You are on page 1of 89

1.

INTRODUCTION TO IWRM
1.1 Definition of IWRM
Integrated Water Resources Management (IWRM), in the simplest application, is a common
sense logical and appealing concept for water resources management. The fundamental
understanding of IWRM is that the many different uses of water resources are inter-
dependent. High irrigation demands and polluted drainage flows from agriculture mean less
freshwater for drinking or industrial use; contaminated municipal and industrial wastewater
pollutes rivers and threatens ecosystems. If water has to be left in a river to protect fisheries
and ecosystems, less can be diverted to grow crops. There are many examples of the basic
theme that unregulated use of scarce water resources is wasteful and inherently unsustainable.
Integrated management means that all the different uses of water resources are considered
together. Water allocation and management decisions consider the effects of each use on the
others. They are able to take account of overall social and economic goals, including the
achievement of sustainable development. As we shall see, the basic IWRM concept has been
extended to incorporate participatory decision-making. Different user groups (farmers,
communities, environmentalists) can influence strategies for water resource development and
management. This approach brings additional benefits since informed users apply local self-
regulation. It can be done in the context of water conservation and catchment protection that
is far more effective than central regulation and surveillance can achieve.
Management is used in its broadest sense. It emphasizes that we must not only focus on
development of water resources but that we must consciously manage water development in a
way that ensures long term sustainable use for future generations.
Integrated water resources management is a systematic process for the sustainable
development, allocation and monitoring of water resource use in the context of social,
economic and environmental objectives. It contrasts with the sectoral approach as applied in
many countries. In some cases, the responsibility for drinking water rests with one agency,
irrigation water with another and waters for the environment with yet another institution. A
general lack of cross-sectoral linkages leads to un-coordinated water resource development
and management, resulting in conflict, waste and unsustainable systems.
Why IWRM?
Water is vital for human survival, health, dignity and a fundamental resource for
development. The world’s freshwater resources are under increasing pressure. Many still
lacks access to adequate water to meet their basic needs. Population growth, increased
economic activity and improved standards of living result in an increased competition for and
conflicts over limited freshwater resources. The following points demonstrate why people
argue that the world faces impending water crisis:
 Water withdrawals have increased more than twice as fast as population growth and
currently one third of the world's population live in countries that experience medium
to high water stress;
 Pollution is further enhancing water scarcity by reducing water usability downstream;

1
 Shortcomings in the management of water, a focus on developing new sources rather
than managing existing ones better and top-down sector approaches to water
management result in un-coordinated development and management of the resource;
 More and more development means, greater impacts on the environment; and
 Current concerns about climate variability and climate change demands improved
management of water resources to cope with more intense floods and droughts
Hence, engineering, economic, social, ecological and legal aspects need to be considered, as
well as quantitative and qualitative aspects, and supply and demand.
Moreover, also the ‘management cycle’ (planning, monitoring, operation & maintenance,
etc.) needs to be consistent.
Integrated water resources management, then, seeks to manage the water resources in a
comprehensive and holistic way. It therefore has to consider the water resources from a
number of different perspectives or dimensions. Once these various dimensions have been
considered, appropriate decisions and arrangements can be made.
The following are the four dimensions that integrated water resources management takes into
account (Savenije and Van der Zaag, 2000):
1. The water resources, taking the entire hydrological cycle into account, including stock
and flows, as well as water quantity and water quality; distinguishing, for example,
rainfall, soil moisture, water in rivers, lakes, and aquifers, in wetlands and estuaries,
considering also return flows etc.
2. The water users, all sectoral interests and stakeholders
3. The spatial scale, including
• The spatial distribution of water resources and uses (e.g. well-watered upstream
watersheds and arid plains downstream)
• The various spatial scales at which water is being managed, i.e. individual user, user
groups (e.g. user boards), watershed, catchment, (international) basin; and the
institutional arrangements that exist at these various scales
4. The temporal scale; taking into account the temporal variation in availability and
demand for water resources, but also the physical structures that have been built to even
out fluctuations and to better match the supply with demand.
Dimension 1: Water Resources
The water resources include all forms of occurrence of water including salt water and fossil
groundwater. An interesting distinction which can be made is between blue and green water.
Blue water, the water in rivers, lakes and shallow aquifers, has received all the attention from
water resources planners and engineers. Green water, the water in the unsaturated zone of the
soil responsible for the production of biomass has been largely neglected but it is the green
water that is responsible for 60% of the world food production and all of the biomass
produced in forests and pasture. It is this resource which is most sensitive to land
degradation. Fossil water, the deep aquifers that contain non-renewable water, should be

2
considered a mineral resource which can only be used once at the cost of foregoing future
use.
Dimension 2: Water Users
There are many different users of water and its functions. Functions can be split into
production functions (for economic production activities), regulation functions (for
maintaining a dynamic equilibrium in natural processes), carrier functions (to sustain life
forms) and transfer functions (as a contribution to culture, religion and landscape). The uses
include: households, industries, agriculture, fisheries, ecosystems, hydropower, navigation,
recreation, etc. Water users consist of consumptive and non-consumptive (often in-stream)
users. Besides on quantity, the users depend largely on the quality of the resource. With
regard to the consumptive use an important concept is that of “virtual” water, where products
are expressed in the amount of water required for its production. This concept is both useful
as a measure for efficiency and for the discussion on food security
Dimension 3: Spatial Scales
Water resources issues are apparent at different levels: the international level, the national
level, the province or district level and the local level. Parallel to these administrative levels
are hydrological system boundaries such as river basins, sub-catchments and watersheds.
Hydrological boundaries seldom concur with administrative boundaries. River basins seem
appropriate units for operational water management but present problems for institutions that
have a different spatial logic.
At all times, two-way interaction should take place between the institutional levels. Different
decisions on water resources management belong at different levels, meaning that the concept
of subsidiarity (decision making at the lowest appropriate level) needs to be a guiding
principle in the development of IWRM. Interests and decisions at lower levels need to be
carried upward to be taken into consideration at higher levels, particularly to the national and
international level. An important element in this process is the participation of stakeholders in
decision-making processes at all levels.
Dimension 4: Temporal Scales and Patterns
Both the water resources themselves and the water uses have distinct temporal patterns. The
temporal distribution of water resources is crucial (floods, droughts, base flows, flooding
patterns) and so is the distribution over time of the demands (peak demands, constant
requirements, cropping patterns, etc.). In water resources assessments the total amount of
water available depends strongly on the possibility to capture flood flows. The staging of
demands (simultaneous or staggered demands) can have a large influence on the development
required.
Integrated Water Resources Management can now be defined as:
Integrated Water Resources Management (IWRM) is a process which promotes the
coordinated 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 (Global Water Partnership, 2000).

3
Integrated Water Resources Management therefore acknowledges the entire water cycle with
all its natural aspects, as well as the interests of the water users in the different sectors of a
society (or an entire region). Decision-making would involve the integration of the different
objectives where possible, and a trade-off or priority-setting between these objectives where
necessary, by carefully weighing these in an informed and transparent manner, according to
societal objectives and constraints. Special care should be taken to consider spatial scales, in
terms of geographical variation in water availability and the possible upstream-downstream
interactions, as well as time scales, such as the natural seasonal, annual and long-term
fluctuations in water availability, and the implications of developments now for future
generations.
IWRM in Ethiopia
Ethiopia is endowed with huge water resource potential (about 122 Bm³) annual surface
runoff and 2.9 Bm3 groundwater) though it is characterized by uneven spatial and temporal
distributions. Most of the rivers in Ethiopia are seasonal and about 70% of runoff is obtained
between June and August. Irrespective of the huge potential, the country’s water resources
have contributed very little to the socio-economic development; on average, access to clean
and safe water supply is about 50%; Irrigation stands at only 6% of the potential and that of
hydropower is at only 2% despite the big potential (2 nd in Africa). Most of the rivers that
originate within the country flow across borders to neighboring countries, and are
transboundary rivers.
The policy environment is highly supportive of IWRM approaches. Ethiopia’s five-year Plan
for Accelerated Sustainable Development to End Poverty (PASDEP) places water
(particularly water supply and sanitation) as a high priority. The Plan references the overall
objective of the National Water Resources Management Policy, which is to enhance and
promote efforts towards an efficient, equitable, and optimally utilized water resources that
would contribute to the country's socioeconomic development on a sustainable basis
(MOWR, 1999).
Ethiopia has adopted the principles of IWRM and has already put in place an appropriate
water policy, legislation, strategy, and development program (including master plans) that
embrace IWRM principles and approaches. Therefore, it could be said that the country has a
National IWRM Plan and is at a stage of implementation. However, there are constraints in
implementation that include capacity limitations, lack of proper coordination/collaboration
among various stakeholders (as sectoral interests dominate), and lack of integrated and
participatory approaches in planning and implementation of water resources.
To complement the Government’s efforts of addressing the above-mentioned constraints,
since 2005 the US Government has supported a project that promotes IWRM in Ethiopia
through the Ethiopia Country Water Partnership. The main objective of the project is to pilot
the principles of IWRM in two selected watersheds, to conduct national level advocacy and
awareness raising on IWRM, and to share lessons from the pilots for scaling up at various
levels.

4
The project was implemented in two river basins, Berki and Mesena. The situation
summarized in Box 1 below illustrates that water is the scarcest resource in the Berki
catchment and cause of conflict among upstream and downstream communities; the
administrative authorities; and local businesses and NGOs. The root cause of the complex
challenge has been identified to be lack of institutional framework for stakeholders’
participation and low level of IWRM awareness among the stakeholders. The biggest
challenge was ensuring sustainable and equitable use and management of the water resources
of the Berki catchment for all the interest groups.

In Messena, most of the inhabitants of the area depend on growing crops for their livelihoods;
however,
Box their The
1: Agulae: farmlands
illusoryare seasonally
border betweenflooded
“scarcitybyandtheplenty”
overflow of the Borkena River.
Problems of drainage mean that communities often lose their crops and are unable to feed
Water is scarcePastoralists
their families. resource in from
the Berki
Oromiacatchment,
Zone of and there Region
Amhara were various water
and from resource
Afar Regional
State, amongproblems,
management others, migrate during
including the dry
conflicts seasons
among in search
upstream and of pasture andcommunities
downstream water for their
cattle. Some people migrate from as far as Bati, more than 100 kms and even Mile which is
and between administrative authorities due to the diverse interests mentioned above. The
more than 200 kms away.
different water use activities have exerted heavy pressure on water availability for
Unlike the Berki Catchment, water in the Mesena watershed is not scarce. The main problems
different purposes,
of the Mesena especiallyare
watershed for high
the downstream
floods from users.
the Inefficient
mountain use dueoftowater, including and
deforestation
application
uncontrolledofgrazing,
inefficient technologies
and drainage was also
of farmlands due tocommon
overflows practice.
from theCommunities
Borkena River.
All the stakeholders
downstream of Agula aretownworking
(outsideindependently without integrating
of the Berki catchment) suffer fromorlack
worrying
of waterabout
due the
sustainability of the system. There is not adequate knowledge about the water resources
to the upstream pumps and diversions.
potential of the catchment, although it is being exploited by many users.
They need to IWRM
Implementing travel long distances
can often seem to access water
overwhelming especially
given the scaleduring dry seasons.
and complexity of the
Moreover, the upstream water users pollute water that is being used by the downstream and
changes needed. This project has supported great experience on how IWRM is a long
participatory process based on an on-going learning process in Ethiopia. The IWRM pilot
users due to their washing and cattle drinking practices. It can be said that water resources
project in the “Berki” watershed for example has shown that IWRM involves many changes
of
to the
thecatchment were being
existing system excessively
through a step-byexploited beyond the
step- approach thatnatural
createslimits of the
a sense of system
ownership
amongst
and all stakeholders.
the ability The offices
of the regulatory case/project demonstrated
to control it. Absencethat
of againing
land usepolitical
plan andsupport
water at
various levelshaveand
ledmulti-stakeholder platforms are crucial for pumps,
the success of the IWRM
regulations to the uncontrolled introduction of private and changes in
process. Building stakeholders’ capacities (tailored trainings, awareness workshops and
cropping
experience pattern
sharingand land use.
activities) Moreover,
also the waterrole
played significant andinother naturaltheresources
facilitating process. of the
catchment were not assessed and decisions were taken without adequate information. Poor
1.2 Benefit of IWRM to sectors
communication among various users and stakeholders and low level of awareness also
Ecosystem
contributed to the problem.
 A voice for environmental needs in water allocation
 Raising awareness among other users of the needs of ecosystems
 More attention to an ecosystem approach to water management
 Protecting upper catchments, pollution control, and environmental flows
 Safeguarding common resources such as forests, wetlands and fishing grounds on
which communities depend
Agriculture

5
 Implications for agriculture of water use by other sectors considered in the management
process
 Rational decision making on water use in which costs and benefits are considered
 More effective use of water within the sector and hence increased returns
 Multi-purpose water resource development and cross-sectoral recycling (e.g. use of
reclaimed municipal wastewater for irrigation)
Water supply
 Increased security of domestic water supplies
 Reduced conflicts between water users
 Increasing recognition of the economic value of water leading to more efficient use
 Increased use of water demand management
 Improved waste management considering environmental effects and human health and
hygiene
1.3 The basic planning unit and concluding remarks
 Terminology:
- Drainage basin/Catchment: the area of land over which the water drains to a single
outlet;
- Watershed: either the equivalent of a drainage basin, or the delineating points
(divide) where water flows to two different outlets; and
- River basin: the drainage basin for a river system.
 Water follows its own boundaries: the river basin, lake-basin, or groundwater aquifer.
 Analyses and discussions of water allocation between users and ecosystem would
make sense when addressed at basin level.
 Hence, a lot of the “integration” in IWRM takes place at basin scale: local catchment
or aquifer, multi-state, multi-country/trans-boundary river basin.
To conclude, Integrated Water Resources Management requires a paradigm shift of thinking,
such as:
• From fragmentation to integration
- Sectoral to cross-sectoral integration.
• From single-risk assessment to multiple-risk assessment
- Water supply and sanitation risks;
- Agricultural risks;
- Aquatic environment risks;
- etc.
• From conflict to prioritizing
- When there is competition for water resources, it brings into the open the need to
justify the allocation of water to one user rather than to another.

6
- This value assessment should take into account both the benefits and the negative
impacts.

2. PRINCIPLES OF IWRM
2.1 Introduction
The two most commonly referred guiding principles of integrated water resources
management are the Dublin Statement and the Rio Declaration.
The four Dublin guiding principles
A meeting in Dublin in 1992 gave rise to four principles that have been the basis for much
of the subsequent water sector reform.
Principle 1. Freshwater is a finite and vulnerable resource, essential to sustain life,
development and the environment.
The notion that freshwater is a finite resource arises as the hydrological cycle on average
yields a fixed quantity of water per time period. This overall quantity cannot yet be altered
significantly by human actions, though it can be, and frequently is, depleted by man-made
pollution. The freshwater resource is a natural asset that needs to be maintained to ensure that
the desired services it provides are sustained. This principle recognizes that water is required
for many different purposes, functions and services; management therefore, has to be holistic
(integrated) and involve consideration of the demands placed on the resource and the threats
to it.
The integrated approach to management of water resources necessitates co-ordination of the
range of human activities which create the demands for water, determine land uses and
generate waterborne waste products. The principle also recognizes the catchment area or river
basin as the logical unit for water resources management.
Principle 2. Water development and management should be based on a participatory
approach, involving users, planners and policymakers at all levels.
Water is a subject in which everyone is a stakeholder. Real participation only takes place
when stakeholders are part of the decision-making process. The type of participation will
depend upon the spatial scale relevant to particular water management and investment
decisions. It will be affected too by the nature of the political environment in which such
decisions take place. A participatory approach is the best means for achieving long-lasting
consensus and common agreement. Participation is about taking responsibility, recognizing
7
the effect of sectoral actions on other water users and aquatic ecosystems and accepting the
need for change to improve the efficiency of water use and allow the sustainable development
of the resource. Participation does not always achieve consensus; arbitration processes or
other conflict resolution mechanisms also need to be put in place.
Governments have to help create the opportunity and capacity to participate, particularly
among women and other marginalized social groups. It has to be recognized that simply
creating participatory opportunities will do nothing for currently disadvantaged groups unless
their capacity to participate is enhanced. Decentralizing decision making to the lowest
appropriate level is one strategy for increasing participation.

Principle 3. Women play a central part in the provision, management and safeguarding of
water.
The pivotal role of women as providers and users of water and guardians of the living
environment has seldom been reflected in institutional arrangements for the development and
management of water resources. It is widely acknowledged that women play a key role in the
collection and safeguarding of water for domestic and – in many cases – agricultural use, but
that they have a much less influential role than men in management, problem analysis and the
decision-making processes related to water resources.
IWRM requires gender awareness. In developing the full and effective participation of
women at all levels of decision-making, consideration has to be given to the way different
societies assign particular social, economic and cultural roles to men and women. There is an
important synergy between gender equity and sustainable water management. Involving men
and women in influential roles at all levels of water management can speed up the
achievement of sustainability; and managing water in an integrated and sustainable way
contributes significantly to gender equity by improving the access of women and men to
water and water-related services to meet their essential needs
Principle 4. Water has an economic value in all its competing uses and should be
recognized as an economic good.
Within this principle, it is vital to recognize first the basic right of all human beings to have
access to clean water and sanitation at an affordable price. Managing water as an economic
good is an important way of achieving social objectives such as efficient and equitable use,
and of encouraging conservation and protection of water resources. Water has a value as an
economic good as well as a social good. Many past failures in water resources management
are attributable to the fact that the full value of water has not been recognized.
Value and charges are two different things and we have to distinguish clearly between them.
The value of water in alternative uses is important for the rational allocation of water as a
scarce resource, whether by regulatory or economic means. Charging (or not charging) for
water is applying an economic instrument to support disadvantaged groups, affect behavior
towards conservation and efficient water usage, provide incentives for demand management,
ensure cost recovery and signal consumers’ willingness to pay for additional investments in
water services.
8
Treating water as an economic good is an important means for decision making on the
allocation of water between different water use sectors and between different uses within a
sector. This is particularly important when extending supply is no longer a feasible option.
The Rio declaration
Two important sections concerning IWRM:
– Integrated Water Resources Management is based on the perception of water as an
integral part of the ecosystem, a natural resource, and a social and economic good,
whose quantity and quality determine the nature of its utilization. To this end, water
resources have to be protected, taking in to account the functioning of aquatic
ecosystems and the perenniality of the resource, in order to satisfy and reconcile needs
for water in human activities. In developing and using water resources, priority has to
be given to satisfaction of basic needs and the safeguarding of ecosystems. Beyond
these requirements, however, water users should be charged appropriately.
– Integrated Water Resources Management, including the integration of land and water
related aspects, should be carried out at the level of the catchment basin or sub basin.
Four principal objectives should be pursued, as follows:
a) To promote a dynamic, interactive, iterative and multisectoral approach to water
resources management, including the identification and protection of potential sources
of fresh water supply, that integrates technological, socio-economic, environmental
and human health considerations.
b) To plan for sustainable and rational utilization, protection, conservation and
management of water resources based on community needs, and priorities within the
framework of national economic development policy.
c) To design, implement and evaluate projects and programmes that are both
economically efficient and socially appropriate within clearly defined strategies,
based on an approach of full public participation, including that of women, youth,
indigenous people, local communities, in water management policy making and
decision making.
d) To identify and strengthen or develop, as required, in particular in developing
countries, the appropriate institutional, legal, and financial mechanisms to ensure that
water policy and implementation are a catalyst for sustainable social progress and
economic growth.
2.2 Water Resources Systems and Sustainability
Water resources systems can be defined as a set of water resources elements linked by
interrelationships into a purposeful whole. In many respects, water resources systems defy
rational description. To an engineer, these systems may be dams and weirs, tunnels, levees,
pipelines, electrical power plants, water treatment and reclamation, spillways and similar
physical works which have been constructed to provide certain benefits. An economist views
them from the point of view of economic efficiency, income redistribution and stimulation of
economic growth. To a lawyer, a water resources system is a device for the implementation
of water rights. To those living in an arid environment, water resources systems mean food

9
and fiber, homes and jobs, laws and politics. To many conservationists, water resources
systems are unwanted interventions, responsible for the destruction of wild rivers, scenic
beauty and wildlife habitat.
Water resources systems indeed include all these points of view which could be physical,
technological, sociological, biological, legal, geological and agricultural.
The elements of a water resources system can be either natural (rivers, lakes, glaciers, etc.) or
artificial (reservoirs, barrages, weirs, canals, hydroelectric power plants, etc.).
The relationships between the elements are either real (e.g., water diversion) or conceptual
(e.g., organisation, information, etc.).
Sustainability of water resources
• “Development that meets the needs of the present generation without compromising
the ability of future generations to meet their own needs.” (WCED, 1987)
• Sustainable development is making efficient use of our natural resources for
economic and social development while maintaining the resource base and
environmental carrying capacity for coming generations.
• This resource base should be widely interpreted to contain besides natural resources:
knowledge, infrastructure, technology, durables and human resources.
• In the process of development natural resources may be converted into other durable
products and hence remain part of the overall resource base.
Sustainable water resource systems are those designed and managed to fully contribute to the
objectives of society, now and in the future, while maintaining their ecological,
environmental and hydrological integrity. (ASCE, 1998)
Sustainable water resource systems are those designed and managed to fully contribute to the
objectives of society, now and in the future, while maintaining their ecological,
environmental and hydrological integrity.
Three types of sustainability are briefly discussed below: physical, economic and
institutional.
Physical sustainability
Physical sustainability means closing the resource cycles and considering the cycles in their
integrity (water and nutrient cycles). In agriculture this implies primarily closing or
shortening water and nutrient cycles so as to prevent accumulation or depletion of land and
water resources: Water depletion results in desertification. Water accumulation into water
logging. Nutrient depletion leads to loss of fertility, loss of water holding capacity, and in
general, reduction of carrying capacity. Nutrient accumulation results in eutrophication and
pollution. Loss of top-soil results in erosion, land degradation and sedimentation elsewhere.
Closing or shortening these cycles means restoring the dynamic equilibriums at the
appropriate temporal and spatial scales. The latter is relevant, since at a global scale all cycles
close. The question of sustainability has to do with closing the cycles within a human
dimension.

10
Economic sustainability
Economic sustainability relates to the efficiency of the system. If all societal costs and
benefits are properly accounted for, and cycles are closed, then economic sustainability
implies a reduction of scale by short-cutting the cycles. Efficiency dictates that cycles should
be kept as short as possible. Examples of short cycles are: water conservation, to make
optimum use of rainfall where it falls (and not drain it off and capture it downstream to pump
it up again); water recycling at the spot instead of draining it off to a treatment plant after
which it is conveyed or pumped back over considerable distances etc.
Strangely enough, economic sustainability is facilitated by an enlargement of scale through
trade in land- and water-intensive commodities (the "virtual" water concept).
The use of virtual water is an important concept in countries where the carrying capacity of a
society is not sufficient to produce water intensive products itself.
The closing of cycles should be realized at different spatial scales:
• The rural scale, implying water conservation, nutrient and soil conservation,
prevention of over-drainage and the recycling of nutrients and organic waste.
• The urban scale, both in towns and mega-cities, implying the recycling of water,
nutrients and waste.
• The river basin scale, implying: soil and water conservation in the upper catchment,
prevention of runoff and unnecessary drainage and enhancement of infiltration and
recharge, flood retention, pollution control and the wise use of wetlands.
• The global scale, where water, nutrient and basic resource cycles are integrated and
closed. The concept of virtual water is a tool for an equitable utilization of water
resources. This requires an open and accessible global market and the use of resource-
based economic incentives such as resource taxing ("Green tax" which taxes the use
of non-renewable or finite resources), as opposed to taxing renewable resources such
as labor, which is the general practice today.
Institutional sustainability
In order to ensure sustainability, the right decisions have to be made. This requires that the
relevant institutions are in place which can facilitate the proper decision processes.
Moreover, institutions need to adequately respond to changing requirements and a changing
environment in which they operate. They should have the capacity to adapt to emerging
circumstances. Their adaptive capacities indicate whether they will prove to be sustainable
institutions. According to Costanza (1994),
A sustainable system is active and able to maintain its structure (organization), function
(vigor) and autonomy over time and is resilient in stress.
Integrated water resources management requires strong institutions; sustainable systems in
Costanza’s sense. Sustainable institutions require good governance; while institutions that are
governed wisely are likely to retain their resilience and will be sustained over time. Thus, it
appears that sustainable institutions and good governance go hand in hand. They need and
presuppose each other.

11
Water Resources Sustainability: How is it measured?
Sustainable development balances the exploitation of natural resources, technology
development, and institutional change to enhance the potential to meet human needs and
aspirations, now and in the future.
To achieve sustainability, all the components in the system must be also in balance.
Water management policies that promote sustainable water resources systems are becoming
more difficult to identify because of environmental considerations, water scarcity, and
climate change.
Recently, strong emphasis has been placed on the adaptive capacity of water resource
systems, which refers to measures that reduce the vulnerability of systems to actual or
expected future changes. Vulnerability is the magnitude of an adverse impact on a system.
Thus, the objective is to look for policies that reduce the adverse impacts of actual and
expected events, and to the extent possible, meet the water requirements for humans and the
environment, now and in the future. To accomplish this goal, it is necessary to have
performance measures or indexes that allow the evaluation and comparison of water
resources systems under different scenarios.
The sustainability index (SI) summarizes the performance of alternative policies from the
perspective of water users and the environment; it is also a measure of a system’s adaptive
capacity to reduce its vulnerability.
If a proposed policy makes the system more sustainable, the index will show that the system
will have a larger adaptive capacity.
The SI is an integration of performance criteria that capture the essential and desired
sustainable characteristics of the basin. The index facilitates comparison of policies when
there are trade-offs among performance criteria.
Performance Criteria
Performance criteria are used to evaluate water management policies and enable the
comparison of alternative policies. Performance criteria can be simple averages, such as
system storage, water supply, evaporation, municipal shortfalls (average deficits), and
outflow of water from a system Probability based performance criteria include time-based
(annual, monthly) and volumetric reliability and resilience
Reliability
Water demand reliability is the probability that the available water supply meets the water
demand during the period of simulation
Resilience
Resilience is a system’s capacity to adapt to changing conditions. Because climate conditions
are no longer steady, resilience must be considered as a statistic that assesses the flexibility of
water management policies to adapt to changing conditions.
Vulnerability

12
Vulnerability is the likely value of deficits, if they occur. Essentially, vulnerability expresses
the severity of failures.
Sustainability Index
An index is a “synthesis of numerous actors into one given factor”
To quantify the sustainability of water resources systems, Loucks (1997) proposed the SI,
with the objective to facilitate the evaluation and comparison of water management policies.
The SI is a summary index that measures the sustainability of water resources systems; it can
be used to estimate the sustainability for water users and to obtain the change in sustainability
by comparing the index among several water policies proposed. Frequently, indexes are
criticized because they are seen as a sum of disparate items, and sometimes in practice,
people in the water sector are reluctant to use indexes.
The SI summarizes essential performance parameters of water management in a meaningful
manner, rather than adding broad factors, and the SI has been used by the scientific
community.
Sustainability by User
Loucks (1997) proposed the following SI for the ith water user:

SI i =Rel i∗Res i∗( 1−Vuli )


The SI has the following properties:
1) Its values vary from 0–1;
2) If one of the performance criteria is zero, the sustainability will be zero also; and
3) There is an implicit weighting because the index gives added weight to the criteria
with the worst performance.
Sustainability by Group
To compare groups of water users, the sustainability by group (SG) was defined as a
weighted average of sustainability indexes (Loucks 1997). The SG is used to calculate the
sustainability for a group k with ith to jth water users belonging to this group:
1= j ∈k
k
SG = ∑ W i∗SI i
i=1∈ k

Where Wi = relative weight for the ith water user, ranging from 0–1 and summing to one. If
the SI of each user is weighted by its annual water demand, the SG for the k th group is
expressed as
i= j∈ k
water demand i
SGk = ∑ k
∗SI i
i=1 ∈k Water demand
i= j ∈k
k
Where Water demand = ∑ water demand i
i=1 ∈k

13
3. WATERSHED MANAGEMENT
3.1 Definition of IWM
The concept of a watershed is basic to all hydrologic designs.  Since large watersheds are
made up of many smaller watersheds, it is necessary to define the watershed in terms of
a point.  This point is usually the location at which the design is being made and is referred to
as the watershed “outlet”.  With respect to the outlet, the watershed consists of all land area
that “sheds” water to the outlet during a rainstorm.  Using the concept that “water runs
downhill”, a watershed is defined by all points enclosed within an area from which rain
falling at these points will contribute water to the outlet. Figure 1 depicts the delineation of a
watershed boundary.

Figure 1: Delineation of a watershed boundary.

14
Integrated Watershed Management (IWM) is the process of managing human activities
and natural resources on a watershed basis. This approach allows us to protect important
water resources, while at the same time addressing critical issues such as the current and
future impacts of rapid growth and climate change.
Why Watershed Approach?
• Watersheds are among the most basic units of natural organization in landscapes.
• The limits of watersheds are defined by topography and the resulting runoff patterns
of rainwater.
• The entire area of any watershed is therefore physically linked by the flow of
rainwater runoff.
• Consequently, processes or activities occurring in one portion of the watershed will
directly impact downstream areas (land or water).
• When detrimental activities like clear-cut deforestation occur, negative impacts are
carried downstream in the form of eroded sediments or flooding.
• Poor agricultural land management activities like excess fertilizer application convey
negative impacts to downstream areas in the form of eutrophication and possible fish
kills.
3.2 Watershed characteristics
The main characteristics which affect behaviors of watersheds and need to be studied for
developing a program of watershed management are: (1) drainage area, (2) shape, (3),
watershed length (4) watershed slope, (5) precipitation, (6) vegetation, (7) geology and soil,
and (8) land use.
Drainage Area
The drainage area (A) is probably the single most important watershed characteristic for
hydrologic design.  It reflects the volume of water that can be generated from rainfall.  It is
common in hydrologic design to assume a constant depth of rainfall occurring uniformly over
the watershed.  Under this assumption, the volume of water available for runoff would be the
product of rainfall depth and the drainage area.  Thus the drainage area is required as input to
models ranging from simple linear prediction equations to complex computer models.
Watershed Length
The length (L) of a watershed is the second watershed characteristic of interest.  While the
length increases as the drainage increases, the length of a watershed is important in
hydrologic computations.  Watershed length is usually defined as the distance measured
along the main channel from the watershed outlet to the basin divide.  Since the channel does
not extend to the basin divide, it is necessary to extend a line from the end of the channel to
the basin divide following a path where the greatest volume of water would travel.  The
straight-line distance from the outlet point on the watershed divide is not usually used to
compute L because the travel distance of floodwaters is conceptually the length of

15
interest.  Thus, the length is measured along the principal flow path.  Since it will be used for
hydrologic calculations, this length is more appropriately labeled the hydrologic length.

While the drainage area and length are both measures of watershed size, they may reflect
different aspects of size.  The drainage area is used to indicate the potential for rainfall to
provide a volume of water.  The length is usually used in computing a time parameter, which
is a measure of the travel time of water through a watershed.

Watershed Slope
Flood magnitudes reflect the momentum of the runoff.  Slope is an important factor in the
momentum.  Both watershed and channel slope may be of interest.  Watershed slope reflects
the rate of change of elevation with respect to distance along the principal flow
path.  Typically, the principal flow path is delineated, and the watershed slope (S) is
computed as the difference in elevation (∆E) between the end points of the principal flow
∆E
path divided by the hydrologic length of the flow path (L):     S=
L
The elevation difference ∆E may not necessarily be the maximum elevation difference within
the watershed since the point of highest elevation may occur along a side boundary of the
watershed rather than at the end of the principal flow path. 
Watershed Shape
Basin shape is not usually used directly in hydrologic design methods; however, parameters
that reflect basin shape are used occasionally and have a conceptual basis.  Watersheds have
an infinite variety of shapes, and the shape supposedly reflects the way that runoff will
“bunch up” at the outlet.  A circular watershed would result in runoff from various parts of
the watershed reaching the outlet at the same time.  An elliptical watershed having the outlet
at one end of the major axis and having the same area as the circular watershed would cause
the runoff to be spread out over time, thus producing a smaller flood peak than that of the
circular watershed. 
Geology and Soil
Geological formation and rock types in a watershed affect water erosion, erodibility of
channel, occurrence of slides and sediments production. Physical and chemical properties of
soils like texture, structure, depth etc. affect infiltration, storage and runoff of water and
erodibility, transportation and deposition of soil. In light textured soils like sands and loamy
sands, infiltration is more resulting less runoff and vice versa in the case of heavy textured
soils like clay.
Precipitation
Precipitation is the major climatic parameter affecting watershed functioning. The amount,
intensity and frequency of rainfall should be studied for watershed management.
Vegetation

16
Vegetation provides a cover on soil in a watershed and, thus, regulates its functioning.
Infiltration, runoff, retention of water and soil erosion and sedimentation of tanks depend on
vegetation. Vegetation in watershed can be temporary like agricultural crops and permanent
type like forest trees, shrubs, grasses and orchards. Watershed behavior depends on density,
composition and nature of the vegetation cover over the watershed.
Land Use
Watershed behaviors also depend on type of land use, its management and areal extent.
Improper land use results deterioration of watershed. The watershed lands should be used
according to its capability. The land use is under the control of watershed inhabitants. It
should be managed judiciously to get maximum benefit.
In addition to above discussed characteristics, geomorphology, orientation, temperature and
ground water also play their role in watershed management.
A number of watershed parameters have been developed to reflect basin shape.  The
following are a few typical parameters:
1. Length to the center of area (Lca):  the distance in miles measured along the main channel
from the basin outlet to the point on the main channel opposite the center of area.
0.3
2. Shape Factor (Ll): Ll=( L Lca ) , where L is the length of the watershed in miles
P
3. Circularity ratio (Fc): F c = , where P and A are the perimeter (ft) and area (ft 2) of
( 4 πA )0.5
the watershed, respectively.
A
4. Circularity ration (Rc): Rc = , where A0 is the area of a circle having a perimeter equal
Ao
to the perimeter of the basin.
2
5. Elongation Ration (Re): Re = , where Lm is the maximum length (ft) of the basin
Lm ( A /π )0.5
parallel to the principal drainage lines.
Generally, the shape factor (Ll) is the best descriptor of peak discharge.  It is negatively
correlated with peak discharge (i.e. as the Ll decreases, peak discharge increases).
3.3 Deriving the watershed parameters
Different watershed parameters can be derived using remote sensing and geographic
information systems.
The Remote sensing and Geographical Information System applications are widely used for
thematic mapping of natural resources for planning, development, management, monitoring
and also in environmental studies. The integration of remote sensing and conventional data
results in decision support system for effective management of land and water resources in a
watershed. The benefits associated with the use of GIS in watershed and hydrologic analysis
include the improved accuracy, less duplication, easier map storage, more flexibility, ease of
data sharing, timeliness, greater efficiency and higher product complexity.

17
The conventional hydrologic data are inadequate for the purpose of design and operation of
water resources systems. In such cases remote sensing data are of great use for the estimation
of relevant hydrological data. RS data can serve as model input for the determination of river
catchment characteristics, such as land use /land cover, geomorphology, slope, drainage etc.
GIS offers the potential to increase the degree of definition of spatial sub units, in number
and in descriptive detail.
3.4 Watershed components and processes
In physical terms, a watershed is an area from which water drains to a common point. This
trait results in a set of physical and biological interactions and processes that causes the
watershed to function as an ecological unit. Watersheds can be considered at a range of
nested scales, beginning with the area contributing to a small first-order stream and
culminating with the world’s great river basins. Ultimately, stream processes that create
habitat integrate the physical and biological processes occurring across the contributing
watershed.
Watersheds are the geographic areas that channel drainage into a river or stream system.
They are defined by topographic boundaries and depending on where they are located might
encompass complex natural ecosystems, highly urbanized landscapes, or elements of both.
The term “watershed processes” refers to the dynamic physical and chemical interactions that
form and maintain the landscape on the scale of a watershed.
Both natural and built environment watersheds have three distinct characteristics: (1) upland
zones that intercept, infiltrate, and transport rain as groundwater and surface water flow, (2)
riparian zones that border surface water bodies, filter surface water runoff, and provide shade
that can lower water temperature, and (3) surface water bodies, such as rivers and lakes, that
provide habitat, food, and water to aquatic and terrestrial species. Naturally functioning
watersheds can also provide migratory corridors and habitat connectivity for birds and
mammals.
Watersheds can assimilate some level of land development and air and water contamination
while continuing to provide the above-stated benefits. Converting natural watersheds to land
uses such as agriculture, housing, and industrial activities can directly interact with natural
forces to change the condition of the land and water as well as the species associated with it.
An example of this type of change is increasing impervious surfaces (such as covering soil
with concrete or asphalt) in urban areas. This leads to decreased water retention and
infiltration, which then leads to increased water runoff. Increased water runoff can in turn
lead to increased erosion and can carry contaminants into surface water bodies, where
increased stream velocities resulting from increased flows can significantly alter the stream
channel and reduce the presence, abundance, and diversity of aquatic species.
3.4.1 Watershed Components
Across landscapes, two controlling factors - climate and geology - create three basic
ecosystem components: soil, vegetation, and water (Note: the effects of animals on soil and
vegetation will be ignored for the sake of simplicity). These components are overlaid on, and
influenced by, topography that is also shaped by climate and geology. Within watersheds,
the interactions of these components result in yields of stream flow and sediment with
patterns of timing, quantity, and quality characteristic of each watershed. These yields of

18
water and sediment, in turn interacting with riparian vegetation (and, in steep, forested
watersheds, large wood delivered from upland sources), form the stream channel and
associated aquatic habitat.
Soil
The soil mantle is a natural storage reservoir for water delivered to the watershed, absorbing
rain or snowmelt and gradually transmitting it down slope. Thus, water stored in the soil is a
primary source of stream flow between storms or periods of snowmelt. The storage capacity
of soil depends on its depth and texture, (i.e., the total pore space available). The rate at
which soil water is delivered to the stream system depends on slope, and soil texture and
structure. Well-developed soils have many sizes of pores with varying degrees of
connectedness. Large pores allow rapid infiltration and drainage of water to and from the
soil mass; small pores absorb water more gradually and retain it longer, making water
available during dry periods for use by plants, or for slow seepage into the stream system.
The development of soil depends upon geology, topography, time, climate, disturbance
factors, and biological agents (e.g., vegetation and soil organisms). The protective vegetative
cover above ground and stabilizing strength of roots below ground are critical to soil
development and stability, particularly on steep slopes.
The rate and magnitude of soil erosion and sedimentation are vital for watershed community
i.e. people and animal. Accelerated erosion:
 Results in a loss of land mass as a natural resource.
 Adversely affects the productivity and production on agricultural horticultural, forest
lands and grasslands; (at higher elevation).
 Results in higher production due to transportation of silt and sediments into the rivers
and reservoirs and then on agricultural land but also results into the reduction of life
and benefits of the reservoir; (in plains).
 results in situation of streams and rivers which causes flash floods;
 results in severe drought due to low soil infiltration, excessive runoff etc.;
 results in converting useful economic assets into useless commodities by formation of
ravines etc., and
 Results in damage to property.
Water
Quantity, quality and timing of water discharged from a watershed are integrated results of
watershed processes. Distributed across the landscape in the form of rain or snow, water is
transported through the watershed, leaving by way of transpiration, evaporation, stream flow,
and groundwater flow. Climate, topography, soil, and vegetation control the processing of
water through the watershed. Because the combination of these factors is unique to each
watershed, the characteristic timing and magnitude of flows through the stream system
constitute the ‘signature’ of the watershed. For example, arid watersheds, in addition to
sparse vegetation, typically have thin, poorly developed soils with low infiltration rates and
little water-holding capacity. Where arid conditions are combined with steep terrain, runoff
tends to occur rapidly following precipitation events, resulting in a ‘flashy’ hydrograph that

19
peaks and declines swiftly. Conversely, where climate supports dense vegetation, an
undeveloped watershed with gentle relief will tend to gradually yield high flows that gently
peak and taper off into strongly sustained base flows. Characteristic elements of the
hydrologic "signature" of watersheds include: 1) high flows - reflecting snowmelt, prolonged
winter rainfall, rain-on-snow events, or intense summer rainstorms, 2) rates of recession from
peak to low flows, and 3) low flows – reflecting groundwater discharge, or water released
from natural storage features such as wetlands and lakes.
In watersheds where rainfall is the dominant form of precipitation, runoff occurs in response
to storm events and the ability of the watershed to store precipitation. To a large degree, this
ability is dictated by soil moisture conditions prior to the onset of the storm. Obviously,
frozen and saturated soils have virtually no storage capability, and rain falling on them will
be quickly delivered to the stream system. Conversely, rainfall delivered at the end of a long,
dry period may do no more than replenish soil moisture, causing little response in stream
flow.
Vegetation
Vegetation performs a variety of functions on the watershed scale. It provides strength and
roughness across the surface of the watershed, thereby slowing the movement of water and
increasing resistance to erosion while promoting the development of deep soils. The
vegetative canopy intercepts precipitation, allowing a portion to evaporate before reaching
the ground, but subsequently inhibiting evaporation from the ground surface. Water use by
vegetation (i.e., evapotranspiration) removes water from the soil. Vegetative litter slows
overland flow and protects the soil surface from raindrop impacts, preventing splash erosion
and the sealing of surface pores. Root channels increase infiltration capacity. The presence of
decayed vegetation and other organic matter characterizes the topsoil, and greatly influence
its properties and structure.
Water use by vegetation reduces total runoff from the associated land areas. However, the
combined influences of vegetation and soil also greatly attenuates the movement of water
through the watershed, dampening peak flows, sustaining stream flow during dry periods,
and maintaining high water quality.
3.4.2 Watershed processes
The types of general watershed processes are given in Table 1.
Table 1: Watershed Functions and Characteristics
Watershed Process Functions and Characteristics

Soils, soil processes, Soil structure, soil organics, soil water repellency, soil moisture,
and infiltration, organic matter mineralization, soil and root
interactions, soil formation, topography, soil transport, and soil
erosion redistribution following land use changes
Nutrient cycling Nitrogen dynamics, phosphate fluxes, nitrogen loads through base
flow, storm flow and underflow, biogeochemistry, nutrient and
metal dynamics

20
Pollutant transport Deposition, surface runoff, impacts on water quality, stream
temperatures, nutrient concentrations
Riparian habitat and
stream Plant canopy, surface plants and litter, shade loss, riparian
vegetation
buffers
Stream morphology Channel characteristics, habitat conditions, stream channel banks,
bank stability, sediment transport, in-channel debris torrents
Hydrology Stream peak flow, rainfall, land cover relationships to stream
hydrology, ground water flow, wet and dry seasons, surface and
subsurface hydrology, groundwater and surface water interaction
Water quality, pH, dissolved oxygen, nitrate, phosphate, turbidity,
Stream chemistry
benthic macro invertebrates

Influence of Disturbance on Watershed Processes


The concept of disturbance is so central to understanding ecosystem functioning that it is
worth providing a definition for the term at this point:
“Any relatively discrete event in time that disrupts ecosystem, community, or population
structure and changes resources, substrate availability, or the physical environment” (Society
of American Foresters, 1996).
Periodic large- and small-scale disturbance is critical to ecosystem functioning, resetting the
‘successional clock’ and preventing the vegetative community from maintaining a
homogeneous climax state. Under natural circumstances, disturbances (e.g., fire, disease,
landslides, and flooding) within an ecosystem occur with characteristic frequencies,
intensities, and extents. Thus, every ecosystem evolves with a particular disturbance regime.
Within a given ecosystem, the variability of size, intensity, and frequency of different
disturbance events creates a mosaic of vegetation at various successional stages.
Over a landscape scale, the disturbance-driven mosaic tends to remain in dynamic
equilibrium during a given climatic period. It is the diversity inherent in this mosaic that
provides diverse habitat. For example, many species of plants and animals are dependent on
early- to mid-successional stages (biological diversity commonly peaks at the mid-
successional stage). The availability of this habitat type limits the populations of numerous
species.
Vegetation, in turn, interacts with the disturbance regime. For instance, among plant
communities, the accumulation, distribution, and type of fuel vary greatly. These are major
factors in fire frequency and intensity, which in turn strongly affects the species composition
and structure of the plant community.
Erosion and Sediment Yield
Erosion is a natural process, made inevitable by gravity, wind, the weathering of rocks (i.e.,
soil formation), and the energy of flowing water. Erosion processes and rates are controlled
21
by climate, topography, soils, and vegetation. Forested landscapes generally undergo little or
no overland flow or surface erosion, with the organic litter on the forest floor sustaining
infiltration rates greater than rates of rainfall or snow melt. In contrast, in arid or semi-arid
landscapes with partially exposed soil, surface erosion may be the dominant erosional
process.
Erosion rates tend to be episodic and linked to disturbance and weather. Substantial surface
erosion occurs following the removal of vegetation with extreme rates occurring after severe
fire consumes the protective organic layer and exposes bare mineral soil. Mass-wasting (i.e.,
landslides, debris flows, etc.) is the result of the gradual accumulation of soil in unstable
locations, combined with a triggering mechanism, such as soil saturation, that activates the
event. Stream bank erosion, the process by which water loosens and wears away soil and
rock from the edge of a stream, generally occurs during high flow events.
Sediment, alternating between moving in brief pulses and being stored in channels or
floodplains, is a major watershed product naturally transported and discharged by stream
systems. In the same way that a given watershed produces a characteristic stream flow
regime, it also has a characteristic sediment budget over time. The budget, consisting of both
sediment quantity and quality (i.e., the distribution of particle sizes transported) is largely a
reflection of the climate, geology, topography, vegetation, and disturbance regime across the
watershed.
Land Use Effects on Watershed Processes
The effects of widespread land use tend to accumulate within watersheds, both over time and
in the downstream direction. Any land use altering one of the three basic watershed
components - soil, vegetation, or water - will affect watershed functioning. Land use (e.g.,
logging, grazing, farming and urbanization) generally alters vegetation, often intercepting and
diverting the movement of water. Land use may also directly affect the soil through
compaction. Road building, in addition to removing vegetation, exposing soils, and creating
impermeable surfaces, can drastically alter the routing of water through watersheds.
Numerous attempts to increase runoff by removing vegetation have had serious unintended
consequences such as greatly increased erosion, earlier, flashier runoff, and correspondingly
decreased base flows (i.e., more water when it is not desired and less water when it is in short
supply).
Reduced vegetation, soil compaction, soil exposure, and increased velocity of water
movement result in increased erosion. Erosional processes, once altered, often accelerate over
time: overland flow across exposed soils creates rills that rapidly develop into gullies; sheet
flow becomes channelized (expanding the drainage network), and more erosive. Expanded
drainage networks reduce soil water storage by capturing water at the soil surface (reducing
infiltration), and intercepting soil water (speeding the drainage of the soil mantle). Soil
erosion in excess of soil formation and compaction that lowers the ability of the soil to absorb
water combine to reduce the water storage capacity of the soil mantle. Severe erosion alters
both soil depth and quality, causing irreversible changes to the vegetation.

22
Quantity, quality, and timing of stream flow are the result of overall watershed processes. In
the absence of climate change, changes to these processes, and by association, to aquatic
habitat, reflect the cumulative effects of land use. A general effect of many land uses is to
reduce the resistance offered to water as it moves through the watershed, speeding runoff,
increasing peak flows and decreasing low flows. Examples of this include intensive timber
harvest, road building, grazing, and urbanization. An exception to this phenomenon occurs
when a significant portion of a watershed undergoes conversion from one plant community to
another that is more water-consumptive. An example of this is the conversion, through fire
suppression, from an open fire-tolerant forest stand to a densely stocked, closed canopy, fire-
intolerant forest stand. In this case, the entire range of flows produced by that land area might
decline.
Ecologically, land use represents a change to the disturbance regime of an ecosystem. Fire
may become much less frequent due to grazing, logging, and fire suppression. The magnitude
and frequency of flooding may change. The effects of droughts may become more severe due
to soil loss, soil compaction, and faster delivery of water to the stream system. Land sliding
may increase due to destabilization of slopes following logging and road building.
Agriculture and urbanization represent major disruptions of native plant communities and
ecosystems; additionally, irrigation and other water uses are inevitably associated with
alterations to stream flow and groundwater.
Aquatic and terrestrial ecosystems evolve within a natural range of disturbance frequency and
intensity. Each system has some resistance to change and some resilience in recovering from
disturbance. If the effects of human activities substantially differ from those of the natural
disturbance regime, the ecosystem will be substantially altered. Ecosystem degradation is the
result of imposing disturbances that are beyond the system’s ability to resist or recover from.
3.5 Watershed management practices
Watershed management is the judicious and scientific management of three basic natural
resources (soil, water and vegetation) and other resources available to its inhabitants for the
wellbeing of the people. The objective of managing a watershed should be clearly defined to
find a correct approach for its development. Watershed development and management should
be designed in such a way that the particular objective for which works are designed is
achieved. The objectives of a watershed management programme may be:
• Prevent further damage to land resources,
• Develop lands already damaged,
• Control productivity of land,
• Use available resources in the watershed for optimum production of
crops/fodder/fuel/timber/fruit,
• Check sedimentation of storage reservoir downstream
• Supply good quality water within or outside the watershed,
• Check flood hazard downstream,
• Exploit land resource to meet population pressure, and
• Recharge groundwater.
Selection of priority areas

23
A large watershed may have different types of sub-watersheds requiring different treatments.
It may not be possible to treat the entire area of watershed with land treatment methods. In
such cases priority area and priority of work should be selected on the basis of survey reports.
For determining the priority of the sub-watersheds for treatment against soil erosion and
sediment yield the preliminary investigation needed are:
a. Reconnaissance surveys,
b. Soil and land use surveys, and
c. Sediment observations.
The purpose of a reconnaissance survey is to get an idea of the relative erosion condition and
extent of damage of sub-watersheds. Soil and land use survey also gives an idea of soil
erosion and land fertility. The measurement of silt load coming from each sub-watershed will
give a clear picture of the erosion extent. Scientific interpretation of aerial photograph
provides good information regarding land, its location, soil texture, drainage characteristics
and vegetation etc. Aerial photographs can give information on the erosion conditions of the
sub-watersheds and can help in determining which watersheds are comparatively under
severe erosion.
Criteria for the determination of priority should be based on curative, preventive and
exploitative measures. The factors like people's participation, aesthetic value, accessibility
and sentimental value attached to watershed should be considered afterwards. The
quantification of these criteria should be based on the severity and frequency of damage
already caused to life and property in past, severity of impending danger and productive
potential of the land resources in the watershed.
The basic criteria which are considered for selection of a watershed for its management are:
• Prospects of solving the problem,
• Severity of problem,
• Potential for overall development,
• Availability of technology,
• Adoptability and people's participation, and
• Availability of infrastructure.
Steps in watershed management
After determining the priority of area based on main objective of the watershed management
programme, certain steps are followed to accomplish the works. These steps are:
i. Surveys and collection of information,
ii. Analysis of information and study of problems,
iii. Development of alternative solution of the problems,
iv. Selection of the best solution,
v. Preparation of work plan,
vi. Execution of works, and
vii. Protection and improvement of works.
The above steps can be divided in four phases of the watershed management programme

24
which are discussed below.
1. Recognition Phase
This involves steps i), ii), iii) and iv). Surveys are conducted to collect information on soil,
geology, land capability, topography, crops, forest, pasture and other vegetation, and socio-
economic condition of inhabitants. Long term climatic records are also collected on rainfall,
temperature, evaporation etc. This information is analyzed to describe the watershed,
ascertain problems both qualitatively and quantitatively and develop solutions to the
problems. Many alternative solutions are designed for individual problem and best one is
selected.
2. Preparation of work plans
Work plans are project proposals of watershed management. It includes description of the
watershed, its problems, details of improvement works, the estimated cost, economic
justification and the responsibilities of the participating agencies. A watershed management
work plan may have following contents:

a) Description of Watershed

It should contain sufficient detail to present clear picture of the watershed. The different
watershed characteristics included are as follows:
Location
Name of the watershed, physiographic region, latitude, longitudes.
Size and shape
Area in hectare or sq. km. shape, length width ratio.
Climate
Rainfall, its seasonal and monthly distribution, rainfall intensity, duration and its areal
distribution. Other climatic factors like mean and extreme air temperature, evaporation,
relative humidity, solar radiation and wind velocity and direction.
Geology
Parent rocks, fractures, faults, extent of outcrop, weathering and ground water recharge area.
Slopes
Mean slope, length of slope, proportion of the watershed in different slope group.
Physiography
Topography, elevation of different watershed areas, information about hills and mountains.
Soils
Major soil groups and their hydrologic groupings. Physical and chemical properties.
Surface drainage

25
Details of principal stream and its distributaries, nature of flow (perennial, intermittent,
spring-fed, ephemeral or seasonal), order of streams, stream density and stream length,
location of lakes and ponds.
Waters uses and needs
Source of water (surface or ground) use for domestic, irrigation, power generation,
recreation; adequacy and future needs.
Land use and cover condition
Existing land use and cover condition including cultivated land, forest land, pasture land,
waste land etc. Details of crops and its management practices, type of forest trees, its legal
status and management. Extent of pasture lands, their major classification, closer and grazing
incidence, and important grasses. Land capability classifications.
Economic condition
Occupations and economic conditions of the people, their dependence on watershed
resources, returns from crops, forest and pasture. Financial agencies and debt on farmers.
Social condition
Family size, social customs, societies etc.

b) Analysis of watershed problems

The problem faced in the watershed should be discussed in this section. Some major
problems of watershed are described below in detail.
Soil erosion
Extent of sheet, rill, gully and channel erosion. Effect of soil erosion on sediment yield and
crop production.
Flood
The amount and value of land improvements and other properties exposed to flood hazard.
Frequency and seasonal occurrence of flood. Land use limitation due to flooding.
Sediment damage
The effects of sedimentation on water supplies, reservoirs, drainage, channels etc.
Problems of water management
The needs of drainage, irrigation and agricultural and non-agricultural water supply.
Poor crop yield
Reduction in yields of crops and other vegetation in the watershed. Crop, fodder, fuel and
timber yield per ha per year.
Socio-economic problems
Income of farmers, social and economic constraints.

Other problems
Land slide, torrents, road erosion, mines etc.
26
c) Proposed management Programme

The basic principle of watershed development is to start planning, design and construction of
works from the outer boundary step wise coming to outlet point down below. This section
includes details of proposed treatment, its cost and expected yield. Different factors to be
described under this section are as follows.
Agricultural lands
Recommendations should be given for each capability class on engineering and agronomic
practices. The engineering measures with plans, designs, calculations, costs estimate should
be given with proper justification and specification. The engineering measures may include
oil and water conservation, irrigation and drainage, and flood protection measures. The
agronomic practices may include crop rotation according to soil class, seed and varieties,
cultural operations, use of manures, fertilizers and implements and insect and pest control.
Present crop yields and expected yields after development works should also be given.
Forest lands
Trees, shrubs and grasses under existing condition and recommended according to the land
and soil capability class should be given with details on its management practices and
availability of seeds and saplings. Agronomic and engineering soil conservation measures
should be given with plans, design and cost estimates. The factors relating to the ownership
of the forest should also be considered.
Pasture and grazing lands
Existing conditions of pasture and grazing lands and package of improved practices should be
described. It includes species, its raising, grazing system, fodder harvesting and hay making.
Other programmes
Complete plans, design and estimate, and costs of individual works on dams, tanks and
structures for water storage, roads, fences and bundings should be described in detail.

d) Phasing of development works

Phasing of each work should be given with time period, expenditure and agency responsible.
Annual phasing of works also required for arranging finance and getting sanctions.

e) Benefits and costs

Estimated returns from individual works and overall watersheds works should be compared
with costs. A benefit-cost analysis should be carried out on the basis of present worth or
equivalent annual worth.
3. Execution Phase

27
Once the problems are recognized and a plan of works is prepared the treatment measures are
applied under this plan. The treatment measures include soil and water conservation works,
improved practices for crop, forest .and pasture management and other works. It is a very
important phase for managing man, money and material in the watershed. It requires
maximum coordination among the different departments/agencies involved in the watershed
management work.
4. Maintenance and improvement phase
This phase takes care of the general health of the treated watershed and ensures normal
working. The protection measures against all factors which might cause deterioration in
watershed are applied and normal functioning of the watershed is maintained. It also involves
overall improvement in the watershed and all lands are covered. Due attention is given
towards improvement of production from agricultural, forest and pasture lands and living of
people. It may also include health, family planning, education, cattle, poultry etc.
Integrated Approach
Watershed management, based on integrated land use, is a multi-disciplinary and multi-
departmental programmes. It involves the agencies dealing with agriculture, forest, animal
husbandry, fisheries, soil conservation, road, dams, irrigation, drainage and buildings. In
addition to these infrastructural facilities like communication, marketing, banking,
cooperatives, panchayats, social and voluntary organizations are needed. The agencies
involved in extension services and supply of inputs have to help and coordinate watershed
development works. An active and timely cooperation and coordination among above
agencies/departments are essential for planning, execution and maintenance of watershed
development programme. A department or agency cannot accomplish a work of watershed
management independently, proper coordination and cooperation are needed from other
related department/agency. For solving any problem in the watershed, there should be an
integrated approach from related individual/department/agency.
Requirements of successful watershed management programme
For successful implementation of a watershed management programme, on large scale there
are three main pre-requisites, i.e. i) availability of technology package, ii) availability of
appropriate infrastructure to implement the programme, and iii) presence of congenial social
and political environment. Technological packages have been developed for many areas after
testing and evaluation of research recommendations under the Operational Research Projects.
The available infrastructure can be simplified into money, men and material. The money
required for watershed management should be adequate but the success of the programme
depends on well trained man power and materials. The planning of manpower and materials
should be done simultaneously with programme planning. The watershed development
requires general awareness and dedication in the society. The political and social
environment should be favorable for watershed management programme.

28
4 PROBLEM DEFINITION AND SCOPING
4.1 Introduction
Issues and problems (issues in contrast to problems: issue = something to consider, a possible
future problem, problem = a present negative situation)  in a watershed context are site-
specific, complex, often interlinked and may be related to ecological, socio-economic,
technical and institutional issues. Awareness of these issues and problems may come from
agency representatives, interested professionals such as consulting engineers or planners, or
from the Government's regulatory directives. However, such perceptions will quite often
come from the community itself, whose members are directly affected by the conditions and
changes of natural resources prevailing in a given watershed.
The participatory identification and thorough analysis of issues and problems in a watershed
context is an important initial element of planning as it will result in a clear understanding of
the issues and problems. Based upon this, scoping can prioritize the key issues and problems
that need to be addressed most urgently, and objectives can be defined that reflect a vision of
what the future conditions within the watershed should look like. It will also enable the
subsequent definition of appropriate strategies to successfully address these issues and
problems, as well as maintain desirable conditions and change currently undesirable
conditions in a given watershed.
As preparatory step for problem analysis and scoping, reconnaissance and base line survey
activities are required to establish necessary information about the economic, social and
environmental settings within the watershed. We will use the term “watershed baseline
inventory” for these survey activities, and “watershed profile” to describe their results.
We will not explicitly examine the watershed baseline inventory. However, the chapter does
introduce the necessary steps and methods to analyze issues and problems, and to be able to
prioritize them and select sub-watershed areas through scoping, and thereby define broad
targets to guide the further planning process. It is important to note here that these steps may
be iterated, and may also iterate with the collection of base line information about the
watershed through the watershed inventory. As for instance, the problem analysis or also the
scoping process may reveal significant gaps in information, which can only be filled through
the sourcing of additional in-depth information.
Once problem identification and analysis have been conducted, there still remains the
challenge of agreeing on the priorities among these issues and problems. The stakeholders
need to determine which uses of water resources and which watershed areas urgently require
interventions, and which of them do not. This process is called scoping. The priority uses and
areas identified during scoping require standards and targets to govern their management, or
even concrete management interventions, while other uses and areas of lower importance will
require them only later or maybe even not at all. This also implies the identification of target
groups (resource users) who will be expected to implement standards or otherwise be
involved in interventions.

29
The first step in watershed planning is to define the scope of your watershed management
plan. A watershed plan must have a clearly defined geographic area within which the
management plan will be implemented. An effective watershed plan must also have clearly
defined environmental impairments so that management efforts can be effectively targeted.
4.2 Watershed assessment methods
4.2.1 Introduction
Watershed assessment is a necessary component of a monitoring program in order to
determine what degraded or impaired areas may exist in the watershed and why. Several
characteristics of a watershed are taken into consideration during the assessment process
including land use, land cover, and hydrology. Land use and cover is considered in a
historical as well as current perspective to determine the types of activities that have occurred
in the watershed and their potential as sources of pollution. It is important to consider the
natural and cultural resources of the watershed as well as the human activities. This
information may be obtained from various sources including topographic maps to determine
drainage area and land features as well as land use data.
A reconnaissance and assessment of watershed character is necessary to:
• Assess watershed conditions to determine the causes and nature of impairment
• Determine feasibility of using restoration or other management options to meet
objectives
In some cases, ecological restoration is the most effective response to impairment; in other
cases, restoration may be one among many candidate tools for achieving objectives. To
determine the appropriate actions, it is necessary to collect, compile, analyze, and interpret
environmental data rapidly to facilitate management decisions and resultant options for
preservation and control or mitigation of impairment. The technical note of watershed
assessment considers watershed and reach reconnaissance techniques that possess the
following principal elements:
 Cost-effective
 Facilitate comparisons among sites
 Quick, yet scientifically valid
 Easily presented to the public
 Environmentally-benign procedures
Assessing a watershed to understand its current condition, and how it got there, is usually the
first step taken in developing a strategy toward improving and protecting the watershed’s
condition. A relevant watershed assessment addresses the sources of watershed impacts rather
than just their symptoms, which key to achieving effective watershed protection and
restoration.
An assessment is far more than an encyclopedic collection of information about the
watershed it must analyze why the watershed is in its current condition. Watershed
assessments can be relatively comprehensive or be focused on several specific issues.
However, watershed practitioners must make a choice between a broad assessment and a

30
focused assessment after thoroughly discussing the advantages and disadvantages of each in
light of the watershed’s needs and the purpose of the assessment.
A watershed assessment is: “a process for analyzing a watershed's current condition and the
likely causes of these conditions.”
A watershed assessment report is: “a report documenting the findings of the watershed
assessment process.”
Within a watershed common zones, often used for management purposes, are: 1) upland (land
above the zone inundated by floods or above the transition between riparian and terrestrial
vegetation), 2) riparian (vegetated area between the water body edge and the upland area), and
3) water body (any stream, river, abandoned channel, pond, lake, wetlands, estuary, or ocean).
The term “watershed” is not synonymous with terms such as “stream” or “riparian corridor” or
another single feature of the watershed.
4.2.2 Watershed assessment: what it is and what it is not
Coming to a common understanding of what watershed assessment is and what it is not is
important for the users. “Watershed assessment” definitions include the following:
• The analysis of watershed information to draw conclusions concerning the conditions in
the watershed. (Nehalem River Watershed Assessment, Washington, 1999).
• A process for evaluating how well a watershed is working. (Oregon Watershed
Assessment Manual, WPN, 1999)
• A process that characterizes current watershed conditions at a coarse scale using an
interdisciplinary approach to collect and analyze information. (NCWAP & CDF 2002)
• The translation of scientific data into policy-relevant information that is suitable for
supporting decision making and action at the watershed level. (Watershed Academy,
U.S. EPA).
Despite their differences, what is common to each definition is a process composed of actions-
analysis, integration, translation that leads to the interpretation of information about the
watershed’s current condition. It is critical that the watershed assessment effort lead to a better
understanding of watershed processes and conditions and why the watershed is in that
condition. That way, the assessment can serve as a compass to help direct further actions.
Your assessment should move beyond a simple description of what a watershed looks like, or
what historical activities took place in the watershed. While those are some of the building
blocks of an assessment, your assessment must connect past and current human activities and
land uses (causes) to watershed processes and current condition (effects). (Watershed
processes refer to the natural processes, such as hydrologic and nutrient cycles, that influence
the waterways’ conditions). With an understanding of dominant watershed processes and
potential causes of watershed condition, watershed practitioners can propose solutions to
problems. Without this understanding, proposed solutions may address only the symptoms. A
successful watershed assessment leads to the implementation of actions that benefit
watershed processes and conditions the ultimate “performance measure”.
What an assessment is

31
• The scientific interpretation of watershed information and data, leading to conclusions
about watershed condition
• An objective problem-solving tool that identifies the potential causes of problems
• An objective problem-solving tool that identifies the potential causes of problems
• A tool to help identify available data or information gaps
• Analysis and findings that can be used to develop appropriate actions
• A component of a watershed management package that leads to planning,
implementation, evaluation, and additional monitoring
• A product that is useful for its audience
What an assessment is not
• Monitoring and data collection only
• A list of data only
• A consolidation or summary of existing information only
• Historical conditions or “baseline” conditions only
• An identification of symptoms of problems only
• A plan
• An endpoint
The following steps are basic process for conducting a watershed assessment, planning and
evaluation
1) Organize the assessment team
2) Define the purpose and develop a plan for the assessment
3) Collect data and information
4) Analyze the data
5) Integrate and report the data to inform decision-making
Step 1: Organize the assessment team
Your organization has decided that it wants to do a watershed-scale assessment. In some
cases, there are conditions and impacts in your watershed that have raised concerns. You are
interested in finding out why these impacts have occurred. In other cases, your watershed
may be relatively pristine and you want to maintain its character. You are interested in
assessing conditions and analyzing potential sources of degradation so you can prevent or
minimize future problems.
Assemble the Team and Committees
No one has all the expertise required to do an assessment, not consultants, agencies,
academics, or watershed groups. Accordingly, your assessment team should include people
with a wide variety of expertise and interests.
Create a Realistic Schedule
32
It’s important to be realistic about how much time it takes to perform a watershed assessment,
but estimating time required can be challenging. Experience has shown that simpler
assessments performed in-house with sufficient expertise and information may take four to
eight months, more complicated or comprehensive assessments or assessments where the
process is not under tight scheduling control can take as long as 36 months. Use milestones to
stay on track. Here are some sample milestones (adapted from Coastal Conservancy 2001):
 Start-up
 Initial project team meeting (define approach)
 Public meeting #1 (review issues, concerns)
 Technical Advisory Committee (TAC) meeting #1 (review strategy)
 Begin assessment
 TAC meeting #2 (mid-progress review)
 Draft assessment complete
 Review results—TAC and Public Advisory Committee
 Release revised draft to public
 Revise and deliver final assessment
Involve the Community
Those who will be making decisions using information contained in the assessment should be
included, consulted, or at least considered when designing an assessment. From start to
finish, the assessment should make clear how and why various steps were taken. This
approach has the benefit of getting all-important buy-in stakeholders and decision-makers are
more likely to trust the assessment’s conclusions if they understand the reasons various
approaches were taken or they were involved in gathering data and information for each step.
Steps
 Assemble the assessment team and committees.
 Appoint a coordinator and seek contractors, if necessary.
 Encourage community participation through public meetings, the media, and
outreach to other relevant local organizations (e.g., Farm Bureau, resource conservations
districts, etc.).
Once you have your watershed team assembled, you can actually begin the work. One way of
organizing a watershed assessment is to break it into four main parts:
 Defining the problem and planning the assessment
 Collecting information and data
 Interpreting results: data analysis and synthesis
 Preparing the report
33
Step 2: Define the purpose and develop a plan for the assessment
The first formal phase of a watershed assessment consists of clearly identifying the issues of
concern, identifying the purpose of the assessment, developing a conceptual diagram of the key
components of the watershed, and developing a plan for carrying out the assessment.
I. Formulate the relevant questions and goals
Watershed assessments may be motivated by one or more influences:
 to evaluate watershed conditions from a neutral perspective, i.e., with no prior
assumptions;
 to address identified watershed issues or problems;
 to meet a particular purpose, e.g., identify conditions that need to be improved in
order to increase drinking water quality;
 To meet a particular goal, such as educating the public about natural and human
features of the entire ecosystem and assist in planning and decision-making.
For many assessments, one or more issue-based questions usually drive the process. The
question may be as generic and general as, “What is the condition of our watershed, and why is
it that way?” More specific questions might be along the lines of, “Why did the salmon stop
spawning in our stream? Why did such a big flood come from such a small storm? Why can’t
we drink the stream water anymore?” or “How can we protect our pristine watershed from the
degradation we see in neighboring watersheds?” Questions based on observations and
community concerns will direct the watershed assessment, which will in turn provide the basis
for addressing the important issues.
Issues to consider
 If there are no fundamental questions or concerns guiding a watershed assessment, you may
wish to make explicit the perceived need for the assessment.
 The questions should be stated clearly enough to capture the prevailing concerns that led to
wanting or needing a watershed assessment.
 Clearly write out the questions and/or issues use them to guide future data collection and
analysis.
II. State the purpose of the assessment
Watershed assessors should develop a clear statement of purpose. A “fuzzy,” or implied
purpose statement that never gets clarified, or an absent purpose statement can lead to bigger
and bigger problems (such as getting off target, or creating misunderstandings due to
different expectations of the product) as the assessment process continues.
It is also important to clearly identify who really wants the watershed assessment, and why
they want it. Otherwise, misunderstandings can occur. For example, the impetus may come
from the local level from a cooperative group (e.g., a watershed council), a local agency (e.g.,

34
a resource conservation or water district), or other private or public stakeholders for a variety
of reasons.
The comprehensive approach assesses the conditions of all processes and features in a
watershed. The advantages are that this broad approach gives an overview of the watershed’s
condition, may expose previously unknown problems in the watershed, and may identify the
interconnections between various problems or issues. On the other hand, “comprehensive”
may sound desirable, but a focused product may prove more useful.
In the focused approach, the assessment process chooses the most critical issues in the
watershed, and then focuses the assessment effort on these. The benefit to this approach is
that it makes the assessment potentially more useful for future decision making about specific
problems or areas. Groups identify upfront the issues of all those possible that most need to
be addressed because the assessment cannot address all issues in depth. The watershed’s
problem(s) drive the assessment. The risks of this approach are that the focus can become too
narrow, miss critical issues, and overlook connections among problems/issues, resulting in a
failure to correctly identify the root cause of problems
State what the watershed assessment will be used for
Assessments generally serve to inform certain functions:
• General watershed management planning with multiple purposes
• Regulatory concerns
• Restoration or enhancement planning
• Monitoring program development
• Management of areas at risk and practices resulting in risk
• Land use activities
III. Define the geographic boundaries of the watershed
Establishing the boundaries of your watershed assessment or the spatial limits of the area to
be analyzed is a critical early step. The only watersheds defined by nature are those with a
low point at the ocean or a closed-basin lake. All others (including those contained within a
“naturally-defined” watershed) are defined by a human choice of the lowest point. Agreeing
on the assessment area at the outset so that everyone knows exactly what piece of ground is
under discussion can head off many problems and arguments.
IV. Develop a basic overview of the watershed (past and present)
When you defined the boundaries of the watershed, you identified the spatial scale of the
assessment. You also need to define the temporal scale. How far into the past and into the
future do you plan to collect data? Ideally, having data that extends over a period of many
years is best because you can get the clearest picture of the changes that have occurred. Also,
year-to-year variability is inevitable. Having data that extends over many years will permit
you to distinguish between natural variability and a real alteration or change.

35
V. Identify the watershed processes and/or valued ecosystem components on which
you will focus
“Watershed processes” refers to the natural physical, chemical, or biological processes that
interact to form the terrestrial and aquatic ecosystems (the water cycle, for example). “Valued
ecosystem components” refers to the things within the watershed that stakeholder’s value,
such as fish, clean water, trees, or open space. In other words, these components may be
structural (population of a certain species of fish) or functional (such as the return frequency
of fire). It is not necessary in every case to directly measure a component or process.
Frequently, surrogates or indicators can be used to get an idea of the condition of the selected
watershed component. Looking at a simple example, you might not be able to directly
measure the population of the split tail fish, but by measuring different habitat and water
quality characteristics (e.g., water temperature), you can get a good idea of whether or not
this fish could survive in these conditions.
There are many possible watershed processes and attributes. These include, for example, the
distribution of benthic macro-invertebrate communities, drinkable water, and presence of a
species of fish or a plant that is important to the stakeholders, or, more generally, the overall
riparian corridor or upland areas.
Some criteria that are often used to select the watershed processes or components that could
be the focus of the assessment are:
• Importance to the health and sustainability of the watershed;
• Related to the assessment’s purposes;
• Sensitive to those activities or factors you suspect might be causing changes in the
watershed;
• Have societal value; in other words, are important to the community, region, or
state.
Additional criteria might include watershed processes for which the natural variability is
known; attributes required by a regulation; and the availability of data, models, or knowledge
about the particular endpoint.
VI. Develop a conceptual model of the relationship between human activities,
processes and conditions, and potential impacts on watershed condition.
A conceptual model is a graphical representation of potential relationships among the
watershed’s components and processes. Once you have identified the watershed processes
you are most interested in, you will need to think about how they are impacted by changes in
regional and watershed processes and the stress, or impacts that may result from human
activities. The relationship between human activities, watershed processes, potential impacts
or sources of stress and the effects on ecosystem function are depicted in the conceptual
model. Watershed assessments typically focus on those alterations that are human-induced
since these are the ones we can influence.

36
Developing a conceptual model is an iterative process. The model you develop at the
beginning of your effort may look different from the one you finally adopt because you will
modify it as your knowledge of the conditions and processes within the watershed grows. The
knowledge needed to develop a conceptual model will come from familiarity with the
particulars of your watershed as well as from general knowledge about watershed science.
Conceptual models can be developed in a variety of ways. These models will help you
understand the possible relationships that are important to consider when you collect and/or
analyze data. The diagram in Figure 2 is an example of a simple conceptual model of an
urbanizing watershed. In this example, the assessors were interested in the health of the
benthic macro-invertebrate community in their local creek. They were concerned that runoff
from agricultural fields and new development was impairing the diversity and viability of the
insects.

Figure 2: Simple conceptual model of an urbanizing watershed

The more complex model shown in Figure 3 is a refinement that reflects more detail as the
group’s understanding of the watershed components and processes increased.
Clearly, the knowledge required to draw accurate conceptual models can be significant. That
is why having a team of people with varied technical backgrounds is very helpful. These are
just two examples of construct a conceptual model. Regardless of how yours looks, the key
point is that the conceptual model diagrams should identify hypothesized relationships
between human activity, changed conditions or processes in the watershed, and the potential
effects of these changed conditions on the selected watershed processes and/or components.
These relationships can serve as the basis for data collection and analysis.

37
Figure 3: More complex conceptual model for an urbanizing watershed

Now that you have figured out what your assessment questions are, developing and
implementing an “analysis plan” for collecting and analyzing information and data is
typically the next step. Data collection and analysis constitutes the heart of the watershed
assessment. The conceptual model or diagram you constructed can serve as a guide.
Accordingly, as you prepare for the analysis phase of your assessment, you should identify
the data and information that must be gathered and outline the process for organizing and
analyzing this material.
The watershed assessment focuses in part on the potential harmful effects of human activities
on watershed components and functions. These effects occur when human activities cause
changes in the watershed’s physical, chemical, or biological characteristics and processes.
• Physical changes include water temperature and flow rate, generation and transport
of sediment, stream channel shape and connectivity with the floodplain, erosion and
incision of the stream bank, and any other physical characteristic that makes up the
habitat on which the watershed processes being evaluated depend.
• Chemical changes include the introduction of pesticides, excess nutrients, oil/grease,
effluent from industry, or other contaminant to the targeted habitat
• Biological alterations that might be associated with harm could include invasive
species and pathogens.
Step 3: Collect data and information
I. Determine the kind of data you need to collect
The data you collect should correspond in type and substance to the questions rose at the
beginning of assessment planning and refined in the conceptual model. Types of data include
spatial or geographic data for understanding things taking place on the landscape (such as
land use) and water quality or quantity data for understanding a waterway.
The conceptual model can serve as a guide to what data you need to collect. The following
list includes key classes of information that typically are useful.

38
• Data on human activities and land uses – the location, type, intensity, areal extent
(acreage), and proximity to or linkage to the waterways (such as via storm drains)
• Data on the physical, chemical, and biological properties and potential sources of impacts
in the watershed – in-stream and riparian habitat characteristics, water quality data,
animal/plant population abundance and diversity, etc.
• Data on alterations in watershed processes – changes in the hydrological cycle, nutrient
cycling, etc., particularly as they are affected by past and current water and land use and
by climate change.
• Data on potential effects of potential impacts on watershed functions
These and other data can come in a variety of forms: digital and non-digital spatial data,
quantitative or qualitative data, and anecdotal information. All these data types can be useful
in a watershed assessment. Specific types of data include quantitative water quality data,
geomorphological surveys, biological surveys, maps, and other similar data. These data may
be presented in various formats and at varying levels of detail of analysis. For example, data
might be presented in spreadsheets, tables, and graphs; as spatial data within or separate from
a computer-based geographic information system; in internal agency memoranda; in field
surveys, in narrative or historic descriptions of a place and past processes and events, such as
floods, landslides, or contaminant spills; and products of computer models developed to
illustrate specific processes (e.g., storm-water runoff). Consider collecting any type of data
that would be useful for the goals of your assessment.
Anecdotal information may be one of the most difficult data types to record and store, but
may also provide knowledge about watershed processes that might otherwise be lacking.
Common types of anecdotal data include:
1) The extent of salmon runs in rivers now lacking these runs due to dams or other
barriers,
2) Increasing turbidity of streams and rivers over time due to upstream activities,
3) Encroachment of roads and human structures into previously undeveloped landscapes,
4) Growth of nuisance vegetation (e.g., benthic algae, riparian weeds, or invasive exotic
weeds), or
5) Increased rate of flooding in river valleys due to landscape modification.
Although not the same as quantitative information, anecdotal data can provide very useful
information and help you develop hypotheses about historic conditions in the watershed and
the effects of human activities
II. Identify sources and collect existing watershed data and information
The following section references sources of watershed information and data that will be
available:
 Waterway data
 Hydrology and Flooding Data
 Riparian Vegetation and Wetlands Data
 Physical Watershed, Channel, and Habitat Conditions Data
III. Identify sources and collect existing spatial data about the watershed
Landscape data are often collected for areas. They may also have been collected initially at
individual survey sites (e.g., soil or vegetation) and then subsequently generalized to areas.

39
Currently, most contemporary data about a watershed landscape is collected with a
geographic reference point. In contrast, historic data may be very valuable, but lack easily
usable or identifiable reference points.
Non-digital Spatial Data
Digital Spatial Data and GIS
IV. Develop a system for archiving and managing your data
As data are collected, they should be organized in a manner that suits the questions being
asked and the users’ needs. Because watershed assessment usually involves the collection of
several different types of data (e.g., maps, water quality, and field surveys), consider
developing file organizational systems for each type of data that conform to a single standard
for categories (e.g., wildlife habitat, water quality, land use). One way to keep track of
information collected is to make a database of the category types. If you will be collecting
data for aquatic and terrestrial systems and of various different types (i.e., from text to digital
spatial data), then keeping track of the types of data and the areas they cover will help in both
organizing the data and describing how much of the watershed they cover.
For most watershed assessments of modest scale, you don’t need to become a database
expert, but you should learn enough to choose an adequate structure for your data needs.
Existing environmental data will be stored in some sort of file structure or database, with the
particular details dependent on the type of data, the agency archiving the data, and the needs
of data users.
V. Identify data gaps and collect new data, when needed
As you collect and organize your data, you will quickly identify important areas of concern
for which you have no or very little data. For example, you might be concerned with
alterations in the hydrological cycle in your watershed, but don’t have any data or
information on stream morphology (% pools and their size, % fine sediment, etc.) except for
anecdotal information. If in-stream habitat is an important factor in your assessment and you
are not in a position to collect this data, you will have to identify it as a data gap in your
report and consider the uncertainty that results when you analyze the data you have. It will be
a limitation in your assessment – but there are limitations on just about everything anyone
does so this may not be a fatal flaw. You might identify the lack of data on stream
morphology as a priority when future funds become available and explain how this
information might help provide a more complete picture of watershed conditions.
Alternatively, if you have the resources and time to fill the identified data gaps, the
uncertainty of the assessment can be minimized with the addition of the new information. If
your budget is limited, many times there are less sophisticated methods to collect the same
information; methods that volunteers or high school students can learn with a short
orientation. Typically these methods won’t be highly quantitative, but at least they can
provide you with a first approximation of the condition about which you have no data.
Frequently, watershed groups will partner with a local community college or university to
collect selected types of data on their watershed. In the best-case scenario, you will have
funds available to collect the information you need to fill in the data gaps.
Step 4 Analyze the data

40
This section suggests ways you can move from the raw data you have collected to
interpreting its meaning and importance. You might be data rich, but information poor staring
at a bunch of numbers that do not yet tell a story. This will assist you in making your
assessment more complete and accurate. In moving from raw data to integration and
interpretation, you may encounter a few stumbling blocks along the way. Suggestions to
overcome these problems can be found in this section.
I. Summarize and explore the data
Before beginning any formal statistical analysis, you should explore the data informally. This
can be done with descriptive statistics. Descriptive statistics refers to simple calculations that
can be done on an Excel spreadsheet and include calculating the mean or average value,
calculating the standard deviation (or range of variation), and making a frequency distribution
if you have sufficient data points. For example, if you’ve collected water temperature data
once a month for three years, you might decide to summarize the data for each month, based
on the value you collected over the three years, by calculating the mean and standard
deviation. You can then construct a graph or table that reflects the average temperature each
month.
A frequency distribution plot is another way to look at data variability. If you collected data
on temperature from 15 sites in the watershed in the month of September, you might plot the
data to see how similar or different the sites are.

Figure 4: Frequency distribution of data

Plotting the data in a frequency distribution (see Figure 4) gives you a visual picture of the
variability in temperature throughout the stream. It helps give more meaning to the average.
Overall, descriptive statistics give you a better feel for the data. These simple statistics are
sometimes all that is needed for the watershed assessment, especially if you have a small
dataset
II. Perform/Decide if statistical analyses are needed or possible with the data
available
Once you have summarized your data, you will need to determine whether it would be useful
to perform a statistical analysis to identify significant changes over time or between different
places within the watershed. Here are some questions to consider to get an overall feel for the
data.

41
• Is the data of sufficient quality to use? Do the data meet appropriate official standards and
practices for collection? Are data collection methods documented adequately so that you
can assess their quality?
• Are the gathered data and information useful for your needs?
• Do all the potential users and detractors of the watershed assessment accept the raw data?
• Do all the stakeholders support the choice of analyses?
• Are you thinking in ranges rather than single values for the data?
• Are you making comparisons to natural variability, which requires determining or
estimating baseline and reference conditions?
• What statistical tests, if any, do you plan to use? Some screening level assessments do
not necessarily require statistics. Also, if the datasets you have collected are limited in scale
(either temporally or spatially) then they might not be suitable for statistical analysis.
You might want to analyze your data using more complex, multivariate methods. These
methods permit you to estimate which factors contribute the most to minimizing variability in
the results. In most cases, those factors that reduce variability in the data are usually the most
important regarding meaningful relationship. Principal components analysis is one method to
determine, for example, which stressor out of six might contribute the most toward the
change in habitat that you might have observed.
Another way to approach data analysis is to do spatial analysis or time series analysis. An
example of an analysis over a spatial scale is the measurement of extent of development (e.g.,
human population or parcel density) in watershed areas that erode more rapidly than other
areas. An example of analysis over a temporal scale is determining whether changes in water
temperature over time are meaningful (or whether they just reflect natural variability).
The methods used to analyze things that change over space are different from those used for
things that change over time. There is an extensive technical literature on how to measure
each of these types of changes, depending on what needs to be measured (e.g., analysis of
trends over time). Two cautionary notes are that most analyses involve assumptions about the
nature of the process being analyzed and that sometimes analysts have employed
inappropriate tools, so copying an approach used elsewhere should be done with caution. In
general, it is wise to consult with someone knowledgeable in statistics to get an informed
opinion and recommendations.
Geographic information systems (GIS) were created to allow calculations for specific places
on the earth. If you have a GIS software program, you can carry out these calculations, too.
Examples of common straightforward analyses are densities of things within a certain area of
the landscape (e.g., abandoned mine density in a sub- watershed), intersection of lines of
different types (e.g., roads crossing streams), and summarizing data for an area (e.g., the
number of people in a watershed).
III. Compare your data to standards, historical, and/or reference conditions

42
Another aspect of analyzing your data is comparing it to standards that are recognized as
supporting the normal functions of biota or watershed processes that you are evaluating. To
objectively do this, compare your data either to that from similar watersheds that are widely
considered to have well-functioning processes and good conditions OR to values for habitat
conditions and water quality standards that are known to be protective for the watershed
processes on which your assessment focuses. One critical issue for this analysis is that for
many processes the standards for comparison will vary by bioregion and by habitat type. So
don’t expect one statewide standard to be available or to fit your needs.
The following information might be useful in identifying sources of information for making
these comparisons.
• Water Quality Standards that Support Aquatic Life
• Sediment Quality Standards that Support Aquatic Life
• Habitat Conditions
Step 5: Integrate and report the data to inform decision-making
Once you have collected all the data needed or available to answer your watershed
assessment questions, you face the challenging step of incorporating or integrating all the
information into a common analysis. Information integration here means combining or
linking information about various watershed processes and attributes in a way that leads to
conclusions about overall watershed condition and the possible causes. You could integrate
information for particular processes, like the movement of sediment from hill-slopes through
waterways until it is deposited and the impacts of that transport and fate, for example. You
could also combine multiple processes and potential impacts in a system using indicators for
potential impacts (e.g., land use), system stressors (e.g., water temperature), and impacts
(e.g., aquatic biota). Without linking individual processes (or separate disciplines or
specialties), watershed assessments may fail to identify potential causes of the watershed’s
condition and important linkages among watershed processes.
In this section, we describe a variety of ways that you can carry out this step, depending on
your needs and available resources. There is no single ‘correct’ way to do this. We give
several examples of approaches that scientists and watershed partnerships have tried. None of
them is necessarily right or always usable; they are listed here to inform you of the range of
choices. The methods range from relatively simple conceptual tools to modeling tools.
The relative condition of watersheds and waterways can be expressed in a variety of ways,
but it is commonly measured using such indicators as drinking water standards, aquatic
community composition, terrestrial and riparian vegetation condition, and constraints on the
free flow of water. A majority of watershed or waterway monitoring and restoration projects
are based upon definitions of “health” that are either explicit (e.g., water quality standards) or
implicit (often expressed as deviation from “historical condition”). Any risk or condition
assessment scheme designed to support monitoring or restoration programs should make
these watershed health definitions explicit so that stakeholders understand and support the
relevance of the findings or products of the assessment activities. Making these overall

43
watershed assessments will require the development of a scheme for integrating the
information.
There are many possible ways to integrate information, from qualitative to highly
quantitative, from informal to formal. Many watershed partnerships have a group of experts
from different disciplines evaluate information and form professional opinions about
watershed condition(s) and the potentially interrelated causes of those conditions. Other
watershed assessments rely on computer modeling for most of information processing and
then base conclusions on the products of these models. Some assessment programs develop
models that return evaluations of watershed condition as the product.
Models are often helpful in this process. When you developed a picture, or conceptual
diagram, of your watershed’s processes and influences, you were modeling, even if the
picture was only in your head. A model in watershed or environmental assessment is a scaled
representation of a system, just as a model boat is a scaled model of a real boat. The term
“model” covers a lot of conceptual and computational territory. You could model using only
mental processes, or you could rely on a physical model intended to represent a system, such
as a watershed.
A model is: A representation of a system; Based on understanding the types and magnitudes
of relationships; Done mentally, visually, or with computers; An aid for evaluation and
decision-making; Dependent on the quality of inputs
A model is not: A replacement for understanding a system; Independent of experts; A
substitute for good science and field work; The answer
With this perspective in mind, there are a number of approaches you can take to analyze your
data and understand the cause-and-effect relationships at work in your watershed. They range
from mental team integration to simple and complex mathematical models.
Regardless of the model or approach used, the overall goal is to identify the link between the
adverse effects and their causes. Being able to attribute a cause to an effect is one of the
major values of doing an assessment. It also can be quite difficult. As noted above, historical,
hidden, or multiple factors can be involved as causes of a problem. Be careful in making
assumptions about cause and effect, even when they might seem obvious: stream bank
erosion caused most of the sedimentation, housing development caused more frequent
flooding, or log jams blocked fish passage. Your data might show instead that roads caused
most of the sedimentation, channel aggradation (from sedimentation due to multiple causes)
increased flooding, and culverts blocked fish passage much more often than log jams.
Option 1: Team Mental Integration: Weighing the Evidence
Most watershed assessments involve convening a team of experts from several disciplines to
discuss the data collected and conclusions reached. The team mental integration method is
really nothing more than the assessment team and appropriate experts systematically
reviewing the data and, using their best professional judgment, assessing the impacts of
various alterations in the watershed on the ecological endpoints on which the assessment
focuses. In many watersheds, a collection of true experts about the watershed may provide
more detailed and accurate knowledge about influential processes than the best computer
44
model. This may be partly due to the absence of adequate data, partly due to the lack of a
model that truly represents the system, and partly because expert knowledge is still pretty
good compared to modeling. Watershed processes are complex, and all models contain
simplifying assumptions, some of which may preclude investigation of relevant issues.
On the other hand, the team mental integration approach has certain limitations. There is not
a single, widely-accepted approach for evaluating the weight of the evidence for an
assessment. Also, it may be difficult to ascertain whether team members have sufficient
knowledge to thoughtfully interpret the data. If your team does not have the right
qualifications, the insight gained from integrating their knowledge and information will be
limited. Competency is best measured by assessing the amount of formal training in one or
more scientific disciplines, field experience, the amount of watersheds like it, and the ability
to see watershed functioning from more than one perspective.
The suggestions in the following list address some of the potential benefits and pitfalls of the
expert team integration approach:
 Record whatever approach you use in a way that will allow a reader of your
assessment, or a future assessor, to understand exactly what you did. This means
describing both the details of the data considered and the analyses chosen and
rejected, as well as providing a summary of the approach your team took.
 The composition of your team determines the quality of your assessment. Include
team members’ qualifications, experience, and training as part of the assessment so
readers can assess for themselves how much confidence to put in the conclusions
drawn.
 Comparing professional judgment can be done in various ways, with the most
common (and possibly easiest) being to turn each set of information into rank values,
using the criteria for establishing cause-and-effect relationships.
 Because you will rarely get a group of experts together again to discuss your
watershed, take advantage of the opportunity and make sure they stretch their brains.
Encourage them to think about novel ways that data and knowledge about individual
processes can be brought together. Record the full spectrum of information, from
speculation with little data to sturdy conclusions based on a lot of data, analysis, and
expertise.
 Find ways to express professional judgment graphically so people can see what the
experts are thinking. This will help make your analysis understandable to a wider
audience.
 Promote diversity in your team by including members from a wide range of
disciplines, backgrounds, ages, and organizational origins. This is bound to lead to
critical questions, a range of approaches, and interesting discussions.
Option 2: Use Statistics
You might wonder how statistics relates to identifying cause-and-effect relationships.
Statistics can help identify associations, the magnitude of differences and other patterns.

45
However, statistics alone cannot determine causation from observational data. The correlation
between two factors does not necessarily mean that one caused the other. Looking for
significant correlations (e.g., regression analysis, r-squared) between various factors (such as
percent of impervious surface vs. peak flows, or population vs. average annual flood) with
available data in your watershed could be performed by following the methods described in
user-friendly books based on watershed research (e.g., Leopold 1994; Gordon et al. 1992;
Center for Watershed Protection 1998). However, a sound statistical approach can be difficult
to apply in a non-research setting due to lack of controls and inadequate data, funding, or
resources. If the analysis you perform is inconclusive or the uncertainty too great, reevaluate
your procedures and your available data. Quite often, sample sizes are just too small to
provide definitive answers. In such cases, if there is no clear alternative means of analysis,
don’t be afraid to admit that you don’t know or that you are unsure or that there is a lot of
uncertainty. It is quite common to have an indefinite outcome from analysis of environmental
data sets where tight experimental control is not feasible or cost-effective. Don’t let it bother
you; just be honest about the limits of the data and analysis and carefully qualify any
conclusions you develop.
Option 3: Relative Risk Model
The relative risk model (RRM) is another method for analyzing watershed data. The RRM is
a simple mathematical method for ranking stressors and altered conditions in a watershed and
the likelihood that they are associated with adverse impacts. It is a useful tool for prioritizing
which factors appear to pose the greatest risk to the ecological endpoints of interest. Most of
the process for conducting a relative risk assessment follows the suggested steps for any type
of environmental assessment, including what has been outlined in this Guide.
The RRM relies on identifying key stressors or altered processes in the conceptual model.
Data related to these factors is compared to data on reference conditions, as previously
described. The integrative aspect of the RRM is that stressors and sources of stress (land
use/human activities) can be prioritized based on their relative rank. To assign ranks:
 Each stressor (altered process or condition) is assigned a rank based on the difference
between the observed value (your watershed data) and the threshold above which an
unacceptable effect is likely to occur.
 Risk is calculated by comparing ranks for all stressors. The assumption here is that
stressors with the highest ranks are more likely than others to be linked to the adverse
effects.
The result of this analysis is a ranking of the types of human activities that are likely to be
linked to the harmful impacts and/or a list of stressors likely to be linked to these impacts.
This model does not prove cause and effect, but suggests tenable hypotheses about likely
causes and effects. In most cases, follow up studies are needed to obtain more definitive data.
However, the value of using the RRM is that it focuses efforts on scientifically credible
hypotheses that can lead to improved decision-making and management activities.
Option 4: Knowledge-base Models: Ecosystem Management Decision Support

46
One process for evaluating watershed condition involves using a new modeling approach
designed both to reflect inexact knowledge about natural processes and to be based upon
expert knowledge of a system. This approach is embodied in the software tool “Ecosystem
Management Decision-Support (EMDS).
The EMDS model is a computer-based model that can be used to compare the observed
conditions to reference values for a variety of watershed components and processes in order
to assess the present conditions. EMDS is an integrative approach to assessment, in that it
combines data about a place or concern. Because of this, the product can be one form of
watershed condition assessment. At the same time, imperfections in data knowledge about
aspects of the watershed processes and features will be reflected in the certainty of the
assessment. This is true of any model.
Option 5: Assessing Cumulative Watershed Effects
Cumulative watershed effects (CWE) or impacts, refers to two or more individual effects
that, when combined together, make a significant, usually adverse change to some biological
population, water quality, or other valued environmental attributes, or that compound or
increase other environmental effects (CEQA Guidelines, Section 14 CCR 15355).
Considering how the effects of human activities may combine to have greater consequences
than the individual effects is central to the watershed approach. Thinking about processes and
impacts in the watershed context usually involves combining individual, seemingly isolated
events. Evaluating CWE typically involves assessing the impacts that might occur in the
future as a consequence of certain human activities or changes in land use. This contrasts
with the previously described methods, which focus on analyzing present conditions that are
the consequences of past activities. Following are a few examples of different approaches you
might consider if you want to analyze cumulative watershed effects.
Step 6: Preparing an assessment Report
A critical component of watershed assessment is describing how you conducted the
assessment, what you found out, and how people can use the information to help them with
decision-making.
A. Watershed Assessment Report
The manual defines a watershed assessment report as: “a report documenting the findings of
the watershed assessment process.” There are several primary components to a well-written
report:
• Concise and accurate descriptions
• Use of structural elements like sub-sections, pull-out boxes, appendices and
indexing
• Visuals (photos, maps, charts)
• Clear distinctions between how you did something, what you found, and what it
means
The report can be published in a variety of ways:

47
• On a CD with hyperlinks to relevant material on the CD and the Internet
• Online, with links to other online resources
With electronic publication, consider associating the watershed assessment report with maps
and other data types on the same Web site. Maps can be served using Internet map server
software. Data can be shared as stand-alone tables for download, or as an online searchable
database. Photographs of parts of the watershed or issues of concern can be linked from a
map or from the report itself.
B Report Evaluation
It is a good idea to build product review into your schedule and budget. Decision-makers and
others using the product may have more confidence in it if it has gone through review. The
outline or framework for the assessment, an interim draft report, and the final draft report can
all be peer or expert reviewed.
Some evaluation criteria that can be used for this process are:
• The flow from assessment questions through data collection and analysis to findings and
recommendations makes sense and is explicit.
• Data are presented clearly and analysis methods are described.
• Conclusions are based on scientific and statistically valid approaches.
C. Make Recommendations
Many watershed assessments make recommendations for particular restoration, management,
policy, or monitoring actions that could or should be taken in a watershed. We recommend
that you keep most action-oriented recommendations in a watershed plan (e.g., a watershed
management plan).
Here are examples of things that the Manual team considers appropriate assessment
recommendations:
• How to deal with data or knowledge gaps
• What assessment findings can be linked to elements in a watershed management plan
• How the assessment could inform future watershed decision-making
Here are examples of recommendations that fit well in a watershed plan:
• Certain restoration actions in specific sub-watersheds would benefit watershed
function.
• Changing specific land and water management policies and implementation would
benefit watershed function and condition.
• Prioritized actions and places for action
• Description of the relative benefits of carrying out specific management and
land/water use actions for watershed function and condition.
D. Use in Decision Making

48
A watershed assessment that is not used in decision-making has lost an important function.
The type of decision may range from “more needs to be learned about the system” to “land
use designation or pollutant discharge must be tied to watershed impacts. “Decisions may be
combined as planned actions in a watershed management plan, or occur separately in
different decision-making venues (e.g., local government). There are many types of decisions
that can be informed by watershed assessments; here are a few:
1) Restoration planning, from action at a single site to changes in permitted land-use within
a watershed, is best done in the context of watershed assessment. Natural and human
processes will affect the efficacy of the restoration action and should be taken into
account by the restoration planner. In turn, a watershed assessment intended to support
restoration planning should make explicit the connections between watershed processes
and/or sub-watershed condition and potential restoration actions.
2) Regulation of human activities on the landscape or in waterways is a critical part of
environmental management. Regulation of these activities should be informed by
watershed assessment when the activities can cause watershed-wide impacts or originate
from large portions of a watershed (e.g., non-point source pollution). Examples include:
permitted discharges from point sources, permitted discharges from diffuse sources (e.g.,
under an agricultural discharge waiver), timber harvest plans, housing development
planning, parcel subdivision and zoning, road or highway construction or enlargement,
water diversion and storage, public lands grazing or logging, and channel or floodplain
modification.
3) Land use planning is carried out by local agencies in California and affects how and
where we impact the environment. General plans, zoning ordinances, and parcel
subdivisions are important land-use decisions that could be informed by watershed
assessment. General plans describe how much new development is desired and where it
will have placed in a city or county, and therefore in a watershed. Zoning decisions show
what kinds of development – industrial, commercial, residential, agricultural – are
permitted in specific areas and thus in sub-watersheds. Subdivision of parcels by
landowners, which must be approved by local governments, affects future development
patterns (including roads, water delivery, and sewage treatment). Aspects of watershed
assessment, such as water quality analysis, erosion modeling, and habitat degradation, are
useful to inform the where and how much of land-use decisions.
4) Water management is a fundamental driver of condition in many watersheds. Water may
be stored, diverted, or pumped from underground. How much of it is moved around,
when it is moved, and where it ends up can all affect the health of waterways. As
California’s population grows and increases pressure on surface and ground water
supplies and as climate change increases the chance of dramatic shifts in weather and the
need for science-based water management increase. Watershed assessment can inform
water management by describing the natural volumes and timing of flows, showing
locations of current and restorable aquatic habitat, and making the links between surface
and sub-surface water quality and quantity.
5) Monitoring of watershed conditions should both inform and be informed by watershed
assessment. Monitoring programs can be designed, or modified based on the findings in a

49
watershed assessment. Monitoring can be allocated to sub-watersheds already under
pressure, or at risk of future pressure from human activities. The type of monitoring
occurring (e.g., water quality, aquatic biology, geomorphology) should be dictated both
by what you find in your assessment and what remained as questions about watershed
functioning. The location of sample sites, the sampling frequency, the parameters chosen,
and the way the data is analyzed can all be informed by your watershed assessment.
4.3 Analysis of watershed
4.3.1 Problem analysis
The planning process typically begins with an identification and analysis of the issues and
problems. Problem analysis is a partly subjective assessment of reality, in a watershed
context, of the current and expected future state of water and related resources.
Clearly identifying and analyzing problems may seem simple, but in practice it is often quite
difficult. It is therefore important to focus the process from the very beginning by setting a
topic, which in a watershed context will naturally be watershed functions, with due
consideration given to both upstream and downstream issues. When establishing watershed
management guidelines, the topic is more concretely water resources related functions that
involve in particular the beneficial economic, social, and ecological uses of water resources
(a beneficial use is any use of the water resource conducive to public benefit, welfare, safety,
health or enjoyment. Beneficial water uses for example comprise domestic water supply,
human primary contact, maintenance of habitat for fish and wildlife and agricultural and
industrial water supply. ) (both on-site and off-site (in general the term off-site means that
something is taking place or is located away from the site of a particular activity.)), and
consideration is also taken of their interaction with related resources such as land.
Key Problems in Watersheds
The problems that affect watersheds are complex and long-term in nature. Watersheds
provide essential livelihoods for their inhabitants, but their natural resources are finite, often
under pressure and at risk of degradation. Degradation caused by unsustainable
exploitation of natural resources is usually the key problem. It leads to poverty, food
insecurity and social conflict. The negative socio-economic consequences of unsustainable
resources use are significant. In a watershed context, degradation can be described as follows:
Watershed degradation is the loss of value over time, including loss of the productive
potential of land and water, accompanied by significant changes in the hydrological behavior
of a river system which results in the inferior quality, quantity and timing of water flow. It is
the outcome of the interaction of physiographic features, climate and poor land use, as well as
other human activities. Watershed degradation accelerates ecological degeneration, reduces
economic opportunities and increase social problems’
Causes and symptoms of degradation may vary from country to country, but they do have
common traits. These include ecological, socio-economic, technical and institutional issues,
which are often interlinked and typically consist of some of the following elements:
• Ecological
• Socio-economic

50
• Technical
• Institutional
Key challenges for watershed management
From an organizational or a managerial point of view, there are some key challenges for
watershed management:
 Impacts of management interventions are difficult to assess
 Areas of decision making are not identical with watersheds
 Watershed management has to face competition and even conflict
 Global climate change
Depending on the given situation, the topics may be broader than just water and related
resources. As for instance, in the establishment of an all-encompassing watershed
management plan, general socio-economic development will also need to be considered. The
setting of a topic will also need to involve the consideration of the time aspect. Planning is
always concerned with future developments, and thus needs to anticipate, forecast, and set
future targets. In order to effectively do this, it is important to be clear about the planning
horizon. Furthermore, it is helpful to split the process of problem analysis into three separate
steps, namely:
• Stakeholder analysis
• Problem identification
• Problem structuring
Stakeholder analysis
The stakeholders will likely be a source of vast historical knowledge of activities that have
taken place in the watershed. Ask them for any information they might have on the water-
shed, including personal knowledge of waste sites, unmapped mine works, eroding banks,
and so on. They might have information on historical dump sites, contaminated areas, places
experiencing excessive erosion, and even localized water quality sampling data. Stakeholders
might be aware of existing plans, such as well head or source water protection plan.
Collecting this background information will help you focus your efforts to identify the issues
of concern and solutions.
Problem analysis requires the consideration of the views of all important stakeholders,
thereby making stakeholder analysis an essential part of the process. It can be conducted prior
to the problem analysis as part of a watershed inventory, or as a parallel procedure. A basic
premise behind stakeholder analysis is that different groups have different concerns,
capacities and interests, and that these therefore need to be explicitly understood and
recognized in the process of problem analysis.
In a watershed context, the state of water and related resources may be a serious problem for
one individual, group or organization, which another does not consider to be a problem at all
or perhaps not one of high priority. Only those individuals, groups or organizations who
perceive the state of water and related resources to be a problem of high priority will be

51
actively engaged in, and supportive of the planning process. Others may be neutral, reluctant,
ignorant, or even counter-productive in their actions.
The main steps involved in stakeholder analysis are to:
 Identify the general development problem or opportunity being addressed / considered
 Identify all those groups who have a significant interest
 Investigate their respective roles, different interests, relative power and capacity to
participate (strengths and weaknesses)
 Identify the extent of cooperation or conflict in the relationships between stakeholders
 Interpret the findings of the analysis and incorporate relevant information into project
design and into intervention measures to help ensure that: (a) resources are appropriately
targeted to meet distributional / equity objectives and the needs of priority groups; (b)
management and coordination arrangements are appropriate to promote stakeholder
ownership and participation; and (c) conflicts of stakeholder interest are recognized and
explicitly addressed in the design.
The following questions are pertinent during a stakeholder analysis:
 Who does and who does not perceive the state of water and related resources as a
problem?
 Who has a basic interest in the use and ownership of water resources and related
resources such as land that impact on those?
 What is the role and power of that user or user group?
 What important upstream and downstream groups of stakeholders exist within the
watershed?
 Who represents the interests of downstream populations outside the watershed and
society at large in consideration of the watershed's contribution to the river basin that it is
a part of?
 Who has a mandate to guide and organize water and land use?
Depending on the appearance of the stakeholder landscape, problem analysis may need to be
structured in several stages that consider the position of different groups of stakeholders. For
instance, the resident population of a watershed may not view off-site (external) impacts as a
problem, because these occur outside of their geographical area of interest. At the same time,
downstream populations or government agencies charged with water resources management
may view off-site impacts as the key problem. This demonstrates the importance of not only
conducting a problem analysis in a watershed context with local stakeholders, as otherwise
there is the danger of ignoring the interests of downstream populations and society at large.
There are a variety of tools that can be used to support stakeholder analysis, which can be
divided into the following two broad categories:
 Tools to analyze the stakeholder landscape, such as the stakeholder analysis matrix
and Venn diagrams

52
 Tools to analyze the situation of individual stakeholders, such as the SWOT analysis
and spider diagrams.
In using any of these tools, the quality of information obtained will be significantly
influenced by the process of information gathering. In this regard, the effective use of
participatory planning and facilitation tools can help ensure that the views and perspectives of
different stakeholder groups are adequately represented and understood.
Problem identification
Problem identification involves the compilation of an overview of basic, existing issues and
problems that characterize a given situation. In the process of establishing watershed
management guidelines, one will need to consider the issues and problems both on-site and
off-site affecting the beneficial economic, social, and ecological uses of water resources as
well as their interaction with related resources such as land. In this context it is particularly
helpful to look at the impairment of beneficial uses.
The term use impairment describes a given situation where a beneficial use is constrained by
the inadequate quality or quantity of water. Inadequate in this sense means that water quality
or quantity is lower than the maximum level possible. Basically, the approach of identifying
issues and problems means stating them in terms of impairments of beneficial uses of water
resources, paying specific attention to the timing and spatial extent of those impairments.
In order to begin identifying use impairments, one needs an overview of the beneficial uses
that are relevant in the watershed. For each of these uses one needs approximate estimates of
the following demand and supply of the water resources that they require:
 Present demand.
 Future demand, which is expected at the end of the planning horizon, as far as
currently foreseeable as a result of realistic forecasts. This implies that one has an idea of
the plans and expectations for future development of actors both on-site and off-site.
 Present supply. This is the highest amount or quality of water resources available for
current usage.
 Future supply that is expected at the end of the planning horizon, as far as currently
foreseeable as a result of realistic forecasts. This is the highest amount or quality of water
resources expected to be available for future use.
 Maximum supply. This is the maximum, either the highest amount or the best quality
that would be available for use under ideal conditions. It reflects the maximum capacity
of the watershed to supply water resources for a particular use.
Furthermore, we need an overview of relevant related resources, such as for example land
and forests – the use of which may have an impact on the beneficial uses of water resources.
For each of these we need approximate estimates of:
 Present use and its impact on the beneficial uses of water resources
 Future use and its impact on the beneficial uses of water resources, as far as currently
foreseeable as a result of realistic forecasts

53
The above information will largely be provided by the watershed assessment. Other portions
of it, such as the impact of land use on beneficial uses of water resources, may require the
statistical evaluation of long-term time series and / or the use of models.
Once the above information has become available, we can in principle begin to identify use
impairments using the following scheme:
A use impairment is given, when supply is lower than demand, whilst at the same time supply
is lower than at its maximum possible level, as is illustrated in cases [A] and [C] in Figure 5.
The use is then considered to be constrained.

Figure 5: Identification of use impairments

A use impairment is obviously not given, when supply is higher than demand, as illustrated in
case [B], or when supply is equal to demand, as illustrated in case [D] in the above chart. No
one would regard such a situation as a problem.
A use impairment is also not given, at least from a management point of view, when supply is
lower than demand, but supply is already at its maximum possible level, as illustrated in case
[E] in the above chart. People might view such a situation as a problem, but it is one that
could not realistically be solved because there would be no means to increase supply.
On a time, scale, there are two basic cases of use impairments:
 A present beneficial use is impaired.
 A future beneficial use is expected to be impaired.
The arrows in the above chart indicate how use impairments can be removed, namely by
increasing supply so that it becomes higher than demand. In order to increase supply, one
firstly needs to investigate the cause of the supply being below its maximum possible level.
The two basic causes can be described as follows:
 The use of a related resource causes the present supply or is expected to cause the
future supply to be below its maximum possible level.

54
 Another competing use of water resources causes the present supply or is expected to
cause the future supply to be below its maximum possible level.
In addition to the above scheme, it is helpful to ask the following sequence of questions so as
to define problems in terms of use impairments:

1. Please create a long list whilst gathering the information: What are all the use
impairments currently observed in the watershed? What are the use impairments
anticipated under future growth conditions until the end of the planning horizon?
2. What is the geographical extent of use impairment? Does it occur locally, or in an
upstream / downstream context within the watershed (i.e. on-site)? Does it occur
outside of the watershed (i.e. off-site)?
3. Why is each use considered to be impaired? What parameters do people use as
criteria in deciding that the use is no longer viable? What standards or objectives for
the use are currently not being met?
4. Which of these parameters can be measured easily? Is existing data available for any
of these? Which can be simulated using simple calculations or computer models?
Which are felt or speculative?
Problem structuring
Problem identification will give an overview of the nature of the issues and problems, and
perhaps also an overview of their relative importance, as well as indicate some directly linked
causes. However, it will not explain how they are interrelated, and show the possible cause-
effect relationships that exist between them. It is for this reason that further and more
elaborated problem structuring is required.
The problem tree approach is a very common tool used in problem structuring. It links the
results of problem identification together with the various issues or factors which may
contribute towards identified problems. It is done by arranging the identified problems and
other accompanying issues or factors into a tree-shaped hierarchy of cause-effect
relationships, and thereby identifying underlying or root causes of the problems. It can also
help to further rank problems according to their relative importance. Those problems at the
top of the hierarchy, which have many other issues and problems as contributing factors are
likely to be those which should be prioritized and for which planning should find solutions.

55
Figure 6: Problem tree designating the overall problem and the root underlying causes of the
PATSIR project (Eyasu, et. al., 2011)
The problem tree can frequently be used at later, more detailed stages of the planning
process. It is particularly used during the objectives analysis stage, when specific objectives
for solving key problems are formulated and related indicators to measure success are
identified. The elements of the problem tree are then re-formulated as an objectives tree. This
is done by re-wording the problems, beginning at the top of the problem hierarchy, as
positive statements or objectives.
Problem tree analysis is helpful in illustrating the linkages between a set of complex issues or
relationships by fitting them into a hierarchy of related factors. It can be utilized in order to:
 Link together the various issues or factors which may contribute to problems
 Help to identify the underlying or root causes of problems
The major assumption underlying a problem tree is the hierarchical relationship between
cause and effect.
The implementation of a problem tree analysis will require an individual to complete the
following process:
• Identify the major existing problems / issues based on available information (e.g. by
brainstorming)
• Select one focal problem for the analysis
• Develop the problem tree beginning with the most substantial and direct causes of the
focal problem.
4.3.2 Scoping the plan
Once problem identification and analysis have been conducted, there still remains the
challenge of agreeing on the priorities among these issues and problems. The stakeholders
need to determine which uses of water resources and which watershed areas urgently require
interventions, and which of them do not. This process is called scoping. The priority uses and

56
areas identified during scoping require standards and targets to govern their management, or
even concrete management interventions, while other uses and areas of lower importance will
require them only later or maybe even not at all. This also implies the identification of target
groups (resource users) who will be expected to implement standards or otherwise be
involved in interventions. Scoping is a stakeholder consultation and negotiation process that
establishes broad agreed priorities and targets, and gives the subsequent detailed planning
process overall direction. Scoping addresses the following fundamental issues:
 The setting of priorities must take place through appropriate consultation and
negotiation processes that involve all important stakeholders. Stakeholder support that is
achieved through the procedure of stakeholders coming to an agreement on the priorities
and course of action is the most important element in watershed management. A lack of
this support, and any disagreement about which issues are to be addressed and which
problems are to be solved, present the biggest obstacles, even if the stakeholders agree on
most issues and problems found in the watershed.
 Any setting of standards or design of interventions have transaction costs, or in other
words will require different kinds of expenditure, whether in terms of time, effort, or
money. Given the situation that these resources are limited, there is the need to set
priorities.
Scoping can be a challenging process because of the differences in perception and the various
priorities that exist among the group of participating stakeholders. It is not usually difficult to
generate a comprehensive overview of use impairments. However, it is much more difficult
to set priorities given these various use impairments. These are issues that must be resolved,
including the decision on which use impairments most urgently demand attention. A decision
must also be reached as to which of them can be deferred until later, and perhaps then could
be addressed during a more affluent time.
The scoping process
The scoping process has two important components:
Focusing (what?): Identification of issues and problems that have an overriding importance
in the watershed and which should therefore be considered in depth as intervention areas,
during the subsequent detailed planning process.
Boundary setting (where? when?): Limitation of the plan to a specific geographical area and
also to a particular time period.
Focusing is in principle a multi-stakeholder decision making process (This term is used to
describe processes which aim to bring together all major stakeholders in a new form of
decision-making structure on a particular issue. These processes are based on the recognition
of the importance of achieving equity and accountability in communication between
stakeholders, as well as involving an equitable representation of stakeholder groups and their
views. They are based on transparency and aim to develop partnerships and strengthened
networks between stakeholders, during which the results of the problem analysis are
reviewed, with particular attention being given to the cause-effect relationships as presented
in the problem tree(s). There is a selection of problems (use impairment of water resources)
57
that are considered to be the most important, and broad targets for their solution are
negotiated among the group. Whilst targets are being negotiated there are a number of issues
that need to be taken into account:
Firstly, there are overarching goals and standards which have been agreed upon by society
and have been set by a particular government that are relevant for watershed management.
These include policies and commitments made by ratifying international or regional
conventions. Additionally, there are goals and targets that have been set in national and sub-
national policies; development plans; water resources sector plans or sector plans for the
development of related resources such as forests.
The existence of national watershed management guidelines or policies, which obviously
need to receive particular attention during the scoping process.
The identification of underlying cause – effect relationships (impact chains) that have brought
about present or are likely to cause future use impairments. A reasonable understanding and
awareness of these cause – effect relationships are essential for the selection of broad
strategies to develop workable solutions.
The consideration of strategic options regarding tradeoffs between different use impairments.
As the different uses of water resources are interrelated, restoring one use may result in the
further impairment of another use.
The review of strategic options regarding standards or restrictions for the use of related
resources. For instance, the question of whether interventions should target conservation, land
use patterns or rather structural measures may need to be negotiated.
Boundary setting is often carried out through the prioritization of catchments using agreed
static or dynamic criteria. This is in principle the application of the critical watersheds
concept at the catchment level. It identifies those catchments where essential watershed
functions are already critically endangered due to human interventions or are likely to
become critically endangered in the near future. In these critical catchments the application of
standards or management interventions are most urgently required.
It would be helpful to ask the following questions during the scoping process:
Which are the most important use impairments for the stakeholders – out of the long list that
was developed during problem analysis?
It may be helpful to aim for a relatively short list of no more than five in order to reduce
complexity. These will then become the focus of the planning initiative.
What specific targets do the stakeholders wish to meet in order to consider an impaired use to
be restored?
What are the specific time periods the stakeholders want to set in order to apply these targets?
Should this be throughout the year or only during the dry season?
What is the size of the area over which they want to apply these targets? Should this be over
the entire area that is currently impaired, or only over a smaller portion of it?

58
The following provides an overview of the major water resources uses and issues to be
considered when setting broad targets during scoping:
Water quantity: Setting targets for water quantity requires an estimation of the water
demands of various users for withdrawal uses (e.g. abstraction for irrigation) or for non-
withdrawal uses (e.g. fish cage culture or recreation). An examination of water quantity needs
to consider total annual volumes, seasonal volumes and water levels that are needed to satisfy
the needs of different water users. The fundamental question is whether or not the watershed
has or will have enough water for all intended uses. The following key questions have to be
answered: What proportion of reliable annual flows is currently required to meet the
withdrawal demands and service non-withdrawal uses? What proportion of estimated
withdrawals is consumed – that is not returned to the hydrological system? How are these
supply-to-demand ratios likely to change in the future considering aspects such as population
growth and the increasing uses by agriculture?
Water quality: Although quality standards are subjective, workable criteria and standards that
can be used to define targets are usually in place at a national level. There are for example
minimum standards for the supply of drinking water or for the operation of hydropower
turbines. These criteria and standards are usually defined by water resources or environment
ministries. Furthermore, there are some rules and procedures in place for Lower Mekong
Basin wide water quality monitoring, which have been agreed upon between the riparian
countries.
Fisheries: Fish stocks are usually a good indicator of a healthy aquatic ecosystem. Setting
targets for watershed fisheries encompasses two principles, namely: (a) desired species
composition and desired stocks of each species, but also (b) secondary conditions in terms of
the habitat conditions required to support a self-containing population. Physical habitat
conditions include water depth, flow, temperature, suitable chemical parameters, and the
availability of appropriate spawning and nursery habitat.
Hydropower generation: In many watersheds hydroelectric power generation is the largest
and most important user in terms of economic value. Hydropower generation imposes special
requirements on watershed management, usually related to maintaining a minimum water
level throughout the seasons, and reducing erosion leading to the silting up of reservoirs and
damage to turbines. Upstream water and land uses affect river sedimentation and flow
regimes, with potential detrimental effects on water volumes, as well as reservoir capacity
and life span. A typical performance target can be to maintain water flow and quality that is
necessary for continuous hydropower generation during a set reservoir's lifetime.
Navigation: Shipping and boating are essential uses of rivers and lakes in many watersheds.
Navigation uses are constrained by water conditions such as channel depth, width,
turbulence, and availability of locks. Commercial navigation depends on the duration of the
navigation season, which depends on water depth. This once again may vary for different
classes of vessels depending on their maximum draft.
Recreation or nature-based tourism: Targets for recreational uses and tourism may be set.
For instance, the beauty of the landscape and a diverse pattern of water bodies and water-

59
related ecosystems such as riparian forests may need to be maintained or enhanced. A direct
target could be land or water surface made available and managed for recreational uses. If
recreational activities include swimming or bathing, clean water bodies are required.
Ecology and conservation: Many watersheds contain elements that are rare or of specific
ecological, aesthetical, recreational, historical or cultural value. It may be appropriate to set
targets as to what degree the natural character, state and function of any specific watershed
ecosystem area should be maintained.
Land use: Land use and related changes may have a profound impact on both water quantity
and quality. For instance, agriculture and other on-farm uses both consume water (e.g.
irrigation) and impact on water quality (e.g. erosion, fertilizer use). Therefore, targets for
certain land uses may need to be set in combination with the targets for water quality and
quantity. This obviously requires a good understanding of the interactions of land use and the
hydrological cycle, and may need to be supported by computer based modelling.
In some cases, the ideal conditions or targets for two water and land uses will differ and may
even be incompatible, such as for example a conflict between irrigated agriculture and
hydropower generation over optimum water levels. Such conflicts, if not already apparent,
are likely to emerge in public negotiations during scoping, and may need to be resolved
through conflict management techniques
The output of the scoping process
In a watershed context, the outputs of successful scoping are:
 A clear prioritization of existing problems (use impairments) to be solved and
anticipated problems to be avoided. This will be a sub-set of the results of initial
problem analysis.
 Targets for maintenance and / or restoration of watershed functions, formulated as
broad targets for the restoration of currently impaired uses of water resources, or
broad targets to prevent future impairment of those uses, both specified in terms of
content, location and timing.
 Broad strategic directions for the management of related resources.
 Focus, contents, methodologies and organization applied during the subsequent more
detailed planning exercise.
 Consensus among stakeholders on the above.

60
5. LAND EVALUATION PRINCIPLES AND APPLICATIONS
5.1 Principles of soil survey and land use capability classification
5.1.1 Introduction to Soil Survey
Soil Survey can be defined as the process of classifying soil types and other soil properties in
a given area and geo-encoding such information
Stages in soil survey program
Stage 1: Pre-Field Activities
Maps and data collection: collection of maps, aerial photographs, satellite imageries and
data’s, existing national, regional, and local topographic map are very essential to get first-
hand information.
Equipment and transport: the leader and the soil survey team should ensure that the
necessary equipment have been acquired, before field work begins.
Preparation of base maps: preparation of base maps to be used in the field and preliminary
delineation of soil from aerial photograph interpretation has to be done in the office.
Preparation of work plan, manpower and survey organization: the work plan for all
activities in office, field and laboratory analysis need to be scheduled on bar charts possibly
on daily, weekly, monthly and annual basis. The responsible expert for each activity has to be
assigned.
Stage 2: Field activities
The main part of activities in the field is checking identified soil classes and boundaries to be
mapped, the mapping legend and actual relations between soil properties and land forms or
surface feature, shown on the preliminary base. This has to be followed by augur observation,
pit description and sampling.
Stage 3: Post-Field Activities

 Submission of the samples to certified soil testing laboratories


 Compilation and analysis of field soil description data’s
 Analysis of laboratory results
 Final soil classification and mapping
 Land evaluation for irrigation
 Land evaluation for selected crops
 Assessment of soil and water conservation aspects
 Investigation of soil problems and recommendation of management options
Kinds of Soil Survey
• Reconnaissance surveys
• Detailed special surveys
• Special purpose surveys

61
Steps in soil survey
Step 1: Delineating soil unit
• Delineating soil unit with respect to topography
• Determination of texture
• Determination of slope – erosion phases
• Vegetation
Step 2: Soil sampling
• 2 types:
a) Free survey: As many as necessary samples are taken within the boundaries
b) Grid survey: Transverses and transects are laid out in a grid pattern, samples are
taken at every intersect. Distance between lines: 50 meter
Step 3: Profile Sampling
• Profile pit is dug to C or R horizon
• 120 to 200 cm if regolith extends beyond 200 cm
• Pit size is enough for a person to go down and turn around
• Expensive & Labor Intensive
• Expressing the physical phases of soil
- Soil Texture
- Soil Structure
- Soil Consistence/Soil Strength
- Soil Color
- Soil Permeability
- Soil Temperature
Step 4: Plotting the soil mapping unit
Collaborating analytical data are coded in term of soil types, phases (stoniness and erosion)
and slope.
Step 5: Correlation
• Previous profile study
• If the profile under study indicates very close similarities with an established profile
existing in another place the soil under study is name after the other soil
Step 6: Soil Mapping
Preparing scaled map showing all soil unit investigated
Step 7: Soil survey report

62
The report embodies all items in the soil survey program
Step 8: Land Capability Classification
• Refers to the productive capacity of the soil for intensive as influence by soil fertility,
moisture supply and depth of solum
• 2 main points about land capability
a) Soil limitation -soil physical aspect that restrict crop production
b) Climatic limitation - adverse climatic problem
5.1.2 Soil mapping
Soil mapping involves locating and identifying the different soils that occur, collecting
information about their location nature, properties and potential use, and recording this
information on maps and in supporting documents to show the spatial distribution of every
soil.
In order to map and identify different types of soil it is necessary to have a system of soil
classification. Soil classification have provided a major challenge, one that has yet to be
adequately solved. There are two international systems, FAO-UNESCO and US Soil
Taxonomy, but a number of other systems have evolved and many countries have produced
their own systems suited both to local conditions of soil formation and to local knowledge.
However, the one uniform feature in all these systems is that they are based on the soil profile
that is the appearance of the section of the soil from the land surface through to the rock or
other rock material that lies below the surface where the investigation is taking place. This
examined thickness is usually about 1 meter but in tropical areas with deeper soil
development it can be several meters, and in steep slope rock landscapes may be much less
than 1 m. soils are allocated to a type according to their profile characteristics including,
abundance of organic matter in the top soil, the texture (combination of the different particle
sizes and their ratio), colors representing aerobic and anaerobic conditions, structure of each
layer, etc.
Traditional soil mapping is conducted with an auger and spade at intervals throughout the
landscape. The intervals between inspections can be according to a pre-determined grid (grid
survey) or more often, are based up on the judgment of the surveyor who uses their
knowledge of the inter-relationship between soil type and landscape, geology, vegetation,
etc., to determine where to make inspections.
Auger borings are supplemented by excavated profile pits at determined points in the
landscape. These profile pits are used to demonstrate lateral changes in the soil as well as
vertical ones, and are important for the full description of type of soils and for the taking of
soil samples for chemical, physical and less commonly, biological laboratory analysis. In this
way a picture is built up of the soil in a region and its relationship to the landscape in which it
lies.
Soils can be mapped at a range of scales from very detailed at 1:1,250 to 1: 5,000 by which
the pattern of soils in individual fields can be identified, through to scales of 1: 500,000 to 1:
500,000,000 which provide only a much generalized picture of the soils of a country.

63
5.2 Principles of land evaluation
The concept of land evaluation
The term land evaluation refers to the process whereby the suitability of land for specific kind
or kinds of use (such as irrigated coffee production, irrigated horticulture, rain fed forestry,
spate irrigated sorghum, livestock production etc.,) is assessed. The main objective of land
evaluation is to select suitable tract of land for development and to find out the suitable
cropping and land management alternatives that would be physically and financially
practicable and economically viable. It helps in making decisions as to whether a certain
project should be implemented or not, and to fairly estimate what benefits or losses are likely
to accrue with its implementation.
Land evaluation is generally presented in the form of land suitability classes which may be
defined in qualitative or quantitative terms, depending on detail of surveys conducted and the
availability of economic data. The qualitative land suitability classes represent relative
suitability for the specified use, while the quantitative land suitability classes are described in
absolute economic terms like “net farm income”, “net returns per ha” or “net incremental
benefits”, etc.
The FAO frame work for land evaluation
The FAO’s framework for land evaluation (FAO, 1976) and the FAO’s guidelines on land
evaluation for rain fed and irrigated agriculture (FAO, 1983 and 1985) recommend land
suitability assessment at the following categorical levels:
a) Land suitability “Order”
b) Land suitability “Class”
c) Land suitability “Subclass”
d) Land suitability “Unit”
At the highest level of “Order”, the land is classified as: S: Suitable, N: Not suitable.
At the second level of “Class”, the order “S” is divided in to three (or four if need arises)
classes, on the basis of degree of suitability for defined use. Each class is identified by
using an Arabic number suffix and defined as below in terms of relative degree of
limitations posed, if any, to sustained application of a certain land use type and expected
reduction in productivity or benefits or increased inputs required to achieve the acceptable
level of outputs from the land.
S Highly No significant limitation; no significant reduction in
1 suitable productivity/benefits or raising of inputs above an acceptable level
S Moderately Moderately severe limitations; considerable reduction in
2 suitable productivity/benefits or input costs raised to the extent that the
use, though still attractive, is not as profitable as of S1 land
S Marginally Severe imitations; much reduced productivity/benefits or input
3 suitable costs raised to the extent that the use is only marginally justified.

64
The order “N” is similarly divided in to two classes on the basis of degree of unsuitability for
the defined use as:
N1: Currently not suitable
N2: Permanently not suitable
At the third level of “Subclass”, the classes “S2”, “S3” and “N1” (not S1 and N2) are divided
in to various subclasses, on the basis of kind(s) of limitation(s) to the defined use, by using
small letter suffixes representing the major limitations as, for example:
S2m: Moderately suitable due to low moisture availability
S3d: Marginally suitable due to impeded drainage
S2t: Moderately suitable due to unfavorable topography
Main procedures followed in land evaluation of a tract of land selected for a development
project include:
1) Preliminary study of the existing information, especially that relevant to the field
appraisals of land conditions and experiences in a fully developed area having
physical, climatic and socio-economic conditions similar to the project area;
2) Selection of the land use alternatives (crops, irrigation methods, soil management
technology, etc.) and description of the land use types (the land use with detailed
definition in terms of all important attributes) for evaluation; in some cases, the land
use types may be obvious from the outset (e.g., irrigated rice, rain fed sisal, irrigated
orchards, rain fed tree crops, etc.,); in some others, these may be already defined in
the terms of reference;
3) Preparation of a land resource inventory (from the land resource surveys already
discussed) of physiography, climate, geology, hydrology, soils, present land use and
vegetation, fauna and socio-economic conditions;
Three sets of data are obtained from the land resource surveys i.e.,
a) Definitions and descriptions of the land units,
b) Maps showing the distribution of these land units and
c) Values of land characteristics of the land units.
The land units are described in terms of all the above characteristics, except for socio-
economic conditions for which the data are separately collected by economists;
4) Listing of land use requirements in terms of the critical limits of land characteristics;
5) Selection of “Class-determining factors” (the variables affecting agronomic,
management, land development, conservation, and environment or socio-economic
conditions, and having influence on the outputs and inputs of a land use type) and
specification of critical limits of the values of land characteristics for designating the
land suitability categories (levels of suitability for individual land use requirements or
limitations, expressed as Sl, S2, S3, n1 or n2) and the weightage given to each "class
determining factor for evaluation;

65
6) Matching the values of characteristics of the land units with the critical limits of land
use requirements or limitations (land suitability categories) and classification of each
land unit to appropriate land suitability subclasses (and further to the land suitability
units, if required) considering the weightage of the class determining factors;
7) Mapping and description of the land suitability subclasses or units, or subclass
associations, depending on the level of investigations made.
Physical vs economic land evaluation
The land suitability classes defined above may be translated into various physical and
economic indices or measures of suitability. Three most practical and convenient measures of
land suitability which lend themselves to progressive applications as more and more data
become available are briefly described below.
i) Land Productivity Index. The land Productivity Index (LPI) may be defined as "the
physical productivity of land under a specified land use (land use type), relative to
that of the best land having similar physical environments". The indices are useful
tools to have rough estimates of the potential levels of productivity for comparison
with the provisional estimates of land development costs at the level of
reconnaissance and prefeasibility studies.
The LPIs may be expressed in terms of “relative yield” (the yield per unit area relative to that
of the best land as a percentage or fraction) or in terms of "absolute yields', (the actual yields
e.g., in kg/ha or t/ha). Both the ranges of values and the single values (which may be mid-
values or estimated values within the ranges) can be used for that purpose. A well estimated
single value within the specific range of the respective land suitability class is preferred in
cases when it can be wisely selected by the economist or planner fully conversant with the
evaluation procedures; in other cases, the mid-values may be used.
The proposed standards of LPIs for different land suitability classes under conditions
prevalent in Ethiopia are as in Table 2.
Table 2: Proposed land productivity indices for Ethiopia
Land suitability Land productivity index
class Range Mid value
Percentage Fraction Percentage Fraction
S1 81 – 100 0.81 – 1.0 90 0.9
S2 51 – 80 0.51 – 0.80 65 0.65
S3 20 – 50 0.20 – 0.50 35 0.35
N1 < 20 < 0.20 10 0.1
N2 0 0 0 0

66
The rough initial calculations based on the LPI may also help in the identification of
alternative land use types, including the possible cropping/conservation and management
systems, and the approximate amounts which can be invested in land development. It is
important to note that the land suitability classes into which the given land unit-land use type
combinations fall may differ abruptly if the "relative yield" is used as the index unit. It may
also be kept in mind that the LPIs do not take into account the prices or the costs of
production.
ii) Net Farm Income: The Net Farm Income is defined as "the value obtained after
subtracting all variable and fixed costs from the gross value of the production got
from the land". It is a suitable method to evaluate land for prefeasibility studies when
the common project costs are not generally known, as these are not taken into account
in estimating the Net Farm Income. The Net Farm Income may be calculated for both
the ‘without project’ and the ‘with project’ situations. It would, however, generally
suffice to base the land suitability classification on the 'with project' Net Farm
Income. The land evaluation by this method does not consider the increase of income
effected by moving from a 'without-project' to ‘with-project' situation. This evaluation
would need further refinement for detailed project planning as it would generally
include some "marginally suitable lands" that would have to be eliminated by further
economic analysis. The further revisions and adjustments in land evaluation may,
however, be curtailed significantly by making estimates in terms of “Net Incremental
Farm Income” (i.e., ‘with project’, Net Farm Income minus ‘without-project’ Net
Farm Income, which is normally assessed for the year of full development).
iii) Net Incremental Project Benefits: The "Net Incremental Project Benefits" may be
defined as "the increase in net benefits from a unit area of land envisaged with the
development under a project plan. It is calculated as:
Net Incremental Farm Income - "Annual Equivalent Value of Common Costs"
(i.e., the project costs calculated on yearly basis; not including the costs of land
improvements like land clearing, drainage, reclamation, levelling etc.) - "Annual
Equivalent Value of Area-specific Land Development Costs".
The "Cut-off value" (the value corresponding to the Net Incremental Project Benefits being
zero) used to define the boundary between the "Suitable" and the "Not suitable" lands is
virtually that represented by the Annual Equivalent Value of Common Costs, and it should
include the project's Annual Operation and Maintenance Costs.
The approaches followed for economic land evaluation
It is generally convenient to follow a two-stage approach to arrive at quantitative (economic)
evaluation i.e., a first approximation of land suitability is made on the basis of physical
criteria, resulting in a qualitative land suitability classification, and then the economic and
social analyses are carried out for each suitability class on the most promising land use
alternatives. The other method, the parallel approach, includes the economic criteria
throughout the process of matching and land suitability classification and the land classes are
based on economic assessment whereby the physical limits are selected to fit the economic
limits.

67
The "two-stage" approach has the advantage of being more straight-forward to follow and
having validity of the physical classification made for a long time, so that the final economic
evaluation can be easily revised in the light of changing economic conditions. Moreover, it is
probably more scientific since it provides a strong basis in the form of land suitability
classification arrived at by using more or less permanent land parameters, which can be
further used to evaluate the land for some more prospective land use types identified in near
future. In comparison, the "parallel" approach directly results in land evaluation which is
valid for a relatively short time and entails repetition of the whole process to revise it
whenever the changed economic conditions necessitate it. A disadvantage visualized in land
evaluation by the "two-stage" approach, however, is that some more promising economic
alternatives may be overlooked in the initial physical evaluation.
5.2.1 Use of aerial photographs for land evaluation
The use of aerial photography to assess and map landscape change is a crucial element of
ecosystem management. Aerial photographs are ideal for mapping small ecosystems and fine-
scale landscape features, such as riparian areas or individual trees, because they often possess
a high level of spatial and radiometric (tonal) detail. Aerial photographs also provide the
longest-available, temporally continuous, and spatially complete record of landscape change,
dating from the early 1930s in some cases. As a result, aerial photographs are a source of
valuable historical information on vegetation cover and condition. Aerial photographs can
reduce costs involved in mapping, inventorying, and planning, and, as such, are used for
applications ranging from forest inventories, disturbance mapping, productivity estimates,
land evaluation, and wildlife management. Thus, many important management decisions are
routinely made on the basis of maps derived from aerial photographs.
Despite the many advantages of aerial photographs, there are specific challenges for using
them, especially with respect to manual aerial photograph interpretation. Although manual
interpretation by highly trained individuals remains one of the most effective and commonly
used approaches for classification of aerial photographs, this technique relies greatly on the
personal experience, knowledge, and expectations of the interpreter for a given location.
Thus, human interpretations are subjective, and are vulnerable to inconsistency and error. In
addition, resource management agencies are beginning to face a shortage of well-trained
interpreters, especially those whose skills have ideally been combined with years spent in the
field. As a result, there is a need for new approaches to reduce or eliminate these difficulties
associated with traditional aerial photograph analysis to help foster its continued and evolving
use.
5.2.2 Land evaluation for irrigated agriculture
Parametric evaluation system for land suitability evaluation for irrigation (The method
of Sys and Verheye, 1974)
The aim of this parametric evaluation system (Sys and Verheye, 1974) is to provide a method
that permits evaluation for irrigation purposes, and that is based on the standard
granulometrical and physico-chemical characteristics of a soil profile.

68
It has been estimated that the soil as a medium for plant growth under irrigation should in the
first place provide the necessary water and plant nutrients in an available form, and in the
most economical way.
The factors influencing the soil suitability for irrigation can therefore be subdivided in the
following four groups:
• Physical properties that determine the soil-water relationship in the solum such as
permeability and available water content both related to texture, structure and soil
depth, also CaCO3 status and gypsum status could be considered here;
• Chemical properties that interfere in the salinity/alkalinity status, such as soluble salts
and exchangeable Na;
• Drainage properties and
• Environmental factors, such as slope.
In this method, the different land characteristics that influence the soil suitability for
irrigation are rated and a capability index for irrigation (ci) is calculated according to [Eq. 1]:
A∗B
∗C
100
∗D
100
∗E
100 Equation 1
∗F
100
∗G
100
CI =
100
Where:
CI =capability index for irrigation
A =rating of soil texture
B =rating of soil depth
C =rating of CaCO3 status
D =rating of CaSO4 status
E =salinity/alkalinity rating
F =drainage rating
G =slope rating
Table 3: Suitability index for the irrigation capability indices (CI) classes
Capability Class Definition Symbol
Index
>80 I Highly suitable S1
60 – 80 II Moderately suitable S2
45 – 60 III Marginally suitable S3
30 – 45 IV Currently not suitable N1

69
<30 V Permanently not suitable N2

The classes II to IV can have following subclasses with regard to the nature of the limiting
factor:
• s-limitations due to physical soil properties (A, B, C, and D);
• n-limitations due to salinity/alkalinity (E);
• w-wetness limitations (F) and
• t-topographic limitations (G).
Factors influencing the soil suitability for irrigation are given in the following tables (Table 4
– Table 9).
Table 4: Rating of soil depth
Soil depth (cm)  Rating for gravity irrigation 
<20  30 
20-50  60 
50-80  80 
80-100  90 
>100  100 

Table 5: Rating of textural classes for irrigation


Textural Rating
class Gravel Fine gravel Coarse gravel
<15 % 15-40% 40-75% 15- 40-75%
40%

70
Clay Loam (CL) 100 90 80 80 50
Silty Clay Loam (SiCL) 100 90 80 80 50
Sandy Clay Loam (SCL) 95 85 75 75 45
Loam (L) 90 80 70 70 45
Silty Loam (SiL) 90 80 70 70 45
Silt (Si) 90 80 70 70 45
Silty Clay (SiC) 85 95 80 80 40
Clay<60% 0-2 μ 85 95 80 80 40
fraction(C) 80 90 75 75 35
Sandy Clay (SC) 75 65 60 60 35
Sandy Loam (SL) 65 65 55 55 30
Clay >60%0-2 μ 55 50 45 45 25
fraction(C)
30 25 25 25 25
Loamy Sand (LS)
Sand (C)

Table 6: Rating of CaCO3 content

CaCO3 (%) Rating for gravity irrigation


<0.3 90
0.3-10 95
10-25 100
25-50 90
>50 80

Table 7: Rating for gypsum content.


CaSO4 content, % Rating
>50 30
25-50 60
10-25 85
0.3-10 100
71
<0.3 90
Table 8: Rating for salinity and alkalinity

ESP Electrical conductivity, EC in dS/m


0-4 4-8 8-16 16-30 >30
0-8 100 95 90 85 80
100 (*) (90*) (80*) (70*) (60)
8-15 95 90 85 80 75
90(*) 80(*) 70(*) 60(*) 50(*)
15-30 90 85 80 75 70
80(*) 70(*) 60(*) 50(*) 40(*)
>30 85 80 75 70 65
70(*) 60(*) 50(*) 40(*) 30(*)

(*) clay, silt clay, sandy clay

Table 9: Rating for drainage classes as related to texture and salinity and depth of
groundwater
Drainage class Rating
Clay, silty clay, sandy clay, Other textures
silty clay loam
Non Saline Non Saline
saline groundwater saline groundwater

72
• Well drained soils gley at
>3 m 100 100 100 100
2-3 m 95 85 100 100
1.2-2 m 90 75 95 95
• Moderately drained with gley
at 80-120 cm 80 50 90 70
• Imperfectly drained with gley
at 40-80 cm 70 35 80 60
• Poorly drained soils with gley
at <40 cm 60 30 65 40
• Very poorly drained reduction
horizon at <40 cm 40 20 65 30

Table 10: Rating of slopes for gravity


Rating for gravity irrigation
Slope class (%)
Non terraced Terraced
0-1 100 100
1-3 95 95
3-5 90 95
5-8 80 95
8-16 70 85
16-30 50 70
>30 30 50

6. WATERSHED TREATMENT MEASURES


6.1 Introduction
Watershed is a fundamental unit for water resources management. The term watershed
implies a domain or system within boundaries. The watershed domain may be further
divided into sub-components of smaller watershed or into sub- processes such as overland
flow and sediment yield.
The term watershed management practice involves changes in land use or land management
practices including structural and non-structural measures. Alternatively, it refers to use of
the land according to its capability for sustainable resource uses. Watershed management
activities involve the following major components;
73
- Stabilizing the top soil,
- Stabilizing stream flow,
- Improving water quality.
Watershed management practices affects upstream and downstream area and usually affect
more than one community, institution and administrative levels. Implemented watershed
management activities yield physical effects (on site; increase productivity; downstream:
reduce sedimentation; increase water flow during critical period, etc.). The key objectives of
watershed management aim to promote both soil conservation and water resources objectives
with upland management strategies that diversify and increase income generation through the
production of agricultural and natural resources.
Thus, watershed management activities/practices involve inter-sectoral activities including
the following:
- Improvement in crop production (rain fed and irrigated),
- Improvement in livestock production,
- Improvement in forestry production,
- Environmental improvement,
- Irrigation and hydropower developments,
The integrated watershed management practices discussed below are very close to each other
and interrelated. The upland soil conservation practices on only arable land can only be
partly effective if not complemented with forage improvement, improved agronomic
practices and others. However, the watershed management practices listed below are not
place-specific, however, variable with local circumstances.
The integrated watershed management practices /activities should be defined as per Land
Class (Chapter 5). Best options from the list should be decided by taking into account
farmer’s needs and priorities. Indeed, the selected management option should have an
immediate and positive impact on farmer’s production or well-being.
6.2 Watershed treatment measures for non-arable land
The use of non-arable land is mainly limited to pasture, forest, wildlife and recreation. These
lands are generally confined to upper reaches of watershed and have an undulating
topography which are susceptible to soil erosion.
Soil at steep slope and lack of vegetative cover, soil erosion is accelerated transporting large
amounts of sediments into streams at the downstream sides. Uncontrolled runoff from the
sloping lands also cause extensive damage in to lower reaches of the watersheds.
In order to prevent degradation of these lands, vegetative and mechanical measures are
employed together and are complimentary to each other. Mechanical measures act like the
foundation of the building whereas the vegetative measures act like a superstructure which
helps in improving the productivity of non-arable lands. These lands have a great potential

74
for producing fodder, fuel, minor forest produce, fruits and low quality timber. It is essential
to treat these lands with suitable location specific soil and water conservation measures.
The practices such as contour trenching, gradonies, installation of temporary and permanent
structures for gully control, construction of sediment retention structures and retaining walls,
reclamation of ravine lands, improvement and management of grass lands and rehabilitation
of mined lands can be adopted for soil and water conservation measures in non-arable areas
of watershed.
Contour trenching
Contour trenching implies excavating trenches along the contour or along a uniform level
across the slope of the land. Bunds are constructed downstream along the trenches with the
material excavated out of them. These trenches break the slope lengths, reduce the velocity of
runoff and retard its scouring action. The rainwater retained in the trenches percolate slowly
in the soil profile and finally increase groundwater recharge. It is very effective and low cost
moisture conservation techniques particularly in rain fed area and hilly region. Size of
trenches generally varies from 0.30 m x 0.30 mx 15 m to 0.50 m x 0.50 m x 15 m and spacing
is kept in between 3 to 5 m. Construction of contour trenches is always started from the ridge
and progressively extended towards the valley. The top soil should be placed on the upstream
side whereas boulders and gravels are staked on the lower side to act as toe to the soil bank.
Gradonies
Gradonies are steeply inward sloping narrow bench terraces constructed on contours.
Usually, gradonies are suitable for afforestation in uniformly steep sloping lands. Based on
the steepness of the slope width of gradonies is decided. Hands tools are usually employed
for the construction of gradonies and digging of trenches for hedge row planting.
Gully control measures
Gully erosion damage the land and also contribute the large amount of sediment load in the
surface runoff. Small rills due to runoff develop in to a deeper crevices depending on the
velocity of flow and slope and result in gully formation. The method of control of gully
erosion are diversion of runoff, construction of temporary structures, construction of
permanent structure and vegetative method.
Construction of temporary structure
These includes brushwood dams and loose rock dams. These are constructed with locally
available materials like rocks, bamboos, woven wire and gabion check dams etc. A series
check dams can be constructed on a stream to recharge the depleted aquifers. The spillway is
designed to convey peak flow with ten years return period. The spillway bottom width
should be equal to bottom width of stream. Many check dams are more beneficial than the
construction of large check dam. Check dams should be so spaced that the crest level of one
will be same as the base level of next dam at upstream.
Construction of permanent structure
In advance stages of gully, the construction of temporary structure may not adequate to

75
check the erosion development process, so strong and permanent gully control structures are
necessary for of stabilization gullies. Permanent structures on gullies block the flow of water
and store water at upstream, which is then used for irrigation purposes. These structures are
constructed with permanent material, namely cement concrete masonry, etc. Permanent
structures are provided with adequate capacity to handle the design runoff and to control the
floods. Permanent structures retain the store water, because of which the energy of the
flowing water is dissipated and the flow of velocity reduced. Ponding helps to stabilize a
gulley by assisting in settling of suspended sediment on the base. Most common permanent
structures are drop spillways erected on the gully beds when the drops are small, drop inlet
spillways used as outlets from water storage reservoirs and inclined spillways installed at
gully heads when the drops are large.
Vegetative measures
Vegetative measures are effective in small gullies in which vegetation is established on the
gullies to control the erosion. The vegetation may consist of grass cover and various types of
plants. Apart from protecting the gully they can generate some income. Smaller gullies of
less than three-meter head and catchment area of less than 10 ha, can be treated with sod
flumes. Grasses in bands are planted at the softest and erodible spots, such as near water fall.
These are providing intermittently as sod bands in the channel and laid out in series across
the slope.
6.3 Conservation measures for arable land
The main aim of conservation measures for arable land is to reduce or prevent erosion while
achieving desired moisture for sustainable production. Conservation measures mainly
depend upon topographic condition and rainfall characteristics. Conservation measures
adopted are broadly classified as:
a) Biological measures and
b) Mechanical measures.
6.3.1 Mechanical measures
Main objective of soil and water conservation measures is to increase the time of
concentration and thereby allowing more runoff water to be absorbed and held in the soil
profile and intercepting a long slope into several short ones, so as to maintain less than a
critical velocity for the runoff water. These includes contour bunding, graded bunding,
terracing and vegetative waterways.

Contour bund
Contour bund is the most widely practiced soil and water conservation measure for mild
slopes (1-6%) and low rainfall (<600 mm) area. Contour bunds are very effective in light
soils in low rainfall conditions so that immediate infiltration of the stored water would
increase the soil moisture for rain fed crops. The main objectives of contour bunds are to
reduce the length of slope for checking soil erosion and to impound water and permit more
seepage of runoff for increasing soil moisture. The bunds made on contour break the length
of slope before rill formation takes place.

76
Graded bund
Graded bunds or graded terrace or channel terrace are laid along pre-determined longitudinal
grade instead of along contour. The rain fed areas where rainwater is not readily observed
either due to high precipitation or low infiltration rate of water into the soil, graded bund is
recommended. A graded bund system is designed to dispose of excess runoff safely from
agricultural fields. The gradient can be either uniform or variable. The uniformly graded
bunds are suitable for areas where the bunds need shorter lengths and the runoff is low. The
variable graded bunds are required where bunds need longer lengths, owing to which
cumulative runoff keeps getting higher towards outlet. Variation in the grade are provided at
different sections of the bund to keep the runoff velocity within the desired limits to retard
erosion process.

Broad base terracing


A terrace is an earth embankment, constructed across the slope, to control runoff and
minimize soil erosion. A terrace act as an intercept to land slope, and divides the sloping
land surface into the strips. Broad base terraces are commonly used in western countries.
These types of terraces consist of ridge which has a fairly broad base and flatter slope so that
farm machinery can easily pass over the ridge. On these types of terraces, even ridge area is
cultivated and no land is lost to agricultural operation because of terracing. These are
basically two types, one is graded terracing and other one is level terracing. Such terraces are
constructed on the land slopes ranging from 5 to 15%.

Bench terracing
Bench terracing are constructed on hill sides which have slope gradients in the range of 15-
30%. Bench terracing transforms a steep land surface into a series of nearly levelled steps
across the slope of the land. These terraces convert the erodible sloping lands into farm lands
that are safe for cultivation. These are again classified as levelled and table top, sloping
inward, sloping out ward and Puertorican type bench terracing. The dimensions of these
terraces depends on the gradient of slope, depth profile of the soil, volume of the earth work
involved and the crops to be grown.

Vegetated waterways
Vegetated waterways are employed for safe disposal of runoff from field, terraced areas,
diversion channels, spillways and other structures. They are constructed in different shape
and dimensions depending upon the discharge, slope, soil type etc. Vegetated waterways are
used for the following purpose (Singh et al. 1990):
• Outlets for diversions and channels
• Outlets for surface and sub-surface drainage systems on sloping land
• Outlets for the farm ponds
• Outlets for the emergency spillways
• Dispose of water collected by road ditches or discharge through culverts
• Carry runoff from natural drains and prevent formation of gullies

77
6.3.2 Biological measures for erosion control
The measures used for controlling soil erosion through crops or vegetation and through
agronomic practices are designated as biological measures. These measures also assist in
increasing infiltration rate and thereby reduce runoff and overland flow through contour
farming, strip cropping, tillage practices, conservation cropping system, mulching and crop
residue management, critical area planting and contour vegetative hedges.
Good management practices such as crop rotations, maintaining the organic matter of the soil
(improved soil structure and tilth), increasing ground cover and variety forms of improved
tillage practices increases the soils resistance to detachment, increases infiltration, and reduce
runoff. The following biological measures should be considered when listing potential
development interventions:
Crop Rotation: A good crop rotation system should have the following characteristics:
 Crops, which are deep feeders, should alternate with shallow feeders. In this way, nutrient
removal occurs uniformly from the various soil layers rather than occurring in only one
layer,
 Crops of the same botanical family (such as potato and tomatoes) are likely to be attacked
by the same pests and diseases, and they should not normally follow each other in the
rotation. Even other crops which suffer from the same pests or diseases should not follow
each other in the rotation;
 One tilled cropped should be included in the rotation for elimination of weeds,
The number of field plots should correspond with the number of years of rotation.
Crop rotations vary with the land and the existing farming system. However, under rain fed
agriculture 3-year rotation of wheat/barley - sorghum/corn - chickpea/lentil or sorghum/corn -
chickpea/lentil - wheat/barley or chickpea/lentil - wheat/barley - sorghum/corn is
recommended.
Organic Matter: Actively decomposing organic matter in the soil added by plant residues
(roots and top growth) and manure, help to stabilize the soil aggregate. Due to critical
shortage of fuel wood and grazing the adoption of these practices by most highland farmers is
hardly, however, if the potential exists this management practice has to be promoted.
Minimum Tillage: It is a useful practice to reduce erosion and conserve soil moisture. It
was proposed for Ethiopia by the Ethiopian Highland Reclamation Study to use chemical
weedicides, in order to minimize excessive tillage operations. Additionally, it relieves
constraints associated with draught ox.
Stabilizing Physical Structures: leguminous trees (such as leuceana and Sesbania sp.) and
grass species used to stabilize the bunds and terraces besides to meeting household needs
(fuel wood, and construction and grazing), will also augment the fertility status of the soil.
However, it is important that the grasses used should not invade the farmlands to the extent
that they interfere with crop production.

78
Alley Cropping: It is an agro-forestry system in which food crops are grown in alleys
between rows of hedges. The hedges follow the contour and consist of trees and shrubs such
as Leucaena or Pigeon peas. It can be practiced for all slope range lands and soil conditions.
Grass Strips: a ribbon like band of grass laid out on cultivated land along the contour.
Usually, grass strips are about 1.0-meter-wide and spaced at 1.0m vertical intervals.
Recommended for all soil types but effective on land slope not exceeding 15%.
6.4 Water harvesting for soil conservation
Runoff from the land is one of the most erosive forms of water leaving the land with rills and
gullies. This runoff can be held on the soil surface and encourage to infiltrate.
Moreover, water harvesting involves optimal and immediate utilization of rain water, flood
water, stream or river water, where all of them are causes for erosion and sedimentation.
Hence, proper water harvesting techniques are not only efficient means of soil and water
conservation but also improve significantly local food security, income levels, crop yields
and standard of living of people in the area.
6.5 Production systems planning
Production systems planning (PSP) is a method of planning for the use of natural resources
under the watershed approach with a focus on ecological characteristics. It essentially
involves spatial allocation of land use for various production systems, namely, agriculture,
horticulture, and animal husbandry, by which conservation goals are met through better
decision-making. PSP is similar to regional land use planning, but it differs from it in that it
depends on ecological characteristics at the watershed level rather than activities at the
regional level. However, in both cases land use is an important element. The role of
production systems in soil and water conservation is evident from the water and soil losses of
catchments under such production systems such as given in table 11
Table 11: Soil and water losses under various production systems
S.NO Catchment coverage Loss of water Loss of soil (Kg/ha)
. (cm)
1 Forest with normal cover 1 1
2 Forest with poor ground cover 3 20
3 Well managed pasture 3 14
4 Agricultural crop 3 3250
5 Grasslands (pasture) 10 130
PSP is a conservation-planning tool useful for watershed management through allocation of
production systems to take care of soil and water conservation. The Indian Council for
Agricultural Research (1973) has proposed a three-tier production system based on the terrain
characteristics alone to serve as a guide for PSP.
Table 12: Terrain Based Production Systems Planning
S.NO Terrain category (slope) Production system

79
.
1 Gently sloping to flat (< 5%) Arable crops
2 Flat to moderate sloping (5- Horticulture crops
30%)
3 Steep (>30%) Forestry

7. WATERSHED PLANNING AND PROJECT FORMULATION


7.1 Scope of watershed planning
The watershed planning implies, the judicious use of all the watershed resources to achieve
maximum benefit with minimum loss/hazard to the natural resources i.e. land, vegetation and
water for the wellbeing of people. The planning should be carried out on the individual
watershed basis. The task of watershed planning includes the treatment of land by using most
suitable biological and engineering measures in such a manner that, the work must be
economical and socially acceptable.
7.2 Objectives and benefits of watershed planning
Objectives: The different probable objectives for watershed management planning may be
cited as under:
 To control damaging runoff and degradation and thereby conservation of soil and
water.
 To manage and utilize the runoff for useful purposes of watershed development
concern.
 To protect, conserve and improve the land of watershed for more efficient and
sustained production.
 To protect and enhance the water resources originating in the watershed.
 To check the soil erosion and reduce the effect of sediment yield on the watershed.
 To rehabilitate the deteriorating lands.
 To moderate the flood peaks at the downstream area.
 To establish watershed management practices and measures.
 To enhance the groundwater recharge, wherever applicable.
 To improve and increase the production of timbers, ranges, and wild life resources.
 To intensify agricultural extension activities.
Benefits: The benefits of watershed planning can be categorized in three aspects-
environmental, social and financial.
1.     Environmental Benefits:
 Improves quality of water for drinking and recreational use.
 Enhances water supply.
 Protects wildlife habitat and improves natural resources.
 Controls flooding by restoring riparian and wetland areas
2.     Community/Societal Benefits:

80
 Directly involves community members in developing a vision for the future of the
watershed.
 Provides opportunities to educate citizens on protecting and fixing the environment
that do not conflict with current and future development.
 Gives citizens an active voice in protecting and restoring natural resources that are
important to them.
 Provides opportunities to cooperate with neighboring communities.
3.     Financial Benefits:
 Reduces costs for meeting regulations and fixing damage that would happen if
sensitive areas are developed.
 Reduces costs for drinking water treatment.
 Improves availability of water for improving cropping intensity and thus the
production.
 Provides a new organization through which to get grants to improve the environment.
7.3 Developing steps of watershed planning
In order to achieve the different objectives selected for watershed planning, it is necessary to
go through the distinct steps:
 Recognition of problems.
 Analysis to determine the causes of watershed problem.
 Development of alternative solutions for the objectives formulated to solve the
problem.
 Selection of best solution.
 Application of selected solution.
 Protection and improvement of works, which have already been implemented.
The above steps can further be grouped in following four phases; i.e. recognition phase,
restoration phase, protection phase and improvement phase.
1.     Recognition Phase
Under this phase, the recognition of watershed problems, their probable causes and
development of alternatives for them, are described, which is carried out by conducting
several surveys such as:
a)     Soil survey
b)    Land capability survey
c)     Agronomic survey
d)    Forest lands under permanent vegetation survey
e)     Engineering survey
f)      Socio-economic survey
These surveys are made to ascertain the watershed’s problems, qualitatively and
quantitatively, to constitute a guide line for deciding the land treatment measures.

81
Furthermore, the compilation of these surveys and collected information are analyzed to
determine the nature of watershed’s problem, causes of problem and effect of the problems
on land unit as human beings, too. All this information obtained so make a basis to select
alternatives for rectification of problems and fulfillment of management objectives.
2.     Restoration Phase
This phase covers the task of selection of best solutions and their applications for watershed
management. In other way, this phase comes after recognized problems, in which treatment
measures are applied to critical areas for the recognized problems, identified earlier during
recognized phase, so that these critical areas can be restored to the pre-deterioration stages. In
forthcoming phase, the proper treatment measures, which will include the biological and
engineering measures, are implemented to all types of land falling under watershed.
3.     Protection Phase
It is third phase of watershed management, in which general health of watershed is taken care
of to ensure normal working. In addition to this, the protection of watershed against all those
factors which cause deterioration is also carried out. The protection is preferably made on the
critical areas, which are restored in the phase of restoration.
4.     Improvement Phase
This is the last phase, has precedential importance in watershed management work. Under
this phase, the overall improvements made during management of watershed are evaluated
for all the lands covered. In addition, attention should be given to make improvement on
agricultural land, forest land, forage production, pasture land and socio-economic status of
the people.
7.4 Formulation of watershed project
Formulation of watershed projects involve careful analysis of available resources, defining
the problem, formulation of objectives, steps wise work plan to achieve the objectives within
defined time and optimum available budget. Detail of these aspects are presented in brief as
below.
Definition/Description of Problem

The problems such as: flood, drought, erosion and sediment damage and other problems
related to the conservation, development, utilization, disposal of water originating in the
watershed etc. are considered under this section. Major problems are outlined as under:

Flood Damage:  The following points are considered to evaluate the flood damage occurred
in a watershed
1. Amount and value of land improvements and other properties exposed to the flood
hazards in the watershed.
2. Frequency of flood occurrence.
3. Significance of small frequent floods or large infrequent floods in total flood
problems.

82
4. Limitations
Sediment Damage: The problems exposed by sediment deposition are considered in
following cases:
1. Problems of reservoir sedimentation
2. Problems of channel silting
3. Drainage problem
4. Irrigation development
5. Loss of agricultural land
Erosion Damage: The problems of erosion damage are studied under the following contents:
1. Extent of sheet, gully and channel erosion.
2. Downstream damage due to sediment deposition.
3. Effect on agricultural production due to erosion.
4. General effects on watershed’s economy.
Water Management Problem: It includes the detail on irrigation needs, drainage, water
supply required for agriculture and non-agricultural uses and other management needs.
Special Problems: The problems such as: land slip, land slide, highway erosion, mines etc.
are counted for preparation of watershed work plan.  
Stepwise Work Plan
Main proposal is divided in different sections.
Section-I
In this section, a brief report about project area is cited, which includes following details:
1. General features
2. Demography
3. Economy
4. Geology
5. Climate
6. Water resources: surface and subsurface water rights and laws.
7. Land resources: soil types, chemical and physical properties of soil and land use
capability classification.
Section-II
In this section, the present status and development potential of the area are explained, which
are outlined with the help of following details:
a)    Present Status
1. Power supply
2. Land use
3. Agricultural production and availability of inputs such as, seeds, fertilizers, money etc.
4. Government policy
a. Incentives
83
b. Financial institutions
5. Marketing facility
6. Infrastructure for transport
7. Growth rate of traditional agriculture
b)    Future Requirement
1. Land preparation
2. Irrigation and drainage requirement
3. Reclamation of saline and alkali soils
4. Farm equipment and supply
5. Land reforms required
c)     Development potential
1. Potential according to land use
2. Aerial photograph for project planning
3. Land use capability
4. Economics of alternative farming methods.
Section-III
a)    Preparation of Development Plan
1. Justification
2. Guideline and concept
3. Objectives and scope of the plan
4. Priorities
5. Economic constraints
6. Stage of development
b)    Main Program
1. Land development
2. Irrigation and drainage
3. Soil conservation measures
c)     Step to be Recommended for Socially Acceptance of Proposal
d)    Evaluation
1. Putting of hydrologic measurement stations
2. Analysis of data
e)     Monitoring of Infrastructures
f)      Development Schedule
Section-IV
Cost Estimation: Capital cost, annual cost, foreign exchange requirement and equivalent
annual cost are considered.

84
Section-V
In this section, the benefits are computed from following sources:
1. Improvement in water quantity and quality
2. Increment in agricultural production
3. Environmental control and recreation
4. Enhancement of economy of area
Section-VI
Economic Analysis
1. Criteria
2. Project cost
3. Tangible and intangible benefits
4. Agricultural and other benefits
5. Benefits-cost analysis
6. Equivalent annual benefit
7. International rate of return
Section-VII
Financial Analysis
1. Cost allocation
2. Payment capacity
Section VIII
Program Implement technique
Section IX
Conclusion and Recommendation

8. Evaluation of Watershed Programs


8.1 Scope of Watershed Program Evaluation

Evaluation is an important aspect of watershed programs. It is a multi-dimensional task which


is generally performed at different times during the implementation of watershed programs.
Until recently watershed program evaluators tended to favor either a quantitative or a
qualitative evaluation.  Typically, quantitative evaluations reflect a simplistic view that
reality takes a single form that can be perceived and measured objectively. On the other hand,
qualitative evaluations reflect a more constructive view, implying that reality can have
multiple versions.

There is a rising interest in mixing both the qualitative and quantitative methods of watershed
program evaluation. This comes from the fact that purely quantitative and purely qualitative
approaches to watershed program evaluation both have limitations. The strengths of each
evaluation often compensate for the weaknesses of the other evaluation.

85
Quantitative Evaluation of Watershed Programs 

The quantitative evaluation of watershed programs attempts to attribute changes in various


outcome variables to a project intervention (i.e., ‘treatment’) and determine whether such
effects are statistically significant.  An experimental approach is often considered as an
acceptable standard for quantitative evaluation of watershed programs. Yet, in many cases the
results of such a study may not extrapolate beyond the watershed projects examined. 

There are many situations wherein an experimental approach to quantitative watershed


program evaluation may not be possible. In such situations, various approaches have been
used, each with their own strengths and limitations. 

The first approach is called a “before/after” study. The evaluator measures the levels of
outcome indicators in a watershed area before and after a watershed treatment. This is a fairly
weak but feasible approach that involves an unlikely assumption that there have been no
other significant changes during the study period. 

A second approach consisting of a “with/without” study, is useful when no baseline data are
available. This is often the case when an evaluation is commissioned after a watershed
project has been implemented.

Cost-benefit analysis has long been the method of choice in economic appraisal of
agricultural development and irrigation projects.  Cost-effectiveness analysis is similar but it
estimates only the costs of alternate approaches of achieving a given objective.  Cost-benefit
analysis aims to evaluate costs and benefits that occur with a project and compare them to
what would happen without the project.  Even if all costs and benefits could be identified and
valued, cost-benefit and cost-effectiveness analysis would give only a single assessment of
overall project performance. However, watersheds consist of multiple users who are affected
differently by the project. A favorable benefit-cost ratio could temporarily mask uneven
distribution of benefits, yet those who do not benefit may be in a position to undermine the
project.

Thus, there are clearly multiple challenges associated with using quantitative evaluation
methods for evaluation of watershed projects. Most challenges are introduced by the fact that
watershed projects are not amenable to the same controlled conditions as in the experiments
which provide the data for a simplistic analysis.   

Qualitative Evaluation of Watershed Programs    

In contrast to quantitative evaluation, qualitative evaluators typically place less emphasis on


measurement and more on context and on understanding the subtle manifestations. In general,
a qualitative approach tends to be flexibly structured and uses open-ended questions in an
inductive fashion. The objective is not to obtain a numerical estimate of some phenomenon,
but to develop an in-depth understanding of an issue by probing, clarifying, and listening to
stakeholders talk about a topic in own words. The in-depth nature of the qualitative approach
means that a study’s scale is usually smaller than that found in quantitative research.

As with quantitative evaluation of watershed programs, sampling issues in qualitative


evaluation also raise questions about biases in data. While quantitative researchers use
random sampling whenever possible, qualitative researchers use several strategies to increase

86
the internal validity of their findings. In qualitative evaluation, data collection and analysis
become inseparable; as such researchers collect much of the data themselves, rather than
relegating this task to field assistants.

Mixed Evaluation of Watershed Programs

Researchers use mixed evaluation of watershed programs for various reasons. Here,
qualitative and quantitative components may be used either sequentially or in parallel or in an
integrated fashion.  When qualitative and quantitative components in a mixed evaluation are
used in an integrated manner, the information and data collected from one activity is used for
the other activities of the evaluation process also.

8.2 Indicators for Watershed Program Evaluation

Watershed program evaluation can be quantified in terms of certain indicators.  These


indicators are the measures of targets or goals of the watershed project implementation,
which facilitate the expected positive change in the watershed projects. They also give an
insight into and quantify the process of evaluation. The various indicators generally used for
watershed program evaluation are discussed in the following
sections:                                             

i) Technical Indicators

Technical indicators in Watershed Program Evaluation include the extent of soil loss and
runoff, amount of discharge in the stream and amount of sediments in flowing water at the
outlet point, increase in the yield of wells and rise in water table, average annual water flow
and flood peak, changes in soil moisture, concentration of suspended sediments, annual
sediment yield, turbidity of water, biological and chemical properties of water, pH, annual
reservoir sedimentation, pesticide concentration, etc.

ii) Common Property Resources (CPR) Use Indicators

Common property resources (CPR) use indicators are productivity of crop, fodder, fuel wood,
pasture land, community forest land and milk. Further information to be collected are areas of
managed agro-forestry, protected degraded forest land by social fencing, unprofitable
cropland and grazing land, unused area with agro-forestry and areas of common property
resources.

iii) Institutional Building and Community Organization Indicators

These indicators include the number of rural development institutions in the watershed and
the coordination among them, financial independence of the institutions, their capacity
building to solve managerial, administrative  and  financial problems, the number of trained
professionals assigned to the project, the number of welfare  and development programs
performed by the institutions, the number of farmers trained in soil conservation and modern
agriculture techniques, the percentage of population willing to adopt appropriate technology
to improve crop, livestock, water harvesting, etc., and the performance of self-help groups,
user groups and watershed development committees (WDCs).                                                          
·

87
iv) Ecological Improvement Indicators

Ecological improvement indicators include the biodiversity and biomass indices, severely
eroded, overgrazed and over-utilized lands, wastelands, lands under shifting cultivation,
stabilized slopes, areas of treated ·gullies, number and depth of gullies, soil fertility and
organic matter content of soil.

v) Economic and Social Indicators

These indicators quantify the change in the living standards, .household savings, household
expenditure and household income, number of families living above poverty line (APL) or
below poverty line (BPL), extent of migration to urban areas in search of .employment and
indebtedness in cash or kind, prevailing wage rate in agriculture and non-farm sectors,
changes in crop production, double cropped areas, agricultural and non-agricultural land
values, number of annual man days generated, number of working women and young people
per year, time spent in fetching and collecting drinking water, annual request for technical
assistance and skill up gradation of rural artisans.

vi) Essential Service Indicators

These indicators include the literary rate, number of schools in operation, percentage of
school attending children and their  age, number  of  primary  school dropouts, percentage of
houses having electricity connection and drinking water facilities, number of dispensaries in
operation per year and the families receiving medical care, annual mortality, percentage of
population of age group 0-16 years receiving immunization, couples protected under family
planning, annual birth rate, number of annual sterilizations, length of motorable road added
per year in kilometers, level of child malnutrition below 1 year age group and availability of
essential commodities.

8.3 Stages of Watershed Program Evaluation

It is a common practice to carry out the watershed program evaluation in four stages with the
help of the six indicators mentioned earlier. These four stages are discussed here:

i) Baseline Evaluation

This is the evaluation in the initial planning stage. The data on the indicators are used as
benchmark for evaluation. A reliable baseline data on hydro-meteorological, economic,
social, physical and biological parameters are provided for this evaluation.

ii) Mid-term Evaluation

This evaluation is done in the middle of the watershed program implementation. In this stage
of evaluation, initial problems in the planning are overcome and the flow of inputs to the
target population is commenced and their response can be observed.  The purpose of such
mid-term ·evaluation is to check on ·the effectiveness of each individual activity. This
evaluation quantifies the short and mid-term benefits of the project.

iii) Terminal Evaluation

88
This evaluation is done at the end of the project economic life. It indicates the efficiency of
project implementation, accuracy of the project estimates, etc.

iv) Post-Terminal Evaluation

This evaluation is carried out after 5 to 15 years of watershed program period. Long-term
effects and impacts become visible in this post-terminal evaluation.

Impact of Evaluation on Watershed Management

The evaluation of watershed management during a normal year and a year under stress
conditions is very difficult and complex. Developmental works in a year under stress decline
significantly. Therefore, the aim of the watershed management should be to focus on
utilization and harnessing of existing resources for the maximum production and benefits.
One of the main thrusts of watershed management programs should be to minimize the
differences in the benefits during a normal year and a year under stress, as far as possible.

89

You might also like