996). It was found that decreasing plant density increased water productivity.
The water productivity of winter wheat (cv. Yecora E) was studied in the plain of Thessaloniki
and it was found that the maximum yield was attained when the amount of applied water
ranged from 400 to 500 mm (Panoras et al., 1997).
Crop water productivity and water use efficiency were studied for sweet sorghum (cv. Keller) by
Dercas et al. (1998). They found a good relationship between dry matter yield and the quantity
of applied water.
The water productivity of cotton was studied and compared with water consumption
(evapotranspiration, ET), simulated by GOSSYM, a simulation model of cotton growth and yield
(Gertsis et al., 1998).
The water productivity of lettuce was studied in the area of Achaia (Peloponnese) and it was
found to be 7.82 kg/m3 (Aggelides et al., 1999).
The water productivity of two varieties of sweet sorghum (cvs Keller and MN 1500) was studied
after two-year experiments in Boeotia. Drip irrigation was applied. The water productivity was
very high (Dercas et al., 2000).
Water productivity of Medicago sativa studied by Lazaridou and Noitsakis (2002) in pure and
mixed crop grown with grass. It was found that water productivity exhibited higher values during
early spring in relation to the measurements of the remaining period and in the mixed crops
despite their smaller productivity compared with pure crop.
Fisarakis (2002) studied the reduction of growth observed at relatively low salinities, often
before the appearance of visible symptoms. The unfavorable effects of salinity on grapevines
were increasing as the level of salinity increased and were demonstrated in terms of a reduced
plant growth, bunch number, berry size and yield.

Sustainability and economic aspects of WUE and WP

This issue contains social, economic, environmental and political measures that add additional
importance to the sustainability of agricultural production. Some of these works are:

Poulovassilis et al. (1993b), studying sea water intrusion in the area of Irria (East Peloponnese),
found the same results as in the case of Argolis.
Poulovassilis et al. (1994), studying sea water intrusion in the area of Argolis, showed that the
salinity status of the soil is correlated with that of the ground water used for irrigation and that
both of them affect adversely crop yields according to their sensitivity.
A comprehensive study has been undertaken for in Argolis (Peloponnese) in an attempt to
identify the factors which caused soil degradation due to salinity and to promote remedial
measures (Poulovassilis et al., 1996).
This study presents the climate change, the water availability, the crop water requirements, the
improvement of water use efficiency and the necessity of high priority in water resources policy
(Chartzoulakis et al., 1997).
Mimides et al. (1997) studied ground water degradation due to the sea intrusion in the region of
Irria (East Peloponnese). Various softwares such as SURFER, MINEQL and DUROV diagram
were utilized.
The plain of Argos in Southern Greece was investigated for sea water intrusion, ground water
pollution and the suitability of ground water for irrigation purposes (Alexandris et al., 1998).
An attempt was made to assess the effect of irrigating tomatoes with saline water in a
greenhouse (Kerkides et al., 1998).
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

98
The influence of low quality irrigation water to agricultural production was studied by
Giannoulopoulos et al. (2002).
Kountouras and van Leeuwen (2002) studied the environmental effects on grapevine grown in
the Nemea region and especially a) the vine water status and physiological parameters and b)
vegetative growth, fruit ripening and yield.
An initiative for Mediterranean water policy is presented to overcome water competition
between agriculture and tourism in Cyclades and especially in Naxos island (Epp et al., 2003).
Chartzoulakis and Psarras (2005) studied the current trends and predictions for the future
climate changes in the area of Crete and discuss the factors that will most probably limit or
enhance photosynthesis and productivity of the most important tree crops in a semi-arid
Mediterranean environment.


REFERENCES

Aggelides, S., G.K. Chardas, P.D. Tsakaleris and G.I. Stamos, 1984. Irrigation of potatoes based on
soil water negative pressure measurements. Agricultural Research, 8:45-55.
Aggelides, S., I. Assimakopoulos, P. Kerkides and A. Skondras, 1999. Effects of soil water potential
on nitrates content and the yield of lettuce. Communication in Soil Science and Plant Analysis, Vol.
30 (1&2).
Alexandris, S., 2000. Reference evapotranspiration estimation with the Penman-Monteith method
using standard meteorological parameters. 5o Hellenic conference of Meteorological Climatology
and Atmospheric Physics. Thessaloniki 2000, pp.439-446.
Alexandris, S. and P. kerkides, 2003. New empirical formula for hourly estimations of reference
evapotranspiration. Agricultural Water Management, 60:157-180.
Alexandris, S., P.M. Allen, I. Black, N. Kalamaras, P. Giannoulopoulos, M. Lemon, T. Mimides, A.
Poulovassilis, M. Psychoyou and R.A.F. Seaton, 1998. Agricultural production and water quality in
the Argolic valley, Greece. The Archaeomedes project, Understanding the natural and
anthropogenic causes of degradation and desertification in the Mediterranean basin. Research
results. EUR 18181. ISSN 1018-5593, pp 281-326.
Ampas, V., 1998. Optimum application efficiency of furrow irrigation. Geotechnical Scientific Issues,
Volume 9, Issue III, pp.4-12.
Anadranistakis, M., A. Liakatas and P. Fragouli, 2000. Parametraization of aerodynamic resistances
in evapotranspiration estimating models. 5o Hellenic conference of Meteorological Climatology
and Atmospheric Physics. Thessaloniki 2000.
Anadranistakis, M., P. kerkides, A. Liakatas, S. Alexandris and A. Poulovassilis, 1999a. How
significant is the usual assumption of neutral stability in evapotranspiration estimating models?
Meteorol. Appl. 6:155-158.
Anadranistakis, M., A. Liakatas, P. kerkides, S. Rizos, J. Gavanosis and A. Poulovassilis, 1999b.
Crop water requirements model tested for crops grown in Greece. Agricultural Water
Management, 45:297-316.
Chartzoulakis, K. and G. Psarras, 2005. Global change effects on crop photosynthesis and production
in Mediterranean: the case of Crete, Greece. Agriculture, Ecosystems and Environment, 106:147-
157.
Chartzoulakis, K., A.N. Angelakis and P. Skylourakis, 1997. Irrigation of Horticultural crops in the
island of Crete, Greece. ISHS Acta Horticulturae 449: II International Symposium on Irrigation of
Horticultural Crops. Chania, Crete, Greece, 1 August 1997.
Dalianis, C., E. Alexopoulou, N. Dercas and Ch. Sooter, 1996. Effect of plant density on growth,
productivity and sugar yields of sweet sorghum in Greece. Biomass for Energy and Environment.
Proc. 9th European Biomass Conference, Ed. Chartier et al., Pergamon Press, Oxford, UK, pp.
582-587.
Dercas, N., G. Kavadakis and A. Nikolaou, 2000. Evaluation of productivity, water and radiation use
efficiency of two sweet sorghum varieties under Greek conditions. 1st World Conference on
Biomass for Energy and Industry, Sevilla, Spain, Ed. James & James (Science Publishers) Ltd.,
pp. 1654-1657.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

99
Dercas, N. and A. Liakatas, 1999. Sorghum water loss in relation to irrigation treatment. Water
Resources Management, 13:39-57.
Dercas, N., C. Panoutsou and A. Liakatas, 1998. Sorghum yield simulation according to irrigation
treatment. IFAC-CAEA 98, Control Applications and Ergonomics in Agriculture, Athens, pp. 179-
184.
Dercas, N., C. Panoutsou and C. Dalianis, 1996. Water and nitrogen effects on sweet sorghum
growth and productivity. Biomass for Energy and Environment. Proc. 9th European Biomass
Conference, Ed. Chartier et al., Pergamon Press, Oxford, UK, pp. 54-60.
Dercas, N., C. Panoutsou, C. Dalianis and Ch. Sooter, 1994. Sweet sorghum (Sorghum Bicolor (L.)
Moench) response to four irrigation and two nitrogen fertilization rates. Biomass for Energy,
Environment, Agriculture and Industry. Proc. 8th European Biomass Conference, Ed. Chartier et
al., Pergamon Press, Oxford, UK, pp. 629-639.
Fisarakis, I.K., 2002. Salinity effects on grapevine. Geotechnical Scientific Issues, No 4/2002, Volume
13, Issue I, pp.73-83.
Gertsis, A., A. Symeonakis, S. Varsami, S. Galanopoulou-Sendouca and G. Papathanasiou, 1998.
Evaluation of water use efficiency with GOSSYM, a simulation model of cotton growth and yield.
7 Hellenic Soil Science Symposium, Agrinio 27-30 May, 1998, pp. 99-108.
Giannoulopoulos P., S. Alexandris, M. Psychoyou and A. Poulovassilis, 2002. Salinization and quality
characteristics of Argolic plain groundwater. Proccedings of 6o Hellenic Hydrogeology symposium,
pp1-12.
Kerkides, P., C. Olympios and M. Psychoyou, 1998. The effect of irrigation with saline water on
greenhouse grown vegetables. 7 Hellenic Soil Science Symposium, Agrinio 27-30 May, 1998, pp.
120-131.
Koundouras, S. and C. van Leeuwen, 2002. Environmental effects on winegrape cv Agiorgitiko (Vitis
vinifera L.) grown in the Nemea region (Greece). 2. Vegetative growth, fruit ripening and wine
chemical and sensory characteristics. Geotechnical Scientific Issues, No 4/2002, Volume 13,
Issue I, pp.28-38.
Lazaridou, M.G. and V. Noitsakis, 2002. Effects of drought on seasonal changes of water-use
efficiency in a mixture crop of Medicago sativa. Geotechnical Scientific Issues, No 3/2002, Volume
13, Issue II, pp.59-66.
Louisakis, A.D., 1994. The water requirements and leaf quantity of tobacco in Greece. Aspects of
Applied Biology. Conference on Efficiency of water use in crop systems. Ed. M. C. Heath. Vol
38:209-216.
Mayropoulos, T.I., 2001. Saving irrigation water in scarcity of water conditions. Geotechnical Scientific
Issues No 2/2001, Volume 12, Issue II, pp.113-126.
Mimides T., S. Aggelides, S. Koutsomitros, M. Psychoyou, G. Kargas and A. Sgoumbopoulou, 1997.
Study of Iria valley aquifer of Argolis with emphasis to hydrochemical processes after the sea
intrusion. Proccedings of 4o Hellenic Hydrogeology symposium, Thessaloniki 11-14 November
1997..
Ministry of Agriculture, 1990. Experimental application of new irrigation methods. Athens, Greece.
Ministry of Agriculture, Land Reclamation Service, 1991. Technical guidelines for conducting
agrotechnical studies for land reclamation works. Athens, Greece.
Ministry of Development, 2002. Master Plan for the Greek Water Resources Management. Athens,
Greece.
National Statistical Service of Greece, 2001. Annual agricultural statistics survey for the year 2001.
Panayiotou, C and S. Aggelides, 1987. Influence of irrigation on the production characteristics of the
Cardinal vine variety. Agricultural Research, 11:393-401.
Panoras, A., A. Giannakaris, M. Dellios, S. Dimov and S. Eneva, 1997a. Water winter wheat yield
relations in Thessaloniki plain. Geotechnical Scientific Issues, Volume 8, Issue IV, pp.45-51.
Panoras, A., A. Giannakaris, M. Dellios, S. Dimov and S. Eneva, 1997b. Corn yield to response to
water in Thessaloniki plain. Hydrotechnica. Volume 7, pp.39-51.
Papamichail, D.M. and D. K. Papadimos, 1996. Efficiency evaluation of the SCS design equations for
border irrigation. Geotechnical Scientific Issues, Volume 7, Issue II, pp.45-57.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

100
Polychronides, M., S. Galanopoulou, S.Aggelides and N. Danalatos, 1998. The effect of irrigation and
fertilisation practice on cotton growth and development under greek conditions. Second
Conference of World Cotton Research, September 6-12, 1998, Athens, Greece.
Poulovassilis, A., M. Anadranistakis, A. Liakatas, S. Alexandris and P. kerkides, 2001. Semi-empirical
approach for estimating actual evapotranspiration in Greece. Agricultural Water Management,
51:143-152.
Poulovassilis, A., P. kerkides, S. Alexandris and S. Rizos, 1997. A contribution to the study of the
water and energy balances of an irrigated soil profile. A. Heat flux estimates. Soil & Tillage
Research, 45:189-198.
Poulovassilis, A., T. Mimides, A. Nikolopoulos, M. Psychoyou, N. Sgoupopoulou, P. Kerkides, S.
Alexandris, S. Aggelides, G. Kargas and P. Giannoulopoulos, 1994. Validity, limits and possible
trends of coastal south mediterrannean traditional groundwater irrigated agriculture. International
Conference on Land and Water Management in the Mediterranean Region. Bari, 4 - 8 Sept.
Poulovassilis, S. Aggelides, P. Kerkides, T. Mimides, M. Psychoyou, S. Alexandris, G. Cargas and A.
Sgoumbopoulou, 1993a. Soil salt accumulation in the valey of Iria- Peloponnese due to irrigation
with brackish groundwaters. First International Congress on the Environment. Abstracts
Geotechnical Chamber of Greece, Athens, March, 21 - 24.
Poulovassilis, A., S. Aggelides, P. Kerkides, T. Mimides, M. Psychoyou, S. Alexandris, G. Cargas and
A. Sgoumbopoulou, 1993b. Sea-water intrusion in the coastal aquifers of Iria Peloponnese due
to overpumping. First International Congress on the Environment. Abstracts Geotechnical
Chamber of Greece, Athens, March, 21 - 24.
Poulovassilis, A., P. Kerkides and S. Aggelides, 1991. Insufficiency of soil water - Crops. Conference
of Hellenic Soil Society: Drought - Productivity of Soil. Centre Helexpo, Thessalonica, 1991.
Sakellariou-Makrantonaki, M.S. and I.N. Vagenas, 2003. Crop water requirements in the prefecture of
Larissa. Hydrotechnika. Journal of the Hellenic Hydrotechnical Association. Vol.13:13-28.
Spanomitsios, G.K. and G. Paraskevopoulou-Parousi, 2001. Strawberry transpiration: comparison of
measurements with the output of two mathematical estimation methods. Geotechnical Scientific
Issues No 2/2001, Volume 12, Issue I, pp.83-93.
Tsanis, K., P.A. Londra and A.N. Angelakis, 1996. Assesment of Water Needs for Irrigation in the
Island of Crete. 2nd International Symposium on Irrigation of Horticulture Crops, 8-13 September,
1996, pp.41-48.





















Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

101
















































Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

102









IRRIGATED AGRICULTURE
AND WATER USE EFFICIENCY IN ITALY



M. Todorovic*, A. Caliandro** and R. Albrizio*
*Mediterranean Agronomic Institute of Bari, CIHEAM-IAMB, Italy
**University of Bari, Faculty of Agriculture, Italy



SUMMARY This document aims to provide a comprehensive review of irrigated agriculture in Italy
with a particular emphasis on the Water Use Efficiency (WUE) and water productivity (WP). The data
presented include the country climatic characterization, water availability and withdrawal, sources of
irrigation water, irrigable and irrigated lands, irrigation methods, irrigated crops and their growing
parameters, experimental data on both biomass and yield WUE. The agronomic data of main crops,
grown principally in Southern Italy environment, are taken into consideration including tree crops
(grapevine, olives, citrus, peach, etc.), field crops (wheat, maize, sugarbeet, sunflower, etc.) and
horticultural crops (tomato, potato, watermelon, beans, spinach, etc.). An analysis of: crop production
functions, application of irrigation methods, and crop water requirements (in terms of estimation of
reference evapotranspiration and crop coefficients), in Italian agriculture is given. The presented crop
coefficient data vary for a crop in respect to local climatic conditions, latitude, altitude, time of sowing
and applied agronomic practices. Moreover, these data differ notable from those presented in
scientific literature: it indicates a necessity for a local calibration and eventual revision of well-known
existing FAO documents on crop water requirements and crop response to water. Finally, some
common agronomic practices for enhancing WUE & WP have been described, focusing mainly on the
situation of Southern Italy. This analysis, based on the evaluation of the national scientific literature
and technical reports, has shown how these strategies should aim at increase of beneficial water
consumption (transpiration) against the non-beneficial losses by: (i) increasing of marketable yield per
unit of water transpired; (ii) maximizing transpiration consumption relative to evaporation losses; (iii)
enhancing effective use of rainfall and water stored in the soil.

Key words: water resources, irrigation, water use efficiency, water saving, Italy.



INTRODUCTION


Italy, with a surface area of 301,277 km
2
, occupies a central location in the Mediterranean basin.
Stretching over 1,200 km between North and South, Italy has shores on four Mediterranean sees (the
Ligurian, the Tyrrhenian, the Jonian and the Adriatic) and it has an exceptionally long coastline of
almost 7,500 km. About 27% of Italian territory (8,136,207 ha) is along the coast line and 73%
(21,997,893 ha) is considered the inland.

The Italian territory can be subdivided naturally into four main physiographic regions:
6. the Alps mountains chain in the North, extending from west to east and reaching up to 4,800
m a.s.l. (with Monte Bianco, the highest peak of Europe);
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

103
7. the lowland of the Po river basin, located on the South of the Alps;
8. the peninsula, including the central Apennine massive with the peaks rising up to 2,900 m
a.s.l. and the coastline, and
9. two large islands, Sicily on the South and Sardinia on the West of the peninsula

The lowlands, flat and valley areas, cover 6,976,373 ha (23.2% of the territory); the mountain
areas occupy 10,611,957 ha (35.25% of the country), while the hill areas cover about 12,542,779 ha
(41.55% of the territory).

The precipitations in Italy are relatively abundant (on average about 1,000 mm/year), but as often,
they are not evenly distributed between seasons and regions, and high evapo-transpiration in coastal
areas causes significant losses. Due to the range of rainfall, hydrological and climatic regimes (from
Mediterranean to continental and Alpine), Italy presents a wide diversity of ecosystems, landscapes
and agricultural practices. In fact, Italy's agriculture is a typical example of the division between the
agricultures of the northern and southern European countries: the northern part produces primarily
grains, sugar-beet, soybeans, meat, and dairy products, while the south is specialized in producing
fruits, vegetables, olive oil, wine, and durum wheat.

Inasmuch as Italian agriculture is very intensive and market oriented it preserves many local
peculiarities especially in the Southern regions. In fact, most farms are small, with an average size of
only 7 ha whereas a large working force (more than 1.5 million) is employed. Irrigation represents a
common practices in all parts of the country due to market oriented agricultural production and strong
variability and uncertainty of climatic factors. However, the cropping pattern, irrigation methods,
agronomic practices and water use efficiency vary significantly from region to region and, also, from
farm to farm. This paper reports the data describing the irrigated agriculture, crop water requirements
and water use efficiency in Italy emphasizing the practices that improve the efficiency of water use
and save water for other purposes.


CLIMATIC CHARACTERIZATION


The Italian climate is highly varied due to variety of hydrographic and orographic factors, its North-
South elongation and exposition to four Mediterranean seas. These factors influence substantial
variation of the main climatic variables as illustrated in Figures 1 and 2. The average temperature in
January (the coldest month) varies from several degrees below zero in the Alpine area to more than
6C in the coastal Mediterranean regions while the average temperature in July (the warmest month)
spans between less than 15C in the Northern Alps and about 30C in the Southern Mediterranean
zones (Fig. 1).

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

104

Fig. 1. Spatial distribution of temperature of the coldest month (on the left) and of the warmest month
(on the right) based 30 years averages (Source: SAIN-UCEA, Rome, 1995)

The sunshine hours cumulated on annual basis ranges between less than 1800 in the Alps to
more than 2200 in the South (Fig 2). The average annual precipitation is relatively abundant and it is
estimated to about 1000 mm per year although it is unevenly distributed among regions and seasons.
In fact, average annual precipitation goes from less than 400 mm in the coastal Southern zones,
receiving almost all precipitation input during the winter season (between October and March), and to
almost 3000 mm in the Northern Alpine areas (Fig. 2). The Southern Adriatic regions receive much
less precipitation than the Tyrrhenian side due to the characteristic movements of the humid air
masses and orographic characteristics of the peninsula. According to the above mentioned
parameters and the Kppen climatic classification, the overall Italian climate can be described as
moist, mid-latitude subtropical although eight climatic zones can be observed moving from the
Northern Alps regions to the South and from the coastal areas to the inner Apennine massive as
illustrated in Figure 3.
Fig. 2. Spatial distribution of sunshine hours per year (on the left) and of the total annual precipitations
(on the right) based 30 years averages (Source: SAIN-UCEA, Rome, 1995)
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

105


Fig. 3. The main climatic zones of Italy according to the Kppen climatic classification (Source:
www.italocorotondo.it/tequila/partner_section/italy_english/)

The coastal zones of Italy are characterized by dry semi-arid Mediterranean climate which passes
to sub-littoral and sub-continental as moving into the inner areas of Apennines. The central Northern
regions, including the Po river valley (the most important Italian river basin where live about 15.5
million inhabitants), are characterized by sub-continental climatic conditions while the climate of the
Alpine mountains goes from cool temperate to cold polar. Some high peaks of Apennines are also
characterized by cool temperate climatic conditions.


WATER RESOURCES AVAILABILITY AND USE IN ITALY

Water resources availability

The analysis of water resources availability in Italy is based on the data coming from several
sources (ANPA, 2001; IRSA CNR, 1999; EUOSTAT, 1998; AQUASTAT, 1998, Blue Plan, 2001) and
a synthesis of results is presented in Table 1.

The precipitation over the Italian territory generates every year a total flow of about 296 km
3
.
However, due to the presence of large areas characterized by semi-arid Mediterranean climate, the
evapotranspiration losses are estimated to 129 km
3
/year while the subsurface flow to the sea is in
average of about 12 km
3
/year. This means that the internal renewable water resources account to
approximately 155 km
3
/year which represents about 52.3% of total flow generated by precipitation.
External runoff is calculated to 7.6 km
3
/year (from Switzerland 51%, from Slovenia 43% and from
France 6%) while spring outflow contribution from local aquifers is estimated to about 3.5 km
3
/year.
This means that the total renewable water resources of Italy are about 166.1 km
3
/year. It is estimated
that only two/thirds of that volume (or about 110 km
3
/year) are technically and economically available
for exploitation.

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

106
The total groundwater availability is about 40 km
3
/year but the greatest part of it (about 30
km
3
/year or 75%) contributes to the recharge of regional aquifers and only 25% (10 km
3
/year)
represents the recharge of local aquifers. Only one/third of it (about 3.5 km
3
/year) is related to the
spring outflow as mentioned previously.

Table 1. A synthesis of water resources availability in Italy (data elaborated from the following
sources: ANPA, 2001; IRSA CNR, 1999; EUOSTAT, 1998; AQUASTAT, 1998, Blue Plan, 2001)
Average precipitation [mm/year] 982
Flow generated by average precipitation [km
3
/year] 296
Average evaporation [mm/year] 428 (438
*
)
Evaporation losses [km
3
/year] 129 (132
*
)
Subsurface flow to the sea [km
3
/year] 12 (9
*
)
Internal renewable water resources [km
3
/year] 155 (=296-129-12)
External runoff inflow from other countries
[km
3
/year]
7.6
Total groundwater availability [km
3
/year] 40
Groundwater recharge of local aquifers [km
3
/year] 10 to12
Spring outflow from local aquifers [km
3
/year] 3.5
Total renewable water resources [km
3
/year] 166.1 (=155+7.6+3.5)
Potentially usable water resources [km
3
/year] 110
*
there is some difference between data coming from different sources


Therefore, the total renewable water resources availability per person can be estimated as about
2914 m
3
/year/capita, or 1930 m
3
/year/capita by means of potentially usable resources. These values
are much more greater than those of the Southern Mediterranean countries (e.g. total renewable
water resources availability in Middle East and North Africa Region is about 1250 m
3
/year/capita, or
about 43% of Italian availability). However, they are significantly lower than the average renewable
water resources of Western Europe countries, which is estimated to about 5183 m
3
/year/capita
(World Resources Institute, 2000).

Furthermore, it is important to highlight that water resources are not regularly distributed over the
Italian territory (Fig. 4): in the Northern part is located about 59.1% of potentially usable water
resources whereas the rest of the country accounts on the 40.9% of resources. This disparity
becomes even more evident when expressed by the availability of potentially usable resources per
capita (Fig. 4b) which indicates that water resources availability per capita in the North is almost 3.5
times greater than in the Islands and it represents about 175% of water availability in the continental
Southern regions. These data emphasize the seriousness of water problems in the Southern regions
especially during the summer months when in those areas water demand is strongly increased due to
important vocation to tourism and consequent high population inflow.

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

107
Fig. 4. Regional distribution of potentially usable water resources in Italy as a percentage of total
resources (a) and as water availability in m
3
/year/capita (b) (Source: IRSA CNR, 1999)


Water withdrawal and sectorial water use

The average water withdrawal in Italy is estimated to about 51.820 km
3
/year which represents
about 31% of the gross annual available water resources and 47% of the water resources technically
and economically available for exploitation. (IRSA, CNR, 1999). This amount, translated to a mean
annual per capita withdrawal of 910 m
3
, is significantly greater than EU average of 662 m
3
/capita/year
and it is, together with Egyptian water withdrawal per capita (however, Egypt uses 100% of
exploitable resources), the greatest in the Mediterranean region. Nonetheless, it is significantly lower
than in some other highly developed countries (e.g. USA - 1873 m
3
/capita/year and Canada - 1736
m
3
/capita/year).
The greatest part of water withdrawal belongs to surface water resources (39.673 km
3
/year or
76.6%) which includes the storage capacity of artificial accumulation reservoirs of about 8.426 km
3
.
The contribution of groundwater is estimated to about 12.147 km
3
/year, which corresponds to 23.4%
of total water withdrawal. Nevertheless, it is important to underline that the knowledge about
groundwater resources is far from accurate due to frequent non-authorized water abstraction for
irrigation especially in the Southern regions.

The water withdrawal varies from year to year between 40 and 56 km
3
/year according to the
availability and demand, and also, it is very variable from region to region. In general, about 65% of
withdrawal belongs to the Northern part of the country, 15% to the Central regions and 25% to the
South and the Inlands. Water withdrawal is the highest in the North-East region of 1975
m
3
/capita/year (even greater than in the USA and Canada), and it the lowest in the Apulia region (220
m
3
/capita/year). In some regions, water shortage is attenuated with the water transfer from other
regions, as it is the case of the Puglia region, which receives more than half of its water demand from
Basilicata region and partially from Campania region. This was possible thanking to the CASSA PER
IL MEZZOGIORNO (Southern Italy Development Fund), promoted and implemented by Italian
North
59.1% Center
18.2%
South
18.2%
Islands
4.5%
2542
1834
1451
1930
743
0
500
1000
1500
2000
2500
3000
North Center South Islands Italy
a)
b)
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

108
authorities during the 50-ties, 60-ties and 70-ties of the last Century. The realization of new
accumulations and water delivery systems is still in progress and, together with an inter-regional
action program for management of common water resources, represents the keystone of strategies
for facing water shortage problems in the South.

The partitioning of water withdrawal between different sectors changes from year to year (Table
2) depending on the overall availability and water demand. Nonetheless, on the basis of average
historical data, can be stated that, in general, about 60 per cent of water withdrawal is used for
irrigation, 25 per cent for industry, and 15 per cent for domestic use (Fig. 5). Certainly, when water
availability is scarce, the reduction is applied primarily to irrigation sector as illustrated in Table 2.

Table 2. Sectorial water use in Italy for a hydrological normal (1991) and a dry (1999) year (Source:
ISTAT, 1991; IRSA-CNR, 1999; MPAF, 2004)
1991 (a normal year) 1999 (a dry year) Sectors of water use
Water use
[km
3
]
Water use [%] Water use
[km
3
]
Water use
[%]
Domestic 8 16 7.9 19.6
Industry 12 24 8 19.7
Energy
*
- - 4.5 11.1
Agriculture 30 60 20.1 49.6
Total 50 100 40.6 100
*
includes only the use of freshwater for thermoelectric plant cooling
Fig. 5. Water withdrawal by sectors in Italy
The use of water for irrigation is not regularly distributed all over the country: 67% of it belongs to
the Northern Italy, 28% to Southern Italy with islands and only 5% to the Central part of the country.
Main sources of irrigation water are rivers (67%), followed by groundwater from wells (27%) and by
reservoirs (6%). The water withdrawal for domestic purposes reaches almost 370 liters/person/year
and it is obtained mainly from groundwater aquifers (50%) and springs (40%) and only marginally
from surface water (10%). Groundwater withdrawals in the Po Basin are considered to have reached
their maximum, with over-exploitation already occurring in some sub-basins (e.g. Lambro-Sveso-
Olana, Parma, Panaro rivers).


IRRIGATED LAND AND IRRIGATION PRACTICES IN ITALY

Agricultural, irrigable and irrigated land

The total utilized agricultural area (UAA) in Italy is estimated to about 131,941 km
2
which
corresponds to 43.8% of total surface area. The agricultural land area is continuously decreasing:
since 1970 the utilized agricultural area diminished by 2.8 million hectares (-16%) according to data
from the most recent survey of farm structures. This is a phenomenon which affects all developed and
Domestic
15%
Industry
25%
Agriculture
60%
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

109
industrialized countries. Between 1991 and 2001, the utilized agricultural area has decreased
progressively by 11.1% per inhabitant, from 0.3 to 0.26 hectares per capita (INEA, 2003), which is in
the range of other EU countries (-10.9%). Land is, thus, becoming an ever-more precious resource,
especially in countries which, like Italy, have a high population density and where national territory is
subject to considerable variation in altitude.

A synthesis of land use in Italy and other European countries is given in Table 3 as a percentage
of total surface area. In general, the data indicate an intensive use of land in Italy and a substantial
difference in respect to other EU Mediterranean countries and to the EU territory. This is probably due
to the fact that Italian territory is exposed to many very different climatic zones (eight), from cold polar
to subtropical, which caused a strong variation in land use. Approximately 37% of the Italian territory
is used for arable agriculture, which is much more greater than EU average of 27%. Nevertheless,
due to many arid and semi-arid zones, the percentage of bare ground is two times greater than the
EU average. Moreover, the urban, unproductive areas, cover about 2.1 million hectares which is 7%
of the country, while the EU average is 5% and average of other EU Mediterranean countries is 4%.

Table 3. Land use in Italy and EU countries (%) (Source: INEA, 2003, on the basis of EUROSTAT
survey)


Italy
Other EU
Mediterranean
countries
(*)

Central EU
countries
(**)


Total EU
(***)

Arable land
(1)
37 33 32 27
Permanent crops
(2)
29 26 32 37
Moorland (areas over 20% covered by shrubs) 8 20 4 8
Permanent meadows and pastures 10 11 20 12
Bare ground 6 5 3 3
Inland waterways and wetlands
(3)
3 1 3 8
Unproductive areas and other land
(4)
7 4 6 5

TOTAL AREA (000 ha) 30,133 72,988 110,172 292,105
(*)
Greece, Spain, Portugal.

(**)
France, Germany, Belgium, Luxemburg, Denmark, The Netherlands,
(***)
Excluding UK and Ireland,
(1)
Including temporary forage crops and set aside.
(2)
Tree and other permanent crops (woods and forests).

(3)
Including glaciers and eternal snow.

(4)
Man-made and industrial settlements, infrastructure, rocks and barren land; ornamental parks and gardens,
roads, railways, etc.


According to the General Agriculture Census carried out in 2000, the irrigable land amounts to
3,887,387 hectares which is equivalent to 29% of total national utilized agricultural area (UAA). A
comparison with the 1990 Census, indicates that irrigable land has remained almost the same
although it varies considerable from region to region (Table 4). The Northern regions, endowed with
significantly greater water resources than Central and Southern regions, could potentially irrigate
about 50% of their UAA.
The average irrigated area is estimated to approximately 2.65 million hectares which
corresponds to 68% of the total irrigable land and to about 20% of UAA. According to the Census, the
irrigated area in 2000 was slightly smaller (2.47 million hectares), with substantial differences
between the regions (Table 4). Slightly less than two thirds of the irrigated area is in the Northern
Regions, involving 34.9% of farms with UAA and with an average area per farm of 6.5 hectares. In the
Centre, only about 17.9% of farms are irrigated, whereas in the South the practice is carried out on
25% of farms with a total area of 758 thousand hectares, equivalent on average to 2.2 hectares per
farm. According to the official data of National Institute for Statistics (ISTAT), about 63.2% of irrigated
land in located in the North, 7.2% is in the Central part of the country while 29.6% is situated in the
South.

There are two main factors limiting irrigation in Italy: the availability of water resources and the
presence of infrastructures for water accumulation and delivery to the fields. Accordingly, the largest
irrigated areas are located in Lombardia Region, covering about 554,382 ha and corresponding to
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

110
almost 80% of UAA. Then, irrigation is fully developed in Piemonte Region (on 335,800 ha), Veneto
(265,253 ha), Emilia Romagna (252,377 ha), and Puglia Region (248,814 ha). Nonetheless, it is
necessary to recognize a drawbacks of official statistics which have difficulties to consider the farms,
located mainly in the South, subjected to non-authorized irrigation from private wells.

Table 4. Irrigable land in Italy and area irrigated in 2000 (Source: ISTAT, 2002)
Region Irrigable land
[ha]
Irrigated land
[ha]
Irrigated/Irrigable
[%]
Piemonte 448,947 335,800 79.25
Valle dAosta 26,212 23,623 90.12
Lombardia 700,140 554,382 79.18
Liguria 11,244 7,191 63.96
Trentino Alto Adige 61,774 57,768 93.51
Veneto 435,845 265,253 60.86
Friuli Venezia
Giulia
91,876 63,202 68.79
Emilia Romagna 565,573 252,377 44.62
Toscana 111,603 47,286 42.37
Umbria 66,927 32,117 47.99
Marche 49,470 25,070 50.68
Lazio 150,088 74,052 49.34
Abruzzo 59,358 29,995 50.53
Molise 20,881 11,812 56.57
Campania 125,305 86,414 68.96
Puglia 389,617 248,814 63.86
Basilicata 80,640 42,325 52.49
Calabria 117,143 66,922 57.13
Sicilia 209,036 161,044 77.04
Sardegna 165,709 62,315 37.6
ITALY 3,887,387 2,467,763 63.48


Irrigated crops

The Census on agriculture, referring to the year 2000, provides the data about irrigated crops in
Italy and a synthesis of elaborations is given in Figure 6. The data indicate that almost 86% of
cultivated citrus crops were irrigated (corresponding to 113,600 ha in respect to total cultivated area
of 132,500 hectares). Then, the irrigation was very intensive in the areas cultivated with vegetables
(70%), potato (67.4%) and maize (58%), followed by fruit-tree crops (38%), sugarbeet (36.2%), soya
(34.5%), vineyards (25.5%), etc.
The maize is the crop which is irrigated on the greatest surface areas in Italy, i.e. on 622,000 ha,
mainly located in the North-West regions. Then, large irrigated areas are covered by forage crops
(267,000 ha), vegetables and potato (217,000 ha), fruit-tree crops (189,000 ha), vineyards (183,000
ha), sugarbeet (81,000 ha), etc. Inasmuch as the cereal cultivation covers the greatest part of UAA
(2,233,00 ha), the cereal crops are irrigated on 99,500 ha which represents only 4.5% of their total
cultivation.

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

111
85.8
36.2
17.5
4.5
38.0
6.7
58.2
70.0
67.4
34.5
25.5
10.0
0
10
20
30
40
50
60
70
80
90
c
i
t
r
u
s

c
r
o
p
s
s
u
g
a
r
b
e
e
t
f
o
r
a
g
e

c
r
o
p
s
c
e
r
e
a
l
s
f
r
u
i
t

t
r
e
e
s
s
u
n
f
l
o
w
e
r
m
a
i
z
e
v
e
g
e
t
a
b
l
e
s
p
o
t
a
t
o
s
o
i
a
v
i
n
e
y
a
r
d
s
o
t
h
e
r

c
u
l
t
i
v
a
t
i
o
n
s

Fig. 6. Irrigated crops in Italy (as percentage of total cultivated area of each crop) according to the
Census in 2000 (Source: ISTAT, 2002)

The citrus crops are almost fully irrigated (up to 95%) in the Southern regions, especially in Sicily
and Basilicata. The fruit-tree crops are irrigated almost completely in Trentino Alto Adige (93%), while
the percentage is lower in other regions: 72% in Veneto, Friuli Venezia Giulia and Basilicata and 61%
in Emilia Romagna. Sugabeet is irrigated principally in Trentino Alto Adige (96%), Sardegna (83%),
Campania (83%) and Umbria (81%). Vineyards are irrigated particularly in Trentino Alto Adige (67%),
Puglia (62%) and Valle dAosta (54%). The irrigation practices are strongly related to the availability of
water resources, especially in the South, where the irrigation strategies and irrigated crops are
selected on the basis of economic parameters and increase of profit. In fact, the irrigated area for the
most crops, except maize and vineyards, has decreased substantially in respect to the census in
1990. The most significant decrease of irrigated land was observed for soya and forage crops, of
about 123,000 ha (60%) and 172,000 ha (40%) respectively. On the other side, an increase of
irrigated land was observed for maize, of about 115,000 ha (23%) and for vineyards, of about 20,000
ha (13%). The irrigated land in 2000 was for about 100,000 ha lower than in 1990.


Irrigation methods

The irrigation methods vary in respect to the irrigated crops, quantity and quality of available
water, size and type of management of irrigated farms, and soil and climatic characteristics. In
general, the sprinkler irrigation method is the most utilized (on 1,047,000 ha), followed by surface and
furrow irrigation (850,480 ha), localized irrigation (366,018 ha) - mainly drippers (290,700 ha), flooding
irrigation (202,000 ha) and other methods (2,300 ha) as illustrated in Fig. 7. During the last twenty
years, there is a general trend of almost all irrigation methods, except localized irrigation, to shrink the
surface area of application.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

112

Fig. 7. Irrigation methods in Italy
The sprinkler irrigation method is frequently used in Emilia Romagna (162,500 ha), in Veneto
(157,500 ha) and in Lombardia (138,500 ha) where field crops as maize, forage crops, sugarbeet, etc
are cultivated.(Fig. 8). The surface and furrow irrigation, characterized by low application efficiency,
high volumes of water supply, well-managed and dense water distribution networks and well-leveled
irrigation fields, are extended mainly for irrigation of herbaceous crops in Lombardia (350,000 ha),
Piemonte (211,500 ha), Veneto (86,000 ha) and Emilia Romagna (45,000 ha). The furrow irrigation
method is utilized also in Campania, on the surface area of 40,000 ha, for irrigation of vegetables. In
this case, short furrows (about 10 m length) with the water flow between 5 and 10 l/s are utilized,
realizing in such a way a sort of flooding by furrows.


Fig. 8. The most utilized irrigation methods in the Italian regions with the highest irrigation surfaces
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

113
Moreover, in Lombardia, where the winter temperatures are very low and frequently below zero,
the surface irrigation is utilized with anti-frost purposes on permanent forage crops in order to have
green forage during the winter season. For this purpose it is necessary to provide an appropriate field
land leveling which permits fast flow of water in the normal direction to the longitudinal axis of
irrigation units. In these cases, the irrigation is performed by using a single or double lateral land
grading (Fig. 9). The slope of land along the axis perpendicular to the longitudinal irrigation unit is 4 to
10% and the length of water course is between 5 and 20 m. In such a way, the time of flow-off is
lower than the time necessary for the conversion of water from liquid to solid state, allowing the
superficial soil layers to have temperature greater than zero and to permit the growing of vegetation
having green forage also during the winter time.

Fig. 9. Surface irrigation method with double (a) and single (b) lateral land grading (Source: Giardini,
2002)
The flooding irrigation method is utilized almost exclusively to irrigate rise, in Piemonte on the
surface area of more than 110,000 ha, in Lombardia on the surface area of about 89,500 ha, and in
Veneto, Emilia-Romagna, Sardegna and Calabria on a total surface area of about 15,000 ha (Fig. 8).

The localized low-pressure irrigation methods (drip, sprayers and capillary sub-irrigation) are
extended mainly in the Southern regions of Italy, and particularly in Puglia (143,000 ha) and in Sicilia
(62,000 ha) while in the North they are utilized prevalently in Emilia-Romagna (38,000 ha). These
methods guarantee a high water application efficiency and they are used mainly for the irrigation of
orchards and vegetables in the areas where water supply is limited.

The sub-irrigation method by regulation of water table depth is used in Veneto, in the areas
where shallow water table is controlled by sub-surface drainage systems, and it is applied as a
supplementary intervention to rise water table when necessary. The capillary subsurface irrigation is
practiced on orchards in Emilia-Romagna, Puglia, Sicilia and Basilicata, burying the dripping laterals
with drippers that release slowly herbicides (Trifluralin) to avoid intrusion of roots into drippers.

Sprinkler irrigation is realized mainly with self-propelled devices which use side-roll laterals with
long jets (sprinklers) which can be substituted sometimes with sprinkling laterals in order to improve
water use efficiency and to reduce the working pressure of the system. These equipments have been
widely used by farmers for irrigation of field crops due to their capacity to adapt at different field
conditions, to move easily, to limit labor requirements and application cost. Recently, there is an
attempt to improve distribution efficiency of the high pressure sprinklers with large wetted diameter in
windy areas through the application of new generation turbine sprinklers with slow return fluctuating
arms and with adjustable angle of the jet until reaching the horizontal position (Fig. 10)

a) b)
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

114

Fig. 10. New generation turbine sprinklers with slow return fluctuating arms and adjustable angle of jet

The devices with mobile and fixed wings (lateral sprinklers) are presented rarely for irrigation of
vegetables while permanents irrigation devices are used prevalently for irrigation of orchards. The
irrigation devices like rangers and center pivots are not frequently used due to small size of farms
and presence of obstacles in the field (trenches, windbreaks, electrical cables, etc.). Surface irrigation
is applied provided that land leveling was done with adequate furrow distances and sometimes by
open ditches 20-30 m far away. This type of lateral infiltration irrigation is used in soils which crack
superficially and water can run laterally over long distances.


Crop yield response to irrigation water

A more significant development of irrigation techniques in Italy coincides with the general
reconstruction of country after the World War 2
nd
. It was particularly relevant in the Southern parts of
the country, where the water shortage problems imposed the construction of dams and water
accumulation lakes. At the same time, an intensive research in the field of irrigation had been
promoted by the National Research Council (Consiglio Nazionale delle Ricerche - CNR). In 1962-63,
these activities resulted in the constitution of a Group for Irrigation Studies (Gruppo di Studio
sullIrrigazione GRU.S.I.) which has been operated up-to-date in an informal way. At the beginning,
GRU.S.I. conducted research prevalently on the yield response to irrigation of herbaceous and tree
crops with the aims to evaluate crop water requirements from the agronomic point of view and to
optimize both the quantitative and qualitative aspects of crop production under different Italian
environments. In fact, it is well known that optimal agronomic crop irrigation requirements do not
coincide with the maximum evapotranspiration.

The research activities on irrigation have been conducted mainly in Southern Italy where the crop
productivity is strongly influenced with limited precipitation, and irrigation represents a fundamental
practice in order to increase and stabilize agricultural production over the years. These researches
have been conducted prevalently on vegetables and field crops (tomato, pepper, bean, sugar-beet,
maize, sorghum, etc.) and, also on the olive trees and vineyards since they are well-adaptable to
water stress conditions.

The results of numerous experimental works highlighted that the seasonal irrigation volume
represents the most important irrigation parameter in the determination of the production of crops
under specific environmental conditions. Accordingly, the crop responses to water are presented in
this document as the variation of yield, expressed as a percentage of the maximum obtainable yield,
in relation to the specific seasonal volume of irrigation. In order to make possible a comparison
between the crop productivity of different cultivars in different years and under different environmental
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

115
conditions, the specific seasonal irrigation volume is expressed as a percentage of the maximum crop
evapotranspiration (ETc).

In most of the experimental works on the crop response to water, the irrigation events have been
programmed using the soil water balance approach with the reference to the maximum crop
evapotranspiration (ETc), estimated with different methodologies and with the crop coefficient values
(Kc) adopted from the literature or defined for the study areas. The methods based on the monitoring
of the soil water content and/or the plant water status have been rarely adopted in the past. In
general, different irrigation strategies have been compared maintaining fixed the irrigation intervals
and changing the volumes of water applied as a percentage of the optimum water supply
corresponding to the 100% of crop evapotranspiration.

An example is given for some herbaceous crops in Figure 11 showing the relations between the
crop yield, expressed as a percentage of the maximum yield obtained during the experimental period,
and the specific seasonal irrigation volume, expressed as a percentage of ETc, obtained in
Metapontino (Policoro, Southern Italy). The relationships reported in Figure 11 are obtained adapting
to the experimental points the Mitcherlich model modified by Giardini and Borin (1985) as:
( )
| |
( )
| |
( )
| |
d b c 1 d b k d b c
m
10 1 10 10 1 y y
2
+
+ =

where: y is crop yield; y
m
is the maximum obtainable crop yield under non-limited supply of the
factor (parameter) under study; c is a coefficient of action (or of increase), indicating the rapidity of the
achievement of the maximum yield; k is a coefficient of depression, indicating the tendency of y to
decrease after the achievement of the maximum value; b is the quantity of the factor under study
available for the crop in natural conditions, and d is the quantity of the factor under study applied
under specific experimental conditions.
Yield
(% of the
maximum)
ETc (%)

Fig. 11. Trend of some herbaceous crops yield expressed as percentage of the maximum obtainable
yield in relation to the seasonal irrigation volumes expressed as percentage of ETc. The
curves have been obtained adapting to the experimental points Mitscherlich model modified
by Giardini and Borin. Negative values indicate the quantities of natural water, from
precipitation, groundwater table and soil water content, utilized by the crops (adapted after
Venezian Scarascia et al., 1987).

The research work was carried out in a deep, silty clay loam soil with moisture levels at field
capacity and wilting point equal to 31.5 and 15% of dry soil weight, respectively; the water table was
between 150 and 200 cm below the ground surface during the rainy season and in the dry months,
respectively. The climate is typically Mediterranean with 600 mm average annual rainfall and 16C as
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

116
mean annual temperature. As an average, 79% of rainfall occurs in autumn-winter season (from
October to March) while the highest averages temperatures (between 22 and 25C) are recorded in
June, July and August; consequently, the dry period extends from the beginning of May to the end of
August.
The crops considered in the study were autumnal and spring-sown sugar-beet, shell bean and
dry bean, tomato, pepper, eggplant, spring and summer-sown grain maize. Utilization of natural water
resources (rainfall and ground water) by crops increased as the cycle extended into the rainy season.
In this regard, figure 11 shows that the amount of natural water actually used by summer-cycle crops
(shell bean, sown in June) is only about 2-3% of ETc and rises to as much as 50% with crops sown in
autumn and harvested in summer (autumnal sown sugar beet): the corresponding water volumes are
150-180 and 3000 m
3
ha
-1
, respectively. Moreover, the yield irrigation water efficiency is much greater
for crops whose cycle extends - at least in part - into the rainy period (maize grown as the main crop,
eggplant, sugar beet whether sown in spring or in autumn) than far spring-summer, or summer crops
(pepper, tomato, maize grown as cash crop); for the first group of crops indeed the curves are
steeper, as compared to the second group, because of the higher values of the action coefficient (c)
which means better water use efficiency (Fig. 11 and Table 5).

Table 5. The parameters of the Mitscherlich equation, flex point coordinates and seasonal irrigation
volumes at 100% of yield
Equation-parameters Flex point coordinates
Ym b c Water volume Yield
Seasonal
irrigation volume
at 100% of

(% of the
max yield)
(% of the
ETc)
(ha /
%ETc10
-3
)
% of
ETc
m
3
/ha
(% of the
max)
(m
3
/ha)
Tomato 78.5 13.9 23.0 29.6 1435 35.3 4734
Pepper 84.9 15.1 20.1 34.6 2004 38.2 5800
Spring maize 91.0 15.4 27.6 20.8 994 41.0 4181
Summer maize 94.9 10.0 16.4 51.0 1603 42.7 3085
Eggplant 82.9 23.1 30.4 9.8 441 37.3 4795
Shell-bean 96.9 2.2 15.8 61.1 1898 43.6 3109
Dry-bean 102.0 4.9 14.9 62.1 2121 45.9 3413
Spring sugar beet 75.9 39.1 32.7 -8.6 - 34.2 7193
Autumnal sugar beet 82.0 49.4 23.1 -6.2 - 36.9 4961

Consequently, the greatest increments in yield were recorded with seasonal irrigation volumes
around 61-62% of ETc (1898-2191 m
3
ha
-1
) for bean (a typically summer crop) between 34 and
20.8% of ETc (2004-994 m
3
ha
-1
) for pepper and maize grown as the main crop (spring-summer cycle
crops), and without irrigation for sugar beet (grown either as spring or autumnal crop): the yields
corresponding to such maximal increments were respectively 43.6, 45.9, 38.2, 41.0, 34.2 and 36.9%
of peak yields recorded during the trial period (Fig. 12 and Table 5 to compare the flex point
coordinates of the curves: the amounts of water and the corresponding yields).

These results stress the fact that yields are less affected by irrigation when dealing with spring-
summer and autumn-summer crops, than with summer crops. Fig. 13 shows indeed that to obtain as
much as 70% of the yield recorded during the trial period the seasonal amount of irrigation water had
to be as high as 75% of the calculated ETc for summer and spring-summer crops and about 25% of
the calculated ETc for autumn-summer or winter-summer crops. Seasonal irrigation volumes
corresponding to 100% of estimated ETc ranged from minimum of 3100 m
3
ha
-1
to a maximum of
7200 m
3
ha
-1
according to the length of the crop cycle and the season of the year during which the
crop cycle develops. The lowest seasonal irrigation volumes were recorded for very short cycle crops
(72 days) including summer crops like shell bean - and also for those crops which crop cycles
develop during seasons with a low evaporative demand of the atmosphere, as it happens in the case
of maize grown as a forage crop. Conversely, the heaviest seasonal volumes were recorded for
longer-cycle crops (more than 150 days) growing during the months when the evaporative demand
increases, such as spring sown sugar beet (Fig. 14).

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

117
Yield
(% of the
maximum)
ETc (%)

Fig. 12. Yield of some herbaceous crops as a function of seasonal water volumes expressed as
percentage of estimated ETc with the indication of the flex points of different curves (adapted
after Venezian Scarascia et al., 1987).


Yield
(% of the
maximum)
ETc (%)

Fig. 13. Yield of some herbaceous crops in relation to the seasonal irrigation volumes, with the
indication of the seasonal irrigation volumes corresponding to the 70% of the maximum yield
(adapted after Venezian Scarascia et al., 1987).
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

118

Fig. 14. Seasonal irrigation volumes of several horticultural crops in relation to the adopted irrigation
regime (adapted after Venezian Scarascia et al., 1987).

In conclusion, a very short cycle of crops makes the best use of irrigation water, irrespective of
the season of their growth cycle. Similar inference can be drown for crops sown in autumn or early in
spring as they make a good use of natural available water resources. When irrigation water is limited
and crops that respond rapidly to irrigation (such as sugar beet and maize grown as main crop) are
grown simultaneously to crops that respond gradually (such as pepper, tomato, maize grown as
forage crop and shell bean), then, the latter group of crops should be irrigated more than the former.


Deficit irrigation strategies

A particular attention has been given to the studies on regulated water stress based on different
crop sensitivity to water supply during various phenological stages and on the crop physiological
mechanisms of response to water stress. Deficit irrigation techniques have demonstrated a high
validity for water saving in the case of various tree crops without particular negative effects on crop
production and farmers income in both Southern and Northern Italy. However, the technique of
controlled deficit irrigation can be applied on the already grown trees since the deficit irrigation can
provoke negative impacts (later start of production and overall decrease of productivity) if applied
during the first three-four years since plantation.

A synthesis of results of the numerous deficit irrigation experiments carried out in Emilia-Romagna
(Northern Italy) on peach tree is given in Fig. 15 subdividing the vegetative cycle of
peach tree in 4 principal phases:
phase 1 from the start of flowering to the formation of small fruits (of 3-4 cm of diameter);
phase 2 from the end of the previous phase until the hardening of the pit;
phase 3 from the hardening of the pit until the harvesting;
phase 4 from the harvesting until the fall of the leaves.

Figure 15 illustrates that the water stress was induced during the phases 2 and 4. A controlled
water stress during the phase 2 does not favour development of shoots which reduces the
competition for assimilates between the shoots and fruits; similarly, during the phase 4 it reduces
vegetative growth and favours the induction of buds to flowers and fruit leader. The overall reductions
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

119
of irrigation volumes in respect to full irrigation in a normal year under Emilia-Romagna climatic
conditions were estimated between 56 and 68% for the medium early and early cultivars and clay soil
and between 20 and 23% for the late cultivars without significant differences related to the soil type
(Table 6). The results (Fig. 16) indicate that the regulated deficit irrigation technique has increased
crop production in respect to traditional irrigation, has maintained the average weight of fruits, has
improved the flowering in the successive years and has reduced the necessity for pruning. Similar
results have been obtained also in the experiments on peach and nectarine trees carried out under
Southern Italy climatic conditions.


Fig. 15. Graphical presentation of the stress thresholds to apply on the peach tree under regulated
deficit irrigation treatments (adapted after Mannini, 2004)

Table 6. Percentage of seasonal irrigation volumes saved by controlled water stress on peach in
respect to normal irrigation regime (Source: Mannini, 2004)
Interspace between rows cultivated Interspace between rows grassy
SOIL Early
cultivars
Medium early
cultivars
Late
cultivars

Early
cultivars
Medium early
cultivars
Late
cultivars
Sandy 44 38 20 38 34 20
Loam 58 59 20 52 46 23
Clay 68 56 22 60 51 23

The studies of regulated deficit irrigation has been done also on the herbaceous crops in
Southern Italy giving different results in respect to those obtained with orchards. In fact, serious drops
of production can be observed even in the cases of limited water reduction during the non-critical
phenological stages.

Four years of investigation on the regulated deficit irrigation of maize have been done in
Southern Italy (Policoro, Basilicata). The experiment was based on suspending one or two irrigations
or doubling irrigation volumes in correspondence to different phenological phases (a when crop has
achieved 1 m height, during the crop growing stage; b at the tassel emission; c at beginning of the
milky stage; d at the beginning of the waxy stage). The results have shown that all phenological
phases demonstrated certain sensitivity to water stress. Anyway, the most sensitive phase almost
always corresponded to the tassel emission and, in particularly dry years, to the phase of intensive
crop growth. These results indicated that maize is not well adaptable to the Southern Italy climatic
conditions where the spring-summer periods are characterized with scarce precipitations and very
high evapotranspiration demand. Consequently, maize should be fully irrigated under these climatic
conditions. In fact, maize is rarely cultivated under Southern Italy climatic conditions because this
crop is very sensitive to water stress and it should not be grown under deficit irrigation practices.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

120
Yield
Fruit set
R.D.I. Test
R.D.I. Test
38
37
36
35
34
33
32
78
76
74
72
70
68
F
r
u
i
t
(
t

h
a
-
1
)
F
r
u
i
t
(
N
.

m
-
1
)
135
130
125
120
3
2
1
0
Prunings
Average fruit size
R.D.I. Test
R.D.I. Test
g
W
o
o
d
(
N
.

m
-
1
)

Fig. 16. Productive and vegetative effects of water stress on peach tree grown as an espalier (by
Chalmers) (adapted after Mannini, 2004).

1000 0 2000 3000
0.0
1.0
2.0
3.0
4.0
5.0
6.0
7.0
8.0
Seasonal volume of irrigation water (m ha
-1
)
G
r
a
i
n

y
i
e
l
d
(
t

h
a
-
1
)
g
g
g
g
q
q
q
q
q
u
u
u
u
x
x
x
x
x x
x
x
A
AB
B
B
B
C
A
B
B
A
AB
BC
C
B
B
B
B
B
B
A
A
(b)
(s)
(s+b)
u u
x x
g g
q q
u u
x x
g g
q q Valenzano 1986
Valenzano 1987
Policoro
Gaudiano

Fig. 17. Variations of wheat production under different irrigation treatments. The values assigned with
the same letter are not significantly different at 0.01 P according to the Newman-Keuls
method. (s) irrigation only at the sowing; (b) irrigation only at the booting phase; (s+b)
irrigation at sowing and booting phase.

Finally, in the Southern Italy environments, characterized with high precipitation variability which
contributes to the instability of agricultural production even of non-irrigated autumn-spring crops (e.g.
wheat), there is a frequent application of supplemental irrigation strategies. This helps in stabilizing
agricultural production and improving the water use efficiency of precipitation. Several experiments
were carried out in Southern Italy on different wheat cultivars grown in deep soils with water
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

121
availability of 8.4% of dry soil weight (Gaudiano di Lavello Basilicata) and of 16.6% of dry soil weight
(Policoro Basilicata) and on shallow soil with water availability of 13.0% of dry soil weight
(Valenzano Puglia). The irrigation strategies included the application of water only during the critical
phenological stages (at sowing, at the booting phase and at both sowing and booting phase) and
during the whole growing cycle with different levels of limitations. The results of these investigations,
shown in Fig. 17, indicated that in particularly dry years one irrigation immediately after sowing
(example of Policoro in 1986) with water volume of 770 m
3
ha
-1
can be sufficient to increase
production from 2.0 t ha
-1
to 4.7 t ha
-1
, while any additional irrigation did not contribute to further
augment of grain yield.


STUDIES ON CROP WATER REQUIREMENTS

The researches on crop yield response to irrigation water required the intensification of the
studies on the adaptability of empirical methods for the estimation of reference evapotranspiration to
different Italian agro-climatic conditions. These studies were necessary in order to estimate and/or
foreseen better crop water requirements for both the irrigation management purposes and the
realization of irrigation projects. A particular attention has been given to the methods indicated in both
the FAO Irrigation and Drainage paper n24 (Dorenboos and Pruit, 1977, 1987) and in n56 (Allen et
al., 1998).

For the implementation of studies on crop water requirements, in many Italian regions have been
constructed the lysimeters of different characteristics by means of both functionality and size. Type,
dimensions and number of lysimeters used in various Italian locations are reported in Table 7, while
the spatial distribution of the lysimetric stations is indicated in Fig. 18.

Table 7. Type, dimensions and number of lysimeters used in various Italian locations
yp , y
Type
Surface area
(m
2
)
Depth
(m)
Presence of
guard
Underground(U)
or
Aboveground (A)
Location and number
l) DRAINAGE
a) groundwater
(70-110)
2x2 = 4 1,30 yes U
Policoro (6) Metaponto (2)
Foggia (4) S. Prospero (4)
Guiglia (4) Gela (2) Roma
2x2 = 4. 2,20 yes U Cadriano (2)
2x2 = 4 1,00 yes U Polignano (2) Cadriano (2)
1,25x1,25 = 1,56 1,40 no A Pisa (6)
1x1 = 1 1,50 yes U Legnaro (20)
b) free percolation 2x2 = 4 0,50 no U Vitulazio (16)
2,75 m*; 5,94 1,50 yes U Sassari (4) **
2) WEIGHING
a) mechanical 2x2 = 4 1,30 yes U
Policoro (2) Rutigliano (1)
Gaudiano (1) Villa d'Agri (1)
b) with loading cells 3 m**; 7,07 2,15 yes U Campo Volturno (4)
* circular
** for tree crops



Water consumption have been valued with drainage lysimeters by using the water balance
equation weekly or 10-days period, whereas it was measured with weighing lysimeters as a difference
in weight at the beginning and the end of the period under consideration, generally on a daily basis,
taking into account natural hydrological inputs, irrigation, and the quantities of drained water
(Tarantino and Onofrii, 1991).

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

122

Fig. 18. Location of some lysimeter stations in Italy

Lysimeters have not been used only for the research on the adaptability of different methods for
reference evapotranspiration estimates under various Italian climatic conditions, but also for the
studies on crop water requirements, or maximum crop evapotranspiration (ETc), during the growing
cycle of numerous herbaceous crops and some tree-crop species. Daily values of ETc measured for
various species have been rationed with the equivalent values of class A pan evaporation (E) and/or
reference evapotranspiration (ETo), calculated with different methods, in order to obtain
corresponding crop coefficients:
E
ETc
Kc = '
ETo
ETc
Kc =
The research locations, corresponding cultivars, years of experiments and some growing and
productive information are given in Table 8, the reference parameters used for the calculation of crop
coefficients (Kc and Kc) are reported in Table 9, while in Tables 10 and 11 are presented the crop
coefficient values (Kc) related to the class A pan evaporation (E). In Figures 19, 20 21 is given the
variation of Kc (derived from the ratio between the measured ETc and ETo calculated by the
Penman-Monteith equation) for some vegetables (muskmelon and eggplant) cultivated under plastic
mulches and without them.

Data reported in Table 10 confirmed that the lowest Kc values, in the range between 0.1 and 0.6,
were observed during the initial growing stage, about 30 days after sowing or planting, when the
water losses are prevalently due to soil evaporation. The highest values, between 0.85 and 1.50,
were observed when the full crop development has achieved and LAI reached the maximum values,
i.e. when the water is almost exclusively consumed in the process of transpiration. The Kc values
were decreasing gradually with the approximation of the end of crop growing cycle, in relation to the
vegetative state of the crops at harvesting.

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

123
Variability of Kc values during the initial crop growing stage is related to the humidity of the
superficial soil layers. In fact, the highest value (0.62) was registered for wheat, an autumnal sowing
crop, when the frequency of precipitation was relatively high and ETo was limited, and, therefore, the
soil water content in the superficial soil layers was pretty elevated. In fact, is well-noted that direct
water losses by soil evaporation increases with the increase of humidity of superficial soil layers.

The Kc values resulted substantially different in the Northern Italy environments, where the crop
growing cycles tend to make longer, in respect to the Southern Italy, where they are shorter: evident
examples are spring sowing tomato and sugar beet (Table 10). Notable differences are also observed
on the values of Kc of maize and sorghum grown under different climatic conditions: higher values
were in the Northern Italy (locations of S. Prospero and Guiglia) and lower in the Southern Italy
(locations of Policoro and Foggia), lower values for the early cultivars (FAO class 200-400), higher
values for the hybrids with longer growing cycle (FAO class 600-700).
Table 8. Crop growing parameters of some experiments on Kc carried out in Italy
Crop Location Cultivar
Years of
experiment
Sowing date
(2)
harvesting Plant density
Yield
(t/ha)
Average ETc
(mm)
On-field growing Herbaceous crops
Sugarbeet
Policoro Monohill 1975-76 March September 10 96.5 669
Cadriano Monogen 1981 March August 10 117.5 652
Artichoke Policoro Locale di mola 1974-75-76 August Decem. April 1 11.4 540
Polignano Locale di mola 1974-75-76-77 August Decem. April 1 29.3 557
Cabbage Broccolo
Summer cultivar Policoro Green duke 1977 September December 6 13.6 120
Winter cultivar Policoro Clipper 1986-87 July Sept. November 4 16.4 374
Cetrioli Policoro Pioner e Bounty 1977-78-79 July October 16 17.0 233
Alfalfa (l) S.Prospero Bresaola 1970-79-80-81-82 April October NR 20.8 (1) 939
Guiglia Bresaola 1970-79-00-81-82 April October NR 16.9 (1) 692
String bean Policoro LIT 551 1977-78-79 April July 65 14.8 276
Bean (type borlotto)
Fresh Policoro Lingua di fuoco 1984 June August 44 7.5 432
Dry Policoro Lingua di fuoco 1984 June September 44 3.0 479
Wheat Policoro Salapia 1985-86 November June 49 ears/m
2
6.7 475
Sunflower 1
st
harvesting Foggia Luciole 1981-82 April August 5.0 3.9 710
S. Prospero Luciole 1981-82 April September 5.1 3.6 571
Guiglia Luciole 1981-82 April September 5.0 3.4 605
Sunflower 2
nd
harvesting
after barley Pisa Mirage 1986 June October 6 3.3 537
after wheat Pisa Mirage 1986 July October 6 2.9 452
maize from granella 1
st
harvesting
Policoro
Dekalb XL 304
FAO 200
1974-75-76 April September 8 10.1 511
Foggia Dedalo 95 FAO 400 1976-77 April September 6 12.0 686
S. Prospero Titano FAO 700 1976-77 April October 6 13.5 589
Guiglia Titano FAO 700 1976-77 April October 6 12.3 587
Legnaro
Dekalb XL 342
FAO 606
1973-7 May September 6 10.0 450
Maize 2
nd
harvesting Pisa Leveret 400 1986 June October 8 13.2 582
Pisa Leveret 400 1986 July October 8 8.6 457
Potato Legnaro Bintje 1978-79 March August 4 50.0 600
Tomato Policoro Ventura 1976-77-78 April September 6 87.0 546
Legnaro Roma VF 1977-78 April September 40 80.0 451
Soya 1
st
harvesting S. Prospero Kig SOY 1983-84-85 May October 35 3.7 861
Legnaro TXR 505 1975-76 May October 30 4.0 500
Cadriano Hodson 78 1984 May October 40 5.2 618
Soya 2
nd
harvesting S. Prospero Arrok 1984 June October 35 2.9 420
Sorghum Policoro
Dekalb XL
FAO 200
1977-78 May October 25 12.7 690
Foggia NK 121 FAO 200 1978-79-80 April September 30 10.6 648
Guiglia 54BR FAO 200 1983-84-85 May September 50 8.2 512
Guiglia NK 180 FAO 400 1983-84-85 May September 35 6.9 589
Guiglia Savanna 5 FAO 600 1983-84-85 May September 35 6.2 630
S. Prospero Savanna 5 FAO 600 1983-84-05 April September 35 8.9 624
Legnaro NK 180 FAO 400 1974- 75 May October 16 9.0 465
Spinach Policoro Seven R 1978 February April 64 31.8 153
Muskmelon mulched 39.4 310
Muskmelon non-mulched
Gaudiano Nabucco 2001-2003 June August 0.5
26.7 257
Muskmelon mulched 30.2 229
Muskmelon non-mulched
Policoro Campero 1999 May August 1.0
27.8 320
Eggplant mulched 96.5 720
Eggplant non-mulched
Policoro Tasca 2003 May July-August 2.0
55.6 703
Tree crops
Orange tree Sassari Washington navel 1987 4 year December - - -
Apricot tree Ponticelli Cafona 1981-82-83-84-85 6 year July-January 400 plants/ha from 4 to 28 380
Olive tree Sassari Tondo di Cagliari 1987 5 year January -
Peach tree Livorno - - - - 1600 plants/ha
1
Yield of alfalfa refers to the total dry matter of 5 years of experiments obtained from 4-5 cutting per year; average annual consumptions refers to the period
May-June.
2
For tree crops it is intended as the years after planting.

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

124
Table 9. Crops and methods used for evaporation measurement and reference evapotranspiration
estimates at different locations in Italy

Evaporation (E) Reference evapotranspiration (ETo)
Crops Locations
Class
A
Wild
Grass
festuca
Blaney-
Criddle
FAO
Radiat.
FAO
Penman
FAO
Epan
FAO
Turc Thornthwaite
Penman-
Monteith
Other
methods
Sugar beet Policoro X X X X X X


Cadriano X X X


Artichoke Policoro X


Putignano X X X X


Cabbage Broccolo
winter
Policoro X


Cabbage Broccolo
summer
Policoro X X


String bean Policoro X


Bean Policoro X X X


Wheat Policoro X X


Sunflower Foggia X X X X X


S. Prospero X X X X X X

X
1

Guiglia X X X X X X X

X
1

Pisa X X


Maize da granella Policoro X X X X X X


Foggia X X X X X


S. Prospero X X X X X X


Guiglia X X X X X X
Legnaro X X X X X
2

Pisa X X
Potato Legnaro X X X X X
2

Tomato Poticoro X X X X X
Legnaro X X X
Soya Cadriano X X
S. Prospero X X X X
1

Legnaro X X X X X
1

Pisa X X X
Sorghum da
granella
Policoro X X X X
Foggia X X X X X
Guiglia X X X X X
1

S. Prospero X X X X X
1

Legnaro X X X X
1

Spinach Policoro X
Eggplant
Pepper
Muskmelon mulched and non
Gaudiano X
Policoro X
Eggplant Policoro X
Apricot tree Ponticelli X
Orange tree Sassari X
Olive tree Sassari X
Peach tree Livorno X


1
Formula of Tombesi-Lauciani.
2
Formula of Blaney-Morin, Hannon, Hargreaves, Hedke, Ivanov, Helse, Loury-Jensen.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

125
Table 10. Measured Kc values (ETc/E ratio) of some crops grown under different conditions in Italy
Days after sowing or planting
Crop Type of crop Location
10 20 30 40 50 60 70 80 90 100 110 120 130 140 150 160 170 180 190 200 210 220 230 240
Sugarbeet Spring Policoro 0.28 0.40 0.49 0.60 0.72 0.81 0.90 0.97 1.01 1.03 1.02 1.00 0.94 0.84 0.68
Cadriano 0.24 0.28 0.32 0.38 0.46 0.54 0.63 0.70 0.78 0.86 0.91 0.96 1.00 1.05 1.08
Artichoke (1) Policoro 0.44 0.52 0.61 0.68 0.75 0.82 0.88 0.93 0.97 1.02 1.05 1.07 1.08 1.07 1.06 1.04 1.00 0.96 0.90 0.85 0.78 0.70 0.65
Polignano . 0.73 0.81 0.88 0.93 0.99 1.03 1.07 1.10 1.12 1.15 1.20 1.20 1.15 1.09 1.02 0.92 0.88 0.80 0.75 0.70 0.65
Cabbage Broccolo
(2)
Winter Policoro 0.31 0.34 0.40 0.50 0.70 0.86 0.95 0.95 0.94
Summer Policoro 0.37 0.42 0.45 0.67 0.71 0.84 0.87 0.86 0.82 0.80
Cetriolo Policoro 0.28 0.45 0.62 0.98 0.90 0.70
String bean Policoro 0.36 0.54 0.70 0.82 0.94 0.88 0.84
Bean Policoro 0.60 0.51 0.53 0.70 0.82 0.86 0.83 0.70 0.43 0.25
Wheat Policoro 0.62 0.62 0.64 0.66 0.68 0.74 0.80 0.85 0.88 0.92 0.95 0.95 0.95 0.90 0.77 0.60 0.57 0.38 0.28
Sunflower 1
st
harvesting Foggia 0.30 0.40 0.53 0.64 0.75 0.87 1.00 1.16 1.25 1.25 1.10 0.95 0.50
Guiglia 0.25 0.28 0.32 0.50 0.77 1.00 1.20 1.50 1.55 1.40 1.20 1.00 0.50
S. Prospero 0.15 0.20 0.38 0.60 0.90 1.20 1.45 1.50 1.20 0.90 0.60 0.50 0.40
2
nd
harvest. after barley Pisa 0.48 0.64 0.86 1.08 1.36 1.46 1.46 1.38 1.30 0.92 0.72 0.68
2
nd
harvest. after wheat Pisa 0.56 0.69 0.84 1.04 1.24 1.36 1.38 1.20 1.00 0.76 0.60
Maize 1
st
harvesting Policoro 0.38 0.42 0.51 0.70 0.87 0.98 1.03 1.00 0.93 0.82 0.71 0.56 0.45
Foggia 0.41 0.47 0.60 0.80 0.95 1.06 1.14 1.14 1.07 0.96 0.85 0.78 0.55
S. Prospero . . 0.40 0.70 1.00 1.20 1.27 1.28 1.25 1.20 1.18 1.05 1.00
Guiglia 0.15 0.36 0.60 0.85 1.02 1.20 1.30 1.35 1.34 1.26 1.10 0.80 0.48
Legnaro 0.40 0.43 0.48 0.57 0.67 0.80 0.91 1.00 1.05 1.07 1.01 0.90 0.75
2
nd
harvest. after barley Pisa 0.54 0.52 0.64 0.85 1.15 1.40 1.60 1.68 1.64 1.46 1.36 1.40
2
nd
harvest. after wheat Pisa 0.52 0.58 0.70 0.86 1.04 1.19 1.30 .1.34 1.31 1.30 1.40
Potato Legnaro . 0.70 0.70 0.72 0.79 0.87 0.93 0.95 0.95 0.90 0.81 0.71
Tomato Policoro 0.35 0.40 0.45 0.55 0.80 0.99 1.07 1.10 1.09 1.01 0.90 0.78 0.69
Legnaro . 0.41 0.42 0.45 0.50 0.55 0.62 0.71 0.82 0.91 1.00 1.08 1.13 1.14 1.14 1.10
Soya 1
st
harvesting Cadriano 0.25 0.47 0.70 0.80 1.00 1.30 1.50 1.45 1.40 1.30
Legnaro 0.34 0.40 0.50 0.60 0.74 0.85 0.94 1.10 1.20 1.30 1.35 1.30 1.18 0.98 0.75
2
nd
harvesting S. Prospero 0.45 0.63 0.92 1.05 1.12 1.15 1.15 1.13 1.00 0.70
2
nd
harvest. after wheat Pisa 0.50 0.56 0.68 0.94 1.22 1.42 1.47 1.44 1.20 0.90 0.64 .
Sorghum 1
st
harvesting early Policoro 0.46 0.49 0.59 0.71 0.85 0.97 1.04 1.07 1.00 1.00 0.88 0.72 0.55
1
st
harvesting medium Foggia 0.43 0.47 0.56 0.65 0.77 0.95 1.00 1.00 0.99 0.96 0.92 0.86 0.78
1
st
harvesting early Guiglia 0.30 0.34 0.45 0.60 0.75 0.85 0.90 0.90 0.85 0.75 0.50
1
st
harvesting medium Guiglia 0.30 0.34 0.50 0.70 0.90 1.10 1.22 1.15 1.12 0.90 0.75 0.60
1
st
harvesting late Guiglia 0.30 0.34 0.55 0.73 0.90 0.95 1.10 1.23 1.20 1.10 0.90 0.75
1
st
harvesting late S. Prospero 0.35 0.45 0.60 0.72 0.85 0.97 1.18 1.24 1.17 0.76 0.63
1
st
harvesting medium Legnaro 0.30 0.43 0.55 0.67 0.80 0.90 0.95 0.97 0.97 0.95 0.91 0.86 0.80
Spinach Policoro 0.38 0.44 0.52 0.61 0.68 0.75 0.82 0.88 0.93 0.98 1.02
1
for artichoke, n of days of the vegetative recover
2
for these crops, n of days after planting
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

126
Table 11. Crop coefficient values of some tree crops related to the class A pan evaporation
Month
Crop Location
Apr May June July Aug Sept Oct
Authors
Apricot tree (cv. Cafona)
Drip irrigation Ponticelli (NA) 0.70 0.33 0.55 0.64 0.68 0.73 0.81 Ruggiero, 1986
Sprinkler irrigation Ponticelli (NA) 0.64 0.52 1.13 0.80 0.80 0.91 0.68 Ruggiero, 1986
Orange tree
(cv. Washington navel;
4th year after planting,
G.C.I. 20%)
Sassari - - 0.17 0.28 0.35 0.38 0.40
Dettori
(unpublished
data)
Olive tree (tondo di
Cagliari;
5th year after planting,
G.C.I. 30%)
Sassari - - 0.47 0.46 0.51 0.52 0.40
Dettori
(unpublished
data)
Peach tree (1) Livorno 0.55 0.81 1.01 1.00 Natali et al., 1984
1
On-field data.
G.C.I. Ground Cover Index

In Table 10 is shown that the peak Kc values of sunflower were anticipated a) in the case of
sunflower intercropping after barley and wheat in respect to the main crop and b) in the case of
growing in a valley in respect to hilly area (S. Prospero in respect to Guiglia). Moreover, the Kc
values of sunflower are higher in the case of cultivation under Northern Italy conditions (Guiglia) in
respect to Southern Italy (Foggia). For soya, the peak Kc values resulted more anticipated and lower
at the second harvesting which is related to the time of sowing and to the local environmental
conditions. The peak Kc values of some herbaceous crops (such as spinach, potato, bean,
cucumber, cabbage, broccoli, wheat and artichoke) were almost always lower than 1.0, except for the
artichoke with the values around 1.1.

The Kc values of tree crops change slightly during the vegetative cycle, although they can vary
notable between the species in relation to the density and the age of plants and applied irrigation
method: the Kc values are greater in the case of irrigation with sprinkler method than with drip
irrigation.

The Kc values obtained under Italian climatic conditions result higher than those recommended in the
FAO Irrigation and Drainage papers, especially for the herbaceous crops during the full development
phase. In fact, the Kc values reported in the FAO document represent the average data from different
environmental conditions and cultivars, while the data given in this document refer to the specific
environmental conditions, agricultural practices, cultivars and irrigation methods which can notable
influence the Kc values.

The Kc values of crops cultivated under plastic mulches (muskmelon and eggplant) have been
obtained for the Southern Italy in Lavello (Potenza) and Policoro (Matera). The data obtained for
muskmelon in the location of Lavello (Fig. 19) indicate that the growing cycle of mulched crops (Fig.
19a) is shorter than of non-mulched crops (Fig. 19b) and that the Kc values at the beginning of the full
development phase (10 days after planting) and immediately after the start of harvesting are greater,
while during almost the whole period of harvesting are lower. On the other side, the Kc values of non-
mulched crops were higher only during the first 10 days after planting. The higher Kc values of
muskmelon grown under plastic mulches during almost the whole growing cycle are related to the
greater vegetative development of mulched crops; it is also confirmed by the greater LAI values.
However, the mulched crops as had a rapid and anticipated development manifested the symptoms
of an earlier senescence of leaves which resulted in a fast reduction of Kc values. Furthermore, these
data indicate how the duration of phenological phases of muskmelon is notable shorter than that
reported in the FAO Irrigation and Drainage Paper n 56 (Allen et al., 1998), independently of
mulching. Moreover, as it is clearly demonstrated in Figure19, the Kc values obtained at location of
Lavello are notably higher than those indicated in the FAO documents.

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

127
Fig. 19. Relation between estimated and measured by lysimeter crop coefficient Kc during
muskmelon cycle cultivated with (a) and without mulch (b) in 2001 and 2003 in Lovello
Southern Italy (from Lovelli et al., 2004).
Fig. 20. Crop coefficient data (ETc/ETo ratio) of muskmelon cultivated under mulches and without
mulches in Policoro Southern Italy (from Cantore et al., 2005).

The Kc values obtained by Cantore et al. (2005) on mulched and non-mulched muskmelon (Fig.
20) grown in Policoro (Southern Italy) are very similar to those obtained in Lavello. In fact, the Kc
values of muskmelon cultivated under mulches are lower during the initial development phase, in the
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

128
first 10-15 days after planting, in respect to the non-mulched crops. However, the mulched
muskmelon reached more rapidly the full development phase and the Kc values for mulched crops
are higher than those for the non-mulched crops. Moreover, the mulched crops have demonstrated
faster and more intensive development as compared to the non-mulched crops, followed by a rapid
and anticipated senescence of leaves.

In Policoro, the Kc values of mulched and non-mulched eggplant (Yared Tesfagaber, 2004) were
very similar to those of muskmelon, although with less remarkable differences. In fact, the Kc values
of mulched crops were slightly lower during the first 20 days after the planting and they were slightly
higher during the successive growing phase, with the very similar phenological phases (Fig. 21). It is
interesting to emphasize that in Policoro, the yield production of both mulched eggplant and
muskmelon crops resulted greater than the yield of the non-mulched crops, although the water
consumption was slightly higher. In fact, in the case of the cultivars grown under mulches, the yield
water use efficiency was higher. Furthermore, the Kc values of these crops grown in South Italy are
higher than those reported in the FAO documents which indicates that they are influenced non only
by the environment in which they are cultivated but also by the cultivars and adopted agronomic
practices.

Fig. 21. Crop coefficient data (ETc/ETo ratio) of eggplant cultivated under mulches and without
mulches in Policoro Southern Italy (from Yared Tesfagaber, 2004).


WATER USE EFFICIENCY AND AGRONOMIC PRACTICES FOR IMPROVEMENT


In the agriculture field, the term Water-Use Efficiency (WUE) was introduced by Viets in the
middle of sixties (Viets, 1962). Since that time, it has been generally used to describe the relationship
between the crop growth development and the amount of water consumed, thus Stanhill (1986) called
it physiological water use efficiency. The physiological water use efficiency is more difficult to be
conceived as a proper efficiency, as it is not a non-dimensional value and it does not represent an
output/input ratio of only one entity. In fact, it describes a process in which water is consumed to
produce new entities (e.g. biomass, yield, etc.), and a maximum value attainable by theory does not
exist for reference (Monteith, 1984). The physiological efficiency is largely utilized by a wide
community of scientists (plant and crop-eco-physiologists, agronomists) and it can be applied at
different space- and time-scales as illustrated in Table 12.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

129

Table 12. Major definitions of water use efficiency terms, as reported by Steduto (1996).
Term
Definition
Time scale Space scale
Photosynthetic WUE
T
A

Minutes, hours Leaf
Carbon Water Flux
Ratio (CWFR)
ET
NCF
f
f
t
t
t
t

0
0

Hour, day, season Canopy
Biomass WUE
(BWUE)
ET
biomass
f
f
t
t
t
t

0
0

Week, season Plant, canopy
Yield WUE
(YWUE)
HI WUE biomass
Season Plant, canopy

In this paragraph is given the state of art of WUE and agronomic practices to improve WUE in
Italian agriculture under field conditions. It is based on the evaluation of the national scientific
literature and technical reports especially focusing on the Southern Italian region. Water use
efficiency values of many field crops, grown under optimal conditions (Table 13) and submitted to
some agronomic techniques, such as irrigation (Table 14), fertilization (Table 15), rotations (Table
16), mulching & early sowing (Table 17) are reported. In all the tables, water use efficiency is
calculated as the ratio of the above ground biomass and/or the yield over the amount of water used,
determined by different methods, and expressed as kg m
-3
.

Table 13 shows as, although all the studies refer to no-limiting environmental conditions and to
environments with similar weather conditions in Southern Italy, there exist a great variability in the
above-ground biomass WUE values among crops. In fact, although it is quite widely acquainted from
the literature the superiority of C
4
species to use water more efficiently than C
3
species, due to the
higher efficiency to fix CO
2
, their values may overlap or overcome those normally found for the C
3
, as
it occurs in the study of Rubino et al. (1999). In this case the very high values of biomass WUE of
sugarbeet (8.0 kg m-3) and rapa (14.0 kg m-3) are explained on the basis of the high net assimilation
rate linked to the high translocation efficiency of yielded sucrose to the roots in the former crop and of
the very low transpiration rate during the winter season in the latter crop. Nevertheless, it is important
to highlight that the biomass WUE value of sugarbeet refers to the total biomass, including the heavy
roots, and consequently it is difficultly comparable with the others. In the same study, very high yield
WUE values are found for celery, lettuce, rapa, pepper and ascribed to the short crop cycles
associated with the very elevated water content (about 85-95%) in the marketable parts of all these
crops.

The results obtained in a recent work carried out by Steduto and Albrizio (2005) to compare
biomass WUE among different crops (sunflower wheat, chickpea and sorghum) indicate large
variability in WUE values, also within the same C
3
group. From this study it is emerged the need to
normalize the amount of water evapotranspired by the climate (vapour pressure deficit and/or
reference evapotranspiration), in order to compare the WUE values of crops grown in different season
and/or year and climatic conditions. Similar conclusions have been reached also by Rubino et al.
(1999).

The effect of irrigation practice on both BWUE and YWUE is not obvious, as it is shown in Table
13 for several crops submitted to different water regimes, including deficit irrigation (ID). Irrigation is
considered among those strategies allowing to increase the water available for the crops: it may
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

130
increase growth and, consequently, WUE, provided that the water supplied by irrigation is transpired
and not lost as evaporation from the soil, drainage and runoff.

Tarantino et al. (1997) compared BWUE and YWUE among six crops and investigated the effect
of four irrigation regimes (rainfed, restitution of 50 and 100% of the crop evapotranspiration, and
deficit irrigation) on both BWUE and YWUE, showing the great variability among BWUE values of C
3

species and a different effect of irrigation regimes on the species. Concerning BWUE, it emerged that:
(i) among all the treatments, the highest values have been obtained on average by sweet sorghum (a
C
4
) and durum wheat (a C
3
); (ii) among the rainfed treatments of all the crops, the highest value was
reached by durum wheat; (iii) among the most watered treatments of all the crops, the highest value
was reached by sweet sorghum. Comparing the effect of water supply on YWUE, the best results
have been obtained by the restoration of minimum 50% of the crop evapotranspiration in sweet
sorghum, kenaf and tomato, while no significant variations have been noticed with increasing
irrigation regimes in sunflower and cotton. Nevertheless, for both crops excellent results have been
reached in the treatment irrigated by deficit irrigation method. Also durum wheat reached high YWUE
values by applying deficit irrigation method, further than without any irrigation. The results achieved in
this study are very important to highlight the importance of deficit irrigation practice for some crops
grown in environments with water restrictions. In deficit irrigation strategy, in fact, water is applied to
create a certain water deficit, which results in a small yield reduction that is less than the consequent
reduction in transpiration, and therefore a gain in WUE per unit water transpired, and possible lower
production costs if one or more irrigations can be eliminated (Kijne et al., 2001).


Table 13. Above-ground Biomass water use efficiency, yield water use efficiency, total water used
of field-grown crops under optimal conditions. Method to determine the water used, experimental
location and reference are also reported.
Crop
Above-ground
Biomass WUE
(kg m
-3
)
Yield
WUE
(kg m
-3
)
Total
water used
(mm)
Determination
of water used
Location Reference
450
477
862
920
180
360
536
161
Durum wheat
Soybean
Spring sugarbeet
Artichoke
Rapa
Broccoli
Pepper
Lettuce
Celery
4.0
-
8.0
*
2.9
**
14.0
4.8
2.0
-
2.3
**

1.7
1.0
11.0
1.4
**
7.8
4.2
7.4
19.5
27.4
**
316
weighing
lysimeter
Policoro,
Matera,
Basilicata
Rubino et
al., 1999
Kenaf 1.8
**
765
weighing
lysimeter
Lavello,
Potenza,
Basilicata
Rivelli et
al., 1998
891
485
230
Sunflower
Grain sorghum
Durum wheat
Chickpea
2.6
5.7
4.5
3.0
***


320
***

canopy
chambers

Valenzano,
Bari,
Puglia
Steduto &
Albrizio,
2005
532
631
Sweet sorghum
Kenaf
Jerusalem
artichoke
4.8
**

2.3
**

2.6
**


556
pan
evaporation &
Kc
Metaponto,
Matera,
Basilicata
Losavio et
al., 1999
*
Roots are included.
**
Avg of more years.
***
Incomplete crop cycle.





Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

131
Table 14. Effect of application of irrigation on above-ground biomass water use efficiency, yield water
use efficiency, total water used of field-grown crops. Method to determine the water used, experimental
location and reference are also reported. I
0
indicates the control; I
33
, I
50
, I
66
, I
100
, indicate irrigation
treatments with 33, 50, 100 percentage of ETc restoration; I
D
indicates treatment supplied by deficit
irrigation method.
Crop
Irrigation
treatment
Above-ground
Biomass WUE
(kg m
-3
)
Yield
WUE
(kg m
-3
)
Total
water used
(mm)
Determination
of water used
Location Reference
I
0
1.0 0.9 213
I
33
0.9 0.8 246
I
66
0.8 0.8 305
Tomato
I
100
0.8 0.8 361
water balance
Gaudiano,
Potenza,
Basilicata
Candido et
al., 2000
I
0
1.3 2.9 195
I
D
2.5 7.3 547
I
50
3.0 8.2 564
Sweet
sorghum
I
100
3.4 7.2 826
pan
evaporation &
Kc
I
0
1.2 0.5 217
I
D
1.8 0.6 464
I
50
1.4 0.6 534
Sunflower
I
100
1.6 0.7 859
weighing
lysimeter
I
0
1.8 0.9 176
I
D
1.8 0.7 438
I
50
1.6 0.7 421
Cotton
I
100
1.7 0.5 546
weight
lysimeter
I
0
3.8 1.2 281
I
D
3.2 1.2 339
I
50
2.6 1.0 454
Durum
Wheat
*

I
100
2.2 0.8 641
pan
evaporation &
Kc
I
0
1.2 2.2 281
I
D
1.3 4.0 697
I
50
1.5 4.9 570
Kenaf
I
100
1.5 5.0 859
weighing
lysimeter
I
0
1.6 8.1 115
I
D
1.5 8.2 359
I
50
1.3 14.4 369
Tomato
I
100
1.3 13.4 635
pan
evaporation &
Kc
Gaudiano,
Potenza,
Basilicata
Tarantino
et al., 1997
I
0
2.1
I
D
1.7 Muskmelon
I
100
1.1
I
0
1.0
I
D
0.9 Pepper
I
100
0.6

water balance
Matera,
Basilicata
Rivelli et
al., 2004
I
0

4.9
435
I
33

2.4
522
I
66

2.1
611
Sunflower
**

I
100

1.8
700
Seasonal
irrigation
volume +
rainfall
Villa dAgri,
Potenza,
Basilicata
Rivelli &
Perniola,
1997
I
70

3.5 0.5
501
No-flood
Rice
I
100

3.4 0.5
594
water balance
Metaponto,
Matera,
Basilicata
Losavio et
al., 2001
A 5.9 446
B 8.9 202
C 9.1 212
D 8.1 276
E 7.4 284
Sunflower
***

F

8.4 306
water balance
Pozzallo,
Ragusa,
Sicilia
Cosentino
et al., 1992
A 7.8
B 9.2
C 10.6
D 11.6
Cotton
****

E

9.2

Seasonal
irrigation
volume +
rainfall
Pozzallo,
Ragusa,
Sicilia
Foti et al.,
1992
*
avg of 2 years;
**
avg of 3 years.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

132
***
Letters indicate 6 irrigation treatments, the same amount of water was given at different stages of crop cycle, as
follow: A: full-irrigated; B: one irrigation at the head visible stage; C: one irrigation at the beginning of flowering; D: two
irrigations at stage of tenth leaf and at the beginning of flowering; E: three irrigations at stage of tenth leaf, at the head
visible stage and at the beginning of flowering; F: four irrigations at stages of tenth leaf, at the head visible stage,
beginning and end of flowering.
****
Letters indicate five irrigation treatments, different numbers of watering were given at different stages of crop
cycle.


Contrasting results with respect to Tarantino et al. (1997) have been shown by Candido et al.
(2000) on YWUE of tomato crop: in fact, YWUE was highest in the control and lowest in the treatment
with 100% evapotranspiration restoration. It is interesting to notice that in the rainfed treatment of
experiment of Candido et al. (2000) the amount of water used is about 2-fold higher than in Tarantino
et al. (1997), while opposite behaviour occurred in the well-irrigated treatment, although both the
studies have been carried out in the same location. It may be due to the very different cultivars used,
but the method utilized to determine the amount of water used plays a crucial role too.

In a recent study of Rivelli et al. (2004) the water use efficiency response of two important
vegetables (muskmelon and pepper), widely cultivated in the Southern Italy, have been compared,
under three different water regimes. The findings have indicated that BWUE was much higher in
muskmelon than in pepper in all the compared treatments, demonstrating a greater efficiency of the
former crop in using water and its better adaptability to tolerate water deficit conditions.


Table 15. Effect of application of fertilizers on above-ground biomass water use efficiency, yield
water use efficiency, total water used of field-grown crops. Method to determine the water used,
experimental location and reference are also reported.
Crop
Above-ground
Biomass WUE
(kg m
-3
)
Yield
WUE
(kg m
-3
)
Total
water used
(mm)
Determination
of water used
Location Reference
- 1.6 489
Grain
sorghum
+ 1.9 466
- 1.3 639
Sugarbeet
+ 1.5 655
- 0.4 309
Soybean
+ 0.6 322
- 1.2 298
Wheat
+

0.8 283
water balance
Foggia,
Puglia
Rizzo et al.,
1990
- 3.3
Sunflower
+

4.2
water balance Viterbo, Lazio
Campiglia &
Caporali,
1992
- 1.3 0.4 399
Soybean
*

+ 1.5 0.5 408
water balance
Foggia,
Puglia
Rinaldi et al.,
1996
- 2.0 0.6 538
Sunflower
+ 2.1 0.6 549
water balance
Foggia,
Puglia
Rinaldi &
Rizzo, 1999
- 2.2 837
Sunflower
+ 2.6

891
- 4.4 485
Grain
sorghum
+ 5.7

510
canopy
chambers
Valenzano,
Bari, Puglia
Steduto &
Albrizio, 2005
*
Avg. of 2 years



Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

133
Both the works carried out by Rivelli and Perniola (1997) and Cosentino et al. (1992) on sunflower,
under different water supplies, show how much it is difficult to compare the results among
experiments. The YWUE values found in these two experiments on sunflower greatly differentiated
from those reported by Tarantino et al. (1997), demonstrating that, for the same crop, YWUE varies
over a wide range. Similar consideration is valid for the findings of Foti et al. (1992) on cotton, as
compared to those of Tarantino et al. (1997). The causes of such great variability may be ascribed to
the application of different methods to determine the total water used and to the use of different
denominators in the WUE ratio. Many times, indeed, as water used by the crop, which represents
the denominator of WUE and/or WP ratios, is not considered the amount of water effectively lost by
transpiration, but the total amount of water supplied by irrigation plus the rainfall. Of course, this
amount is not all necessarily used by the crops for transpiration.

Table 16. Effect of crop rotations and intercropping on above-ground biomass water use efficiency,
yield water use efficiency, total water used of field-grown crops. Method to determine the water used,
experimental location and reference are also reported.
Crop
Above-ground
Biomass WUE
(kg m
-3
)
Yield
WUE
(kg m
-3
)
Total
water used
(mm)
Determination
of water used
Location Reference
Sorghum-Wheat 1.5 463
Sorghum-
Wheat+Soybean
1.9 466
Sugarbeet-Wheat 1.3 691
Sugarbeet-
Wheat+Soybean
1.5 655
Sunflower-Wheat 0.5 466
Sunflower-
Wheat+Soybean
0.6 487
Sugarbeet-
Wheat+Soybean
0.5 327
Wheat+Soybean 0.5 385
Sunflower-
Wheat+Soybean
0.5 344
Sorghum-
Wheat+Soybean
0.6 323
Wheat 0.5 266
Wheat+Soybean 1.2 286
Wheat+Sorghum 1.2 284
Sugarbeet-Wheat 0.7 242
Sugarbeet-
Wheat+Soybean
0.9 260
Sunflower-Wheat 0.8 258
Sunflower-
Wheat+Soybean
0.9 267
Sorghum-Wheat 0.7 275
Sorghum-Wheat+Soybean 0.7 258
water balance
Foggia,
Puglia
Rizzo et al.,
1990
Soybean
*
as main crop 1.0 0.4 861
Soybean
*
as catch crop
after barley
1.3 0.7 420
drainage
lysimeters
Modena,
Emilia
Romagna
Costantini
& Melotti,
1991
Sunflower-Wheat
**
2.0 0.5 246
Sunflower-
Wheat+Soybean
**

2.1 0.6 239
water balance
Foggia,
Puglia
Rinaldi &
Rizzo, 1999
*
Avg. of 3 years

*
Avg. of 12 years


Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

134
Application of fertilizers may not only result in increased growth but also in increased WUE, as it is
shown in Table 15. Fertilizers use may increase slightly the total amount of water used (e.g. Rinaldi et
al., 1996; Rinaldi and Rizzo, 1999), but the main effect is to increase early canopy growth so that it
shades the surface and therefore reduces the evaporation as a proportion of the total water that is
lost (Steduto and Albrizio, 2005). However, the positive effect of fertilizer in increasing both growth
and water used, and reducing the evaporation is not universal. In fact, in the study of Rizzo et al.
(1990) on wheat and grain sorghum an opposite behaviour in water use was observed between
treatments with either low or high application of fertilizers, despite large positive effect of fertilizers on
biomass production.

A proper choice of the crop rotation is of fundamental importance for an appropriate use of water,
and it affects the length of crop cycle (to be chosen), the efficiency for water uptake, the amount and
the quality of crop residuals, the number and type of soil tillage practices. All these factors influence
some important physical properties of the soil, such as the porosity, the water retention, the infiltration
rate and the evaporation from the bare soil. Consequently, also the WUE and WP result to be strongly
affected by the crop rotations, as it is shown in Table 16.

Rizzo et al. (1990) compared YWUE among rotations of wheat cultivated in monoculture, with or
without catch crop of soybean or sorghum, and three two-years rotations (sugarbeet-wheat;
sunflower-wheat; sorghum-wheat, with or without catch crop of soybean). For wheat the best YWUE
was reached in the monoculture with the catch crop of soybean or sorghum. For both sorghum and
sugarbeet as main crops, the best results were obtained with soybean as catch crop, while the YWUE
of both sunflower and soybean did not significantly differentiate among rotations. Also in the
experiment of Rinaldi and Rizzo (1999) all the investigated parameters (BWUE, YWUE and the water
used) for sunflower did not significantly varied in the rotation sunflower-wheat as compared to the
same rotation, but with soybean as catch crop.

Costantini and Melotti (1991) compared both BWUE and YWUE and the water requirement of
soybean cultivated for three years as main crop (spring sowing) and as catch crop (summer sowing).
From this study, it is emerged that the amount of water used by soybean as main crop was nearly
double in comparison with that used by soybean as catch crop, as consequence of a longer crop
cycle and the highest temperatures during the summer months. Soybean as main crop produced
more biomass and yield dry matter, as compared to the catch crop, but it showed a lower BWUE and
YWUE.

The effect of mulching and early sowing on BWUE, YWUE and the amount of water used is shown
in Table 17. Mulching practice is a common way to reduce evaporation from the soil surface, further
than decrease the soil temperature. In terms of water conservation, the main effect of mulches is to
reduce the rate of evaporation when the soil surface is damp and then to extend the duration of this
stage (Gregory, 2004). In a recent study of Cantore et al. (2005) the use of plastic mulches positively
affected both biomass and yield WUE of muskmelon; this effect was mainly due to the reduction of
about 40% of the evapotranspiration, as both the evaporation from the soil and the length of the crop
cycle were strongly reduced in the mulching treatment.

Early sowing of crops is a very important mean of maximizing crop yield and WUE. In fact,
increasing the early growth of the canopy when the soil surface is usually damp and the vapour
pressure deficit is low has proved effective in increasing WUE. Bonari et al. (1989) found that an early
sowing of ten days increased the yield of 54, 35 and 17% for maize, soybean and sunflower,
respectively. Hence also biomass and yield water use efficiencies increased significantly in all the
crops except of sunflower, although the water use in early sowing was higher than in the normal
sowing. Differently, Rivelli and Perniola (1997) dealing with sunflower found that the increase in yield
water use efficiency was strictly linked to the decrease in the amount of water used, as effect of a
reduced evaporation from the soil.







Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

135

Table 17. Effect of both mulching and early sowing on above-ground biomass water use efficiency,
yield water use efficiency, total water used of field-grown crops. Method to determine the
water used, experimental location and reference are also reported.
Crop
Above-ground
Biomass WUE
(kg m
-3
)
Yield
WUE
(kg m
-3
)
Total
water used
(mm)
Determination
of water used
Loca
tion
Reference
Muskmelon 1.7 8.7 320
Mulching Muskmelon 2.8 13.2 229
weighing
lysimeter
Policoro,
Matera,
Basilicata
Cantore et al.,
2005
Normal
sowing
0.7 487
Sunflower
Early
sowing
1.0 385
Seasonal
irrigation
volume +
rainfall
Matera,
Basilicata
Rivelli &
Perniola,
1997
Normal
sowing
4.0 1.9 457
Maize
Early
sowing
4.5 2.3 582
Normal
sowing
2.0 1.0 457
Soybean
Early
sowing
2.3 1.2 547
Normal
sowing
1.8 0.6 452
Sunflower
Early
sowing
1.7 0.6 537
Drainage
lysimeter with
variable water
table
Pisa,
Toscana
Bonari et al.,
1989


CONCLUSIONS

During the last 20-30 years, irrigated agriculture has been expanded over the whole Italian territory
assuring a more stable agricultural production. In the same period, an important development of
various irrigation techniques and agronomic practices have been occurred and followed by numerous
research activities especially in two relevant agricultural regions: Puglia region in the South, and
Emilia Romagna in the North in the delta of river Po.

The research activities on water saving practices in irrigation have been conducted mainly in
Southern Italy where the crop productivity is strongly influenced with limited precipitation, and
irrigation represents a fundamental practice in order to increase and stabilize agricultural production
over the years. Accordingly, a particular attention was given to the research related to crop water
requirements (estimation of reference evapotranspiration and crop coefficients), crop production
functions and application of irrigation methods and practices that improve water use efficiency. A
review of published data on crop water requirements revealed that the lowest irrigation volumes were
recorded for short crop-cycle crops (e.g. shell bean) while the highest volumes were observed for the
long-term crops especially if their growing cycle coincides with the summer season (e.g. spring sugar
beet). The presented data on the crop coefficients pointed out a large divergence between the data
measured under Italian environmental conditions and those published in the FAO Technical
documents. This is specially true in the cases of application of specific agronomic practices (e.g.
mulching) when the length of growing season and corresponding Kc values substantially differ from
those published in the literature. Therefore, further research is needed to revise the existing data and
to match better the modern agricultural practices, new varieties and recently adopted standard
method for reference evapotranspiration estimate (FAO Penman-Monteith approach).

This document reports the most important data related to irrigated agriculture in Italy and biomass
and yield water use efficiency values found in many experiments carried out mainly in Southern Italy.
Inasmuch as a large amount of data on WUE is available, there is a difficulty to compare them. In fact,
it has clearly emerged how for each particular crop both BWUE and YWUE vary over a wide range.
Possible reasons of it are: (i) the application of different methods to estimate the water used; (ii) the
use of different nominators/denominators in WUE and WP ratios. In such perspective, more efforts
should be done by the scientific community to make data comparable, using standardized procedures
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

136
and units of measurements. Certainly, a more clear conceptualisation of WUE and WP terms is
necessary at national and regional scale.

Agronomic practices to improve WUE rely on the improvement of water use efficiency (WUE)
defined in terms of the yield or the biomass per unit area divided by the amount of water used (or
transpired) to produce that yield or biomass. Hence, WUE can be enhanced from either crop
improvement that increases yield per unit of water transpired (increased transpiration efficiency), or
from crop management practices that minimize transpiration relative to other losses, or both. While
transpiration efficiency is strictly linked to crop species and it varies among cultivars, there is a wide
range of management practices that can reduce the loss of water by evaporation from the soil surface
(such us mulching, application of fertilizers, early sowing and the choice of cultivars with rapid early
growth) and/or increase the amount of water available to a crop (such as irrigation, fallowing and
suitable rotations, weeds control, and the choice of cultivars having deep roots). The success of these
practices at specific locations depends on soil properties, crop characteristics and climatic factors.


REFERENCES

Allen R.G., Pereira L.S., Raes D. and Smith M., 1998. Crop evapotranspiration Guidelines for
computing crop water requirements. Irrigation and Drainage Paper 24, Food and Agricultural
Organization of United Nations, Rome, 300 p.
Angus J.F. and van Herwaarden A.F., 2001. Increasing water use efficiency in dry land wheat. Agron.
J., 93: 290-298.
Campiglia E. and Caporali F., 1992. Effetto della disposizione spaziale degli individui, delle modalit
di semina e della concimazione azotata sulla consociazione girasole (Helianthus annuus L.)
cece (Cicer arietinum L.). Nota II. Complementariet per luso della risorsa acqua. Riv. di Agron.,
26, 4: 508-516.
Candido V., Miccolis V., Perniola M., Rivelli A.R., 2000. Water use, water use efficiency and yield
response of long time storage tomato (Lycopersicon esculentum MILL.). Proceedings 3rd ISHS
on: Irrigation Hort. Crops, Ferreira and Jones (Eds), Acta Horticolturae, 537: 789-797.
Cantore V., Boari F., Albrizio R., De Palma E., 2005. Influenza della pacciamatura sui consumi idrici e
sullefficienza duso dellacqua del melone. Convegno SIA, su: Ricerca ed innovazione per le
produzioni vegetali e la gestione delle risorse agro-ambientali, Foggia, 20-22 Settembre.
Cosentino S., Sortino O., Litrico P.G., 1992. Risposta produttiva, temperatura radiativi della copertura
vegetale e stato idrico della pianta nel girasole (Helianthus annuus L.) in secondo raccolto con
differenti regimi irrigui. Riv. di Agron., 26, 4: 633-640.
Costantini E.A.C. and Melotti M., 1991. Consumi idrici e risposte quanti-qualitative allirrigazione della
soia in coltura principale e intercalare nella bassa pianura emiliana. Riv. Irr. e Dren., 38, 1: 23-32.
Doorenbos, J. and Pruitt. W.O., 1977. Guidelines for predicting crop water requirements, Irrigation
and Drainage Paper 24, Food and Agricultural Organization of United Nations, Rome, 179 p.
Foti S., Copani V., Guarnaccia P., 1992. Interventi irrigui in momenti significativi del ciclo biologico del
cotone (Gossypium hirsutum L.) per una pi efficace valorizzazione dellacqua. Riv. di Agron., 26,
4, 663-670.
Giardini, L., 2002. Agronomia generale ambientale e aziendale. Patron Editore, Bologna, 742 p.
Giardini, L. and M. Borin, 1985. Proposta metodologica per lesame delle curve di risposta produttiva
allirrigazione. Riv. di Agron., XIX, 4, 239-250.
Gregory P.J., 2004. Agronomic approaches to increasing water use efficiency. In: Bacon M.A. (Eds.),
Water use efficiency in plant biology. Blackwell Publishing Ltd, CRC Press, 327 p.
Hatfield J.L., Sauer T.J., Prueger J.H., 2001. Managing soils to achieve greater water use efficiency:
A review. Agron. J., 93: 271-280.
INEA Istituto Nazionale di Economia Agraria, 2003. Italian Agriculture in Figures. Italian Ministry of
Agricultural Policies and Forestry, National Institute for Agricultural Economy, Rome, 172 pp.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

137
ISTAT, 2002. Census on Agriculture, ISTAT, Rome
IRSA-CNR, 1999. Un futuro per l'acqua in Italia, Eds. M. Benedini, A. Di Pinto, A. Massarutto, R.
Pagnotta, R. Passino, Quaderni IRSA-CNR, Roma.
Kijne J.W., Barker R., Molden D., 2003. Water productivity in Agriculture: Limits and Opportunities for
Improvement. CABI Publishing, UK, 332 pp.
Kijne J.W.,Tuong T.P., Bennett J., Bouman B., Oweis T., 2001. Ensuring food security via
improvement in crop water productivity. Available on: http://www.iwmi.cgiar.org/challenge-
program/pdf/paper1.pdf
Losavio N., Ventrella D., Vonella A.V., 1999. Consumi idrici, efficienza delluso dellacqua e della
conversione dellenergia in biomassa: parametri per valutare lintroduzione di nuove colture
nellambiente mediterraneo. Riv. Irr. e Dren., 46, 2: 34-38.
Lovelli S., Pizza S., Caponio T., Rivelli A.R., Perniola M., 2005. Lysimetric determination of
muskmelon crop coefficients cultivated under plastic mulches. Agric.Water Manag., 72:147-159.
Mannini P., 2004. Le buone pratiche agricole per risparmiare acqua. I supplementi di Agricoltura 18.
Regione Emilia-Romagna, Assessorato Agricoltura, Ambiente e Sviluppo Sostenibile, p.178.
Monteith J.L., 1984. Consistency and convenience in the choice of units for agricultural science. Exp.
Agric., 20: 105-117.
MPAF (Ministero delle Politiche Agricole e Forestali), Italian Ministry of Agricultural and Forestry
Policies, 2004. Irrigazione sostenibile la buona pratica irrigua (P. Scandella and G. Mecella, eds.).
Editorial project Panda, LInformatore agrario, 300pp.
Passioura J., 2004. Increasing crop productivity when water is scarce. From breeding to field
management. Proceedings 4th International Crop Science Congress on: New directions for a
diverse planet. Brisbane, Australia, 26 September 1 October. Published on CDROM. Web site:
www.regional.org.au/au/cs
Perniola M., 1994. Ecophysilogical parameters and water relations of sweet sorghum, cotton and
sunflower during drought cycles. Proceedings of the International Conference on: Land and water
resources management in the Mediterranean region. Mediterranean Agronomic Institute,
Valenzano, Bari, Italy. 4-8 September.
Rinaldi M. and Rizzo V., 1999. La coltura del girasole inserita in due avvicendamenti e sottoposta a
due livelli di input agrotecnico: produzione e uso dellacqua. Riv. di Agron., 33: 265-273.
Rinaldi M., Ventrella D., Fornaro F., 1996. Analisi di crescita, bilancio idrico e produzione di soia
(Glicine max (L.) Merr.) in secondo raccolto dopo frumento duro (Triticum durum Desf.) sottoposta
a due livelli di input agrotecnico. Riv. di Agron., 30, 2: 160-167.
Rivelli A.R. and Perniola M., 1997. Effetti del regime irriguo e dellepoca di semina su alcune cultivar
di girasole (Helianthus annuus L.) in tre ambienti della Basilicata. Riv. di Irr. e Dren., 44, 1: 17-25.
Rivelli A.R., Albrizio R., Lovelli S., Perniola M., 2004. Water use efficiency response of field-grown
muskmelon and pepper to environmental water status. Proceedings 4th International Crop
Science Congress on: New directions for a diverse planet. Brisbane, Australia, 26 Settembre 1
Ottobre 2004. Published on CDROM. Web site: www.regional.org.au/au/cs
Rivelli A.R., Perniola M., Tarantino E., Disciglio G., 1998. Consumi idrici e irrigui del kenaf (Hibiscus
cannabinus L.) in un ambiente meridionale. Riv. Irr. e Dren., 45, 3: 29-36.
Rizzo V., Castrignan A., Stelluti M., Ventrella D., Carlone G., 1990. Prime valutazioni sui bilanci idrici
di colture in rotazione. Ann. Ist. Sper. Agron., Bari, XXI, suppl. 2, 95-106.
Rosegrant M.W., Cai X., Cline S.A., 2002. Global water outlook to 2025: averting an impending crisis.
Food Policy Report 2020 Vision. Washington, D.C., IFPRI, 26 pp.
Rubino P., Cantore V., Mastro M.A., 1999. Studio dellefficienza delluso dellacqua di alcune specie
erbacee in un ambiente dellItalia meridionale. Riv. Irr. e Dren., 46, 2: 39-46.
Stanhill G., 1986. Water Use Efficiency. Adv. in Agron., 39: 53-85.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

138
Steduto P. and Albrizio R., 2005. Resource use efficiency of field-grown sunflower, sorghum, wheat
and chickpea. II. Water Use Efficiency and comparison with Radiation Use Efficiency. Agric. and
For. Meteor., 130: 269-281.
Steduto P., 1996. Water Use Efficiency. In: Pereira, L.S., Feddes, R.A., Gilley, J.R., Lesaffre, B.
(Eds.), Sustainability of irrigated agriculture, NATO ASI Series E: Applied Sciences. Kluwer
Academic Publ., Dordrecht, pp. 193-209.
Tarantino E., Rivelli A.R., Perniola M., Nardiello I., 1997. Efficienza nelluso dellacqua di alcune
colture erbacee sottoposte a differenti regimi irrigui: valutazione a livello di pieno campo. Riv. di Irr.
e Dren., 44, 1: 8-16.
Venezian Scarascia, M.E., Caliandro A., Giardini L., Quaglietta Chiaranda F., Rubino P., Giovanardi
R., Losavio N.and dAndria R., 1987. Yield response to different amounts of irrigation water for its
best utilization. Thirteenth International Congress on Irrigation and Drainage, Rabat, Morocco,
September 1987. Transaction actes, Volume I-D: 189-224.
Viets F.G., Jr., 1962. Fertilizers and the efficient use of water. Adv. Agron., 14: 223-264.
World Resources Institute, 2000. World Resources 2000-2001. People and ecosystems: The fraying
web of life. United Nations Development Programme, United Nations Environment Programme,
World Bank, Washington DC, 400pp.
Yared Tesfagaber L., 2005. Studio dellefficienza duso dellacqua di colture erbacee in Italia
meridionale. Tesi di Laurea in Agronomia Generale, Facolt di Agraria, Universit degli studi di
Bari.


























Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

139



EFFECTS OF DEFICIT IRRIGATION ON YIELD AND WATER USE EFFICIENCY
OF SOME CROPS UNDER SEMI-ARID CONDITIONS
OF THE BEKAA VALLEY OF LEBANON



F. Karam* , K. Karaa**, N. Tarabey***
* Lebanese Agricultural Research Institute, Dept. of Irrigation and Agro-Meteorology, P.O. Box 287,
Zahl, Lebanon
** Litani River Authority, Dept. of Rural Development, P.O. Box 3732, Bechara El Khoury, Beirut,
Lebanon
*** Lake Share Communities Union, Association of Irrigation Water Users in South Bekaa Valley,
Lala, Lebanon


SUMMARY - A six-year experiment (1998-2003) was conducted at Tal Amara Research Station in
the Bekaa Valley of Lebanon to determine water use, yield and water use efficiency in four annual
crops with contrasting response to deficit irrigation (DI); maize (1998-1999) a determinate species
with a limited capacity to adjust grain yield in response to water availability; soybean (2000-2001), an
indeterminate species with a high capacity to compensate the effects of early water stresses; cotton
(2001-2002), an indeterminate species with a larger capacity to adjust the number of dehiscent bolls
under stressful conditions, and sunflower (2002-2003), a determinate species with a single
inflorescence and an aptitude to tolerate moderate water stresses. Crop evapotranspiration (ET) was
measured using drainage and weighing lysimeters. In the plots, ET was measured using a simple soil
water balance model. Yield and its components were determined in sampling areas reserved for
harvest. Water use efficiency at grain (WUE
g
) or seed (WUE
s
) basis was calculated as the ratio of dry
yield to crop evapotranspiration (Y/ET), while water use efficiency at biomass-basis (WUE
b
) was
calculated as the ratio of dry biomass to ET (B/ET). For cotton, water use efficiency (WUE
l
) was
calculated as the ratio of dry lint yield to ET. Furthermore, the relationships between yield (Y) and
biomass (B) in one hand, and crop evapotranspiration (ET) in the other hand were examined using
linear models. Results of the experiments showed that corn seasonal ET reached on the lysimeter
952 mm in 1998 and 920 mm in 1999. Furthermore, grain-related water use efficiency (WUE
g
) varied
with corn treatments from 1.34 kg m
-3
to 1.88 kg m
-3
, while at biomass-basis (WUE
b
) the values varied
from 2.34 kg m
-3
to 3.23 kg m
-3
. For soybean, seasonal ET totaled 800 mm in 2000 and 725 mm in
2001. Seed-related water use efficiency of soybean (WUE
s
) varied from 0.47 kg m
-3
to 0.54 kg m
-3
,
while WUE
b
varied from 1.06 to 1.16 kg m
-3
. For Cotton, seasonal ET was 641.5 mm in 2001 and
669.0 mm in 2002. Average WUE
l
values varied among treatments from 0.43 kg m
-3
to 0.64 kg m
-3
,
while WUE
b
varied from 1.82 to 2.16 kg m
-3
. For sunflower, average across years of
evapotranspiration attained 827 mm. WUE
s
of sunflower varied among treatments from 0.64 kg m
-3
to
0.86 kg m
-3
, while at biomass-basis WUE
b
varied from 3.23 kg m
-3
to 4.8 kg m
-3
. The results also
showed that yield and biomass have positive, though weak, relationships with ET in corn and
soybean, while in cotton and sunflower the relationships are negatives.


Key words: Drainage lysimeter, weighing lysimeter, crop evapotranspiration, yield, biomass, water
use efficiency.


INTRODUCTION

Water is fast becoming an economically scare resource in many areas of the world. The need for
more efficient agricultural use of irrigation water arises out of increased competition for water
resources and increasing environmental concern (Doorenbos and Kassam 1988).

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

140
The best use of water must be made for efficient crop production and higher yields. Therefore,
agriculture under unfavorable climatic conditions and limited water resources can not be profitably
practiced unless on-farm water management techniques are designed to meet the present growing
demands of water for increased food production (Oad et al., 2001).
It is necessary therefore to develop new irrigation scheduling approaches, not necessarily based
on full crop water requirement, but ones designed to ensure the optimal use of allocated water. Deficit
irrigation (or regulated deficit irrigation, RDI) is one way of maximizing water use efficiency (WUE) for
higher yields per unit of irrigation water applied (English et al., 1990; English and Raja, 1996; Kirda et
al., 1999). The crop is exposed to a certain level of water stress either during a particular growth
period or throughout the whole growing season, without significant reductions in yields. The main
objective of deficit irrigation is to increase the WUE of a crop by eliminating irrigations that have little
impact on yield. The resulting yield reduction may be small compared to the benefits gained through
diverting the saved water to irrigate other crops for which water would normally be insufficient under
traditional irrigation practices.

In some irrigation schemes, the system is designed to deliver full irrigation in order to meet full
crop water demands (Walker and Skogerbe, 1987). In others, the system is designed upon minimum
allowable soil water depletion (James, 1988; Keller and Bleisner, 1990). Cuenca (1989) introduced for
the first time the concept of partial irrigation within an irrigation scheme, and suggested that under
some circumstances the designer might allow for greater soil water depletion, which could result in
reduced yields as an economic tradeoff against the higher costs of intensive irrigation water.

Deficit irrigation is a common practice in many areas of the world (English and Raja, 1996). A
number of researchers have analyzed the economics of deficit irrigation in specific circumstances and
have concluded that this technique can increase net farm income (Martin et al., 1989; English, 1990).
The potential benefits of deficit irrigation derive from three factors; increased irrigation efficiency,
reduced costs of irrigation and the opportunity costs of water. Four levels of applied water could be
defined as optimal in one sense or another (English et al., 1990):
The level of applied water at which crop yields per unit of land are maximized;
The level at which yields per unit of water are maximized;
The level at which net income per unit of land is maximized;
The level at which net income per unit of water is maximized.

The optimum level of applied water for a particular situation will be that which produces maximum
profit or crop yield, per unit of land or per unit of water, depending on whether the goal is to maximize
profits or food production and whether the most limiting resources is water or land. The two other
levels of applied water are the deficit levels at which net returns will be equal to those which would be
realized by full irrigation (English and Raja, 1996).

Irrigation management of crops involves a balance between vegetative and reproductive growth.
Excessive vegetative growth can delay maturity and reduce final yield. One of the irrigation strategies
that could be implemented to reduce excessive vegetative growth maintain yield and reduce water
use, leading to an improvement in water use efficiency, is regulated deficit irrigation (RDI), which may
be implemented during part of the growing season by regulating moisture within a desired deficit
range. RDI aims to optimize water use efficiency and therefore maximize the yield returned per unit of
water applied. Any minor yield loss which may result from the implementation of a mild moisture
deficit/stress under RDI is offset by the benefits of reduced water use leading to a reduction in
excessive vegetative growth (Kirnak et al., 2002). A variety of crops have been found to benefit from a
RDI strategy including maize, wheat, sunflower, potatoes, tomatoes and cotton. Irrigation using drip is
typically able to apply smaller quantities of water more frequently, and is better able to maintain soil
moisture at the mild deficit required to implement RDI.

The most desirable benefits associated with implementing RDI strategy in crops are:
The reduction in excessive vegetative growth
Maintenance of soil moisture in the most agronomical desirable range
An increase in water use efficiency, and
The ability to better capture and use in-season rainfall events after an irrigation event due the
maintained deficit.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

141

Where water scarcity exists at regional level, irrigation managers should adopt the same approach
to sustain regional crop production, and thereby maximize income (Stegman et al., 1990). Burt et al.,
(1997) defined irrigation efficiency (IE) as the proportion of irrigation water input to the farm for crop
production that was used by the crop as evapotranspiration (ET) over the growing season.
Tennakoon and Milroy (2003) defined crop water use efficiency (CWUE) as crop production per unit
of ET. CWUE quantifies the efficiency with which economic yield is produced as a function of water
used by the crop in the field. Furthermore, Tennakoon and Milroy (2003) defined farm water use
efficiency (FWUE) as the amount of yield produced per millimeter of total seasonal water input at the
farm level. Oad et al., (2001) calculated irrigation water use efficiency (IWUE) from the marketable
fruit yields and amount of water applied to the plants. Moreover, they calculated total water use
efficiency (TWUE) as the ratio of marketable fruit yields to water use.

The objectives of this study were to determine water use and yield in four annual crops with
contrasting response to deficit irrigation (DI); corn, soybean, cotton, and sunflower, and to examine
the existing relationships between yield and biomass, in one hand, and evapotranspiration in the
other hand.


MATERIAL AND METHODS

Field studies aiming at examining the response of maize (Zea mays L.), Soybean (Glycine max L.
Merril), cotton (Gossypium hirsutum L.) and sunflower (Helianthus annus L.) to deficit irrigation stress
were conducted during the period 1998-2003 at Tal Amara Research Station in the Central Bekaa
Valley of Lebanon (33 51' 44'' N lat., 35 59' 32'' E long., altitude 905 m a.s.l). Tal Amara has a well-
defined hot, dry season from May to September and very cold for the remainder of the year. Long-run
data indicate an average seasonal rain of 592 mm, with 95% of the rain occurring between November
and March. Crops were grown on deep and fairly-drained soil, characterized by high clay content
(44%). Measured field capacity (-0.33 bar) and permanent wilting point (-15 bars) averaged 29.5%
and 16.0% by weight. Extractable plant water is estimated at 190 mm for 1 m rooting depth and a bulk
density of 1.41 g cm
-3
.

Hybrid corn (cv. Manuel) was sown on 19 May in 1998 and 25 May in 1999 at 10 plants m
-2
.
Soybean hybrid (cv. Asgrow 3803) was sown on 10 May 2000 and 25 April 2001 at a density of 12
plants m
-2
. Cotton (cv. AgriPro AP 7114) was sown on 5 May in 2001 and on 13 May in 2002 at a
density of 10 plants m
-2
. Sunflower (cv. Melody) was sown on 2 June 2003 and 10 June 2004 at a
density of 8 plants m
-2
.

For corn, crop evapotranspiration (ET) in both years was measured using a set of two drainage
lysimeters of 4 m2 surface area (2 m 2 m) by subtracting the volume of drainage from the irrigation
amount. The lysimeters, 1.2 m deep, 24 m apart, aligned N-S, are situated in the middle of 1-ha field
(200 m N-S by 50 m W-E) (Karam et al., 2003).

For soybean, ET was measured by a weighing lysimeter of 16 m surface area (4 m 4 m) and 1.2
m deep, containing the same clay soil as in the drainage lysimeters. Watering of the lysimeter was
made upon a 30% soil depletion of the available water in the 0-100 cm soil layer. The weight loss of
the lysimeter due to soil evaporation and plant transpiration was measured with load cells and
recorded at a 15-minute interval on a computer located near the lysimeter. Water was supplied to the
lysimeter when the weight loss reached a threshold value. Data were transferred via telephone
modem to the irrigation laboratory, 500 m from the lysimeter. ET was determined as the difference
between lysimeter weight gains (irrigation and/or rain or dew) and the lysimeter weight loss (from soil
evaporation and plant transplantation) divided by the lysimeter surface area (16 m
2
), so that day/night
ET from midnight to midnight was computed as the average of 96 readings (one reading for each 15-
minute time scale). The lysimeter has an ET accuracy and resolution of 1 kg, which corresponds to
0.062 mm of water for a surface area of the soil in the lysimeter. Water percolating through the soil
mass of the lysimeter was collected and measured in a drainage reservoir located at the bottom of the
lysimeter, so that drainage was accounted for in the water balance calculation of the lysimeter (Karam
et al., 2005).

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

142
For cotton and sunflower, crop evapotranspiration (ET) was estimated using the FAO method
(Doorendos and Pruit, 1977) by multiplying reference evapotranspiration (ETrye-grass) as measured
in rye-grass lysimeters by the corresponding crop coefficients (Kc), which were derived for the
different growth stages (Doorenbos and Kassam, 1988):
ET = ET
rye-grass
K
c

Reference evapotranspiration (ETrye grass) was measured in a set of two rye-grass drainage
lysimeters of 4 m surface area (2 m 2 m) and 1m depth. The lysimeters are 24 m distant, aligned
W-E, and located inside the weather station (40 m 40 m), 50 m apart of the experimental plots.

Water was distributed in the field uniformly and simultaneously at 100% of field capacity using line
source drippers, 16 mm in diameter, 40 m long, aligned W-E and spaced 70 cm apart. The dripper
spacing was 40 cm, each delivering 4 l hr-1 of irrigation capacity at 100 kPa pressure.

For corn, deficit irrigation was applied continuously during the growing cycle upon the measured
crop evapotranspiration (ET). Water was then applied at 100% (I-100 treatment) and 60% (I-60
treatment) of ET. For soybean, cotton and sunflower, deficit irrigations were made by cutting-out
irrigation or for a two-week period during one or more of the different growth stages. For soybean,
deficit irrigations were applied at full bloom (R2 stage), at seed enlargement (R5 stage) and at mature
seeds (R7 stage). For cotton, deficit irrigations were applied at first open boll, at early boll loading,
and at mid-boll loading. For sunflower, irrigation was withheld for a two-week period prior to flowering
(E2 stage), at mid flowering (F1 stage), and at the beginning of seed formation (M0 stage) and at
seed ripening (M2 stage). For all crops, a control was fully-irrigated throughout the growing period.
Table 1 illustrates the deficit irrigation treatments for the crops under study.

At sowing, crops were irrigated to keep water content at 100% of soil available water. Weeds and
insects were adequately controlled. Each species was grown in a different section in a 2-ha
contiguous experimental field. Irrigation treatments were laid out within each crop in a block design
with three or four replications.

In the plots, evapotranspiration was calculated using a simple soil water balance model
(Doorenbos and Kassam 1988):
ET = I + P D
r
R
f

s

Where ET is evapotranspiration, I is irrigation application, P is effective rainfall, Dr is drainage
water, Rf is amount of runoff, and s is change in the soil moisture content determined by gravimetric
sampling. All terms in this equation are expressed in mm.

Moisture content in the 0-90 cm soil profile was measured gravimetrically before irrigations. Since
there was no observed runoff during the experiment and the water table was at 4 m depth, capillary
flow to the root-zone and runoff flow were assumed to be negligible in the calculation of ET. Drainage
below 90 cm, after a number of soil-water content measurements, was considered as negligible. So
the above equation was reduced to:
ET = I + P
s

Soil moisture in the plots was also measured using a Sentry 200-AP TDR (Time Domain
Reflectometry). The TDR was calibrated to the soil at Tal Amara over a wide range of soil moistures
(Sentry, 200-AP, 1994). Four access PVC tubes, 50 mm in diameter and 1.0 m in length, were
inserted in the middle rows of each plot. Readings were taken one day before irrigation and 2-to-3
days after irrigation during the growing seasons at 0-15, 15-30, 30-45, 45-60, 60-75 and 75-90 cm of
the soil profile. Readings were then converted to soil moisture content values using a locally-
calibrated equation. Gravimetric measurements and TDR readings were used to estimate seasonal
ET in the plots using a water balance model as indicated above.

At physiological maturity, all individual plants in the sampling areas were harvested to determine
above ground biomass production (B) and yield (Y). For corn, grain number per m2 and the 1000-
grain weight were determined. For soybean and sunflower, seed number per m2 and the 1000-seed
weight were also determined. For cotton, yield was determined by weighting lint at dry basis in the
sampling areas.

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

143
Our study is largely based on the linear relationships between Y and B in one hand, and ET in the
other hand. For this reason, we fitted linear models to the data:
Y = a1 (ET) + b1
B = a2 (ET) + b2
These models are the simplest and more often used models to describe the relationships between
yield, biomass and evapotranspiration. These relationships are appropriate frameworks to investigate
the pattern of water use efficiency, i.e. WUE = YET-1, or WUE = BET-1. This concept is widely used
in agronomic research. Departure from linearity can be tested through regression of log Y or log B on
log ET (Thompson et al., 1991). However, as this test can produce misleading results when the y-
intercept differs from zero, polynomial regressions were preferred, as in Thompson et al. (1991).

In corn, soybean and sunflower, water use efficiency at grain or seed-basis (WUEg,s) was
calculated as the ratio of yield at dry basis to the amount of crop evapotranspiration (Y/ET), while
water use efficiency at biomass-basis (WUEb) was calculated as the ratio of biomass at dry basis to
ET (B/ET) (Foroud et al., 1993; Howell et al., 1998). In cotton, water use efficiency at lint-basis
(WUEl) was calculated as dry lint yield to the amount of water evapotranspired from the crop
(Tennakoon and Milroy, 2003). WUE was expressed in kg m-3 (1 kg m
-3
= 1 g m
-2
mm
-1
).


Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

144
Table 1: Deficit-irrigation treatments of the various crops under study.
Crop Period Treatment Period of irrigation cutout
I-100 No irrigation restriction during the growing period 1998
I-60 Deficit irrigation at 6-leaf stage (from d.o.y 164 to d.o.y 255)
I-100 No irrigation restriction during the growing period
Corn
1999
I-60 Deficit irrigation at 6-leaf stage (from d.o.y 170 to d.o.y 250)
C No irrigation restriction during the growing period
S-1 Irrigation cutout at full bloom (R2 stage) (from d.o.y 215 to d.o.y 227)
S-2 Irrigation cutout at seed enlargement (R5 stage) (from d.o.y 236 to d.o.y 250)
2000
S-3 Irrigation cutout at mature seeds (R7 stage) (from d.o.y 250 to d.o.y 264)
C No irrigation restriction during the growing period
S-1 Irrigation cutout at full bloom (R2 stage) (from d.o.y 197 to d.o.y 211)
S-2 Irrigation cutout at seed enlargement (R5 stage) (from d.o.y 218 to d.o.y 232)
Soybean
2001
S-3 Irrigation cutout at mature seeds (R7 stage) (from d.o.y 232 to d.o.y 246)
C No irrigation restriction during the growing period
S-1 Irrigation cutout at first open boll (from d.o.y 217 to d.o.y 259)
S-2 Irrigation cutout at early boll loading (from d.o.y 231 to d.o.y 259)
2001
S-3 Irrigation cutout at mid boll loading (from d.o.y 245 to d.o.y 259)
C No irrigation restriction during the growing period
S-1 Irrigation cutout at first open boll (from d.o.y 232 to d.o.y 274)
S-2 Irrigation cutout at early boll loading (from d.o.y 246 to d.o.y 274)
Cotton
2002
S-3 Irrigation cutout at mid boll loading (from d.o.y 260 to d.o.y 274)
C No irrigation restriction during the growing period
S-1 Irrigation cutout prior to flowering stage (from d.o.y 232 to d.o.y 274)
S-2 Irrigation cutout at mid flowering stage (from d.o.y 246 to d.o.y 274)
S-3 Irrigation cutout at the beginning of seed formation (from d.o.y 260 to d.o.y 274)
2003
S-4 Irrigation cutout at mid seed ripening (from d.o.y 260 to d.o.y 274)
C No irrigation restriction during the growing period
S-1 Irrigation cutout prior to flowering stage (from d.o.y 232 to d.o.y 274)
S-2 Irrigation cutout at mid flowering stage (from d.o.y 246 to d.o.y 274)
Sunflower
S-3 Irrigation cutout at the beginning of seed formation (from d.o.y 260 to d.o.y 274)

2004
S-4 Irrigation cutout at mid seed ripening (from d.o.y 260 to d.o.y 274)





Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

145

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

146


RESULTS AND DISCUSSION

Evapotranspiration, yield and water use efficiency

Table 2 shows the values of evapotranspiration (ET), yield (Y), biomass (B) and water use
efficiency of the various crops under well and deficit irrigation conditions.

Corn

Seasonal ET reached on the drainage lysimeter amounts of 952 mm and 920 mm in 1998 and
1999, for total growing cycles of 128 days and 120 days, respectively. In the plots, crop
evapotranspiration totaled in the well-irrigated treatment (I-100) 863 mm and 833 mm in 1998 and
1999, respectively, while in deficit-irrigated treatment (I-60) ET totaled 575 mm and 556 mm in 1998
and 1999, respectively (Karam et al., 2003).

Grain yield on a dry basis declined in 1998 from 1520 gm-2 on the lysimeter to 1450 gm
-2
on the
full irrigated treatment (I-100) to 1080 gm
-2
on the deficit-irrigation treatment (I-60). In 1999 these
reductions ranged from 1340 gm
-2
on the lysimeter to 1280 gm
-2
and 1040 gm
-2
on I-100 and I-60,
respectively. Total aboveground biomass at harvest was also reduced by deficit irrigation. In 1998, a
reduction of 130 gm
-2
was observed in I-100 in comparison with the lysimeter, while the reduction on
I-60 exceeded 800 gm
-2
when compared to I-100. In 1999, these reductions were 100 gm
-2
and 400
gm
-2
, respectively.

Grain-related water use efficiency (WUEg) of lysimeter grown corn was 1.52 kg m
-3
in 1998 and
1.34 kg m
-3
in 1999. However, fully-irrigated corn had a WUEg of 1.68 kg m
-3
in 1998 and 1.54 kg m
-3

in 1999. Higher WUEg values of 1.88 kg m
-3
and 1.87 kg m
-3
were obtained in 1998 and 1999,
respectively, from the I-60 treatment. On a biomass basis, I-100 treatment had values of water use
efficiency (WUEb) of 3.16 kg m
-3
and 2.46 kg m
-3
in 1998 and 1999, respectively, while the I-60
treatment had values of 3.23 kg m
-3
and 2.97 kg m
-3
, respectively. On the lysimeter, these values
were 3.0 kg m
-3
and 2.34 kg m
-3
, respectively.

Soybean

Total evapotranspiration (ET) as measured by the drainage lysimeters in 2000 totaled 800 mm for
a total growing period of 140 days. However, when ET was measured by the weighing lysimeter in
2001, it was 725 mm during a growing period of 138 days (Karam et al., 2005).

Average seed yield was 3.2 t ha
-1
in the control treatment, compared to 3.5 t ha
-1
in the lysimeter,
whereas total aboveground biomass productions were 7.3 t ha
-1
and 8.1 t ha
-1
, respectively. Deficit
irrigation during R2 stage (S1) reduced biomass production by 16% (P<0.01) with comparison to the
control (C). Deficit irrigation at R5 stage has resulted in significant reductions (P<0.01) of 28% of seed
yield in the S2 treatment, while the reduction in aboveground biomass was only 6%, with respect to
the control, while deficit irrigation at R7 stage (S3) reduced only by 6% (P<0.05) seed yield in the S3
treatment.

Seed-related water use efficiency (WUEs) of the well-irrigated treatment was 0.47 kg m
-3
, showing
no consistent difference with the lysimeter grown soybean. Apparently in this experiment, WUEy of
the deficit-irrigated treatments S1 and S3 were 13% and 4% higher than the control. However, the S2
treatment had a WUEs value 17% lower than the control. For the biomass-basis, water use efficiency
(WUEb) of the control averaged 1.06 kg m
-3
, whereas WUEb of treatments S2 and S3 were 6% and
9% higher, respectively. No significant difference was found between treatment S1 and the control.

Cotton

Lysimeter measured crop evapotranspiration (ET) totaled 641.5 mm in 2001, while when
estimated with the FAO method in 2002 it averaged 669.0 for total growing periods of 134 days in
2001 and 140 days in 2002.

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

147
Average across years, the highest lint yield was obtained in S1 treatment, where it amounted
638.7 kg ha
-1
, followed by S2 (576.9 kg ha
-1
) and S3 (546.7 kg ha
-1
), while the control produced the
lowest yield (457.0 kg ha
-1
).

The highest lint water use efficiency (WUEl) was encountered in S1 treatment, and averaged 0.62
kg m
-3
, followed by S2 (0.50 kg m
-3
), S3 (0.46 kg m
-3
) and the control (0.36 kg m
-3
). These values are
very close to those obtained by Gilham et al., (1995). At biomass basis, WUEb varied from 2.07 kg m
-
3
in the control, to 1.97 kg m
-3
in S1 treatment, to 1.96 kg m
-3
in S2 and 1.93 kg m
-3
in S3.


Table 2: Crop evapotranspiration (ET), yield, biomass and water use efficiency at grain or seed basis
(WUE
g,s
) and water use efficiency at biomass basis (WUE
b
) of the various treatments

Crop Variety Period Treatment ET Yield
*
Biomass WUE
y
*
WUE
b
*

(mm) (t/ha) (t/ha) (kg m
-3
) (kg m
-3
)
Lysimeter 952.0 15.2 28.6 1.60 3.00
I-100 863.0 14.5 27.3 1.68 3.16
1998
I-60 575.0 10.8 18.6 1.88 3.23
Lysimeter 920.0 13.4 21.5 1.46 2.34
I-100 833.0 12.8 20.5 1.54 2.46
Corn Manuel

1999
I-60 556.0 10.4 16.5 1.87 2.97
Lysimeter 800.0 3.38 7.96 1.95 4.61
C 720.0 2.82 6.88 1.81 4.43
S-1 596.0 2.50 5.66 1.94 4.40
S-2 632.0 1.76 6.21 1.29 4.55
2000
S-3 647.0 2.57 6.64 1.84 4.75
Lysimeter 725.0 3.65 8.23 2.33 5.26
C 652.0 3.59 7.65 2.55 5.43
S-1 541.0 3.65 6.53 3.12 5.59
S-2 580.0 2.93 7.38 2.34 5.89
Soybean Asgrow
3803
2001
S-3 567.0 3.43 7.50 2.80 6.12
Lysimeter - - - - -
C 577.4 0.4233 2.47192 0.34 1.98
S-1 473.9 0.6534 1.90098 0.64 1.86
S-2 537.6 0.5682 2.11622 0.49 1.82
2001
S-3 542.6 0.5398 2.16691 0.46 1.85
Lysimeter - - - - -
C 602.2 0.4906 2.80900 0.35 2.16
S-1 482.9 0.6239 2.16020 0.61 2.07
S-2 531.8 0.5856 2.40480 0.50 2.09
Cotton AgriPro
AP7114
2002
S-3 569.6 0.5535 2.46240 0.44 2.00
Lysimeter - - - - -
C 827.0 4.96 21.05 6.00 25.45
S-1 676.0 5.43 25.25 8.03 37.35
S-2 664.0 5.59 24.64 8.42 37.11
S-3 686.0 5.16 22.59 7.52 32.93
Arena 2003
S-4 726.0 5.31 21.10 7.31 29.06
Lysimeter - - - - -
C 827.0 5.60 32.35 6.77 39.12
S-1 676.0 5.82 37.22 8.61 55.06
S-2 664.0 5.83 39.43 8.78 59.38
S-3 686.0 5.81 37.61 8.47 54.83
Sunflower
Arena 2004
S-4 726.0 5.43 31.68 7.48 43.64





Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

148

Sunflower

Year-to-year sunflower evapotranspiration reached a total of 827 mm, as calculated from the FAO
method. Average seed yield averaged in the full-irrigated control was 4.96 t ha
-1
. Deficit irrigation
increased seed yield by 9.5% in S1 treatment, 12.7% in S2 treatment, 4.0% in S3 treatment and 7.0%
in S4 treatment, with comparison to the control. Moreover, average seed-related water use efficiency
(WUEs) of the fully-irrigated control an average of 0.60 kg.m
-3
, while WUEs values of the deficit-
irrigation treatments were 0.80, 0.84, and 0.75 kg m
-3
, in S1, S2, S3 and S4, respectively.


Relationships of yield, biomass and evapotranspiration

Corn

The yield of field-grown corn tended to increase linearly with ET from a range of 500-600 mm (I-60
treatment) up to about 800-900 mm (I-100 treatment). Beyond these values, in the range of ET
between 900 and 1000 mm, additional water consumption did not increase significantly yield, as
marked in Figure 1. As a result, the slope of yield response curve to ET of corn was relatively
constant up to 800-900 mm, and then it decreased for ET values higher than 900 mm. The average
slope was equal to 10.5 kg ha
-1
mm-1. However, the coefficient of correlation was relatively high (R2
= 0.88). At biomass-basis, the slope of the response curve of biomass versus ET was equal to 22.2
kg ha
-1
mm
-1
, and the coefficient of correlation was 0.64.

Soybean

A linear relationship between seed yield and ET, with a weak but positive slope, was found in
soybean plants (Figure 2). This weak correlation indicates that soybean can compensate the effects
of early water stresses, as described earlier in this paper. Moreover, the coefficient of correlation was
also found to be very low (R2 = 0.05). In the range of ET variation between 500 and 800 mm, the
slope of the response curve of seed yield versus ET was 1.6 kg ha
-1
mm
-1
for seed yield and 6.2 kg
ha
-1
mm
-1
for aboveground biomass.

Cotton

Excessive irrigation application in cotton promotes vegetative growth at the expense of
reproductive development and lint yield. This was observed in Figure 3, where lint yield decreased
with increased ET in the range of 500-600 mm. Negative linear regression relationship was found
between cotton yield and evapotranspiration. The slope of the regression curve was -1.6 kg ha
-1
mm
-
1
, however, high coefficient of correlation was observed (R2 = 0.87). Deficit irrigation at first open boll
(S1 treatment) caused greater yield increase than at early boll development (S2 treatment) and mid-
boll development (S3 treatment). Deficit irrigation at first open boll seemed to limit vegetative growth
and led to a good fruit set and higher yields.

Sunflower

Figure 4 shows that seed yield of sunflower decreased with increased ET in the range of variation
of 500-900 mm. Negative linear regression relationship was found between seed yield and biomass in
one hand, and evapotranspiration in the second hand. The slopes of the regression curves was -3.4
kg ha
-1
mm
-1
for yield versus ET (R2 = 0.46) and -35.8 kg ha
-1
mm
-1
for biomass versus ET. Deficit
irrigation at the beginning of seed formation (S3 treatment) and at mid seed ripening (S4 treatment)
caused greater yield increase than at flowering stages (S1 and S2 treatments), and consequently
WUE values were higher in the former treatments than the later treatments.






Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

149

Corn (1998 - 1999)
y = 10.492x + 4632.9
R
2
= 0.8864
0
2000
4000
6000
8000
10000
12000
14000
16000
400 500 600 700 800 900 1000
ET (mm)
Y
i
e
l
d

-

k
g
/
h
a
(a)
Corn (1998 - 1999)
y = 22.192x + 4787
R
2
= 0.6423
0
2000
4000
6000
8000
10000
12000
14000
16000
18000
20000
22000
24000
26000
28000
30000
400 500 600 700 800 900 1000
ET (mm)
B
i
o
m
a
s
s

-

k
g
/
h
a
(b)


Fig. 1. Relationship between yield (a) and biomass (b) and evapotranspiration of corn (data points are
means of five quadrates of 1m
2
each per treatment 1 S.E; The lines represent linear
regression trend lines)



Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

150


Soybean (2000-2001)
y = 1.5797x + 2005.3
R
2
= 0.0494
0
500
1000
1500
2000
2500
3000
3500
4000
400 500 600 700 800 900
ET (mm)
Y
i
e
l
d

-

k
g
/
h
a
(a)
Soybean (2000-2001)
y = 6.1554x + 3087.7
R
2
= 0.3924
0
1000
2000
3000
4000
5000
6000
7000
8000
9000
10000
400 500 600 700 800 900
ET (mm)
B
i
o
m
a
s
s

-

k
g
/
h
a
(b)


Fig. 2. Relationship between yield (a) and biomass (b) and evapotranspiration of soybean (data
points are means of five quadrates of 1m
2
each per treatment 1 S.E; The lines represent
linear regression trend lines)
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

151

Cotton (2001-2002)
y = -1.6214x + 1427.1
R
2
= 0.8707
0
100
200
300
400
500
600
700
800
400 450 500 550 600 650 700
ET (mm)
Y
i
e
l
d

-

k
g
/
h
a
(a)
Cotton (2001-2002)
y = 5.4139x - 610.59
R
2
= 0.8032
0
500
1000
1500
2000
2500
3000
400 450 500 550 600 650 700
ET (mm)
B
i
o
m
a
s
s

-

k
g
/
h
a
(b)


Fig. 3. Relationship between yield (a) and biomass (b) and evapotranspiration of cotton (data points
are means of five quadrates of 1m
2
each per treatment 1 S.E; The lines represent linear
regression trend lines)
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

152

Sunflower (2003-2004)
y = -3.4422x + 7666.7
R
2
= 0.462
0
1000
2000
3000
4000
5000
6000
7000
400 500 600 700 800 900 1000
ET (mm)
Y
i
e
l
d

-

k
g
/
h
a
(a)
Sunflower (2003 - 2004)
y = -35.793x + 54776
R
2
= 0.3172
0
5000
10000
15000
20000
25000
30000
35000
40000
45000
0 200 400 600 800 1000
ET (mm)
B
i
o
m
a
s
s

-

k
g
/
h
a
(b)


Fig. 4. Relationship between yield (a) and biomass (b) and evapotranspiration of sunflower (data
points are means of five quadrates of 1m
2
each per treatment 1 S.E; The lines represent
linear regression trend lines)





Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

153


CONCLUSIONS

Clearly, water use efficiency is not only dependent on the total water applied but also when it is
applied. Given that maximum yield was observed at a given ET value, WUE shows a general decline
above this value. Therefore, an optimum seasonal ET varying between 500 and 900 mm was
indicated for each crop, where the highest yield was observed. This provides high yield potential with
maximum WUE, and may provide a useful guide for retrospectively assessing irrigation strategy.

The relationship between yield and ET is an appropriate framework to investigate the pattern of
deficit irrigation. Furthermore, these two variables bring forth the variable water use efficiency, i.e.
WUE = YET-1, a concept widely used in agronomic and irrigation research. Thus, our study was
largely based on the relationships between Y and ET. Additionally, we investigated the relationship
between biomass and ET.

The linear models are the simplest and more often used models to describe the relationship
between Y and B, in one hand and ET, in the second hand. Departure from linearity can be tested
through regression of log Y on log ET, or log B on log ET. However, as this test can produce
misleading results when the y-intercept differs from zero, polynomial regressions were preferred, as in
Thompson at al. (1991). Moreover, these two models have important implications for the water use
efficiency, either at grain or seed basis, or biomass basis. Depending on whether the slope is
constant or variable, and whether the intercept is zero or negative, the expected relationship between
Y and B and ET can be outlined.

It is worthwhile noting that the parameters a1 and a2 were positives in corn and soybean, with
intercepts b1 and b2 different from zero, indicating thus the relationships between Y and B and ET
were positively correlated. However, the coefficient of correlation was much higher in both
relationships for corn than for soybean. These results demonstrate that corn has a limited capacity to
adjust grain yield in response to water availability, while soybean can compensate the effects of early
water stresses. On the contrary, the parameters a1 and a2 were negatives in cotton and sunflower,
with intercepts b1 and b2 different from zero, indicating thus the relationships between Y and B and
ET were negatively correlated. However, the coefficient of correlation was much higher in cotton than
in sunflower.

Maximum WUE values at grain basis were close to 1.88 kg m
-3
in corn, 0.55 kg m
-3
in soybean,
0.64 kg m
-3
in cotton, 0.62 kg m
-3
in sunflower. Our study demonstrated that WUE stability strongly
varied among deficit irrigation treatments of the same species.


REFERENCES

Burt, C.M., Clemmens, A.J., Strekoff, T.S., Solomon, K.H., Bliesner, R.D., Hardy, L.A., Howell, T.A.,
Eisenhauer, D.E., 1997. Irrigation performance measures: efficiency and uniformity. J. Irrig.
Drain. Eng. ASCE 123: 423-442.
Cuenca, R.H. 1989. Irrigation System Design: an Engineering Approach. Prentice Hall, Engle wood
Cliffs, NJ.
Doorenbos J., Kassam A.H. 1988. Yield response to water. Irrigation and Drainage Paper No. 33,
FAO, Rome.
Doorenbos, J., Pruit, W.O., 1977. Guidelines for prediction of crop water requirements United Nations
Food and Agriculture Organization, Irrigation and Drainage Paper 24, Rome, Italy, 144 pp.
English M.J., Musich, J.T., Murty, V.V.N. 1990. Deficit irrigation. In: G.J., Hoffman, T.A., Howell and
K.H., Soloman (Eds.). Management of Farm Irrigation Systems. ASAE, St. Joseph, MI.
English, M., and Raja, S.N. 1996. Perspectives on deficit irrigation. Agriculture Water Management,
32: 1-14.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

154
English, M.J., 1990. Deficit irrigation. I: Analytical framework. J. Am. Soc. Civil Eng., 116 (IR3): 399-
412.
Karam, F. J. Breidy, C. Stephan, and Joe Rouphael. 2003. Evapotranspiration, Yield and Water Use
Efficiency of Drip Irrigated Corn in the Bekaa Valley of Lebanon. Agricultural Water
Management, 63: 125-137.
Karam, F., R. Masaad, T. Sfeir, O. Mounzer and Y. Rouphael. Evapotranspiration and seed yield of
field grown soybean under deficit irrigation conditions. Agricultural Water Management, 75(3):
226-244.
Foroud, N., Mundel, H.H., Saindon, G., Entz, T., 1993. Effect of level and timing of moisture stress on
soybean yield components. Irrigation Science 13: 149-155.
Gilham, F.E.M., Thomas, M.B., Tijen, A., Matthews, G.A., Rumeur, C.L., Hearn, A.B., 1995. Cotton
production prospects for the next decade. World Bank, Technical Paper Number 287. World
Bank, Washington, DC, USA, 275 pp.
Howell, T. A., Tolk, J. A., Schneider, A. D., Evett, S. R., 1998. Evapotranspiration, yield and water use
efficiency of corn hybrids differing in maturity. Agronomy Journal, 90: 3-9.
James, L.G., 1988. Principles of Farm Irrigation System Design. Willey, New York.
Keller, J., and Bleisner, R.D., 1990. Sprinkle and Trickle Irrigation. Van Nostrand Reinhold, New York.
Kirda, C., Moutonnet, P., Hera, C., and Nielsen, D.R. 1999, Crop yield response to deficit irrigation,
Kluwer Academic Publishers, Dordrecht.
Kirnak, H., Tas, I., Kaya, C., Higgs, D. 2002. Effects of deficit irrigation on growth, yield, and fruit
quality of eggplant under semi-arid conditions. Aust. J. Agric. Res., 53: 1367-1373.
Martin, D., van Brocklin, J., and Wilmes, G., 1989. Operating rules for deficit irrigation management.
Trans. ASAE, 32 (4): 1207-1215.
Oad, F.C., Soomro, A., Oad, N.L., Abro, Z.A., Issani, M.A., and A.W. Gandahi, A.W. 2001. Yield and
Water Use Efficiency of Sunflower Crop under Moisture Depletions and Bed Shapes in Saline
Soil. Online Journal of Biological Sciences, 1 (5): 361 362.
Sentry 200-AP. 1994. Users Manual of Operation and Instruction. Troxler Electronic Laboratories,
Inc. and Subsidiary, Troxler International, Ltd. Cornwallis, USA.
Stegman, E.C., Schartz, B.G., and Gardner, J.C. 1990. Yield sensitivities of short season soybeans to
irrigation management. Irrigation Science, 11: 111-119.
Tennakoon, S.B., and Milory, S.P. 2003. Crop Water Use and Water use efficiency on irrigated cotton
farms in Australia, Agriculture Water Management 61: 179-194.
Thompson, B.K., Weiner, J., Warwick, S.I. 1991. Size-dependent reproductive output in agricultural
seeds. Canadian Journal of Botany, 69:442-446.
Walker, W.R., and Skogerboe, G.V., 1987. Surface Irrigation. Theory and Practice. Prentice Hall,
Englewood Cliffs, NJ.
















Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

155



























































Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

156


WATER USE EFFICIENCY
AND WATER PRODUCTIVITY IN MALTA



G. Attard, J. Mangion, P. Micallef
Institute of Agriculture, Malta Resources Authority, Water Services Corporation, Malta


SUMMARY This paper is focused on the water use efficiency and water productivity of agricultural
sector in Malta. Irrigation water demand is discussed in the view of limited water resources. The
presents status of agricultural water use is presented pointing out the main factors of inefficiency in
the use of irrigation water such as: a) losses in the storage reservoirs; b) losses in the conveyance of
irrigation water; c) low efficient on-farm irrigation methods, d) inefficiency related to the irrigation
systems setup and e) losses due to inadequate irrigation practices. Water productivity is analyzed
from an economic point of view through a study on economic considerations regarding markets for
water in the Maltese Islands. Two scenarios (wet and dry season) are elaborated by means of crop
water requirements, costs and income of agricultural production. The cultivation of crops which have
low crop water requirements but have a high economic return should be encouraged together with an
extensive use of treated sewage water for agriculture.

Key words: water use efficiency, water productivity, economics, Malta.


INTRODUCTION

Agricultural activity accounts for 3% of Maltas Gross National Product and contributes to 2%
employment. The total territory of the Maltese Islands encompasses 315 km2, of which approximately
one-third is available for cultivation. The 10-11,000 hectares are shared by some 15,000 producers,
mostly part-timers. Agricultural land census (2000) indicates that about 8-11 % of the whole
agricultural land is irrigated whereas the rest is rain fed. The average annual rainfall is approximately
550mm and the effective precipitation is 300mm per year. Taking into consideration that the rainy
season extends from September till March, vegetables and summer crops cannot be grown unless
supplementary irrigation water is available. In a region where the availability of water is normally less
than the water demand, aspects in water management such as use efficiency and productivity play a
major role. The paper will address the current situation in the use of water efficiently and its
productivity in Malta with special reference to agriculture.


PRESENT STATUS

Irrigation demand vs water availability

Mild wet winters and hot dry summers typifies a normal climate scenario in Malta. The
unpredictable patterns of rainfall distribution and intensity make it difficult to forecast. With the bulk of
precipitation falling during the month of October, water abundance is present when demand is
relatively low, (germination stage). On the other hand, during peek growth and maturation natural
moisture is almost completely absent. Irrigation is a must to ensure decent crop performance and
yields. Summer crop are totally dependant on the availability of irrigation water. Mitschoff (1990)
carried out an extensive report on water requirements of most crops cultivated in Malta. Using
evapotranspiration and effective rainfall data at the time, he established that the total annual water
requirements based on 30% cropping intensity was 12,319 m
3
per hectare of land. Thus, the total
annual irrigation demand amounts to approximately 123 Mm
3
. On a national level, assuming that 4.5
Mm
3
come from the sewage treatment plant, about 30Mm3 from rain water, about 89Mm3 must be
coming from private groundwater extraction.



Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

157


Water Use Efficiency

In theory, agriculture competes with other water users and this competition should induce the
agriculture sector to enhance water use efficiency and productivity. Since water used for irrigation is
still very affordable and does not carry a realistic price, water use in agriculture is used in a very
careless manner. The main use of water in agriculture is without any doubt for irrigation. Sources of
irrigation water range from rainwater to groundwater to treated sewage effluent (all three sources
carry negligible costs to the farmer). The indiscriminate use of such waters has put an appreciable
stress on the economy of Malta as all types of irrigation water have, to different extent, a monetary
factor attached to it. Unfortunately to-date, irrigation methods are not so efficient. Several factors lead
to inefficiency in the use of irrigation water. These include:

1) Inefficiencies in the storage of irrigation water.

There are three main ways in which irrigation water is stored for eventual use. The method of
construction and operation of these water-bearing structures have led to water loss and hence
resulted in an ineffective utilization of this rare resource.

a) Underground reservoirs and cisterns situated in fields and other rural areas:

The majority of these structures are excavated and constructed in bedrock. There are a good
number of these are not sealed and/or partially plastered. Therefore water is lost slowly by infiltration
into the rocks. Slow ground movements and expansion in concrete render those that are plastered
liable to develop fissures and hence resulting in water loss. Silting results from heavy storms
especially during the first storms of September and October. Unfortunately, the lack of maintenance
of these reservoirs (re-sealing and cleaning from silt) mainly due to financial problems and lack of
awareness for water conservation has led to reduction in the capacities of these reservoirs. Moreover,
many reservoirs, especially cisterns situated in fields and other rural areas have been abandoned
following the introduction of drilling boreholes and pumping groundwater directly from the aquifers.

b) Above ground reservoirs/soakpits (open top):

These are used mainly for the storage of water pumped from the aquifers, storage of water from
the sewage treatment plant (in the South East area of Malta) and catching of runoff storm water. In
general, these are kept in better condition since their maintenance is easier. Water loss from these
reservoirs is mainly due to evaporation. (Figure 1).

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

158
Fig. 1. Above ground water storage reservoir in Malta

c) Dams:

Over the years, many dams have been constructed by the government in many valleys to trap run-
off rainwater with the intention to retain soil as well as to store water for irrigation and for aquifer re-
charge. A good number of these have never been maintained and/or cleaned since then, and today
have silted up almost completely and therefore do not hold water. Others need extensive repairs
especially on the sides from where water escapes downstream (Figure 2).


Fig. 2. Wied is-Sewda Dam in Malta


2) Inefficiencies in the conveyance of irrigation water.

a) Gravity systems:

Some gravity systems are still in use, especially where there are natural springs and in the South
East area for the distribution of the treated effluent from the SantAntnin Sewage Treatment plant.
Water is conveyed by means of open channels. These are constructed from a number of channel
pieces, each one hewn out of stone or made from pre-cast concrete, placed end to end. Mortar or
cement is used to bond the pieces of channels together. The mortar often gets broken due to
subsidence of the channel support or by being accidentally hit, and water leaks out of the channels.
Overflowing and evaporation is also responsible for a good percentage of the water losses (Fig. 3).

b) Pressurized systems:

Galvanized pipes and/or polyethylene (PE) pipes are normally used to convey irrigation water. The
galvanised pipes are now rarely used. In general, the PE pipes used are of good quality and have a
long life. Unless the pipes are corroded or punctured, the chances of water losses from these
systems are lower when compared to gravity systems. On the other hand, if punctured, water losses
could be high since the system would be under pressure. Generally leaks are observed near the
joints and the connections between fittings.

3) Inefficiencies in the irrigation systems.

a) Furrow irrigation:

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

159
This method is still practiced in some places, but generally on small scale irrigation systems.
Water use efficiency is considered medium to low, since a large volume of water is supplied in a
relatively short time. This leads to considerable deep percolation and also evaporation.

b) Sprinkler irrigation:

This is the second most popular irrigation method used in Malta. It is mainly used for the irrigation
of potatoes, fresh vegetables and fodder. The two main factors that make sprinkler irrigation not very
suitable to Malta are the prevailing windy conditions and the fact that most of the fields are small and
of irregular shape. This makes it very difficult to control the wetted area and this result in water losses
due to it being thrown outside the field boundaries. Sometimes we observe that very powerful spray-
guns are used. Apart from depositing part of the water outside the field, the soil is unable to absorb
the water fast enough, and results in surface runoff. Irrigating with sprinklers around midday also
results in large evapotranspiration losses.

c) Drip irrigation:

This is the most common irrigation system used in Malta. The water is used much more efficiently
but we still observe over-irrigation in some places.

One can also note other problems associated with the setup of irrigation systems which eventually
lead to water use inefficiencies.

4) Inefficiencies related to the irrigation system setup.

This is mainly the result of lack of technical information.
a) Lack of proper design of the irrigation system, mainly to undue consideration of pressure losses
along the pipes, leading to problems in uniformity distribution. Not very acute in Malta since most
fields are small and flat.
b) Improper layout of the irrigation system, especially sprinklers, which can also lead to uneven
uniformity distribution and water losses out of the field boundaries (very common since fields in Malta
are small and of irregular shape).
c) Omission of certain accessories from the system due to lack of know how or due to financial
restrictions affecting the overall performance of the system, e.g., leaving out or using an improper
filter from a drip irrigation system (because groundwater looks clean!), resulting in clogging of the
drippers. Pressure gauges are usually installed the first time, but then neglected.
d) Not using the proper irrigation system and equipment suitable to the size of the field, type of crop
and type of soil. E.g. using powerful spray guns in small fields.
e) Using cheap material or excessive recycling of certain items. Here again, the main cause is the
financial aspect. E.g. re-using drip tape over and over again, even though many emitters had been
clogged.
f) Lack of maintenance of the system, e.g. cleaning of filters and repairing leaks.

5) Inefficiencies due to improper irrigation practices.

This is mainly the result of lack of information on crop-water characteristics, soil moisture and
evapotranspiration. The following are the main causes for these inefficiencies:
a) High or low application rates due to lack of know-how on the infiltration characteristics of the
soil, resulting in surface ponding and runoff. This is especially true when high capacity drippers are
used on clay soils and on the other extreme, deep percolation if low capacity drippers are used on
sandy soils for a long time.
b) Lack of soil moisture monitoring and/or irrigation scheduling. Very few use sensors to check the
state of moisture in the soil and to decide when and how much to irrigate. Irrigation scheduling among
the farmers is still very primitive, usually based on past experience and by intuition.
c) Losses due to over irrigation during the early stages of growth when the crop uptake is low and
the root zone depth is shallow, resulting in excessive evaporation from the wetted soil and deep
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

160
percolation. Evaporation can be reduced by the use of mulching, but deep percolation cannot be
stopped.
d) Missing water metering. Most farmers irrigate on a time base rather than on water volume
needs.
e) Improper selection of crops according to water availability and water salinity.

Water Productivity

Water is a scarce resource that has a marked effect on the economics in any line of production
and economic sector. However, its use and eventual possible re-use is constrained by economic,
technological, physical, spatial and temporal conditions. Controlling its use will affect outputs and
therefore on revenue.

The concept of economic efficiency when applied to agriculture can be termed as water
productivity in agriculture. This is defined as the return obtained from one m
3
of water expressed in
terms of added value of agricultural production. This is a result of the currently accepted concept that
water has an economic value and as such any end user making use of water must perform in order to
effectively compete for water.

In the study Economic Considerations regarding markets for water in the Maltese Islands, by Delia
(2004), the share of water cost in the total production was obtained. The study was based on billed
consumption only. It transpired that even if one excludes the cost of water for irrigation (14.5 million
m
3
) from secondary sources, the biggest consumer of water for production purposes remained
agriculture with 2.4 % of the total input value. Electricity and the service and tourist industry come
after with 1.63 % and 1.46 % . (Table 1)

Table 1. Water requirements per LM100 of output (Source: National Statistics Office, National
Accounts of the Maltese Islands, Input Output Tables



Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

161
However, though water is an essential input, when one comes to decide what to produce, other
issues must be considered. Infrastructure, labour and marketing risks must also be considered.
Farmers need not only to consider the volume of water used to grow their crops but also should take
into account the net benefit they eventually get from the crops. Needless to say, this line of action can
be extended to the other water competitors.

In an exercise carried out by Borg Victor P.,1997, were evaluated the following considerations:
absorption per crop throughout the year,
the profitability per crop after accounting for prices fetched on the market, and
total costs

It should be noted that in actual fact, the prices which could be obtained to-date for agricultural
products were in a sheltered market environment i.e. quotas and tariffs are used to protect the local
produce. Moreover, the relatively low cost of water used for irrigation has also a marked effect on the
low total production costs.

In the above mentioned report by Borg, two water requirements scenarios were considered
namely:

Scenario 1: during the wet season, when the evapotranspiration is lower than rainfall and
most, if not all of the crop water requirements are met by rainfall (Fig. 4);

Scenario 2: during the hot and dry summer period when the evapotranspiration is higher
than rainfall, rainfall meets, if any, little of the crop water requirements and
hence, other sources of irrigation water are required (Fig. 5).

Data indicate that considering scenario 1, melons and watermelons would require the highest
amount of water to grow whilst onions the lowest. However, under scenario 2, the broad beans
consume more water than melons with onions still needing the least.

Fig. 4. Water consumption (in liters per kg of crop) according to Scenario 1


When considering costs and the net benefit per crop, under scenario 1, cauliflowers and onions
yielded the better value for the efforts inputted in the growing of the crops (Fig. 6). Strawberries were
seen to be grown at a loss. When the same analysis was made during the dry season (scenario 2),
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

162
courgettes, melons and watermelons, onions and cabbages and cauliflowers gave the best values
(Fig. 7).


Fig. 5. Water consumption (in liters per kg of crop) according to Scenario 2




Fig. 6. Percentage of low gross income for Scenario 1


When the consumed water and the net benefit per crop were integrated, it became evident that in
a low water requirement scenario (scenario 1 Fig. 8), the crop with most net benefit was onions.
However, when the water demand was high (scenario 2 Fig. 9), courgettes is the crop with the
higher net benefit.


Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

163


Fig. 7. Percentage of low gross income for Scenario 2



Fig. 8. Valuing irrigation water: water requirements and profit for Scenario 1

From these analyses, it was evident that prior to any water and agricultural management policies
are formulated, other studies specific to different time periods need to be carried out on different
commodities.


CONCLUSIVE REMARKS

The problems mentioned in this paper regarding the inefficiencies in the use of irrigation water
must be dealt with both on the national as well as on the local level. With regards the local level, the
farmers must start adopting irrigation practices which give more importance to water-use efficiencies
and hence water savings. This would eventually add up to less stress on the countrys need to
provide more and more water. Hence, the National Economy will benefit.

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

164



Fig. 9. Valuing irrigation water: water requirements and profit for Scenario 2

With the introduction of new and larger sewage treatment plants, more irrigation water will be
made available. The locations of these plants do not necessarily match or coincide with the water
requirements and proposing large distribution systems to meet all requirements is definitely not
recommendable. Transporting treated sewage effluent via road water tankers (which already
transport groundwater all over Malta and Gozo anyhow) would be a novel thing and definitely would
render a three-fold benefit namely:
involve the private sector in water distribution,
eliminate the need to allocate large amounts of capital in a water distribution system as well
as associated operating costs related to monitoring, pumping, water accounting, theft etc. and
last but not least by so doing most if not all water tankers which presently are used to
transport groundwater which is being extracted uncontrolled and illegally, will be used solely
to transport irrigation water originating from the treatment plants.

With regards to water productivity in agriculture, in areas such as Malta where water is scarce but
has a high economic value, the cultivation of crops which have low crop water requirements but have
a high economic return, should be encouraged. It should be noted that in Malta, approximately 56% of
the total water production for potable purposes is produced by very expensive desalination plants. It
has been estimated that 80% of this ends up in the local sewers. This implies that artificial water
available for irrigation has an even higher economic value. Moreover, as mentioned previously, the
excessive extraction of groundwater for irrigation and hence increased salinities has exacerbated
matters and the economic value of water is even higher. Thus, water productivity in agriculture is of
utmost importance.

Producers have to be made aware so as optimal use of water is practiced and that utilizing water
efficiently could be attractively be linked to profitable returns (value per unit output). Efforts should
concentrate at channeling water to be used on crops that generate the highest value added. This
could actually act as the incentive to eventually reduce water consumption and hence water
production costs through increased water use efficiency and water productivity.


Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

165

REFERENCES

Water and Agriculture in a Competitive Environment, Proceedings APS Seminar, 2001
Upgrading and Modernisation of San Antnin Sewage Effluent Irrigation System. Prepared by Haiste
International Ltd based on Dr. Josef Mitschoff
Economic Considerations regarding Markets for Water in the Maltese Islands, A report prepared for
FAO on water resources review for the Maltese Islands, Delia Carmen, 2004
WSC and Department of Agriculture internal reports and databank




Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

166
IRRIGATION RESEARCH RESULTS
IN THE SYRIAN ARAB REPUBLIC


A. Kaisi
*
, Y. Mohammad
**
, Y. Mahrouseh
***

*
Ministry of Agriculture and Agrarian Reforms, General Commission for Scientific Agricultural
Research, Damascus, Syria, E-mail: ak-gcsar@scs-net.org
**
Ministry of High Education, Dmascus, Syria, E-mail: yasser-m@scs-net.org
***
General Union of Peasants, Syria



SUMMARY Low-efficient traditional surface irrigation prevails over more than 80% of irrigated area
in Syria, resulting in a loss of more than 50% of available water resources at a time these resources
are facing growing pressure from all sectors. Accordingly, the Syrian government accorded an
increasing consideration to the implementation of water research programme applicable for most
strategic crops and farming, aiming at improving on-farm water management and, consequently,
improving WUE and reduction of water loss in agricultural sector. Research and technical results
showed that the use of modern irrigation methods on cotton, for example, leads to irrigation water
saving by 41% as compared to traditional irrigation, average yield 30%, and improvement of WUE
from 0.23 to 0.53 kg/m
3
. In economic point of view, modern irrigation methods were superior to
traditional irrigation, and drip irrigation ranked the first in terms of revenues and net profits per unit
area. Depending on technical and economic results of modern irrigation methods the government
made several decisions for transferring to modern irrigation. Areas transferred to modern irrigation
estimated to 260 thousand ha during the period 2000 2004.

Key words: Traditional surface irrigation, research programs, modern irrigation techniques, irrigation
water saving, water use efficiency, technical and economic results, irrigated area.


INTRODUCTION

Limited available water resources in Syria and their growing demand for different purposes
(agricultural industrial domestic) as well as the restricting constraints of use regulation imply the
necessity of making a radical change in the current low-efficient irrigation methods and systems, and
following-up ways and means that properly use water in agriculture at minimum level with high
productivity.

According to the above-mentioned considerations, the scientific plans included an implementation
of specialized research programmes on water management and use rationalization (water duty
irrigation methods and techniques salt-land reclamation and drainage methods the use of
mathematical model for studying salt movement water harvesting and spreading techniques
supplemental irrigation).

The experiments and research proved the feasibility of modern technique application in irrigation
projects in particular and agriculture in general, and at all levels regarding the following two key
indicators: (1) water saving and increase of available amount for horizontal agricultural expansion;
and (2) increase WUE kg/m
3
or SP/m
3
.

Research programmes dealt with water study of most strategic crops (cotton wheat maize
sugar beet) as well as water consuming vegetables and other kinds of fruit trees.

Due to the growing strategic importance of cotton over the recent years where cotton grown area
ranged 250 270 thousand ha i.e. 18 19% of irrigated area in Syria, and since cotton is a summer
crop consuming larger portion of irrigation water ranging 3 4 billion m3 i.e. 20 23% of annual
water supply and the prevailing traditional (flooding) irrigation which has the following disadvantages:
low application efficiency not more than 50% i.e. losses 50% at least; and high water table and soil
salinization as in Down Euphrates Basin as a result of over-irrigation and on-farm waste, Ministry of
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

167
Agriculture/General Commission for Agricultural Scientific Research gave special importance to
cotton irrigation research, aiming at improvement of water uses and irrigation efficiency, and
consequently water saving and cotton yield increase by the introduction of modern irrigation methods
and techniques.

Hence, this paper addressed the important cotton irrigation outcomes obtained at irrigation
stations in Hama, Aleppo, Raqqa, Deir Ezzor and Hassakeh (cotton producing- governorates) which
has a diversity in climate, soil and water quality.


IRRIGATION SYSTEMS AND PRACTICES APPLIED IN SYRIA

Traditional irrigation system

Traditional surface irrigation (flooding irrigation ) prevails in irrigated agriculture because it is very
early used since it is low-cost, easily implemented and doesnt need skilled labor or advanced
techniques. Traditionally irrigated lands are estimated to 82% of total irrigated area amounting more
than 1.4 million ha, considering that the total engineering efficiency of water uses, expressing the
relation between plant consumption from water for physiological processes and water withdrawal from
the source, is not more than 50% at best as the water is taken from irrigation systems (government &
private) by gravity or pumping from wells or rivers via earth canals unsuitable for water conveyance.
Gravity irrigated area of governmental irrigation systems constitutes 20% of total irrigated area.

Average water application per hectare is estimated to 14 thousand m3, and this average
considerably varies from one region/basin to another depending on WUE which is identified by
conveyance and delivery efficiency and on-farm irrigation techniques.

Project irrigation efficiency is related to its components. If it is possible to achieve canal
conveyance and delivery efficiency 80 95% , this figure will decline to 40 50% or less under
surface irrigation which has several negative features:
Wasting a large portion of irrigation water in conveyance and delivery canals.
On-farm irrigation water loss due to low field-irrigation application efficiency.
High water table level and soil salinity as in Down Euphrates basin.


Modern Irrigation System

It comprises modern irrigation techniques (sprinkler localized) in addition to improved surface
irrigation. Using modern irrigation methods started as individual initiatives, then the government gave
consideration to the introduction of these techniques and encouragement of farmers to possess and
use these techniques through the national programme for transferring to modern irrigation which was
started late 2000 (technical findings of irrigation methods and techniques). This programme aims to
transfer the whole irrigated area from traditional methods to modern ones during a specific period.
The government has developed several decisions for facilitating transfer process and ironing out the
financial and administrative constraints facing the implementation of this programme, together with
their promotion of research centers for preparing the necessary research plans, conducive to
reduction of traditional irrigated area and increase of modern irrigation area.


RESEARCH PROGRAMMES

The research programmes aim at improving water resource management, on-farm use
optimization and WUE. A research plan was prepared in 1990 including the following four research
programmes:
Water requirements, irrigation scheduling and supplementary irrigation, including the
following:
o Study of water requirements for different crops;
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

168
o Study of applying different irrigation rates on strategic crops;
o Study of supplementary irrigation for winter crops.
Irrigation methods and techniques, including the implementation of the following researches:
o Use of different irrigation techniques for irrigating field crops and vegetables, and
making comparison;
o Use of different irrigation techniques for irrigating fruit trees and making comparison.
Salinity and drainage, including:
o Study of salinity development phenomenon using different irrigation systems on
crops.
o Study of the effect of several levels of soil and water salinity on the yield of main
crops (wheat cotton sugar beet maize).
Rainwater harvesting and spreading, including:
o The use of contour farming for rainwater harvesting.
o Assessment of productive efficiency of collected rainwater.
o Runoff improvement in the Syrian steppe.
o Vegetation regeneration and rangeland rehabilitation.

A synthesis of main research programmes and corresponding irrigation and water use stations in
Syria is given in Figure 1.


TECHNICAL RESULTS

Research Results and Their Technical Feasibility

Four research programmes were implemented on a large number of field crops (including wheat,
cotton, sugar beet, maize), in addition to vegetables and fruit trees. In this document, a special
emphasis is given on cotton because:
It occupies 250 270 thousand ha, i.e. 18 19% of total irrigated area.
It is summer water-consuming crop.
It is a strategic crop playing a significant role in socio-economic relations.


Results of water requirement and irrigation scheduling research programme
Research implementation period: 1990 1995
Implementation place: Irrigation research stations in the following cotton growing
governorates (Hama Aleppo Hassakeh Raqqa Deir Ezzor), which have a diversity of
pedological and climatic conditions.
Experimental treatments:
o 60 70 80% of field capacity in addition to the control (as farmer irrigates).
Equations used to estimate ETo:
o Epan Ivanov Blaney Kriddle Modified Penman
Equipment and instruments
o Equipment for soil analysis (physical chemical).
o Evaporation basin.
o Neutron probe for soil moisture measurement.
o Various irrigation systems.
Varieties:
o Locally-produced varieties suitable for each governorates
Irrigation methods: Basin irrigation
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

169
Research Programmes
Water Harvesting
and Spreading
Drainage and
Salinity Irrigation
Methods
Irrigation Methods
and Techniques
Water Requirements
and Supplementary
Irrigations
Irrigation and Water Use Stations
Quneitra
Rural
Damascus
Homes
Hama
Idleb
Aleppo
Deir Ezzor
Hassakeh
Lattakia
Tartous
Deir Ezzor
Hassakeh
Mehasseh Research
Center For Natural
Resource Development
Land - Leveling Training Field days Revolving Fund
Technology Transfer



Fig. 1. Main research programs and corresponding irrigation and water use stations in Syria

Results as an average (years and regions):
Treatment of 80% field capacity was superior to 60 70% and control treatments.
Water requirement ranged 8600 9950 m
3
/ha.
Number of irrigations ranged 16 18 at irrigation rate of 400 600 m
3
/ha/irrigation.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

170


Results of irrigation methods and techniques research:

Programme objectives:
Study irrigation efficiency and water uses, using modern irrigation techniques (drip sprinkler
improved surface) vs. traditional surface irrigation.
Make technical and economic comparison between applied irrigation methods vs. traditional
irrigation.
Assess gained net income as compared to traditional surface.
Assess gained national production per unit area and additional national income resulting from
the use of modern techniques in irrigation.

Implementation areas:
The following governorates: Hama Aleppo Hassakeh Deir Ezzor
Implementation period: Two phases: (1991 1997) and (1996 2000).
Varieties: Research was conducted on adopted varieties adaptable for each governorates.

Experimental treatments:
Water treatment 80% of field capacity was adopted.
a. Drip irrigation:
Different models of drippers at different discharges 2 8 l/hr were adopted.
Filtrating pipes with discharge 7 8 l/hr/ linear meter.
Two engineering layouts:
o Irrigation line x plant row
o Plant row & irrigation line x two plant rows.
b. Sprinkler irrigation:
Different engineering designs (sprinkler spaces) at 9 x 9 , 9 x 12, 12 x 12 were adopted, and there
were three experimental treatments:
Sprinkler irrigation throughout the season;
Sprinkler irrigation: During germination, growth and flowering buds;
Surface: During flowering and nut formation.

It is very important to underline that sprinkler irrigation research was conducted in impermissible
ecological zones to avoid the outbreak of fungi diseases.
c. Improved surface irrigation:
It means the use of long furrow irrigation after laser land-leveling, and identification of proper
slope, soil nature and composition, and discharge of on-site available water. The applied treatments
were:
Furrow length 100 150 200 m.
Discharge 0.75 1 1.5 l/sec.
Traditional irrigation



Results
a. Statistical:

Localized irrigation research: Drip irrigation method (one line x one row) with discharge 4 l/sec was
the best at statistical function 1%.

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

171
Sprinkler irrigation research: Sprinkler irrigation method at sprinkler spacing 9 x 12 was the best in
statistical function 1%.

Surface irrigation research: Long-furrow irrigation method 200 m with discharge 1.5 l/sec was the
best in function of 1%.

b. Technical:
As an average for a number of years and regions (at country level):
the localized irrigation vis--vis traditional surface gave the following results:
o Applied water was 6113 m3/ha and 14446 m3/ha for localized and traditional
irrigation respectively.
o Yield increase 35%.
o Irrigation water saving 58%.
o WUE increase from 0.23 to 0.74 kg/m3.
o Application efficiency increase from 25 50% for traditional to 88.5% for localized.

Anyway, applying the research results over the whole irrigated area estimating to 250 thousand ha
grown with cotton for the year 2004 gave the following:
Total irrigated area 250 thousand ha
Total applied water by traditional irrigation 3.5 billion m
3

Total applied water by localized irrigation 1.5 billion m
3

Irrigation water saving 2 billion m
3

Irrigated area to be increased as a result of water saving 327 thousand ha
Foreseen production of additional area 1.48 million ton

Sprinkler irrigation vis--vis traditional irrigation led to the following:
Applied water by sprinkler was 8920 m3/ha, while it was 14446 m
3
/ha for traditional.
Yield increase 31%.
Irrigation water saving 38%.
WUE increase from 0.23 to 0.49 kg/m3.
Application efficiency increase from 45 50% for traditional to 78% for sprinkler.

Applying the results over the whole area gave the following:
Total applied water by traditional irrigation 3.5 billion m
3

Total applied water by sprinkler irrigation 2.225 billion m
3

Irrigation water saving 1.275 billion m
3

Additional area can be irrigated 143 thousand ha
Foreseen production of additional area 616 million ton

Improved surface irrigation vis--vis traditional surface gave the following results:
Applied water by improved surface was 10612 m
3
/ha, while it was 14446 m
3
/ha for traditional;
Yield increase 18%;
Irrigation water saving 27%;
WUE increase from 0.23 to 0.37 kg/m
3
;
Application efficiency increase from 45 50% for traditional to 62% for sprinkler.
Applying the improved surface irrigation over the whole cotton area:
Total irrigated area 250 thousand ha
Total applied water by traditional irrigation 3.5 billion m
3

Total applied water by improved surface irrigation 2.555 billion m
3

Irrigation water saving 2 billion m
3

Irrigated area to be increased as a result of water saving 94 thousand ha
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

172
Foreseen production of additional area 376 million ton

The effects of modern irrigation methods on the increase of Water Use Efficiency, yield and
irrigation water saving in relation to traditional surface irrigation on cotton are given in Table 1.

Table 1. The effect of modern irrigation methods on the increase of WUE, yield and irrigation water
saving vis--vis traditional surface irrigation on cotton
Statement Improved surface Drip Sprinkler
Traditional
surface
Actual irrigation m
3
/ha 10612 6113 8920 14446
Number of irrigations 12 17 14 11
Water saving % 27 58 38 -
Yield kg/ha 3952 4516 4376 3337
Yield increase % 18 35 31 -
WUE kg/m
3
/ha 0.37 0.74 0.49 0.23
Application efficiency % 62 88.5 78 45 50%
Hama 94 99 96 2000 95 2000 95 2000
Aleppo 94 99 91 2000 95 2000 91 2000
Deir Ezzor 94 99 91 97 91 96 91 97
Research
area
Hassakeh 94 99 91 - 97 92 - 98 92 - 98


ECONOMIC RESULTS

Economic analysis of the research results of applied irrigation methods and
techniques (improved surface sprinkler drip traditional surface method)

This analysis aims at the interpretation of research results to physical data by estimating the
revenues and costs per unit area as a results of productivity change and reduction of irrigation water
costs. The following parameters are analyzed:
Actual results of water requirements obtained through research;
Achieved yields vs. water requirements of executed research;
Actual status of production costs and product value international prices once the economic
analysis is made;
Calculation of irrigation water costs through economic analysis, considering irrigation water in
Syria is used according to the established laws. Irrigation water value includes water delivery
cost from the source to the field, and it was fixed according to:
o Actual cost paid by the farmer as irrigation fees (operation and maintenance) that is
estimated to 3500 SP/ha/year as a financial value, while the economic value of
irrigation water from governmental irrigation projects was calculated by the
depreciation of the capital used in irrigation projects according to the international
prices and repair maintenance of these projects.
o Financial and economic costs of irrigation water from other sources such as pumping
from rivers and pumping from groundwater at different depths according to the
current situation of groundwater were calculated financially and economically through
their actual costs, divided into:
Fixed costs (depreciation interests maintenance and repair)
Consumable costs: fuel, fats and labor wages.

The costs of well drilling and casing, pump unit and surface irrigation systems, if any, are included
in this calculation. Irrigation water value was accurately calculated using pumping units at capacity 40
and 60 m
3
/hr.
Land rent was at 15% from product value.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

173
For accurate analysis, the productivity for farmers adopted averagely 80% of research
results.

The study team wanted to give a clear picture on the economic analysis, comprising several
cases:
Results of several research stations.
Economic tests on four irrigation methods: traditional, post land-leveling surface,
sprinkler and drip irrigation.
Crop economic test on all irrigation sources, namely:
Irrigation from government projects.
Pumping from rivers.
Pumping at 5 100 200 m deep.
Excluding other affecting factors such as fertilization and varieties, considering they
are the same for all irrigation techniques for showing the impact of irrigation
techniques.
Economic test according to two kinds of pumping 40 and 60 m3/hr, which are widely
used in Syrian.
Economic test via the value of costs and revenues according to international prices.

The economic costs of one m
3
irrigation water by water source (river of well pumping), irrigation
method (traditional surface, improved surface, sprinkler and drip) and pumping unit capacity are given
in Table 2.

Table 2. Economic costs (in SP) of one m
3
irrigation water by water source, irrigation method
and pumping unit capacity
Well pumping
Statement
Pump
discharge
Government
Irrigation
projects
River
pumping
50
m
100
m
150
m
200
m
40 m
3
/hr 0.13 0.05 0.10 0.26 0.20 0.27 Depreciation and
maintenance of
structures and pumping
units (SP)/m
3

60 m
3
/hr 0.13 0.4 0.80 0.16 0.16 0.17
40 m
3
/hr 0.84 0.17 0.34 0.71 1.08 1.44
Fuel and oil/SP
60 m
3
/hr 0.84 0.23 0.47 0.98 1.46 1.95
40 m
3
/hr 0.84 0.75 1.49 2.50 3.44 4.22
Total cost of traditional
surface irrigation/SP
60 m
3
/hr 0.84 0.68 1.18 2.08 2.93 3.74
40 m
3
/hr 0.74 0.68 1.37 2.22 3.14 3.85
Total cost of improved
surface irrigation/SP
60 m
3
/hr 0.74 0.62 1.05 1.76 2.64 3.14
40 m
3
/hr 0.65 0.80 1.30 2.33 3.29 4.25
Total cost of sprinkler
irrigation/SP
60 m
3
/hr 0.65 0.80 1.22 2.13 2.98 3.80
40 m
3
/hr 0.85 1.11 1.55 2.60 3.55 4.51
Total cost of drip
irrigation/SP
60 m
3
/hr 0.85 1.11 1.47 2.40 3.25 4.39




Economic comparison of drip irrigation to traditional surface irrigation

Drip irrigation trials were carried out at four research station (Hassakeh Aleppo Deir Ezzor
Hama), focusing on cotton growing in Syria.

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

174
Average results at these stations for the entire period showed that applied water was 6113 and
14446 m
3
/ha and yield was 4516 and 3337 kg/ha for drip and traditional irrigation respectively. And
this reflected on the practical situation of economic costs and revenues and the below results were
obtained:

Results of economic analysis

Economic results do not considerably differ from financial results in terms of comparing drip
irrigation with traditional surface. A synthesis of results obtained per one hectare are given in Table 3.

Table 3. Production value, cost value, net profit and cost/profit ratio for tradition and drip irrigation
methods in Syria
Irrigation from
government projects
Pumping at 100m
deep
Pumping at 200 m
deep
Statement
40 m
3
/hr 60 m
3
/hr 40 m
3
/hr 60 m
3
/hr 40 m
3
/hr 60 m
3
/hr
Production value
SP/ha traditional
irrigation
133480 133480 133480 133480 133480 133480
Production value
SP/ha drip
irrigation
180640 180640 180640 180640 180640 180640
Cost value SP/ha
traditional irrigation
78929 78929 113310 96842 114321 112154
Cost value SP/ha
drip irrigation
74202 74202 100549 88873 101770 106527
Net profit SP/ha
traditional irrigation
54551 54551 20170 36638 19159 21326
Net profit SP/ha
drip irrigation
106436 106436 80091 91767 74113 87870
Cost/profit ratio
traditional irrigation
69 69 30 38 17 19
Cost/profit ratio
drip irrigation
143 143 101 103 70 78


Economic comparison of sprinkler irrigation vs. traditional surface

Examination of trial results at country level showed that water applied is estimated to 8920 and
14446 m
3
/ha and yield 4376 and 3337 kg/ha under the same conditions for sprinkler and traditional
irrigation respectively. Adoption of these results, prevailing prices, costs and revenues for traditional
irrigation and sprinkler method are estimated and presented in Table 4.

Table 4. Production value, cost value, net profit and profit/cost ratio for tradition and sprinkler irrigation
methods in Syria
Traditional irrigation from
government projects
Sprinkler irrigation from
government projects
Statement
40 m
3
/hr 60 m
3
/hr 40 m
3
/hr 60 m
3
/hr
Increase in
sprinkler
irrigation %
Production value SP 102613 102613 134562 134562 8
Cost SP 73381 73381 74816 74816 2
Net profit SP 29232 29232 59746 59746 51
Profit/cost SP % 40 40 80 80 40


Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

175
In general, economic superiority of sprinkler irrigation method in respect to traditional irrigation is
evident as compared to traditional irrigation. The following economic results are analyzed also by
means of production value and net profit considering different pumping depths and the results of this
analysis are summarized in Table 5.




Table 5. Production value and net profit for traditional and sprinkler irrigation methods considering
different pumping depths

Government
irrigation projects
Pumping
50 m
Pumping
100 m
Pumping
200 m
Statement
40
m
3
/hr
60
m
3
/hr
40
m
3
/hr
60
m
3
/hr
40
m
3
/hr
60
m
3
/hr
40
m
3
/hr
60
m
3
/hr
Production value
SP/ha traditional
irrigation
133480 133480 133480 133480 133480 133480 133480 133480
Production value
SP/ha sprinkler
irrigation
175040 175040 175040 175040 175040 175040 175040 175040
Net profit SP/ha
traditional irrigation
54551 54551 45161 49640 30751 36638 19159 21326
Net profit SP/ha
sprinkler irrigation
100572 100572 88976 89690 79788 81572 62662 66676


Improved surface vs. traditional irrigation

Irrigation method in land leveled lands was superior to that of non-leveled lands, and this
superiority came from two: First, reduction of applied water per unit area, and the second productivity
increase. The results of economic analysis showed clear superiority of improved surface over
traditional irrigation as it is reported in Table 6 for different pumping depths.


Table 6. Yield, production value, irrigation water use, irrigation value, costs, net profit and profit/cost
ratio for traditional and improved surface irrigation methods considering different pumping
depths
Government
irrigation projects
Pumping
50 m
Pumping
100 m
Pumping
200 m
Statement
Pump
capacity
(m
3
/hr) Tradit. Improv. Tradit. Impro. Tradit. Impro. Tradit. Improv.
Yield kg/ha 3194 3952 3194 3952 3194 3952 3194 3952
Production
value SP
127760 158080 127760 158280 127760 158080 127760 158080
Irrigation
water m
3
/ha

14446 11616 14446 11616 14446 11616 14446 11616
40 12135 8538 21525 15914 36115 25788 47527 44722
Irrigation
value SP/ha
60 12135 8538 17045 12197 30048 20444 45360 36474
40 78929 77293 88319 93207 102909 103081 114321 113015
Costs SP/ha
60 78929 77293 83840 89480 96842 97737 112154 113767
40 48831 80787 39441 64873 30751 54999 13439 45065
Net profit
SP/ha
60 48831 80787 43920 68600 30918 60343 15606 44313
40 62 105 45 70 30 53 12 40
Profit/cost %
60 62 105 52 77 31 62 14 49

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

176


Economic comparison among different irrigation techniques on cotton:

It is essential to make a comparison among different irrigation methods by different sources, so a
table including this comparison was prepared for purpose.



Relative Significance of Irrigation Techniques at Local Prices:

Through data on relative significance of irrigation techniques using financial analysis, the following
can be concluded:
a. Drip irrigation ranked the first for the irrigation technique economics in terms of
revenues, net profits per unit area and profit/cost ratio, followed by sprinkler irrigation
for all prevailing irrigation sources in Syria.
b. Differences among irrigation techniques increase by increasing irrigation expenses
i.e. high pumping level as traditional surface costs increases as a result of using
higher amount of water.
c. Net profit of irrigation techniques is increasing forward with high irrigation costs as
compared to traditional irrigation. For example, profit ration of drip to traditional was
153% from government irrigation projects, and it increased to 325% from well
pumping at 50 m.


CONCLUSIONS AND RECOMMENDATIONS

This paper shows that low-efficient traditional surface irrigation prevails over more than 82% of
irrigated area. Depending on the technical and economic findings of modern irrigation methods, the
government, starting from mid-2000, took several decisions toward the movement from traditional
irrigation to modern one and the necessary facilities were provided for this movement besides its
support that made irrigated area exceeding 260 thousand ha. Several research programmes are
applicable for most strategic crops and different farming in Syria, aiming, as a whole, at improving on-
farm water management, water use efficiency and water loss reduction in the agricultural sector. This
works focuses on irrigation research on cotton as it is a socio-economic crop occupying a large part
of irrigated area and one of the largest water-consuming crops. It also referred to the technical and
economic results that can be applied in extensive farming in farmers' lands, and these results showed
the superiority of modern irrigation methods over traditional ones in the following fields:
WUE increased from 0.23 0.25 kg/m
3
/ha for traditional surface to 0.37, 0.74 and 0.49
kg/m
3
/ha for improved surface, drip and sprinkler irrigation respectively.
yield increased by 18 25%, 35% and 31% for improved surface, drip and sprinkler irrigation
respectively, as compared to traditional surface. Irrigation water saving ranged 24 26% for
improved surface irrigation.
Drip irrigation ranked the first in terms of revenues, net profits per unit area and profit/cost
ratio. In case of using pumps at discharge 40 m
3
/hr, the cost/profit ratio was:
o Government irrigation projects: 143% for economic analysis.
o Well pumping at 50 m: 116% for economic analysis.
o Well pumping at 100 m: 101% for economic analysis.
o Well pumping at 200 m: 70% for economic analysis.
Once pumps at discharge of 60 m
3
/hr are used, the cost/profit ratio was as follows:
o Government irrigation projects: 143% for economic analysis.
o Well pumping at 50 m: 117% for economic analysis.
o Well pumping at 100 m: 103% for economic analysis.
o Well pumping at 200 m: 78% for economic analysis.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

177

The following important recommendations can be drawn from this work:
1. There is a necessity to gear to the conjunctive use of groundwater as far possible by establishing
regular irrigation projects on groundwater, conducive to water use rationalization on one hand,
and stakeholders' potential control of groundwater abstraction on the other.
2. Encouraging the establishment of private and common companies in the field of laser land-
leveling is of fundamental importance in order to improve the water use efficiency of surface
irrigation method.
3. It is necessary to encourage the local manufacturing of modern irrigation equipment according to
international standards.


REFERENCES

Agricultural Statistical Abstracts 1970 - 2004, Directorate of Statistics and Planning, Ministry of
Agriculture, Damascus, Syria.
ANRR 's Research Plan 1990 2003, GCSAR, Syria.
Terminal Report of Improved Water Resource Project SYR/90/001, Directorate of Irrigation and Water
Use, 1997.
Terminal Report of Regional Supplementary Irrigation Project, Directorate of Irrigation and Water Use,
1996.
Research Programmes at Irrigation Research Stations in Hama, Aleppo, Raqqa, Hassakeh and Deir
Ezzor, 1999 2004.
National Plan of Movement toward Modern Irrigation/Notes, 2000 2002.
Proceedings of the Governmental committees, 2005.
































Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

178









WATER USE EFFICIENCY IN TURKEY



R. Kanber
*
, M. Unlu
*
, E.H. Cakmak
**
, M. Tuzun
***

*
Univ.of Cukurova, Agric. Struc. and Irr. Dept., Adana, Turkey
E-mails: kanber@cu.edu.tr, munlu@cu.edu.tr
**
Middle East Technical University METU, Turkey, E-mail: cakmake@metu.edu.tr
***
South-Eastern Anatolia Project Regional Development Administration GAP-RDA, Turkey



SUMMARY - Efficiency in the use of water for irrigation consists of various components and takes
into account losses during storage, conveyance and application to irrigation plots. Identifying the
various components and knowing what improvements can be made is essential to making the most
effective use of this vital but scarce resource in Turkeys cultivated areas. Enhancements in water use
efficiency (WUE) depend on productivity gains, depicted by consistent increases in outputs per unit
inputs and the irrigation techniques. Improved water use efficiency in agriculture is important not only
for water conservation, but for obtaining high yields. Modern irrigation technologies, such as sprinkler
and micro irrigation, are highly efficient and have the potential to increase yields substantially.
Unfortunately, the high costs may prevent small farmers from using the systems. Thus, the use of
modern irrigation techniques may be restricted to production of high value crops so that the systems
may be financially viable. In this work, it is given some experimental results on water use efficiency of
cotton, orange, Lemon, strawberry, watermelon using different irrigation methods in the
Mediterranean and Southeastern regions of Turkey.

Key words: Water Use Efficiency, Irrigation Methods, Cotton, Orange, Strawberry, Turkey


INTRODUCTION

In semiarid areas of the world, soil water deficits and excessively high temperatures are probably
the most common yield-limiting factors in crops. To improve yields, many scientists are seeking
means of reducing the effects of drought and making agricultural water use more efficient. Water use
efficiency (WUE) generally describes crop production per unit of water use during the growing
season. Irrigation accounts for up to 80% of consumptive use of fresh water in Turkey. Most feasible
water development projects have already been undertaken. As urban populations grow and industrial
and municipal water needs increase, a decrease in irrigation consumption is required to meet needs.
At the same time, irrigation accounts for large percentages of agricultural production and thus food.
There is a clear need for greater understanding of crop water use and water use efficiency as affected
by irrigation method, climate, variety, soils, water quality and management. Improvements in water
management and water use efficiency are key to reducing consumption while maintaining production.
Current research efforts in Turkey focus on improved understanding of crop water use and its
prediction for use in water management and irrigation scheduling; improved understanding of key
indicators of crop water status and their use in irrigation scheduling and control; and improved
technologies for irrigation scheduling and control for improved water use efficiency.

Techniques and inputs that improve WUE have been extensively researched and tested on
research stations and at farm or field levels. The dry areas comprise a region of heterogeneous land
and variable weather. It is expected that variability of experimental results caused by uncontrolled
factors such as physical and biotic environment, will greatly exceed the variability due to controlled
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

179
factors. Therefore, localized research results alone can not produce an optimal strategy for
maximizing WUE. Dealing with water issues on a basin-wide level will present many problems and
constraints that are not so apparent on the farm or research station level. This holistic approach has
not yet been given enough attention. When developing water-harvesting projects runoff is often
intercepted at the upstream reaches of the catchments, thus depriving potential downstream users.
The lack of balance in these systems for equitable water allocation among upstream and downstream
users often causes social, economical and environmental problems. Moreover, in many areas socio-
political considerations override any optimization of the management of these resources (El- Baltagy,
1997).

The introduction of technologies to increase yield and/or to improve WUE at the farm level not only
affect water availability at other parts of the basin in term of quantity, but - maybe more important - the
quality as well. This has been reflected in the existing literature, but no approach has been found for
assessing the consequences of implementing these technologies and optimizing their use at the
macro level. As water becomes scarcer, the need to have a methodology for optimizing water use at
the macro level becomes more important. Designing research work at a basin level in combination
with appropriate agro-ecological characterization and within a modeling prospective would help in
making the results less site specific, more transferable, and environmentally safer (El- Baltagy, 1997).

Unless innovative solutions that satisfy these considerations and optimize the use of water are
developed, overexploitation of water resources will continue to threaten the sustainability of the
groundwater-based agriculture development.

Turkey is taking the lead in developing a system-wide water-management research initiative for
improving WUE in the dry areas. This goal will be achieved by implementing strategies and
techniques that have been developed and approved as effective on-farm means for increasing crop
production, saving water and potentially increasing water-use efficiency. Among these proven
effective WUE-boosting techniques are supplemental irrigation and water harvesting. Other available
strategies are related to crop varietals selection, cropping pattern, cultural practices and farm inputs.
These techniques and inputs have been tested on research station and/or farm levels. The challenge
is to extend the available on-farm techniques for improving water use efficiency to the basin-wide
level. To this end, a number of recommendations may be made.

In the following section of the study, brief information about general indicators of geography and
climate, soil and water resources, are presented. In the third section, the concept of water use
efficiency and results for different regions and different studies are investigated. Methods used and
comparable results for different conditions, systems and regions are pointed out in order to evaluate
the water use efficiency status in Turkey.


WATER USE EFFICIENCY CONCEPT

Due to rapid increase in the worlds population, urbanization, income and consumption choices,
and as a result of this process, because of demand for water and deterioration of its quality, today lots
of country is faced with important water problems. About 18.25 m
3
per year covers basic human water
requirements such as drinking, sanitation, bathing and food preparation. It is estimated that over a
billion people had access to less than 50 liters of water a day. According to 2000 year, agriculture
accounts for about 75 %, industry uses 15% and municipal domestic uses 10%. It is projected that
one third of the countries in water-stressed regions of the world are expected to face water shortages
in this century. Decreases of water usage in agriculture can be provided only by increasing the
efficiency it means that irrigation water use will result in large water savings.

Agricultural irrigation is important in terms of the increasing the productivity in Turkish arable lands,
accelerating the economic growth and decreasing the migration from rural to urban areas. Thus, the
efficiency of water use and water conservation in agriculture comes into prominence.

The term efficiency is generally understood to be a measure of the output obtainable from a given
input. Irrigation and water-use efficiency can be defined in various ways, depending on the nature of
the inputs and outputs considered. For example, one may attempt to define as an economic criterion
of efficiency the financial return in relation to the investment in the water supply. One problem is that
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

180
costs and prices fluctuate from year to year and vary widely from place to place. Another problem is
that some of the costs of irrigation, and certainly some of the benefits, can not easily be quantified in
tangible economic or financial terms, especially in places where a market economy is not yet fully
developed. Often, only the short-term costs and immediate benefits are discernible, whereas the long
term advantages or disadvantages are unknown a priori. How can be assign monetary value, fro
instance, to the possibility that an irrigation project might save the population of a region from the dire
effects of a drought if the frequency or probability of droughts of varying degrees of severity can not
be determined?

Quite different from the strictly economic criterion of efficiency is the physiological one i.e. the plant
water use efficiency. The criterion here is the amount of dry matter produced per unit volume of water
taken up by the plant from the soils. As most of the water taken up by plants in the field is transpired
(in arid regions-99% or more) while generally only a small fraction is retained, the plant water use
efficiency is in effect the reciprocal of what has long been known as the transpiration ratio, defined
as the ratio of the amount of water transpired to the amount of dry matter produced (tons per ton).
That ratio can run as high as 500 or even 1000 in regions and seasons of high evaporability.

The technical efficiency is what irrigation engineers call water use efficiency. It is generally
defined as the net amount of water added to the root zone divided by the amount of water taken from
some source. As such, this criterion of efficiency can be applied to complex regional projects, or to
individual farms, or to specific fields. In each case, to difference between the net amounts of water
added to the root zone and the amount withdrawn from the source represents the seepage and
evaporative loses incurred in conveyance to the crop, as well as the loses due to deep percolation
below the root zone within the field and to runoff from the field.

From the point of view of water use, some large-scale irrigation projects operate in an inherently in
efficient way. In many of the surface irrigation schemes, one or few farms may be allocated large
flows representing the entire discharge of a lateral channel for a specified period of time. Where water
is delivered to the consumer on a fixed schedule and charges are imposed per delivery regardless of
the actual amount used, customers tend to take as much water as they can. This often results in over
irrigation, which not only wastes water but also causes project-wide problems connected with the
disposal of return flow, water logging of soils, leaching of nutrients, and elevation of the water table
requiring expensive drainage. Although it is difficult to arrive at reliable statistics, it has been
estimated that the average irrigation efficiency in such schemes is probably well below 50% (and may
be as low as 30%). Since it is a proven fact that, with proper management, it is possible to achieve
irrigation efficiencies as high as 85% or even 90%, there is an obviously much room for improvement.

Particularly difficult to change are management practices which lead to deliberate waste not
necessarily because of insurmountable technical problems or lack of knowledge but simply because it
appears more convenient, or even more economical in the short run, to waste water rather than to
apply proper management practices of strict water conservation. Such situations typically occur when
the price of irrigation water is lower than the cost of labor or of the equipment needed to avoid over
irrigation. Very often the price of water does not reflect its true cost but is kept deliberately low by
direct or indirect government subsidy, which can be self-defeating.

Open and unlined distribution ditches are used, uncontrolled seepage and evaporation, as well as
transpiration, can cause major losses of water. Even pipeline distribution systems do not always
prevent loss. Leaky joints resulting from poor workmanship, corrosion, ill-maintained valves, or
mechanical damage by farm machinery may cause large losses. Sometimes the damage is not
immediately apparent, as when a buried pipe under pressure fails at night, with no one in attendance.

Surface runoff resulting from the excessive application of water ideally should not occur. Sprinkler
irrigation systems should be designed to apply water at rates which never exceed soil infiltrability. In
the case of gravity irrigation systems, however, it is often virtually impossible to achieve uniform water
distribution over the field without incurring some runoff (tail water). Only when provision is made to
collect irrigation and rainwater surpluses at the lower end of the field and to guide them as controlled
return flow can this runoff water be considered anything but a loss.

Evaporative losses associated with water application include any evaporation from open water
surfaces or border checks or furrows, evaporation of water droplets during their flight from sprinkler to
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

181
ground surface, wind drift of droplets away from the target area, and evaporation from wetted crop
canopies (foliage) or from the wet soil surface immediately after irrigation. While some of these water
losses cannot be totally eliminated, most can be greatly reduced. Transpiration by weeds is also
largely preventable loss.

In the open field, little can be done to decrease transpiration by the crop if the conditions required
for high yields are to be maintained. Attempts to use of windbreaks to control wind movement above
and through a crop stand does not always produce the desired effect economically.

It appears at present that the greatest promise for increasing water use efficiency lies in allowing
the crop to transpire freely by alleviating any water shortages while at the same time controlling all
other processes of water loss and obviating the other environmental constraints to attainment of the
full productive potential of the crop. This is particularly important in the case of the new and superior
varieties which can attain their full potential yields only if water stress is eliminated and such other
factors as soil fertility, aeration, salinity, and soil cultivation are optimized. Plant diseases and pests
may depress yields without a proportionate decrease in transpiration and water use. All management
practices can thus influence the efficiency of water use in irrigation, so the practice of irrigation should
not be regarded merely as the provision of water to thirsty crops, but more comprehensively as an
integrated production system designed to maximize the efficiency of land, water, manpower,
machinery, and energy utilization.

In many parts of the world, far greater returns can be obtained from intensification of production in
existing irrigation systems, i.e. by improving methods of water, soil, and crop management, than by
building ever new irrigation projects on the basis of the same antiquated and inflexible design. Since it
is difficult to convert traditional systems to modern irrigation scheduling, it is important to make
decisions affecting irrigation frequencies and quantities in the early stages of planning new projects,
before the distribution system is designed and installed and future irrigators are thereby locked into an
inefficient pattern (Hillel, 1987).


PRINCIPLES OF WATER USE EFFICIENCY

The term "water use efficiency" originates in the economic concept of productivity. Productivity
measures the amount of any given resource that must be expended to produce one unit of any good
or service. Thus, for example, labor productivity in a steel mill would be the amount of labor required
to produce a tone of crude steel. In a similar manner, water productivity might be measured by the
volume of water taken into a plant to produce a unit of the output. In general, the lower the resource
input requirement per unit, the higher the efficiency. Throughout this book, improved water use
efficiency in its simplest form means lowering the water needs to achieve a unit of production in any
given activity (Donald, 2000).

In an environmental resource context, however, the efficiency concept must be extended to
include considerations of quality. Any effort to improve water use efficiency should be consistent with
maintaining or improving water quality. Taking both quantity and quality into account, therefore, the
following definition applies:

Water use efficiency includes any measure that reduces the amount of water used per unit of any
given activity, consistent with the maintenance or enhancement of water quality.

Water use efficiency is closely related to, and in several cases overlaps, other basic concepts of
current environmental resource management. The best established of these related concepts,
perhaps, is water conservation. Water conservation is any socially beneficial reduction in water use or
water loss. Put in this manner, water use efficiency is of central importance to conservation. At the
same time, the conservation definition suggests that efficiency measures should, in addition to
reducing water use per unit of activity, make sense economically and socially.

Finally, water use efficiency has a clear role to play in sustainable development, in other words,
the use of the earth's resources by today's inhabitants while assuring that future generations have
sufficient capacity to meet their own needs. Improving the efficiency of resource use comprises one
means of meeting sustainable development goals.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

182

The importance of efficiency in water use clearly varies across regions and nations, as well as
through time. Geographically, for instance, water availability will condition the manner in which use
patterns develop. Other things being equal, arid and semi-arid regions require a greater efficiency of
water use than humid ones. But simple geographical patterns mask several equally important factors.
Economic conditions will often lead to greater or lesser water use efficiency. Many regions in the
world have been assisted in their development through public financing of water development. While
the benefits or costs of such projects in efficiency terms are often debatable, the main point here is
that economic factors can influence water use efficiency. Further, in some cases, where water
developments have supported new settlements in dry areas, industrial processes and technologies
that use water more efficiently than elsewhere may develop. An example would be the development
of recirculation technologies or process changes. Social conditions may also be important in
examining the efficient use of water resources. The literature reveals many areas where public
education has led to conservation and better use of available water supplies.

Water-use efficiency measures are commonly used to characterize the water-conserving potential
of irrigation systems. Alternative efficiency measures reflect various stages of water use and levels of
spatial aggregation. Irrigation efficiency, broadly defined at the field level, is the ratio of the average
depth of irrigation water beneficially used (consumptive use plus leaching requirement) to the average
depth applied, expressed as a percentage. Application efficiency is the ratio of the average depth of
irrigation water stored in the root zone for crop consumptive use to the average depth applied,
expressed as a percentage.

Crop-water consumption includes stored water used by the plant for transpiration and tissue
building, plus incidental evaporation from plant and field surfaces. Leaching requirement, which
accounts for the major difference between irrigation efficiency and application efficiency, is the
quantity of water required to flush soil salts below the plant root zone. Field-level losses include
surface runoff at the end of the field, deep percolation below the crop-root zone (not used for
leaching), and excess evaporation from soil and water surfaces. Conveyance efficiency is the ratio of
total water delivered to the total water diverted or pumped into an open channel or pipeline,
expressed as a percentage. Conveyance efficiency may be computed at the farm, project, or basin
level. Conveyance losses include evaporation, ditch seepage, operational spills, and water lost to
non-crop vegetative consumption.

Project efficiency is calculated based on farm irrigation efficiency and both on- and off-farm
conveyance efficiency, and is adjusted for drainage reuse within the service area. Project efficiency
may not consider all runoff and deep percolation as loss since some of the water may be available for
reuse within the project.


RESULTS ON SOME IRRIGATION EXPERIMENTS

For activities of WASAMED which is a thematic network in Mediterranean countries, the results
from all research activities on irrigation carried out in Turkey have been tried to collect, however, the
results of all studies conducted, published data and other activities could not be obtained, because of
the deficiencies in our archives system, assessment to all the conducted studies is limited. Statistical
aspects of the collected results for the last 10-15 years are given with aim to give information and
knowledge in experiences on irrigation science and assessment of past and existing experiences and
identifications of relevant gaps and problems in Turkey.

Some studies conducted, in the Mediterranean and Southeastern regions of Turkey, under open-
field conditions, many crops, such as strawberry, lemon, orange, and banana, were irrigated using
various methods, including the drip technique. The results of these experiments are presented in
Table 1.

The first crop, used for drip irrigation experiments, was strawberry in the Mediterranean region of
Turkey. The first half of the 1970s was the time of adaptation of drip irrigation techniques in Turkey.
Kanber and Dervis (1975) conducted the preliminary work on drip irrigation for strawberry. This
experiment was very primitive and made the drip-system irrigation network using ordinary techniques.
However, it involved taking no scientific results, only a demonstration for the farmers. Another
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

183
experiment on strawberry took place in the Irrigation Engineering Department of Cukurova University,
using two cultivars, Pochantos and Aliso. The yields of strawberry in the Adana experiments did not
show any difference between irrigation methods (Tekinel et al. 1984), but the marketable yield from
trickle was higher than that from other methods. An experiment in Tarsus was, again, on strawberry
(Kanber et al. 1986). However, in this study, the effects of irrigation methods on yield were statistically
different. Water use from trickle methods was 34% more than that in furrow. Yield from trickle
irrigation was 12% more than in furrow. However, unmarketable yield decreased 17% with drip
irrigation. WUE in trickle irrigation is higher, so it may be argued that higher yield was obtained from
trickle method with less water (Table 1).

Table 1. Irrigation water use and yield with various irrigation methods (Source: Kanber and Dervis
1975; Tekinel et al. 1984; Kanber et al. 1986; Cevik et al. 1982; Ozsan et al. 1983 )
Crop Irrigation method Irrigation water
(mm)
Yield (t/ha) IWUE
Kg/da-mm
Furrow 7.5
Drip - 11 -
Strawberry
(Adana)
Sprinkler 9
Furrow 400650 12-13 3020 Strawberry
(Tarsus)
Drip 300400 13-15 43
Furrow 460575 24.5-36.7 5060
Drip 151299 20.1-37.3 130120
Orange (Adana)
Sprinkler 344430 31.0-42.4 90100
Drip 115445 5.9-7.6 5020 Orange (Tarsus)
Sprinkler (under tree) 670844 13.6-13.3 2016
Furrow 10021336 2.2-2.8 2
Drip 184277 2.2-2.5 109
Sprinkler (over tree) 1001 2.5-3.4 23
Lemon
Sprinkler (under tree) 10641463 2.5-2.8 21

In orange experiments (Table 1), we used Magnum Bonum cultivar. Experimental trees in Adana
were 25 years old. The results from Adana indicated that sprinkler irrigation increases the yield in
compression with the drip and furrow methods (Cevik et al. 1982). Trickle irrigations results were
insignificant because of the root-development properties resulting from surface irrigation carried out
for a long time. The values for WUE of sprinkler method were not higher than those for drip, and they
varied 01.0. The Tarsus experiments used young trees, observing the growth of trees in trials during
19781988. The effects of irrigation methods on growth were found to be insignificant (Eylen et al.
1988). Only 1 or 2 years after planting, trickle method increased the growth. WUE values were low
with both irrigation methods. However, they were higher than those from Adana. In contrast, trickle
systems were profitable only for areas of more than 50 decares.

Ozsan et al. (1983) also studied effects of irrigation methods on the lemon yield and growth in the
same citrus irrigation program. Although tree growth was not affected by the irrigation techniques, the
trickle irrigation positively increased the pomological properties. The highest yield resulted from use of
the over-tree sprinkler, and the lowest resulted from planting in furrow. WUE was highest in drip
irrigation.

After 1988, in the Mediterranean and southeastern regions of Turkey, work started on drip
irrigation of some important crops, such as cotton. Cukurova is the place where cotton growing is the
widest spread in Turkey.

Yavuz (1993), conducted a detailed experiment to determine suitable irrigation methods for cotton.
Yavuz tested three irrigation methods, namely, furrow, drip, and sprinkler.

In addition, this study included various management techniques for each irrigation method. For
instance, furrow irrigation comprised ponded alternative furrows (PAF), free-end furrows (FEF), and
ponded continuous-flow furrows (PCF). Drip irrigation used two emitter spacings (30 and 60 cm) and
two planting techniques, traditional and double row in a single planting bed. Accordingly, the study
used four drip irrigation treatments. In sprinkler irrigation, Yavuz evaluated various final irrigation
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

184
dates and levels. Soil water observations in the free-end furrows determined the irrigation times for
furrow and sprinkler methods. Yavuz calculated the amount of irrigation water to use in the plots on
the basis of the amount needed to replenish the soil water deficit to field capacity. Drip irrigation used
an irrigation interval of 7 days. Yavuz calculated the amount of the irrigation water for drip plots from
cumulative free water evaporation, measured from class-A pan between the irrigation intervals.
Yavuz (1993) has tested different surface irrigation methods and compared to performances
obtained from drip and sprinkler irrigation methods. Table 2 shows some efficiency components such
as application (Ea), requirement (Er), infiltration (Ei), tail water ratio (TWR), deep percolation ratio
(DPR), uniformity of Christiansen coefficient (UCC), distribution uniformity (DU), and water use
efficiency (WUE) calculated for different irrigation methods.

Table 2. Performances of various cotton irrigation methods (Source: Yavuz, 1993).
Method Zi Ea Er Ei TWR DPR UCC DU WUE
PAF 375 80 81 100 20 89 90 0.49
FEF 653 67 69 100 33 94 62 0.40
PCF 722 77 75 100 23 91 94 0.35
SI 834 92 85 100 8 100 100 0.27
0.39
Da Eu PELQ AELQ Dn
DTd
2
8 90 91 78 7 0.39
DTd
1
16 82 74 71 12 0.36
DDd
2
9 70 63 61 6 0.54
DDd
1
15 76 68 66 11 0.43
AELQ, application efficiency; Da, average application depth; DD, double-row drip irrigation; Dn,
minimum application depth; DT, traditional drip irrigation; DU, distribution uniformity; Ea, application;
Ei, infiltration; Er, requirement; Eu, emission uniformity; FEF, free-end furrows; PAF, ponded
alternative furrows; PCF, ponded continuous-flow furrows; PELQ, potential application efficiency; SI,
sprinkler

In the Table 2, infiltrated water estimated from net infiltration opportunity time, which were obtained
flow advance and recession data during irrigation event, was also given. The irrigation methods
differed in their performances. The highest application efficiency was in sprinkler irrigation, at 92%.
The ponded alternate furrow followed sprinkler irrigation, with 80%. The application efficiency of FEF
was 67%, an acceptable value. All irrigation methods had acceptable efficiencies for cotton irrigation.

Cetin (1997) conducted another detailed experiment on irrigation of cotton in SanlurfaHarran
Plain to determine the effects of various irrigation methods (furrow, stationary sprinkler, stationary
drip, mobile sprinkler, mobile drip, and low-energy precision application [LEPA]) and irrigation water
levels on yield, quality, and WUE for cotton between from 1991 to 1994. Cetin estimated the applied
water for the methods of drip and furrow using cumulative pan evaporation of 50 5 mm and 100
10 mm at varying time intervals and adjusted coefficients of 0.61.8 as increased 0.3 increment (for
furrow and drip). For sprinkler irrigation, Cetin calculated the amount of water given to the plots close
to lateral line, using 100 10 mm cumulative pan evaporation measured in a time interval and
coefficient of 1.8. The results showed that these irrigation methods have significant effects on the
yield. Stationary drip gave the highest cotton yield, and the lowest yield from stationary sprinkler
(Table 3). Amount of irrigation water stands to cotton yield in a quadratic relation (Figure 1). This
figure shows that the yield increased to a peak and then decreased with irrigation water. However,
WUEs in drip irrigation were high among the treatments at all water levels. The lowest values were
from furrow methods at all water levels.

Table 3. Average irrigation water, yield, and WUEs for various cotton irrigation methods (Source:
Cetin, 1997).
Furrow Stationary sprinkler Stationary drip IWUE (kg/ha per mm)
IR
(mm)
Yield
(kg/decare)
IR
(mm)
Yield
(kg/decare)
IR
(mm)
Yield
(kg/decare)

Furro
w Sprinkler Drip
624 254 328 216 341 207 4.07 6.59 6.07
937 363 735 291 619 346 3.87 3.96 5.59
1248 385 1106 328 898 438 3.08 2.97 4.88
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

185
1561 397 1432 338 1144 489 2.54 2.36 4.27
1872 364 1664 350 1408 490 1.95 2.10 3.48
1917 333 1.74
Note: IR, irrigation water; WUE, water-use efficiency.

a
1 decare = 0.1 ha.

Maximum yield for cotton was 438, 363, and 328 kg/decare from drip, furrow, and sprinkler,
respectively, with 898, 937, and 1106 mm of irrigation water. Cetin calculated the amounts of water,
using pan evaporation coefficients of 0.87, 0.90, and 1.07. According to these results, the yield from
drip irrigation method was 34 and 24% more than those from furrow and sprinkler methods,
respectively, and the yield from furrow was 11% more than that from sprinkler. Generally, the mobile
irrigation systems (mobile drip and LEPA) gave lower cotton yields.


Fig. 1. The relationships between cotton yield and amount of irrigation water for various irrigation
methods (Source: Cetin, 1997).

In Cukurova region, Ertek (1998) carried out an experiment to develop a suitable program for drip
irrigation of cotton, as well as studying the possibility of using drip systems to irrigate cotton. The
study took place in 1994 and 1995. Ertek used Cukurova-1518 variety cotton. The laterals were at
0.7-m intervals (a lateral for every crop row). Ertek determined the amount of irrigation water on the
basis of free surface evaporation from a screened class-A pan. The treatment comprised two
irrigation intervals (5 and 10 days), three plant-pan coefficients (0.75, 0.90, and 1.05), and two wetting
percentages (0.70 and the cover percentage of crop). Ertek applied the first irrigation when the
available soil moisture was at 40% in the 120-cm depth of the profile.

Average seasonal irrigation water varied 336439 mm; seasonal evapotranspiration varied 468
580 mm; and the cotton yield varied 269320 kg/decare (Table 4). Although the effect of irrigation
interval and wetting percentage on cotton yield was not significantly different for the first- and second-
year plant-pan coefficient, the interaction of wetting percentage and crop-pan coefficient was
significantly different at 5% between the treatments.

Table 4. Some results from drip irrigation of cotton on the Cukurova Plain (Source: Ertek, 1998).
Treatment
a,b
IR
(mm)
IR
(%)
ET
(mm)
Yield
(kg/decare)
TWUE
(kg/decare
per mm)
IWUE
(kg/decare
per mm)
IR/ET
(%)
I
1
Kcp
1
P
1
336 76 468 269 0.58 0.89 68
I
1
Kcp
2
P
1
360 82 490 279 0.58 0.86 69
I
1
Kcp
3
P
1
383 87 525 297 0.57 0.87 69
I
1
Kcp
1
P
2
370 84 496 283 0.59 0.85 71
I
1
Kcp
2
P
2
401 91 536 318 0.61 0.88 71
I
1
Kcp
3
P
2
431 98 564 299 0.53 0.75 73
I
2
Kcp
1
P
1
336 76 471 287 0.62 0.93 67
I
2
Kcp
2
P
1
360 82 491 292 0.61 0.90 69
I
2
Kcp
3
P
1
383 87 524 307 0.61 0.89 69
I
2
Kcp
1
P
2
376 86 516 311 0.62 0.92 68
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

186
I
2
Kcp
2
P
2
407 93 562 320 0.59 0.88 70
I
2
Kcp
3
P
2
439 100 580 317 0.56 0.79 72
Note: ET, evapotranspiration; I, interval; IR, irrigation water; IWUE, irrigation water-use efficiency;
Kcp, crop-pan coefficient; TWUE, total water-use efficiency.
a
I
1
, 5 days; I
2
, 10 days.
b
K
cp1
, 0.75; K
cp2
,
0.90; K
cp3
, 1.05.
c
1 decare = 0.1 ha.

There were significant relationships between plant height, leaf-area index, and development of
plant covers, dry matter with both irrigation waters and evapotranspiration. Depending on the
treatment, effective root-zone depth for cotton varied 88111 cm. Total WUE and irrigation WUE
varied 0.580.62 kg/decare per mm and 0.750.93 kg/decare per mm, respectively. The ratio of
irrigation water to evapotranspiration was 6873%. The salt accumulation at 15 cm from the dripper
increased in the upper layer and gradually decreased toward the bottom. At 30 cm from the dripper,
salt accumulation increased to near the wetted front.

Senyigit (1998) conducted an experiment on watermelon. This experiment studied various
irrigation methods (sprinkler and drip), nitrogen forms (liquid and granule) and amounts (based on
applied line source sprinkler), and two varieties of watermelon (Paladin and Madera). Senyigit carried
out the study at the Research and Production Farm of the Agricultural Faculty of Cukurova University,
during the 1996 and 1997 growing seasons.

Generally, Senyigit irrigated the plants at 512-day intervals. Free water-surface evaporation
determined the amount of irrigation water. Senyigit estimated the irrigation water in the plot with drip
irrigation based on an assumed irrigation of 70% per volume of the soil. Only the treatments with
sprinkler and liquid nitrogen and with sprinkler and granule and liquid nitrogen had three nitrogen
levels, providing a gradient during the irrigation season. The yield losses and WUE for watermelon
with various irrigation methods and nitrogen types are given in Table 5. The yield losses as a
proportion of marketable yield showed differences between total and marketable yield of watermelon.
The highest loss occurred for Madera with sprinkler and granule and liquid nitrogen, at 32%. The
lowest loss was for both varieties with drip irrigation.

Table 5. Yield losses and WUE (Source: Senyigit, 1998).
Yield loss (%) WUE (kg/decare per mm)
Treatment Madera Paladine Madera Paladine
SG 31 23 9.78 7.56
SGL 32 27 9.67 7.18
SL 26 27 8.85 8.59
DL 16 17 12.92 11.11
Note: DL, drip with liquid N; SG, sprinkler with granule N; SGL, sprinkler with granule and liquid N; SL,
sprinkler with liquid N; WUE, water-use efficiency. 1 decare = 0.1 ha.

Average WUEs ranged 7.1612.92 kg/decare per mm. WUEs under drip irrigation were higher
than under sprinkler irrigation by an average 27% and 29% for Madera and Paladin varieties,
respectively. Similarly, values for Madera were 17% higher than those for Paladin under sprinkler
irrigation. Yield-response factor was 1.07 for total yield and 1.49 for marketable yield.

In the experiment, corn was irrigated 6 and 7 times in 1993 and 1994, respectively, and a total of
752 mm to 823 mm or irrigation water were applied to I100 irrigation treatment, in which water use
was determined as 999 mm and 1052 mm in 1993 and 1994, respectively. Grain yield obtained from
the I100 treatment, 1001.5 kg/da in the first year and 1003.5 kg/da in the second year of the
experiment. Yield obtained from the I80 treatment, which received 20% less water as compared with
I100, was not significantly different from the full irrigation treatment. Beyond the I80 level, deficit water
application resulted in significant yield reduction by affecting both seed mass and kernels per ear.
Significant second power and linear relationships were found between grain yield (Y) vs seasonal
irrigation (I), and grain yield vs water use (ET), respectively. In the first and second year of the
experiment, the yield response factor (ky) was determined as 1.08 and 1.61, respectively. Irrigation
water use efficiency (IWUE) and water use efficiency (WUE
ET
) were found to be between 1.02.43
kg/da mm and 0.221.25 kg/damm, respectively for the treatments studied (Gencoglan and Yazar,
1999).

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

187
Determination of irrigation interval and evapotranspiration of Sanlurfa pepper was carried out in
Harran University Research Area in 2001. In this study three different irrigation intervals used with
three crop pan coefficients (Kcp1=1,25, Kcp2=1 and Kcp3=0,75). Irrigation water amount in the
treatments varies between 652-1010 mm and seasonal water consumption is between 726-1069 mm.
The yield reached between 2444-4703 kg/da in the study. Total water use efficiency (WUE) varies
between 2,75-5,22 kg/da/mm, irrigation water use efficiency (IWUE) 3,03-5,81 kg/da/mm were
detected (Tas, 2002)
In Turkey, early irrigation experiments on cotton were conducted in 1940s in Cukurova Region
(Alap, 1958). Deficit irrigation of cotton was first proposed by Tekinel and Kanber (1979) who
investigated crop water requirement and yield production functions of cotton. Their irrigation
treatments were such that 60 % of available water was allowed to decrease to start irrigations in the
control treatment. Other treatments received a given fraction less water than control (Table 6). They
found that a second degree polynomial relation could adequately describe yield response of cotton,
which showed that as much as 30 % reduction in irrigation water application did not appreciably
hinder cotton yield.
Table 6. Average Yield, IR, ET and Crop Water Use Efficiency (Source: Tekinel and Kanber, 1979).
Treatment IR
mm
ET
mm
Yield
kg/ha
Yield
%
IR/ET TWUE
kg/ha-mm
A (Control) 660 828 3590 a 92 0.80 4.3
B: 80%A 528 728 3840 a 98 0.72 5.3
C: 60%A 394 618 3900 a 100 0.63 6.3
D: 40%A 267 478 3820 a 98 0.80 8.0
E: Dry - 118 1650 b 42 -

Another experiment were done in Harran Plain in Southeast Anatolian region by also Kanber et al.,
(1991) for getting the convenient irrigation program for cotton using the free water evaporation. Here
again, 3 irrigation intervals (I
1
: 7, I
2
: 14 and I
3
: 21 days) and four crop-pan coefficients (Kpc
1
: 0.7;
Kpc
2
: 0.9; Kpc
3
: 1.1; Kpc
4
: 1.3) were tested regarding to obtain the suitable coefficient and irrigation
interval for irrigation of cotton. Irrigation water amount given to the plots was estimated using the
cumulative pan evaporation occurred during the aforementioned irrigation interval. They found that
irrigation water varied from 619 to 1112 mm, whereas evapotranspiration from 1075 to 1504 mm
(Table 7).

According to results, the evapotranspiration of cotton was very high and effected by the advection
from widespread bare soils placed surrounding of the experimental area. Other side, it was
determined that the free water surface evaporation can be used for the irrigation scheduling of cotton.
For this purpose, cotton must be irrigated at the 7 days interval and irrigation water amount to be
applied to soil can be calculated using crop-pan coefficient of 1.4. In some places where the
evaporation losses are very high, chemicals were applied to reduce evapotranspiration of cotton. In
this study, the effects of irrigation intervals and antitranspirant doses on evapotranspiration, yield, and
water use efficiency of cotton were investigated on the field plots in Harran Plain for 4 years (Kanber
et al., 1992). Different irrigation intervals (I1: 7, I2: 14, and I3: 21 days) and four antitranspirant doses
(D0: 0; D1: 40 g/ha; D2: 80 g/ha; and D3: 160 g/ha) were tested. The antitranspirant that contains N,
N, N-tributtill-3- (trifluoromethyl) benzene methananium chloride as the effective substance was used
in sub-plots of the experiment. The antitranspirant application was done in the two times in which the
reddish color on the main stem of cotton 5-7 cm reach to the top bud (as the first application) and at
the 5-7th days of ball formation (as the second application) during the growing season. The irrigation
programs were begun after the first application of antitranspirant and 90 cm soil depth was wetted in
irrigation events.

Table 7. Average yield, IR, ET and crop water use efficiency (Source: Kanber et al., 1991)
Treatments No. Of
Irrig.
IR
mm
ET
mm
IR/ET Yield
kg/da
TWUE
kg/decare-mm
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

188
I
1
Kcp
1
I
1
Kcp
2
I
1
Kcp
3
I
1
Kcp
4

I
2
Kcp
1
I
2
Kcp
2
I
2
Kcp
3
I
2
Kcp
4
I
3
Kcp
1
I
3
Kcp
2
I
3
Kcp
3


I
1
Kcp
4

9
9
9
9
6
6
6
6
3
3
3
3
671
819
965
1112
619
752
884
1016
671
818
966
1112
1091
1256
1381
1481
1079
1198
1260
1369
1075
1218
1386
1504
0.61
0.65
0.70
0.75
0.57
0.63
0.70
0.74
0.62
0.67
0.70
0.74
233
285
341
376
206
237
277
316
203
238
227
235
0.21
0.23
0.25
0.25
0.19
0.20
0.22
0.23
0.19
0.19
0.16
0.16

Results show that the frequent irrigation increased evapotranspiration (ET) and net irrigation water
requirement (IR). The maximum ET and IR values were found to be 1670 and 1555 mm, respectively
in treatment I
1
(Table 8). The highest WUE values, although not statistically significant, were obtained
from I
2
as 2.41 and 2.69; and from D
1
as 2.34 and 2.60.


Table 8. Results from experiment of antitranspirant doses and irrigation program (Source: Kanber et
al., 1992)
Average Values
Treat
No
of
rr.

IR
mm

ET
mm

IWUE


TWUE Yield*
kg/da
No of
irr.

IR

ET

IWUE

TWUE
I
1
D
0

I
1
D
1

I
1
D
2

I
1
D
3

I
2
D
0

I
2
D
1

I
2
D
2

I
2
D
3


I
3
D
0

I
3
D
1

I
3
D
2

I
3
D
3

13
13
13
13

7
7
7
7

5
5
5
5
1555
1555
1555
1555

1113
1113
1113
1113

894
894
894
894
1670
1670
1670
1670

1234
1234
1234
1234

1019
1019
1019
1019
2.45
2.55
2.39
2.35

2.62
2.76
2.65
2.74

2.45
2.48
2.57
2.38
2.28
2.36
2.23
2.18

2.34
2.48
2.36
2.46

2.15
2.18
2.27
2.11
384a
394a
361a
376a

295b
302b
298b
304b

223c
224c
227c
209c
8 (D
0
)
8 (D
1
)
8 (D
2
)
8 (D
3
)
1201 (D
0
)
1182 (D
1
)
1172 (D
2
)
1196 (D
3
)
1322 (D
0
)
1310 (D
1
)
1290 (D
2
)
1312 (D
3
)





2.51 (D
0
)
2.60 (D
1
)
2.54 (D
2
)
2.49 (D
3
)



2.44 (I
1
)
2.69 (I
2
)
2.47 (I
3
)


2.26 (D
0
)
2.34 (D
1
)
2.29 (D
2
)
2.25 (D
3
)



2.26 (I
1
)
2.41 (I
2
)
2.18 (I
3
)


* = x S 20.94 and 9.77; the yield groups were statically obtained by the orthogonal comparison
methods.

The application of various antitranspirant doses had no significant effect both on seasonal ET and
WUE values. The irrigation intervals have significant effect on the yield and quality of cotton. The
maximum cotton yield was obtained from frequent irrigations. Frequent irrigation applications
increased lint length, whereas, infrequent irrigations and antitranspirant doses resulted in shorter and
thicker lint.


CONCLUSIONS

The subject of water use efficiency is quite complex and often misunderstood both within and
outside the scientific communities. The information presented herein has identified the major factors
contributing to improvements in WUE in both the Turkeys irrigated and non-irrigated agriculture
sectors. One of the sources of future growth in crop production in Turkey is enhanced efficiency of
irrigation and water use. In this paper the investigated studies reflects that the water use efficiency
differs by a wide range of supply-side water efficiency practices, such as better system integration,
conjunctive use of surface and groundwater supplies and other measures that can stretch existing
supplies even further. With the increasing population the water availability is an increasingly critical
constraint to expanding food production in many of the world's agroecosystems. In this way because
of the water limitations water use efficiency comes into prominence.

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

189

REFERENCES

Alap, M., 1958. Cotton report. Soil and Water Research Institute, Tarsus, Turkey.
Cetin, O. 1997. The effects of different irrigation methods on the cotton yield. The research yearbook
of soil and water sources of 1996. Research, Planning and Coordination Department of General
Directorate of Village Affairs. No. 102, Ankara, Turkey. pp. 280294.
Cevik, B.; Tekinel, O.; Ozsan, M.; Tuzcu, O.; Yurdakul, O. 1982. A research on comparison of furrow,
sprinkler and drip irrigation methods for citrus growth under Cukurova conditions [in Turkish]. Doga
[Turkish Journal of Agriculture and Forestry], 6(3), 9198.
Donald, M.T. 2000. Principle of water use efficiency. Inaugural Ceremony International Seminer on
Efficient Water Use Chapter 2, 14p. www.unusco.org.uy
El Beltagy, A. 1997. West Asia and North Africa: A regional vision. ICARDA-023/May 1997, Aleppo,
Syria: ICARDA.
Ertek, A. 1998. The possibilities of irrigation of cotton with drip systems [in Turkish]. Irrigation and
Drain Engineering Department, Institute of Natural Science, Cukurova University, Adana, Turkey.
PhD thesis. 140 pp.
Eylen, M.; Tok, A.; Kanber, R. 1988. Water consumption, growth and yield of oranges irrigated by
mini-sprinkler and trickle methods under Cukurova conditions. Soil and Water Research Institute,
Tarsus, Turkey. General Pub. No. 136, Report Series No. 80, pp. 3560.
Gencoglan, C and Yazar, A. 1999. The Effects of Deficit Irrigations on Corn Yield and Water Use
Efficiency. Turkish Journal of Agriculture and Forestry 23 (1999) p. 233-241. Ankara, Turkey.
Hillel, D., 1987. The Efficient Use of Water in Irrigation. Principles and Practices for Improving
Irrigation in Arid and Semiarid Regions. World Bank Technical Paper, ISSN 0253-7494; Number
64. Washington, DC, USA.
Kanber, R. and Dervis, O. 1975. Comparison of drip and furrow irrigation methods on strawberry
under Cukurova conditions [in Turkish]. Soil and Water Research Institute, Tarsus, Turkey.
General Pub. No. 74, Report Series No. 29, pp. 126132.
Kanber, R.; Eylen, M.; Tok, A. 1986. Water consumption and yield of strawberry irrigated by furrow
and trickle methods under Cukurova conditions. Soil and Water Research Institute, Tarsus,
Turkey. General Publication No. 135, Report Series 77, pp. 735.
Kanber , R., Tekinel, O., Baytorun, N., Onder, S. 1991. The Possibilities of free water surface
evaporation for using irrigation program and evapotranspiration of cotton under Harran Plain
conditions. Prime-minister of Turkish Republic, Chairmanship of Southeastern Anatolia Project
Authority, Pub.44, Adana, 38 pp
Kanber, R., Yazar, A., Koksal, H., Oguzer, V., 1992. Evapotranspiration of Grapefruit In The Eastern
Mediterranean Region Of Turkey. Scientia Hort., 52:53-62
Ozsan, M.; Tekinel, O., Tuzcu, O., Cevik, B. 1983. A research on determination of the most efficient
irrigation method for lemon growth under Cukurova conditions [in Turkish]. Doga [Turkish journal
of Agriculture and Forestry], Series D2, 7(1), 6370.
Senyigit, U. 1998. The effects of different irrigation methods and different nitrogen dozes and forms
on the yield and quality of watermelon. International Center for Advanced Mediterranean
Agronomic Studies, Bari, Italy. 60 pp.
Tas, I. 2002. Determination of Evapotranspiration and Irrigation Interval of Sanlurfa Pepper. Institute
of Natural Science, 91 p. Sanlurfa, Tukey
Tekinel, O., Kanber, R., 1979. Determination of cotton evapotranspiration and yield by appling limited
water quantity under Cukurova condition. Soil and Water Research Institute, General No. 98,
Report. No. 48, Tarsus, 39 p.
Tekinel, O., Kaska, N., Din, G., Yurdakul, O. 1984. A research on comparison of furrow, sprinkler
and drip irrigation methods upon early strawberry growth under ukurova conditions [in Turkish].
Doga [Journal of Agriculture and Forestry], Series D2, 8(1), 4855.
Yavuz, M.Y. 1993. The effects of different irrigation methods on the yield and water use efficiency of
cotton [in Turkish]. Irrigation And Drain Engineering Department, Institute of Natural Science,
Cukurova University, Adana, Turkey. PhD thesis. 196 pp.






Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

190















































OTHER CONTRIBUTIONS












Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

191




























































Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

192





















WATER USE EFFICIENCY IN C
3
CEREALS UNDER MEDITERRANEAN
CONDITIONS: A REVIEW OF SOME PHYSIOLOGICAL ASPECTS



E.A. Tambussi
*
, J. Bort
**
, J.L. Araus
**

*
Instituto de Fisiologa Vegetal (INFIVE), Universidad Nacional de La Plata, cc 327, 1900 - La Plata,
Argentina, Tel. + 54 221 4236618; Fax: + 54 221 4233698; E-mail: tambussi35@yahoo.es
**
Unitat de Fisiologia Vegetal, Facultat de Biologia, Universitat de Barcelona, Agda. Diagonal 645 (E-
08028)- Barcelona, Spain



SUMMARY In this review we will discuss physiological traits of C
3
cereals related to water use
efficiency (WUE) in Mediterranean environments, from leaf (instantaneous WUE) to crop level (water
productivity). Carbon isotopic discrimination (
13
C) is the main approach used to estimate
instantaneous WUE in C
3
plants, and the negative correlation between both parameters has been
confirmed by several works. However, the relationship between
13
C and grain yield is more complex,
and may differ among environments. In Mediterranean irrigated conditions, a positive correlation
between
13
C and grain yield is found in barley and wheat, whereas in stored-water crops (such as
in some regions of Australia), lower
13
C (i.e. higher WUE) is associated with higher grain yield,
particularly in more stressful conditions. These apparent inconsistencies and their possible
implications in plant breeding are discussed.
One physiological trait that has received minor attention is the role of the ear photosynthesis to
improve instantaneous WUE. Several works carried out by our group (using gas exchange analysis
and carbon isotopic discrimination) have reported that, ears of barley and durum wheat have a higher
instantaneous WUE than the flag leaf, both in well watered as in drought conditions. The underlying
causes for the higher WUE of ears are not completely known, but their refixation capacity (i.e., the
capacity of re-assimilate respired carbon dioxide) could be important. Although the genotypic
variability of this trait has not been extensively studied, some data support the idea that such variability
does exist.
At the crop level, decreasing soil evaporation is a crucial factor to ameliorate the WUE in
Mediterranean conditions, and, in this sense, a fast initial growth of the crop (early vigour) seems to be
a relevant trait. A positive correlation between higher early vigour and grain yield was found in some
studies. However, early vigour may not confer an advantage in other environments, and this issue is
discussed.
Finally, because the water losses by the epidermal cuticle (residual transpiration) may be important in
droughted plants, we will briefly comment on possible future research directions in this field, as a
possible way to improve WUE in cereals.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

193


Key words: barley, carbon isotope discrimination, cereals, ear photosynthesis, Mediterranean, water
use efficiency, wheat


INTRODUCTION

Water is the main abiotic factor limiting plant production in several regions of the world, with crop
growth and economic yield being severely affected by water availability (Araus et al. 2002). The water
use (WU. i.e. the water consumed) and water use efficiency (WUE, in general terms, the efficiency of
this consumed water to produce biomass or grain yield) are critical parameters where water is scarce,
as in semi-arid regions with Mediterranean climate.

Mediterranean climate (e.g. Mediterranean basin in Europe, North Africa, West Asia, Western
Australia), is characterised by relatively scarce and erratic precipitation, with wet winter and dry and
hot summers (e.g. Acevedo. et al. 1999), and spring being the main growing season (Fischer and
Turner 1978). The vegetative growth of C
3
cereals takes place at low vapour pressure deficit (VPD)
and (eventually) with well soil moisture conditions. Grain filling, by contrast, may be affected by
terminal water stress episodes, in particular associated with high irradiances and high temperatures
(i.e. drought). World cereal demand is growing at the present (for wheat, ca. 2% per year)
(Skovmand et al. 2001), and more water will be required in expanded agriculture of the future. In this
context, saving water during the crop cycle through higher WUE may have strong impact at local and
regional scale. In short, the use of less water to achieve high yield is a major objective of the modern
agriculture (Richards et al 2001).

In this article we will focus the discussion to C
3
cereals, in particular durum (Triticum turgidum L.
var durum) and bread (T. aestivum L.) wheat as well as barley (Hordeum vulgare L.)., three widely
cultivated crops in Mediterranean regions. In the last decades, several morpho-physiological traits of
these cereals associated with high (or eventually, low) WUE have been studied. We will discuss some
of these issues in the present review.


WATER USE EFFICIENCY: ONE TERM FOR DIFFERENT CONCEPTS

In general terms, the efficiency of one process is the ratio between the obtained product (the
numerator) and the energy or resource invested in the process (denominator). In the context of WUE
the product is the assimilated carbon and the inversion is the used water (the resource). Numerator
and denominator of this ratio may be considered at several levels, and consequently, different
definitions of WUE can be made. At leaf scale, WUE may be defined as the net CO
2
assimilated by
photosynthesis (A), divided by the water transpired in the same time period (symbolised as E or T),
being an instantaneous definition of WUE (WUE
instantaneous
) (v.g. Polley 2002) as measured by gas
exchange analysis. Intrinsic water-use efficiency (WUE
intrinsic
), on other hand, is the ratio between A
and stomatal conductance (i.e. A/g). Since WUE
intrinsic
is not influenced by VPD (the driving force of
transpiration rate), this parameter is used in comparative studies where different evaporative
demands could be present (e.g. Morgan et al. 1991; Johnson 1993; Abbad et al. 2004).

Agronomists define WUE rather from an integrative approach, i.e. the accumulated dry matter
divided by the water used by the crop in the same period (e.g. Abbate et al. 2004). It must be noted
that the term integrative has two components, thus, a temporal and spatial dimension. Firstly,
compared with WUE
instantaneous
, the dry matter accumulation takes place throughout a longer time (at
least, several days or months) than instantaneous photosynthesis rate. Secondly, dry matter
represent an integrative parameter at space scale, because may include organs (leaves, stems,
roots) and process (v.g. respiration in heterotrophic tissues) at several plant levels. Thus, although
gas exchange and integrative WUE can be related, we must keep in mind the different spatial and
temporal scale of both concepts (see below in the section Scaling-up: from instantaneous to crop
WUE).

In broad sense, assimilated dry mater can be considered as the total biomass (commonly, above-
ground parts) or, alternatively, as the accumulated dry matter partitioned the economical product (for
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

194
cereals, the grains). Thus, it may be defined a WUE for the biomass (WUE
biomass
) and for the grain
yield (WUE
yield
) (v.g. Hatfield et al. 2001; Huang et al. 2005). Although strongly linked, WUE
biomass
and
WUE
yield
may indicate different concepts. As we will discuss later, whereas both may be severely
affected by growth, WUE
yield
is also influenced by the partition of assimilated, i.e. the harvest index of
the crop.

On the other hand, the denominator of the WUE (i.e. the consumed water) can be considered in
different ways. Water losses may include, despite of water transpired by the plant, the direct
evaporation from the soil (symbolised as E
s
; e.g. Oweis et al. 2000).

Because of the existence of several scales and uses of the term WUE, this concept should be
accurately defined in each particular study. For instance, whereas that many authors consider the
term transpiration efficiency as CO
2
assimilated by photosynthesis, divided by the water transpired in
the same time period (i.e, a equivalent term to WUE
instantaneous
; e.g. Araus et al. 2002), another
researchers have used the same term to design the dry matter produced per unit transpiration (for
instance Angus and van Herwaarden 2001; Richards et al. 2001). In order to remove the confusion
arises from the misuse of the term water use efficiency, which may indicate different meanings,
another terms has been introduced, such as water productivity (Pereira et al. 2002; Passioura 2004)
equivalent to WUE
yield
. In summary, from the above considerations in this review we will use the
following definitions of WUE:


Gas exchange WUE Integrative WUE
WUE
instantaneous
= A / T WUE
biomass
= Dry Matter / (evapo)transpired water
WUE
intrinsic
= A / g WUE
yield
(water productivity) = Grain Yield / (evapo)transpired water


Instantaneous and intrinsic WUE

A higher WUE
instantaneous
can be achieved either through lower stomatal conductance (e.g.; Van den
Boogard 1997; Ashraf and Bashir 2003) or higher photosynthetic capacity or a combination of both
(Morgan and LeCain 1991; see references in Condon et al. 2002). At first, if a high assimilation rate is
linked to a higher stomatal conductance, WUE will not ameliorate. However, photosynthetic rate could
be increased by another ways. The reduction of photorespiration of C
3
plants (e.g. increasing of
affinity of the enzime Rubisco by CO
2
) has been postulated, but scarce exit has been achieved until
the present (Reynolds et al. 2000). No increases of the photosynthetic rate were found in association
with breeding, at least in durum wheat (see references in Araus et al. 2002). Only recently, increases
in photosynthetic rate in wheat cultivars have been reported, although at expenses of a parallel rise of
stomatal conductance (Fischer et al. 1998).

One postulated way to increase WUE
instantaneous
is by a higher specific leaf weight (SLW, the ratio
between weight and leaf area), since a higher SLW represents an increase in photosynthetic
machinery per leaf area. The correlation between SLW and WUE
instantaneous
has been reported,
although seem to be low (Morgan and LeCain 1991). On other hand, an increase of mesophyll
conductance is linked to higher photosynthetic rates, without increasing stomatal conductance. In this
situation, WUE
instantaneous
will be increased. In synthetic hexaploids wheat-derived populations,
mesophyll conductance accounted the 85% of variation in photosynthetic rate (del Blanco et al.
2000), suggesting that simultaneous increase in photosynthetic rate and WUE
instantaneous
are possible.
If a high WUE
instantaneous
is associated with a high photosynthetic capacity, positive correlation with
growth rate could be found. This seems to be the case of some legumes (such as peanut), where
variation in photosynthesis account for most of variation of WUE
instantaneous
(Condon et al. 2002 and
references cited therein).

However, the increases in WUE
instantaneous
seem to be associated rather with de reduction of
stomatal conductance. In this case, (as in several species in Mediterranean conditions; for instance
Bolger and Turner 1998), the growth rate may be reduced. Thus, high WUE
instantaneous
can be
associated with conservative water use (and consequently, a lower growth; Condon et al. 2002) and
this may have penalties in terms of grain yield (see Araus et al. 2002). However, Van Den Boogaard
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

195
et al. (1996), analysing two cultivars, of wheat reported that higher water use efficiency was
independent of growth rate. In fact, A/T (WUE
instantaneous
) ratio was positively correlated with LAR (leaf
area ratio). Since a higher LAR is commonly associated with higher growth rates (Poorter 1989), the
results reported by Van Den Boogaard (1996) indicate that a higher WUE
instantaneous
and high growth
rate could not to be mutually excluding.

Gas exchange analysis carried out in bread wheat showed that there are genotypic difference in
WUE
instantaneous
(Morgan and LeCain 1991; Abbad et al. 2004), although the greatest genetic variation
were observed in the comparison of hexaploid wheat with early progenitors (see Morgan et al. 1993
and references cited therein).

Higher stomatal conductances have been reported in the progress of yield analysing cultivars of
Northwest Mexico (Fisher al. 1998). Cultivars with higher transpiration rate and, thus, greater
transpiration rate and canopy temperature depression (CTD), show higher grain yield. This positive
correlation between CDT and grain yield has been reported elsewhere, in particular in warm regions,
where heat tolerance associated to a cooler canopies could be important (Reynolds et al. 2001).
Consequently, if the increase of WUE
instantaneous
is associated to a reduction of stomatal conductance,
penalties in terms of heat tolerance (and virtually in grain yield) could take place in crops of warm
regions.




Instantaneous

WUE and high growth rate: are they mutually exclusive?

The influence of physiological traits on WUE depends on the balance between the effects on
growth and WU. At first, plant traits that increase WUE
instantaneous
might have penalties in terms of
growth rate (Van den Boogard et al. 1997). In other words, it has been argued that a high WUE could
be genetically linked to a low growth rate. By one side, relative growth rate (RGR) is split into the net
assimilation rate (NAR, the increase in plant weight per unit leaf area) and the leaf area ratio (LAR,
the ratio between the total leaf area and the total plant weight) (Poorter 1989). On the other hand,
plant transpiration rate is the product between the LAR and the transpiration rate per leaf area unit
(E). Consequently, if a morphological or physiological plant trait increases the WUE affecting the LAR,
the RGR could decrease. Moreover, a lower transpiration rate per unit leaf area could lead to a
decrease in NAR if a lower stomatal conductance and photosynthetic rate area were implicated.
However, some reports show no association between a higher A/T ratio (i.e. WUE
instantaneous
) and slow
growth (Van den Boogaard et al. 1997). In this study -where 10 cultivars of bread wheat were
analysed- a positive correlation between LAR and the A/E ratio was found. Both assimilation rate and
stomatal conductance showed negative correlation with the LAR, but the slope was lower in the
former, leading to an increase of the A/E ratio at higher LAR (Van den Boogaard et al. 1997). In
summary, an amelioration of the WUE
instantaneous
could be possible without a concomitant decrease of
growth rate, in particular if LAR was not affected.


Instantaneous WUE and water stress

The broadest known effect of the water stress in plants is the stomatal closure and the increase of
WUE
instantaneous
. The increase of WUE
instantaneous
under water stress is a well documented phenomenon
(e.g. Johnson 1993; Morgan et al. 1993; Rekika et al. 1998; Van den Boogaard et al. 1997), although
the response of WUE
instantaneous
is dependent of the severity of the water stress. In bread wheat, for
instance, a moderate water deficit, can lead to a relevant increase of WUE
instantaneous
(e.g. Morgan et
al. 1993; Rekika et al. 1998; Van den Boogaard et al. 1997), but a decrease of WUE
instantaneous
has
also been reported in plants subjected to severe stress (for instance El Hafid et al. 1998; Shanguan et
al. 2000). This behaviour is not surprisingly, because at severe water deficit, photosynthesis may be
decreased by metabolic causes (i.e. non-stomatal limitation Lawlor 2002). Thus, in spite of the
decrease of transpiration rate when stomata are partially closed, the WUE
instantaneous
may drop at
severe drought if photosynthetic capacity is affected. In the last years, the increase of WUE
instantaneous

under moderate drought is used in management systems where a deficit irrigation or partial root
drying is imposed to the crops. This issue is discussed in section WUE and agronomic practices
(see below).
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

196


Scaling-up: from instantaneous to crop WUE

Several efforts has been conduced to understand mechanistic (genetical and physiological) basis
of the stomatal response and development, in order to ameliorate the WUE
instantaneous
of crops (for a
genetic approach, see Chaerle et al. 2005). However, in moving between scales, it is important to
take into account that the water relations of plant canopy are distinctively different than would be
predicted from lower organisation levels, such as individual leaves. Thus, differences observed in
WUE
instantaneous
could not be reflected at canopy or yield levels. For instance, it has been pointed out
that differences ca. 24% in instantaneous WUE, can drop down to 5% at canopy level (Lambers et al.
1998). Moreover, Bolger and Turner (1998), reported that, whereas significant differences in A/g ratio
between Mediterranean annual pastures were observed in glasshouse experiments, these
differences seem to be eliminated in the field. The subjacent causes of this phenomenon are diverse,
but it may summarise in two points: (a) the dominance of boundary layer resistance in transpiration
rate and (b) the effects in thermal balance of the leaf associated to the lower stomatal conductance.

(a) If the boundary layer resistance is high, as it is the case of dense canopies of wheat and barley
(Kang and Zhang 2004), the stomatal opening could exert a lower control over the transpiration
rate. Considering that stomatal conductance is commonly measured by porometry -in which the
leaf boundary layer is eliminated-, observed differences in stomatal conductances could have
lower impact than expected in transpiration rate if the former is the dominant factor.
(b) The potential gain in WUE
instantaneous
at crop level may be lower than expected, for instance, if low
stomatal conductance is linked to higher leaf temperature and, thus, increased transpiration per
stomatal conductance unit (Condon et al. 2002). Additionally, the increase of leaf temperature
could have penalties in the cases where the evaporative cooling effect of transpiration seems to
be important (Reynolds et al. 2001).

In summary, both phenomena (i.e. boundary layer and the increase of leaf temperature) may limit
the scaling-up between WUE
instantaneous
of leaf and crop level. For instance, it has been pointed out
that modern irrigation systems in which partial water stress is applied (improving WUE by partial
stomatal closing) could have a lower effect than expected in crops with dense canopies (see below;
Kang and Zhang 2004).


Decreasing futile losses of water: leaf temperature and residual transpiration

Considering that a leaf temperature is a component of the driving force of the transpiration rate, a
lower leaf temperature might have a important impact in WUE
instantaneous
. Low chlorophyll content has
been associated with lower leaf temperatures in barley. Landraces adapted to Mediterranean dry
conditions in Syria have leaves with pale colour, due to a decrease in chlorophyll content per unit leaf
area. This reduction in chlorophyll did not cause any change in photosynthetic capacity (Tardy et al.
1998). In the Mediterranean landrace Tadmor, for instance, lower chlorophyll content is associated
with lower leaf temperatures, in particular when stomata are closed (Havaux and Tardy 1999). Lower
leaf temperature under water stress could mitigate the heat stress associate to drought, and
concomitantly, reduce the losses of water across the cuticle and improve WUE.

The losses of water trough the cuticle (residual transpiration) are futile, because it is not paired
with CO
2
influx into the leaf. Residual transpiration may be quantitatively important during the day, in
particular in dry and hot climate and substantial genetic variability has been report for wheat (see
references in Richards et al. 2001). For instance, measurements carried out on varieties and
landraces from the Middle East, North Africa, INRA and CYMYT, Araus et al. (1991) and Febrero et
al. (1991) found considerable differences (ca. 50%) in epidermal (residual) conductance among
genotypes. Although the heritability of this traits is unknown (Richards et al. 2001), the results
mentioned above suggest that WUE of cereals could be improved by this way. It must be noted that a
better WUE linked to a lower cuticular transpiration do not have penalties in terms of photosynthetic
rate, and consequently, in growth performance.

In addition, some transpiration occurs at night, through abnormal stomatal closure and the leaf
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

197
cuticle. A recent report showed that night-time transpiration is not trivial in several C
3
and C
4
species,
because it may achieve 20% respect diurnal rates in some cases (Snyder et al. 2003). In this work,
nocturnal transpiration was positively correlated with diurnal rates, i.e. higher night time losses of
water are found in species with the higher diurnal transpiration rates. The impact of this phenomenon
has not been extensively analysed in crop. Warmer conditions at night (e.g. during the grain filling of
cereals in semi arid- Mediterranean regions) could lead to significant water losses, decreasing the
WUE of the crop.


Leaves versus ear photosynthesis: an opportunity to ameliorate WUE
instantaneous
in
Mediterranean regions?

The WUE of the ear of C
3
cereals has been reported in some studies (Teare et al. 1972; Araus et
al. 1993; Bort et al. 1994; Abbad et al. 2004). The higher WUE of the ear has been estimated by
different approaches, including instantaneous measurements of the photosynthesis/transpiration ratio
(i.e. gas exchange analysis) and by isotopic composition of
13
C or
13
C (Araus et al. 1992). As it was
mentioned above, the relationship between WUE and
13
C is well known. In C
3
cereals,
13
C is
positively related to CO
2
levels in intercellular spaces and (given a constant vapour pressure deficit),
negatively related to WUE (Farquhar and Richards 1984; Hubick and Farquhar 1989). In fact, triticale
(x Triticosecale) has been reported as having a progressively higher
13
C isotopic composition (
13
C,
i.e. a lower
13
C) from flag leaf to glumes and glumells (lemma plus palea) (Araus et al. 1992). Ear
parts show lower
13
C compared with the flag leaf of bread wheat (Araus et al. 1993), suggesting a
higher WUE (Hubick and Farquhar 1989). This observation is supported by gas exchange analysis
(i.e., direct measurements of A/T ratio; Araus et al. 1993; Abbad et al. 2004). The higher WUE
instantaneous

in the ear might come from the capacity to recycle respired CO
2,
which has been well documented
(Bort et al. 1996; Gebbing and Schnyder 2001).

Genotype variability (and their possible implication in breeding) of the ear photosynthesis and WUE
has been scarcely analysed. However, there are reports that suggest that some variability could exist
(see Table VII in Abbad et al. 2004). In this study, the authors analyse the performance of flag leaf and
the ear in photosynthetic rate and the A/T ratio (i.e., WUE
instantaneous
) in six cultivars, either in well-
watered or water-stressed treatments. Under both conditions, significant differences in A/T ratio
between cultivars were found, suggesting that this trait could be explored in future research. In another
work in bread wheat and barley, Bort et al. (1996) showed that refixation capacity of the ear (a
parameter possibly linked to ear WUE) seem to be genetically fixed.

In summary, future investigations should be carried out in order to explore this field research, in
particular the ear traits (e.g. refixation capacity) related to a higher photosynthetic rate and
WUE
instantaneous
.


Decreasing soil evaporation: early vigour in Mediterranean conditions

Soil evaporation is an important component of total water losses, in particular in sparse canopies
at the beginning of the crop growth (e.g. Kato et al. 2004). In semi arid region (for instance Northern
Syria), soil evaporation may be a major component of total water use, (value >50% are reported; e.g.
Corbeel et al. 1998), in particular in non-fertilised crops (Gregory et al. 2000). A better seedling
emergence and a more early vegetative growth (i.e. early or seedling vigour) has been reported as an
important trait in terms of WUE in cereals in Mediterranean conditions (Richards et al. 2001; Richards
et al. 2002). Evaporation from the soil is negatively correlated with fractional area of shaded soil (e.g.
Passioura 2004). For instance, compared with wheat, barley achieve higher leaf area at early stages
of the crops (e.g. Lpez-Castaeda et al. 1995; Rebetzke et al. 2004), decreasing the loss of water by
soil evaporation. The presence of a canopy can decrease soil evaporation by three main mechanisms
(Gregory et al. 2000) (a) reduction of net radiation absorbed by the soil (b) humidification of the air,
increasing the aerodynamic resistance to the transfer of water vapour from the soil and (c) the uptake
of water from the roots near the soil surface reduces the hydraulic conductance (Gregory et al. 2000).
In wheat canopies, the evaporation from the soil is negatively correlated with the fractional shaded
area (Passioura 2004) and it is well documented that is lower in barley than wheat (Siddique et al.
1990). In this study, cultivars with higher SLA and better early vigour reduce soil evaporation.
Moreover, compared with old cultivars, modern cultivars of wheat had lower soil evaporation rates
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

198
early in the growing season despite that transpiration efficiency for dry matter production were similar
for all cultivars. Moreover, the growth took place under low VPD, decreasing the total transpired water
by the crop, and consequently, increasing the WUE (Lpez-Castaeda et al. 1995). However, it has
been reported that a closed canopy increases the water losses by interception and evaporation from
the canopy (Leuning et al. 1994).

In summary, the fast growth of leaf area has little benefit in regions where soil evaporation is a
small component of total crop water use (Condon et al. 2004). Similar consideration may be pointed
out in areas where evaporation is high, but it is limited by the movement of water in the soil (and non
by canopy density; Gregory et al. 2000; Yunusa et al. 1993). According to estimations derived from
simulation models, the reduction of soil evaporation by a higher SLA and early vigour seem to occur
only in Mediterranean-type environments and when high nitrogen doses are applied (Asseng et al.
2003).

Finally, it must be mentioned that an early growth may be achieved by applying some agronomic
practices. Early sowing date (Oweis et al. 2000; Richards et al. 2002), non-tillage management (Klein
et al. 2002) or small irrigation at the first stages of the crops (Tavakkoli and Oweis 2004), for instance,
ensure early germination, seedling establishment and a fast growth. In addition, sedding pattern may
be relevant. The use of narrow row spacing and adequate plant population would help conserve water
and hence increase the WUE. Works carried out in semi-arid region of Morocco, for instance,
revealed that WUE were increased when row spacing was reduced (Karrou 1998).

Although the study of the early vigour has received considerable attention in Mediterranean
climate, it could be relevant in other dry regions. In the Loess Plateau in China (a semi-arid region),
the use of straw mulch (to reduce soil evaporation) has been proposed to improve WUE in bread
wheat crops (Huang et al. 2005). In this context, thus, a fast growth at early stage may replace or
complement the mulching practice.


Early vigour in non-competitive modern cultivars

In wheat, the widely use of semi-dwarf cultivars (GA -giberellic acid- insensitive) which it has
increased the harvest index of modern cultivars with lower plant height (Austin 1999), is associated
with short coleoptiles (Richards et al. 2002). Tall cultivars have longer and wider leaves and produce
higher biomass than semi-dwarf genotypes (v.g. Rht) at early stage (Rebetzke et al. 2004). These
traits of semi-dwarf genotypes lead a poor seedling establishment and, consequently, higher soil
evaporation at the beginning of the crop. At first, it could suggest that a high potential yield and high
early vigour are mutually exclusive, i.e. that modern cultivars are non-competitive ideotype (Blum
1996). However, a recent report showed the existence of dwarfing genes that promote short shoot but
no small coleoptiles, opening the possibility to explore in this way (see references in Passioura 2004).
Researches carried out by Richards group found that GA-sensitive lines have good partitioning
characteristic (i.e. high harvest index) and, at the same time, long coleoptiles (Richards et al. 2002).
Consequently, obtain wheat varieties with high yield potential and high WUE at early stage of the
crops is feasible.


WATER USE EFFICIENCY, GRAIN YIELD AND ISOTOPIC DISCRIMINATION: BREEDING FOR
HIGH O LOW WUE?

The Passioura equation as conceptual model

In water limited environments, grain yield could be modelled by Passioura equation, which it has
been largely discussed in several works (e.g. Blum 2000; Araus et al. 2002):

GY = W x WUE
biomass
x HI

where W is the water used by the crops (evapotranspiration) and HI is the harvest index. Accordingly
with this model, grain yield could be increased by (a) the capacity of capture more water, (b) the
efficiency for producing dry matter per unit of used water and (c) the ability to devote more assimilates
to the grains (Araus et al. 2002). Although at first view this model is attractive, - and sensu Passioura
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

199
(2004) the three components of the equation are sufficiently independent to make it worthwhile
considering them one by one- the terms are not independent and their interrelationship may be
complex. A higher WUE may be related with a lower water use (and virtually, lower growth and grain
yield) under drought conditions. In addition, although harvest index may be drought-independent in
some cases, drought-dependent HI is often a function of post-anthesis W (Richards et al. 2002 and
references cited therein; see references in Araus et al. 2002).

In short, is the WUE a suitable parameter to use in breeding programs? The answer is complex
and not universal, because it seems to depend on the target environment. As it was pointed out by
Blum, WUE is therefore a misleading parameter when applied to plant breeding for water-limited
environments where soil water extraction capacity is important (Blum 2001). Those traits that could
confer a higher water extraction capacity, and then, a higher water use, such as osmotic adjustment
(Blum et al. 1999), could have the opposite effect in WUE if higher stomatal conductances were
involved. We discuss this point in the following section.


Carbon isotope discrimination and their relationship with WUE and grain yield

Discrimination of the stable isotope
13
C (
13
C) has been widely accepted as an indicator of WUE
(see review of this issue in Araus et al. 2001; Pate 2001). In short,
13
C in C
3
plants is determined by
the following equation:

13
C = a + (b-a).(ci/ca)

where a = 4.4 represents the isotope fractionation associated with differential diffusivities of
13
C
versus
12
C, b = 27 is the fractionation by Rubisco carboxilation and ci and ca are the intercellular
and ambient CO
2
concentration respectively (Pate 2001). The ci/ca ratio is determined by the balance
between stomatal conductance and photosynthetic rate, thus, related to the A/g ratio (the demand
and supply of CO
2
respectively). Since
13
C and ci/ca are partially determined by the A/g ratio,
measurements of
13
C provide a relative index of WUE
intrinsic
or WUE
instantaneous
for given VPD
conditions (e.g. Pate 2001). In fact, a negative correlation between
13
C and WUE has been
contrasted in several works (e.g. Hubick and Farquhar 1989; Morgan et al. 1993; Johnson 1993). It
must be noted that from a mechanistic point of view,
13
C is related with instantaneous A/E ratio (or
rather, with A/g ratio), but it can provide a time and spatially integrated estimate of water use
efficiency. However, the correlation of
13
C with WUE at other scales could be lower (for instance,
Shaheen et al. 2005).

Although the negative relationship between
13
C and WUE is widely consistent throughout several
studies (see references above), the sign and magnitude of the correlation with grain yield in C
3

cereals is complex, and may be strongly influenced by several factors. In Mediterranean conditions,
numerous studies reported a positive correlation between kernel
13
C and grain yield of bread wheat
(Morgan et al. 1993) durum wheat (Merah et al. 1999; Merah et al. 2001, Araus et al. 1997; 2003;
Fischer et al. 1998; Clay et al. 2001; Royo et al. 2002) and barley (Voltas et al. 1998). In a recent
study Monneveux et al. (2005) confirms these previous finding, although a consistent positive
correlation between grain
13
C and yield was observed only under post-anthesis water stress
conditions. Under pre-anthesis and limited residual moisture stress, the correlation was weaker. At full
irrigation, by contrast, no correlation was found (Monneveux et al. 2005). Summarising the former
considerations, Royo et al. (2002) suggests that selecting for higher
13
C (lower WUE) in breeding
programs in Mediterranean basin could take advantage only under wet or irrigation conditions.

A high
13
C in the grains may involve different phenomenon, such as (a) a greater access to soil
water (for instance, related to a deeper root systems) or higher water extraction capacity (e.g.
occurrence of osmotic adjustment) (b) higher remobilization of stems reserves of pre-anthesis
assimilates which may have a lower isotope signature and (c) an earlier flowering (see Condon et al.
2002 and references cited therein). For these reasons, it has been pointed out that kernel
13
C may
be a more cryptic parameter than leaf
13
C (Condon et al. 2004). However, the correlation between
leaf
13
C and grain yield are poor in some cases (Merah et al. 2001).

Several works carried out by CSIRO group (Condon et al.) showed that breeding by low leaf
13
C
increased the grain yield under rainfed conditions in Australia (Rebetzke et al. 2002). In fact, the first
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

200
genotypes with low
13
C and high grain yield were released during the 2002 and 2003 (v.g. Drydale
and Rees cultivars; see web page of CSIRO: www.csiro.au; Condon et al. 2004). In according with
this studies, the advantage of low carbon isotope discrimination is higher at low environment mean
yields (i.e., when water stress is more severe). The improvement of low
13
C selection, however,
decline at higher mean yield, where higher season rainfall are present (Rebetzke et al. 2002).

The apparent discrepancies between studies carried out in Mediterranean climate (e.g. in
Southern Europe) versus some regions of Australia seem to arise from the source of water used
during crop cycle. In Mediterranean zone, rainfall (although scarce) is present during the crop growth
and stored water may be a minor component of the water used by the crop. In this context, genotypes
with higher water extraction capacity will have higher grain yield and higher
13
C. On other hand, in
some regions of Australia with summer rainfall, stored water is a main part of the total water used by
the crop. In those environments, rainfall are scarce after seeding, and the saving water could be a
critical factor to avoid terminal severe water stress (Passioura 2004). In this context, cultivars with
conservative strategy may have advantage respect water spender ones. The negative correlation
between the advantage in grain yield of low
13
C varieties and rainfall (mentioned above) support this
idea (Rebetzke et al. 2002). Simulation of the effect on yield incorporating higher instantaneous WUE
(low
13
C) showed that advantage was significant in environments where dominate stored water (with
summer rainfall). In environments with Mediterranean climate (i.e. winter rainfall), by contrast,
improved WUE did not confer advantage in yield (Condon et al. 2004). In this case, early vigour
seems to be more important (see above).

Araus et al. (2003) -analysing the environmental factors determining
13
C in durum wheat under
Mediterranean conditions in several trials with 25 genotypes- found that the correlation between
kernel
13
C and grain yield was steady a positive when mean yield of the trial above 2500 kg ha
-1
. By
contrast, trials with mean yield below 2000 kg ha
-1
showed low correlation between
13
C and grain
yield. Thus, in more stressful environments,
13
C may be a poor indicator of grain yield, as it was
suggested by previous reports (e.g. Royo et al. 2002). Where additional water is not available to the
crop, to increase WUE (selecting by low
13
C) appears to be an alternative strategy (Araus et al.
2002 and references therein).


Cereals yield progress and WUE

In a retrospective study of wheat Siddique et al. (1990) found that WUE
yield
increased substantially
from old to modern cultivars, with little difference among modern cultivars. WUE
biomass
, by contrast,
was similar between cultivars. Improved WUE
yield
in modern cultivars was associated with faster
development, earlier flowering, improved canopy structure and higher harvest index (Siddique et al.
1990). More recently, in a study of several cultivars of bread wheat released in Mexico, Fischer et al.
(1998) found an increase in photosynthetic rate and stomatal conductance in modern varieties: the
rise in the last parameter (g) was higher than the former (A), leading to a decrease in WUE
intrinsic
of
modern cultivars (calculated from date of Fischer et al. 1998).

One question to answer is the range of genotypic variability in WUE in cereals. The range in
carbon isotope discrimination among cereals cultivars is around 4. However, in Aegilops geniculata
(closely related to Triticum) the range seem to be higher (ca. 7), suggesting that WUE of wheat
could be improved by introgression in hybridisation programs (Zaharieva et al. 2001).


WUE AND AGRONOMIC PRACTICES: IRRIGATION AND FERTILIZATION

Agronomic options for improving rainfall-use in dry land regions were extensively reviewed by
Turner (2004): at least, half of the increase of rainfall-use efficiency may be attributed to improved
agronomic management. The adoptions of practices such as minimum tillage, appropriate fertilization,
timely planting, in conjunction with new cultivars, has the potential to increase rainfall use efficiency of
dry land (such as Mediterranean) crops. Here, we will discuss irrigation and fertilization practices in
relation to physiological aspects of WUE.


Irrigation
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

201

The amount, timing and frequency of irrigation have strong impacts in WUE
biomass
and WUE
yield
. In
fact, WUE may drop (e.g. Huang et al. 2005) or increase (Brandypadhyay et al. 2003) at higher
irrigation levels. Irrigation may increase the WUE
yield
without any effect in WUE
biomass
, due to the
increase of postanthesis water use, which results in a higher harvest index and better grain yield
(Zhang et al. 1998). This author pointed out that the lack of an effect in WUE
biomass
could result from
two counteracted forces: the decrease of Es/T ratio (related with early growth) may be
counterbalanced by the decrease of transpiration efficiency (in this context, the ratio between dry
matter and plant transpiration), which could result from a greater leaf area and higher stomatal
conductance in irrigated treatments (Zhang et al. 1998).

Oweis et al. 2000 reported that WUE
yield
of bread wheat under Mediterranean conditions was
higher at 2/3 of irrigation requirements (compared with at full irrigation, where WUE
yield
was lower).
Similar results are reported by Tavakkoli and Oweiss (2004). As these authors pointed out, the
common practice of supplemental irrigation is not the most efficient in terms of WUE
yield
for
Mediterranean environments. Considering that water is the main limiting resource in dry areas, the
loss of grain yield due to deficit irrigation may be negligible compared with the saving in water (Oweis
et al. 2000). Deficit irrigation is a strategy under which crops are deliberately allowed to sustain some
degree of water deficit and yield reduction (Pereira et al. 2002). In Northern Syria, for instance, the
maximum in water productivity of wheat are achieved with some grain yield reduction (see references
in Pereira et al. 2002). In fact, increases of WUEyield under water limitation are reported in several
studies and climatic conditions (e.g. Abbate et al. 2004 and references cited therein).

In the last years, considerable attention have received the irrigation systems where a part of root is
subjected to dry conditions (known as CAPRI or Controlled Alternate Partial Root Irrigation). This
technique is based in two assumptions (a) fully irrigated plants usually have widely opened stomata
and (b) roots in the drying soil can respond by sending a root signal to the shoots, where the stomata
may be partially closed, increasing WUE
instantaneous
(Kang and Zhang 2004). Although these systems
could be implemented in several crops, their use in cereals could be more doubtful. As it was
mentioned earlier, in dense canopies (such as cereals crops) boundary layer resistance may be high,
and exert a main control over the transpiration. Additionally, the increase of leaf temperature coupled
with lower transpiration might eliminate any advantage of stomatal close. However, the advantages of
deficit irrigation systems (see above) suggest that stomatal control is a useful feature to improve
WUE at a crop level.

In summary, WUE
yield
in Mediterranean regions may be incremented with some irrigation, in
particular at the beginning of the crops, which improves early growth and decreases soil evaporation.
Additionally, some degree of water deficit may improve WUE
yield
, which it could be explained to some
extent by stomatal control on the transpiration (Abbate et al. 2004).


Fertilization

Fertilizer use has a remarkable effect on crop yield and WUE. There are several studies reporting
the increase of cereal WUE with N application (e.g. Zhang et al. 1998). At first, nitrogen fertilization
may increase the CO
2
assimilation rate capacity, i.e. WUE
yield
may be ameliorated due to the
improvement in WUE
nstantaneous
. In addition, and probably more importantly, fertilization increases the
early growth and the crop cover, protecting the soil from evaporation and, consequently, increasing
the proportion of transpired water by the plant (see references above). Additionally, WUE is also
ameliorated because more crop growth takes place with lower VPD in early spring (Zhang et al.
1998). Nitrogen and phosphorous nutrition have been shown to increase the early growth in
Mediterranean environments (see above; references in Turner 2004). In wheat, nitrogen fertiliser
input reduced soil evaporation, increasing the WUE
yield
(Asseng et al. 2001; Sadras 2002).

For barley in Mediterranean conditions, higher WUE
biomass
and WUE
yield
were achieved by nitrogen
application (Cantero-Martnez et al. 2003). According to this study, the higher WUE was explained by
an increase of preanthesis WUE. WUE was increased by nitrogen fertilization only when water was
not limiting: an excess of nitrogen may increase the water use, without improving WUE (Cantero-
Martnez et al. 2003). Gregory et al. 2000 also reported an increase in WUE
biomass
in fertilised barley.
Fertilization with N and P increased the shoot dry weight, the T/(Es + T) ratio and the WUE
biomass
.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

202
Total water use (Es + T) was not modified, but the fertilization treatment increased the WUE changing
the partitioning between Es and T (Gregory et al. 2000).

Furthermore, although there are several studies that showed the correlation between vigorous
growths for fertilization (at early stages) and grain yield (see above), this seem not to be the case in
some environments. In dry regions of Australia, a higher vegetative growth can lead to a reduction of
yield (phenomenon named as haying off) in particular when vigorous growth is followed by a terminal
severe water stress (Angus and van Herwaarden 2001). Although at first this yield reduction could be
explained by the lack of soil water during grain filling -because more water was used in producing
additional vegetative material- another causes have been proposed. For instance, the decrease of
soluble carbohydrate available for retranslocation associated with nitrogen availability (see Angus and
van Herwaarden 2001 and references cited therein). In spite of the subjacent causes, the association
between more vigorous growth by nitrogen fertilization and higher WUE seem to be not universal, and
peculiarities of the environment target must be considered.

Finally, it must be noted that there is a trade-off between a higher WUE
instantaneous
and
photosynthetic nitrogen-use efficiency, because negative correlations between both parameters have
been reported (e.g. Van Den Boogaard 1997 and references cited therein). As it was pointed out by
this author, improving WUE
instantaneous
may be suitable in environments where water is the limiting
factor, but the cost of a less efficient use of nitrogen should be considered.







REFERENCES

Abbad H, El Jaafari SA, Bort J, Araus JL. 2004. Comparative relationship of the flag leaf and the ear
photosynthesis with the biomass and grain yield of durum wheat under a range of water conditions
and different genotypes. Agronomie 24: 19-28.
Abbate PE, Dardanelli JL, Cantarero MG, Maturano M, Melchiori RJM, Suero EE (2004) Climatic and
water availability effects on water-use efficiency in wheat. Crop Science 44: 474-483.
Acevedo EH, Silva PC, Silva HR, Solar BR (1999) Wheat production in Mediterranean environments.
In Wheat: ecology and physiology of yield determination (Eds Satorre EH and Slafer GA). Food
Products Press, New York, pp. 295-323.
Angus JF, van Herwaarden AF (2001) Increasing water use and water use efficiency in dryland
wheat. Agronomy Journal 93: 290-298.
Araus JL, Amaro T, Casadess J, Asbati A, Nachit MM (1998) Relationship between ash content,
carbon isotope discrimination and yield in durum wheat. Australian Journal of Plant Physiology 25:
835-842.
Araus JL, Febrero A, Vendrell P (1991) Epidermal conductance in different parts of durum wheat
grown under Mediterranean conditions: the role of epicuticular waxes and stomata. Plant, Cell and
Environment 14: 545-558.
Araus JL, Bort J, Ceccarelli S, Grando S (1997) Relationship between leaf structure and carbon
isotope discrimination in field grown barley. Plant Physiology and Biochemistry 35: 533-541.
Araus JL, Brown HR, Febrero A, Bort J, Serret MD. 1993. Ear photosynthesis, carbon isotope
discrimination and the contribution of respiratory CO2 to differences in grain mass in durum wheat.
Plant Cell and Environment 16: 383-392.
Araus JL, Santiveri P, Bosch-Serra D, Royo C, Romagosa I. 1992. Carbon isotope ratios in ear parts
of Triticale. Plant Physiology 100: 1033-1035.
Araus JL, Casadesus J, Bort J (2001) Recent tools for the screening of physiological traits
determining yield. In Application of physiology in wheat breeding (Reynolds MP, Ortiz-Monasterio
JI, McNab A eds.), CIMMYT, Mxico DF. pp. 59-77.
Araus JL , Slafer GA, Reynolds MP, Royo C (2002) Plant breeding and drought in C3 cereals: what
should we breed for? Annals of Botany 89: 925-940.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

203
Araus JL, Villegas D, Aparicio N, Garca Del Moral LF, El Hani S, Rharrabti Y, Ferrio JP, Royo C
(2003) Environmental Factors Determining Carbon Isotope Discrimination and Yield in Durum
Wheat Under Mediterranean Conditions. Crop Science 43: 170-180.
Ashraf M, Bashir A (2003) Relationship of photosynthetic capacity at the vegetative stage and during
grain yield of two hexaploid wheat (Triticum aestivum L. ) cultivars differing in yield. European
Journal of Agronomy 19: 277-287.
Asseng S, Turner NC, Keating BA (2001) Analysing of water- and nitrogen-use efficiency of wheat in
a Mediterranean climate. Plant and Soil 233: 127-143
Asseng S, Turner NC, Botwright T, Condon AG (2003) Evaluating the impact of a trait for increased
specific leaf area on wheat yields using a crop simulation model. Agronomy Journal 95: 10-19.
Austin RB (1999) Yield of wheat in the United Kingdom: recent advances and prospects. Crop
Science 39: 1604-1610.
Bolger TP, Turner NC (1998) Transpiration efficiency of three Mediterranean annual pasture species
and wheat. Oecologia 115: 32-38.
Bort J, Febrero A, Amaro T, Araus JL. 1994. Role of awns in ear water-use efficiency and grain
weight in barley. Agronomie 2: 133-139.
Blum A (1996) Yield potential and drought resistance: are they mutually exclusive? In Yield potential
in wheat: breaking the barriers (Reynolds MPS, Rajaram, Mc Nab A eds.). Mexico DF CIMMYT,
pp. 90-100.
Blum A, Zhang J, Nguyen HT (1999) Consistent differences among wheat cultivars in osmotic
adjustment and their relationship to plant production. Field Crops Research 64: 287-291.
Blum A (2000) A molecular perspective of crop adaptation and production under drought stress. V
Simposium Hispano-Portugus de relaciones hdricas - Sevilla
Bandyopadhyay PK, Mallick S (2002) Actual evapotranspiration and crop coefficients of wheat
(Triticum aestivum) under varyiing moisture levels of humid tropical canal command area.
Agricultural Water Management 59: 33-47.
Bort J, Brown HR, Araus JL. 1996. Refixation of respiratory CO2 in the ears of C3 cereals. Journal of
Experimental Botany 47: 1567-1575.
Cantero-Martnez C, Angas P, Lampurlans J (2003) Growth, yield and water productivity of barley
(Hordeum vulgare L.) affected by tillage and N fertilization in Mediterranean semiarid, rainfed
conditions in Spain. Field Crop Research 84: 341-357.
Chaerle L, Saibo N, van Der Straeten (2005) Tunning the pores: towards engineering plants for
improved water use efficiency. Trends in Biotechnology 23 (6).
Chen C, Payne WA, Smiley RW, Stoltz MA (2003) Yield and water-use efficiency of eight wheat
cultivars planted on seven dates in Northeastern Oregon. Agronomy Journal 95: 836-843.
Clay DE R. Engel RE, Long DS, Liu Z (2001) Nitrogen and Water Stress Interact to Influence Carbon-
13 Discrimination in Wheat. Soil Sci. Soc. Am. J. 65: 1823-1828.
Condon AG, Richards RA, Rebetzke GJ, Farquhar GD (2002) Improving intrinsic water-use efficiency
and crop yield. Crop Science 42: 122-131.
Condon AG, Richards RA, Rebetzke GJ, Farquhar GD (2004) Breeding for high water-use efficiency.
Journal of Experimental Botany 55 (407): 2447-2460.
Corbeels M, Hofman G, Van Cleemput V (1998) Analysis of water use by wheat grown on a cracking
clay soil in a semi-arid Mediterranean environment: weather and nitrogen effects. Agricultural
Water Management 38 (2): 147-167
del Blanco IA, Rajaram S, Kronstad WE, Reynolds MP (2000) Physiological performance of synthetic
hexaploid wheat-derived population. Crop Science 40: 1257-1263.
El Hafid R, Smith DH, Karrou M, Samir K (1998) Physiological responses of spring durum wheat
cultivars to early-season drought in a Mediterranean environment. Annals of Botany 81: 363-370.
Farquhar GD, Richards RA. 1984. Isotopic composition of plant carbon correlates with water-use-
efficiency of wheat genotypes. Australian Journal of Plant Physiology 11: 539-552.
Febrero A, Vendrell P, Alegre L, Araus JL (1991) Epidermal conductance in flag leaves and ears of
several landraces and varieties: morphological and anatomical characteristics involved. In
Physiology/Breeding of Winter Cereals for Stressed Mediterranean Environments. Les Coloques
55: 143-157.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

204
Fischer RA, Turner NC (1978) Plant productivity in arid and semiarid zones. Annual Review of Plant
Physiology 29:277-317.
Fischer RA, Rees D, Sayre KD, Lu ZM, Condon AG, Larque Saavedra A (1998) Wheat yield progress
associated with higher stomatal conductance and photosynthetic rate, and cooler canopies. Crop
Science 38: 1467-1475.
Gebbing T, Schnyder H (2001) 13C labeling kinetics of sucrose in glumes indicates significant
refixation of respiratory CO2 in the wheat ear. Australian Journal of Plant Physiology 28: 1047-
1053.
Gregory PJ, Simmonds LP, Pilbeam CJ (2000) Soil type, climatic regime and the response of water
use efficiency to crop management. Agronomy Journal 92: 814-820.
Hatfield JL, Sauer TJ, Prueger JH (2001) Managing soil to achieve greater water use efficiency: A
review. Agronomy Journal 93: 271-280.
Havaux M, Tardy F (1999) Loss of chlorophyll with limited reduction of photosynthesis as an adaptive
response of Syrian barley landraces to high-light and heat stress. Australian Journal of Plant
Physiology 26: 569-578.
Huang Y, Chen L, Fu B, Huang Z, Gong J (2005) The wheat yields and water-use efficiency in the
Loess Plateau: straw mulch and irrigation effects. Agricultural Water Management 72: 209-222.
Hubick KT, Farquhar GD (1989) Carbon isotope discrimination and the ratio of carbon gains to water
lost in barley cultivars. Plant, Cell and Environment 12: 795-804.
Johnson RC (1993) Carbon isotope discrimination, water relations, and photosynthesis in tall fescue.
Crop Science 33: 169-174.
Kang S, Zhang J (2004) Controlled alternate partial root-zone irrigation: its physiological
consequences and impact on water use efficiency. Journal of Experimental Botany 55 (407):
2437-2446.
Karrou M (1998) Observations on effect of seeding pattern on water-use efficiency of durum wheat in
semi-arid areas of Morocco. Field Crops Research 59 (3): 175-179.
Kato T, Kimura R, Kamichika M (2004) Estimation of evapotranspiration, transpiration ratio and water-
use efficiency from a sparse canopy using a compartment model. Agricultural Water Management
65: 173-191.
Klein JD, Mufradi I, Cohen S, Hebbe Y, Asido S, Dolgin B, Bonfil DJ (2002) Establishment of wheat
seedlings after early sowing and germination in an arid Mediterranean environment. Agronomy
Journal 94: 585-593.
Lawlor DW (2002) Limitation to photosynthesis in water-stressed leaves: stomata vs. metabolism and
the role of ATP. Annals of Botany 89: 871-885.
Leuning R, Condon AG, Dunin FX, Zegelin S Denmead OT (1994) Rainfall interception and
evaporation from soil below a wheat canopy. Agricultural and Forest Meteorology 67(3-4): 221-
238
Li ZZ, Li WD, Li WL (2004) Dry-period irrigation and fertilizer application affect water use and yield of
spring wheat in semi-arid regions. Agricultural Water Management 65: 133-143.
Lpez-Castaeda C, Richards RA, Farquhar GD (1995) Variation in early vigor between wheat and
barley. Crop Science 35: 472-479.
Merah O, Delens E, Moneveaux P (1999) Grain yield, carbon isotope discrimination, mineral and
silicon content in durum wheat under different precipitation regimes. Physiologia Plantarum 107:
387-394.
Merah O, Delens E, Souyris I, Nachit M, Monneveux P (2001) Stability of carbon isotope
discrimination and grain yield in durum wheat. Crop Science 41: 677-681.
Monneveux P, Reynolds MP, Trethowan R, Gonzlez-Santoyo H, Pea RJ, Zapata F (2005).
Relationship between grain yield and carbon isotope discrimination in bread wheat under four
water regimes. European Journal of Agronomy 22: 231-242.
Morgan JA, LeCain DR (1991) Leaf gas exchange and related leaf trait among 15 winter genotypes.
Crop Science 31: 443-448.
Morgan JA, LeCain DR, McCaig TN, Quick JS (1993) Gas exchange, carbon isotope discrimination,
and productivity in winter wheat. Crop Science 33: 178-186.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

205
Oweis T, Zhang H, Pala M. (2000) Water Use Efficiency of Rainfed and Irrigated Bread Wheat in a
Mediterranean Environment. Agronomy Journal 92: 231-238.
Passioura J (2004) Increasing crop productivity when water is scarce- from breeding to field
management. In New directions for a diverse planet Proceedings of 4th International Crop
Sciences Congress, Brisbane, Australia.
Pate JS (2001) Carbon isotope discrimination and plant water-use efficiency. In Stable Isotope
Techniques in the Study of Biological Process and Functioning of Ecosystems (Unkovich et al.
eds), Kluwer Academic Publishers, The Netherlands, pp 19-36.
Pereira LS, Oweis T, Zairi A (2002) Irrigation management under water scarcity. Agricultural Water
Management 57: 175-206.
Polley WH (2002) Implications of Atmospheric and Climatic Change for Crop Yield and Water Use
Efficiency. Crop Science 42: 131-140.
Poorter H (1989) Interspecific variation in relative growth rate: on ecological causes and physiological
consequences. In Causes and consequences of variation in growth rate and productivity of higher
plants? (Lambers et al. eds.), Academic Publishing, The Hague, The Netherlands, pp.45-68.
Rebetzke GJ, Condon AG, Richards RA, Farquhar GD (2002) Selection for reduced carbon isotope
discrimination increases aerial biomass and grain yield of rainfed bread wheat. Crop Science 42:
739-745.
Rebetzke GJ, Botwright TL, Moore CS, Richards RA, Condon AG (2004) Genotypic variation in
specific leaf area for genetic improvement of early vigour in wheat. Field Crop Research 88: 179-
189.
Reynolds MP, van Ginkel M, Ribaut JM (2000) Avenues for genetic modification of radiation use
efficiency in wheat. Journal of Experimental Botany 51: 459-473.
Reynolds MP, Nagarajan S, Razzaque MA, Ageeb OAA (2001). Heat tolerance. In Application of
physiology in wheat breeding (Reynolds MP, Ortiz-Monasterio JI, McNab A eds.), CIMMYT,
Mxico DF. pp. 88-100.
Richards RA, Rebetzke GJ, Condon AG, van Herwaarden (2002) Breeding opportunities for
increasing the efficiency of water use and crop yield in temperature cereals. Crop Science 42:
111-121.
Richards RA, Condon AG, Rebetzke (2001) Traits to improve yield in dry environments. In
Application of physiology in wheat breeding (Reynolds MP, Ortiz-Monasterio JI, McNab A eds.),
CIMMYT, Mxico DF. pp. 88-100.
Poorter H (1989) Interspecific variation in relative growth rate: on ecological causes and physiological
consequences. In Causes and consequences of variation in growth rate and productivity oh higher
plants (H Lambers ed.), SPB Academic Publishing, The Hague, The Netherlands, pp. 45-68.
Rekika D, Nachit MM, Araus JL, Monneveaux P (1998) Effects of water deficit on photosynthesis rate
and osmotic adjustment wheats. Photosynthetica 32 (1): 129-138.
Richards RA, Condon AG, Rebetzke (2001) Traits to improve yield in dry environments. In
Application of physiology in wheat breeding (Reynolds MP, Ortiz-Monasterio JI, McNab A eds.),
CIMMYT, Mxico DF. pp. 88-100.
Royo C, Villegas D, Garca del Moral LF, Elhani S, Aparicio N, Rharrabti Y, Araus JL (2002)
Comparative performance of carbon isotope discrimination and canopy temperature depression as
predictors of genotype differences in durum wheat yield in Spain. Australian Journal of Agricultural
Research 53: 561-569.
Sadras V (2002) Interaction between rainfall and nitrogen fertilization of wheat in environments prone
to terminal drought: economic and environmental risk analysis. Field Crop Research 77: 201-215.
Shaheen R, Hood-Nowotny RC (2005) Effect of drought and salinity on carbon isotope discrimination
in wheat cultivars. Plant Science 168: 901-909.
Shangguan ZP, Shao MA, Dyckmans J (2000) Nitrogen nutrition and water stress effects on leaf
photosynthetic gas exchange and water use efficiency in winter wheat. Environmental and
Experimental Botany 44: 141-149.
Siddique-KHM, Tennant D, PerryMW, Belford RK (1990) Water use and water use efficiency of old
and modern wheat cultivars in a Mediterranean-type environment. Australian-Journal-of-
Agricultural-Research 41(3): 431-447.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

206
Skovmand B, Reynolds MP, Delacy IH (2001) Searching genetic resources for physiological traits
with potential for increasing yield. In Application of physiology in wheat breeding (Reynolds MP,
Ortiz-Monasterio JI, McNab A eds.), CIMMYT, Mxico DF. pp. 17-28.
Snyder KA, Richards JH, Donovan LA (2003) Night-time conductance in C3 and C4 species: do
plants lose water at night? Journal of Experimental Botany 54 (383): 861-865.
Tardy F, Crach A, Havaux M (1998) Photosynthetic pigment concentration, organization and
interconversions in a pale green Syrian landrace of barley (Hordeum vulgare L. Tadmor) adapted
to harsh climatic conditions. Plant, Cell and Environment 21: 479-489.
Tavakkoli AR, Oweis TY (2004) The role of supplemental irrigation and nitrogen in producing bread
wheat in the highlands of Iran. Agricultural Water Management 65: 225-236.
Teare ID, Sij LW, Waldren RP, Goutz SM. 1972. Comparative data on the rate of photosynthesis,
respiration and transpiration of different organs in awned and awnless isogenic lines of wheat.
Canadian Journal of Plant Science 52: 965-972.
Turner NC (2004) Agronomic options for improving rainfall-use efficiency of crops in dryland farming
systems. Journal of Experimental Botany 55 (407): 2413-2425.
Van Den Boogaard R, Alewinjnse D, Veneklaas EJ, Lambers H (1997) Growth and water-use
efficiency of 10 Triticum aestivum cultivars at different water availability in relation to allocation of
biomass. Plant Cell and Environment 20: 200-210.
Van Den Boogaard R, Veneklass EJ, Lambers H (1996) The association of biomass allocation with
growth and water use efficiency of two Triticum aestivum cultivars. Australian Journal of Plant
Physiology (now Functional Plant Biology) 23 (6): 751-761.
Voltas J, Romagosa I, Muoz P, Araus JL (1998) Mineral accumulation, carbon isotope discrimination
and indirect selection for grain yield in two-rowed barley grown under semiarid conditions.
Yunusa IAM, Sedgley RH, Belford RK Tennant D (1993) Dynamics of water use in a dry
mediterranean environment I. Soil evaporation little affected by presence of plant canopy.
Agricultural Water Management 24 (3) 205-224.
Zhang H, Oweis TY, Garabet S, Pala M (1998) Water-use efficiency and transpiration efficiency of
wheat under rain-fed conditions and supplemental irrigation in a Mediterranean-type environment.
Plant and Soil 201: 295-305.
Zaharieva M, Gaulin E, Havaux M, Acevedo E, Monneveux P (2001) Drought and heat responses in
the wild wheat relative Aegilops geniculata Roth: potential interest for wheat improvement. Crop
Science 41: 1321-1329.

























Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

207




























































Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

208






























PRODUCTIVITY OF THE POTATO CROP
UNDER IRRIGATION WITH LOW QUALITY WATERS



N. Ben Mechlia
*
, K. Nagaz
**
, J. Abid-Karray
*
, M.M. Masmoudi
*

*Institut National Agronomique de Tunisie.
**Institut des Rgions Arides, 4119 Mdenine, Tunisia.


SUMMARY - Opportunities to enhance agricultural productivity in arid Tunisia have been the domain
of active investigation during the last years. Many field experiments were carried out on the potential
of using limited amounts of saline water from wells to grow profitably crops. This paper concerns the
performance of the potato crop gown in commercial farms under relatively stressful conditions. The
FAO-Water Balance calculation method was applied for irrigation scheduling over three contrasting
seasons. Results obtained from full and deficit irrigation treatments show that under optimum
irrigation, marketable yields varied from about 40 t/ha to 20 t/ha , and water productivity from 11.7 to
9.1 Kg/m
3
, respectively for the spring and autumn periods of the year. We propose here a simple
linear relationships between yield (Y, t/ha) and total water application (Q, mm) to be used for quick
appraisals of attainable yields of the potato crop when waters of about 3-3.5 dS/m are used. Within a
range of 150 to 350 mm total water supply after planting the models are [Y= 0.1Q+5] for the spring-
summer season and [Y=0.1Q-5] for the autumn-winter season. These models could be valuable tools
for estimating yield gaps and potential for water savings concerning the potato crop grown under arid
environments.

Key words: irrigation scheduling, potato yield, water productivity, salinity.


INTRODUCTION

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

209
Potato is considered relatively susceptible to salinity (Maas and Hoffman, 1977) and normally is
not suited for stressful conditions. However its cultivation has been gaining popularity during the last
decade in arid areas as a cash crop because temperature and radiation conditions allow for cropping
over the spring, fall, and winter seasons.

In the Southern part of Tunisia where mean annual rainfall is less than 200 mm, potato is
cultivated primarily on shallow wells in private farms where water is limiting not land. Irrigation is
typically applied on a routine basis without scheduling. Surveys carried out recently in the governorate
of Medenine show that potato yield vary usually between 10 and 20 t/ha. Inadequate management of
irrigation has been identified as an important limiting factor (Nagaz and Ben Mechlia, 2003, Nagaz et
al. 2004). In farms where cultivation is practised under drip irrigation water savings by drip seem to be
forfeited with inappropriate scheduling.

Yield is greatly influenced by timing, amount and frequency of irrigation applied, therefore, precise
knowledge of the amount of water required by the crop and the proper timing for supply is essential.
Scheduling based on crop water requirements and soil characteristics allows for applying irrigation
water when needed during the growing season. However, its application is only possible when water
supply and irrigation amounts can be managed independently by farmers (Smith, 1985). In areas
where potato is irrigated with well waters, accurate scheduling is manageable. This is precisely the
case of our area, therefore chances to achieve tangible productivity improvements are high.

Field trials were implemented with the objective to evaluate the applicability of representative
irrigation scheduling methods for drip systems. Basically, the investigation had to quantify yield, water
use efficiency and soil salinity when the soil water balance method is used instead of the prevailing
common approach. With the expectation to enable growers to incorporate more appropriate irrigation
scheduling methods in their usual production practices, all field work was conducted with farmers
participation.



MATERIALS AND METHODS

Irrigation-scheduling experiments were carried out during the years 2000 to 2004 in commercial
farms situated in the south-eastern part of Tunisia near the Institut des Rgions Arides, IRA-
Mdenine. The potato cultivar "Spunta" was used following standard cultural practices as reported by
Nagaz et al. (2004) over three distinct growing periods: spring, autumn and winter seasons.

For the irrigation-scheduling experiment, the soil water balance (SWB) methodology was adopted.
A spreadsheet program estimates the day when the soil readily available water (RAW) would be
depleted and estimates the amount of irrigation water needed to replenish the soil profile to field
capacity. The target depletion threshold was set to 35 % of total available water in the root zone
(TAW).

The program calculates the soil water depletion with potato root depths estimated on daily basis.
The soil depth of the effective root zone is increased with the program from a minimum depth of 0.15
m at planting to a maximum of 0.60 m in direct proportion to the increase in the potato crop
coefficient. Once the maximum root depth is reached, it is held constant.

The soil is of a sandy type with low organic matter content. The total soil available water,
calculated between field capacity and wilting point for an assumed potato root extracting depth of 0.60
m, is 75 mm.

The crop evapotranspiration (ETc) was estimated for daily time step by using reference
evapotranspiration (ET0) combined with a potato crop coefficient (Kc), following the FAO-56 method
given in Allen et al. (1998). Weather data used in this experiment (Tmax, Tmin and u2) were collected
daily from a nearby meteorological station located at IRA-Medenine.

The development of quantitative relationships between yield and water supply was an integrated
part of this study. To produce experimental data, the layout included four distinct water treatments
with decreasing amounts in relation to maximum crop evapotranspiration (ETc). Full irrigation is the
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

210
treatment that received 100 % of accumulated ETc. Deficit treatments were irrigated at the same
frequency as the control, but with quantities equal to 80, 60 and 40 % of accumulated ETc .

The producer method consisted in the supply of fixed amounts of water of about 17 mm to the crop
every 5 days from planting till harvest. This method corresponds to irrigation practices traditionally
implemented by local farmers using drip irrigation.

A randomised block design with four replications was used as trial layout. Water, obtained from a
well with a conductivity of 3.25 dS/m, was applied by means of polyethylene dripper lines with an
emission rate of 4 l h
-1
. For each block, it passed through a water meter, gate valve, before passing
through laterals placed in every potato row. A control mini-valve in the lateral permits use or non-use
of the dripper line.

Before planting, the soil was set to field capacity over a depth of 0.6 m i.e. all treatments received
a start amount of 75 mm. Our investigation was concerned with the effect of different water
management tactics corresponding to decreasing irrigation water supplies on yield, water saving and
soil salinization.

Potato yields were estimated from samples of ten plants per row within each plot. Soil samples
were collected after harvest and analyzed for ECe.


RESULTS

The SWB scheduling techniques based on limited weather information and standard data provided
in FAO guidelines was effective for managing irrigation under ordinary farming conditions. Maximum
yields obtained under this scheduling technique (Yo) were respectively 39.7, 30.4 and 22.7 t/ha for
spring, autumn and winter crops.



Yield water supply relationships

Reduction of irrigation amounts resulted in proportional reduction of yield (Figure 1). Because all
treatments were uniformly at field capacity at the beginning of the cropping seasons, the impact of
irrigation reduction varied slightly from one season to the other, depending on the contribution of soil
reserves to the overall water supply to plants. Intuitively one can assume that the least irrigated
treatment would rely more on soil water initial store, provided that sufficient roots were developed
over the wetted profile. The little amounts of rainfall waters although relatively limited can also affect
the plant water status and therefore mitigate to different extents the effect of irrigation restrictive
regimes.


Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

211


Fig. 1. Relative yield decrease of the potato crop in relation to reduction of irrigation, for different
production seasons in Southern Tunisia.

Figure 2 represents yields obtained under situations of full and deficit irrigation regimes with saline
waters. Two simple linear relationships are identified for the spring and for the autumn-winter
seasons. The proposed quantitative models are:
Y = 0.1 Q + 5, for spring productions (1)
Y = 0.1 Q - 5, for the autumn-winter season (2)
with Y the yield of fresh potato tuber in t/ha, and Q the total water supply during the growing
season, including the little amounts of rainfall. As empirical relationships, these equations correspond
to situations where applied water is comprised between 150-350 mm limits, water salinity about 3
dS/m and rainfall input representing less than 20 % of the total supply. It is worthy to notice that these
results were obtained on sandy soils not significantly affected by salinity (ECe less than 4 dS/m).

The obtained models have different intercepts because of differences between growing conditions
in spring and Autumn seasons. Temperature and radiation regimes affect the partitioning of dry
matter between tuber and vegetative pert. High temperatures and low radiation are known to be
favorable to vegetative development (Autumn), whereas cool conditions associated with important
radiation loads (spring) are suitable for more tuber production. Our results show that the ratio
between dry matter of tubers and total dry matter biomass is 10-13% higher for spring productions
than for the other two seasons.




Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

212


Fig. 2. Potato yields obtained with irrigation waters having an ECi of 3.25 dS/m over three contrasting
growing seasons in Medenine, southern Tunisia. The soil is sandy with ECe lower than 4 dS/m
and rainfall represented about 6 to 20% of the total supply. Prior to planting, all treatments
received 75 mm to put the top 60 cm soil layer at field capacity.

Marginal gains in fresh tuber yield associated with increased water supply seem to be similar for
the various cropping periods. The slope of the idealized linear relationships indicates an overall water
productivity of 100 kg ha
-1
mm
-1
. This value fall within ranges commonly reported in the literature
(Bowen 2003).


Water use efficiency and potential for water saving

Amounts of irrigation water and total water supply for the two growing seasons are presented in
Table 1. With the producer method more water was used than the SWB. Surplus was about 63 mm.

The water use efficiency (WUE) expressed as the ratio of potato yield to water supply from
planting to harvest varied typically from 11 to 9 Kg/m
3
(Table 1). These values are comparable to
those obtained in other field studies (Bowen, 2003; Fabeiro et al., 2001; Ferreira et al., 1999).
Maximum values were observed during the spring season (11.7 Kg/m
3
) under full irrigation with the
SWB method. Since water used to prepare the field for planting (75 mm) is not included in the
calculation of total water supply, caution should be used in analyzing productivity under deficit
irrigation. It would be normal that with differential water reduction levels variable amounts of the soil
readily available waters are used.

Actually water use efficiency (WUE) was significantly affected by deficit irrigation treatments. The
highest values occurred in the 100 % and 80 % treatments. It was also observed that significant
reduction in marketable tuber size (tuber weight) were associated with the 60 % and 40 % regimes.

The lower yields and WUE values obtained by the producer are attributable to the fact that water
was applied regardless of plant changing ETc. Irrigation occurrences relates to days after planting
rather than to crop growth stages progress and to ETo values.

The relatively high yields and water use efficiency values noted under full irrigation in both
seasons indicate the high potential of the potato crop to valorize irrigation waters of limited quality,
provided that good management is applied.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

213

Table 1. Water supply from planting to harvest and water use efficiency (WUE) measured under SWB
scheduling and grower's practices.




Soil Salinity

The electrical conductivity (ECe) values measured before planting were, respectively, 1.35 and
3.45 dS/m (0-60 cm depth of soil) for spring and autumn seasons. Under the different irrigation
scheduling methods, ECe in both seasons remained lower than ECi of the irrigation water. Low ECe
levels were observed after rainfall events. The leaching by rain occurred mainly during September,
October and December in autumn season (72 mm), and in April and May in spring season (29.7 mm).
Under the prevailing conditions leaching seemed to be sufficient to control the build-up of soluble
salts in the profile of this well-drained soil. Singh and Bhumbla (1968) observed that the extent of salt
accumulation depended on soil texture and reported that in soils containing less than 10 % clay the
ECe values remained lower than ECi.

Differences between treatments within a given season were also observed. With SWB full
irrigation, the average ECe value was equal to 1.0 dS/m, beneath the emitter in autumn and to 0.75
dS/m in spring season. The zone of highest ECe was moved out to 20 cm from the emitter. With the
producer method values of 1.9 and 1.7 dS/m were recorded below the emitter, respectively, in the
spring and autumn seasons.


CONCLUSION

The potential of irrigation scheduling to improve yield and to save water has been demonstrated in
this work. Results, obtained under actual farming conditions, support the practicality and usefulness
of using the Soil Water Balance (SWB) scheduling method as simplified by FAO to optimise irrigation
in arid regions.

Using well waters having an ECi of 3.25 dS/m, it was possible to produce potato at about 20 to 40
t/ha, depending on the period of the year. Scheduling can ensure a water productivity of 10 Kg/m
3
of
water, in spite of the stressful environmental conditions.

The proposed quantitative linear relationship between yield and water applied from planting to
harvest could be used as a tool for selecting the appropriate irrigation tactic to respond to irrigation
water shortages after planting.
Y = 0.1 Q + 5, for spring productions
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

214
Y = 0.1 Q - 5, for the autumn-winter season
These model can be also used to investigate the causal factors for potato productivity variations
and yield gaps in areas similar to Southern Tunisia.


REFERENCES

Allen R.G., Perreira L.S., Raes D., Smith M., 1998. Crop evapotranspiration: Guidelines for computing
crop water requirements. Irrigation and Drainage Paper N 56, FAO, Rome, Italy, 300p.
Bowen W. T., 2003. Water productivty and potato cultivation. In Water Productivity in Agriculture :
Limits and Opportunities for Improvement. Ed. J. W. Kinje, R. Baker and D. Molden. CAB
international, pp.229-238.
Fabeiro C., Martin De Santa Ollala F., De Juan J.A., 2001. Yield and size of deficit irrigated potatoes.
Agriculture Water Management, 48, pp.255-266.
Ferreira T.C., Malheiro A.N.C, Freixo F.A.M.F.P., Bernardo A.A.S, Carr M.K.V., 1999. Variation in the
response to water and nitrogen of potatoes (Solanum tuberosum L.) grown in the highlands of
Northeast Portugal. In: Proceedings of 14th Triennial Conference of the European
Association for Potato Research, Sorrento, Italy, May, 2-7, 1999, 410-411.
Maas E.V., Hoffman G.J., 1977. Crop salt tolerance: Current assessment. J. Irrig. Drain. Div. Am.
Soc. Civ. Eng., 103,pp.115-134.
Nagaz K., Ben Mechlia N., 2003. Caractrisation de la conduite de pomme de terre en irrigu dans
les primtres privs sur puits de surface. Unpublished data.
Nagaz K., Masmoudi M, Ben Mechlia N., 2004. Etude de la Rponse de la pomme de terre (Solanum
tuberosum L.) de saison, de primeur et darrire-saison aux rgimes dirrigation leau sale
dans la rgion de Mdenine. 4th AgroEnviron 2004 int. Symp., Udine-Italy October 20-24,
2004.
Singh B. and Bhumbla D.R. 1968. Efeect of quality of irrigation water on soil properties. J. Res.
Punjab Agri. Univ., 5, 166.
Smith M., 1985. Irrigation scheduling and water distribution. In: Les besoins en eau des cultures.
Actes de Confrence Internationale, INRA, Paris, France, 11-14 Septembre 1984, pp.497-
514.
























Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

215

USE OF THE HEAT DISSIPATION TECHNIQUE
FOR ESTIMATING THE TRANSPIRATION OF OLIVE TREES



J. Abid Karray
*,**
, M.M. Masmoudi
*
, J.P. Luc
**
, N. Ben Mechlia
*

*
Institut National Agronomique de Tunisie
**
Institut de Recherche pour le Dveloppement, UR-DIVHA.


SUMMARY - The heat pulse method is used on olive trees for determination of transpiration. It
consists of generating heat pulses at a point in the trunk and detecting their effect in another point. In
spite of its usefulness its precision has not been demonstrated since the integration from one point
measurement to the trunk cross section area could result in large errors. The more recently
developed heat dissipation method, which is based on measuring temperature differences between
heated and non heated probes, has been used on many forest species. Because it uses long probes
(20 mm to 80 mm) it allows for better integration of the radial flow. The objective of this work is to
investigate the practicality of using this technique for sap flow measurements of olive trees. To this
end an experiment has been carried out on twelve year olive trees grown in central Tunisia. Four
trees were selected for transpiration monitoring under changing water regime. On each tree three
probes placed at different exposures were used, North (N), South East (SE) and South West (SW).
Measurements of all twelve probes were collected every thirty minutes, over a complete growing
season. Results obtained over periods where all probes were operating properly show a good
coherence between sap flow estimates and environmental conditions. For the 95 days, during which
all sensors were operational, daily values of sap flow density vary from 10 to 120 l dm-2 d
-1
, about
one to five ratio is observed between sensors. The mean value of sensors installed on each tree is
highly correlated with the 12 sensors average values. When considering each direction separately, it
seems that flux density is not related to the direction of the sensor, correlation between sap flow
density of a sensor placed in a given direction and average value within that direction is variable.
Variability of sap flow density between probes seems to be related rather to sapwood area
heterogeneity. Regression equations relating sap flow density of each probe to the 12 probes average
values were established and correlation coefficient exceed 0.85. With such calibrated equations,
estimation of olive tree transpiration from a single probe and reconstitution of missing data becomes
possible.

Key words: olive tree, transpiration, sap flow, heat dissipation technique, model.


INTRODUCTION

Estimation of transpiration is required for appropriate irrigation management, particularly for crops
with variable planting densities. In orchards, quantifying water used by trees cannot be performed
easily by methods based on water balance and micrometeorological measurements. However sap
flow methods seem to have the potential of estimating the course of transpiration flux in a continuous
manner. Presently a variety of methods are used successfully on many forest and fruit tree species.
These methods can provide direct measurements and are easily automated, so continuous records of
plant water use with high time resolution can be obtained. Sap flow measurements are also versatile
because complex terrain and spatial heterogeneity does not limit their applicability. Sap flow can be
estimated by heat-pulse, heat-balance or thermal dissipation methods. All of these techniques use the
heat as a tracer for sap movement, but they are fundamentally different in their operating principals
and each one have its merits and drawbacks.

On olives, a modified heat pulse technique (Compensation Heat Pulse Velocity, CHPV) was first
used by Moreno et al (1996) and Fernandez et al. (1997) on either roots to study hydraulic behavior
or trunks to estimate whole-tree water consumption respectively. A good agreement was found
between the transpiration determined by CHPV system and that predicted by the Penman-Monteith
equation. However, Fernandez et al (1998) reviewed the performance of the CHPV technique and
outlined its advantages and limitations. Indeed, this system is reliable and of low maintenance, it
provides information on the dynamics of both water uptake by roots and tree transpiration and
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

216
determines the effect of meteorological conditions and soil water status on both processes.
Nevertheless, the capability of the technique to estimate the tree transpiration is limited by the
considerable heterogeneity of the conductive area in mature olive trees. Because this system uses
point sampling, radial variability of sap flow results in large errors when scaling up to the whole tree.
Consequently, many probes have to be installed at different depths and azimuths to account for radial
variations.

Girorio and Giorio, (2003) used the CHPV with six gauges and found a strong correlation between
the sap flow measured by each gauge and the average of the remaining ones. Estimation of average
sap flow is therefore possible using measurement of a single probe. However, the validity of such
models requires appropriate calibration over long period to provide a high resolution response to the
physiological and environmental factors.

The heat dissipation technique developed by Granier, (1987) is based on the measurement of
temperature differences between a heated and unheated probes. Use of probes with high
temperature conductance allows integration of radial variability of flow along the probe length, but
does not solve problems related to natural temperature gradients and sap flow density across the
trunk as reported in the literature. The original method using continuous heating was therefore
modified by Do et Rocheteau (2002) to avoid natural gradient effect on measurements.

The heat dissipation technique using alternative heating is presently used in two sites in Tunisia
having contrasting climates. This paper concerns the work carried out in the arid environment of
central Tunisia on four olive trees. The objective is to investigate the variability of measurements
among probes on a given tree and the inter-variability of sap flow density among trees. Relationships
between single probe measurements and averaged values will be analyzed. It assumed that with
quantitative inter-relationships between probes and appropriate calibration good quality data could be
derived with fewer sensors and therefore an easy to manage monitoring system could be obtained.


EXPERIMENTAL LAYOUT

The experimental work was carried out in a commercial orchard at Chebika near the city of
Kairouan in central Tunisia (latitude 3537N, longitude 955W, altitude 110 m). The climate is arid
with an average annual rainfall of 280 mm. Four olive trees (Olea europea L., cv Chemlali) planted in
1993 and spaced 1111 m, having similar shape (canopy and trunk diameter) have been chosen to
be representative of the plot. Three probes are placed in each tree at three exposures: North (N),
South East (SE) and South West (SW).

The sensors, associated electronics and worksheets software have been developed locally
(Masmoudi et al, 2004). Two data loggers (Campbell CR10X) have been used for continuous
monitoring and recording of sensors signals. Measurements of the twelve sensors were collected
every fifteen minutes, over a complete growing season from 01/01/2003 to 31/12/2003.

The sensors are two cylindrical probes with 2 mm diameter and 20 mm length equipped with
heating resistance and thermocouple and associated electronic modules for current regulation and
control. Probes are inserted radially in the xylem and spaced vertically by 8 to 10 cm (Granier, 1987).
The upper probe is heated with a constant power with an alternative 15 mn heating and cooling
cycles while the lower probe is unheated. Sap flow density is calculated according to the Do and
Rocheteau equation recalibrated for olive trees.


RESULTS

A sample of 95 days of data has been selected and used for investigation. As it was mentioned
previously measurements on all 12 sensors were taken every 15 minutes, and half an hour sap flow
density were calculated and daily cumulative values were derived. A typical course of sap flow density
during a summer day is given in Figure 1.



Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

217




Fig. 1.Sap flow density (Fd) measured on four trees (T1 to T4) on three directions: North (N), South
East (SE) and South West (SW) on 31/08/05

Sap flow given by all sensors increased rapidly after 7h15 to peak at about 09h15 and decline
from 17h45 when solar radiation decreases. While there is a general consistency between the pattern
of hourly measurements, absolute values of sap flow density obtained by individual sensors varied in
high proportions. For instance, maximum value given by South Est (SE-T2) oriented sensor of tree 2
was 1.4 l dm
-2
day
-1
while the North oriented one on the same tree (N-T2) indicated 4.1 l dm
-2
day
-1
.

In order to characterize variability between sensors, daily values for individual sensors (Fd
i
) were
compared to average values of all sensors (Fd
m
). Figure 2 shows that the sensors have the same
behavior for the wide range of transpiration levels as observed over the 95 selected days.



Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

218
Fig. 2. Sap flow density (Fd
i
) of individual probes (12) in relation to mean sap flow density (Fd
m
), data
relates to 95 days covering a large range of environmental conditions.

Overall sap flow patterns seem to be related to the probe orientation or position and to the local
conditions of sap flow within the trunk. Consistency in the sensors relative change exclude any of
random variation. Indeed, Table 1 gives the regression coefficients and shows that apart from
sensors on tree 2 (T2) the regression coefficient is higher than 0.6.

Table 1. Values of slope (a), intercept (b) and R
2
of linear regression between sap flow density of
each probe (Fd
i
) and 12 probes average (Fd
m
)



As transpiration is related to solar radiation, the existence of preferential orientation of the flow
related to sun position is tested. To this end, regression equations have been established for each
tree between values corresponding to the different directions (N, SE and SW) and the average sap
flow of the considered tree. Figure 3 give results concerning the sap flow density on the three
directions (Fd
i
-T4) versus mean sap flow density for tree 4 (Fd
m
-T4).



Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

219
Fig. 3. Sap flow density for the three directions : North (N), South East (SE) and South West (SW) for
tree (Fd
i
-T4) versus mean sap flow density of tree 4 (Fd
m
-T4)

Table 2 gives the slope coefficients and R
2
of all regressions. Correlation between individual
sensors values and 12 sensors mean exceeds 0.7 except for tree 2 (T2). The slope of regression
equation indicates that the relative weight of directions is quite random, there is no influence of the
direction of the probe on average value of transpiration. The variability observed between signals
from different probes is not due to the effect of exposure but to the variability of sap flow conductivity
within the trunk section area.

Table 2. Slope coefficients (a) and R
2
of regressions between sap flow density of a given direction
(North, South Est and South West) and mean sap flow density of the correspondent tree (T1, T2, T3,
T4)



With increased number of sensors, average values will better represent the effective sap flow and
take into account the heterogeneity of flow within the cross section. As shown in figure 4 a good
correlation is observed between mean sap flow of each tree (Fd
m
-Ti) and the value obtained with all
sensors (Fd
m
). The number of three sensors seems to be a good compromise between the validity
and the complexity of the measurements.


Fig. 4. Mean sap flow density of each tree (Fd
m
-Ti) as related to mean sap flow density of all sensors
(Fd
m
)


CONCLUSION

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

220
The use of heat dissipation technique for measurement of sap flow is known as a technique which
reduces the effect of radial sap flow heterogeneity as it integrates radial variability of flow along the
probe length. The present work concerns the representatiness vity of a single or a set of sensors
when installed on one tree or on several trees.
Continuous measurements of 12 sensors installed on 4 similar trees grown under the same
conditions showed that sap flow densities obtained are coherent with climatic conditions but large
differences are observed between absolute values produced by different sensors.

It was found that behavior and relative importance of a single sensor was consistent with
reference to overall average for a wide range of transpiration. Mean values are well correlated to
individual sensor outputs. There is no apparent relationship between orientation of sensors and the
flow distribution within the trunk cross section.

A number of three sensors seems to be adequate to produce good estimates of effective sap flow
and cover heterogeneity of flow in the trunk cross section. Methodology using regression equations
seems to be appropriate and could be used for reconstitution of missing values in case of failure of
sensors or to reduce the number of sensors for long term monitoring. However the use of such
models needs calibration for local ranges of transpiration.


REFERENCES

Do F. and Rocheteau A., 2002. Influence of naturel temperature gradients on measurements of xylem
sap flow with thermal dissipation probes. 1. Field observations and possible remedies. Tree
Physiology, 22, p641-648.
Do F. and Rocheteau A., 2002. Influence of naturel temperature gradients on measurements of xylem
sap flow with thermal dissipation probes. 2. Advantages and calibration of a noncontinuous
heating system. Tree Physiology, 22, p649-654.
Giorio P. and Giorio G., 2003. Sap flow of several olive trees estimated with the heat-pulse technique
by continuous monitoring of a single gauge. Environmental and Experimental Botany, 49, p9-20.
Granier A.,1987. Mesure du flux de sve brute dans le tronc du Douglas par une nouvelle mthode
thermique. Annales des Sciences Forestires, 44 (1), p1-14.
Masmoudi M. M., Mahjoub I., Charfi-Masmoudi C., Abid-Karray J and Ben Mechlia N., 2004. Mise au
point d'un dispositif de mesure du flux de sve xylmique chez l'olivier. Revue des Rgions Arides,
ns 2004, p242-251.
Moreno F., Fernndez JE., Clothier B.E., Green S.R., 1996. Transpiration and root water uptake by
olive trees. Plant and soil, 184, p85-96.
Fernndez JE., Palomo M.J., Daz-Espejo A. and Girn I.F., 1997. Calibrating the compensation
heat-pulse technique for measuring sap flow in olive. Acta Horticulturae, 474, 2, p455-458.
Fernndez JE., Palomo M.J., Daz-Espejo A., Girn I.F. and Moreno F., 1998. Measuring sap flow in
olive trees: potentialities and limitations of the compensation heat-pulse technique. Proceedings of
the 4th Workshop on Measuring Sap Flow in Intact Plants. Zidlochovice, Czech Republic, 3-4
November, p.16-22.












Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

221



WATER RESOURCES MANAGEMENT AT THE RIVER BASIN LEVEL:
AN INSTITUTIONAL ANALYSIS



A. Billi*, A. Quarto*, E. Zini**
*University of Rome La Sapienza
** Graduated at the University of Rome La Sapienza



SUMMARY - Water resources management at the river basin scale is, under the perspective of
institutional analysis, a decentralization process aiming at maximizing socio-economic benefits related
to water resources and minimizing transaction costs, in both the implementation of new institutional
arrangements and the adaptation of existent arrangements to changing situations. After describing
the two key-principles of decentralization and participation, the current work analyzes a number of
important characteristics, factors and variables influencing the successful implementation of river
basin management decentralization processes. The intent of this paper is to provide a practical tool
for researchers and policy-makers, useful to describe on a case-by-case basis the performance of
water resources management institutional arrangements at the river basin level.

Key words: river basin management, water resources institutional arrangements, decentralization,
participation.


INTRODUCTION

Integrated water resources management at the river basin level is - since the beginning of 90s -
one of the key-themes of the world debate on water issues. The paradigm merges the two concepts
of integrated management and of river basin management.

The first, defined as integrated water resources management (IWRM), broadly refers to the
integration of the natural system and the human system into the management of water resources: the
integration is aimed at reconciling the aggregate supply and demand for water through structural and
non-structural measures, and to achieve sustainable water management under the economic, the
social and the environmental profile.
2


The second concept refers to the management of water resources at the river basin level, based
upon the consideration that the river basin area constitutes a single inter-connected natural system
(even if a complex system) hence it needs a corresponding coordination of collective decisions on
that scale. For this reason, integrated river basin management is considered a particularly suitable
form of implementation of the IWRM paradigm.

The institutional perspective for the analysis of river basin management aims to identify the
institutional arrangements associated with concrete sustainable outcomes.

Moreover, the institutional asset and the dynamics of the processes of institutional change reveal
the complex relations between managing water and the management of agriculture, socio-economic
issues, and the natural environment. Thus, it is essential to consider the sustainable integrated
management of water resources on a basin scale as a dynamic phenomenon, consisting of the
process of balancing decisions through different needs and objectives, in a specific, uniquely-tailored
way in every single scenario.


2
For the definition of the IWRM concept, see GWP (2000).
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

222
The implementation of water management at the basin scale means activating a process of
coordination/cooperation between institutions and stakeholders. The process aims to introduce or
reinforce institutional arrangements for water management at the basin level not necessarily to
create or strengthen an ad hoc institutional subject (such as a river basin organization).
The general target of the process is the maximization of socio-economic welfare related to water
resources, through coordination of multiple activities and resolution of conflicts arising between
different users about the limited resource base. It is then fundamental to identify case-by-case the
main problems to be properly faced at the basin level, and especially those connected with negative
externalities at the basin scale (e.g. water pollution, allocative problems).

Indeed, basin management is the result of interactions between different institutional subjects
and different stakeholders in facing a specific set of problems, hence it has to be identified as a highly
context-specific process.

For this reason, since it is difficult to formulate a standard setting of institutional arrangements
(that is, an institutional model) expected to fit the situation of each basin,
3
the recent literature
mainly based upon a transaction costs approach
4
aims at establishing a set of factors and variables
related to successful outcomes, as result of wide case-study experiences of institutional performance
at the river basin level in different countries.
5



KEY-PRINCIPLES: DECENTRALIZATION AND PARTICIPATION

Under the perspective of institutional analysis, river basin management is a process of
decentralization in decision-making, to be implemented with an adequate involvement of the
stakeholders in the decision-making process itself.

Thus, the two fundamental elements are:
i) The achievement of decentralization at the maximum appropriate extent
6

ii) The activation and strengthening of participation among all the different stakeholders, as
main component of the same decentralization process

As regards the first element, decentralization means organizing or re-organizing the institutional
arrangements towards managing water resources on a basin scale, a process that generally implies a
transition of powers and functions, from the central to the local level. But it does not univocally
correspond to devolution. It has to be noticed that, in implementing a decentralization process in a
specific setting, several functions can be opportunely provided by the central authorities and not
devolved to the basin level.
7
On this purpose, the literature refers to the aspects of water
management characterized by a public good nature (e.g. weather monitoring and forecasting, but
even to some extent hydrological / environmental research, or flood control).

As regards the second element, participation can take various and different forms: transparency
and accountability (participation as openness of decision-making to the public), consultation of
stakeholders, negotiation (active involvement of stakeholders in decision-making), full transition in
powers and functions from central administrative authorities to stakeholders.
8
An important role is
played by the relationships and the different situations among the stakeholders, given that strong

3
This topic is analyzed by Alaerts, with regard to river basin organizations. See Alaerts, Le Moigne (forthcoming).
4
About the transaction costs approach and his relevance for the institutional analysis of water resources
management, see Saleth, Dinar (2004).
5
See the findings of the World Bank-supported study Integrated River Basin Management and the Principle of
Managing Water Resources at the Lowest Appropriate Level When and Why Does It (Not) Work in Practice?as
reported in Blomquist, Dinar, Kemper (2005), Blomquist, Ballestero, Bhat, Kemper (2005), Bhat, Ramu, Kemper
(2005).
6
The maximum level of appropriate decentralization is defined as the need for decisions to be taken at the
lowest appropriate level (ICWE,1992).
7
As observed in Mody (2004) ..the case for decentralization as against central control is not unambiguous..
given that the devolution of functions can accompany benefits with counter-effects.

8
See on this topic Massarutto (2005).
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

223
asymmetries in resources endowment (e.g. financial or political asymmetry) may be associated with
an heterogeneous distribution of incentives to cooperative action.

As can be clearly noticed, the two elements of decentralization and participation are deeply inter-
related.

The following paragraphs will detail single aspects connected with these two fundamental
concepts, in order to provide an analytical framework to describe the institutional performance of
water resources management processes at the local/basin scale.


ANALYTICAL FRAMEWORK FOR INSTITUTIONAL ARRANGEMENTS

In order to provide an analytical framework for assessing water management institutional
arrangements in a local setting, four main categories of factors and variables will be considered.
9
The
framework will be completed with a schematic table and a set of specific questions, intended as a tool
for documental research and in-field survey of a case-study.


Contextual factors and initial conditions

The success of a decentralization process is influenced by various pre-existent factors, configuring
a smaller or greater aptitude, in each specific context, to improve the management of water resources
at the river basin scale.

As regards the early stages of the decentralization process, the first contextual factor to be
observed is the level of economic development at both the central
10
and at the local/basin level.

The level of economic development is connected with the potential activation of financial
commitment to the decentralization process from the central government and the local stakeholders:
sustainable outcomes are linked to the financial viability of the institutional arrangements for water
management, and this target will be easier to achieve where economic well-being, at both the central
and at the local level, allows the bearing of transition and ongoing costs of the process.

The central financial commitment to the decentralization process is an important starting factor,
while it is possible to observe that it is not a necessary factor (in theory, a decentralization process
can be initiated even with the sole financial commitment of local stakeholders). The central authority
can activate funding for initial implementation, in the forms of financing the devolution and/or financing
the maintenance of a set of water management-related functions considered to be better managed
centrally rather than locally.

The local financial commitment to the decentralization process is a significant factor, as connected
with the financial autonomy at the local level, one of the main components of successful
implementation of a decentralization process.

Another initial factor influencing the development and implementation of a decentralization process
is the distribution of resources between basin stakeholders. When resources are asymmetrically
distributed (in terms of financial power, rights over the water resources, or also political influence over
water allocation) it is possible that a cooperative arrangement will be less attractive for the better-
situated subjects than a non-cooperative option.
11
This element acts in a complex way, because the
most endowed stakeholders may assume, if attracted by the future benefits deriving from the basin
management option, a leadership role and strengthen the process itself rather than making it more
difficult. This leadership role can take the form of a strong financial commitment for decentralization

9
According to the theoretical framework presented in Blomquist, Dinar, Kemper (2005).
10
The central level in a decentralization process can correspond, depending on the context, to a nation, a state
(in federal countries), a region or a province.
11
When a cooperative arrangement generates significant gains for the group as a whole but results (or is
perceived as) less favourable of a non-cooperative scheme for some actors, a redistribution or compensation in
benefits can occur. The problem, with regard to the management of international river basins, has been analyzed
in Sadoff, Grey (2002).
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

224
by the main stakeholders. Thus, if extreme inequality may be detrimental to the decentralization effort,
the assumption of a leadership role by the better-situated stakeholders can foster the process itself.
The described factor, to be taken into account, behaves in a multi-directional way.

Also the social and cultural distinctions at the local/basin level are a significant contextual factor,
because they can affect communication between stakeholders, as well as trust and aptitude to
cooperation. All other factors being equal, it is expected that the greater and more contentious are
these distinctions, the more difficult it will be to develop and maintain institutional arrangements for
the management of water resources on a basin scale.

Furthermore, a contextual factor to be considered is the existence of previous experiences of
governance at the local level. It is expected that water management decentralization initiatives will be
more likely to achieve successful results in settings where there is a local experience in governing
and managing other resources and services in a cooperative way. While the main challenge is, in this
regard, to strike a balance between the central role and the local role in organizing water
management at the lowest appropriate level, this ability will also depend on the skills previously
developed in other areas of social life.


Characteristics of the decentralization process

A decentralization process implies the devolution of authority and responsibility from the centre,
and the acceptance of authority and responsibility at the local level. The occurrence of both depends
upon the way in which the decentralization takes form, and this form may affect the achievement of
successful results.

A first characteristic to be considered is the nature of the decentralization initiative. If, in theory, a
decentralization process may start from an exclusive top-down initiative (by the central government)
or, at the other extreme, bottom-up initiative (by local stakeholders), it is expected that most of the
actual settings lie in-between these two extreme examples. A decentralization initiative will be more
likely to achieve successful results where devolution is a mutually desired process, shared by basin
stakeholders and central government officials.

Furthermore, one of the main targets of a decentralization process is to obtain a deep involvement
of the stakeholders into the making of decisions. For this reason, incorporating existing local
institutions and practices is another important characteristic of the decentralization initiative. Where
the traditional institutions are involved, they play a participating and legitimating role for basin
management towards the stakeholders. Moreover, it is to be expected that the transaction costs (in
terms of time and effort) to basin stakeholders will be smaller in existing organizational forms than in
an additional set of organizational arrangements. All other things being equal, decentralization
initiatives are more likely to succeed where they involve existing governance institutions and
practices.

Another significant characteristic of the decentralization initiative is the continuity in central level
commitment for the decentralization policy. Usually, a decentralization initiative includes a transition of
authority from the central government to the local level.12 In these situations, an important element is
how the decentralization policy can survive any changes of power that may occur at the central level
during the process. Thus, all other things being equal, when there is a lack of continuity in the central
level commitment to the decentralization policy, it will be harder to achieve a successful
implementation of the process.


Central government and basin-level relationships and capacities

Coordination of central and local actions is an essential element of a successful decentralization
process. The respective capacities of central government and basin stakeholders, and the
relationships between them, are key to achieve this target.

12
Other cases, expected to be rare, are when the decentralization doesnt include any kind of transition in powers
by the central authority to the local level. In these cases, the mentioned factor is not important.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

225

In this direction, a significant factor to be considered is the extent of actual devolution from the
central to the local level. Indeed, a central government, while formally pursuing the implementation of
a decentralization policy, may act substantially only in a symbolic or even in an abandoning way for a
real devolution of authority and responsibility at the local/basin level.
13
This behaviour may also
undermine stakeholders commitment to the decentralization process. It is reasonable to observe that
the degree of actual devolution in resource management responsibilities to the local level is
associated with more or less successful results for the decentralization process.

The achievement of financial autonomy at the local/basin level is another factor to be considered
for a successful decentralization process. Financial autonomy will be better achieved when there is a
balance between the central and the local authorities in financing the process, and in managing the
financial resources. On one side, a form of financial autonomy is needed at the basin level but, on the
other, a complete transfer of financial responsibilities from the central to the local level may be
dangerous for the process itself. All other things being equal, favourable prospects of success will
occur where there is a balance in funding and control between the central government and the basin
level.

A third significant characteristic is the basin-level authority to create and modify institutional
arrangements. A decentralization process is highly context-specific, and the functions of governing,
financing, and monitoring water resources, as well as coordinating the infrastructure construction and
maintenance, have to be tailored to the specific settings of the basin area. Moreover, sustainability in
efficiently managing these functions in the long run necessitates the power to modify the institutional
arrangements in response to changed conditions. These activities will be better performed by the
local authorities, for two principal reasons: one is the high requirement of information needed; another
is the potential of this form of local autonomy to attract stakeholders and foster their involvement into
the process. For these reasons, it is expected that successful and sustainable implementation of a
decentralization process could occur where stakeholders are empowered to create and modify the
institutional arrangements. Another important element related to what mentioned is the power of local
authority to set and modify any form of cross-jurisdictional arrangement useful to efficiently implement
the process: the relevance of this factor is high in the case of water management, given that in many
cases the administrative boundaries dont match the basin or sub-basin boundaries.

With regard to central/local relationships, the distribution of central-level political influence among
local stakeholders is another significant factor of a decentralization process. In a specific context, it is
possible that the better-situated stakeholders have a stronger access than others to central
government influence: the exercise of this influence, consisting in a block or overturning of disagreed
local level decisions, can erode the stakeholders collective commitment to the decentralization
process. All other things being equal, a more successful implementation of decentralized
management will occur in settings where there is a relative symmetrical political influence of the
stakeholders upon the central government.

Another important factor to be observed is the characteristics of the water rights systems (formal
or informal rights, recognized as binding among stakeholders). The water rights can be defined at the
local level, but it is more likely that at least in some aspects these rights are defined at the central
level (as national, state or provincial rules). The nature of these systems of rights, by which the
central and the local level relate, can change the commitment of the local stakeholders to the
agreements needed by the collective action.

Furthermore, under a transaction costs profile, the adequate time for implementation and
adaptation to new institutional arrangements is a significant aspect of a decentralization process
related to central and local institutional capacities. Longevity, as well as adaptability to change in a
trial-and-error learning process are both important factors for the success of a basin management
process, although is difficult to generally establish after how much time a given arrangement can be
considered apt to achieve its expected results or to be substituted by another. If this aspect can be

13
This is distinct from the above-mentioned factor of continuity in the central action; in this case the problem lies
in the gap between formal pursuance and actual implementation of a decentralization policy by the central
authority.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

226
opportunely analyzed only inside a specific case, it can be observed that time is an influencing factor
for the successful implementation of a decentralization process.




Internal configuration of basin-level institutional arrangements

The possibility to achieve a successful implementation of decentralized water resources
management will depend on the characteristics of the institutional arrangements configured at the
basin level.

Among these characteristics, a necessary component is the presence of institutional
arrangements for the basin-level governance, by which stakeholders articulate interests, share
information, communicate and take collective decisions. Nevertheless, it should remain clear that
basin-level governance, or the presence of institutional arrangements to enable stakeholders actions
at multiple levels, doesnt imply the creation or strengthening of an ad hoc river basin organization.

Another significant characteristic of institutional arrangements at the local level is the clarity of
institutional boundaries and their matching with the basin boundaries,14 given that decentralized
water management at the basin scale is a process of collective decision-making. Unclearly defined or
mismatched boundaries create a lack of efficiency and effectiveness of collective decisions (e.g.
inadequate information, mismatch between decisions and users involved or excluded). All other
things being equal, it is reasonable to expect that successful implementation of decentralized water
management will take place where basin-level institutions have clearly defined boundaries and where
these boundaries are well-matched to the basin boundaries.

Furthermore, an important characteristic of basin-level arrangements is the recognition of sub-
basin communities of interest. With regard to this characteristic, it can be observed that the basin
system naturally configures an inter-relation of interests among users or groups, but water users and
groups have different interests: interests are likely to be different among users/groups in the various
sectors of activity (agriculture, industry, hydropower generation) or among users/groups differently
situated in the basin area (downstream/upstream users). Recognition of communities of interest can
include only representation (guaranteed participation to decisions) or even assurance that decisions
are the results of agreements reached between the different communities of interest. The recognition
of sub-basin communities of interest is not a costless practice: transaction costs are expected to
increase for the recognition of each sub-basin community, to the extent that beyond a given
threshold additional recognitions may become counter-productive. Nevertheless these recognitions,
supporting trust and reciprocity between stakeholders, are an important factor to the emergence and
sustainability of basin-level arrangements.

Among the characteristics of basin-level institutional arrangements, a significant role is
represented by the availability of fora where stakeholders can communicate and resolve conflicts.15
These instruments will function as a means to strengthen both cooperation and participation between
the different actors.

Information sharing and communication between stakeholders are important elements of water
resources management, because they reduce information asymmetries and differences of
interpretation, thus fostering cooperation between stakeholders. For this reason, the presence of
regular fora for information sharing and communication is expected to be, all other things being equal,

14
Matching the administrative (political) boundaries and the basin (natural) boundaries, is one of the most
challenging issues in water resources management. In many cases, a solution has been found in creating river
basin organizations. Nevertheless, the necessary sustainability of such institutions in the long period suggests the
opportunity to consider into single cases (especially in less developed settings), the relation between the benefits
to be obtained from a new ad hoc institutional organization, and the costs included transaction costs of
creating and managing it.
15
If Alaerts indicates how the forum is one of the most specific features of river basin organizations and, in some
cases, the essence of such organizations in Alaerts, Le Moigne (forthcoming) occasions for ensuring
participation of stakeholders at the basin level can be organized even in absence of an ad hoc river basin
organization.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

227
a factor contributing to the successful implementation of decentralized water management at the
basin level.

Moreover, the sustainability of decentralized water management depends on the presence of fora
for conflict resolution, since disagreements between stakeholders might arise in any conceivable
water resources management setting. Thus, all other things being equal, decentralized water
management is more likely to achieve sustainable results where there are fora for conflict resolution.

TABLE OF RECAPITULATION: CATEGORIES, FACTORS AND RELATED QUESTIONS

The following table represents schematically the described categories and factors as a set of
elements useful to evaluate the institutional performance of the decentralization process of water
resources management at the basin level. Related to the single factors, a number of questions are
purposed as a case-study tool for documental research and in-field survey.

Table 1. Main categories, single factors and related questions useful to evaluate the institutional
performance of the decentralization process of water resources management at the basin level
Main categories Single factors Related questions
Economic
development at the
central level
Does the national/state/regional/provincial level of
economic development allow a financial commitment to
the basin management process from the central
authorities?
Economic
development at the
basin level
Does the level of economic development at the basin level
allow a financial commitment to the basin management
process from the basin stakeholders?
Distribution of
resources between
local stakeholders
Are there consistent asymmetries in financial (or other
kind of) resources endowment among local stakeholders?
If yes, are these asymmetries expected to weaken (or
eventually to foster) the decentralization commitment?
Socio-cultural
background
Is the basin area shared by different
cultural/ethnical/religious groups?
If yes, are the relations between these groups expected to
allow cooperation or are they expected to raise
conflictuality in a local water governance process?
Contextual factors
and initial conditions
Previous
experiences of
local governance
Are there previous successful experiences of institutional
arrangements for governance at the local/basin level?
If yes, have these experiences developed local capacities
expected to be useful for water resources management?
Top-down / Bottom
up / Mutually
desired devolution
Does the decentralization process start from a top-down
central government officials initiative or from a bottom-up
local stakeholders initiative (or in-between the two cases)?
If from a top-down initiative, is this initiative likely to
activate an adequate stakeholders involvement at the
local level?
Incorporation or
involvement of
existing local
governance
arrangements
Does the decentralization initiative adequately
incorporate/involve pre-existing local institutions in the
process?
If yes, is this involvement expected to enhance
participation of local stakeholders into the process?
Characteristics of
the decentralization
process

Consistent central
government policy
commitment
Does the decentralization initiative include a transition of
authority from the central to the local/basin level?
If yes, is (or is expected to be) the decentralization policy
commitment at the central level continuous or
discontinuous with regard to changes of the political
situation ?
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

228
Extent of actual
devolution
Is there an adequate degree of actual devolution of
responsibilities from the central level to the local/basin
level, or is the central government decentralization role
only formal?
Financial resources
and autonomy at
the basin level
Is there a balance in funding/managing funds between the
central and the local/basin level?
Basin-level
authority to create
and modify
institutional
arrangements
Is the local level empowered to create institutional
arrangements / to modify them in response to changed
conditions (especially in regard to cross-jurisdictional
arrangements at the basin or sub-basin level)?
Distribution of
central level
political influence
among
stakeholders

Is there a relative asymmetrical access to central level
political influence among basin stakeholders?

If yes, are these asymmetries expected to configure
blocks or overturns of local level decisions?
Characteristics of
the water rights
system

Are there (formal or informal) local systems of water rights
and rules defined at the central level?

If yes, are these rights and rules perceived as certain and
clear by local stakeholders (otherwise there can be a lack
of stakeholders commitment to the decentralization
process)?
Central government
and basin-level
relationships and
capacities
Adequate time for
implementation and
adaptation
Is there an adequate longevity and/or adaptability to
change of the institutional arrangements?
Presence of basin-
level governance
institutions
Are there adequate institutional arrangements to enable
stakeholders collective decisions for water management
at the basin level?
Clarity of
institutional
boundaries and
match to basin
boundaries

Do the water management institutional arrangements act
into clearly defined boundaries?

Do these boundaries match the basin boundaries?
Recognition of sub-
basin communities
of interest
Is there an adequate recognition of different communities
of interest in taking decisions at the local level?
Availability of fora
for information
sharing and
communication
Are there regular fora for information sharing and
communication between stakeholders at the basin level?
Internal configuration
of basin-level
institutional
arrangements
Availability of for a
for conflict
resolution
Are there regular fora for conflict resolution at the basin
level?


REFERENCES

Alaerts, G., and G. Le Moigne (Editors), Forthcoming, Integrated Water Management at River Basin
Level: An Institutional Development Focus on River Basin Organizations. G.Alaerts, Chapter 18 -
Institutions for River Basin Management: A Synthesis of Lessons in Developing Cooperative
Arrangements
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

229
Billi, A., Y. Meroz, and A. Quarto, 2004, Water Sector and Institutional Reform, 3rd WASAMED
Workshop on Non- conventional Water Use, Cairo, Egypt, 7 11 December 2004
Bhat, A., K. Ramu, and K. Kemper, 2005, Institutional and Policy Analysis of River Basin
Management. The Brantas River Basin, East Java, Indonesia, World Bank Policy Research
Working Paper 3611, May 2005
Blomquist, W., A. Dinar., and K. Kemper, 2005, Comparison of Institutional Arrangements for River
Basin Management in Eight Basins, World Bank Policy Research Working Paper 3636, June
2005
Blomquist, W., M. Ballestero, A. Bhat, and K. Kemper, 2005, Institutional and Policy Analysis of River
Basin Management. The Trcoles River Basin, Costa Rica, World Bank Policy Research Working
Paper 3612, May 2005
GWP, 2000, Integrated Water Resources Management, Global Water Partnership Technical
Committee, TEC Background Paper 4, Global Water Partnership
International Conference on Water and the Environment (ICWE),1992, The Dublin Statement on
Water and Sustainable Development, International Conference on Water and the Environment,
26-31 January 1992
Massarutto, A., 2005, Partecipazione del Pubblico e Pianificazione nel Settore Idrico, presented at
the Conference La partecipazione pubblica nellattuazione della Direttiva Quadro europea sulle
acque, Universit Bocconi, 30 June 2005, Milano
Mody, J., 2004, Achieving Accountability through Decentralization: Lessons for Integrated River
Basin Management, World Bank Policy Research Working Paper 3346, June 2004
Sadoff, C.W., and D. Grey, 2002, Beyond the River: The Benefits of Cooperation on International
Rivers, Water Policy n.4-2002, pp. 389-403, Elsevier
Saleth, R.M., and A. Dinar, 2004, The Institutional Economics of Water. A cross-country Analysis of
Institutions and Performance, Edward Elgar and The World Bank


























Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

230

















































Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

231









THE ECONOMICS OF WATER EFFICIENCY: A REVIEW OF THEORIES,
MEASUREMENT ISSUES AND INTEGRATED MODELS



A. Billi
*
, G. Canitano
**
, A. Quarto
*

* University of Rome La Sapienza, Italy
** University of Florence, Italy


SUMMARY - The present paper focuses on the economics of water efficiency, in its theoretical
foundations, statistical developments, and interactions with hydrology. After defining the various
concepts of water use efficiency and the main techniques for its valuation, the paper presents the
main indicators that can be used for measuring economic efficiency in water use and the basic
characteristics of an integrated hydrologic-economic water accounting system. Finally, the main
issues at stake in integrating different perspectives into a comprehensive model are presented, in
order to give suggestions on how to design appropriate policies for water resources planning and
management.


Key words: water economics, river basin modelling, water efficiency.


INTRODUCTION (Section 1)

The pressures of societies and the economy, combined with traditional approaches to water
supply and management, have led to the unsustainable use of worlds freshwater resources. Indeed,
in some areas of the Mediterranean basin, it is essential to implement appropriate water-savings
policies, in order to avoid shortages, ecological degradation, and even permanent economic and
social consequences. With the growth in population and the demand for economic development,
increasing the efficiency of water use is certain to become more and more important.

Improving efficiency and increasing conservation are the cheapest, easiest, and least destructive
ways to meet future water needs. They are also the most politically- and environmentally-responsible
measures to be implemented in the sector. However, there are dissimilar perceptions on what the
determinants of water efficiency are in the end, and what policies should be implemented to pursue
the goal of efficient and wise water management.

Two main approaches, the hydrological/engineering approach and the economic/institutional
approach, have usually confronted on what methodologies and performance ratings to use to
measure water efficiency. The former approach focuses on the abstraction, storage, distribution,
treatment and disposal activities related to the hydrological cycle and its variability. These models
give rise to the recommendations of implementing supply-side measures, such as infrastructure
expansion and investment in reduction of leakages.

On the other hand, the economics of water deals with the ways to improve social gains form the
use of a scarce resource. It uses optimization techniques under alternative institutional policies, in
order to maximize the benefits of an allocation of an exogenous amounts of water in the economy.
This approach follows from the joint analysis of production and environmental costs, and of demand
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

232
conditions, and is at the basis of the introduction of demand-management policies, such as cost-
recovery, environmental taxes, water use permits tradable on special markets.

In the last decade or so, a trend has emerged toward integrating economic and institutional
considerations into complex hydrological modeling, and to represent hydrological relationships to
determine the available amount of water into economic models. Such analyses are based on the
recognition that there complex interactions exist within a territory of reference, between the users
system, represented by the economy, and the water resource system, represented by the
hydrological cycle. The ongoing integration of different perspectives into a comprehensive model is a
fundamental task in designing appropriate policies for water resources planning and management.

The present paper focuses on the economics of water use efficiency, in its theoretical foundations,
statistical developments, and contacts and interactions with the hydrologic science. Water use
efficiency includes any measure that reduces the amount of water used per unit of any given activity,
consistent with the maintenance or enhancement of water quality. Anyhow, it is when water prices
reflect the full social costs of developing supplies, that the incentives are created to use the resource
efficiently and rationally. Hence, when resources are correctly valued, reflecting its contribution to
production, the incentive exists, through the forces of supply and demand, to use those resources
efficiently though the introduction of technological change. The achievement of economic efficiency in
resource use is a major economic policy aim, for it means that the economy is approaching its
maximum in the context of available resources. The economics of water resource addresses these
issue, both on a sector basis through stand-alone analyses, and in a comprehensive manner, through
multi-objective approaches.

The purpose here is illustrative, not comprehensive or definitive. Nonetheless, these aspects are
considered of special interest to the WASAMED project, since they give suggestions on how to
integrate the different approaches in water resources modelling, planning and management in the
Mediterranean. From table 1, it is evident that total renewable waster resources pre capita vary widely
between participating countries. Moreover, water use efficiency is especially important in this region,
since many of the Mediterranean countries suffer high levels of water stress, as shown by the last two
indicators in table 1, displaying the pressure on water resources from agriculture. In some instances,
the dependency ratio, is equal to the part of the renewable water resources which originates outside
the country, is worryingly high.

The rest of the paper is organized as follows. Section 2 introduces the different meanings of
efficiency in water resources use, focusing on economics but stressing the ways the various
approaches can complement each other. It also introduces the basic concepts of water economics,
especially in the agriculture sector. Section 3 briefly analyses the basic principles of water supply and
demand, carrying out examples of stand-alone economic analyses of water resources. Section 4
examines the techniques used to assess the value of water use, focusing especially on the
agricultural production functions that are used in complex modeling. Section 5 reviews water
accounting principles and the techniques for modeling the interactions between the hydrologic and
the economic conditions and the institutional framework. It also examines the main indicators
developed to measure water efficiency and productivity in related sectors, and how they can be used
for policy-making. Section 5 draws some policy conclusions and proposes directions for future
empirical research.


A MULTI-FACED APPROACH TO WATER EFFICIENCY (Section 2)

It is commonly intended that achieving water efficiency consists of optimizing water use. Indeed,
different points of view should be considered when investigating water use efficiency. Absolute or
physical efficiency means using the least possible amount of water for any activities. Economic
efficiency seeks to derive the maximum economic benefit for the society. Institutional efficiency
qualifies the functions of an institution regarding its water-related tasks. Social efficiency strives to
fulfill the needs of the user community. Environmental efficiency looks at natural resource
conservation. Finally, technological efficiency refers to the process of finding ways for extracting more
valuable products from the same resources. Depending on the conditions of each users system,
these non-exclusive definitions of water use efficiency can be achieved simultaneously. In any case, it
is clear that efficient water use should be approached in a multi-objective, cross-sectional and
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

233
comprehensive manner. In particular, it should include the management of both supply and demand,
assigning an economic value to water resources (Garduo and Corts, 1994).






Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

234
Table 1. Basic water data WASAMED countries (Source: FAO AQUASTAT, 2003, unless specified)

Groundwater:
produced
internally (10^9
m
3
/yr)
Surface water:
produced
internally (10^9
m
3
/yr)
Overlap:
surface and
groundwater
(10^9 m
3
/yr)
Water
resources:
total internal
renewable
(10^9 m
3
/yr)
Water
resources:
total internal
per capita
(m
3
/inhab/yr)
Water
resources:
total renewable
(actual) (10^9
m
3
/yr)
Water
resources:
total renewable
per capita
(actual)
(m
3
/inhab/yr)
Dependency
ratio (%)
Ag water
withdrawal
(1998, as % of
total renewable
water
resources)
Total water
withdrawal
(1998, as % of
total renewable
water
resources)
Algeria 1.70 13.20 1.00 13.90 429.80 14.32 442.8 2.93 27.51 42.39
Cyprus 0.41 0.56 0.19 0.78 965.30 0.78 965.3 0.00 21.79 30.77
Egypt 1.30 0.50 0.00 1.80 24.53 58.30 794.4 96.91 101.20 117.20
Germany 45.70 106.30 45.00 107.00 1,297.00 154.00 1,866.0 30.52 6.05 30.55
Greece 10.30 55.50 7.80 58.00 5,284.00 74.25 6,764.0 21.89 8.42 10.46
Italy 43.00 170.50 31.00 182.50 3,182.00 191.30 3,336.0 4.60 10.46 23.19
Jordan 0.50 0.40 0.22 0.68 121.10 0.88 156.8 22.73 86.36 114.80
Lebanon 3.20 4.10 2.50 4.80 1,294.00 4.40 1,189.0 0.76 20.88 31.31
Malta 0.05 0.00 0.00 0.05 127.50 0.05 127.5 0.00 19.80 100.00
Morocco 10.00 22.00 3.00 29.00 933.60 29.00 933.6 0.00 37.97 43.45
Palestine - - - - - - - - - -
Portugal 4.00 38.00 4.00 38.00 3,773.00 68.70 6,821.0 44.69 12.82 16.39
Spain 29.90 109.50 28.20 111.20 2,704.00 111.50 2,711.0 0.27 21.74 31.96
Syria 4.20 4.80 2.00 7.00 384.10 26.26 1,441.0 80.26 72.09 75.97
Tunisia 1.49 3.10 0.40 4.19 422.20 4.595.00 462.4 8.70 47.12 57.45
Turkey 69.00 186.00 28.00 227.00 3,139.00 213.60 2,953.0 1.52 13.05 17.57
Mean 14.98 47.63 10.22 52.39 1,605.00 63.46 2,064.0 21.05 33.82 49.56
Std. Dev 21.52 64.45 14.87 72.41 1,614.00 72.38 2,162.0 30.73 29.87 35.78




Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

235

Such approach is required because both population and water are unevenly distributed, such that
different areas experience differing degrees of water stress. Moreover, keeping it constant in terms of
quantity, absolute water supply is diminishing in terms of quality. Hence, water allocation is a highly
controversial subject, particularly in arid and semiarid zones. Most water planners assign priority to
water use in the following hierarchical order: human consumption, food production and industrial
production. However, this criterion has often caused conflicts because in many countries the priority
of development strategies is not necessarily to improve the quality of life through sanitation and good
health, but rather through the development of industry and exports (food and finished products).
Above all, peoples fundamental right to clean water and sanitation at and affordable price should be
recognized. But in the past, the prevailing lack of awareness about the economic value of water has
led to its being wasted and to its use with a negative impact on the environment. Moreover, indirect
benefits such as environmental and psychological ones have not been included, as they are difficult
to quantify. In practice, the value water is considered inferior to its real worth and this gives rise to
inefficient use.

For much of the time, water management has focused on manipulating water supplies from natural
source to where it was needed. In this supply management, water is thought of as a requirement to
be met, not as a commodity the demand for which can be altered. Thus, water use efficiency has
often meant satisfying all possible demands for the resource. It is only comparatively recently that the
focus has shifted on how demand can be satisfied without massive supply developments. In
particular, the concept of allocation of water based on its values in alternative uses has become
increasingly important, and, with it, consideration of water use efficiency.

Water use in most socioeconomic activities can vary widely depending upon the interplay of many
factors. Many of these factors, such as pricing policies, comprise products of public decision-making.
Others, such as the selection of production processes, are private decisions, but again are the
product of many forces that change through time. Thus, policies and practices which lead to improved
water use efficiency bring about an array of possible choices for adapting to local circumstances. The
establishment of a dynamic balance between interventions in water supply and demand, taking into
account the variability of supply in time and space, the changes in demand, and the limits and
opportunities of technology, is the final goal of the society. This will enable to dispose of water supply
in adequate amounts for human groups with positive growth rates (increasing population) and provide
responses that are better suited to availability and demands.


Economic and institutional issues in water use efficiency

Any method of increasing water use efficiency should be subjected to a technical evaluation in
order to obtain an estimate of the actual reduction in water demand or discharge resulting from using
the method. However, economic and environmental factors are of foremost importance. Where water
development costs and competition for available capital are rising, the concept of physical or
engineering efficiency is limited by its inability to address the value of any specific use of water in
relation to alternative uses for the same water.
16
An exclusive emphasis on improving the engineering
efficiency of a given water use, therefore, may lead to unproductive expenditures if the value of that
use is less than the value of some other use of the same water. Hence, the economic efficiency
concept should be taken into account.

While the hydrological cycle is at the basis of physical calculations of water efficiency, the concept
of the factors of production is the starting point for analyzing economic efficiency. Three generalized
factors underlie all productive activities: natural resources, labour and capital. In particular, natural
resources are added to varying amounts of labour and capital, to bring about production of goods and
services.


16
Physical efficiency means using the least possible amount of water for any activities. Such technical
evaluations basically measure the ratio between water pumped into a system and water delivered to consumers
or end uses. More on physical water efficiency is presented in section 5 on water accounting and related
indicators.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

236
This general paradigm applies not only to the goods and services traded in markets (or sometimes
by other means), but also to some which may be quite far removed from market processes, like
recreation. In any event, these three factors of production combine in varied combinations to yield
products for consumption.

Economic efficiency in a production setting involves technical and allocative components.
Production is technically efficient when the maximum possible output is generated with a given set of
inputs, or when a selected output is produced at minimum cost. Allocative efficiency involves the
optimal combination of inputs and occurs when the marginal prices of each of the factor inputs are
equal. In turn, the way in which the factors combine depends upon their relative prices.
17


The economic approach to deciding the most desirable allocation of water is to use the principles
of economic efficiency in order to ensure that water is supplied to its most valuable uses. In an
economic sense, two principles or rules are the criteria which guarantee the greatest efficiency in the
allocation of a resource. These are (Agudelo, 2001):

The principle of equimarginal value, which means that the marginal benefit, or incremental value,
per unit of resource used should be equal across all uses. When equality of marginal values is
achieved, further redistribution of water can make no sector better off without making another sector
worse off. The principle, then, is that the resource should be allocated in such a way that all users or
consumers derive equal value in use from the marginal (the last) unit used or consumed. It should be
noted that this principle presupposes an homogeneous good, which is not really the case with water;
surface water, for instance, cannot be interchanged with ground water.

The principle of marginal cost pricing, which means that the marginal benefit of use of the
resource should be equal to the marginal cost of its supply. Whether an enterprise is private and
unregulated, private and regulated, or public, the condition that the price set should be equal to
marginal cost is the desired situation from the point of view of economic efficiency considerations
(provided the principle of equimarginal value is also met).

Where free competition in the economic sense exists, market processes tend to automatically
bring about this optimum. If government policies dominate price-quantity determination because of
public ownership or regulation, political processes replace market processes. When water prices are
low relative to the costs of other inputs and in relation to the costs of developing supplies, the
resource will be overused and efficiency of use will be correspondingly low.
18


Three general considerations emerge. First, the level of attention paid to water use efficiency is
directly proportional to the prices charged for water servicing. Second, rising prices lead to increasing
attention to water use characteristics, and, over the long run, to more efficient water use, improved
productivity and reallocation among users. Finally, when water prices reflect the full social costs of
developing supplies, incentives are created to use the resource efficiently and rationally, reflecting its
value in production or in its various other uses. In other words, rising prices generate powerful
incentives for increasing water use efficiency.

This last point leads directly to the issue of water institutions. The legal systems of societies are
endlessly complex, and beyond the scope of this paper.
19
Nonetheless, a few characteristics can be
pointed out which clearly affect water efficiency decisions. First of all, most nations employ systems of
building codes, which specify minimum standards that must be met in new or renovation construction.
Until recently, the matter of water efficiency has rarely formed part of these codes. However, until

17
Firms or consumers normally will tend to use relatively more of the cheaper inputs, and relatively fewer of the
more costly ones. If any of the required inputs has a very low, or zero price, to the user, then as much as possible
of that input will be used. Here lies one of the fundamental problems of water resources management. As outlined
earlier, water supplies have been cheap historically in most areas of the world, even in semi-arid areas. This
basic factor plays a major role in explaining why water usage per unit of production is high, why recycling rarely
reaches its full potential and why water usage per capita is higher in some countries than in others.
18
Basic pricing considerations also play a major role in explaining why pollution occurs. Waste removal, in the
majority of cases requires the use of environmental resources, such as water. When this input is available free of
charge, it is invariably cheaper than any other option for waste disposal. The resulting overuse leads directly to
water pollution problems.
19
For an analysis of the institutional implications of water management, see Billi, Meroz, and Quarto (2004).
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

237
codes and standards are modified, improved water efficiency will be very difficult to achieve. Similarly,
the ability to charge self-supplied water users royalties for the use of water constitutes formal legal
arrangements, which can be manipulated to induce adequate incentives.

Above all, the fundamental institutional issue underlying water use efficiency is that of the regime
of property rights. The rights to natural resources of any kind display varying degrees of ownership on
the spectrum from public to private. At the public end of the scale, access is completely open to all
citizens. The resource is essentially free for the taking. With open access, no incentive exists to
manage the resource in a conserving, efficient manner, except through moral suasion. At the other
end of the spectrum, where private ownership pertains, access to the resource belongs exclusively to
its owner, is enforceable under law and is both divisible and transferable. Under such conditions,
positive incentives do exist for effective management and efficient use.
20


The point is that water typifies common property resources, with non-exclusivity, non-enforceability
and low prices. Under these conditions, little incentive exists for conserving, efficient resource use.
Indeed, in many cases, the potential for overuse and abuse is strong, and management becomes a
very complex and difficult undertaking. But the theory goes further by suggesting that externalities,
under such conditions, will rise to socially unacceptable levels, and that, over time, the development
of private or quasi-private arrangements of rights will develop. Currently, in some parts of the world,
the development of water markets for re-allocating water supplies, and the fledgling use of effluent
discharge fees and tradable permits for pollution control reflect the growing reformation of property
rights to water. Under such conditions, the development of increasingly efficient water use practices is
an accompanying trend. The principle emerging here is that water use efficiency is partially a
response to the property rights prevailing in a society. The greater the degree of private ownership,
the greater the use of water efficient practices.

The aforementioned three key concept of water use efficiency (physical, economic and
institutional) can be seen as related to each other in a sequential way. Figure 1 exemplifies this
relationship and gives evidence of the fact that the story of changing social uses of water forms a
spiral movement, oscillating between a perceived scarcity of the natural resource, and the
implementation of the means required to overcome such scarcity. While at the beginning, engineering
solutions were put in place to overcome water scarcity, and technical productive efficiency was
sought to ameliorate the use of water, nowadays we have moved to demand management, in term of
allocative and institutional efficiency. This process has been called by Ohlsson and Turton (2000) the
turning of a screw and is displayed in Figure 1. The movement to economic and institutional
considerations has been key to facing properly water problems, since besides physical scarcity, it is
necessary to overcome the conflicts generated by competing uses, that is the so-called social water
scarcity.


Other aspects of water use efficiency

Social and political realities, technological innovations and environmental constraints in different
regions or nations of the world also play important roles in the use of water, and therefore, in
efficiency considerations. In a sense, the economic factors singled out above form a subset of these
realities. Socio-political factors are embedded into the fabric of societies. Many of them are subtle and
indirect in their effects on efficient water use, and the best that can be done is to deal with those
factors that seem to have particular importance to the issue at hand.


20
Demsetz (1967), and later Pearse (1988), have illustrated that the progression from common property to
private ownership of resources reflects a response to social cost externalities. When resources are plentiful
relative to demands, no incentive exists to develop property rights systems, and common property characteristics
apply. However, as population growth and economic growth occur, conflicts over access to the resource rise in
number and seriousness. There comes a time when the social costs, or externalities, of such conflicts rise to such
a degree that it becomes worthwhile to reform the basic property rights (a costly undertaking in itself) to bring
about increasing degrees of private ownership.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

238

Management contents Management phases Management objectives
Demand management 2
Less water
More produce
per unit of water
More value
per unit of water
User-pay,
polluter pay
Physical efficiency
Productive efficiency
Economic efficiency
Institutional efficiency
Demand management 1
Supply management 2
Supply management 1


Fig. 1 The different phases of water management strategies


The other caveat is that the social factors are quite complex, and would, if dealt with
comprehensively, merit a detailed treatment of their own. It is thought, however, that even a brief
treatment will yield a few principles and observations. The discussion that follows deals briefly with: a)
effects of social tastes and preferences, b) effects of technological innovation, and c) introduction of
environmental considerations.

The issue of social tastes and preferences is deeply embedded in societies, and may be a major
influence on the ways in which individuals and groups view the need for water use efficiency. For
example, water supply abundance creates general attitudes that water is very plentiful, and thus the
need to conserve is not felt. This makes efforts at water efficiency more difficult than in less water-
abundant areas. A further example relates to a characteristic commonly termed green lawn
syndrome. This term refers to the attitudes that residential landscaping should be green, with healthy
lawns, trees and shrubbery. These attitudes have led in the past to excessive water demands,
particularly in drier areas, with subsequent overcapitalization of water infrastructure. In drier areas,
the use of water efficient landscaping is gradually being accepted as an alternative to the green lawn
syndrome. The point here is that deeply ingrained attitudes, tastes and preferences are important
considerations in moving towards increased water use efficiency.

Technological change also has a great impact on water use efficiency. On the supply side, the
progression of technology has vastly increased the resources available, through the discovery of new
reserves and stocks. Supplies have been expanded even more by advances enabling the use of less
accessible resources, of lower quality, and lesser concentrations. Even land, though limited in the
spatial sense, has been augmented enormously in its capacity to produce crops. On the demand
side, technology has progressively reduced and eliminated our dependence on particular resources
for particular purposes. Technological innovations have more than offset the depletion of resources
through consumption: notwithstanding the economic growth of the past century, the demand for
almost all natural resources and for food has risen slower than the supply.

For present purposes, it is important to understand the forces driving all the creative technological
effort that has overcome the limits of natures endowment. Owners of land and natural resources are
constantly striving to generate the greatest possible value from them. And those who need these
resource commodities are constantly searching for cheaper sources of supply, alternative materials
that are less costly, and ways of using them more efficiently. Both suppliers and demanders, driven
by the financial incentives created by resource commodity markets, direct their creativity towards
overcoming scarcity. The lesson for water use and its efficiency is that, when resources are correctly
valued, commensurate with their contribution to productivity, the incentive exists to use those
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

239
resources efficiently though the introduction of technological change. This principle relates back to
earlier points made about the economic forces underlying water use efficiency.

Finally, the environmental approach to water use efficiency takes a broader view of the issue,
emphasizing the need for integrated approaches to its management. This is where water quality
considerations become important, as opposed to concentrating on the quantitative aspects. An
environmental view highlights the often-ignored fact that water quantity and quality are tightly
interlinked, such that actions that affect one dimension have inevitable effects on the other.
21


This principle, as suggested at the outset of the section, means that water use efficiency should
only be considered when they maintain or enhance water quality. The extent to which efficient water
use can forestall or even prevent development would be considered under environmental appraisal.


Why an economic perspective for analyzing water use efficiency?

While the classical concept of water efficiency are helpful in describing changes in the volume and
quality of water available, they are not sufficient for describing all of the economic implications of
current or alternative allocation practices. This task requires understanding in economic terms the
direct and indirect impacts of water use decisions, including opportunity costs and externalities.

The primary objective of the economic analysis of water resources is, thus, demand management
and efficient allocation among its various uses. Under growing scarcity, valuing water appropriately
and allocating it to the uses in which it has the most value, promotes rational use of scarce resources
and greater overall societal net benefit. Allocation mechanisms should resolve trade-offs and balance
competing demands, both within and between sectors, as well as between countries and regions.

A second domain in which water economics plays a fundamental role is the study of the
interdependencies of the water sector with the wider economic and social domain. This
interdependence generates externalities, which are uncompensated effects of one agent (or group) to
another. The fact that someone could be harmed by another, or get a benefit from him, without a
counterpart transaction, create inefficiencies. This calls for a careful analysis of externalities.

A third objective of the economic analysis of water is cost recovery. This means pricing water at its
full long-run marginal cost, which includes O&M costs, capital costs, opportunity costs, and costs of
economic and environmental externalities. Finally, economics is the basis for calculating financing
needs. In this sense, economic estimation of financing requirements is a precondition for rehabilitation
and improvement of sector performance, and for identifying and mobilizing additional financial
resources.

The first and second of the aforementioned objectives assumes particular importance.
22
The
economics of water has devoted much time to the study of allocative trade-offs and externalities.
While traditional measures of water efficiency can even indicate that after a certain threshold there is
little scope for saving water, there may still be significant opportunities to increase the net value
generated by limited resources. This is the role of the economic analysis of water allocation, which
aims at both improving the distribution of water among its uses, and reducing the negative external
effects of one use on the others. This helps identifying opportunities and designing policies to improve
water management practices.

In what follows, a straightforward microeconomic framework of irrigation activities, composed of a
production possibilities frontier, is presented to demonstrate how externalities and opportunity costs

21
The common wisdom is that a reduction in water use without an accompanying decline in waste generation will
cause wastes to increase in concentration. The consequences of the latter can vary in their effect on quality.
Such increases in waste concentration may overwhelm the ability of existing treatment plants to operate
effectively. On the other hand, increased waste stream concentrations might actually enhance the operation of
waste treatment systems. The point here is not so much the actual answer to this issue, which, in any event,
probably depends on local conditions, but rather the illustration of an additional principle of water use efficiency.
22
The third objective is the specific focus of section 3.1, while the fourth objective is overlooked in the present
work.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

240
can prevent a region or nation from achieving economic efficiency, even when irrigation is described
by high measures of technical water efficiency (Wichelns, 2002a).

The key concept of water productivity is to be introduced now. It is the quantity of produce (crops
or other goods) that can be obtained per each unit of water used (Molden, 1997). Water productivity
can be increased by improvements in agronomic practices, varying crop varieties, and supply and
demand management, both regionally and at the farm level. In economic words, increasing water
productivity means increasing the technical efficiency of production.

Technical and allocative efficiency can be jointly represented in a graphical setting such as that in
Figure 2. The continuous convex curve depicted in Figure 2 is the production possibilities frontier, i.e.
the locus of technically-efficient combinations of outputs that can be obtained with a given set of
resources. Points above the frontier are not feasible, while points below it could be achieved with
fewer inputs and are thus inefficient. For example, point E is technically efficient, while point F is not
feasible and point I is inefficient. The frontier also describes the technical trade-off that must be
considered when choosing crop combinations. The shape of the curve reflects diminishing
incremental returns in the production of each crop, that is the opportunity cost of augmenting the
production of one crop at the expense of the other.

Allocative efficiency describes the maximization of the net benefit, for example the revenue from
crop production, by the allocation of a resource. This is achieved in the single point where the ratio of
output prices is equal to the rate at which the output of one product (for example, cotton) must be
reduced in order to increase the output of another product (for example, rice). In Figure 2, this occurs
when the line PP, describing the ratio of output prices of cotton (PC) and rice (PR) is tangent to the
production possibilities frontier.23 Farmers interested in maximizing revenues will choose to produce
C1 units of cotton and R1 units of rice, and they will respond to any changes in output prices by
adjusting crop choices and moving along the frontier.
24




I
E
F
P
P
Cotton production, C
Rice production, R
R
1

C
1

C
2

E
2



Fig. 2. Technical and allocative efficiency in a production possibilities frontier



23
The slope of PP is P
R
/ P
C
.
24
Similarly, a production possibilities frontier can also be used to describe production opportunities throughout a
region or nation with multiple uses of water resources.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

241
The economic value of production at point E is both technically and allocatively efficient.
Conversely, an inefficient point such as I might be the result of water rationing, degraded water
quality, inability to obtain water when needed, unreliable supply, overexploitation by head-end users.

The production possibilities frontier is also useful in analyzing the effects of externalities caused by
upstream users to a downstream activity. When a negative externality occurs along the water flow, it
affects both crops and the result is the dotted production possibilities frontier depicted in figure 1. This
new frontier exemplifies a situation where waterlogging and salinization, caused by upstream
irrigation activities, reduce the feasible set of downstream production alternatives. This is reduced by
the area between the original and the shifted frontiers. The maximum value of crop that can be
produced with a given set of resources is less. For example, it is no more possible to produce the pair
(C1, R1); if the quantity of rice is to be the same, water should be diverted and cotton production must
be reduced to C2.

When it is the cultivation of one crop that generates a negative externality on another, the
production possibilities frontier assumes the shape represented in Figure 3. The rotation of the
frontier indicates that the maximum achievable production of cotton is reduced at all levels of rice
production. The revised feasible set includes the inner curve and the line segment BA. The maximum
possible output of cotton, point A, can be achieved only if rice is not produced.

B
E
1

A
Cotton production, C
Rice production, R
R
1

C
1

C
2

E
2




Fig. 3. Efficiency constraints in a production possibilities frontier


The analysis of negative externalities is a particularly important, yet very difficult, aspect of the
economic analysis of water. As noted by Perry et al. (1997), while surface runoff and subsurface drain
water can be used beneficially by downstream farmers, agricultural return flows usually are lower in
quality than the water diverted originally, due to higher concentration of salt, pesticides or nutrients. In
addition, return flows may not be available to downstream farmers at the time when the water can be
used most productively. The intensity of external effects in water use is perhaps greater than in any
other sector of the economy. This is why a careful assessment of water values requires accounting
for externalities, as explored in more detail in section 4.



Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

242




MAJOR COMPONENTS OF THE ECONOMIC ANALYSIS OF WATER RESOURCES (Section 3)

The supply and demand for water can be traced back to the so-called hydro-social water cycle
(Merrett, 1997). This relates the natural hydrological cycle to the production and consumption
activities related to water. The cycle starts from the natural flow in a catchment, which is variable both
during the course of a year, and from a year to another.
25
Then, there are the seven main phases of
supply: abstraction, storage, treatment, distribution, wastewater collection, treatment, and disposal.
These seven phases are the sources of supply costs and require major construction works. Return
flows are then released into the freshwater network, such as surface sources, ground reservoirs, and
the sea. Between freshwater distribution and wastewater treatment, consumption takes place in
several sectors of the economy.

Other sources of supply costs, that require engineering interventions, are internal and external re-
use, and recycling. Internal re-use refers to the process of treating internally the water used in the
production process, in order to employ it in the same or related processes. External re-use means
making wastewater suitable for being utilized in other uses. Recycling includes the activities related to
releasing wastewater to the freshwater network, supplementing the natural downstream flows.

After this brief overview of the hydro-social water cycle, the remaining of the section is intended to
give some other elements of the economics of water resources. The starting points are the basic
economic concepts of supply and demand. Supply is mainly analyzed in this section, while section 4
gives special focus to valuation techniques of demand components.


The economic perspective of water supply

The supply of water, from an economic perspective, is driven by the costs of constructing and
operating the infrastructure, the opportunity cost of these resources in alternative uses, and the
correction for external side effects on economic agents beyond the transaction in analysis. Figure 4 is
the well-known representation of Rogers et al. (2002) of the total cost to societies of supplying water.
Environmental externalities are given evidence as fundamental non-economic supply costs, then they
are not given special treatment in this paper.


25
The variability of the natural flow calls for the economic analysis of supply reliability, alternative sources and
storage facilities. On these issues the paper will return later.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

243


Fig. 4. Components of water supply cost

Supply costs are one of the focuses of the economic analysis of water; on these costs will be
based the discussion in this sub-section. In the literature they are usually referred to as use costs
and are distinguished in headworks costs, incurred in abstraction, storage and treatment, and
network costs, for distribution, wastewater collection and disposal. Water supply costs can be fixed or
variable (Merrett, 1997). When costs are converted into a monetary measure, one can single out
those capital expenditures that are incurred for the purchase of resources whose expected life is
greater than 12 months. They are called fixed costs and include the main construction works.
Conversely, current expenditures, paid for the purchase of resources used up routinely in the
production process. These variable costs include those for freshwater itself, materials, chemicals,
labor, and so on, required to operate and maintain the system.

The total cost function is the sum of the two and is expressed as quantitative relationship
describing the cost of supplying output in any time-period at each scale of output, from zero to the
systems theoretical capacity. It is function of the quantity of water supplied to the economic system.
In the analysis of the water industry, supply can refer, as appropriate, to any single stage of the
hydro-social cycle, or to the system as a whole. Unless otherwise specified, in the following the total
cost of water supply will be referred to. Total water costs functions are technical relationships that for
economic analyses can be approximated by a quadratic function, usually expressed in the following
form:
c bQ aQ Q TC + + =
2
) ( (1)
where TC are total costs, Q is the quantity of water, and a, b and c are the parameters of the
relationship, estimated through regression analysis. Average costs are equal to total costs divided by
the unit of water produced, that is AC = TC / Q. However, costs that are looked upon by economists
are usually expressed in marginal terms, i.e. the resources that have to be employed if capacity
needs to be expanded to produce another unit of water. The focus on marginal costs is explained by
the fact that it is the appreciation of incremental costs of getting one more unit of water, instead of the
absolute cost paid for each unit, that contributes to determining the right incentives to proper use.
Marginal costs are expressed as
Q TC Q MC / ) (
and, in the short-term, are strictly positive,
due to scarcity or capacity constraints. Hence, marginal costs tend to be increasing in the short term,
at least after a threshold.

Combining the three cost concepts gives a measure of economies of scale in water supply. This is
given by the output elasticity of total costs, which is defined as the percentage change in total costs
per unit percent change in quantity. In symbols:
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

244
ATC
MC
Q
TC
Q
TC
Q
Q
TC
TC
Q TC
=

=
,
(2)
The elasticity
Q TC,

can be lower or higher than unity, depending on whether there are


economies or diseconomies of scale respectively; or it may be equal to unity, if costs are constant all
along the relevant values. When average costs are falling, as happens in those phases of the value
chain where there are economies of scale (treatment and network operation), marginal costs are less
than average costs. For raw water abstraction, the opposite is true, since usually the closest,
cheapest sources are those which are used first. The slope of the costs curve of abstraction is,
therefore, strictly positive, since marginal costs are greater than average costs. The total cost of
supplying water can exhibit several slopes, each depending on the relative strength of the two
opposite effects. Determining whether the first case or instead the second prevails is an empirical
mater and is essential in characterizing each alternative sources of supply, whenever different
degrees of scale economies are displayed.

Another key aspect is the distinction between short- and long-run costs. In the former, increased
daily output is possible through operational changes or by organizational innovations demanding new
procedures. In the long term, additional capital equipment is required, either in new projects or for the
expansion of existing infrastructure and plants. Whenever economic analysis is used for determining
the societal cost of providing water to the whole system, the long-run perspective is required and,
therefore, the long-run marginal cost should be calculated and compared to water demand. This is
because, in order to produce even modest levels of output, major works are necessary, as long as the
infrastructure is operating close to its maximum capacity. This is called the indivisibility character of
water provision, and allow for higher elasticity of supply in the long run. This perspective is
accordingly used throughout the rest of the paper.

Increasing long-run marginal costs give rise to an upward-sloping supply schedule, since higher
costs are incurred by producers to expand the quantity of water.


Opportunity costs of water supply alternatives

The opportunity cost is defined as the value of a resource in its highest-value alternative use
(Briscoe, 1996). In this sense, a water supply project imposes two opportunity costs: that of water
resources, which could be used in alternative supply projects, and that of other resources used up in
the proposed project, that can be used elsewhere. While the latter is beyond the scope of this paper
and of many water projects, few remarks can be accounted for in this discussion about the
opportunity cost of water itself.

First, the opportunity cost of water depends on the specific attributes of the supply, in terms of
location and hydraulic connections. If water can be used only by the proposed project, it has a very
low opportunity cost, since other uses elsewhere are prevented in any case by the physical
characteristics of the source. Conversely, the opportunity cost is maximum when transfers of water
from one use to another are relatively easy to implement, and grow up as a source becomes more
densely used.

Second, the relative importance of opportunity costs is determined by the regime of property rights
enforced in each context. Where transfers of water are prohibited by law or customs, then the
opportunity costs are close to zero. On the contrary, where private markets are free to operate,
opportunity costs assume relevance since welfare maximization requires that the best supply project
alternative is chosen.

Third, opportunity costs are different in each single water use: high-valued uses impose a lower
cost on low-valued ones, than what is the case in the opposite situation.

Where opportunity costs are high, conflicts among users may arise, hence proper inter-sector and
intra-sector allocation choices have to be made, and even rationing may be an option. This is
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

245
essential to assure that water flows could be exploited to the benefits of the best alternative use. The
estimation of opportunity costs entails calculating water benefits in other alternative uses, hence it will
be explored in section 4.


The demand for water

The consumption side of the economics of water resources is represented by three major types of
economic agents: households, farmers and other firms. The residential sector is composed of
households that use water in final consumption, whereas in the agriculture and industrial sectors,
water is a raw input required for the production process. In this lays the key difference between the
residential demand, which is direct, and agricultural and industrial demand, which is indirect and
derived as requirement for the production of other final or intermediate goods. Relevant demand-side
economic sectors other then agriculture are hydropower, navigation and waste dilution.

In the present context, the focus is on water demand in the agriculture sector, which is the focus of
the WASAMED project, and is the major consumer in Mediterranean countries, as evident from the
table 2, which shows the latest available data on sectoral water withdrawals.







Table 2. Water withdrawals, total pre-capita and by sectors (Source: FAO AQUASTAT)
1998-2002 Households (%) Agriculture (%) Industry (%)
Total water
withdrawal per
capita (m
3
/inhab/yr)
Algeria 21.91 64.91 13.18 194.1
Cyprus 29.17 70.83 0.00 301.5
Egypt 7.76 86.38 5.86 968.7
Germany 12.35 19.79 67.86 570.9
Greece 16.34 80.44 3.22 708.3
Italy 18.19 45.10 36.71 771.9
Jordan 20.79 75.25 3.96 189.5
Lebanon 32.61 66.67 0.72 383.8
Malta 79.21 19.80 0.99 128.5
Morocco 9.76 87.38 2.86 419.0
Palestine
Portugal 9.59 78.24 12.17 1,121.0
Spain 13.44 68.03 18.52 869.5
Syria 3.31 94.89 1.81 1,148.0
Tunisia 13.83 82.01 4.17 271.4
Turkey 14.81 74.23 10.95 533.7
Mean 20.20 67.60 12.20 571.9
Std. Dev 18.08 22.63 18.12 342.9

Each demand type has particular estimation techniques, which will be analyzed in section 4. Here
it suffices to specify that the demand for water is based on the monetary evaluation of the benefits
that an additional unit of water provided to each agent. The inverse demand curves, in which the
quantity is a function of the price, are downwards sloping, since the benefits of an additional unit of
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

246
water are decreasing. Each sector is characterized by a specific relation between water quantity and
the benefits derived, hence the demand curves have different slopes. The total demand schedule is
the horizontal sum of the marginal benefit of water use in all the relevant sectors.

Taking agriculture as the focus sector, a generic equation for the inverse water demand schedule
of all farmers is be the following (Tsur and Dinar, 1995):

=
=
n
i
i i
p q L p D
1
) ( ) ( (3)
where the aggregate water demand D is the sum of each farmers optimal demand of water per unit of
land, qi ( p ), derived form water being an input in the production process, multiplied by the land
endowment, Li.


Welfare analysis of alternative sources of supply

In institutional settings characterized by non-competitive markets, the demand and supply
schedules can be partially independent of costs and value considerations, since other political and
social factors may play a major role in determining the price at which water is sold. Conversely, in
competitive markets, strict economic efficiency is guaranteed by the price mechanism. This can be
demonstrated by using partial equilibrium analysis of market clearing, carried out by relating water
quantity to its price in an agriculture market. The difference between increased costs and benefits
gives the net benefits to society. Efficiency equilibrium is attained at a price where supply and
demand meets. If the price is lower then the costs required of meeting the current demand, then there
is a deadweight loss, that is a decline in net benefits for the society.
26


The purpose here is to use partial equilibrium analysis to explain briefly the advantage of
estimating the supply of water form alternative sources, following Zekri and Dinar (2003). In figure 5,
the demand for water in agriculture, D, and different curves of water supply are shown, which reflect
both use and opportunity costs of supplying water. Public supply, Sp, at service level Qp is provided
at price Pp. There, the quantity of water demanded is Qp, but this quantity cannot be provided by
public supply, which is therefore exogenously fixed. The part of consumers that are not satisfied
would be willing to pay up to P1 to improve the service. Hence, there are incentives to introduce
alternative supply projects, represented by the supply schedules SA1 and SA2. Given the current
demand, the two alternative projects can provide the same quantity Q2 of water at the same price P2,
but are characterized by different elasticity of supply to price:
) / )( / ( ) / ( ) / ( Q P P Q P P Q Q
S
= = (4)
The same formulae applies to the elasticity of demand,
D

. Under public supply, social surplus is


given by the area abcQp. Social surplus with SA1 is equal to the area hbf. Hence an alternative
supply is socially justifies whenever hbf abcQp > 0, or alternatively cdf > ahdQp. Social surplus with
SA2 is equal to the area cfd. With two alternatives, the best project is given by the comparison of the
net benefits with SA1, | cdf ahdQp | with those with SA2, | cfd adQp |.


26
We point to the works of Young (1996) and Briscoe (1996) for the analytics of these results.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

247

f
c
b
P
1

P
2

Pp
h
a
Qp
a
Q
1
Qp
g
d
d
S
A2

S
A1

Sp
P
Q


Fig. 5. Partial equilibrium analysis of alternative sources of water supply (Source: Zekri and Dinar
(2003)


Estimation of supply and demand elasticities makes it possible to empirically calculate the areas
under supply and demand functions. When quantities are known, assuming an invariable demand
function, one can derive the incremental benefits. The observed quantities of public and alternative
supply give the incremental quantities,
Qp Q Q =
1
. Similarly,
Pp P P =
2
. Therefore,
the estimated elasticity can be used to specify supply and demand functions and perform welfare
comparison. In particular, the area V under the demand curve between two points of consumption,
say Qp and Q1, can be calculated in discrete terms as follows:
| | | |
) 1 (
1 1
) 1 ( ) 1 (
1
1 1
D D D
Q Q Q Q Q P V
p p D p

= (5)
Box 1. Case study of alternative supply sources in rural Tunisia
The estimation of elasticity of supply and demand for water has been carried out by Zekri and Dinar
(2003) for rural Tunisia, where the public supplier, SONEDE has expanded production in rural areas,
traditionally supplied by associations of joint use called ACI. Their main results are summarized in
table 3. According to authors estimations, the water sector in Tunisia presents conflicting results. In
1996, ACIs were more efficient than SONEDE form a cost perspective. The price-to-cost ratio, which
indicates the cost recovery rate of the supply agency or instead the level of subsidization, clearly
shows that ACIs recovered a larger proportion of their O&M costs. Moreover, while ACI members
paid 21% of the total costs, SONEDE customers paid only 18%.
However, both systems were highly subsidized, in that ACIs receive substantial public subsidies and
SONEDE used cross-subsidies among regions. Furthermore, both systems operated at a sub-optimal
scale (declining marginal costs), what is evident from the negative values of the producer surpluses.
The high consumer surpluses indicate instead that consumers would be willing to pay more for
improvements. Hence, there is scope for additional steps, such as price increases, in order to expand
supply to a more efficient scale.


Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

248
Table 3 Welfare analysis of supply alternatives in rural Tunisia, 1996 (Source: Zekri and Dinar,2003)
SONEDE ACI
Total cost per unit of water (TD/m3) 0.422 0.929
Average O & M cost of water (TD/m
3
) 0.241 0.148
Average charge per consumer (TD/m
3
) 0.190 0.195
Total bill per consumer (TD) 44.6 13.0
Price-to-cost ratio (%) 45 132
Cost elasticity of supply -1.42 -0.78 / -0.42
Price elasticity of demand -0.24 - 1.30
Total consumer surplus 2.45 5.04
Total producer surplus -1.91 -10.64
Total social welfare 0.54 -5.60


METHODS FOR DETERMINING THE VALUE OF WATER (Section 4)

In adopting new strategies for water management, a central issue concerns effective evaluative
procedures. In the definition of water conservation, two criteria have suggested that the methods
adopted must reduce water use or consumption and must also be socially beneficial. This section
examines the demand side of the economics of water, exploring the evaluation criteria that can be
used in assessing various water demand management measures.

Economic or allocative efficiency addresses the value of scarce resources available to society.
Thus, concern with the economic efficiency of water use creates a concern about net values of water
in alternative uses and whether existing institutions are flexible enough to permit the allocation of
existing supplies in such a way that society as a whole derives maximum value from those supplies.

In an economically efficient resource allocation, the marginal benefit of the employment of the
resource is equal across uses, and thus social welfare is maximized. Hence, there is a case to
understand the underlying economics of water demand and value in various economic sectors. The
starting point is figure 6, which shows the various components of water value identified by the well-
known study of Rogers, Bhatia and Huber (1998). As evident, in order to arrive at determining the full
value of water to societies, several components have to be included in the calculation.
27


A common aspect to valuation techniques is therefore the need to consider many possible benefits
that water produces. The economic value of water, in each location, each use and each time, is given
by the sum of:
o value to users, calculated on the basis of marginal value product, which is an estimate of per
unit output for a unit of water used;
o net benefits form return flows, derived from aquifer recharge during irrigation, or downstream
benefits of water diversion during hydropower generation;
o net benefits form indirect uses, derived when the water diverted for one purpose is used for
another purpose;
o adjustment for societal objectives, in order to take into consideration the wider considerations,
such as poverty alleviation, gender empowerment and food security.



27
The issue of water valuation has been also discussed during the Expert Group Meeting on Strategic
Approaches to Freshwater Management, held in Harare in 1998. The Meeting recommended considering the
value of water within the broader context of Integrated Water Resources Management (IWRM). Hence, in what
follows, a basin-wide approach to economic valuation will be used, although priority is given to valuing water for
agriculture.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

249


Fig. 6. Components of water demand value (Source: Rogers, Bhatia, Huber, 1998)


The authors also refer to the intrinsic value, reflecting environmental, social and cultural benefits,
in order to arrive to the full value of water. Nonetheless, the present paper focuses on the economic
value and presents an overview of the main techniques employed for estimating the marginal benefits
to users. In fact, when competition among uses arises, a careful examination of marginal benefits in
each use could help identifying large disparities and aid pressure for legal change in allocation rules.
In addition, marginal benefits of water use should be compared to marginal costs of water supply
proposals, in the interest of promoting economic efficiency and fiscal responsibility (Gibbons, 1986).
In general, the economic equilibrium can be achieved through the operation of price signals in a
competitive market place. Nevertheless, in the absence of market-clearing prices, there are a number
of alternative means of estimating the value of a resource.


Classification of major water use values

A long-standing debate on how to value water has led to recognizing the need to have a clear
analysis of what this means. In fact, it is widely recognized that water has traditionally been regarded
as a free resource of unlimited supply with zero cost at supply point and, at best, water users have
been charged only a proportion of the cost of extraction, transfer, treatment, and disposal. All
associated externality costs of water have been ignored and users have been offered very little
incentive to use water efficiently and not waste it (WWAP, 2003, pp. 327-8).

However, water has a value to society in all its uses. To be exact, it has several values, each
specific to each location, each use and each time. Hence, valuing water is an exercise that should be
undertaken systematically and consistently in water resource planning and management. The key
issue is the cost imposed on others by a particular use of the resource, which is called opportunity
cost.

The economic valuation of water resource starts form the premise that water can be considered a
natural asset, the value of which resides in its ability to create flows of goods and services over time
(Agudelo, 2001). Values derived from water can be mainly divided into use values, also know as
extrinsic or direct values, and nonuse values, called intrinsic or passive or existence values.
28


28
Option values, that is the desire of individuals who do not actually use water to preserve nonetheless its
integrity for future eventual uses, is also a nonuse value.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

250

In the present paper, the interest is centered around use values, that is the economic benefits
derived from direct use of water by consuming it or its services. Comparing marginal values between
sectors, in fact, allows assessing the economic efficiency of allocations among them. For a fair
comparison, adjustments are required to express values in commensurate terms of place, time,
source and quality. This can be done by further specifying the different categories under which water
use values can be classified, as in Figure 7.



Fig. 7. Specifications of water use by three criteria (Source: Agudelo, 2001)

According to their subtractability, water use values can be divided into consumptive and non-
consumptive values. If the former case, as in the main sectors of the economy, water is no longer
available for further uses.
29
Considerations about competition are paramount, since consumption by
one economic agent generate a negative externality on others, in terms of opportunity costs in
alternative projects. Complementarities also assume relevance, as water can be used repeatedly or
even simultaneously for different uses. Finally, changes in water quality are important too, since they
may differently affects beneficial uses elsewhere.
30
Conversely, non-consumptive use values include
the benefits received in the activities that leave water and its properties essentially intact. Hydropower
and navigational uses are the main examples.

Another breakdown of water uses is by location. Those uses occurring in a watercourse and
dependent on its flow characteristics are called instream uses. Those uses for which water has to be
removed from the watercourse are called offstream uses. The distinction assumes relevance since
water is a bulk commodity, that is its transportation is extremely costly per unit of water. Hence, the
location becomes a crucial element in assigning values to water, because adjustments have to be
made to reflect the site-specific nature of offstream uses. Important offstream uses are those related
to the three main economic sectors, while the main instream uses are navigation, hydropower and
waste dilution.

Finally, by their economic role, water values can be distinguished according to the level of the
value chain in which they appear, so as to distinguish between intermediate or final water uses. Water
is a final good for households and uses in navigation and waste dilution, while it is an intermediate
good in all other economic sectors, which use water as an input in the production process. This is
reflected in the different methods that can be used to estimate water use values in the different
sectors.

The focus of the paper is on agricultural uses. These can be classified, according to the
aforementioned scheme, as offstream consumptive uses of an intermediate good. Coherently, the

29
In this sense, it is essential to keep in mind the distinction between withdrawal and consumption, due to the
possibility of reuse and recycling of water withdrawn.
30
Accordingly, the waste dilution properties of water can properly be classified as consumptive uses.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

251
estimation technique used mainly in agricultural research are explored in the next sub-sections.
Beforehand, in Box 2 an overview of the trends in water uses in the two other main sectors of the
economy is provided for OECD countries.

Table 4 instead presents some indicators related to a broad estimation of potentials for expanding
water uses in agriculture in WASAMED countries. It can be firstly noted that the weight of the
agriculture sector, in terms of area cultivated and population employed, is clearly dependent on total
land and morphological characteristics. More importantly, irrigation potentials are substantially
different, implying different scopes for irrigation extensions and diverse opportunity costs of doing so.
On the other hand, those countries whose cultivated areas are close to be fully covered by irrigations
services, would face higher incremental costs of expanding irrigation. Finally, the percentage of
irrigated area on total cultivated land highlight the infrastructure-poor countries and show where
potential improvements are likely to cost less at the margin. Hence, a close economic investigation of
water resources availability, supply costs and willingness to pay, that take all those elements into
account, would be especially helpful to guiding water policies.

Box 2. Global trends in water use by sector
Registered trends in water use vary among countries and, within countries, among sectors.
Globally, agriculture is responsible for about 69% of total freshwater abstraction. The corresponding
figure for OECD countries is 45%. Agricultural demand for water is projected to increase substantially
over the next few decades, as much of the additional food that will be needed to feed the worlds
growing population is expected to come from irrigated land.
Over the past 20 years, there has been a continuous upward trend in water use for irrigation in many
OECD countries, associated with an increase in irrigated land area that has been mainly encouraged
by government investment in irrigation infrastructure and by irrigation water subsidies. For most
countries, irrigation water represents over 80% of total agricultural water use, with much of the
remainder being accounted for by livestock farming.
While agriculture is likely to remain the primary abstractor of freshwater in the near future, industry will
be the fastest growing water user overall, largely due to rapid industrialization in many non-OECD
countries. Industry is the fastest growing user of freshwater resources worldwide, and demand from
this sector is expected to more than double over the next two decades.
Industry accounts for 23% of global water abstraction, weighted towards the OECD countries but with
industrial use in developing countries growing. The most water-intensive industries include pulp and
paper, chemicals, and food and beverages. Another important emerging trend in many OECD
countries is the growing use of freshwater for cooling in electricity production.
The remaining 8% of global water abstraction is used by households.
Source: OECD (2003)

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

252

Table4. Main agriculture water indicators in WASAMED countries in 1998 (Source: FAO AQUASTAT)

Cultivated area
(arable land and
permanent
crops) (1000 ha)
Total
economically
active population
in agriculture
(1000 inhab)
Drained area as
a percentage of
cultivated land
(%)
Irrigation
potential (1000
ha)
Area equipped
for irrigation: full
control - total
(1000 ha)
Part of area
equipped for
irrigation actually
irrigated (%)
Area equipped
for irrigation as
perc of irrigation
potential (%)
Area equipped
for irrigation as
percentage of
cultivated land
(%)
Algeria 8,265 2,660 0.74 510 513 79.61 111.60 6.89
Cyprus 113 31 37
Egypt 3,400 8,475 4,420 3,422 100.00 77.42 100.00
Germany 11,997 923 485 4.043
Greece 3,846 753 1,422 36.97
Italy 11,064 1,220 2,698 24.39
Jordan 400 192 85 77
Lebanon 313 43 1775
Malta 10 2 2
Morocco 9,283 4,274 6.97 1,664 1,417 97.50 86.70 15.54
Palestine 19 12 0
Portugal 2,358 609 632 26.80
Spain 18,715 1,220 3,640 19.45
Syria 5,421 1,563 1,250 1,267
Tunisia 4,908 958 4.01 560 367 99.75 70.36 8.03
Turkey 28,523 14,697 8,500
Mean 6,790 2,508 3.91 1,721 1,329 75.37 86.52 26.97
Std. Dev 7,885 4,027 3.12 2,731 1,269 42.98 18.00 29.60





Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

253
Values inferred from the markets: are they reliable?

In section 3.4, a technique has been presented that has been used to select the best source of
water supply when one or more alternatives are possible. The analysis therein calculated the water
demand making use of regression techniques that use data from actual transactions in irrigation water
markets.

The market approach accomplishes the estimation of the value of water by means of the
observation and analysis of market transactions (rentals and sales) of either or both water rights and
land properties with irrigation facilities. In the latter case, the value is implicitly packaged in the value
of the property and is given by the difference between the price of irrigated land and the price of
comparable non-irrigated land.

It turns out that, for estimations to be reliable, well-functioning water markets and irrigation land
property markets should be in place. This possibility is ruled out if water markets are poorly
functioning or are absent at all, as seems to be the case in many countries (Briscoe, 1996; OECD,
2003). The study of the performance of water markets is outside the scope of this paper, being
already analyzed in a previous paper of the WASAMED project (Billi, Meroz, and Quarto, 2004).
However, it is worth mentioning the limitations, some technical and one theoretical, of the technique
used to estimate water values from market observations, when these market are in place.

From a technical point of view, it should be firstly noted that observed rental prices are usually
referred to the short term, whereas observed prices for perpetual water rights give a more correct
picture of long-term incremental marginal value of water. However, for the purposes of planning
horizon, the price for perpetual water rights has to be converted in annual values, using capitalization
formulas. This in turn requires the selection of appropriate planning period and interest rates, which is
a non-trivial task. Moreover, the assumption implicit in capitalization formulas that the annual value is
constant over time is also unrealistic from a long-term perspective, since inflation and variation of real
values of water can influence expectations and profitable opportunities.

From a theoretical point of view, water values estimated from market transactions may be
unreliable since they tell little about the actual shape and elasticity of demand to an expansion of the
sources of supply. In figure 8, P0 is the price observed from market transactions. However, the
historical estimated demand is known only up to this market price. Before undertaking a project, the
willingness to pay for more quantities of water is unknown. Hence, form that point onwards, the water
demand D may take either of the different shapes D1, D2 or D3 in figure 8, so the actual prices would
vary greatly. This effect is not taken into consideration in market estimates, which usually assume an
invariant demand.


D
D
1
D2
D
3
S
2
S
1
Price
Quantity
P1
P
2

P3
P
0


Fig. 8. Limitation of current market prices as a measure of water value Source: Agudelo (2000)
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

254

What precedes imply that even when prices are easily discoverable, they are an imperfect and
frequently bad-performing estimations of the values of water. This suggests that market estimations
should be complemented by valuation techniques that rely on other kind of data.
31



Non-market estimation of the value of water as an intermediate good

The framework for valuing water should link changes in the physical characteristics (quantity and
quality) of the resources to changes in the level of services or uses of water, and to how each society
value these changes. The methods usually assess on-site water values with point estimates. That is,
water valuation is a highly-focused activity that should be carried out at least at the level of the single
service or basin. Moreover, for off-stream values to be compared to in-stream values, the former
values have to be reduced by the cost of transporting or pumping and distributing water to end
users.
32


Valuation techniques are based on the concept of shadow pricing. Shadow prices are dollar
values a resource display in a given situation, when all internal and external economic (and frequently
social and environmental) factors are taken into account. Non market valuation techniques are
evolving, in response to social and environmental pressures and the desire of policy-makers to adopt
more informed decisions. Shadow pricing of water is base on four alternative approaches: the
residual imputation; the change in net income; the value added; and the alternative cost. The second
approach is a variant of residual imputation. Both are base on the estimation of the farmers
production function and use the same analytical apparatus developed in the present paper. They are
referred to as the farm budget approach. In what follows, this approach is emphasized.
33


The basic procedure estimates the farmers production functions and then infers the demand
function from the analysis of optimal water use patterns, utilizing mathematical programming
methods. The technique starts form the theory of cost-minimizing producers that use an input up to
the point in which its dollar-value marginal contribution to the production (value of marginal product
VMP) equals its price or marginal cost of supply. This is the basic point of benefit-cost analyses of
development projects.

In a model agricultural production function, a vector of m crops, Yj, is produced out of employing
irrigation water, whose quantity is QW, and n other factors of production, whose quantities are
expressed by the vector Xi. Given that, in a competitive equilibrium, the price of each input is equated
to its returns at the margin, then the price of the single non-traded input may be derived by making it
the residual claimant of the total value product.34 Therefore, solving for the price of water, its shadow
price,
*
W
P
, is equal to its value of marginal product, VMPW, which is given by:
( )
W
n
i
i i
m
j
j j W W
Q P Q P Y P VMP
(
(

= =

= = 1 1
*
(6)

The shadow price given by equation (5) is interpreted as the on-site maximum average willingness
to pay for water for that combination of crops. To derive the equivalent marginal value, a refinement
has been introduced by the change in net income approach. The value of water is the change in the
net benefits attributable to the project, which in turn is given by the difference between the with- and

31
Other aspects that prevent observation from market transactions to be a good indicator of water values are the
following: externalities are not accounted for in the individuals transactions; prices do not reflect in-stream values
or other environmental considerations; there may be imperfect competition between suppliers; imperfect and
asymmetric information may prevent efficient decision-making; and concerns about equity and conflict resolution
are left out of the analysis.
32
From a higher perspective, the value-flow concept integrates these point estimates in space and time, yielding
values of water in the different stages throughout its flow. See Seyam, Hoekstra and Savenije (2003).
33
The discussion draws heavily from Young (1996).
34
In this case, the Eulers theorem, that allows the total value of product to be divided into shares, is assumed to
hold.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

255
without-project net incomes derived from producing that combination of crops. In symbols, calling
( )

=
i i j j
P X P Y Z
the net income, with inputs is also including water, then the change
in net income is equal to
0 1
Z Z Z =
, where the subscripts 1 and 0 refer respectively to the
with- and the without-project water situations.

The main problem of using the aforementioned approaches involves specifying the production
functions, especially in the long term. This leaves room for both omitting relevant variables, and
forecasting inaccurately the levels of output associated with a given increase of an input. However,
this technique remains a useful instrument for public planning, which is usually oriented towards the
short and, at best, the medium terms. This is especially true today since computer technologies have
made available sufficient power to run mathematical programming models that make use of several
production functions. These models are important instruments where a large set of activities compete
for scarce resources. The models find the profit-maximizing set of activities given the resources
constraints. They can trace out a set of net total benefit points, from which a set of marginal values
can be derived.
35


For the sake of completeness, it is worth describing briefly the basic characteristics of the two
other approaches mentioned at the outset of this sub-section. The value added approach starts from
the net payments to primary economic resources,
36
in order to build up an input-output matrix,
organized on a sector-basis, providing a static picture of production processes in the economy. The
estimated value of water is referred to a broad sector, such as agriculture, and is equal to the residual
value added per unit of the resource, imputed after subtracting for the value added of all other
economic resources. The approach gives broad estimations of average values, hence it is less useful
in broad basin-wide efficiency analysis, than what it is in planning allocations within a sector.

Finally, the alternative cost approach is a variant of cash-flow analysis, aimed at determining the
cost of producing an output in the next-best project, and then attributing that cost as the water of
water for the proposed project. This technique is better suited for determining the least-cost option of
a supply project, than for carrying out complex analyses of efficient allocation in a river-basin context.


Economic analysis of water allocation policies

Having defined the main tools for estimating the incremental benefits of water projects, one can
turn out to analyzing a proposal to implement a particular agriculture policy, such as the expansion of
an irrigated area. The aim is to single out the true cost of the proposed project by including the
opportunity cost of water that can be used alternatively by other agents in the same sector. Many
other techniques can be applied, as highlighted before. Here, the aim is informative, hence a simple
case is analyzed, and then the results of an empirical test conducted in Egypt is acknowledged in Box
3.

The starting point is 9, which represents the optimal allocation choice between to competing
projects. The evaluation of water in each single project gives the estimations of the value of marginal
product of water inputs, which are decreasing in the quantity. The optimal choice is taken when Q1
and Q2 water is allocated to project 1 and 2 respectively. With such allocation, the marginal benefits
of water are equated across the economy and take the value of W. Societal benefits are therefore
maximized form an economic efficiency perspective with such an optimal allocation.






35
This aspect, integrated with geographic information technologies, may provide visual comparisons of the net
benefits of alternative policy alternatives. This aspect is further explored in the conclusions.
36
That are wages, capital, depreciation, rents to primary natural resources, and payments for government
services.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

256

Water volume
Q
1
Q
2

?
W

?
W

VMP
W
in project 1
VMP
W
in project 2
Water
value
Water
value


Fig. 9. Optimal allocation of a limited supply of water



Box 3. Allocation scenarios in Egypt with a farm budget approach
Wichelns (2002b) estimated the true costs of a water project in Egypt that proposes to divert water
away from the delta region. Through a small-scale simulation model of production regions, the author
calculates empirically the optimal allocation of irrigation water under different policy objectives
between three irrigation expansion projects, using a farm budget approach to maximize total net
revenues from agriculture. The main results are summarized in table 5.
The net revenue impacts of alternative water allocation policies vary with the total volume of water
available for irrigation in the three regions. That volume is expected to decline in the future, as
municipal and industrial demands increase. Two sets of scenarios are examined, pertaining to two
policy objectives: a) the supply is allocated so as to maximize the sum of net revenues generated in
the three regions; b) as supply is reduced, the volume of water is held constant in Toshka and Sinai
and reduce in the Delta. The optimal irrigation depths in table 5 are used to show the kind of
economic analysis that can be done with farm budget estimations.
Under the first policy objective, when water supply is reduced, net revenue is maximized by reducing
irrigation depths in all three regions and reducing irrigated area in Sinai, where the marginal
productivity of water is the smallest. When water supply is reduced by 10%, it is no longer optimal to
irrigate land in Sinai (the case is not reported in the table). The second policy scenario, where all
reductions in irrigated area and irrigation depths occur in the Delta, is far less attractive. In this case,
the sum of net revenues for each reduction is smaller than under the first policy scenario. The impact
of the second policy objective is reflected in the marginal value of water, which increases in the Delta,
where the net revenue from crop production declines more sharply. Reallocating water to the Delta is,
therefore, welfare improving, since it is the most productive region.
For a similar estimation of water values in Namibia, see MacGregor et al. (2000).


Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

257
Table 5 Optimal allocation of water in three district in Egypt (Source: Wichelns 2002b)
Policy
objectives
a) Maximizing the sum of
net revenues
b) Maintain full production in
Toshka and Sinai regions

Area in
prod.
(ha)
Applied
water
(cm)
Net unit
revenue
(US$/ha)
Area in
prod.
(ha)
Applied
water
(cm)
Net unit
revenue
(US$/ha)
Regions Water supply not limiting
Delta 8.3 94.9 711
Toshka 1.2 9435 435
Sinai 0.5 93.9 160
Water supply reduced by 5%
Delta 8.3 90.7 707 8.3 89.2 704
Toshka 1.2 89.6 431 1.2 94.5 435
Sinai 0.4 88.2 155 0.5 93.9 160
Water supply reduced by 20%
Delta 7.7 77.4 642
Toshka 1.2 94.5 435
Sinai 0.5 93.9 160


ACCOUNTING FOR AND MODELING EFFICIENT WATER USE (Section 5)

The starting point of basin-wide water accounting and modeling is that the water sector interacts
with all other sectors of the economy, and could potentially become a binding constraint on economic
expansion and growth. This is especially true because while the amount of renewable water resource
is practically fixed, water demands will continue to grow and diversify. Thus, the economic challenge
is to maximize social and economic benefits under varying circumstances, by efficiently using the
available resources.

Water accounting is based on the water balance approach and focused on the hydro-social cycle
described at the outset of section 3, that is on the interactions between two systems within a territory
of reference: the users system represented by the economy and the water resource system. The
territory of reference can be a country, a region or a river basin. Figure 10 describes these interaction,
both between the water system and the economy for a given territory, and their interaction with the
economies and the environment of other territories.

Water accounting is not intended to give an explicit valuation of water. Instead, it registers the
flows of water over time in and out of the physical system and the economy. Of special interest here
is the fact that water is considered a physical asset as all other economic assets, whose stock should
be calculated at the beginning and at the end of each given period. The objective is to provide a large
amount of data by disaggregating water inflows and outflows per supply source and demand sector,
so that specific water accounting indicators can be calculated, as well as more sophisticated
estimations of water use efficiency and productivity can be performed.

Modeling differs form accounting in that it explicitly deals with the issue of estimating water values
in use and productivity in alternative uses, for the purpose of optimizing its allocation at the basin
level. Models developed to this aim use mathematical programming for solving complex equation
systems, that usually involve both a simulation model for the hydrologic components, and an
optimization model for the economic components.

The main advantages and disadvantages of each of the two techniques are presented in turn in
what follows. The underlying thesis is that both tools are necessary for appropriate water planning
and management. Economic indicators that can be derived from those methods are acknowledged as
well.


Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

258
Territory of Reference
Economy
Physical Water Resource System
A
b
s
t
r
a
c
t
i
o
n
A
b
s
t
r
a
c
t
i
o
n
R
e
t
u
r
n
s
RoW
Economy
RoW
Economy
Exports
Sea
Evapo-
transpiration
R
e
t
u
r
n
s
Collection, purification
and distribution of water;
Transport via pipeline
Households
Other Industries
(incl. Agriculture)
Imports
A
b
s
t
r
a
c
t
io
n
Sewage and refuse
disposal...
R
e
tu
rn
s
Sea
Atmosphere
In situ use of
precipitation
A
b
s
t
r
a
c
t
i
o
n
R
e
t
u
r
n
s
R
e
t
u
r
n
s


Fig. 10. Main flows of water within the economy (Source: UNDESA/UNSD)


Basic features of water accounting techniques

As evident from figure 11, the first element to consider carefully in water accounting is the water
available for abstraction. This is not the net inflow of water, which is the gross inflow (effective rainfall
plus
37
surface and sub-surface flows) plus change in storage. In order to obtain the quantity of water
available for supply in a given year, the net inflow must be reduced by the amount of water flowed out
of the system because already committed to other uses, such as downstream rights or minimum
stream flows for the environment. An interesting extension, following Merrett (1997), is to allow for the
possibility of water recycling. The supply of water augmented with recycled water would then be:

=
+ =
n
p
P P C
I E S
1
(7)
where SC is the total supply corrected for recycling, E is the net inflow, n is the number of discharge
points P, IP is a specific flow of recycled water at a defined discharge point, P is a parameter that
reflect the distance of that particular discharge location from the sea, divided by the distance from the
sea of the point where that flow was abstracted. This formulae allows taking into consideration the
location of the recycling point, such that those plants that collect wastewater downstream and
discharge it treated upstream, can be attributed a higher role in total supply. This is an important
aspect when considering the valuation of the net economic benefits of a project intended to provide
treatment facilities.

Of the available water resources, part flows out of the system in both usable and non-usable form,
while part is depleted, either beneficially in a productive activity or non-beneficially, through
evaporation, deep percolation into saline aquifers, or waterlogging. In turn, beneficial depletion is
composed of both the amount depleted in the process of producing goods and services of value, or
the amount that is naturally depleted through evaporation and other causes (Molden and
Sakthivadivel, 1999).

37
Effective rainfall is equal to total rainfall in an area minus transpiration and evaporation.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

259



Fig. 11. Main water accounting relationships Source: Molden and Sakthivadivel (1999)

Water accounts can be compiled at the basin, service or use levels. Maintaining a consistency
between these various levels is an ambitious, but unavoidable task of future applied research. It is a
promising approach in that it gives evidence of the main water uses in the economy, By
disaggregating water inflows and outflows per supply source and demand sector. However, the
technique is limited in that, in order to provide meaningful economic information, it needs estimating
water values through other techniques. It can therefore be considered as a complement, not a
substitute, for more sophisticated analysis of integrated models, as those that follow.


Integrated water basin modeling: rationale and classifications

Water resources management modeling represents the most advanced tool for optimal water
efficiency, reliability and cost-effective use. Such characteristics as the inherent intricacy of aquifers
flows, the host of hydrologic uncertainties, the stochastic nature of water flows, the possibility of using
various supply sources simultaneously, the conflicts and complementarities that arise in water use, all
make complex mathematical modeling the necessary means to effective planning and efficient
management. Results from these models can be interpreted to reveal opportunities for improvement
over the status quo.
38


The hydrologic approach to water resource modeling is mainly concerned with simulating the
functioning of a hydrological system and/or optimizing water flows under a technically-inferred
objective function. On the other hand, the aim of integrated modeling is to characterize, not only the
natural and physical processes, but also the proposed projects and institutional strategies, and to
optimize for the maximum net benefit to society under a politically-inferred multiple-objective function,
whose weights reflect the underlying social values (McKinney et al., 1999).


38
Lee and Dinar (1995) discuss at long the limitations and weaknesses of this sort of models. The results depend
on the model assumptions embedded in the objective function. Hence, the degree of accuracy and specification
is usually limited. Data limitations, qualitative factors and subjective inference also play a role. Some of these
critiques are further explored in the conclusions.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

260
The task of these models is to allow generating both physical savings of water and economic
gains, by increasing the output per unit of water used. This involves both using water where it has the
most value (an allocation decision), and reducing water losses due to evapotranspiration, salinization,
pollution and percolation (an evaluation of a proposed projects benefits).

Water resources management models can be referred to the short or the long term. Short-term
models estimate the optimal combination of water quantity and quality for one year or single irrigation
season. Long-term models also account for the effects of salt accumulation in the soil profile and, in
the extended versions, in the groundwater.

Simulation model is the preferred technique to assess water resources systems response to
extreme, non-equilibrium conditions. Optimization models are based on an objective function and a
set of constraints that can include social values and institutional settings. However, integrated
hydrologic-economic optimization models always need a simulation component, in order to
characterize the hydrologic regime, although usually at a considerably simplified level.

Integrated models can also follow a compartment approach or a holistic approach. Under the
former, the main relationships that characterize the hydrologic and the economic systems are
represented as stand-alone systems of computable equations, whose output data are transferred
between the two components. In the holistic approach, there is one single unit that contains both
components, which are tightly connected to a consistent analytical framework.

Finally, it is at the basin level that hydrologic, agronomic, and economic relationships can be best
integrated into a comprehensive modeling framework. At this level, in fact, allocation decisions have
the widest economic implications. As a result, it is recommended that the design of policy
instruments, to make more economically-rational water use, is carried out at this level. The basin as
appropriate unit of analysis has long been acknowledged in international fora.


The analytical framework of hydrological-economic optimization

A river basin system is made up of three types of components: (1) sources, such as rivers, canals,
reservoirs and aquifers; (2) off-stream and in-stream demands; and (3) intermediate facilities such as
treatment and recycling plants. The conceptual framework is developed as a node-link network, with
nodes representing physical entities and links the connection between these entities (Rosegrant et
al., 2000). The nodes are the sites of the three aforementioned components. At each agricultural
demand site, water is allocated to a series of crops, according to their water requirements and
economic profitability. Figure 12 shows as example the node-link representation of the Maipo river
basin in Chile. The graphical representation also makes it easier to detect the spatial relationships
among the various elements.

Water demand is determined endogenously based on empirical agronomic production functions.
Water supply is determined through the hydrologic water balance in the river basin with corrections
for distribution to the irrigated crop fields at each irrigation demand site. Water demand and supply
are then balanced based on social objective functions, such as maximizing net benefits to water use.
The calculation of the salt concentration allows the endogenous consideration of this important
externality with respect to upstream and downstream irrigation districts.

The major component of integrated hydrologic-economic models is the representation of the
production functions for agriculture that include water as an input, and demand functions for domestic
and industrial uses, in order to estimate the uses and values of water by sector. Other types of water
demand, such as hydropower, recreational and environmental demands, can also be included,
though at the cost of complicating the analysis. The specification of the theoretic agronomic
production relationships can be done in several ways, some of which have been described in section
4. Here the concept is operationalized in terms of specific econometric estimation models drawn from
the literature.

The basic framework is a farm model of constraint optimization, where constraints are given by
physical characteristics of water and soil (given by the hydrologic model), institutional policies (prices,
property regime), and financial/investment restrictions. The objective function is maximizing the
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

261
present value of farmers profits, PA, at each demand site, over a chosen time horizon (Varela-Ortega
et al., 1998). The farmers net profits are given by crop-water functions, which relate crops yields to
water applications. One of the most used crop-water function is that of Dinar and Letey (1996), which
for each demand site defines:
( )
( )


=
= =
+
+
=
T
t
t
Wt Wt
m
j
j j j
m
j
j j j
r
P Q Tc Fc A P Y A
PA
1
1 1
1
(8)
where Yj, Pj, QW and PW indicate as before the jth crop quantity and price and applied water quantity
and price, whereas Fcj and Tcj are respectively the fixed and technological water application costs of
crop jth,
39
and the subscript t = (1, .., T) refers to the each time period. Fixed and technological costs
can be specified in technical terms and be referred to biophysical determinants, other inputs costs,
and externalities such as salinity in return flows.



Fig. 12. Schematic representation of a river basin (Maipo, Chile) (Source: Rosegrant et al., 2000

It is important to stress that PA is to be referred to each single demand site, in order to reflect
differences in hydrologic conditions. Ideally, it should also account for several cropping patterns. This
can be done by further specifying the underlying economic functions, by at the expenses of greater
model complexity, more rigidity of the hydrologic model, or its complete absence, such as the analysis
that has been applied to the case of a water-scarce basin in southern Spain, as described in Box 4.

39
One technological cost parameter is the Christiensen Uniformity Coefficient (CUC), that is used as a proxi for
both irrigation technology and management activities. The choice of water application technology can be
determined endogenously. See Rosegrant et al. (2000).
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

262
In a complex system that accounts for all water uses, the objective function solves simultaneously
for all the sectoral benefit function.
40
Moreover, the objective function may account for climate
uncertainty and varying water storage capacity, using stochastic techniques,
41
or employing more
sophisticated tools for incorporating spatial and temporal distribution of the aquifer response to
external stresses.
42


The institutional rules are the very key policy variables. They can dictate, on a command-and-
control basis, specific allocations of water resources among irrigation districts, among sectors, and
among in-stream and off-stream uses. Institutional can be set so as to resemble more or less the
market mechanisms, in order to introduce efficiency elements into the system. An even the
introduction of private property rights, as well as markets for permanent allocations and/or rental
markets, can be tested out of these models.

Box 4. Accounting for different cropping patterns in water-scarce regions
Reca et al. (2001a) consider three resolution levels in an optimization model or an irrigation system,
where water scarcity is taken as an exogenous external constraint. The objective function is to
maximize the benefits of consuming a volume of water today, or instead store it and consume more
water the uncertain future. The model is limited to a detailed economic component, hence the
hydrologic component is taken as given in the process of economic optimization.
The first sub-problem studies the optimum irrigation timing that maximizes a single crop yields. The
second component derives optimal aggregated economic functions associated to each irrigation area,
by analyzing optimal land and water allocation for all cropping patterns. Finally, the third level
addresses optimal water allocation for a complex distribution system, such as the irrigation system of
an entire basin, by taking into account the economic functions for each irrigation area.
The authors in a companion paper (2001b) apply the model to the Bmbezar system in the
Guadalquivir river basin in southern Spain, using data produced ah hoc from field irrigation
evaluations. A deterministic analysis has been carried out in order to compare optimum water and
cropping patterns management with actual ones, in a stochastic environment that determines water
availability.
The authors conclude, among others, for the superiority of water markets in reducing consumption
levels.


INDICATORS OF WATER USE EFFICIENCY AND WATER PRODUCTIVITY

The overview of water economics principles and techniques developed in the previous sections
would not be given full significance unless remarks would be made about the appropriate combination
of water data and modeling outputs. The purpose of the present section is providing the reader with
insights about the most useful indicators of water efficiency used in irrigation analysis and policy
making. Indicators are usually expressed as fractions, that is to say ratios between physical measures
of water inflows and outflows. Several comments and remarks arisen in the literature are provided.


Technical indicators of efficient water use

The first class of ratios considered herein as a benchmark for reference are the technical irrigation
efficiency, which is adimentional (Palacio-Vlez, 1994). Efficient water use, Ef, is defined as the ratio
between the actual volume of water used for a specific purpose, Vu, and the volume extracted or
diverted from a supply source for that same purpose, Ve. Functionally expressed:
Ve
Vu
Ef = (9)

40
See for example the water policy analysis of the Mekong river basin in Ringler et al. (2004).
41
See the work of Mejas et al. (2004) applied to an irrigation district in southern Spain.
42
See for example Faisal et al. (1997).
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

263
Efficiency in the use of water in irrigation may be separated into three components: storage,
conveyance, and irrigation efficiency. Storage efficiency, Es, is the ratio between the volume of water
diverted for irrigation, Vd, and the volume entering a storage reservoir for that same purpose, Ve. In
symbols:
Ve
Vd
Es = (10)
Conveyance efficiency, Ec, is the ratio between the volume of water diverted to irrigation plots, Vp,
and the volume diverted from the supply source, Vd, that is:
Vd
Vp
Ec = (11)
The classical irrigation efficiency, Eu, is defined as the volume of water beneficially used net of
evapotranspiration, Vu, divided by the volume of water diverted, Vd, or analogously:
Vd
Vu
Eu = (12)
When large volumes of surface runoff or deep percolation are generated during irrigation events,
the value of irrigation efficiency Eu tends to be low, even if a proportion of the drainage water is used
by other farmers. This generates potentials for misinterpreting those low values.
43
Keller et al., 1996
introduced the concept of effective efficiency, EE, as:
Vp
Et
E
E
= (13)
where Et is the net crop evapotranspiration and Vp is the net volume of water diverted to a field that
reaches the irrigation plots.

Water that becomes usable surface runoff and deep percolation is subtracted from the total
volume delivered. Effective efficiency can be also adjusted to reflect water quality. An aggregate
version of the effective efficiency is the concept of basin or global efficiency, whose value increases
when farmers reuse drainage water (Seckler, 1996).

The total efficiency of water use for irrigation, Ei, is given by the product of (10), (11) and (13), that
is:
E c S i
E E E E = (14)
A common deficiency of physical indicators of water efficiency is that they may generate
misperceptions among policy analysts, who may interpret higher values as preferable to lower values,
without examining the economic implications of alternative allocation scenarios. Yet, without
considerations of economic variables it is not possible to determine if higher irrigation efficiency
generates greater net economic value to society.


Economic indicators of water efficiency and productivity

The economic indicators of water use efficiency and water productivity take their rationale from the
discussion of the previous sections. They are usually expressed as ratios of physical and economic
variables referred to a specific time period, though many economic indicators are based on more
complex statistical elaborations and actualization of multi-period relationships.

Several indicators have already been given in the previous sections, in particular the measure of
economies of scale in water supply (equation (2)) and the elasticity of supply and demand (equation
(4)), which have an explanatory meaning in themselves, apart from being used in more sophisticated
analyses.


43
When a unit of water in a water basin is diverted from a source to a particular use, three basic things happen to
it. First, a part is lost to the atmosphere because of evaporation or evapotranspiration. Second, part of the
diverted water may drain to the sea, a deep canyon, or a similar sink where it cannot be captured and reused, in
which case it is truly lost to the system. Otherwise, the drainage water flows back into a stream or to other surface
and subsurface areas where it can be captured and reused as an additional source of supply. This water is not
lost or wasted in physical terms (Seckler, 1996).
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

264
The first other key indicator that can be taken as an element in the objective function is water
productivity. In principle, water productivity can be calculated for any sector or sub-sector of the
economy and is equal to a sectors dollar-value net benefits per unit of time, divided by water used
measured in cubic meters per unit period of time. Water productivity can also be calculated at the
basin or service level, in order to measure the overall impact of waster policies. Maximizing water
productivity across sectors and services within a basin is a major policy objective. Thus, recalling that
Zj represents the net benefits for user/sector j and that QWj is the quantity of water used up in the jth
production process, water productivity PRW can be written as:

=
=
m
j Wj
W
Q
Zj
PR
1
(15)
Other economic ratios come from water accounting techniques, which relate annual output to land
and water, and provide the basis for comparison of irrigated agriculture, are: output per cropped area,
Yca, output per unit irrigation supply, Yds, and output per unit water consumed, Ywc. Respectively,
the following ratios express these relationships (Sakthivadivel et al., 1999):
A
Y
Yca = (16)
Vd
Y
Yds = (17)
Et
Y
Ywc = (18)
where Y is the output of the irrigated area in terms of gross dollar value of production at local prices,
A is the sum of the irrigated areas (in ha) under crops during the time period of analysis, Vd and Et
have the usual meaning of net volume of water diverted to a field, and evapotranspiration. For
international comparison, it is useful to convert all values of production in a standardized term, in
order to reflect differences in local prices and of tastes throughout the world. To obtain this value,
equivalent yield is calculated based on local prices of the crops grown, compared with the local price
of the predominant, locally grown, internationally traded base crop, such as wheat. The Standardized
Gross Value of Production, SGVP, is equal to:
world
crop b
i
i i
P
P
P
P A SGVP
|
|
.
|

\
|
=

(19)
where Yi is the yield of crop ith, Pi is its local price, Ai is the area cropped with crop ith, Pb is the local
price of the base crop, and Pworld is the value of the base crop traded at average world market price.

More detailed analyses can be performed when a comprehensive modeling framework is put in
place. In this case, on the basis of the irrigation water demand functions developed in the previous
sections, the profit per unit of water consumed at each demand site, PUWdm, and the same at the
basin level, PUWb, can be calculated as follows (Rosegrant et al., 2000):


=
time
t dm t dm
dm
dm
RF WD
PA
PUW
, ,
(20)
WDP
PA
PUW
dm
dm
b

= (21)
where PAdm is the net profit from irrigation given by equation (8) of section 5.3, whereas WDdm and
RFdm are respectively water withdrawal and return flows from demand sites, and WDP is total
irrigation water depleted, all taken from a water accounting matrix.

Finally, some indicators can be useful in assessing the degree of cost recovery and financial
sustainability of basin water systems, such as irrigation services. Two main indicators assume
relevance: the Cost Recovery from Water Billing, RCR, and the Cost Recovery Rate, CRR (RUB,
2002). The first indicator relates the total revenues from billing, TR, to the estimated theoretical cost
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

265
of reaching sustainability of technical systems, Csts. The second indicator relate the total revenues,
TR, minus the subsidy provided, S, to the total actual cost of providing the service, TC. In symbols:
100
|
.
|

\
|
=
sts
C
TR
RCR (22)
( ) 100

=
TC
S TR
CRR (23)


Policy-related water indicators and indices

Indicators and indices for water policy-making are instruments of simplification in that they
summarize large amounts of measurements to a simple and understandable form, in order to
highlight the main characteristics of a system. Information is reduced to its elements, maintaining the
crucial meaning fro the questions under consideration. Though the aggregation causes a loss of
information, if the indicator is planned properly, the loss will not gravely deform the results. A
fundamental different exists between indicators and indices (WWAP, 2003).

An indicator, comprising a single data (a variable) or an output value from a set of data
(aggregation of variables), describes a system or process such that it has significance beyond the
face value of its components. It aims to communicate information on the system or process. The
dominant criterion behind an indicators specification is scientific knowledge and judgment.

An index is a mathematical aggregation of variables or indicators, often across different
measurement units so that the result is dimensionless. An index aims to provide compact and
targeted information for management and policy development. The problem of combining the
individual components is overcome by scaling and weighting processes, which will reflect societal
preferences.

A plenty of indicators exist for evaluating the effectiveness and impact of water policies.
44
The
focus here is put on a tool that uses the water accounting technique and poverty analysis to examine
water use in relation to specific social goals. The Water-Poverty Accounting Framework (WPAF) is
one such tools that gathers the different aspects of water management and use, in specific relation to
poverty. The WPAF expands on the water accounting technique outlined in section 5.1, in order to
account for how water is used to meet social and economic goals, in particular poverty alleviation.
The approach has a bias towards water used for agriculture, sanitation and nutrition, since these are
the most significant source of employment for the poor (Biltonen and Dalton, 2003).

Using the WPAF, water allocations required to meet different poverty dimensions can be analyzed
for each specific use. The desired and actual situations are then compared to determine options for
reallocating water to meet social goals. A set of indicators, based on current and target allocations,
has been developed to show the efficiency of water use to meet different demands.

There are two classes of indicators:
- adequacy ratios, which indicate how well either current or future needs are being
met; and
- bias indicators, which show the bias of water allocations either toward or away from
meeting certain social goals.

The indicators show where surplus water is available for reallocation and where additional water is
required to meet other goals. It is possible to compile a set of indicators for any area, country or
region.
45


CONCLUSIONS AND PROPOSALS


44
For a comprehensive review, see WWAP (2003) and WaterStrategyMan Project (2004).
45
For the details about the indicators, we point to the original work of Biltonen and Dalton (2003).
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

266
This concise review has tried to show how fundamental the economic analysis of water is, and
could further be, in order to improve water use efficiency and water productivity. The discussion
started from the fundamental reasons at the basis of water economics, going ahead through the
major techniques developed to assess water supply costs and water demand values, with a focus to
the agriculture sector, which is the main topic of the WASAMED project. The paper concludes for the
superiority of the choice to push the accounting and modeling analyses of water resources, since they
can generate a large amount of useful information and can be integrated into user-friendly policy
tools.

There are barriers to the effective use of integrated river basin management modeling, including
informational, physical, and application barriers (Lee and Dinar, 1995). Insufficient data, limitations,
and poor information about the cultural, social and political norms often limit the development of an
effective planning model. Moreover, because basins are irregular and receive flows from multiple
sources, difficulties arise when attempting to dived a basin into discrete, manageable subunits.
Temporal and special variability also complicate matters. Application barriers are due to the fact that
such models are usually formulated and applied to a particular area, they are finely calibrated to
address specific problems, and any modifications can be undertaken with additional costs.

However, models are by definition intended to abstract from reality, so the best model is the most
adapted to available data and the most transparent in terms of social values embedded in the
objective function. Trade-offs exist in model specification. But while an overly simple specification
may yield insufficient information to address the problem or unreliable results, a more complex model,
that would require extensive data collection, is not always a more preferred option then a simple,
cost-effective modelling approach.

However, given the demonstrated usefulness of integrated models, two complementary strategies
might be suggested. On the one hand, regular data collection should be improved, to allow for the
progressive introduction of more complex models. Agricultural extension services can be a useful tool
for finding information on farm possibilities and constraints, in so helping specifying production
functions to be used in estimations of water use values. Setting up a comprehensive Water
Information System in each country, that uses environmental-economic accounting practices, is a key
tool. This in developing countries may require foreign aid, in the form of training, technical assistance
and financing of collection stations.

On the other hand, it is necessary to refine and adapt existing tools, and create new models
tailored on specific circumstances. Future directions in integrated modelling include the combination
of Decision Support Systems (DSS) and Geographic Information Systems (GIS) into comprehensive
Spatial Decision Support Systems (SDSS). DSS are interactive programs with a graphical interface,
which embed simulation and optimization models to support users in problem solving. GIS offer a
spatial representation of water resources systems using existing datasets. SDSS integrate spatial
representations and modelling capacity into a single operational framework. Even the Water-Poverty
Accounting Framework, referred to in section , 6.3, can be integrated into a geographical information
system, with indicators separated into a range of categories and assigned graduated color codes. In
this manner, maps can be constructed that demonstrate the current conditions as related to any of the
preferred indicators.

Reinforcing these trends would be a decisive step towards helping decision makers with problems
that have a spatial dimension, such as water allocation. The combination of reliable datasets, robust
models, maps, and statistical analysis components, in fact provides water planners with effective and
comprehensive support for taking informed decisions.


REFERENCES

Agudelo, J. I., 2001, The economic valuation of water. Principles and methods, Value of Water
Research Report Series No. 5, IHE Delft, The Netherlands, August
Allan, T., 1999, Productive efficiency and allocative efficiency: why better water management may
not solve the problem, Agricultural Water Management, Volume 40, Issue 1, 1 March, Pages 71-
75
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

267
Billi, A., Meroz, and A. Quarto, 2004, Water sector and institutional reform, 3rd Workshop on Non-
conventional Water Use, Cairo, Egypt, 7 11 December 2004
Biltonen, E., J. A. Dalton, 2003, A Water-Poverty Accounting Framework: Analyzing the Water-
Poverty Link, Water International, Vol. 28, No. 4, pp. 467-477, December
Boscha, D. J., V. R. Eidmanb, and L. K. Oosthuizen, 1987, A review of methods for evaluating the
economic efficiency of irrigation, Agricultural Water Management, Volume 12, Issue 3, April,
Pages 231-245
Dinar, A., and J. Letey, 1996, Modeling economic management and policy issues of water in irrigated
agriculture, Westport, Connecticut: Praeger Publishers
Faisal, I. M., R. A. Young, and J. W. Warner, 1997, Integrated Economic-Hydrologic Modelling for
Groundwater Basin Management, Water Resources Development, Vol. 13, No. 1, pp. 21-34
Garduo, H., and F. Arregun-Corts, eds., 1994, Efficient Water Use, Proceedings of the
International Seminar on Efficient Water Use, Mexico, October, UNESCO Regional Office for
Science and Technology for Latin America and the Caribbean, available at:
http://www.unesco.org.uy/phi/libros/efficient_water/tapaefus.html
Gibbons, D. C., 1986, The Economic value of Water, Resources from the Future, Washington D.C.
Gleick, P., et al., 2003, Waste Not, Want Not: The Potential for Urban Water Conservation in
California, Pacific Institute, November
Grimble, R. J., 1999, Economic instruments for improving water use efficiency: theory and practice,
Agricultural Water Management, Volume 40, Issue 1, 1 March, Pages 77-82
Keller, A., J. Keller, and D. Seckler, 1996, Integrated Water Resource Systems: Theory and Policy
Implications, Research Report 3, International Water Management Institute, Colombo, Sri Lanka
Lee, D. J., and A. Dinar, 1995, Review of Integrated Approaches to River Basin Planning,
Development, and Management, World Bank Policy Research Working Paper, No. 1446, April
Mac Gregor, J., S. Masirembu, R. Williams, and C. Munikasu, 2000, Estimating the Economic Value
of Water in Namibia, paper presented at the 1st WARFSA/Waternet Symposium Sustainable Use
of Water Resources, Maputo, 1-2 November 2000
McKinney, D. C., X. Cai, M. W. Rosegrant, C. Ringler, and C. A. Scott, 1999, Modeling Water
Resources Management at the Basin Level: Review and Future Directions, SWIM Paper No. 6,
International Water Management Institute, Colombo, Sri Lanka
Mejas, P., C. Varela-Ortega, and G. Flichman, 2004, Integrating agricultural policies and water
policies under water supply and climate uncertainty, Water Resources Research, Vol. 40,
W07S03
Merrett, S., 1997, Introduction to the economics of water resources. An international perspective,
Rowman & Littlefield Publishers
Molden, D., and R. Sakthivadivel, 1999, Water Accounting to Assess Use and Productivity of Water,
Water Resources Development, Vol. 15, No. 1-2, pp. 55-71
Morris, B L, Lawrence, A R L, Chilton, P J C, Adams, B, Calow R C and Klinck, B A., 2003,
Groundwater and its Susceptibility to Degradation: A Global Assessment of the Problem and
Options for Management, Early Warning and Assessment Report Series, RS. 03-3. United Nations
Environment Programme, Nairobi, Kenya
OECD, 2003, Improving Water Management. Recent OECD Experience, OECD Publications, Paris
Ohlsson, L., and A. R. Turton, 2000, The Turning of a Screw. Social Resource Scarcity as a Bottle-
Neck in Adaptation to Water Scarcity, ECD News, Swedish International Development
Cooperation Agency (Sida), available at: http://www.edcnews.se/Reviews/Turningofascrew-1.html
Palacio-Vlez, E., 1994, Water Use Efficiency in Irrigation Districts, in H. Garduo and F. Arregun-
Corts, 1994, op. cit.
Perry, C. J., M. Rock, and D. Seckler, 1997, Water as an Economic Good: a Solution or a
Problem?, Research Report 14, International Water Management Institute, Colombo, Sri Lanka
Reca, J., J. Roldn, M. Alcalde, R. Lpez, and E. Camacho, 2001a, Optimization model for water
allocation in dficit irrigation systems. I. Description of the model, Agricultural Water Management,
Vol. 48, pp. 103-116
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

268
Reca, J., J. Roldn, M. Alcalde, R. Lpez, and E. Camacho, 2001b, Optimization model for water
allocation in dficit irrigation systems. I. Application to the Bmbezar irrigation system, Agricultural
Water Management, Vol. 48, pp. 117-132
Ringler, C., J. von Braun, and M. K. Rosegrant, 2004, Water Policy Analysis for the Mekong River
Basin, Water International, Vol. 29, No. 1, pp. 30-42
Rogers, P., R. Bhatia, A. Huber, 1998, Water as a Social and Economic Good, Technical Advisory
Committee (TAC), Technical paper n 2, Global Water Partnership (GWP)
Rosegrant, M. W., C. Ringler, D. C. McKinney, X. Cai, A. Keller and G. Donoso, 2000, Integrated
economichydrologic water modeling at the basin scale: the Maipo river basin, Agricultural
Economics, Volume 24, Issue 1, December, Pages 33-46
RUB, 2002, Methodology Report on the Quantitative Analysis of Water Systems, WaterStrategyMan
Project, Deliverable 7, prepared by Ruhr-University Bochum, November
Sakthivadivel, R., C. De Fraiture, D. J. Molden, C. Perry, and W. Kloezen, 1999, Indicators of Land
and Water Productivity in Irrigated Agriculture, Water Resources Development, Vol. 15, No. ,
pp. 161-179
Seckler, D., 1996, The New Era of Water Resources Management: From Dry to Wet Water Savings,
Research Report 1, International Water Management Institute, Colombo, Sri Lanka
Spulber, N., and A. Sabbaghi, 1994, Economics of Water Resources: From Regulation to
Privatization, Kluwer Academic Publishers
Sullivan, C., 2002, Calculating a Water Poverty Index, World Development, Vol. 30, No. 7, pp.
11951210
Tate, D. M., 1994, Principles of Water Use Efficiency, in H. Garduo and F. Arregun-Corts,
Efficient Water Use, Proceedings of the International Seminar on Efficient Water Use, Mexico,
October
Tsur, Y., and A. Dinar, 1995, Efficiency and Equity Considerations in Pricing and Allocating Irrigation
Water, World Bank Working Paper, No. 1460, May
Varela-Ortega, C., J. M. Sumpsi, A. Garrido, M. Blanco, and E. Iglesias, 1998, Water pricing policies,
public decision making and farmers' response: implications for water policy, Agricultural
Economics, Volume 19, Issues 1-2, September, Pages 193-202
Wang, H., C. Wang, J. Wang, and D. Qin, 2004, Investigations into the Effects of Human Activities on
the Hydrological Cycle in the Yellow River Basin, Water International, Volume 29, Number 4,
December, Pages 499509
WaterStrategyMan Project, 2004, Indicators and Indices for decision making in water resources
management, Newsletter, Issue 4, Jan-Mar
Wichelns, D., 2002a, An economic perspective on the potential gains from improvements in irrigation
water management, Agricultural Water Management, Vol. 52, pp. 233-248
Wichelns, D., 2002b, Economic analysis of water allocation policies regarding Nile River water in
Egypt, Agricultural Water Management, Vol. 52, pp. 155-175
WWAP, 2003, World Water Development Report, World Water Assessment Programme, UNESCO.
Young, R. A., 1996, Measuring Economic Benefits for Water Investments and Policies, World Bank
Technical Paper No. 338, The World Bank, Washington, D.C.
Zekri, S., and A. Dinar, 2003, Welfare consequences of water supply alternatives in rural Tunisia,
Agricultural Economics, Vol. 28, pp. 1-12












Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

269










PROPOSAL FOR THE INTEGRATION OF IRRIGATION EFFICIENCY
AND AGRICULTURAL WATER PRODUCTIVITY



B. Blmling
*
, H. Yang
**
, C. Pahl-Wostl
***

*
Institute of Environmental Systems Research, Barbarastrae 12, 49076 Osnabrck, Germany;
Bettina.Bluemling@usf.uni-osnabrueck.de
**
EAWAG, Dbendorf, Switzerland; Hong.Yang@eawag.ch
***
Institute of Environmental Systems Research, Barbarastrae 12, 49076 Osnabrck, Germany;
Pahl@usf.uni-osnabrueck.de


SUMMARY In this paper, we provide a concept for the integration of the engineering and agronomic
definitions of irrigation efficiency into the concept of Water Productivity. After Water Productivity has
entered the water policy and research arena, there has been some confusion in its use and
delineation from Efficiency. We will therefore first make a clear differentiation between the terms,
and then actually integrate the different kinds of efficiency into what we call Agricultural Water
Productivity. Agricultural Water Productivity then sets the boundaries within which efficiency
indicates the smoothness of the water use process which itself is directed towards high Agricultural
Water Productivity. The latter denotes at which points a process has to be efficient in order to get the
highest overall value out of water. Applying this system perspective of Water Productivity to
agriculture allows going beyond yield as the only output from irrigation water use, but considers
different outputs with differing values. The conceptualisation of Agricultural Water Productivity
provides a sound basis for a harmonized application of irrigation and water use efficiency and water
productivity to decision making.

Key words: Water Productivity; Agricultural Water Productivity; Classical Efficiency; Irrigation
Efficiency.


INTRODUCTION

The importance of increasing and securing food production for a growing world population, while
at the same time limiting agricultural water use, has been extensively discussed among practitioners
and researchers (see for example Rosegrant 1997; IFPRI 2001; FAO 2003; Qadir et al 2003; SIWI-
IWMI 2004). The debate for a long time focussed on (agricultural) water use / irrigation efficiency as
the core concept to indicate the successfulness of water policy that aims at increasing the crop per
drop ratio. In the course of the discourse, many researchers have made an appeal to change the
perspective on and thereby modify the conceptualisation of dealing with water resources in agriculture
(e.g. Carruthers et al 1997; Perry 1999; Gleick 2000; Molden et al 2001b; Postel 2003). Moldens
concept of water productivity was one response to this plea, and was added to the discussion in
1997. With this concept he framed the idea of a group of researchers who thought that efficiency
underlies a conceptual blindness since, what is waste from the efficiency point of view may be
used beneficially elsewhere in the hydrological system. When water is used, not all of it is lost but
parts return to the system and may provide input to other uses.

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

270
There were also other attempts to label the idea which stands behind the water productivity
definition by Molden. Carruthers et al in 1997 proposed a differentiation into real water savings and
paper water savings for the same phenomenon. Seckler (1996) previously advocated for the
terminology of dry and wet water savings. Keller et al (1996; 1998) put forward the term of
effective irrigation efficiency. Taking a look at literature, water productivity is the concept which
some researchers have tried to improve or to apply to their research and therewith seems to be
accepted by the community (see for example Sakthivadivel et al 1999; Renault et al 2000; Droogers &
Kite 2001; Hamdy et al 2003; Peranginangin et al 2004; Dong et al 2004; Bessembinder et al 2004;
Ahmad et al 2004; Kijne et al (without date); Kijne et al. 2004).

At the same time, there seems to be a differing use of the term Water Productivity. For some, it is
just a new name for what was originally referred to in literature as water use efficiency (Zwart &
Bastiaanssen 2004: 116 in their thorough literature review). But the two terms do have different
underlying etymologies and concepts, and the one may not just be re-named into the other
46
. We will
come to this point under the paragraph about classical efficiency in relation to water productivity.

Not only that Water Productivity is used in two different ways, but one may also ask what the one
or other meaning may add? Why introducing a term like Water Productivity when there are already
irrigation efficiency and water use efficiency which are widely applied? In parallel to Water
Productivity entering the water policy and research arena, efficiency remains the indicative term to
other researchers for the evaluation of water use in agriculture (see for example Skaggs and Samani
2005, Rosenzweig et al 2004, Mo et al 2004; Hatfield et al. 2001)
47
.

There hence seems to be a need for clarification. We will in the following make a clear
differentiation between irrigation and water use efficiency on the one hand and Water Productivity on
the other. We then will integrate efficiency into Water Productivity. The system perspective of Water
Productivity as defined by Molden is proven to be very useful for meeting the new challenges in
agricultural water policy, and applying the concept to irrigation water use allows going beyond yield
as the only output from irrigation water use. This Agricultural Water Productivity, as we call it, sets
the boundaries within which efficiency indicates the smoothness of the water use process which itself
is directed towards high Water Productivity. By integrating efficiency into the concept of Water
Productivity as defined by Molden, we add to the latter and at the same time clarify the difference
between Water Productivity and irrigation and water use efficiency.

Whereas Molden focuses on definitions of Water Productivity depending on the scales of
investigation and their interlinkages (Molden et al 2003), we remain on the scale of a field and make
Water Productivity, through the integration of efficiency, a more operational term. Our incorporation of
the system perspective refers to a single water user, whereas Molden focuses on various system
users and their interrelations on different scales. We will start with outlining three kinds of irrigation
efficiency, the so-called Classical Efficiency, Water Use Efficiency and Irrigation Water Use
Efficiency. We then contrast one of it, Classical Efficiency, with the concept of Water Productivity.
Since the upcoming of Water Productivity to some extent can be regarded as a reaction to Classical
Efficiency, we will focus on the comparison of these two concepts. We nevertheless will also discuss
the relation of Water Use Efficiency to Water Productivity.

After a detailed description of the concept of Water Productivity, we come to its modification for
irrigation water use. We will show that the different concepts of efficiency and that of Agricultural
Water Productivity do not compete against each other, but can be used synergistically. This
conception, as we will see, provides a very helpful link to policy making.

46
It is moreover interesting that a definition which previously had been named out of an engineering terminology,
now should be replaced by a term out of an economic context. This may also allow for a discussion about who
takes the lead of discourse in the domain.
47
According to the advocates of the water productivity approach, when speculating on the reasons for the
persistence of what they call classical efficiency (CE), they regard it as a matter of training of current irrigation
practitioners, the orientation of their professional interests and positions around CE, and also the institutional
establishment around CE, as well as the fact that CE serves the interest of other professions and groups as well.
Economists can use low CE as justification for pricing water and water markets; and environmentalists can use it
in their battles against large dams, transbasin diversions and other water-development projects. (Seckler et al
2002: 47f)

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

271


DEFINITION OF TERMS - IRRIGATION EFFICIENCY AND WATER PRODUCTIVITY

Engineers as well as agronomists use the term irrigation efficiency, but denoting two different
meanings. The concept of Water Productivity can be seen as a response and critique to the definition
by irrigation engineers and practitioners, which is often referred to as classical efficiency or CE
(see Wichelns 2002; Seckler et al 2002). Some agronomic definitions seem to be very close to
Water Productivity, but, as we will see, only in their parameters, not in their conceptualisation.



Two Terms of Efficiency

The classical irrigation efficiency, as for example defined by ICID, at each stage of an irrigation
scheme relates the volume of incoming water to the volume coming out of the scheme
48
. For the
whole irrigation scheme, the amount of water stored in the root zone is related to the amount of water
delivered for irrigation. Across different scales, irrigation efficiency is defined for: irrigation
conveyance (farm supply/main system supply), farm irrigation efficiency (field application/farm
supply), field irrigation efficiency
49
(rootzone storage/field application), and overall irrigation efficiency
(rootzone storage/main system supply) (Kassam, Smith 2001: 15). To its users, the term has an
operational function; it is a management ratio which can be taken for management decision support
50
.

For agronomists, there are various definitions of irrigation efficiency. Basically, efficiency relates
the agricultural yield to water consumption. Therefore, whatever may be integrated into the definition
of efficiency as used by agronomists, at the core of it lays (Crop) Water Use Efficiency. It is the ratio
of crop yield to the water consumed to produce the yield, that is, evapotranspiration or, better,
transpiration.

This definition is still widely used (Viets 1962; Hatfield et al 2001; Kang et al 2002
51
; Yuan et al
2003
52
; Zhang et al 2004
53
). The only difference in its use lays in the framing of the nominator,
whether yield may be crop dry matter (either total biomass or aboveground dry matter), the economic
yield (including the crop price), etc. For the denominator, evapotranspiration is often taken, since the
calculation of transpiration is considered difficult.

Water Use Efficiency varies with crop species, available energy from sunlight, atmospheric
pressure, etc. This definition hence expresses the property of a plant at a certain location, that is, the
characteristic of a crop, and therewith is much related to plant breeding. Water Use Efficiency in its
strictest sense does not take into account the role of irrigation. It hence is a genuinely agronomic
term.

48
The International Commission on Irrigation and Drainage uses about the same definition: The water used in
irrigation passes through successive stages of storages (possible), conveyed up to the head of the area,
distributed among the fields and finally applied to each field. During each stage, there is loss of water and the
volume coming out is less than the volume entering. The efficiency at each stage is equal to the ratio: volume
coming out/volume entering (ICID 2000)
49
which often is also referred to as application efficiency
50
We thank M.G. Bos for this remark, who also emphasized that efficiency was never intended to carry out, and
therewith did not include parameters of a water balance. It would be a term which currently would be disused,
ratio may taking the lead.
51
WUE, defined as the ratio between grain yield and total growing season evapotranspiration (Kang et al 2002:
204)
52
Water use efficiency (WUE) is the relation between yield or dry matter produced and the quantity of water
consumed (Yuan et al 2003: 164)
53
Water use efficiency is generally defined in agronomy as the ratio of crop yield (usually economic yield) to
water used to produce the yield (Zhang et al 2004: 113) They set the yield in relation to evapotranspiration.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

272

Some agronomists include opportunity costs into the definition of what they then call economic
efficiency, focussing on financial aspects of irrigation. Economic efficiency of irrigation water use
refers to the economic benefits and costs of agricultural water use in agricultural production. As such,
it includes the cost of water delivery, the opportunity cost of irrigation and drainage activities, and
potential third-party effects or negative (and positive) externalities [...]. Economic efficiency can be
expressed in various forms, for example, as total net benefit, as net benefit per unit of water, or per
unit of crop area and its broader approach compared to physical efficiency [which is here referred to
Classical Efficiency, the authors] allows an analysis of private and social costs and benefits (Cai et al
2001: 6). Since irrigation plays a role only as a cost factor, economic efficiency may go too much
into the direction of economics.

The definition of Irrigation Water Use Efficiency by Howell seems to be more suitable from an
agronomic perspective. It specifies the above Water Use Efficiency in order to take the benefits of
irrigation into account. Irrigation Water Use Efficiency (Howell 2003: 471; Howell 2001: 285) is
calculated by first subtracting the yield which would be achieved without irrigation from the yield which
is produced with the help of irrigation. The same applies for the water fraction in the denominator
where evapotranspiration of precipitation input during the growing season is subtracted from
evapotranspiration of irrigation water input.



This definition of irrigation efficiency incorporates agronomic aspects of plant characteristics as
well as the management of irrigation (e.g. irrigation scheduling or irrigation system).

When further referring to the agronomic definition of irrigation efficiency, we will refer to this
Irrigation Water Use Efficiency. We will understand the genuinely agronomic Water Use Efficiency
(WUE) as integrated in Irrigation Water Use Efficiency (IWUE). Whereas WUE considers the variation
in the yield of different species of a crop, or even among different crops, under the same input of
water, Irrigation Water Use Efficiency looks at the variance of the yield of the same specie / crop
under different applications of water. This integration will be of importance when we further below will
set the two terms in the frame of Agricultural Water Productivity. Irrigation Water Use Efficiency and
Classical Efficiency are relating to Water Productivity in different ways, which we will see in the
following section.


Efficiency in Relation to Water Productivity

As mentioned before, there exists mix-up in the naming of Water Use Efficiency as Water
Productivity, and we will first shortly address this question under the chapter about Water Use
Efficiency in relation to Water Productivity. But since Classical Efficiency is the definition which is
criticized by the advocates of Water Productivity, we will in the following focus on this discussion of
distinguishing Water Productivity and Classical Efficiency. Irrigation Water Use Efficiency actually to
some extent has the same conceptual blindness as CE in that it focuses on the in- and outputs of
agricultural production only. The critique under the chapter about Classical Efficiency in relation to
Water Productivity hence does not only address Classical Efficiency but implicitly is also related to
WUE / IWUE.


(Irrigation) Water Use Efficiency in Relation to Water Productivity

Some of the confusion in the definition of Water Productivity comes from the fact that researchers
use it interchangeably with Water Use Efficiency (see before Zwart & Bastiaanssen 2004). Belder et
al (2004) define Water Productivity as the amount of harvested product per unit water use (Belder et
al 2004: 170), and also Cantero-Martinez (2003) talk about the water productivity of barley, when
actually referring to Water Use Efficiency in their article. Cabangon et al (2004) differentiate between
irrigation water productivity (WPI, kg/m3), and calculated from grain yield divided by the volume of
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

273
irrigation water input during the crop season (Cabangon et al 2004: 197) and water productivity with
respect to the total water input (WPI+R, kg/m3), the denominator was the total water input (I + R)
(ibid). This shows that what previously was, and actually still is, defined as Water Use Efficiency, has
been renamed in Water Productivity.

Though the concept of yield per defined unit of water is very useful, the double naming of Water
Productivity in two different, yet related scientific communities suggests to return to Water Use
Efficiency when it comes to the relation of yield to water consumption, since this term for a long time
has been proved useful, as well as the renaming in Water Productivity actually does not add much to
it. Applying Water Productivity as defined here in contrast will enhance the concept of irrigation
water use and agricultural produce. This is why we encourage returning to Water Use Efficiency and
leaving Water Productivity for taking a new perspective on irrigation water use.


Classical Efficiency In Relation to Water Productivity

Taking a step back and looking at the semantic meaning of the term, productivity focuses on the
result of an action. Being productive implies yielding or furnishing results
54
, while the term at the
same time has a positive connotation of resulting in or providing a large amount or supply of
something
55
. Productivity is a term which is used in an economic context where it means The rate
at which goods or services are produced especially output per unit of labor
56
. In being defined as the
"rate of output per unit"
57
, that is by referring to a unit, productivity incorporates system boundaries in
its definition, as well as it has the notion of getting the most out of a defined limited base, that is the
notion of (profit-) maximization. In summary, productivity is result-orientated and focuses on the
maximization of output based on a certain unit of input.

Efficiency as the quality of being efficient may also be expressed as being productive without
waste
58
or acting or producing effectively with a minimum of waste, expense, or unnecessary
effort
59
. In this sense, the term focuses on the quality of a process, like using water well without
wasting any. Efficiency is expressed in a ratio, that is, as the ratio of the effective or useful output to
the total input in any system
60
. In irrigation, Classical Efficiency stems from an ideology of
technological process optimization. For irrigation engineers, since efficiency is per definition related
to comparing input with output during a given process, the same units for input and output should be
applied (van Dam and Malik 2003: 13). Focussing on making the process within the system
smoother (or: less wasteful), Classical Efficiency therefore does not necessarily have a fixed
reference unit like water productivity which relates the yield to, for instance, a cubic meter of water
(output of a system per unit of input
61)
. At any scale of Classical Efficiency, one may increase the ratio
by means of technological innovation and better irrigation practices, but, and this is the critique by
those using water productivity, without referring to system boundaries (see Molden 1997: 2; Perry

54
Websters Revised Unabridged Dictionary, 1996, 1998 MICRA, Inc.;
http://dictionary.reference.com/search?q=productive, viewed December 2004
55
Cambridge Advanced Learner's Dictionary, Cambridge University Press 2004,
http://dictionary.cambridge.org/define.asp?key=63151&dict=CALD, viewed December 2004 see also for
productive: producing or capable of producing (especially abundantly); productive farmland; his productive
years; a productive collaboration [...] marked by great fruitfulness; fertile farmland WordNet 2.0, 2003
Princeton University, http://dictionary.reference.com/search?q=productive, viewed December 2004
56
The American Heritage Dictionary of the English Language: Fourth Edition, 2000.
http://www.bartleby.com/61/12/P0581200.html, viewed December 2004; see also: The ratio of the quantity and
quality of units produced to the labor per unit of time. Ultralingua.Net.
http://www.ultralingua.net/index.html?service=ee&text=productivity, viewed December 2004
57
Online Etymology Dictionary, November 2001 Douglas Harper,
http://www.etymonline.com/index.php?term=productive, viewed December 2004
58
Merriam-Webster Online Dictionary, viewed December 2004
http://www.m-w.com/cgibin/ dictionary?book=Dictionary&va=efficiency, viewed December 2004
59
The American Heritage Dictionary of the English Language: Fourth Edition. 2000
http://www.bartleby.com/61/98/E0049800.html, viewed December 2004
60
The American Heritage Dictionary of the English Language: Fourth Edition. 2000.
http://www.bartleby.com/61/95/E0049500.html, viewed December 2004
61
see also: productivity, in economics, the output of any aspect of production per unit of input. [] Output can
be measured in output per acre for land, per hour for labor, or as a yearly percent... Encyclopedia.com 2004,
http://www.encyclopedia.com/searchpool.asp?target="productivity", viewed December 2004
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

274
1999: 46f). An increase of efficiency simply implies more total water savings within the system. The
expansion of the system boundaries, the irrigation scheme, that is, the expansion of efficient
irrigation, can only be welcomed. Increasing efficiency in some cases then may even lead to the
overexploitation of the resource. A common example for this is that farmers, when increasing their
application efficiency by utilising irrigation efficient technologies, most likely extend their irrigated land
surface and therewith the overall water use
62
.
The main differences between Water Productivity and Classical Efficiency are provided in Table 1.
In general, Water Productivity takes the hydrological system as a reference unit to set the system
boundaries, whereas Classical Efficiency refers to the irrigation scheme, the infrastructure, as the
system boundaries, within which efficiency shall be increased.

Table 1: Comparison of Water Productivity and Classical Efficiency


Thus, efficiency as being able to function without wasting resources may not be a concept
integrative enough for dealing with the potential conflicts around scarce water resources, since
actually, what is left over may be productively used by other stakeholders. Water Productivity then
again seems to be a suitable terminology in the discussion of how much food or value may be
secured based on limited water resources. It underlies a more integrative view on water resource use.


THE IDEA BEHIND WATER PRODUCTIVITY

Coming from the semantics of efficiency and productivity, we now want to present Moldens
(1997) concept of Water Productivity and dwell on two points which we consider important. On the
one hand, this is the system perspective (see chapter Differentiation of a River Basins In and
Outflows) which allows differentiating diverse products out of irrigation water use. The latter then

62
(Perry 1999: 48) gives an example of a water scarce country in the Middle East [in which] on-farm investments
were made to increase measured efficiency from 40 50 % to 60 70 %, releasing water for further expansion
of the irrigated area. Measurements to date show that the improved technology results in increased crop yields
and increased water consumption a direct confirmation of the many existing studies showing the positive
relationship between yield and evapotranspiration, but not the hoped-for saving in water!
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

275
have to be valuated out of the context of a river basins water scarcity and water use situation in the
chapter Getting the Highest Value Out of Water.

Differentiation of a River Basins In- and Outflows
The Water Productivity terminology was developed in parallel with the emergence of the IWMI
water resources paradigm (see Perry 1999)
63
. The IWMI-paradigm states, among others, that water
is not lost to a larger system though it may be lost to a smaller system. As Seckler et al put it: One of
the cardinal features of water use is that, when water is used, not all of it is used up. Most of the
water remains in the hydrological system, where it is available for reuse or recycling. As water is
recycled through the hydrological system, the efficiency of use increases. Thus, while every part of
the system may be at low levels of water-use efficiency, the system as a whole can be at high levels
of efficiency (Seckler et al 2002: 37f). Molden and de Fraiture give the following example: In some
cases, when 18 irrigation efficiency is improved downstream users (often the environment) can
actually get less water because water gained from farm-level efficiency increases is used upstream.
In southern Sri Lanka, cement lining of canals led to reduced groundwater recharge and consequently
several shallow drinking-water wells dried out [...]. These shallow wells provide better quality drinking
water than fluoride-laden deep wells in the area (Molden and de Fraiture 2004: 9).

This more integrative view on water use as being situated in an overall context of a river basin
certainly is a main improvement by the concept of Water Productivity. The idea actually had been
existing for some time (see for example Palacios-Vlez 1994
64
), but had not been formulated into a
concept. It results in new ways of assessing water use: water accounts are proposed across sectors
within a defined hydrological unit, the receipts and the outgoings of the balance being the inflows
to and the outflows from the water body.

On the sides of the outflows of a water use, according to Molden (Molden 1997: 7), Classical
Efficiency as well as the agronomic definition of irrigation efficiency would only take into consideration
water which is lost to the system through transpiration or evapotranspiration during the growth stages
of plants. Water Productivity, on the other hand, would in addition integrate the water fractions to the
outflows which, though even allocated to irrigation, are not consumed by the crop. The conventional
definitions of irrigation efficiency would consider this part as something which is left when irrigation
efficiency would not be high, that is, as a loss for the system. But according to the Water Productivity
concept, water is only then lost from the system when it e.g. is deteriorated in its quality, flows to
saline sinks or evaporates into the air. The output of a water use, the left over according to the
efficiency definition, thus is much more differentiated: All fractions of water which deplete from the
irrigation site should be integrated; no matter whether they further may render benefits or not.

Looking at the inflows to a system, the critique addresses the same conceptual blindness.
Seckler et al (Seckler et al 2002: 43) criticize that in all the definitions of efficiency up to this point,
precipitation only enters the analysis as effective precipitation (Pe). The difference between total
precipitation (P) and Pe(P-Pe) the amount of ineffective precipitation, as it were is lost; it simply
vanishes from the system, much like the water losses in CE. This is unacceptable in terms of the
water balance of the hydrological system as a whole.

The Water Productivity concept hence integrates different kinds of in- and outflows into the water
use balance for a defined hydrological unit. Molden structures the flows into a water flow diagram
which is reproduced in the illustration below (Fig. 1). We will shortly explain each fraction.

Available water is the amount of water available to a service or use, which is equal to the inflow to
the system less the share of outflow which is committed to other uses. The inflow is split into gross
and net inflow. Gross inflow is defined as the total amount of water flowing into the system
(precipitation, surface and subsurface inflows), whereas net inflow includes changes in the storage of

63
See also (van Dam and Malik 2003: 13) The International Water Management Institute (IWMI) has started a
strong lobby to change the nomenclature from water use efficiency into water productivity, which is now also
followed by other Consultative Group on International Agricultural Research (CGIAR) institutes and the Food and
Agricultural Organization of the United Nations (FAO).
64
Palacios-Vlez (1994) who states in a talk at a seminar held in 1991: In many cases, however, part of that
water can be reused, either in the same system or downstream in another system () when considering actions
to improve water use efficiency, care must be taken that such actions do not have harmful effects in other parts of
the system
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

276
the system, that is: net inflow adds water to the gross inflow if water is removed from the storage, or it
subtracts water from the gross inflow if water is added to the storage. Depleted water comprises that
share of water of the net inflow which becomes unavailable for further use by the system. Molden
distinguishes between process depletion and non-process depletion. The former refers to water which
gets unavailable for further use in the system during its processing for the production of a certain
good (e.g. when it comes to irrigation water use: transpiration during plant growth and water
incorporated into plant tissues). Non-process depletion comprises depletion of water from the system
without fulfilling a specified use (e.g. evaporation of water from the soil and free water surfaces; water
flowing into the sea or into saline groundwater which makes it further unavailable to the system, or
water as much polluted that it is not usable anymore). Non-process depletion is further subdivided
into beneficial or non-beneficial. A beneficial non-process depletion may be that of fruit trees
consuming irrigation water. A non-beneficial non-process depletion may comprise water which is lost
to sinks as well as water rendered unusable because of pollution. But it non-beneficial non-process
depletion can also be caused by weed which is using up water for evapotranspiration. Hereby, the
consideration of how beneficial the depletion may be is defined by the stakeholders in the system. If
stakeholders may find out that a plant which was previously considered a weed has beneficial uses in
their agriculture or as an herb, the water depletion may then be considered as beneficial. The water
still remains defined as non-process depleted since depletion by these plants was not the main
reason why water was diverted from the system.





Fig. 1: Water Flow Balance (after Molden 1997: 5)


The outflow is additionally split into committed and uncommitted outflow. This distinction is
important for the integration of the context into the evaluation of the water productivity of a water use.
Committed outflow comprises the fraction of water which is allocated to further uses in the system.
Uncommitted outflow is further divided into utilizable and non-utilizable. Outflow is utilizable if existing
facilities or the improved management could make further use of it but actually doesnt. Non-utilizable
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

277
uncommitted outflow then is the fraction which leaves the system since the facilities in any case could
not capture it for further use.

Water input to, e.g. an irrigation scheme, thus is considered to have many outflows, and not only
the one to which it is allocated. Several products out of an allocated water input result, and these
products have to be given a value.

Getting the Highest Value Out of Water

Since the term of productivity concentrates on what may result out of a water use, the labelling of
this output is of additional importance. It opens up the discussion about what the value of a produce
actually is to different stakeholders, like for different sectoral products that originate from the same
source of water such as industrial products, bird habitats or tourism.

In Classical Efficiency, since in- and output are of the same entities, no further value would be
given to the output. The result of Classical Efficiency is always that the processing is more or less
efficient. In Irrigation Water Use Efficiency, the value of agricultural produce of course differs for, e.g.
farmers and the government. But the comparison of the value remains within the sector. For Water
Productivity, in its broadest sense, an increase means obtaining more value from each drop of
waterwhether it is used for agriculture, industry or the environment. Improving agricultural Water
Productivity generally refers to increasing crop yield or economic value per unit of water delivered or
depleted(Molden, de Fraiture 2004: 9). But even within one sector, there are different understandings
of what the value may be. Should agricultural water use be set into relation with economic values,
should Water Productivity be a nutritional concept indicating how much nutritional value is produced
out of a certain amount of water (see Renault, Wallender 2000), or should one simply refer to the
yield, without assigning a nutritional or economic value? Seckler et al (2002) decide this question by
making three distinctions of the Water Productivity terminology. Pure physical productivity would be
defined as the quantity of the product divided by the quantity of AWS [available water supply, the
authors], diverted water or depleted water, expressed as kg m
-3
(Seckler et al 2002: 46). Economic
productivity would be the net present value of the product divided by the net present value of the
amount of available water supply, or the water which is diverted or depleted, which can be defined in
terms of its value, or opportunity cost, in the highest alternative use
20
(Seckler et al 2002: 47). And as
a hybrid definition, combined physical and economic productivity is defined in terms of the net
present value (NPV) of the product divided by the amount of water diverted or depleted. Thus, the
quantity of the product is productivity times the amount of AWS or water depleted (Seckler et al 2002:
47). According to Kijne et al, the question of the output of Water Productivity can be dealt with flexibly.
Within one context of water productivity (physical or economic), the choice of the denominator
(depleted or diverted water) may vary with the objectives and domain of interest of the study (Kijne et
al 2003: 5). As we understand it here, the choice of the denominator will be subject to the interests
and values of the respective groups or organisations having a stake in water use, not so much of the
study itself. If the value which different stakeholders denote to certain agricultural products differs,
new issues of conflicts of interest enter the water policy arena.


THE APPLICATION OF WATER PRODUCTIVITY TO IRRIGATION WATER USE

The concept of Water Productivity up until now has been applied to whole river basins (see, for
example, Molden et al 2001a, Peranginangin et al 2004) but not to the detailed analysis of a single
users Water Productivity. But the systematisation of a river basins Water Productivity is also valuable
for the single water use irrigation. As we will see, Agricultural Water Productivity proves to be very
useful to encompass different kinds of benefits out of irrigation water use under the respective natural
resources conditions.

We will in the following make Modifications in the theoretical concept of Water Productivity for its
applicability to irrigation water use. A general adjustment addresses the differentiation of system

20
water used in one place has an opportunity cost in terms of the value of its use in another place within the
system. The concepts of efficiency and productivity need to reflect the values of all the uses and alternative uses
within the system. (Seckler et al 2002: 45)
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

278
flows, while the later integration of the two kinds of efficiency into the concept of Water Productivity
improves it considerably for its application to irrigation water use.


Adding Non-Beneficial Process Depletion To The Water Balance For The Case of Irrigation
Water Use

Whereas Molden splits the fraction non-process depletion further into non-beneficial and
beneficial in his definition of water productivity, the fraction process depletion is not. Process
depletion in the context of irrigation water use for us describes the water rendered unusable to
the system in the process of crop growth. As soon as water is used by other processes or leaves the
field and then is used by other processes, it is rendered to the fraction of non-process uses. From this
perspective, process depletion is the minuend for the calculation of non-process depletion. After the
subtraction of non-process-depletion, water is returned to the fractions of committed or uncommitted
outflow.

Under beneficial process depletion, we then understand the portion of water that is lost to the
system because of a special function it fulfils, that is, in the case of irrigation water use, the water
transpired by the crop. Non-beneficial process depletion would comprise the fraction of water which
e.g. evaporates from the soil surface.

Figure 2 illustrates the application of this water productivity systematization to the farm level. On
this scale, available water defines the Gross Inflow. Inflows at the field level are irrigation
application, precipitation, subsurface contributions, and surface seepage flows. The storage change
in the hydrological system is expressed in soil moisture change in the active root zone. Beneficial
process depletion at the field level is set equal with crop transpiration. Non-beneficial process
depletion comprises the fraction that evaporates from the soil surface, or water rendered unusable
due to the degradation of quality. Non-beneficial non-process depletion for example comprises weed
evapotranspiration, beneficial non-process depletion comprises the evapotranspiration from useful
plants.





Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

279
Fig. 2: Water flow balance for irrigation water use
21


The subdivision of water fractions is not necessarily something new for irrigation water use. The
American Society of Civil Engineers On-Farm Irrigation Committee in 1978 defined irrigation
efficiency as the ratio of the volume of water which is beneficially used to the volume of irrigation
water applied. Beneficial uses would e.g. include crop evapotranspiration, deep percolation for salt
control, crop cooling, frost control, or would take place in combination with pesticide or fertilizer
applications. The denominator also in this case represented the total volume (which means beneficial
as well as non-beneficial uses) of irrigation water. By extending the range of beneficial uses,
efficiency still remains high though the water is not used for transpiration only. Still, the definition does
not see agriculture as embedded in a context of other water users and therewith does not allow for
the valuation of the water fractions.





Problems With The Differentiation of Beneficial / Non-Beneficial Process Depletion

Whereas evaporation and transpiration in this conceptualisation are indicated as the fractions
non-beneficial process depletion and beneficial process depletion, respectively, they generally are
integrated into evapotranspiration, that is, a crop is considered together with certain management
practices under which a certain amount of water evapotranspirates. But we need to make a
differentiation here since it is important to know how much water the crop itself takes to grow
(transpiration), to then examine in how far soil and water conservation practices may change the
evaporation of water from the soil. Different human decision making processes and activities are
linked to the respective depletion fraction, and, as we will see, efficiency and water productivity are
linked to them in different ways.

Crop transpiration basically is a result of plant breeding, and on the farm level a consequence of
crop and species selection. A change in the amount of transpiration more likely requires making
decisions about which crop to grow and which species to select (since transpiration can change
across the species of a crop with their respective production-biomass ratios, the length of the growing
season, etc.). A farmer here basically has to decide whether he wants to grow a crop which he can
use for his livelihood or not, whether he wants to take the risk of producing it (which is also linked to
the selection of certain species of a crop), whether he wants to take the time to manage it, whether it
is easy to sell etc. Since transpiration stands for how much water a crop needs to produce yield, it can
be indicated by Water Use Efficiency as defined before
22
.

Evaporation is always a component related to crop specific growth, tillage and water management
practices Zwart, Bastiaanssen 2004: 116), thus, a reduction in evaporation requires an alteration in
management practices, and here, a farmer most likely has to spend more time on agriculture, or
invest in infrastructure to reduce this actual water loss. The surplus value out of the management then
likely is to play a role if a farmer is a main-income farmer, but maybe not so much if he is a side-
income farmer. Evaporation then is a side-effect / -loss during the course of a water use, that is, non-
beneficial process depletion occurs in connection with an allocation of water to a water use
23
. The
potential amount of water loss from soil evaporation may best be indicated with a ratio of evaporation
to (potential) evapotranspiration (E/ET).


21
terms taken from Molden 1997 and Kijne et al., 2003
22
Actually, also here exists confusion of terms. What we call Water Use Efficiency is also labelled Transpiration
Efficiency by researchers of a biological science / plant breeding community (Byrd 1997; Turner 2004; Condon et
al 2004). To them, Transpiration Efficiency is the the weight of dry matter or biomass produced per unit of water
transpired. Since this paper addresses researchers dealing with irrigation, we will stick to Water Use Efficiency as
yield per water consumed by transpiration.
23
Precipitation of course does not happen for the sake of watering the crop. Still, non-beneficial process depletion
also in the case of precipitation input describes the part of the rain which evaporates from the soil in the proximity
of the crop after a precipitation event. It is lost to further use, though it could have been captured by soil and
water conservation methods.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

280
To support our argument for a separation of evaporation and transpiration despite the common
integration into evapotranspiration, we will give examples of Zwart and Bastiaanssens thorough
literature search by which the variety of evapotranspiration values depending on crop management
practices should get clear
24
. They found out that the variability of the yield for actual crop
evapotranspiration of wheat ranges between 0.6 and 1.7 kg m
-3

25
. The values with the most efficient
water use were found by Jin et al (1999) where the application of manure allowed for higher
production and straw mulching again improved soil water and soil temperature conditions, reducing
evaporation
26
. The variability of cotton lint yield again ranged between 0.14 to 0.33 kg m
-3
. The best
values were found in China and Israel, and for China, they were the result of experiments in which
cotton was planted in furrows and the soil was covered with plastic leaving holes for the infiltration
near the plants (see Jin et al. 1999). By this method, soil evaporation was reduced and the soil water
status of the root zone was improved.

From these results we conclude that a differentiation into evaporation and transpiration makes
sense in two regards: the differentiation shows in how far there is scope for a reduction of the non-
beneficial process-depletion fraction, as well as it provides the opportunity to denote which kind of
action the respective farmers should take in order to achieve this reduction, may it be by changing to
a different crop (in case of a high transpiration), or by investment in irrigation technologies or effort for
crop-, water- or soil-management practices in case of a naturally given high ratio of evaporation.


Integration of Terms In a Concept of Agricultural Water Productivity

As stated above, irrigation efficiency as a single indicator for water use may not respond to
contemporary requirements of harmonized water use planning in water scarce river basins. But the
fact that the two terms of irrigation efficiency may not match current needs does not mean that they
are not important. In fact, Water Productivity can set the boundaries within which efficiency indicates
the smoothness of the process which itself is directed towards high Water Productivity. Water
Productivity then denotes at which points a process has to be efficient in order to get the highest
overall value out of water. Setting this benchmark does not so much orientate at agriculture, but looks
at the context of water use in order to evaluate the respective Water Productivity of agriculture (see
Figure 3).

Setting the benchmark for Classical Efficiency defines the optimal relation of the water outflow
fractions to each other. If we take the above mentioned example of Molden and de Fraiture (2004: 9),
high Water Productivity under these conditions may imply that agriculture facilitates the percolation of
irrigation water to groundwater, so that the fraction process depletion which leaves the system
through drainage would be beneficial and its share in overall irrigation input should accordingly be
high. The same is true for the case of excessive accumulation of salts in the soil. In arid areas,
excess irrigation water is used to leach salt from the root zone. In this case flushing salts with
additional water guarantees future fertility of the soil. From a CE perspective, efficiency would
consequently be called low, but from a Water Productivity point of view, the inefficiency may be
rather valuable. In cases in which water from agriculture would otherwise flow to sinks (like when
agriculture is located close to the coast), high Water Productivity would imply increasing overall
irrigation efficiency. The same holds true if excess irrigation water would leach soluble chemicals
below the root zone, as well as if nitrate is carried below the root zone. Often, in arid and semi-arid
areas, a gradual salinization occurs due to rising water tables where proper drainage has not been
provided and too much water leached underground.

This distribution of outflow water fractions from incoming irrigation water is determined by how the
process of irrigation is managed. To indicate the smoothness of the process of irrigation water use at
the field level, we set evapotranspiration and drainage below the root zone in relation to irrigation

24
As mentioned above, Zwart and Bastiaanssen use water productivity interchangeably with water use
efficiency, which in their research is measured as amount of yield per amount of evapotranspiration. We therefore
describe the results listed in their article with yield per unit evapotranspirated. We take actual
evapotranspiration since their listed examples deal with limited water supply.
25
The value of the FAO study by Doorenbos and Kassam ranged between 0.8 1.0 kg m
-3

26
The CWP for the experiment with straw mulching was 2.67 and 2.41 kg m
-3
for a combination of straw
mulching and manure (Zwart, Bastiaanssen 2004: 118).
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

281
water input. As can be seen from Figure 3, these ratios make up the Evapotranspiration Fraction
and the Drainage Fraction respectively.

A water balance would actually additionally incorporate soil moisture change and runoff. But we
simply neglect soil moisture change which finds only expression in evapotranspiration, and, - if water
content is beyond field capacity drainage from the soil. Runoff would be left as a non-beneficial
water fraction. In Figure 3, we additionally make the simplifying assumption that the irrigation water
input does not leave the field so that the entire outflow is made up of process-depletion fractions and
does not contain non-process depletion fractions.

Additional to the Drainage and Evapotranspiration fraction, we think that the E/ET-ratio is an
important indicator (see Figure 3). The ratio requires special attention since evaporation in every case
is lost to the system without benefit. Especially if a lot of the irrigation water input is allocated to the
Evapotranspiration fraction, the E/ET-ratio becomes important because it shows whether there is
scope for reducing its size through preventing evaporation. The more the E/ET-ratio approaches 1,
the smaller the scope for action will be to reduce evapotranspiration by minimizing evaporation.





Fig. 3. Integration of Water Productivity and Irrigation Efficiency


These above efficiency indicators all relate to Classical Efficiency since they indicate the
processing of different water fractions.

Setting Water Use Efficiency then into the frame of Water Productivity allocates an output to one
water fraction (transpiration) already, which would have to be given a value. Efficiency in this case
denotes whether this value is achieved with a high or low input of water depending on the cultivar.
Irrigation Water Use Efficiency then shows the importance of irrigation to the crop. Table 2 shows the
main function of the above mentioned efficiency indicators.

As Figure 3 as well as our above explanations show, the system perspective on irrigation water
use allows for more products than only agricultural yield. Therewith, also the value of irrigation water
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

282
use does not only relate to agricultural produce. Setting the benchmark for the efficiency of water
processing out of the context in which irrigation water use is embedded thus shall provide value to the
different outflow water fractions, to then allot the size of the fractions within the given water input. We
will come to the point of valuation in our discussion.














Table 2. Efficiency indicators
Efficiency Indicator Function
Evapotranspiration Fraction shows how much of the irrigation water input meets its primary
purpose
Drainage Fraction
indicates how much of the irrigation water input drains below the
root zone and can potentially be used for other purposes than
agricultural production. Whether this water fraction can be named
beneficial or non-beneficial depends on the context.
E/ET-ratio
shows how much of the evapotranspiration is actual loss since it
evaporates without returning a benefit. The ratio also indicates if
there is scope for water saving through soil and water
management: if the ratio is low, evaporation in overall
evapotranspiration is high, so that soil and water conservation may
reduce the size of the evapotranspiration fraction, considerably.
Water Use Efficiency
indicates how much yield a crop returns out of transpiration. It
mainly depends on crop breeding (but also agronomic practices
like fertilizer input, but this is beyond our subject). If WUE is low,
as well as the E/ET-ratio is high, the scope for increasing the
beneficial use of water would be rather limited. In this case, a
change to other crop species or another crop may be advisable.
Irrigation Water Use Efficiency
shows how much of the total water consumption by a crop can be
attributed to irrigation. It shows the dependency on irrigation by the
cultivar.


DISCUSSION AND OUTLOOK

In this conceptualisation of agricultural water productivity as well as irrigation and water use
efficiency, water productivity directs the process of optimization, that is efficiency, towards the highest
overall value of agricultural water use. Process optimization applies at the system flows of the water
source (Classical Efficiency), as well as at the increase of output out of the water transpired
((Irrigation) Water Use Efficiency). Efficiency thus is integrated into the framework of water
productivity and therewith relates to the boundary conditions of water use.

A farmer can impact the above described efficiencies in different ways. In Figure 4, we have linked
the respective efficiency indicators with two main activities: Crop and crop species selection will
have an impact on how much a crop depends on irrigation (Irrigation Water Use Efficiency), as well as
how much yield a farmer may gain out of transpiration (Water Use Efficiency). Soil and water
management again will influence the size of evaporation in evapotranspiration (E/ET-ratio), and how
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

283
much of the irrigation input may be used for evapotranspiration and drainage respectively
(Evapotranspiration Fraction and Drainage Fraction).

Depending on the value of the agricultural produce in relation to the valuation of the drainage
outflow from the field, farmers may take one or the other way in order to increase or decrease the size
of the respective water fractions, and therewith raise their Agricultural Water Productivity. The
valuation of the outflow will depend on the natural as well as the socio-economic conditions in which
agricultural water use is embedded. The valuation of agricultural produce depends on whether it is
sold or not. If it is not sold, the valuation becomes difficult since non-monetary aspects come into
play. If it is sold on the market, it has a value to the consumers of a local (or global) market.

If for example the outflow is valued high since otherwise the soil may turn saline and will make
future agriculture less possible, the Drainage Fraction would have to increase. Soil and water
management would have to be adjusted accordingly. If a farmer in the respective context values his
produce high and does not want to change it, a comparison of the Water Use Efficiency of his crop as
well as the E/ET-ratio can indicate whether agronomic practices may better be changed or the
respective soil and water management, in order to allow for a big Drainage Fraction and achieve a
high Agricultural Water Productivity. If a farmer did not have an interest to stay with a certain crop, he
may also change to another with a lower Irrigation Water Use Efficiency so that more water is set free
for drainage. The dimension of Agricultural Water Productivity then would depend on the price of the
crop and the monetary value of the drainage.



Fig. 4. Field Management Practices and their impact on Agricultural Water Productivity


This example also shows that an interesting point is who may set the value for the products of
irrigation water use in a river basin. Different stakeholders will provide different meanings and hence
values to outcomes, and often, these values may not be easily compared, or monetarised. Even
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

284
within a sector like agriculture, stakeholders will give different meanings and values to agricultural
produce.

Our conceptualisation of Agricultural Water Productivity makes it possible to operationalize this
integrative view on irrigation water use as providing several outputs into an indication of how the
different water use flows may achieve a high overall Agricultural Water Productivity. For this purpose,
we made use of different kinds of efficiency indicators. We think that this conceptualisation provides a
good basis for the integration of irrigation efficiency and water productivity to respond to current
needs of dealing with limited water resources and increasing water demands from different sectors,
as well as it provides links for policy makers to inspire the optimization of the process of water use in
direction of high water productivity.

The integration nevertheless still is in its conceptualisation phase. The most prevailing question is
how to really integrate values provided by the river basin to the outflow water fraction in order to guide
water use flows, and which implications this may have to agriculture.






REFERENCES

Ahmad, M.D.; I. Masih, H. Turral (2004). Diagnostic analysis of spatial and temporal variations in crop
water productivity: A field scale analysis of the rice-wheat cropping system of Punjab, Pakistan.
Journal of Applied Irrigation Science 39.1: 43 63
Belder, P; J.H.J. Spiertz, B.A.M. Bouman, G. Lu, T.P. Tuong (2004). Nitrogen economy and water
productivity of lowland rice under water-saving irrigation. Field Crops Research 93.2-3: 169-185
Bessembinder J.J.E.; P.A. Leffelaar; A.S. Dhindwal; T.C. Ponsioen (2004). Which crop and
which drop, and the scope for improvement of water productivity. In: Agricultural Water
Management (in press)
Byrd, G.T. (1997): Transpiration efficiency, specific leaf weight, and mineral concentration in peanut
and pearl millet. Crop Science 3.1
Cai, Ximing; Claudia Ringler; Mark W. Rosegrant (2001): Does efficient water management matter?
Physical and economic efficiency of water use in the river basin. EPTD Discussion Paper No.
72. Washington, USA: IFPRI
Cabangon, Romeo J.; To Phuc Tuong, Ernesto G. Castillo, Lang Xing Bao, Guoan Lu, Guangho
Wang, Yuanlai Cui, Bas A. M. Bouman, Yuanhua Li, Chongde Chen, Jianzhang Wang (2004):
Effect of irrigation method and N-fertilizer management on rice yield, water productivity and
nutrient-use efficiencies in typical lowland rice conditions in China. Paddy and Water
Environment 2: 195 206
Cantero-Martinez, C.; P. Angas, J. Lampurlans (2003). Growth, yield and water productivity of
barley.(Hordeum vulgare L.) affected by tillage and N fertilization in Mediterranean semiarid,
rainfed.conditions of Spain. Field Crops Research 84: 341 357
Carruthers, Ian; Mark W. Rosegrant; David Seckler (1997). Irrigation and food security in the
21st.century. In: Irrigation and Drainage Systems 11: 83 101
Condon, A.G.; R.A. Richards; G.J. Rebetzke; G.D. Farquhar (2004): Breeding for high water-use
efficiency. Journal of Experimental Botany 55: 2447 2460
Dong, B.; D. Molden; R. Loeve; Y. H. Li; C.D. Chen; J.Z. Wang (2004): Farm level practices and water
productivity in Zhanghe Irrigation System [online]. In: Paddy and Water Environment. Published
on-line 10 November 2004. Heidelberg: Springer Verlag
Droogers, P; G. Kite (2001): Simulation Modeling at Different Scales to Evaluate the Productivity of
Water. In: Physics and Chemistry of the Earth, 26.11-12: 877-880
FAO (2003): Agriculture, Food and Water. Rome: FAO
FAO (2003): Unlocking the Water Potential of Agriculture. Rome: FAO
Gleick, Peter H. (2000): The Changing Water Paradigm A Look at Twenty-first Century Water
Resources Development. In: Water International 25.1: 127 138
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

285
Hamdy, A.; R. Ragab, Elisa Scarascia-Mugnozza (2003): Coping with water scarcity: Water saving
and increasing water productivity. Irrigation and Drainage 52: 3 20
Hatfield, Jerry L.; Thomas J. Sauer; John H. Prueger (2001): Managing Soils to Achieve Greater
Water Use Efficiency: A Review. Agronomy Journal 93: 271 280
Howell (2001): Enhancing Water Use Efficiency in Irrigated Agriculture. Agronomic Journal 93: 281
289
Howell (2003): Irrigation Efficiency. Encyclopedia of Water Science: 467 472
IFPRI (2001): Sustainable Food Security for All by 2020. Proceedings of an International Conference
September 46, Bonn, Germany
Jalota, S.K.; S.S. Prihar (1998): Reducing Soil Water Evaporation with Tillage and Straw Mulching.
Ames: Iowa State University Press
Kang, Shaozhong; Lu Zhang; Yinli Liang; Xiaotao Hu; Huanjie Cai; Binjie Gu (2002): Effects of limited
irrigation on yield and water use efficiency of winter wheat in the Loess Plateau of China.
Agricultural Water Management 55: 203 216
Kassam, Amir; Martin Smith (2001): FAO Methodologies on Crop Water Use and Crop Water
Productivity. Paper No. CWP-M07, submitted to the Expert Meeting on Crop Water
Productivity, Rome, 3 to 5 December 2001
Keller, A. ; J. Keller, D. Seckler (1996): Integrated water resource systems: Theory and policy
implications. Research Report 3. Colombo, Sri Lanka: International Irrigation Management
Institute (IIMI)
Keller, A. ; J. Keller, G. Davids (1998): River basin development phases and implications of closure.
Journal of Applied Irrigation Science 33.2: 145 163
Kijne, Jacob W.; Randolph Barker, David Molden (2003). Water productivity in agriculture: limits and
opportunities for improvement. Wallingford: CAB International
Mo, X.; S. Liu, Z. Lin, Y. Xu, Y. Xiang, T.R. McVicar (2004). Prediction of crop yield, water
consumption and water use efficiency with SVAT-crop growth model using remotely sensed
data on the North China Plain. Ecological Modelling
Molden, David (1997). Accounting for Water Use and Productivity. SWIM Paper 1. Colombo, Sri
Lanka: International Irrigation Management Institute
Molden, David; R. Sakthivadel; Christopher J. Perry; Charlotte de Fraiture; Wim H. Kloezen (1998).
Indicators for Comparing Performance of Irrigated Agricultural Systems.
Molden, David; R. Sakthivadel; Zaigham Habib (2001a). Basin-Level Use and Productivity of Water:
Examples from South Asia. Research Report 49. Colombo, Sri Lanka: International Water
Management Institute
Molden, David; F. Rijsberman; Y. Matsuno; U.A. Amarasinghe (2001b). Increasing Productivity of
Water: A Requirement for Food and Environmental Security. Dialogue Working Paper 1.
Colombo, Sri Lanka: Dialogue Secretariat
Molden, David; Hammond Murray-Rust, R. Sakthivadivel; Ian Makin (2003). A Water Productivity
Framework for Understanding and Action. In: Jacob W. Kijne, Randolph Barker, David Molden:
Water productivity in agriculture: limits and opportunities for improvement. Wallingford: CAB
International
Molden, David; Charlotte de Fraiture (2004). Investing in Water for Food, Ecosystems and
Livelihoods. Comprehensive Assessment of Water Management in Agriculture, Blue Paper
Discussion Draft. Stockholm: IWMI
On-Farm Irrigation Committee of the Irrigation and Drainage Division. (1978). Describing irrigation
efficiency and uniformity. Journal of Irrigation and Drainage Div., ASCE: 104(1):35-42.
Palacios-Vlez, Enrique (1994). Water Use Efficiency in Irrigation Districts. In: Hctor Garduno, Felipe
Arreguin-Corts: Efficient Water Use. International Seminar on Efficient Water Use, Mexico
City, October 21-25, 1991.
Peranginangin, Natalia; Ramaswamy Sakthivadel; Norman R. Scott; Elise Kendy; Tammo S.
Steenhuis (2004). Water accounting for conjunctive groundwater/surface water management:
case of Singkarak-Ombilin River basin, Indonesia. In: Journal of Hydrology 292: 1 22
Perry, C.J. (1999). The IWMI water resources paradigm definitions and implications. In: Agricultural
Water Management 40: 45 50
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

286
Postel, Sandra L (2003). Securing water for people, crops, and ecosystems: New mindset and new
priorities. In: Natural Resources Forum 27.2 (May 2003): 89 98
Qadir, M.; Th.M. Boers; S. Schubert; A. Ghafoor; G. Murtaza (2003). Agricultural water management
in water-starved countries: challenges and opportunities. In: Agricultural Water Management
62: 165 185
Renault, D; W.W. Wallender (2000). Nutritional water productivity and diets. In: Agricultural Water
Management 45: 275 296
Rosegrant, Mark W. (1997). Water Resources in the Twenty-First Century: Challenges and
Implications for Action. Food, Agriculture, and the Environment Discussion Paper 20.
Washington: International Food Policy Research Institute
Rosenzweig, Cynthia; Kenneth M. Strzepek, David C. Major, Ana Iglesias, David N. Yates, Alyssa
McCluskey, Daniel Hillel (2004). Water resources for agriculture in a changing climate:
international case studies. In: Global Environmental Change 14: 345 360.
Ruthenberg, H. (1980). Farming systems in the tropics. New York: Oxford University Press
Sakthivadivel, R.; Charlotte de Fraiture; David D. Molden; Christopher Perry ; Wim Kloezen (1999).
Indicators of Land and Water Productivity in Irrigated Agriculture. In: Water Resources
Development 15.1/2: 161 179
Seckler, D. (1996). The New Era of Water Resources Management: From Dry to Wet Water
Savings. Issues in Agriculture 8, April 1996. Consultative Group on International Agricultural
Research. http://www.cgiar.org/publications/pub_issues.html
Seckler, D.; D. Molden; R. Sakthivadivel (2002). The concept of efficiency in water resource
management and policy. In: Water productivity in agriculture: Limits and opportunities for
improvement, ed. J.W. Kijne. Wallingfort, UK: CABI
SIWI-IWMI (2004). Water More Nutrition per Drop. Stockholm International Water Institute.
Stockholm
Skaggs, R.K.; Z. Samani (2005). Farm size, irrigation practices, and on-farm irrigation efficiency.
Irrigation and Drainage 54: 43 57
Turner, Neil C. (2004). Agronomic options for improving rainfall-use efficiency of crops in dryland
farming systems. Journal of Experimental Botany, Vol. 55, No. 407: 24132425
van Dam, J.C.; R.S. Malik (eds) (2003): Water productivity of irrigated crops in Sirsa district, India -
Integration of remote sensing, crop and soil models and geographical information systems.
Final Report of the WATPRO project
Viets, F.G., Jr. (1962). Fertilizers and the efficient use of water. Advances in Agronomy 14:223-264
Wallace, Jim S.; Peter J. Gregory (2002). Water resources and their use in food production systems.
Aquatic Sciences 64: 363 375
Wallace, J.S.; C.H. Batchelor (1997). Managing water resources for crop production. Philosophical
Transactions of the Royal Society of London: Biological Sciences 352: 93747
Warkentin, B.P. (1994). Protection of groundwater quality through efficient irrigation. In: Hctor
Garduno, Felipe Arreguin-Corts: Efficient Water Use. International Seminar on Efficient Water
Use, Mexico City, October 21-25, 1991.
Wichelns, Dennis (2002): An economic perspective on the potential gains from improvements in
irrigation water management. In: Agricultural Water Management 52: 233 248
Yuan, Baozhong; Soichi Nishiyama, Yaohu Kang (2003). Effects of different irrigation regimes on the
growth and yield of drip-irrigated potato. Agricultural Water Management 63: 153 167
Zhang, Yongqiang; Eloise Kendy, Qiang Yu, Changming Liu, Yanjun Shen, Hongyong Sun (2004).
Effect of soil water deficit on evapotranspiration, crop yield, and water use efficiency in the
North China Plain. Agricultural Water Management 64: 107 122
Zwart, Sander J.; Wim G.M. Bastiaanssen (2004). Review of measured crop water productivity values
for irrigated wheat, rice, cotton and maize. In: Agricultural Water Management 69: 115 133






Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

287



















EFFECTIVENESS OF INDICATORS FOR SUSTAINABLE WATER USE IN
AGRICULTURE


G. zerol
Institute for Environmental System Research (USF) - University of Osnabrck
Barbarastr. 12 - 49069 Osnabrck, Germany
gulozerol@yahoo.com


SUMMARY Water scarcity is a prominent issue about water sustainability in many countries, in
particular in the Mediterranean Region. Given the fact that agriculture has a dominant share among
the water user sectors and water scarcity is a threat to water sustainability, it is expected that
sustainable use of water resources in the agricultural sector is crucial for dealing with the water
scarcity problem. Accordingly, water use in agriculture is the topic under discussion. Being among the
tools for assessing sustainability, indicators can foster actions that can contribute to implementing
sustainability. As with all the indicators, the effectiveness of the indicators for water use in agriculture
is desirable. By effectiveness it is understood that through monitoring and evaluation, indicators can
have an effect on the behaviour of stakeholders, who develop, manage and use water resources, in
order to engage them in collective action for sustainable water use. The results of a field study in
Harran Plain, a region in the southeast Turkey, demonstrate that lack of stakeholder participation
during the development of indicators for water use in agriculture is among the reasons for the resulting
ineffectiveness of the indicators. It is also observed that lack of participation of stakeholders during the
development of indicators has implications about the resulting outcomes, which indicate the lack of an
integrated approach for managing water resources and a resulting collective action problem of
unsustainable water use in the field study region. It is concluded that integrated approaches to water
management, which adopt participation as a core principle and reflect on the economic, social,
institutional and ecological dimensions of sustainability, can bring about not only effective indicators of
sustainable water use, but also long-term achievements.

Key words: sustainable water use, indicators, stakeholder participation, collective action.


INTRODUCTION

Water scarcity is a prominent issue about water sustainability in many countries, in particular in
the Mediterranean Region. Scarcity of water is related to availability and use - or consumption -. On
the one hand, the availability is limited since total water supply available for human use cannot be
increased substantially given the economic, ecological and social constraints on the development of
non-conventional water resources. On the other hand, water consumption continuously increases due
to several reasons, mainly including warmer climate and population rise.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

288

There is a need to ponder on how to deal with water scarcity and ensure water sustainability
through the integration of economic, social, ecological and institutional dimensions of sustainability.
Integrated approaches, which consider these four dimensions, integrate the perspectives of related
disciplines and respect the interests of all stakeholders, are essential. However this is not a simple
issue given the existence of several factors, which put water resource under pressure. These factors
mainly include population rise, increased economic activity, improved living standards, social inequity,
lack of pollution control measures and management of water resources through sectoral approaches
and top-down institutions (Gleick, 1995; Gardner-Outlaw and Engelman, 1997; GWP-TAC, 2000).
Furthermore, the dominant approach to the management of water resources has been supply-
oriented, meaning that increasing the water supplies is the priority issue, whereas the need for
managing the corresponding demand is neglected.

Within this context, integrated water resources management (IWRM) is relevant. IWRM is
defined as follows:
a process, which promotes the co-ordinated development and management of water, land and
related resources, in order to maximize the resultant economic and social welfare in an equitable
manner without compromising the sustainability of vital ecosystems (GWP-TAC, 2000:22).

It is emphasized in the definition of IWRM that success in ecological, economic and social
dimensions, depends on coordination of development and management activities, which constitutes
the institutional dimension. Indeed the causes of the problems about water resources are attributed to
inefficient governance and increased competition for the finite resource (GWP-TAC, 2000: 9). The
inefficiency of governance can be due to disintegrated and uncoordinated approaches for water
management, which rely on sectoral approaches and top-down institutions (Jaspers, 2003). Such
institutions might not create a common understanding about the need for ensuring sustainability and
neglect, or exclude, the knowledge and perspectives of stakeholders, who are not represented or
involved in the existing institutions. The increased competition can be considered as the result of the
allocation of available water resources to users, or user sectors, each of which claim their own right to
use water.

It is important to ensure the participation of stakeholders in efforts towards water sustainability,
whatever the scale and the issue. In the general context of water policy implementation and water
resources management, stakeholder participation is stated among the key requirements for success
(Seppala, 2002; Mostert, 2003). IWRM also takes participation as an approach to be adopted for
water management. It is suggested that all the relevant stakeholders should be a part of the decision
making process, so as to create the balance of top-down and bottom-up institutions and to ensure
that all the stakeholders are aware of the water sustainability as their common issue at stake (GWP-
TAC, 2000; Dungumaro and Madulu, 2003; Jasper, 2003). These premises require that the
stakeholders have the capacity to participate and represent their interests, which calls for the need to
build capacity for participation.

Justifying the emphasis put on participation, there are various benefits expected from
stakeholder participation in the management of resources. Firstly, participation enables an
understanding of the impacts of the individuals actions on the current state of the resource that they
use (Marshall, 1999). These impacts can be costs and benefits of resource use decisions both at the
individual and collective level (Johnson, 1997). Through participation, the stakeholders can realise
these costs and benefits, and having information about the impacts of their actions on the resource
that they use, they are more likely to arrange their actions for the interest of the collective.

Participation provides information about the gap between the institutions -or the rules- and the
actions of the stakeholders (Johnson, 1997). Stakeholders can have a better understanding of the
rules by obtaining information about them and they can become aware of and reduce the gap
between their perception about the rules and the resulting costs and benefits of following or breaking
the rules. They can also agree on and make commitments so as to act in compliance with the rules.
Having information about the actions of other stakeholders would also be useful for the stakeholders
to make sure that the commitments are kept by every stakeholder. In that respect, participation can
be useful for making use of the experiences of stakeholders (Johnson, 1997; Marshall, 1999). Each
participant can monitor his own actions as well as the actions of others, which creates an environment
that binds the individuals to each other and to the outcomes of their actions. Such an environment
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

289
can enhance the perception of the stakeholders that the sanctions are assessed on those who break
the rules (Ostrom et al., 1994).

Participation has additional benefits in terms of enabling the decision making process to be at the
same time a social learning process (Pahl-Wostl, 2002). This means that, through participation, a
common understanding about the situation can be built; exchange of ideas among stakeholders, who
might have diverse interests and knowledge, can be possible and the stakeholders can become more
likely to act for the common goal of sustainability.

The final, and very crucial, benefit that can be expected from participation is the ownership that it
raises about the resource used by the stakeholders. Participation demonstrates to the stakeholders
that each of them has a stake in the state of the resource that they use and their knowledge and
perspectives are important for the betterment of the resource (Johnson, 1997; Marshall, 1999).
Consequently, based on the added-value of the knowledge of stakeholders, the benefits of
stakeholder participation can be summarised as follows:
Local knowledge is often valuable for devising rules, decision procedures and monitoring and
sanctioning mechanisms that take equity as well as efficiency considerations into account, and
therefore are likely to gain broad support from local citizens or resource users (Baland and Platteau,
1996, cited in Marshall, 1999).

In addition to the participation of stakeholders, the assessment of the progress towards or away
from sustainability is a major issue for ensuring sustainability. Developed and used at different levels,
sustainability indicators constitute a major tool for assessing sustainability. The emphasis put on
indicators for the assessment of sustainability is attributed to the expected contribution of indicators to
sustainability, which is possible through the monitoring and evaluation. Monitoring enables the
quantification and communication of the information and in turn improves stakeholders knowledge
about the system. It also establishes the basis for description of the system. Through monitoring,
system characteristics and dynamics of the system can be better understood by the stakeholders. If
the measured values of indicators are also evaluated by taking implications into consideration, then
indicators can become tools to assess the success of the actions and policies for implementing
sustainability and to improve their performance.

With regard to the assessment of water sustainability, the importance of indicators should be
acknowledged as well. In particular regarding the water scarcity problem, the lack of stakeholders
awareness of the water scarcity problem is mentioned as a challenge for water resources and this
ignorance is attributed to the perception that water will always be abundantly available (Abu-Zeid,
1998). Therefore, developing and using indicators that address the problem of water scarcity is
relevant since they enable the assessment of the state of water resources and the impacts of the
actions of stakeholders.

Given the fact that the availability of water resources is limited and human population
continuously rises, a competition among different user sectors is inevitable and agricultural sector
might also be affected from the situation. Therefore agricultural sector can gain from developing and
using indicators for water use, since it has the largest share among all water user sectors, especially
in the Mediterranean region (Araus, 2004). Justifying the argument that indicators are useful for
assessing water use in agriculture, in many countries indicators are developed and used within the
agricultural sector, e.g., indicators for irrigation (Bos, 1997; Molden at al, 1998; Lorite at al., 2004;
Kellett at al., 2005).

Within the context of water use in agriculture, the effectiveness of the indicators is desirable. In
this paper, the effectiveness of an indicator is defined the degree of adoption and utilisation of the
indicator by the stakeholder in taking actions that contribute to sustainable water use in agriculture. It
is argued that effectiveness of indicators requires stakeholder participation during indicator
development. Through a participatory approach the knowledge and perspectives of the stakeholders
can be incorporated when the indicators for sustainable water use in agriculture are developed and
hence the indicators can be perceived as relevant and useful and hence their effectiveness can be
improved. It is acknowledged that the indicator development requires the multiple concerns and,
among others, stakeholder participation is considered as a principle during the development of
indicators (Hardi and Zidan, 1997; Bell and Morse, 2004; McCool and Stankey, 2004).

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

290

MATERIALS AND METHODS

With the aim of investigating the role of stakeholder participation in the effectiveness of indicators
for sustainable water use, an analytical framework was developed and applied using empirical
findings from a field study.


Analytical Framework

The analytical framework was developed through the incorporation of the relationship of the
stakeholder participation during indicator development and effectiveness of indicators to the theory on
collective action for common-pool resources (Ostrom, 1990; Ostrom et al., 1994). After a review of the
theory on collective action and its application to the use of common-pool resources, it was inferred
that sustainable water use requires collective action, since water for agriculture is a common-pool
resource that needs to be used collectively by all relevant stakeholders. Several relevant aspects of
collective action for sustainable water use and the relationships among these aspects were
examined. These aspects include the context for water use in agriculture, the stakeholders of water
use in agriculture, the indicators used by the stakeholders, the institutions -or the rules- for water use
in agriculture, the actions of the stakeholders in terms of following the rules, using water for
agriculture, and developing and using indicators, and the outcomes of these actions.

It is suggested that the actions of the individuals take place in three different levels, namely the
operational, collective-choice and constitutional-choice level (Kiser and Ostrom, 1982 cited in Ostrom
et al., 1994) Corresponding rules are devised and used for each level and for the interactions
between the levels (Ostrom, 1990; Ostrom et al., 1994). The rules at the constitutional-choice level
are about the governance and they indirectly affect the operational level actions since they define the
way that the collective-choice rules are made and changed. Similarly collective-choice rules also
affect the operational level actions indirectly. They are the rules about how the operational rules are
made and by whom the operational rules can be made and changed. Collective-choice rules mainly
include the rules about policy making and management. Finally the operational rules include the rules
that affect the daily actions of the individuals about appropriation, provision, monitoring and
enforcement. Given the scope of the research, the focus of the analysis of the empirical data is on the
collective-choice and operational rules and the corresponding actions of the individuals. The
constitutional-choice rules are discussed to the extent that there are findings about the rules and
actions at this level.

With regard to the position of indicators for sustainable water use in the framework, it is
acknowledged that indicators can contribute to collective action, since they have the potential to
influence the behaviour of the stakeholders in a way to engage in collective action. Such behaviour of
stakeholders is associated with the adoption and utilisation of indicators by the stakeholders and
indicators are considered effective to the extent that they fulfil this potential by affecting the behaviour
of stakeholders, and in turn their actions. Based on the assumption that stakeholder participation
during indicator development is among the reasons for the effectiveness of indicators, stakeholders
and indicators were examined thoroughly in terms of their interactions and their relationship to
collective action for sustainable water use.


Field Study

The field study was carried out through individual interviews with the representatives of the
stakeholders and through the review of written documents, which are extracted from different sources
of information.

Harran Plain was chosen in order to carry out the interviews with the stakeholders from regional
and local level. Harran Plain is a region in anlurfa province, located in the south-eastern part of
Turkey and included in the GAP (Southeastern Anatolia Project). GAP is a regional development
programme, aiming at, among others, increasing agricultural production, employment and income in
the South-Eastern Anatolia Region of Turkey (nver, 1997). The region has a semi-arid climate with
very low precipitation levels. Within the content of the GAP, investments have been made for water
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

291
development and as a result irrigated agriculture is practiced for 10 years in the region. Before that
time, rain-fed agriculture was practiced.

Currently, water for agriculture is abundant due to the water transferred from the Euphrates River
through anlurfa Tunnels and main irrigation canals. However the area of the irrigated land gradually
increases through the completion of canals that carry water to the regions other than Harran Plain.
This means that the same amount of water will be shared by higher number of users, which implies a
potential for water scarcity when the water is used excessively. There are studies, which address the
problem of excessive water use in the region, having also adverse effects on the soil quality and
water table levels (Kendirli et al, 2005). Therefore the field study was fruitful in terms of eliciting the
perception of stakeholders about water scarcity and indicators for water use in agriculture as tools to
prevent the potential impacts of irrigated agriculture.

The specific circumstances, under which the indicators for water use in agriculture are developed
and used, were also investigated for Turkey in general, and for the field study region, in particular. It is
expected that the use of indicators in terms of monitoring and evaluation, can be made on the basis of
different water user sectors. However the development of indicators cover several related processes
about water, i.e., planning, investment, management, etc., which are usually done for all sectors
through several overlapping laws and regulations. Making the system more complicated, the
responsibility and authority regarding different uses of water are distributed among different
stakeholder organisations. Therefore it is meaningful to have an understanding of the national setting
for water use without a narrow focus on agricultural water use. Stakeholders of water use in
agriculture include the governmental organisations, environmental non-governmental organisations
and water user organisations. The activities of each stakeholder and the distribution of responsibility
and authority among the stakeholders at different levels and were explored.

The two major stakeholders at regional level are State Hydraulics Office (DSI) and Water User
Organisations (WUOs). DSI is the governmental organisation responsible for the planning,
development and administration of water resources at national level. It has a general directorate and
26 regional directorates (DSI-RD), which function at regional level. Two officers of DSI-RD, which is
responsible for Harran Plain, were interviewed. WUOs are the legal entities formed by the
representative of farmers. They are responsible for the appropriation, i.e., distribution of water to the
farmers and provision, i.e., operation, maintenance and repair of the irrigation facilities. Since the
WUOs are active in all Turkey, with a total number of around 780, only the situation in Harran Plain is
investigated and five out of the eleven WUOs were contacted in the region. In addition to the
scheduled interviews, conversations were made with farmers and the relevant observations and
impressions from these conversations are incorporated.

Given the fact that the aim of interviews was to elicit the perception and the knowledge of the
stakeholders, it was found appropriate to conduct semi-structured interviews. Most of the questions
were prepared as open-ended questions so as to enable the interviewees to talk freely on the issue
addressed by the questions. While most of the questions are the same for all the interviewees, one or
two questions were included or excluded for several interviewees, according to the major tasks and
specialisation fields of the interviewees.


Analysis of Findings

The findings from the field study were utilised as empirical data for the application of the
framework. For this purpose, the context, rules, actions and outcomes, which are related to water use
in agriculture, were identified and analysed. Firstly, the examination of the context for water use in
agriculture was made, which includes description of water for agriculture as a common-pool resource,
the producers, providers and the appropriators of water for agriculture, and the social and cultural
conditions of the region.

Secondly, the rules about water use were identified and discussed. Utilising the empirical data,
the rules at constitutional-choice, collective-choice and operational levels were identified. Given
limited empirical data, a thorough analysis of the constitutional-choice rules could not be made and
these rules were mentioned in order to give a general idea about the background for the collective-
choice and operational rules.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

292

Thirdly, the outcomes of water use in agriculture were identified at the system level and the
actions of the stakeholders, which might lead to these outcomes, were investigated. The analysis of
outcomes is focused on the collective-choice and operational levels with the purpose of tracing how
the interactions of the context, rules and actions might lead to the outcomes. From the analysis of
context and the rules, several implications have been made. The resulting set of implications is
classified under two groups; the implications of the context and the constitutional-choice rules and the
implications of the collective-choice and the operational rules, most of which are related to the
monitoring and enforcement of rules for allocation of water.


RESULTS

The interactions among the context and the rules were investigated with the purpose of revealing
how these interactions might lead to the actions and outcomes of water use in agriculture in the field
study region. For this purpose, firstly, the outcomes observed at the regional level were identified.
According to all interviewees from DSI-RD and the WUOs, there is excessive water use in the
region, which results in the following outcomes:
- rise in water table
- degradation of soil quality; mainly in the form of salinisation and becoming barren
- high quantity of water that goes to discharge

The above outcomes are negative externalities of water withdrawals by water users. These
externalities are experienced by all the water users in the region and they imply that the individuals do
not act for the achievement of common welfare or reflect on the ecological impacts of their individual
interest. Hence it can be stated that there is a collective action problem in the form of unsustainable
water use. There is no evidence that DSI or the WUOs experience problems about the irrigation
facilities. This situation might mean that the focus of the WUOs and DSI are more on the irrigation
facilities. Therefore it can be inferred that the actions of the stakeholders do not lead to provision
problems, but rather appropriation problems in the form of the outcomes above.

It is argued that excessive water use can be a result of the context in the region, as well as the
lack of monitoring and enforcement of the rules. The factors, which might lead to above outcomes,
are identified and explained below.


Context and Constitutional-choice Rules

Water as a new resource

Water for irrigation is available since ten years. Before that time, there was no irrigated
agriculture and almost no water at all. Hence, irrigation water is new for the farmers. Therefore it can
be expected that the farmers are not experienced in using water for agriculture. Another expectation
is the lack of awareness about the limited water availability and the need for using water efficiently.
This might cause a distorted perception. Since water continuously flows in the canals, the farmers
might have a perception that water is not abundant. Similarly the farmers, inexperienced in irrigation,
might think that using more water brings about higher yield. Unlike the farmers, all the WUOs state
that the water resources are not abundant and they try to enforce the farmers to use water efficiently.
Hence, it can be expected that the farmers awareness of water availability and use might be a critical
factor on their actions and resulting outcomes.

Governmental support for cotton production

The national agricultural policy, which supports cotton production, might also have an impact on
the actions of the farmers. The governmental support on cotton production might induce the farmers
to cultivate cotton with the expectation that they can sell what they raise and guarantee their income.
Currently, the share of cotton is about 80-90% of the whole cultivated land in the region. This
proportion had been foreseen as 25% in the GAP Master Plan (nver, 1997). The impact of this high
proportion is a water demand much higher than expected, since cotton consumes more water than
many other crops. The support of cotton production has an impact on the type of the crop cultivated
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

293
by the farmers. The individual interests of farmers about increasing their income and the outweighing
short-term benefits can be the factors that lead to this choice of crop, which in turn affects
substantially the quantity of water used by the farmers.

Size and composition of the WUOs

There is no rule about the limit to the number of farmers to be included in a WUO. This situation
results in the fact that the WUOs are not homogeneous in terms of their size and the area of land that
they manage. Having a high number of members can make it difficult to build continuous relationships
and trust among the farmers. Furthermore, the existence of leasehold farmers increases the diversity
of the members of the WUOs. The resulting heterogeneous and unstable composition of the WUO
can lead to short-term relationships among the farmers and lack of a shared interest for a common
future with other farmers in the WUO. It is also mentioned that the leasehold farmers might give less
consideration to the externalities of water use in agriculture and take only the short-term benefits into
account. Combined with the size of the WUOs, the existence of leasehold farmers can also lead to
problems about monitoring the actions of farmers in terms of water use and enforcing the rules for
allocation and provision. The farmers who withdraw water from the same canal even might not know
each other.

Organisation of the Water User Organisations

The WUOs find their institutional capacity low in terms of the outputs of their activities. They
attribute it again to the lack of a WUO law, as well as the lack of awareness among farmers, lack of
coordination among higher level organisations, limited financial and administrative resource and lack
of qualified staff. Furthermore, within the organisational structure of the WUOs, the secretary general
is the only person who has to deal with the administrative, operational and legal issues. This situation
makes secretary general the central person in the WUOs and in turn creates dependency on the side
of other personnel as well as the member farmers of the WUO.


Monitoring and Enforcement of Collective-choice and Operational Rules

Monitoring and Enforcement by DSI

Monitoring of water use for agriculture at the regional level is carried by DSI-RD through
monitoring the quantity of water allocated to each WUO and the water use per hectare for the region.
However most of the WUOs indicate that there is no comparison of among the WUOs with respect to
each other or to the regional level.

Furthermore, each year, WUOs submit three documents to DSI, namely, monitoring and
evaluation report, inspection report and water distribution plan, all of which have the same structure
for all WUOs, meaning that they are not tailored according to specific conditions of the WUOs.
Inspection report and water distribution plan are used for declaring respectively the provision and
distribution activities. Monitoring and evaluation report is used in order to monitor and evaluate all
activities of the WUOs. Since none of the abovementioned outcomes are related to provision
problems, the use of inspection report is not analysed, whereas the issues about monitoring and
evaluation report and water distribution plan are discussed in detail.

Monitoring and evaluation report is a very comprehensive document conveying many types of
data to DSI. It can be expected that the data in this document are evaluated by DSI and the
performance of the WUOs are assessed in terms of their past vs. present outputs or against other
WUOs. With regards to the uses of the monitoring and evaluation report, most of the WUOs indicate
that they could not make use of it in terms of a contribution to their success. All the WUOs indicate
that they receive no regular or formal feedback from DSI about the monitoring and evaluation report.
This situation can imply that the monitoring and evaluation reports, which the WUOs send in the
previous years, are not regularly evaluated by DSI. It is mentioned by all the interviewees from WUOs
and DSI that DSI informally evaluates the performance of the WUOs, e.g., by talking about the
outputs of the monitoring and evaluation reports during their meetings. However there is no evidence
about a regular performance evaluation of the WUOs.

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

294
Absence of feedback from DSI implies a lack of common understanding about the uses of the
monitoring and evaluation report as well as inspection report and water distribution plan. It is also a
communication problem between the DSI and WUOs due to the one-way flow of information from
WUOs to DSI. Some WUOs mention that they expect DSI to evaluate their performance. It can be
expected that WUOs require a performance evaluation from DSI, since they share the water from the
same canal and each of them would like the others to perform well. However the WUOs complain
about their legal and organisational incompetence, which constitutes a barrier for them to take such
actions.

With regards to the water distribution plan, it is used for planning the allocation of water to the
WUOs. It is mentioned that the demanded value on the water distribution plan are negotiated and it
can be decreased by DSI. However, during the irrigation season, water is continuously available and
the weekly demands of WUOs are met. Hence, even if DSI monitors the quantity of water allocated to
each WUO, the WUOs do not make their allocation rules according to the quantity of water allocated
to them.

Monitoring and Enforcement by the WUOs

As observed from the operational rules, most of the decisions made by the WUOs are based on
two variables, namely, the area of the land irrigated and the type of the crop cultivated on the land. A
water related variable is not used by the WUOs when they make decisions, nor they monitor the
quantity of water used. Distribution of water to farmers and contribution of farmers to provision are
made based on area of the land and the crop type.

WUOs plan the distribution of water based on the area of the land to be irrigated and the type of
the crop cultivated on the land. The amount of water to be distributed increases with increasing area
of land and is also dependent on the type of the crop. According to the distribution plan, the water is
allocated to each farmer for a predefined number of days. Thus the WUOs do not monitor the quantity
of water withdrawn by each farmer; instead they rely on the duration of irrigation. Only one of the
interviewed WUOs is an exception to this case. This WUO was founded in 2004. It is planned by the
WUO that water is withdrawn through valves instead of siphons and irrigation is made with piped
irrigation network and a closed system instead of open canals. It is stated that existence of valves
also enables the metering of the quantity of water taken by each farmer. Currently, some of the fields
have been equipped with the valves and it is mentioned that they will monitor the water use through
the valves. It is also mentioned that this method decreases the amount of water that goes to
discharge.

With regards to the method for the determination of the irrigation fee to be paid by each farmer,
most WUOs use the area of the irrigated land and the type of the crop cultivated as the two criteria.
This implies that the irrigation fee is independent of the quantity of water actually used by the farmers.
Indeed, most of the WUOs indicate that there is no metering of the water withdrawn by each farmer;
instead the farmers take the water from the tertiary canals through the siphons, which withdraw the
water from the canal. It can be expected that collecting the irrigation fee based on quantity is not
adopted by the WUOs since its justification would more difficult, given that they do not monitor it.

Finally, all the WUOs indicate that they are responsible for monitoring the level of water table in
their region, but most of the do not make it. It is also mentioned by some of the WUOs that currently
DSI monitors the level of water tables, but there is an impression that WUOs are not successful in
terms of communicating these values with the farmers and convincing the farmers not to use
excessive water.

When the monitoring and enforcement actions of WUOs are considered as a whole, three
factors, which might have an effect on the actions of the farmers, in terms of monitoring their own
water use, are identified as follows:
- determining the irrigation fee and water distribution according to crop type and irrigated area
- using the irrigation duration as the unit of water distribution
- not monitoring the quantity allocated to each farmer

Monitoring by the Farmers

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

295
Following from the rules, which are based on the above three factors, it can be inferred that the
farmers do not have an incentive to monitor how much water they use or what the impacts of their
water use are. In that respect, the interactions between three variables, namely the type of the crop,
quantity of water and duration of irrigation, and two main issues that result from these interactions are
relevant about the actions of the farmers and the abovementioned outcomes.

Firstly, despite the fact that WUOs assume 24 hours of irrigation per day, it is mentioned that
some of the farmers do not apply night-irrigation; they irrigate the land during the day and let the
water go to the discharge during evening and night. Hence, insufficiency of night irrigation is the main
reason of the high quantity of water that goes to discharge. It is also indicated that quantity of
discharged water should be as low as possible, and zero in the ideal, however, it is not possible with
open canals and siphons. It can be expected that because the distribution of water is on the basis of
time, duration of irrigation is critical for the farmers and they should adopt night irrigation. Indeed, it is
mentioned that night irrigation has several advantages. Firstly, if it is applied as complementary to
day-light irrigation, the duration that the water is withdrawn from the canal can be extended to 24
hours/day and irrigation can be finished earlier. Furthermore it is likely that the water allocated to the
WUO is used more efficiently, since the amount of water that goes to discharge reduces significantly.
Secondly, if it is applied either as a substitute (or complementary) to day-light irrigation, the water loss
due to evaporation is decreased since the crop consumes the water, not (only) during the warmer
hours of the day, but cooler hours of the evening and night. Thirdly, the consumption of water by most
of the crops is easier when the crop is irrigated during evening or night. But still the WUOs cannot
effectively implement night irrigation, implying that the farmers might not have enough motivation to
apply it.

Secondly, the quantity of water that they use is not a priority of farmers when they irrigate the
land. The WUOs takes into account the type of the crop, since each crop needs different amount of
water. However, the farmers do not monitor the quantity of water that they use; instead the relation
between the type of the crop and the amount of water that the crop needs is reflected by a heuristic
adopted by the farmers. It is mentioned that the farmers base their actions about water use according
to the total number of times that they irrigate the crop. It is stated by a farmer that in the previous
years, the farmers used to irrigate the cotton ten times; now they irrigate six or seven times.

Even if they do not monitor their own water use, it could be expected that the farmers could have
an interest in the water use at WUO level and follow the outcomes of the monitoring activities carried
out by the WUO. However it is mentioned that the farmers do not have a concern about these
monitoring activities. Consequently, it is observed from the actions of farmers that using the irrigation
duration as the unit of water distribution and planning it according to crop type and irrigated areas are
not reflected on the side of the farmers actions. This situation has several implications about the
WUOs and farmers. These issues imply that the rules used by the WUOs do not relate the quantity of
water, needed by the crops, to the duration of irrigation. Therefore the farmers might not adopt the
night-irrigation and monitor the quantity of water at the farm or local level. Several secretary generals
explain the reasons for the lack of monitoring by the farmers and the farmers awareness of the
impact of their actions on the abovementioned outcomes using a common proverb. This implies that,
since most of the outcomes are observed with a time lag after their actions, the farmers might not be
considering what the outcomes will be. It can be expected that lack of knowledge, experience and
awareness of water for agriculture, and the outweighing short-term benefits, such as earning income,
can lead to these actions.


DISCUSSION

Remember that the condition for the indicators to be effective is that they are adopted and
utilised by the stakeholders in taking actions that contribute to sustainable water use in agriculture.
Hence the actions of DSI-RD, WUOs and farmers, in terms of monitoring and evaluating the
indicators, are of utmost importance in evaluating the effectiveness of the indicators. Analyses of
findings address the following actions of DSI-RD, WUOs and farmers:
- Neither DSI-RD nor WUOs monitor the quantity of water withdrawals by the farmers.
- Farmers monitor neither the quantity of water that they use nor the impacts of irrigation on soil
quality and water table level.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

296
- WUOs do not monitor the realisation of their water distribution plan in terms of the allocation
of the quantity of water according to it.
- WUOs monitor many indicators, use them to prepare the monitoring and evaluation report
and submit it to DSI annually, but they receive no regular feedback about it and most of the WUOs
make little use of the monitoring and evaluation report in their success.
- WUOs do not have an ownership of indicators that they monitor for the monitoring and
evaluation report; several WUOs perceive monitoring and evaluation report as irrelevant to their
activities and having no contribution to their performance.
- DSI monitors indicators about water use by WUOs, but the performance of the WUOs is not
evaluated explicitly and little communication occurs between DSI and WUOs about the outcome of
the evaluation.

Above findings suggest that the indicators are either not monitored at all, or monitored with a lack
of evaluation of their value, implying the lack of utilisation of indicators by the stakeholders.
Furthermore there is an impression that the adoption of the indicators by the stakeholders is very low,
which is reflected by the lack of ownership about indicators and motivation to monitor and evaluate
the indicators. All these findings are considered as strong evidences for the existence of ineffective
indicators and it is concluded that the indicators for water use in agriculture are ineffective in the field
study region.

In order to verify whether there was lack of stakeholder participation during indicator
development, the functioning of WUOs and the method, with which the indicators for water use in
agriculture had been developed, are discussed. The role of the WUOs in creating a participatory
context is considered important, since the WUOs act like a bridge between DSI and the farmers. It is
acknowledged that, by carrying out the operation and maintenance of the irrigation facilities and by
distributing the water on their own, WUOs constitute a useful collective-choice entity for the
participation of farmers in the management of the irrigation system. However, participation of WUOs
remains at operational level. There is no evidence from the empirical data that WUOs have
participated during the design of many constitutional-choice rules that in turn affect their decisions at
both collective-choice and operational level. For instance, with regards to the monitoring and
evaluation of the activities of the WUOs, there are several collective-choice actions, in which the
WUOs could be involved. Among others, contributing to the design of monitoring and evaluation
report, tailoring it according to their context, communicating its content to their farmers and in turn
adapting it by considering the changes in conditions would be forms of participation at collective-
choice level. However there is little evidence that WUOs have been involved in such actions.

The methods that had been used for the development of indicators monitored by DSI and WUOs
have direct implications about the participation of stakeholders during indicator development. At this
point, a differentiation can be made with the formal and informal indicators, which imply the indicators
used for monitoring and enforcement of formal and informal rules. The formal indicators mainly
include those monitored for the monitoring and evaluation report. These indicators had been imposed
by DSI with a unique format to be used by all the WUOs in Turkey. As explained above, there is no
evidence that the WUOs or farmers have contributed to the design or preparation of the monitoring
and evaluation report or adapted it to their specific conditions.

With regards to the informal indicators, WUOs have the initiative to choose the method of
distribution of water to the farmers. Accordingly the WUOs develop and use indicators, in particular
for monitoring water distribution. However, the informal indicators that are used by the WUOs do not
match with the formal indicators of DSI-RD. On the one hand, DSI-RD takes weekly water demands
from the WUOs and monitors the quantity of water allocated to each WUO. On the other hand, the
WUOs distribute the water to farmers on the basis of irrigation durations and number of irrigation
turns, which are different indicators to monitor water distribution. The choice of WUOs to use
durations and number of turns as indicators is their own decision, but not a result of the participation
of DSI-RD or the farmers. Furthermore, the secretary general is the only person responsible for the
fulfilment of administrative and technical tasks of the WUOs. This situation might imply lack of sharing
of responsibility with other members of the WUO and lack of a participatory approach for the
establishment, monitoring and enforcement of collective-choice rules devised and used by the WUOs.

Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

297
Consequently, the development of neither formal nor informal indicators was made using a
participatory approach and lack of a common understanding, about the monitoring and enforcement
of collective-choice and operational rules, is observed, implications of which are experienced in the
form of mismatching formal and informal indicators. Thus, it can be concluded for the field study
region that there is a lack of stakeholder participation during the development of indicators for water
use in agriculture.

Having concluded that indicators for water use in agriculture are ineffective and there was lack of
stakeholder participation during indicator development, it can now be discussed whether the latter is
among the reasons of the former. For investigating this relationship, several factors are discussed
based on the interactions among the context and the rules at different levels. Firstly, the
constitutional-choice level decisions can affect the behaviour of the farmers. These decisions mainly
include the governmental support for cotton production as part of the agricultural policy and the
supply-oriented approach of DSI. Governmental support for cotton production is likely to effect the
decision of the farmers in terms of their crop choices. Since the cultivation of a crop, which can be
sold, is rational for the farmers, assuring their income, i.e., the short-term benefits of water use can
become dominant for the farmers. Therefore short-term benefits of using water without monitoring its
quantity or impacts might outweigh the concern for common interests, which require monitoring and
evaluation. Furthermore the supply-oriented approach of DSI, through the increase of water supply for
irrigation by constructing necessary infrastructure and diversion of water from the rivers is relevant.
This approach can result in a stimulus, on the side of the WUOs and farmers, that water use in
agriculture and its impacts are not a priority issue, since water is continuously available and irrigation
is possible throughout the year, no matter what the indicator values are. Given also the lack of
awareness about irrigation methods and potential impacts of irrigated agriculture, it becomes more
likely that indicators are not monitored or evaluated by the WUOs and the farmers.

Secondly, the socio-economic conditions of the region require that some farmers cultivate the
land through leasing it temporarily. It is observed from the analysis of the findings that the existence of
leasehold farmers both increases the diversity of the WUOs and makes it more likely that indicators
are not monitored. The leasehold farmers might have different preferences than the owners, in
particular about the sustainability of soil and water resources the region. Since the priority of a
leasehold farmer can be to earn as much as income, without caring for the water use levels and
impacts of irrigation, it can be possible that the leasehold farmers do not monitor and evaluate the
indicators. Hence, the ineffectiveness of indicators can be attributed to the ownership patterns of the
farms, too.

Thirdly, the perceptions of the stakeholders about water for agriculture are important. The
context for water use in agriculture in the region shows that water for agriculture is a new resource for
the farmers. This situation makes it more likely that the farmers are inexperienced about irrigation
methods and ignorant of the availability of water for agriculture and the impacts of irrigation. Despite
the fact that both DSI-RD and the WUOs are aware of the limited availability of water, all the
problematic outcomes at regional level are attributed to excessive water use. However, the transfer of
irrigation water from the Euphrates River conceals the potential water scarcity problem, which would
be experienced in the absence of the irrigation canals. This supply-oriented and top-down approach
does not enforce DSI-RD and WUOs to allocate water without causing excessive use. In the absence
of such enforcement, the awareness of DSI and WUOs are not put in practice and it is not
communicated to the farmers, either. Under these conditions, the farmers become critical actors for
water use; they are the people closest to water and their perceptions and actions have direct effect on
the quantity of water used. Thus, existence of an action like excessive water use might indicate that
the farmers do not perceive water as a scarce resource. Such distorted perception of the farmers
might imply lack of farmers awareness about limited availability of water. Lack of farmers awareness
is attributed to the fact that farmers the communication of farmers with DSI-RD and WUOs about
limited availability of water is not sufficient and the farmers were not trained about irrigation methods
and water use, i.e., they did not participate in the design of the rules, which include the monitoring
and evaluation of the currently used indicators. This means that lack participation of farmers can also
be among the reasons of ineffective indicators.

Fourthly, the limited capacity of the WUOs to train the farmers for monitoring, to communicate
the results of the monitoring activities and to involve the farmers in the management of the WUOs can
also lead to a lack of motivation on the side of the farmers to monitor and evaluate their water use
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

298
behaviour. This reasoning applies to the relationship of DSI and the WUOs, too. There is little
evidence that the WUOs are motivated to monitor and evaluate the indicators. This problem is
associated with the lack of communication between DSI and the WUOs, especially in terms of the
design of the allocation and provision rules, the evaluation of the performance of the WUOs and the
feedbacks of the monitoring activities. Additionally, as mentioned above, the mismatch between the
indicators monitored by DSI and the WUOs is a result of the lack of participation of the other during
the development of own indicators. These communication and cooperation problems bring about lack
of adoption and utilisation of the indicators. Thus, the lack of participation by DSI, WUOs and farmers
to develop the indicators is a reason for ineffective indicators.

The third and fourth factors imply that lack of stakeholder participation during the development of
indicators is among the reasons for ineffective indicators. It is also acknowledged that the
constitutional choice level rules, e.g., governmental support on cotton production, and the context
related factors, e.g., the socio-economic conditions of the region and existence of leasehold farmers,
also have an effect on the ineffectiveness of indicators. However it is expected that their impacts on
the outcomes at regional level could be mitigated if the farmers would have been involved when the
indicators had been developed. Through participation it could be more likely that the farmers adopt
and utilise the indicators; since that would have had the opportunity to be aware of the scarcity of
water resources, the impacts of their actions on the sustainability of soil and water resources, the
trade-off between individual and common interest and the benefits of monitoring and evaluating the
indicators.

It is concluded from the above discussion that indicators for water use in agriculture are
ineffective, there is lack of stakeholder participation during the development of indicators and lack of
stakeholder participation during the development of indicators for water use in agriculture is among
the reasons for having ineffective indicators in the end.

Consequently, the overall situation in the region addresses a collective action problem due to
excessive use of water and its adverse impacts on the soil quality and water table level. The previous
barrier on agricultural production, i.e., the arid climate that causes water scarcity, has been overcome
through bringing water from Euphrates River. However, excessive use of water indicates that the
management of water resources, which is practiced since the construction of irrigation system ten
years ago, has not adopted an integrated approach and the outcomes indicate a collective action
problem of unsustainable water use. It is expected that an integrated approach would not only
improve social and economic conditions by creating employment and income through increased
agricultural production, but also build institutions, which enable participation of WUOs and farmers,
and in turn ensure that the adverse ecological impacts of irrigated agriculture are prevented, or at
least minimised.


CONCLUSIONS

From the review of literature it was observed that indicators are developed and used as tools to
support the efforts towards implementing sustainability and the participation of stakeholders during
indicator development is considered among the major factors that bring about effective indicators. In
this paper, an attempt was made in order to explore the relationship between the lack of stakeholder
participation during the development of indicators for water use in agriculture and the ineffectiveness
of the indicators.

The empirical findings demonstrate that indicators for sustainable water use in agriculture are
essential tools for dealing with collective action problems. When they are not monitored and
evaluated by the stakeholders, the outcomes at system level are more likely to be the symptoms of a
collective action problem, which is experienced as excessive water use in the field study region. Lack
of stakeholder participation during the development of indicators proves to be among the reasons for
the ineffectiveness of the indicators, since the stakeholders do not have an ownership about the
indicators and lack a common understanding about the benefits of monitoring and evaluating them.
Furthermore, the affects of the context and constitutional-choice rules are also to be acknowledged.
In that respect, the socio-economic conditions in the region and decisions made regarding the
national water, energy and agricultural policies affect the perceptions of the stakeholders to create a
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

299
priority for short-term benefits and can result in a lack of motivation to monitor and evaluate the
indicators.

The results presented in this paper demonstrate a static view. However the national context is
rapidly changing due to several reasons. For instance, new institutional arrangements are foreseen at
the constitutional-choice level, e.g., a water law and a WUO law (DPT, 2001). The priorities of all
governmental organisations may also change according to the course of events in the EU accession
process.

The limited availability of empirical evidence is a reason for making conclusions only for Harran
Plain, in which the interviews were carried out with the WUOs. At the national level, a more thorough
analysis could be made if there were more empirical data available. For instance, data could be
collected for more than one region and the results could be compared. If more data could be collected
at the national level, the effects of the constitutional-choice rules could also be discussed in more
detail. At this level, two issues were analysed, namely the irrigation projects in the region, which made
water available for agriculture, and the effect of governmental support on farmers crop choice and in
turn water use decisions. However, there is not enough evidence from the field study so as to discuss
all the rules and actions at this level. This situation is due to relevance of many other stakeholders,
who could not be included in this study, and national policies, which affect the constitutional-choice
rules and actions about energy, agricultural production, and water planning and development. Given
the scope and aims of this paper, it is hoped that the presented findings and discussions constitute a
comprehensive picture of the situation for water use in agriculture.


REFERENCES

Abu-Zeid, M.A., 1998, Water and sustainable development: the vision for world water, life and the
environment, Water Policy, 1, 9-19.
Araus, L. J., 2004, The problems of sustainable water use in the Mediterranean and research
requirements for agriculture, Annals of Applied Biology, 144, 259-272.
Bell, S., Morse, S., 2004, Experiences with sustainability indicators and stakeholder participation: a
case study relating to a Blue Plan project in Malta, Sustainable Development, 12(1), 1-14.
Bos, M.G., 1997, Performance indicators for irrigation and drainage, Irrigation and Drainage
Systems, 11, 119137.
DPT (State Planning Organisation), 2001, Long-term strategy and eighth fiveyear development
plan, 2001-2005, Ankara, Turkey.
Dungumaro, E.W., Madulu, N.F., 2003, Public participation in integrated water resources
management: the case of Tanzania, Physics and Chemistry of the Earth, 28, 10091014.
Gardner-Outlaw, T., Engelman, R., 1997, Sustaining water, easing scarcity: A second update,
Population Action International, Washington DC.
Gleick, P. H., 1995, Human population and water: To the limits in the 21st century, Pacific Institute
for Studies in Development, Environment, and Security, Oakland, California.
GWP-TAC (Global Water Partnership - Technical Advisory Committee), 2000, Integrated Water
Resources Management, TAC Background Papers no.4, Global Water Partnership.
Hardi, P., Zdan, T. (eds), 1997, Assessing sustainable development: Principles in practice
International Institute for Sustainable Development, Winnipeg, Manitoba, Canada.
Jaspers, F.G.W., 2003, Institutional arrangements for integrated river basin management, Water
Policy, 5, 77-90.
Johnson, C. A., 1997, Public participation and sustainable development: Counting the costs and
benefits, TDRI Quarterly Review, 12(2), 25-32.
Kellett, B. M., Bristow, K. L., Charlesworth, P. B. 2005, Indicator Frameworks for Assessing Irrigation
Sustainability, Land and Water Technical Report No. 01/05, Commonwealth Scientific and
Industrial Research Organisation, Australia.
Kendirli, B., Cakmak, B. Ucar, Y., 2005, Salinity in the Southeastern Anatolia Project (GAP), Turkey:
Issues and options, Irrigation and Drainage, 54, 115122.
Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

300
Lorite, I.J., Mateos, L., Fereres, E., 2004, Evaluating irrigation performance in a Mediterranean
environment: Model and general assessment of an irrigation scheme, Irrigation Science, 23, 77-
84.
Marshall, G.R., 1999, Economics of incorporating public participation in efforts to redress
degradation of agricultural land, 6th Annual Conference of the New Zealand Agricultural and
Resource Economics Society, Christchurch, New Zealand.
McCool, S.F., Stankey, G.H., 2004, Indicators of sustainability: Challenges and opportunities at the
interface of science and policy, Environmental Management, 33(3), 294-305.
Molden, D. J., Sakthivadivel, R., Perry, C.J., de Fraiture, C., Kloezen, W. H., 1998, Indicators for
comparing performance of irrigated agricultural systems, Research Report 20, Colombo, Sri
Lanka: International Water Management Institute.
Mostert, E., 2003, The challenge of public participation, Water Policy, 5, 179-197.
Ostrom, E., 1990, Governing the commons: The evolution of institutions for collective action,
Cambridge: Cambridge University Press.
Ostrom, E., Gardner, R., Walker, J., 1994, Rules, games and common-pool resources Ann Arbor:
The University of Michigan Press.
Pahl-Wostl, C., 2002, Participative and stakeholder-based policy design, analysis and evaluation
processes, Integrated Assessment, 3, 3-14.
Seppala, O.T., 2002, Effective water and sanitation policy reform implementation: need for systemic
approach and stakeholder participation, Water Policy, 4, 367-388.
nver, H.O., 1997, Southeastern Anatolia Project (GAP), Water Resources Development, 13(4),
453-483.



































Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

301




































Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

302
CIHEAM
Centre International de Hautes Etudes Agronomiques Mditerranennes
International Centre for Advanced Mediterranean Agronomic Studies
Conseil d'Administration /
Comit Scientifique Consultatif
Governing Board
/ Scientific Advisory Committee
Prsident / : Moun HAMZ
Membres /
Chairman
Members
Albanie / Albania: Sali METANI Liban / Lebanon: Hady RACHED
Algrie / Algeria: Mohamed Foued RACHEDI Malte / Malt a: Salvino BUSUTTIL
Egypt e / Egypt : ABD EL-AZIM EL-TANTAWI BADAWI Maroc / Marocco: Fouad GUESSOUS
Espagne / Spain: Vicent e FLORES REDONDO Port ugal : Jos Manuel ABECASSISEMPIS
France : Philippe BARR Tunisie / Tunisia: Abdelaziz MOUGOU
Grce / Greece: Elefterios TJAMOS Turquie / Turkey: Vedat MIRMAHMUTOGULLARI
It alie / Italy: Giuliana TRISORIO LIUZZI
Prsident / Chairman
Membres / Members
Velesin PEULI
Faculty of Agricult ure, Agri-environment and Ecology Depart ment
Agricult ural Universit y of Tirana, ALBANIE
Jacques BROSSIER
Prsident du Cent re INRA Dijon
Dijon, FRANCE
Mohamad TALAL FARRAN
Agricult ural Research and Educat ion Cent er (AREC)
American Univerity of Beirut, LEBANON
George ATTARD
Inst it ut of Agricult ure
Universit y of Malt a, MALTA
Teodoro Massimo MIANO
Dipart iment o di Biologia e Chimica Agro-forest ale e Ambient ale (DIBCA)
Universit di Bari, ITALIA
Foued CHEHAT
Inst it ut Nat ional Agronomique dEl Harrach
El Harrach, ALGERIE



Options mditerranennes, Series B, n57 Water Use Efficiency and Water Productivity

303
OM is a CIHEAM series devoted to the development
of Mediterranean agriculture. Having appeared in the
form of a periodical from 1970 to 1976, the title OM
has been given to the "Etudes" series from 1981 to
1989. To date, OM includes three series:
(Ser. A), (Ser. B)
and .
Sminai res
Mditerranens Etudes et Recherches
Cahi er OM
OM est une collection du CIHEAM ddie au
dveloppement de lagricul ture mditerranenne.
Publ ie en forme de pri odique du 1970 1976, l e
titre OM a t donn la col lection du 1981
au 1989. Actuellement, OM comprend trois
col lecti ons:
et
Etudes
Sminaires Mditerranens (Ser. A),
Etudeset Recherches(Ser. B) Cahier OM.
I NTERNATI ONAL CENTRE FOR ADVANCED MEDITERRANEAN AGRONOMI C STUDIES
CENTRE DEHAUTESETUDES AGRONOMIQUES I NTERNATI ONAL MEDI TERRANEENNES
Options Mditerranennes
ISSN : 1016-1228
ISBN : 2-85352-355-1
WASAMED (WAter SAving in MEDiterranean agriculture) isa Thematic Network
funded by the European Commission (INCO-MED Programme). The main
objective of WASAMED is to establish a platform for effective Mediterranean
dialogue on water saving in agriculture, contributing to improved management of
limited water resources and sustainable development in the Mediterranean Region.
Specific objectivesof the Project are:
To improve regional co-ordination of present and future actionsin water savi ng;
To establish a Mediterranean-wide convention to strengthen communication and
sharing of experience among relevant researchers, decision and policy makers,
and end-users;
To develop water saving research projectsand actions that meet with the needs and
concerns arising from the different Mediterranean contexts;
To facilitate accessof different stakeholdersto an easy-to-use knowledge-base;
To create a framework and seek consensusto assist regional planning and EU-
funding in water resources management for the Mediterranean Region.
WASAMED Thematic Network involves 42 partners from decision-policy making
Institutions, Universities and End-Users associations of 16 countries: Algeria,
Cyprus, Egypt, Germany, Greece, Italy, Jordan, Lebanon, Malta, Morocco, Palestine,
Portugal, Spain, Syria, Tunisia, Turkey.
The Workshop of Amman is the fourth of a seriesof five Workshopsplanned between
2003 and 2006, each of them addressing respectively different issues of Water
Saving: Participatory Irrigation Management and Cultural Heritage (in Turkey),
Irrigation Systems Performance (in Tunisia), Use of non-conventional waters (in
Egypt), Water Use Efficiency and Water Productivity (in Jordan), Integration and
Harmonisation of technical water saving options and policies (in Lebanon). A final
International Conference in Bari (Italy) on the whole context of Water Saving
perspectivesin the Mediterranean will close the project in February 2007.

50,00



INTERNATIONAL CENTRE FOR ADVANCED MEDITERRANEAN AGRONOMIC
CENTRE INTERNATIONAL DE HAUTES ETUDES AGRONOMIQUES

S-ar putea să vă placă și