You are on page 1of 95

Technical Guidelines for the

Construction, Rehabilitation of Drilled


Water Wells

January 2020

This document represents the official guidelines on the issue of the construction and
rehabilitation of boreholes. It was compiled through a collaborative effort led by the
Somalia Wash Cluster, with key inputs from sector stakeholders including, but not limited
to, UNICEF, GSA, STC, WOCCA, ACTED and the Ministry of Energy and Water Resources.
Glossary
Anthropogenic – man-made
Aquifer – an underground layer of water-bearing material from which groundwater can be usefully
extracted via a water well.
Borehole – generally used to refer to a small diameter water point constructed by drilling.
Cone of depression – when water is pumped from a well, the water table (in the case of an
unconfined aquifer) or piezometric surface (in the case of a confined aquifer) near the well
is lowered. This area is known as a cone of depression. The land area above a cone of depression is
called the area of influence.
Confined aquifer – has a confining layer of clay or low-permeability rock that restricts the flow of
groundwater from one formation to another. A confined aquifer is thus not in direct contact with
the atmosphere.
Derogation is the effect of pumping a well on the seasonal flow from springs, the drawdown in
nearby wells, or the drying of wetlands.
Drawdown refers to water-level lowering caused by groundwater pumping.
Hydraulic conductivity (K) is the rate of movement of water through a porous medium (e.g. soil or
an aquifer). It is defined as the flow volume per unit cross-sectional area of porous medium. The
units are usually m3/m2/day or m/d. Sometimes hydraulic conductivity is referred to as permeability
but permeability refers to all fluids, not just water.
Geophysical surveys (or techniques) measure the physical properties of rocks (resistivity,
conductivity, magnetic fields and sonic properties). The measurements are interpreted in relation to
geological features that are expected to facilitate groundwater storage and movement.
Interference – the effect that pumping from a well has on the drawdown in neighboring wells.
Impermeable material – a material such as clay through which water does not readily flow.
Permeability – see hydraulic conductivity.
Potentiometric surface is the level to which water in a confined aquifer will rise in a well. In an
unconfined aquifer the potentiometric surface is the water table.
A rock formation is an identifiable body of natural earth material. It may be unconsolidated (e.g.
loose sand or gravel) or consolidated (e.g. a sandstone or granite).
Storativity is the amount of water that can be removed from the aquifer for a given lowering of
water level.
Transmissivity is the product of hydraulic conductivity and saturated aquifer thickness.
An unconfined aquifer (also known as a water table or phreatic aquifer) does not have a confining
layer which separates it from the surface. In other words, the upper boundary is the water table, or
phreatic surface. The aquifer is in direct contact with the atmosphere.
The unsaturated zone is the formation in which water occurs but all the pores are not completely
filled (saturated) with water. This zone is above the water table.
A water table is the free water surface in an unconfined aquifer.
Well is either used to refer to a hand-dug shaft, or it is used more generically to mean any small-or
large-diameter vertical groundwater abstraction point other than a spring, regardless of method of
construction.
A well field is a cluster of wells supplying water for a large-scale need such as a town, an irrigation
scheme, or a refugee camp.

2
Abstract
Groundwater is frequently chosen as the most suitable source of drinking water, supplies of which
are brought to the surface by rehabilitating existing boreholes or drilling new ones. However,
constructing, or repairing, boreholes requires specialized knowledge and technical expertise, much
of the specialized knowledge and technical expertise needed for this purpose can be gained from the
standard literature. However, field operations in remote areas or in difficult conditions often require
flexibility and imagination in avoiding or solving technical problems. These guidelines are intended
mainly as a practical tool and therefore contain a minimum of theory. They are aimed at water and
habitat engineers working in the field who are undertaking or supervising borehole drilling or
rehabilitation programs. The end result should be a cost-effective facility capable of supplying
drinking water for many years.

3
1. Introduction
1.1 Background
Groundwater is undeniably a very important freshwater resource for drinking water supply,
irrigation and industry, in addition to its natural role of sustaining river flow and aquatic ecosystems.
Sustainable groundwater development is fundamental in order to provide universal access to safe
drinking water. The lack of understanding of groundwater resources in much of Somalia undermines
its potential to contribute to poverty reduction and economic development, and threatens its
environmental sustainability.

Over the past two decades, Somalia has witnessed a significant increase in drilled water wells, or
boreholes. These are financed by development programs as well as investments by water users and
local businesses. Use of groundwater in Somalia is on the rise; use of groundwater for irrigation in
Somalia is expected to grow significantly.

Boreholes are one of the best means of obtaining clean water in field conditions. However,
constructing, or repairing, boreholes requires specialized knowledge and technical expertise, much
of which can be gained from the standard literature; but field operations in remote areas or in
difficult conditions often require flexibility and imagination in avoiding and solving technical
problems. If boreholes are not properly sited, designed and constructed in the first place, supplies
cannot be maintained, and investments are wasted. And in the longer term, if groundwater
resources are not properly managed, there is risk of over-abstraction and pollution and massive
failure of drinking water supplies. This cannot be allowed to happen.

There is growing evidence of major weaknesses in how borehole drilling initiatives are undertaken in
Somalia. Over 25 major implementing organizations are actively involved in building and funding
water points in Somalia. These are in addition to smaller national nongovernmental organizations
(NGOs), local companies, local communities, and private persons who are also doing the same.
Different agencies use different designs and modes of construction without any clear guidelines or
standards. If there is to be any chance of meeting the Sustainable Development Goal (SDG) target for
drinking water, this needs to change. As the main UN agency supporting the SDG drinking water
target in rural and peri-urban areas, projects need to demonstrate professional groundwater
development and support effective groundwater management.

It is against this background that Somalia National WASH Cluster decided to publish Technical
Guidelines for the Construction and rehabilitation of Boreholes in Somalia. The use of these
guidelines should harmonize borehole construction practices among all stakeholders operating in
the country in line with international best practices.

This technical guideline is aimed at project coordinators, water engineers, and technicians. It is
intended to be of assistance to everyone, from planners in offices to on-site personnel, in the
making of technically correct and cost-effective decisions in the field when the drilling or
rehabilitation of boreholes is required. An attempt has been made to orient the contents towards

4
problems that might be encountered in the field. Nevertheless, some consideration of theoretical
information has been necessary, because engineers will not be able to function without it. The
authors hope that they have struck a balance between the practical and the theoretical, a
combination that is required in professional water engineers.

This guideline builds on existing guidelines in use by various organizations in Somalia. All
stakeholders operating in the country are expected to adhere to the provisions of the guideline. The
guideline set out the minimum standards that are expected in the construction of borehole for
village water supply, self-supply, and semi collective water supply. Ensuring compliance with the
guideline is the responsibility of the Government. Some of the provisions of the Technical Guideline
document are mandatory. However, due to the variety of local conditions and socioeconomic
circumstances across the country, as well as the need for communities and households to select
options that work best for them, some provisions are recommendations, and their adoption is left to
discretionary judgment.

The guideline begins with an overview of the benefits of utilizing groundwater and a consideration of
various drilling methods, techniques are compared and details of the drilling equipment associated
with each are provided to assist the user in selecting appropriate equipment. The guideline focuses
on mud and air rotary drilling, as they are the most common methods of borehole drilling found in
the Somalia. Details on borehole construction, design and development using these two methods
are found in chapter two. Key factors influencing borehole deterioration and aspects of monitoring
and maintenance are outlined in chapter four. When borehole deterioration reaches a stage where
production is severely hampered, rehabilitation becomes unavoidable: this subject is treated in
chapter five.

1.2 Groundwater and the Advantages of Boreholes


Easy access to safe, potable water is a basic human need, important for health and quality of life.
Groundwater is of huge importance in Somalia. Apart from the areas along the Juba and Shabelle
Rivers, all regions depend on groundwater for domestic water supply, livestock and small scale
irrigation. There is very low effective rainfall and no perennial surface water across most of the
country. Groundwater is accessed through boreholes, shallow wells and springs. Most boreholes are
between 90 m and 250 m deep, but in some areas reach over 400 m deep and most of the shallow
wells are less than 20m deep.

The successful and sustainable development of groundwater resources in Somalia is critical for
future safe water supplies, economic growth and food security in the country. Doing this successfully
relies on good hydrogeological understanding - but much of the data and information that already
exists about groundwater in Somalia is not available to the people who could make use of it.

1.2.1 Exploiting Groundwater

The principal source of inland groundwater is rainfall. A proportion of rain falling on the ground will
percolate downwards into an aquifer if the conditions are right. A great deal of rain water ends up as
run-off in streams and rivers, but even here there is often a direct hydraulic connection with a local
aquifer. Indeed, in arid areas like Somalia with ephemeral streams, high groundwater levels may be
able to sustain surface flow along drainages. It is obvious that a hole dug or bored into a saturated

5
‘sponge’ will release water from storage. This water can be sucked or pumped out, and all being
well, more water will enter the hole to replace that which has been withdrawn. This is the basic
principle behind a water borehole.

- Geological Constraints

The Earth’s crust has often been compared to a sponge, in that it can soak up and hold water in pore
spaces, fractures and cavities. This ability to store water depends very much upon geological
conditions and on the host formation. For example, fresh, un-fractured, massive granite – a
crystalline rock – has virtually no space available for water, whereas unconsolidated, or loose, river
gravel and highly weathered cavernous limestone can store large quantities of groundwater and are
capable of releasing it relatively freely. Sandstone and mudstone may be able to hold significant
groundwater resources, but because of differences in grain size – and hence porosity – will release it
at different rates. One may be a good aquifer, the other a poor one. The rate at which water flows
through a formation depends on the permeability of that formation, which is determined by the size
of pores and voids and the degree to which they are interconnected. Permeability and porosity
should not be confused, porosity being the ratio between the volume of pores/voids to the bulk
volume of rock (usually expressed as a percentage).

The three principal characteristics of aquifers are transmissivity, storage coefficient, and storativity.
Transmissivity is a means of expressing permeability, the rate at which water can flow through the
aquifer fabric. Storage coefficient and storativity express the volume of water that can be released
from an aquifer.

Hydrogeology is the science of groundwater, and it is the job of a hydrogeologist to assess the
groundwater resources in any given area. This is accomplished through the use of maps
(topographic, geological), satellite images, aerial photos, field observations (geological mapping,
vegetation surveys, etc.), desk studies (literature, field reports, etc.) and ground geophysics. Ground
geophysical surveys are now quite effective in locating water-bearing formations at depths down to
around 100 metres. Methods include resistivity (vertical electrical profiling), natural-source self-
potential and electromagnetic methods (such as VLF), magnetic methods, and micro-gravity
surveying.

- Borehole siting

Choosing a borehole site is a critical part of the process of providing a safe and reliable supply of
groundwater. The best sites are those in which catchment (natural water input) may be maximized.
Such locations are not necessarily those that receive the highest rainfall (which may occur in upland
watersheds). ‘Bottomlands’ – such as river valleys and lake basins – tend to be areas of maximum
catchment as both surface water and groundwater migrate towards them under gravity. Fracture
zones, although not always directly related to bottomland, can also be good reservoirs for
groundwater, and may be located by ground observation or satellite images/aerial photographs, and
by geophysical methods.

Another aspect of borehole siting that demands careful consideration in populated areas is the
potential for contamination by livestock and pit latrines or other waste disposal facilities. Because
near-surface groundwater migrates downslope, a shallow dug well or a borehole tapping shallow

6
groundwater should be sited as far away as possible (while bearing in mind the human need for
proximity to a source of water) and upslope of potential sources of pollution (latrines or sewage
pipes, for instance). Deeper aquifers confined by impermeable layers are at less risk of
contamination from surface pollutants. One final consideration is the nature of the shallow aquifer.
If the host formation is made of fine or medium-grain-sized sand, it will act as a natural filter for
particulate pollutants, whereas fissured limestone, with a high rate of water transmission
(transmissivity) will carry away pollutants faster and to greater distances from the source. It is
estimated as a rule of thumb that most microorganisms do not survive more than 10 days of
transportation by underground water.

- Types of geological formation

Figure 1.1 shows a hypothetical geological situation in which different sources of groundwater may
(or may not) be tapped by dug wells or boreholes.

Figure 1.1: A hypothetical hydrogeological scenario

A) Perched aquifers:

At site A, a shallow dug well may provide a little water from a ‘perched’ aquifer in the weathered
zone above relatively impervious (low porosity) mudstones. If this well was extended into the
mudstones it might produce very little additional water. A perched aquifer is normally limited in size
and lies on an impervious layer higher than the area’s general water table.

B) Shallow unconfined aquifers:

The term ‘unconfined’ refers to an aquifer within which the water is open to atmospheric pressure:
the so-called piezometric surface (pressure head level) is the same as the static water level (SWL) in
the borehole. At site B, a borehole extracts water from an unconfined sandstone aquifer, the SWL of
which is somewhat lower than the level of flow in the river. This sandstone aquifer is in a good
catchment area because of recharge from the river.

C) Confined aquifers

7
A ‘confined’ aquifer may hold groundwater under greater pressure, so that when punctured by a
borehole, the SWL rises to the higher piezometric level. If the piezometric surface happens to be
above ground level (which is not uncommon), water will flow out of the borehole by itself: this is
known as ‘artesian water’. Deep borehole C intersects the sandstone aquifer and a deeper confined
aquifer in fissured limestone; because of overpressure in the limestone aquifer, the SWL in C may be
at the same or higher elevation than in B. The limestone aquifer may have no source of
replenishment; so the water in it is ancient, or ‘fossil,’ and could be exhausted if over-exploited.

D) Fracture zone

Borehole D, which has been drilled into fractured granite (shaded area), finds water held in the
fracture zone. Fracture zones develop during geological times as a result of the severe mechanical
stress, caused by tectonic movements, that is exerted on non-plastic formations.

E) Hydrogeological basement

Site E, a borehole sunk into massive granite on top of a hill, is dry. In this situation, it would be a
waste of time and money to extend a deep borehole (such as C) into the metamorphic basement,
which is generally known as the ‘hydrogeological basement’ or ‘bedrock.’ The bedrock marks the
level below which groundwater is not likely to be found.

1.2.2 Groundwater Extraction

A water borehole is not just a hole in the ground. It has to be properly designed, professionally
constructed and carefully drilled. Boreholes for extracting water consist essentially of a vertically
drilled hole, a strong lining to prevent collapse of the walls, which includes a means of allowing clean
water to enter the borehole space (screen), surface protection, and a means of extracting water.
Drilling by machine is an expensive process, and boreholes require professional expertise for both
their design and their construction. There are, however, compensations: this method of extracting
water has a number of significant advantages.

The common alternatives to drilled boreholes, available to everyone with basic knowledge and
simple tools, are surface water sources, springs, and dug wells. Where shallow groundwater
emerges at a seepage site or at a spring, water catchment systems can be constructed to provide
water of reasonable quality. Catchment, that include sand or stone filters, and collector sumps are
extremely effective means of collecting water. Gravity may be used to effect pipe network
distribution from upland springs. Shallow dug wells usually exploit near-surface groundwater. Wells
down to a depth of five metres are relatively simple to construct (given time and willing local
labour), and there are many publications describing this process. Furthermore, because of their
relatively large diameter, wells provide valuable storage volume. The water supply can be protected
by lining the well, covering it with a lid, and fitting a hand-pump to it.

However, dug wells, and surface water catchment in general, are very vulnerable to contamination
caused by agricultural activities, animals, poor sanitation and refuse. In addition, surface or shallow
groundwater catchment is vulnerable to poor rainfall and declines in water level caused by drought,
because it usually taps the top of the aquifer. Borehole water, by contrast, often requires no
treatment and is less susceptible to drops in water level during periods of drought or limited rainfall.

8
- Advantages of drilled boreholes

If they are properly designed and maintained, drilled boreholes:

 Are less vulnerable to drought or drops in water level when drilled into deep water-
bearing formations.
 Can be designed to exploit more than one aquifer (when individual aquifers are
vertically separated and not hydraulically connected)
 Are less vulnerable to collapse
 Are less vulnerable to contamination
 Are, if properly sited, capable of producing large yields; so, mechanically or
electrically powered pumps can be used
 Are amenable to quantitative monitoring and testing, which enables accurate
assessment of aquifer parameters (as in aquifer modeling), water supply efficiency,
and optimal design of pump and storage/distribution systems
 Can be used to monitor groundwater levels for other purposes, e.g. environmental
studies or waste disposal

- Disadvantages of drilled boreholes

 High initial material costs and input of specialized expertise, i.e. construction,
operation, and maintenance may require skills and expensive heavy equipment.
 Vulnerable to irreversible natural deterioration if inadequately monitored and
maintained
 Vulnerable to sabotage, can be irreparably destroyed with little effort if
inadequately protected
 Require a source of energy if water extraction pumps are used (unlike gravity feed
systems)
 Do not allow direct access, for maintenance or repairs, to constructed parts that are
underground

1.3 Hydrogeological Classification of Aquifers in Somalia


Groundwater is of huge importance in Somalia. Apart from the areas along the Juba and Shabelle
Rivers, all regions depend on groundwater for domestic water supply, livestock and small scale
irrigation. There is very low effective rainfall and no perennial surface water across most of the
country. Groundwater is accessed through boreholes, shallow wells and springs. Most boreholes are
between 90m and 250m deep, but in some areas reach over 400m deep. Most of the shallow wells
are less than 20m deep.

Yields vary from one aquifer to another, but most shallow wells yield between 2.5 and 10m³/hr,
compared to the typical range of borehole yields of between 5 to 20m³/hr. Groundwater quality is a
major issue. Most groundwater sources have salinity levels above 2,000µS/cm. Many of the shallow
wells are also unprotected and vulnerable to microbiological and other contamination.

9
The knowledge of the hydrogeological classification of aquifers in the various parts of Somalia and
the quantity and quality of the groundwater in the various aquifers is vital, for making an informed
decision on the design of the water treatment system to be used (Annex 1). According to
hydrogeological information, Somalia’s hydrogeological units are divided into three main groups:

1. Porous rocks of relatively high to low hydrogeological importance


2. Fractured rocks of relatively medium to low hydrogeological importance, (this unit is
characterized by local aquifers restricted to fractured zones. It could be unconfined or
confined. The permeability varies and it is generally low. The water quality is generally good.
Thermal saline waters may occur. Its relative importance is medium to low and the potential
is generally low).
3. Porous or fractured rocks with very low hydrogeological importance

Practical problematic situations that are related to groundwater in different parts of Somalia are
expressed in terms of high salinity, fine sand, loss of circulation, running sand & caving, thick mud-
stone and other problems that will occur during drilling of boreholes like the presence of boulders.

Figure 1.2: The geology & hydrogeology map shows a simplified overview of the geology and the
main aquifers in Somalia.

10
1.4 Methods of Drilling Wells
Once a suitable site has been selected and borehole drilling decided on, the proper drilling method
must be chosen.

Another primary consideration in project planning is the availability of existing water sources and
water points. There may be completed dug wells and boreholes already in the area. Are they in use?
If not, can they be rehabilitated to augment water availability or to reduce the cost of the program?

Drilling equipment, such as compressors, can be used to bring disused boreholes back into use; the
question of rehabilitation will be addressed in this guideline. This section outlines the factors that
must be considered when choosing a drilling method.

1.4.1 Common Drilling Methods

Essentially, a drilling machine consists of a mast from which the drilling string components (tools
plus drill pipes or cable) are suspended and, in most cases, driven. Modern systems are powered
rotary-driven, but it is probably worth a short digression to describe some methods of manual
drilling for water. Simple, low-cost methods include:

- Hand-auger drilling

Auger drills, which are rotated by hand, cut into the soil with blades and
pass the cut material up a continuous screw or into a ‘bucket’ (bucket
auger). Excavated material must be removed and the augering continued
until the required depth has been reached. Auger drilling by hand is slow
and limited to a depth of about 10 meters (maximum 20 meters) in
unconsolidated deposits (not coarser than sand, but it is a cheap and
simple process.

- Jetting

A method whereby water is pumped down a string of


rods from which it emerges as a jet that cuts into the
formation. Drilling may be aided by rotating the jet or by
moving it up and down in the hole. Cuttings are washed
out of the borehole by the circulating water. Again,
jetting is useful only in unconsolidated formations and
only down to relatively shallow depths, and would have
to be halted if a boulder is encountered.

- Sludging

This method, which may be described as reverse jetting, involves a


pipe (bamboo has been successfully employed) being lowered into
the hole and moved up and down, perhaps by a lever arm. A one-
way valve (such as someone’s hand at the top of the pipe) provides
pumping action as water is fed into the hole and returns (with

11
debris) up the drill pipe. There may be simple metal teeth at the cutting end of the pipe, and a small
reservoir is required at the top of the hole for recirculation. The limitations of sludging are similar to
those of the previous two methods.

- Percussion drilling

Drilling by percussion is done by simply dropping a heavy cutting tool, of 50 kilograms or more,
repeatedly in the hole. The drilling tools are normally suspended by a rope or cable; and – depending
on the weight of the drill string, which, for manual operation, is obviously limited – it is possible to
drill to considerable depths in both soft and hard formations. Basic percussion drilling systems are
still widely used in Abudwaq and Galgadgob districts of Somalia to drill shallow boreholes. They
consist of a strong steel tripod, cable and power winch, percussion tools, and a baler.

These systems are seriously hindered when the ground is hard, and can accidentally change
direction along weaker zones, causing boreholes to become crooked or tools to jam. Unconsolidated
materials, although easy to drill with cable tool, become very obstructive when boulders are
present. Sticky shales and clays are also difficult to penetrate with cable tool rigs, and loose sand
tends to collapse into the hole almost as fast as it can be bailed. These manual shallow drilling
techniques might be used as low-cost alternatives in groundwater investigations for dug well sites,
particularly if geophysical surveys prove to be ineffective, unavailable or impracticable because of
ground conditions. In such instances, when the drilling is done solely for the purpose of prospecting,
only small holes are drilled, rapidly.

- Rotary drilling

Most borehole applications in the field will require rotary drilling. True rotary drilling techniques
allow much deeper boreholes to be constructed, and use circulating fluids to cool and lubricate the
cutting tools and to remove debris from the hole.

Figure 1.3: A mud and Air rotary machine (Combine DANDO Rig)

12
Circulating fluids usually take the form of compressed air or of pumped water with additives, such as
commercial drilling muds or foams.

Table 1.1: Comparison of drilling methods

Drilling Methods Advantages Disadvantages


Manual construction Simple technology using Shallow depths only
(Hand dug wells and hand cheap labour
drilling)
Percussion drilling Simple rigs, low-cost Slow, shallow depths only
operation
Rotary drilling, direct Fast drilling, no depth limit, Expensive operation, may need large
circulation needs no temporary casing working space for rig and mud pits,
may require a lot of water, mud cake
build-up may hamper development
Rotary DTH, air circulation Very fast in hard Generally not used in soft, unstable
formations, needs no formations, drilling depth below
water, no pollution of water table limited by hydraulic
aquifer pressure
Rotary, reverse circulation Leaves no mud cake, rapid Large, expensive rigs, may require a
(not described in text) drilling in coarse lot of water
unconsolidated formations
at large diameters

1.4.2 Drilling Equipment

Once a drilling method has been selected, you must decide on the type of drilling equipment or rig
that best suits your situation. This section discusses the various types of rig available and their
suitability and also provides an overview of drilling rig parts.

- Choosing a drilling rig:

The type of rig chosen may be determined on the basis of the site geology, the anticipated depths of
the boreholes, and their expected diameters. Access is an important consideration. All drilling
machines, except the smallest units capable of being dismantled and reassembled on site, require
transportation: a road may have to be cut through bush to reach a location. For the largest truck or
trailer-mounted rigs this can be a significant problem during rainy seasons in remote areas. Heavy
rigs are notorious for becoming stuck in mud, and in such difficult conditions they should be used
only if rain is not expected, or if there are means of pulling the rig out of trouble.

- Drilling rig components


a) Drill bit

No single type of drill bit can cope with all possible drilling conditions and formations. Some typical
examples are shown in Figure 1.4: Drag bits consist of three or four serrated blades that shear the
formation when the bit is rotated; they can penetrate softer formations such as poorly consolidated
or stiff clays and mudstones rapidly.

13
Roller cone bits (or tricone rock bits), which can be used with air or liquid flushing, are popular with
the oil industry. They can be used to penetrate both soft and relatively hard formations.

Figure 1.4: Two common types of drill bits: left drag bits; right, roller cone bits

b) Hammer

In air-circulation drilling, if a formation is too hard for penetration by a drag bit, a DTH hammer is
generally employed. This tool was developed for the mining and quarrying industry. The ‘business’
end – the button bit – is studded with hemispherical tungsten carbide ‘buttons,’ and with channels
built in to allow the passage of compressed air. When the hammer is pressed against the ground, the
bit is forced into a pneumatic hammer action (like a road drill) by compressed air fed down the drill
pipes. Then, as the tool is rotated in the hole, the buttons act across the entire base of the borehole.
Most hammers rotate at speeds of 20 to 30 revolutions per minute, and blows can be struck at rates
of up to 4000 per minute. Debris is normally flushed (blown) out of the hole at the end of each drill
pipe. DTH hammers are most effective in hard rock formations such as limestones or basalts; soft,
fine-grained formations tend to clog the air ducts or jam the piston slides.

Nonetheless, DTH hammers are extremely cost-effective and hence very popular with commercial
drillers.

Figure 15: DTH hammer button bits.

14
2. Siting of Drilled Water Wells
Proper siting of an improved groundwater supply (particularly borehole and hand-dug water wells)
provides a foundation for its success and long-term sustainability. Determining the best site for wells
requires consideration of a number of interrelated aspects, as shown in Figure 2.1. The prevailing
geology and available groundwater resources are fundamental since they determine what is
possible. The use or uses of the water and the users themselves are important factors as they
strongly affect the location where water is needed. The impacts and risks of a new well need to be
considered so as not to adversely affect existing and new abstractions. Last but not least, there is
need to consider access to the source, not only for the drilling itself, but also in the long term.

Figure 2.1: Combining Different Aspects for Site Selection

Thus, sound knowledge and practice of well-siting procedures have important implications for the
economics of water development, as well as for the environmental and functional sustainability of
groundwater abstraction. There are many techniques for well siting, each requiring different skills as
well as investments in technology. These techniques are appropriate, either singly or in combination,
for different geological and user settings. It is important that the appropriate techniques are utilized.
Of particular mention is the facts that water-well siting in Somalia tend to use geophysical surveys,
even though they are not always necessary. Based on extensive experience, we strongly recommend

15
that a stepwise approach (Fig 2.2) be followed for the comprehensive planning and management of
a successful siting assignment. The field note is structured according to the stages in Figure 2.2, and
each of the subsequent sections focuses on one stage.

Figure 2.2: Work Flow for Water-Well Siting

1. Conceptual Model

In order to plan and implement effective water-well siting, it is necessary to gather available
knowledge of local groundwater occurrence and conditions, including climate data and the effect of
pumping and groundwater recharge. It is useful for this in-formation to be set out as a simple
conceptual model illustrating the understanding of groundwater conditions.

The model may comprise a basic map of geology (plus information on rivers, settlements, land use).
It can be interpreted with the use of existing knowledge, supplemented wherever possible by
information obtained on reconnaissance visits. Are-as of good and poor groundwater availability and
water quality constraints can be highlighted.

Sufficient specific knowledge of the prevailing modes of groundwater occurrence and uses in the
project area is needed from the planning stage so that adequate well-siting capacity can be built into
programs from the beginning.

2. Requirements of the Water Well(s)

16
Prior to siting one or more wells, the requirements in terms of water use, population to be served,
water quantity and quality as well as water-lifting and distribution mechanisms need to be
considered. This section sets out the implications of two different types of water supply on siting.

- Rural water supply

Siting of wells (Hand dug or hand pump) needs to take careful account of both physical access by
users and access according to socio-economic class, caste, disability and other factors which can
exclude certain user groups. It is necessary to clearly define the user population, and to identify any
sub-groups within that population which differ in terms of wealth, power, position or influence. It is
almost inevitable that the poorest and those from lower caste or class tend to be marginalized by
those with more power or influence.

In many situations, the most powerful individuals try to exert undue influence to have the well sited
for their own convenience. As it is mainly girls and women who are engaged in carrying water for
domestic uses, they are the ones who are the most affected by the newly installed water supplies
and therefore by the siting. To overcome such problems of marginalization, some organizations
deliberately site wells in low-income sections of the com-munity in order to positively discriminate in
their favour, whilst not excluding wealthier or more powerful users. Participatory decision-making
methods can be used for this.

- Small town water supply and camps

In the case of small town supplies from groundwater, two extra considerations apply: (a) the greater
likelihood of groundwater contamination beneath urban centers, and (b) the greater demands for
water as a consequence of high population densities. For both these reasons, well fields serving
small towns tend to be located out-of-town in high-yielding aquifers where groundwater
contamination is less problematic, with transmission mains to bring water to the consumers. In
camps for refugees and internally displaced persons (IDPs), population densities may be as high as in
towns, but water supply requirements are usually significantly lower (the Sphere standard is 15 litres
per person per day, whereas in a town with piped water to the home, a figure nearer 100 litres is
more usual). Abstraction points may be sited much closer to the users, and this makes them
vulnerable to contamination especially in long-term settlements.

Table 2.1: gives some very rough estimates of the daily water abstraction depending on the type of
settlement to be served.

Water use Scale Approximate demand Average pump rate


[m3/day] [l/sec]*
Rural water supply Single well for 100-300 2-6 0.1-0.3
persons
Small town water Single well for 2,000 – 500 -2000 2-10
Supply 10,000 persons
Assumptions for consumption:
Rural water supply – 20 litres/person/day
Small town water supply - 40 litres/person/day
*Assumes that water is pumped for ten hours a day.

17
3. How to Determine the Most Suitable Site?

Determining the best site for the water well requires consideration of technical, environmental,
social, financial and institutional issues. The siting process should show which groundwater
conditions are dominant in the project area and enable the well design to be specified. Professional
siting involves desk and field reconnaissance, and makes full use of existing data. The actual process
of well siting requires (i) consideration of key factors and adequate use of sensible combinations of
(ii) information sources and (iii) siting techniques. This section examines these three aspects in turn
and provides a logical approach to well siting.

3.1 Key Factors for Consideration

In order to determine the best location for a well, ten factors are of particular importance:

1. Sufficient yield for the intended purpose

The groundwater aquifer should have a sufficient yield for a rural water supply (around 0.1-0.3
l/sec), for a small town water supply (2-10 l/sec), or for a larger scale need such as a significant
irrigated area. This information is sometimes available from existing documents or maps or can be
derived by performing a pumping test on an existing borehole

2. Sufficient renewable water resources for the intended purpose

Although a well may be capable of delivering a certain yield in the short to medium term, if the
groundwater is not regularly replenished by infiltration from rainfall or river flow, then that yield will
not be sustained over the long term. It is therefore important to evaluate the likely recharge to the
aquifer, and how this might vary with time. This estimate can be based on a calculated water
balance of an area.

3. Appropriate water quality for the intended purpose

Different water uses impose different water quality requirements. Domestic water must be free of
disease pathogens (which are carried in human excreta) and low in toxic chemical species such as
arsenic or fluoride. When using groundwater for irrigation, the level of salinity should be checked.
Well siting must therefore take account of knowledge of the occurrence of such undesirable
substances. The quality of the water from the completed and developed well should be compared to
national standards. Where these are not available, the WHO guidelines for drinking water (WHO
2008) may be used.

4. Avoidance of potential sources of contamination

It is essential to avoid point contamination sources such as pit latrines, septic tanks, livestock pens,
burial grounds and solid waste dumps. There are international guidelines on separation distances or
groundwater protection zones.

5. Community preferences, women’s needs and land ownership

Engagement with the community to agree on the well location is essential. It requires some
negotiation to explain technical constraints whilst taking community preferences into account. Full

18
consideration of the needs of women, who tend to be responsible for water collection, is essential.
Land ownership issues also need to be considered to avoid subsequent disputes between the land
owner and water users. Formal agreements regarding land ownership and access to the supply may
be required.

6. Proximity to the point of use

Within the constraints of geology, groundwater resources and groundwater quality, wells should
ideally be sited as close as possible to the point of use. This means that walking distances to collect
water from rural point sources (e.g. hand pump wells), energy costs and piped supplies should be
minimized. Walkover surveys should be undertaken to prepare a map of the community. Interviews
with householders will help to understand the community’s preference for well location. In general,
the community would be expected to indicate three preferred well sites in their locality, in order of
priority.

7. Access by construction and maintenance teams

In the case of wells constructed by heavy machinery, access by drilling rigs, compressors and support
vehicles is crucial. Even when lighter equipment is used, vehicle access for construction and for
maintenance is important. Site selection must therefore take account of these needs.

8. Avoidance of interference with other groundwater sources and uses

In areas where some groundwater development has already taken place, the construction of a new
well can lead to increased drawdown in existing sources. This in turn can lead to greater pumping
costs in both the existing well and the new well, reduced yields, changes in groundwater quality and
potential conflict between users. In an early phase of the siting process, possible interference and
risks of derogation should be described and discussed. This means that the radius of influence of
existing wells should be calculated and new wells located outside this zone. In high risk situations,
possible alternative siting areas should be evaluated.

9. Avoidance of interference with natural groundwater discharges.

In a similar way, construction of a well too near to natural springs, watercourses or wetlands can
lead to a reduction of water levels, potentially drying up these important water sources and
ecosystems and affecting uses and users dependent upon them. The intrusion of saltwater due to
too high abstraction of groundwater near the coast could lead to irreversible decline of water
quality.

10. Risk

As part of water-well siting, the risk of drilling a dry borehole should be categorized (e.g. high,
medium and low risk). In the case of wells which are to be fitted with hand-pumps in areas with
known hydrogeology, geophysical techniques (e.g. resistivity, conductivity) are rarely required so
long as a desk study has been undertaken of the general hydrogeology of the area. Drilling small
diameter exploratory wells (e.g. with a small hand auger) can also be a suitable siting method for
shallow wells. However, this hole should be properly sealed afterwards to avoid aquifer
contamination.

19
3.2 Information Sources

It is essential to use the information sources described below to support the siting techniques
described in the following section, and in conjunction with the considerations of different water uses
and user group requirements as mentioned in section 2 of this field note.

Maps: Topographic maps provide the most basic information for undertaking a well siting program.
While community names and locations may not always be correct, the terrain and the rivers are
likely to be accurate and provide general indications of the situation of the land and accessibility.
Geological and hydrogeological maps present and summarize a great deal of complex information in
a succinct visual form. Hydrogeological maps at similar scales are much more rarely found. The map
legend is as important as the map itself, as it contains much descriptive information which is
necessary to get the most from the map itself. More recently there have been useful approaches to
the production of maps of groundwater development potential, often at a more local scale. On
these, existing hydrogeological information and borehole data are depicted and interpreted to
indicate where there is good potential for using groundwater and where there are likely to be
significant constraints in terms of both quantity and quality to provision of water supplies.

Documents: A wide variety of documents, including project reports, master plans, geological survey,
consultants’ and drillers’ reports, NGO project documents and academic studies and meteorological
data (i.e. rainfall), can provide useful information about the areas where groundwater development
is proposed.

Field visits and interviews can provide considerable information from the community on
groundwater resources, including seasonal fluctuations. In addition, relevant information on source
preferences, water uses, gender issues and economic interests can be collected. These will all
influence the selection of a suitable location for the well.

Drilling records, databases and data exchange: Often, the most reliable information on local
geology and hydrogeology comes from the field experience of previous drilling and well-digging
activities. Ideally, such experience is encapsulated in drilling logs and geologists’ logs which are held
in national databases. In reality, however, such records are often not kept (especially if the well is
unsuccessful), not submitted (especially in the case of NGOs and individuals), or not collated
(especially when Government resources are limited, and higher priority is placed on new
construction than on record keeping). In the best cases, such logs, records and databases can be
extremely useful sources of information.

The extra cost in a well construction program of collecting such data is relatively small, and the
incremental benefits for siting from the cumulative knowledge, improved interpretation and
enhanced conceptual model of local groundwater occurrence are very large. It would be even more
beneficial if drilling results were subsequently compared to the specific well-siting techniques used
in a systematic, site by site and overall project evaluation. Such an assessment of siting “success” of
course reflects on the operator as well as the technique, and this is very rarely done and almost
never published. It is also made difficult because siting and construction are often undertaken by
different organizations. If these are private contractors or consultants, they may treat the siting data
as commercially confidential information which they are unwilling to share.

20
3.3 Techniques for Siting

Many techniques are available for siting drilled water wells, and the most commonly used ones are
summarized in this section. There is no ideal single technique that fits best to every condition.
Instead, the use of different techniques should always be tailored to the local conditions and rock
types, and in particular to the three situations of increasing hydrogeological complexity set out in
section 3.4. Some of the techniques may look simple, but considerable skill and experience is
required to understand and interpret the results correctly. Therefore, it is strongly recommended
that the techniques always be used by a trained hydrogeologist or technician and carried out under
the supervision of one. The most often applied techniques are summarized below.

Remote sensing: The use of aerial photography, side-looking airborne radar (SLAR) and satellite
imagery has a powerful role in identifying geological boundaries and hydrologically significant
features (such as deep fractures) which may not be visible on the ground. Such remote techniques
always require an independent check at the ground surface by field reconnaissance, by geophysical
survey or by drilling (known as ground truth) to have confidence in the findings obtained remotely.
Remote sensing can be very useful, but its use should always be determined by realistic expectations
of what it might or might not indicate.

The most likely applications are firstly in the planning and reconnaissance stage and secondly for
narrowing down target areas or locating specific features for geophysical survey, for locating and
delineating communities requiring water supplies and identifying existing supply sources.

Hydrogeological field surveys: If indications of the potential for using groundwater have been
obtained from maps, documents, satellite images, hydrogeological field reconnaissance provides the
opportunity to check this. Thus, the mapped geological formations should be confirmed from rock
exposures (for example in river beds and road cuttings). Local topography and geomorphology can
influence groundwater occurrence, storage and flow and enhance groundwater recharge and help to
produce favourable sites. These should be observed and noted. Vegetation cover can reflect
geological conditions and indicate the presence of shallow groundwater.

The field reconnaissance should locate and examine existing dug and drilled wells to verify the
information about yields and water levels already collected from secondary sources. Their operating
status and condition should be noted, together with any visual evidence of water quality constraints
such as iron staining or fluoride impacts, and any likely sources of pollution. These observations
should be supplemented by information from local communities who are likely to be very
knowledgeable about their local environment and their water sources. Older members of the
community, for example, may be able to indicate scoop holes that have dried up, or they may recall
vegetation patterns prior to deforestation. Such information should include the normal seasonal
variations and more severe drought impacts on yields and groundwater levels in both traditional and
improved sources as well as information about water quality constraints. All of the information
collected in the field reconnaissance should be carefully recorded (using a logbook, drawings,
photographs, GIS, other field equipment). To collect such information effectively, field surveys need
the participation of trained community workers who can converse in the local language.

If the survey finds sound evidence of groundwater potential, then sites can be selected without the
need for additional investigations using geophysics. This is likely to the case for most of the areas

21
with simple hydrogeological conditions of unconsolidated sedimentary materials and shallow
groundwater. The survey should also enable a choice to be made on the hydrogeological suitability
of areas or locations for dug wells or boreholes where both are envisaged within a program.

Establishing a siting approach based on a combination of existing information, remote sensing and
field survey can be highly cost effective for these conditions, but will only be successful if it is led by
an experienced hydrogeologist. Moreover, qualified personnel with hydrogeological knowledge are
needed to decide when such an approach cannot be applied with confidence and additional
investment in geophysical surveys is needed, and then to plan, implement and interpret the surveys.

Geophysical surveys: Geophysical surveys are by far the most commonly used techniques in well
siting. These techniques measure the physical properties of rocks, such as their resistivity and
conductivity, magnetic fields and sonic properties. Most cannot directly detect the presence of
water. Instead, the contrasts in sub-surface (rock and water) properties are interpreted in relation to
geological features that are expected to facilitate groundwater storage and movement. In favorable
circumstances, these techniques can detect vertical fractures in hard rock, layering in horizontal
formations, and contrasts between dry and wet rock and between fresh and saline water. As with
remote sensing, ground truth is needed, and use of geophysics should always be determined by
realistic expectations of what can be achieved.

Although geophysical surveys can assist in locating productive sites, they are often included
automatically in a tender in the hope that they will produce something useful. This approach rarely
rewards the effort put in, and there are many cases of geophysics having added nothing to the
reliability of well siting. While there are many geophysical techniques, those most commonly used
for well siting are the electrical resistivity and electromagnetic (EM) methods, with seismic refraction
and magnetic techniques also having some applications.

The resistivity method has been used for many years and can be employed in two distinct ways. The
first is a Vertical Electrical Sounding (VES) in which depth variations in subsurface resistivity at fixed
point can be interpreted in terms of a sequence of geological layers. The electrodes are expanded in
an array about this central point (Fig 2.3). The second is a constant separation traverse in which the
electrode array is moved across the ground to provide qualitative information about lateral changes
in subsurface rock types and structures. Resistivity profiling has largely been replaced by the
electromagnetic (EM) methods, which provide better information about lateral changes in resistivity
much more quickly and cheaply.

Figure 2.3: Equi-potential surface and associated current lines two current electrodes

22
Figure 2.4: VES Measurement in Somalia

Electrical resistivity and electromagnetic methods are the two most widely used geophysical survey
methods. The equipment is relatively inexpensive, robust and not difficult to operate in the field. A
widespread consequence of this is the routine use of such equipment by field technicians who do
not have geological training. However, in order to obtain the best results, the interpretation of data
from these (or any other) geophysical methods requires experience and triangulation with local
hydrogeological knowledge. All the information obtained is entered into a computer that analyses
the measurements with special software. The range resistivity’s are very large. The values given in
figure 2.5 are only informative: the particular conditions of the site may change the resistivity values.
For example, dry coarse sand or gravel may have a resistivity like that of igneous rock, whereas
weathered rock may be more conductive than the soil that overlaying it.

Figure 2.5: Electrical resistivity and conductivity of earth materials

23
Since the resistivity of the soil or rock is controlled primarily by the pore water conditions, there are
wide ranges in resistivity for any particular soil or rock, such that resistivity values cannot be directly
interpreted in terms of soil type lithology. Commonly, however, zones of distinct resistivity can be
associated with specific soil or rock units on the basis of local outcrops or borehole information. It is
the enormous variations in rock and mineral electrical resistivity that makes resistivity techniques
attractive. In fact, siting which is carried out by inexperienced operators and analysts can actually
reduce the likelihood of finding water. If the interpretation is poor, it may be better to drill at
random.
3.4 A logical approach to well siting

Siting is not solely about applying the science of groundwater, but also encompasses social,
economic and institutional aspects as well as consideration of the management of the water supply.
The user aspects highlighted in section 3.1 (i.e. community preferences, women’s needs and
proximity to the point of use) are important. However, this needs to be balanced with the need to
select sites using hydrogeological criteria which ensure the best chance of obtaining adequate and
sustained yields of good quality water. These can limit what is possible. Many investigation
techniques are available and often well proven to assist in well siting. However, none are
consistently useful at all times, and their success or failure depends on them being used correctly
and applied in appropriate situations. The challenge is to match the effort, intensity and costs for
investigation to the complexity and uncertainty of the hydrogeological conditions and the scale of
user requirements. A logical and systematic approach to well siting is recommended which:

- Identifies features on the ground that may be favorable for groundwater occurrence.
- Selects the geophysical method or methods most suited to the task of locating them;
- Plans the survey fieldwork and interpretation accordingly;
- Provides adequately qualified and experienced staff to undertake the fieldwork and
interpretation;
- Provides adequate funding and resources for the work.

Three situations shown in Figure 2.6 are discussed to illustrate this logical approach:

 Scenario 1: the hydrogeology is well-known and not difficult; groundwater is relatively easy
to find, and wells can be sited almost anywhere;
 Scenario 2: the local hydrogeology is largely understood and reasonably consistent but
challenging; it needs some effort to find reliable groundwater resources;
 Scenario 3: the local hydrogeology is less known and understood; it is complex and
uncertain; there is a high risk of failure to get boreholes with enough water.

Figure 2.6: Three complexity scenarios for siting: Source: MacDonald et al 2005
Scenario 1: Scenario 2: Scenario 3:

24
In these three scenarios, successful siting requires increasing levels of technical capability,
manpower resources and costs, as set out in Table 2.2.

Table 2.2: Siting scenarios and resources needed

Resources Scenario 1 Scenario 2 Scenario 3


Technical–organizational issue
Hydrogeological Desk Study   
Field visit   
Risk analysis   
Geophysical survey - ? 
Social issues
Social structure and community preferences   
Time and Costs
Time needed low medium high
Costs low medium high
 to be undertaken ? depends on level of risk - not necessary

The case study in below provides an example of the siting procedure and techniques used in Somalia
by GSA organization operating in the conditions of scenario 2.

Case Study of Siting by GSA:

Gurmad for Sustainable Aid (GSA), in Somalia follows the following siting procedures:

Phase 1 is a planning and reconnaissance study which includes mobilization of equipment and
personnel, collection and interpretation of existing data and preliminary selection of target areas for
detailed investigations. It normally also includes field data collection, site specific data analysis and
verification of results of desk studies and preliminary site selection. Activities to be carried out
during this phase are as follows:

 Liaise with community;


 Collect and review hydrogeological reports and literature for the areas of interest;
 Collect and study maps – (topographic, geological and hydrogeological);
 Collect and study drilling information and records;
 Visit field to determine field conditions, accessibility to preferred sites and community
readiness to participate;
 Use a GPS to locate sites on a topographic map.

This will provide information on where to expect to find water and whether the quality is expected
to be good.

Phase 2 comprises the hydrogeological investigations, including topographic map analysis and
detailed geophysical surveys of the areas of interest. These mainly use the resistivity technique to
characterize the different formations. Since most of Somalia is underlain by hard rocks, the approach
uses an ABEM Terrameter SAS 1000 or 4000 for traversing and vertical electrical soundings. These
methods provide a) an estimation of the thickness of the regolith, b) an indication of horizontal
changes in aquifer properties and c) the locations of any vertical geological boundaries.

25
Existing data for nearby water sources are collected and used for calibration. Where data are not
available, calibration resistivity soundings are made at existing drilled water wells to characterize the
underlying geology in terms of resistivity and groundwater potential. The results of the calibration
measurements guide final interpretation of the data. Traversing is carried out to assess lateral
variation whenever this is found to be necessary based on the local hydrogeological environment.
The anomalies identified from the profiling are further investigated by soundings to ascertain
hydrogeological variation with depth. Initial resistivity profiles are always run perpendicular to the
inferred fracture zone.

4. Preparing Tender and Contract Documents

A siting assignment will most probably form the basis for the subsequent procurement of a drilling
enterprise to undertake water-well construction. Ideally, the siting and drilling are undertaken as
two separate assignments, with the water well drilling undertaken after the siting. In some cases,
the siting consultant will subsequently be responsible for supervision of the drilling assignment.
There are cases where siting and drilling are undertaken as one combined contract. In fact, some
contracts place the full risk of drilling a non-productive well with one contractor who is responsible
for siting and drilling. However, such practice should only be undertaken in circumstances where the
risk of drilling a dry borehole is fairly low (i.e. scenario 1 in section 3.4). Prior to preparing tender
documents, it needs to be decided if siting will be undertaken under separate contract or if it should
be combined with the drilling. In order to prepare tender and contract documents for well siting it is
essential to:

 Establish the number of wells that are to be sited and ultimately drilled and their geographic
distribution;
 Have an understanding of a conceptual model of the way groundwater occurs;
 Consider the requirements of the water well as well as key social, technical and
environmental factors;
 Make full use of all available information sources and;
 Undertake a pre-selection of suitable siting techniques. Note that the logical approach to
well siting set out in section 3.4 provides guidance as to the selection of siting techniques.

The tender document should include the following:

 Objective of the siting.


 A general description of the hydrogeological conditions in the siting areas and the
challenges to be expected.
 Information about the techniques that are considered suitable for investigation and siting.
 A clear explanation of the deliverables, including a definition of the specific set of geological
and hydrogeological data that should be investigated and verified during the siting (such as
depth of water bearing layers, depth to water table, transmissivity).
 The number and approximate location of the sites expected, water use, yield and water
quality requirements.
 Overall timeframe of the work, deadlines and milestones.
 Clear definition of roles and responsibilities.

26
 An assumption of the number of meetings with clients. Siting assignments can lead to an
iterative process of interpretation, investigation and detailed assessments. Meetings with
the clients are necessary.
 An assessment of the risks associated with the assignment. The work could be disturbed by
weather conditions, difficult road access or social unrest. These may prevent the siting team
from performing the contract as planned. The tender documents should define a clear
procedure to be followed in such circumstances. There are also risks of failure to find a
suitable drilling site due to the complexity of the geological and hydrogeological conditions.
 Clarification of the payment scheme. Very often, payment for a siting assignment takes
place after a debriefing meeting with the client, including presentation of the results, and
after submission of the final documentation. Alternatively, some part of the payment could
be withheld until the driller has finished work. If unsuccessful drilling occurred because of
wrong siting, some of the payment for the siting assignment could be permanently withheld.
Such procedures have to be clearly and explicitly defined in advance in the tender
documents for the siting.

A rough cost estimate should be generated for the siting work based on this information. This can
subsequently be used to compare with the prices quoted in the tender offers submitted.

The quality of the siting directly influences the quality and cost of work of the drillers. Therefore, the
Terms of Reference (TOR) in the tender document need to be precise, complete and clearly written.
The TOR should define at least the objectives of the assignment, the services executed by the
consultant or organization which undertakes the siting, the tasks of the client, the deliverables
including the format of the data, the timeframe and quality standards.

It is recommended to append a draft contract to the tender documents.

5. Procurement and Contract Award

Enterprises interested in undertaking the work submit tenders based on the information and
requirements of the tender documents. The tender offers should:

- Specify the composition of the team.


- Provide details of equipment and methods that will be used (even including alternatives to
those proposed in the tender)
- Set out the experience of the enterprise,
- Provide a rough risk analysis with mitigation measures and
- Set out a draft time schedule with tasks to be completed.

Drilling companies should prepare offers for their work based on realistic prices. In order to prepare
a financially reasonable offer, the consultant or company bidding for the work should be aware of all
formal deadlines, eligibility and selection criteria, the Terms of Reference and other contract issues.
Any areas which are unclear in the tender documents should be clarified prior to submission of the
tender.

The procurement procedure should allow the client to select the best eligible offer according to
specific eligibility and selection criteria defined by the client. In order to make sure that the process
is fair, a clear procedure, as defined by the client or donor organization, should be followed. The

27
specific eligibility, selection criteria and procedure to be followed should be transparent. These
should also be set out in the tender documents. Formal standards and procedures for evaluation for
public procurement exist in most countries.

For the evaluation of the bids, the client will first check whether the bidder has fulfilled the eligibility
criteria (e.g. license, registration or other prequalification requirements). If these are fulfilled, the
offer will be evaluated according to predefined selection criteria and price. The tender evaluation
should focus on the experience and expertise of the key personnel of the team and their
presentation of their methods rather than on analysis of the price alone. Very rarely is the cheapest
offer the best offer. Generally, the best offer is the one with the best quality/price relationship.

In cases of complex siting assignments, it is recommended to involve experienced advice for the
tender evaluation process. Following the tender evaluation, the siting contract is awarded, and the
work can commence.

6. Field Work and Contract Management

After the signing of the contract, both the NGO and the drilling company will plan how the
assignment is to be undertaken, including communication between the two parties and field visits to
the community. The siting company will subsequently commence the assignment. The NGO will
introduce a project manager who is also responsible for quality assurance. The siting company/team
in particular will collect and analyses available information, contact sub-contractors (e.g. for
geophysics) and organize staff and logistics. During the entire siting assignment, there should be an
organized exchange of information, decisions and documents between the NGO and siting company.
The process of siting often follows an iterative working process which includes:

- Critical analysis of data (deskwork);


- Compilation of a first conceptual hydrogeological model of the area (deskwork);
- Field visits, local knowhow, test drilling;
- Optional: refined conceptual model, verification including water quality data;
- Geophysical field measurements, interviews with water users, land owners, other
actors/stakeholders;
- Verification, refined conceptual model, risk analysis;
- Recommendation of sites;
- Documentation (including face to face debriefing).
- Depending on the complexity of an assignment, some of these steps could be combined.

In case of questions, constraints and problems, it is advisable to contact the client as soon as
possible.

The roles and responsibilities for the different stakeholders involved in borehole construction are as
follows:

 The Community members are the end users of the water supply. They must be included in
the process of technology selection and siting so that the finished water point can meet their
needs. There are cases where the Community is involved in supervision, but they should not
be responsible for technical or contractual details unless their capacity has been built
extensively.

28
 The Client is the organization, company, household or community that is contracting out the
borehole construction. Their responsibility is to fulfill regulatory requirements and ensure
that they have well trained Supervisors present on site for the full duration of drilling
operations.
 Note that even if district local government is not the client, it is still important for them to
be involved in the process. District local government should attend a pre-mobilization
meeting, commencement of drilling and the end of the construction, including pump testing.
 The Funding Organization pays for the borehole. It may be the Client or another
organization such as an international development partner or NGO. The funding
organization should not impose conditions that create perverse incentives or undermine the
long-term sustainability of the finished borehole (e.g. by insisting that the cheapest bid is
accepted regardless of quality). It should work within national or local government systems.
 The Regulator (once established) issues permits and licenses for siting, supervision, drilling
or abstraction. Legal requirements should be established by the Client early on to avoid
delays.
 The Project Manager is responsible for a wider project. The siting and drilling will usually be
just one component within a project, comprising community training/mobilization, pump
technology choice, water point design and construction, and establishing or strengthening a
rural water supply service.
 The Supervisor is sometimes called the ‘Rig Inspector’. Supervision is usually done either by
the Client’s staff or by a consultant. The Supervisor may be a hydrogeologist, an engineer, or
a technician. Although the Driller and the Supervisor work together to deliver the product,
their roles are different. The Supervisor’s responsibility is to ensure that the Driller adheres
to the technical specification, makes all the required measurements, keeps all re -cords
accurately and ensures that health and safety procedures are adhered to.
 The Driller, or Contractor, is the organization that physically does the drilling. Sometimes,
this will be an independent private sector company. In other cases, it will be an in-house
team working for a government agency or NGO. The Driller’s responsibility is to drill the
borehole as specified. Each Driller should have a designated ‘Record Taker’ who should
remain on site at all times, with the duty to collate all the measurements and complete all
the forms.

7. Payment, Follow-up and Documentation

Payment for siting services has to adhere to the signed contract. Often, payments will be released
after milestones have been passed and the results have been approved by the NGO. In the case of
non-compliance, an arbitrator may be required. The NGO should build up internal capacities and
resources to manage relevant data from the siting assignment and submit it to the relevant
authority. The results of the siting assignment form the basis for the procurement of a drilling
contractor.

29
3. Borehole Construction
Once the hydrogeological survey is done and drilling method and the equipment have been chosen,
you will be required to observe and monitor the construction of the borehole. You may also be
charged with the responsibility of supervising (in terms of quality control) the drilling of boreholes by
drilling contractors. Quality control of drilling operations requires knowledge and confidence, which
are acquired only by experience hydrogeologist. This section outlines key considerations for
borehole construction using the mud and air rotary drilling methods.

Borehole construction has been arranged into four stages, with each stage requiring a set of actions
(table 3.1). Some of the stages are mentioned only briefly in this document to avoid repetition as
they will be covered in detail in specific guidelines devoted to the particular stage.

Table 3.1: Stages in Borehole Construction

Stage Action
Stage 1: Community sensitization, - Stimulating demand
Technology choice & mobilization - Community contribution
- Setting up post-construction monitoring
Stage 2: Site selection - Community consultation
- Site selection to ensure health and
hygiene as well as a productive borehole
Stage 3: Borehole construction - Mobilization; appointment of contractor
and supervisor
- Drilling, Casing, development, pump
installation, Test pumping, Borehole
completion
Stage 4: Water facilities construction - Elevated water tank, kiosk, animal
troughs

3.1 Construction Considerations:


Large drilling rigs are equipped to ensure that a borehole is started true and vertical. Maintaining
verticality and straightness can be difficult during the early stages of drilling, but as the drill string
weight increases, this problem tends to dissipate unless highly heterogeneous drilling conditions are
encountered (in the form of boulders or cavities). Straightness is particularly important for water
boreholes in which long strings of casing and screens may have to be installed with a gravel pack
filter.

As drilling proceeds, drill pipes are screwed together. This allows tools and pipes to be rapidly
attached and screwed together on the rig. A blast of air is sent through each pipe to remove
blockages, and the string is tightened with heavy-duty spanners on the rig. Taller drill masts can
obviously handle longer drill pipes – six meters is the normal length, except for smaller rigs (see
above) – which speeds up bit lowering (‘tripping-in’) and raising (‘tripping-out’) times.

30
Figure 3.1: Schematic section of an example of temporary borehole completion.

Reaming – the enlarging of an existing hole – can be carried out either with a drill bit of any kind or
with specially designed reaming tools. Drilling companies often devise tools for special use,
sometimes in the field, and they are often very ingenious.

It should be borne in mind that after drilling has begun, the sides of the upper part of the borehole
are likely to suffer erosion by circulation fluid and cuttings, which causes an irregular enlargement of
the borehole, reducing up-hole fluid (air or mud) velocity. This can be dealt with by installing
conductor pipe as described below.

As drilling proceeds, the amount of water leaving the borehole will – it is hoped – be seen to
increase, reaching a point at which it becomes clear that the borehole will provide the required
supply. Even then, the borehole may have to be deepened further to provide sufficient pumping
drawdown. However, if the borehole is found to be wanting, it may be advisable to stop drilling early
(unless a hand-pump is acceptable at that location) or carry on in the hope of a greater water strike
(here some knowledge of local geology would be very useful). If fragments of basement rocks start
appearing in the cuttings, and the penetration rate decreases significantly, the ‘hydrogeological
basement’ has, in all likelihood, been reached and it would probably be futile to continue to deepen
the hole.

Penetration rate through each zone or formation in the borehole may be determined simply by
timing the progress of one drill pipe or a fixed distance marked by two chalk marks on the drill pipes
as they pass through the table. Penetration rate can provide an estimate of formation consolidation
or hardness, and also show precisely when an aquifer was crossed.

The question, then, is: When to stop drilling? The supervisor normally has an idea, from the project
specifications, of how much water is required from a borehole; a hand-pump, for instance, does not
demand a large supply (0.5 litre/sec is more than enough), whereas pump supplying a storage tank
for a village, a refugee camp, or a facility such as a school requires a significantly greater yield. When
drilling is finally stopped by the supervisor (who normally bears this responsibility), it is advisable to
allow a few minutes for the water level in the borehole to recover and to then measure it with a
cable dip meter.

31
Field hydrogeologists and water engineers working on borehole drilling projects in the commercial
or humanitarian sectors are most likely to encounter rotary drilling machines (of whatever size)
using mud circulation or compressed air. For this reason, the discussion in this review is limited to
those techniques most commonly used in water borehole drilling: mud rotary and air rotary, as
cable-percussion drilling, auger drilling, and other methods are becoming increasingly rare.

I. Mud rotary drilling

Besides the cooling and lubrication of drilling bits, which has already been mentioned, the addition
of special muds or other additives to circulating water provides the following significant advantages
when drilling in unstable formations:

- By using fluids of a density higher than that of water itself, significant hydrostatic pressure is
applied to the walls of the borehole, preventing the formation from caving in
- The liquid forms a supportive ‘mud cake’ on the wall of the borehole, discouraging the
collapse of the formation
- The liquid hol ds cuttings in suspension when drilling is halted for the addition of drill pipes
- The liquid removes cuttings from the drill bit, carries them to the surface, and deposits them
in mud pits

Drilling mud – a partially colloidal suspension of ultrafine particles in water – fulfills these functions
by virtue of its properties of velocity, density, viscosity, and thixotropy (ability to gel or freeze when
not circulated). Water by itself exerts hydrostatic pressure at depth in a borehole, but at shallow
depths this may not be sufficient. Among additives for increasing the density of water, salt is one of
the most convenient; but one of the most widely used is a natural clay mineral known as bentonite
(calcium montmorillonite), which swells enormously in water. A slurry consisting of water and
bentonite combined in the proper proportions has a higher viscosity than water and forms a mud
cake lining in the borehole. However, a major disadvantage is that the mud needs to be mixed and
left for some 12 hours before use to allow the viscosity to build up.

The normal bentonite mud mix is 50 kilograms per cubic meter of water (a 5% mix), or 70
kilograms per cubic meter, if caving formations are expected.

Natural polymers provide a more practical solution for water boreholes, but they are relatively
expensive, so should be used with care. One example of such a polymer, used in oilfield and water
drilling, is guar gum, an off-white colored powder extracted from guar beans. It is an effective
emulsifier used in the food industry, so is biodegradable, and will lose its viscosity naturally after a
few days. Polymers are best mixed by sprinkling the powder into a jet of water, to prevent the
formation of lumps. The polymer mud should be mixed during the setting-up stage – a minimum of
30 minutes is usually required – so that it has time to ‘yield’ (build up viscosity).

The normal mix for guar gum polymer is one kilogram per cubic meter of water; for drilling in clay
formations, use up to 0.5 kilogram per cubic meter, and for caving formations, use one to two
kilograms per cubic meter.

Besides the usual mud properties, polymer drill fluids also coat clay cuttings, preventing the
formation of sticky aggregates above a drill bit (known as ‘collars’), which can hold up drilling while

32
they are removed (a simple remedy for clay aggregation is to add salt to the drilling fluid). Another
advantage of polymers is that they make it possible, when it is clay that is being drilled through, to
distinguish genuine formation samples from the mud. Degradation of polymer muds is accelerated
by high ambient temperatures, acidity, and the presence of bacteria (using the polymer as a food
source): polymer-based mud might last only two or three days in tropical conditions, and can cause
bacterial infection of the borehole. It could be that natural polymer powders have a limited shelf life,
and this should be checked before purchasing stock from a supplier. Food-grade bacterial inhibitors
have been used as additives to prevent the breakdown of polymer-based muds. When using
polymers, observe the manufacturer’s guidelines. Foaming agents are also widely used as drilling
fluid additives, normally in air drilling.

- Checking the viscosity of drilling mud

Every mud additive (bentonite, mud, salt, etc.) must be mixed into the circulating water to provide
the correct viscosity. This can be done initially in a specially prepared pit, but as drilling proceeds,
and especially if groundwater is struck, the mud will become diluted, and more mud or additive
powder will have to be added. Too low a viscosity may result in fluid seeping into the formation, and
it may later be difficult to remove the fine mud particles from the wall of an intersected aquifer,
reducing the efficiency of the borehole. ‘Thin’ mud may also cause cuttings to fall back onto the drill
bit, causing it to stick in the hole. The viscosity of drilling mud can be easily and frequently checked
by means of a simple viscometer known as a Marsh funnel.

Extremely porous or fissured formations can cause a loss of drilling fluid (mud); it is possible that the
entire mud circulation might disappear into a cavity. This could put a stop to drilling altogether, if
increasing fluid viscosity by adding more additive has no effect. If the area from which fluid is being
lost is not likely to be part of an aquifer, fibrous materials such as sawdust, dried grass, or cow-dung
could be introduced into the mud, while ensuring that a pumpable circulation is maintained. Such
additives can block large pores and cavities permanently, which is why they should not be used to
cure losses in a water-bearing zone.

- Mud pits

To mix the mud, as described previously, mud pits are required. This can be combined with a
‘suction pit’ or sump from which a mud pump will take the circulation supply. Second, a larger,
‘settling’ pit is essential, in which mud returning to the surface from the borehole’s annular space
will be allowed to drop its load of drill cuttings. The two pits and the borehole are usually connected
by shallow channels or ditches and a weir; a typical arrangement is shown schematically in Figure 12.
Mud pits are most commonly dug in the ground alongside the rig, but some contractors can supply
steel tanks, which are their equivalent.

If dug in soft soil, pits may be lined with plastic sheeting, clay or cement. Mud circulation through
pits must be slow and steady, to settle the cuttings and to make collecting formation samples
(normally taken from a channel close by the borehole) easier. The mud pump inlet and strainer are
held by rope above the bottom of the suction pit, so that mud that is as clean as possible can be
recirculate into the borehole via the drill pipes. Optional extra ‘swirl pits’ may be included between
the borehole and the settling pit to further aid settlement of debris. The capacity of the suction pit

33
should be roughly equal to the volume of the hole being drilled; the capacity of the settlement pit
should be at least three times that.

To roughly calculate pit volumes, given hole diameter


D in inches (drill bit size):
Borehole volume and suction pit volume = D2H/2000 in cubic metres
(or D2H/2in litres), where H is depth of hole in metres.
Settlement pit volume should be ~0.002D2H cubic metres (or 2D2H litres).

II. Compressed air rotary drilling

Using compressed air as the circulation medium does away with having to prepare and inject liquids
into a borehole (although water and additives may be introduced for special purposes). In some
cases, the use of air drilling may be essential: for example, when constructing observation holes for
pollution studies, where groundwater contamination should be kept to a minimum. Even then, a
formation may become contaminated by oil particles from the compressor. The principal features of
air drilling may be summarized as follows:

 The use of a low-density circulation medium (air) requires high fluid velocities to lift debris
out of the borehole. Thus, for large-diameter boreholes, large-capacity compressors are
required.
 Dry formations present few obstacles for air drilling, but a water strike at depth requires that
the air pressure overcome hydrostatic pressure to a significant degree, to operate the DTH
hammer and carry water and cuttings to the surface. Damp formations can, however, cause
problems, such as the accumulation of sticky cuttings above the drill bit (like the clay ‘collar’
referred to earlier).
 Air provides very little protection from borehole collapse, other than dry or damp pulverized
rock powder that smears the wall of the borehole. Because softer formations are easily
eroded, it is vital to protect the looser upper section of the borehole by inserting a suitable
length of steel tubing known as a ‘conductor pipe,’ which is a little larger in diameter than
the drill bit used when ‘spudding in’ (the very moment drilling starts at surface level). The
conductor pipe should protrude a little above ground level – but not so much that it
interferes with the rig drilling table – leaving space for cuttings to blow clear. Boreholes are
drilled with larger bits at first, reducing diameter at depth, after installing temporary steel
casing (protective lining inserted inside the conductor pipe) to protect areas of unstable
formation.
 While air drilling, up-hole airflow rates should be within the range 900 to 1200 cubic
meters/minute.

Temporary casing may be particularly difficult to insert through a horizon containing stones or
boulders (such as coarse river channel deposits), but unfortunately such formations often host good
aquifers. DTH hammers can break hard rock boulders (or partially fragment them), but there is
always the risk of the hammer diverting and becoming wedged, or lumps of rock falling behind the
bit and jamming it in the hole. The best way to deal with boulders is to install a simultaneous casing
system, which is supplied by most DTH hammer manufacturers. This allows steel casing to be pushed

34
or pulled down a borehole, directly behind the hammer, to prevent the walls from caving in. The
hammer has a large diameter bit that is used to make the hole for the casing; the bit can be
mechanically reduced in size and retrieved through the casing.

Such systems require rigs with strong masts and the power to handle heavy casing insertion in
difficult drilling conditions. During air drilling, foams can be added through the drill pipes to
eliminate dust emerging from a dry hole, to keep the borehole clean, and to prevent fine particles
from clogging any small water-bearing fissures that may be intersected.

Furthermore, soap bubbles help lift debris out of the borehole. However, foams do not provide any
hydrostatic support for collapsing boreholes; they also make it difficult to collect samples at any drill
depth.

3.2 Borehole Logging:


For a borehole to be properly logged, the driller and supervisor need to know its exact depth at all
times. This is necessary for the calculation of drilling charges, and while designing the borehole. First,
make a note of the length of the drill bit and of any other tools that may be used to drill the hole.
Put the bit on the ground and make a chalk mark, ‘0,’ on the first drill pipe against a suitable fixed
point on the rig and at a known height above ground level, such as the drilling table (which
centralizes the drill pipes in the hole). From then on, marks can be made on the drill pipe at regular
intervals – say, every half meter – to record the depth of drilling and to assist in the logging of
penetration rates.

Formation samples need to be obtained as drilling proceeds: the usual sampling interval is one
meter. These are obviously highly disturbed samples, having been sheared or broken from their
parent formation, so should not be used to infer characteristics such as bedding, texture, porosity, or
permeability. There will be a slight delay as formation fragments are lifted to the surface by the
circulating mud, but a rough estimate of the up-hole velocity should enable one to calculate the
actual depth at which cuttings were derived. Keep in mind that if mud viscosity is too high, or if
formation collapse occurs (viscosity too low), some fragments could return to the borehole, with the
potential of causing confusion. Cuttings obtained from the shallow mud channel near the borehole
should be washed in water to remove mud, and laid out in order (by the depth at which each was
acquired) on the ground or in a sample box with separate compartments for each sample. They can
then be logged by the supervisor or site geologist and bagged if required. Samples should, of course,
be labeled correctly with all information relevant to the job in hand.

The main attributes of a borehole log are accuracy and consistency; a good set of logs can be a
useful resource when planning future drilling programs. Drillers must keep their own logs and notes
and, as is often stipulated in contracts, these should be accurate; however, in practice, they cannot
always be relied upon, especially if the supervisor is absent from the site for a period. All geological
samples and water strikes should be logged by the drillers and the supervisor, as this important
information will be required for designing the borehole and the equipment to be installed.

Full borehole logging may also include geophysical logging, which is normally carried out only after a
well has been completed. Annexs gives a typical example of a drilling log sheet, which is applicable
for both mud and air drilling, and which should be kept by the supervisor. The driller’s log should

35
also include information on drilling or other work time, standing (waiting) time, and downtime
(breakdowns).

- Geophysical logging

Information about structural features and geological formations in a borehole can be remotely
obtained by geophysical borehole logging techniques. The object of well logging is to measure the
properties of the undisturbed rocks and fluids they contain. Geophysical logs can provide
information on lithology, the amount of water in a formation, formation density, zones of water
inflow, water quality, and other in situ parameters that cannot be derived from highly disturbed
drilling samples. A suite of geophysical log data, including deep-penetration methods, will more or
less complete the technical description of a borehole, but geophysical logging is a specialized field
best left to geophysical contractors or hydrogeological consultants.

A logging unit consists of a power supply, a receiver/data processing unit, and a cable on a powered
winch that lowers special sensor probes (‘sondes’) into the borehole to measure various properties.
The cable contains multi-conductors that transmit signals to the receiver console. Data, processed by
computer, can be shown as a geophysical record on a graphic display, which should consist of a
number of different structural, formation, and fluid logs. Specialized software packages enable
manipulation, interpretation, and comparison of data. Multiple-sonde geophysical (‘suite’) logging
can provide a substantial amount of information about the sub-surface conditions in and around a
borehole.

Figure 3.2: Borehole geophysical logging and logs

3.3 Borehole Construction Design:


As water is pumped out of a borehole, the water level in the hole falls. It may fall by an amount
known as the ‘pumping drawdown,’ which eventually stabilizes for that rate of extraction. If the
water level does not stabilize and continues to drop until the borehole is ‘dewatered,’ the hole is

36
being over-exploited. In this discussion it is assumed that boreholes are designed with the intention
of maximizing yield and efficiency, the normal requirements for everything other than hand-pump-
equipped holes.

- The maximum yield of a borehole is defined as that yield which the borehole can sustain
indefinitely before drawdown exceeds recharge from the aquifer.
- Borehole efficiency is technically defined as the actual specific capacity (yield per unit of
drawdown: say, liters per second per meter) divided by the theoretical specific capacity,
both of which can be derived from a pumping test. Specific capacity declines as discharge
increases.

 Borehole casing

Boreholes are constructed by inserting lengths of protective permanent casing. These are lowered or
pushed into the hole by the drilling rig to the required depth; the lengths of casing may be joined
together by means of screw threads, flange-and-spigot, gluing, riveting, or welding. Casing normally
extends up to the surface, with a certain amount (say 0.7 meter) standing above ground level.
Lengths of casing may be obtained in mild steel, stainless steel, and plastic (such as UPVC, ABS,
polypropylene, and glass-reinforced plastics).

Plastic casings are more fragile and deformable than steel casings (especially the screw threads), and
so should be used mainly for low-yield and shallow boreholes. The casing should be capable of
withstanding the maximum hydraulic load to which it is likely to be subjected, that is, about 10
kilopascals (kPa) for each meter that extends below the water level down to the maximum expected
drawdown.

Steel casing is available in a variety of grades and weights. Low-grade casing can be used for shallow
tube-wells, but heavy-duty, high-grade steel should be used for deeper boreholes (especially those
more than 200 meters deep) and when ground conditions hamper insertion (such as coarse
gravel/boulder formations). Special types of casing that can resist aggressive waters are also
obtainable, but stainless steel is the best means of combating corrosion. Casing is usually supplied in
standard lengths already equipped with screw threads or other jointing methods.

 Borehole well screens

When a borehole has been dug alongside a water-bearing zone, the casing installed in it must have
apertures that allow water to enter as efficiently as possible while holding back material from the
formation. These perforated sections are known as borehole or well screens; they come in sizes and
joints similar to casing, so can be interconnected with suitable plain casing in any combination, or
‘string.’ Screens can also be obtained with a variety of aperture (slot) shapes and sizes, from simple
straight slots to more complex bridge slots and wire-wound screens made with V-cross section wire.
Screen slots should be of a regular size, aperture, and shape because they might have to efficiently
prevent all particles of a certain size from getting through. Plain plastic casing can be easily slotted
with a saw or special slotting machine, but beware again of drilling contractors cutting irregular,
messy slots in steel casing with grinders or oxyacetylene torches. The open area of factory-made
plastic screens commonly exceeds 10% of total surface area, but rough-cut holes in mild steel casing
rarely take up more than 2 or 3%. Screen slots should be slightly smaller than the average grain size

37
of the aquifer fabric, and should allow water to enter the borehole at a velocity within the range 1 to
6 centimetres/second (0.01 to 0.06 meter/second). Entrance velocity is defined as the discharge rate
of the well divided by the effective open area of the screen. Too high an entrance velocity may lead
to screen incrustation, excessive well losses, and other damaging consequences of turbulent flow
conditions.

Formula to calculate the open area of a screen:


Open area of screen per meter of screen in cm2 = l*w*n/10, where l is length of slot in cm, w is
width of slot in mm, and n is number of slots per meter length.

For example, a minimum screen open area of 100 cm2 provides, roughly speaking, the minimum
entrance velocity for a yield of about 0.3 litres/second (about 4gallons/minute). In practice,
additional screen lengths should be included to allow for variations within the aquifer (which is
unlikely to be homogeneous) and in the borehole.

The most efficient well screens are the well-known ‘Johnson screens’ – continuous-slot types
manufactured with V-wire wound spirally around a cage of longitudinal support rods. The whole
structure may be composed of stainless steel or low-carbon galvanized steel. These wire windings
have been constructed such that the slots widen inwards, which significantly reduces rates of screen
clogging.

The effective open areas of Johnson screens are more than twice that of conventional slots, which
allows more water to enter per length of screen. Slot sizes of 0.15 to 3 mm, diameters of 1 ½" to 32",
and screen lengths of 3 meters and 6 meters are available. The different grades of screen are
suitable for a variety of borehole depths; the ends are plain (for welding) or screw-threaded.
Johnson screens allow yields of about 5 to 6 litres/second per meter length, so that a 6-metre-length
can give about 30 to 35 litres/second and a 12-metre-length twice as much. Most projects, and
especially those involving shallow or low-yield boreholes, require only basic PVC casing and screens
to be installed.

 Gravel pack

After the casing and screen string have been inserted, natural material will tend to fall from the walls
of the borehole into the annular space, forming a natural backfill or ‘gravel pack’ that helps to filter
incoming water. The screen slot sizes should be such that only the finer content of this backfill is
allowed into the borehole; this can be washed out during development, leaving the coarser portion
behind to act as a filter. Thus, an aquifer is suitable for the development of a natural gravel pack if it
is coarse-grained and poorly sorted, as many alluvial gravels are (a relatively rare situation). A
borehole drilled into an unstable aquifer formation, or into one that is well sorted, and with a high
proportion of fines (which would be apparent from the drill samples), will require an artificial gravel
pack around the screens. When the only screens available on site are of a slot size larger than the
average grain size of the aquifer, then a gravel pack should be installed. Unfortunately, time and
other constraints do not normally allow a detailed grain-size analysis of the aquifer fabric to be
carried out in the field; so, a degree of intuition is required here. If there is any uncertainty, install an
artificial gravel pack.

- Artificial gravel pack

38
Ideally, an artificial gravel pack should consist of clean, rounded, quartz ‘pea’ gravel supplied in bags;
grains the size of small household peas are generally suitable. Coarse, well-worn river sand is often
ideal; the grains should be a little larger but no more than twice the screen slot size. Being smooth
and spherical, the grains should run easily down into the annular space without clumping and
leaving gaps of air (a little water often helps).

The standard practice is to produce a 3 to 4-inch wide annular space for the gravel pack (say, a 6"
screen in a 12" hole); the casing/screen string must be centered in the borehole. Most boreholes are
not perfectly straight, so the casing will almost invariably be in contact with the wall in some places
unless it is centralized. This is achieved by using manufactured centralizers (such as flexible ‘wings’)
or some other suitable alternative.

Before pouring gravel pack material into the annular space, which must be done smoothly and
without haste, calculate the volume of annular space (it is reassuring to see that the correct volume
of gravel has been installed). Again, an accurate log of borehole size changes is required here.
Pouring gravel pack into a borehole with a high water level usually results in displaced water rising in
the hole and overflowing. Water overflow will abruptly stop as the screen becomes covered by
gravel. Continue to pour gravel until you are certain that the top of the pack is well above the top of
the screen.

Annular volume between borehole diameter D and casing/screen diameter d (D and d both in
inches), length L (in meters) = ~½L(D2-d2) in liters.

Thin gravel packs (less than 50 mm, or 2", thick) may be installed to act as a formation stabilizer only
in conditions such as those associated with a fractured or slightly weathered consolidated aquifer. It
should also be noted that gravel packs greater than 150 mm (6") thick will make borehole
development more difficult, especially if a drilling mud lining has to be removed.

Introducing a natural or artificial gravel pack into a borehole will reduce the effective open area of
the screen, because now the open area (porosity) of the system at the aquifer/screen interface will
be limited by that of the packing in the annular space rather than the screen. Well-rounded grains of
uniform size (as in the ‘ideal’ gravel pack) have among the highest primary porosities (around 40%)
and permeability’s (20 or more meters/day) in unconsolidated sediments; in practice, the figures are
probably much lower. The effective open area of the adjacent aquifer is more likely to be around
10%. The result is that one should assume the effective open area of a screen, even with a gravel
pack of good quality, to be roughly half the actual screen open area. The recommended minimum
actual open area of any installed screen is around 10%.

The essence of borehole design is deciding the combination of plain casing and screens to be
inserted and the type of screen to be used, and whether a gravel filter pack (or thin formation
stabilizer) is required. Table 3.2 attempts to summarize these decisions for a variety of ground
conditions that are likely to be encountered during drilling. An ‘open hole’ design is one in which no
screen or gravel pack is used in the area of the aquifer, but all boreholes that this writer has
encountered have required casing, to at least stabilize the superficial soils or weathered zone. Open
holes are suitable mainly for hand-pumps, because of the danger of a powerful motor pump sucking
in debris even from a stable hard rock formation. If a stable formation is encountered some way

39
below a water strike, reducing the penetration rate, some extra borehole depth, to act as an open-
hole sump, can be created a little more quickly by reducing drill bit size.

Table 3.2: Choice of screens and gravel pack for various ground conditions

Aquifer Crystalline (narrow Consolidated Unconsolidated Stable but with


characteristics fissures (small fissures/Caverns
or joints) voids/porosity)
Thin (<100m) No screen or gravel No screen or gravel Screen with high Screen with high
pack normally pack normally open area and open area.
required (open required (open gravel pack Gravel pack
hole). Screen plus hole). Screen plus required. Might required if
formation stabilizer formation stabilizer develop natural caverns contain
might be necessary might be necessary gravel pack if loose sediment.
if formation if formation aquifer
fractured. fractured. homogeneous.
Thick (>100m) Long screen with Long screen with Long screen (or Long screen (or
small open area small open area multiple multiple screens)
(10%). No gravel (10%). No gravel screens) and and gravel pack
pack (except, pack (except, gravel pack required if
possibly, formation possibly, formation required. Might caverns contain
stabilizer if stabilizer if develop natural loose sediment.
formation formation gravel pack if
fractured). fractured). Aquifer
homogeneous.
Deep (>200m) No screen or gravel No screen or gravel Strong (steel) Strong (steel)
pack required pack required casing/screens casing/screens
(except, possibly, (except, possibly, and gravel pack and gravel pack
formation stabilizer formation stabilizer required. At required if
if formation if formation depth, a natural caverns contain
fractured). fractured). gravel pack loose sediment.
might be less
likely to
develop.
Corrosive water As above, but use As above, but use As above, but As above, but
(e.g. high plastic or stainless plastic or stainless use plastic or use plastic or
salinity, low pH, steel casing/screen steel casing/screen stainless steel stainless steel
High (s). (s). casing/screen casing/screen
temperature) (s). (s).
Encrusting As above, but use As above, but use As above, but As above, but
water (e.g. iron/ high open area high open area use high open use high open
carbonate screen (s) to screen (s) to area screen (s) area screen (s)
enriched) reduce entrance reduce entrance to reduce to reduce
velocities. velocities. entrance entrance
velocities. velocities.

3.4 Pump selection


One also needs to consider what type of pump is to be used in the borehole as well as its size and
the casing and screen diameters needed to house it. Relatively low-yield boreholes intended for

40
hand-pumps with submerged pistons are rarely more than 30 meters deep and need not be of large
diameter (4 inches can be enough). Certain types of hand-pump can be fitted with counterweighted
operating handles, to make drawing from deeper water levels less exhausting (mainly for women in
developing countries). Mechanical reduction of the submerged piston stroke can make drawing
water easier, but as less water is drawn by each stroke of the pump, filling a container takes longer.
Shallow boreholes often tap weathered zone aquifers above hard, impervious clays or bedrock (such
as granite). The standard practice is to place the pump inlet slightly above the upper end of the
screen. For hand-pumps, this does not matter much, the pump inlet could be lowered into the upper
part of the open hole/sump (by attaching extra pipes) if there is a significant fall of static water level
in the borehole because of prolonged drought.

More powerful types of pump, such as electric submersibles and rotary positive displacement
pumps, must always be installed inside protective plain casing: screens are not appropriate for this
purpose. The internal diameter of the pump chamber should be at least 5 centimeters (about 2
inches) greater than the external diameter of the pump.

3.5 Sealing the Borehole


The borehole structure must be sealed at the top of the casing with a ‘sanitary seal’; it can also be
sealed at the bottom to completely eliminate the possibility of material entering through any means
other than the screen(s). With plastic casing this can be done by attaching a ‘closing cap’ to the
bottom end.

It will be readily seen that producing a point at the lower end will assist lowering of the casing into
the borehole, especially if difficult drilling conditions such as coarse gravels, stones, or boulders are
encountered. Furthermore, closing off the casing ensures that gravel pack material cannot enter the
borehole from the annular space.

Figure 3.3: Sealing the bottom end of mild steel casing by the welded ‘saw-teeth’ method. Typical
length of teeth 0.5 to 1m

41
3.6 Borehole Development
At this stage it might seem almost impossible that clean, potable water could emerge from the mess
being pumped out of a new borehole, but provided it has been properly constructed, good water
will indeed appear. Whether a borehole has been drilled using mud or air circulation, it will have to
be cleaned out. After the installation of the permanent casing, screens, and gravel packs (if any),
dirty water, mud, crushed rock, oil (from the drilling machinery), and perhaps other debris will be
left in the hole. Development can also repair damage done to the adjacent aquifer by the drilling
process, develop the aquifer (increase transmissivity), and enhance the performance of the
borehole.

Development has two broad objectives: (1) repair damage done to the formation by the drilling
operation so that the natural hydraulic properties are restored, and (2) alter the basic physical
characteristics of the aquifer near the borehole so that water will flow more freely to a well.”

Drilling tools smear crushed rock and clay all over the walls of the borehole, and the drilling process
forces dirty water and clay into the rock matrix around the hole, sealing off many water entry points
from the aquifer. If these matters are not remedied, the borehole’s performance will be very poor; it
should also be noted that the filthy water that would be pumped out – were things to be left
unaltered – would damage a pump quickly. Mud rotary drilling leaves a cake of firm clay (‘mud cake’)
on borehole walls, often of a thickness up to one centimeter, which can effectively choke aquifers.
Removal of this layer is not easy, requiring ‘violent’ or ‘aggressive’ methods, and should not be
hurried or abandoned prematurely. Moreover, since mud-caked walls will have been partly isolated
from the borehole internal space by casings, screens, and possibly even gravel packing, cleaning up
will be even more difficult.

Development also encourages a gravel filter pack to settle properly, eliminating voids, which may
necessitate topping up the gravel pack a little. The process should continue until the water being
discharged from the borehole is, in the judgment of the supervisor, as clean as possible. Small
particles of sand might occasionally issue from the borehole, but cloudiness (turbidity) of the water
should have disappeared before development is stopped. Excessive production of sand might be
caused by voids in the gravel packing or by damaged casings or screens. Turbidity is generally caused
by colloidal clay or micro-organic particles; and can, in the latter case, result in unpleasant tastes and
odors and in organic growths like slimes.

- Development methods:

In most cases, development entails the surging and blowing of compressed air; the process may be
helped along by the use of additives, which can assist in breaking down drilling mud.

a) Mud dispersants

Bentonite-based muds are particularly difficult to remove; organic polymers are biodegradable and,
in theory, are destroyed by bacteria with which wells can be seriously infected. Chlorine-rich
compounds are effective mud dispersants, as well as bacterial disinfectants. Bentonite is more
efficiently dispersed using polyphosphate compounds like Calgon (brand name for water softener).
This chemical consists of granular or powdered sodium hexametaphosphate, a hygroscopic (water

42
attractant) material that destroys the cohesiveness and plasticity of clay particles. Granular Calgon
should be dissolved in hot or boiling water (about one kilogram of the chemical in 40 to 50 litres of
hot water): the usual dosage is 10 to 50 kilograms of Calgon per cubic meter of water estimated to
be in the borehole. Try to leave the dispersant in the hole overnight to allow the solution time to
permeate into aquifer formations. Higher concentrations are needed to remove the more resistant
mud cakes. Leftover mud is more easily removed when the borehole is washed out.

b) Acid treatment

Development of wells drilled into calcareous (limestone, chalk, or dolomite) formations can be aided
by the use of certain acids (such as hydrochloric acid, HCl) that dissolve pulverized carbonate smear
on borehole walls. Acid treatment can widen and clean carbonate aquifer fissures even when they
are tens of meters away from the borehole.

c) Surging

The technique of surging (or surge pumping) consists of forcing water up and down a borehole and,
more importantly, back and forth through the screens, gravel pack, and adjacent aquifer matrix.
Surging can be conducted with pumps, but this is not advisable because of the possibility of
damaging debris entering the pump. Furthermore, a powerful pump could dewater a borehole if the
screen is blocked by mud cake and cause the screen to collapse inwards as a result of hydraulic
pressure in the annular space. In practice, surging is almost invariably accomplished using
compressed air – that is, air-lift pumping using the drill pipes on the rig (with the drill bit removed)
and a compressor.

d) Blowing yield

Between bouts of surging, air is blown into the borehole and its pressure adjusted so that the
outflow of water is more or less equal to the inflow: this is known as the ‘blowing yield.’
Measurement of this flow gives an indication of the performance of the borehole, and helps to
design a subsequent pumping test Blowing yield can be easily measured if the drillers arrange that
all the water (in practice, most of the water) being ejected from the casing can be led along a
shallow channel or pipe in the ground into a measuring device, such as a bucket in a pit or a V-notch
weir. Filling of a bucket of known volume can be timed to give the discharge, which, in the opinion of
this writer, is a much simpler and more reliable method, less prone to error. A V-notch is an opening
or weir in the form of an inverted triangle in the side of a tank or reservoir; it is used to determine
surface- water flows. Water whose discharge is to be estimated is allowed to flow over the notch
weir, and the rate of flow can be calculated by measuring the depth of the water over the apex of
the notch. Various publications and websites give rates of flow tables for notches of different apex
angles (the most common is 90°).

e) Air-lift pumping

Basic air-lift pumping suits well development because it does not involve mechanical parts, which
can be damaged by debris. The air line from a compressor may or may not be inside a rising main
with a discharge outlet, and the end of the air-line is at such a depth that at least 50% of its length is
submerged in the borehole water. Pumping can be stopped and started at intervals by shutting off
the air supply at the compressor, which induces surging. Violent releases of pressure pull water in

43
through the screen (and gravel pack); then, when the valve is closed, pressure rises to force water
back out through the screen. So the pump/non-pump (back and forth) cycle acts like a surge block,
and material is effectively loosened and removed from the borehole/aquifer boundary. Normally,
the drill pipes are lowered to the bottom of the borehole and surging commences from there, with
dirty water being blown vigorously out of the casing at the surface. The pipes should be raised and
lowered at intervals so that different parts of the screen are subjected to surge action.

f) Jet washing

After the mud drilling of what may be a low-yield borehole, or if there is no compressor, another
method that may be used is jetting: washing of the well face with high-pressure water jets. The mud
pump or a separate jetting pump is used to inject clean water into the borehole down the drill pipes
from a source such as a river, lake, bladder tank, or mobile bowser. At the bottom of the drill string
is a jetting nozzle tool, which produces the high-pressure jets; ideally this can be raised, lowered,
and rotated in the borehole.

The jets are directed horizontally at the screen slots or the borehole wall. This method can be used
alongside an air percussion rig, but a separate jetting rig will be required, either piped into the drill
string, or with a separate injection pipe lowered down the borehole. Jetting is an effective cleaning
system for screens and sections of open hole, but is less effective at penetrating the aquifer matrix
than surging. Of course, a blowing yield cannot be obtained from jet development, because water is
being pumped into the borehole.

Figure 3.4: Borehole development using compressor.

44
3.7 Borehole Completion
a. Sanitary seal

With borehole cleaning completed, the final job is the construction of a sanitary seal, which, as the
name suggests, seals the borehole from surface contamination. This should also be the responsibility
of the drilling contractor, and written into the work agreement. At least the uppermost two meters
or so of annular space (probably that section formerly protected by the conductor pipe during
drilling) should be cleaned out and dug into a fresh larger hole – perhaps square in shape –
surrounding the permanent casing.

Below this, the borehole annular space above the gravel pack will have been backfilled with a plug of
ordinary gravel, chippings, bentonite granules, or even just cuttings from the borehole. The fresh
hole for the sanitary seal can then be filled with concrete grout up to, or preferably slightly above,
ground level. For boreholes with high static water levels, capped by permeable superficial soils, little
more can be done other than to take care not to spill wastewater around the borehole. Finally, the
supervisor should confirm the completed borehole’s total depth and static water level with a plumb
line and a dip meter.

Then the top of the casing must be sealed with a locked cap or welded plate, on which the borehole
identification number may be inscribed.

b. Pumps and test pumping

After drilling has been completed and the sanitary seal put in place, borehole test pumping is carried
out. It has the following objectives:

 To measure the performance of the borehole


 To determine the efficiency of the borehole, or variation of its performance under different
rates of discharge
 To quantify aquifer characteristics, such as transmissivity, hydraulic conductivity, and
storativity.

In remote locations, a supervisor or hydrogeologist may require some indication of a borehole’s


performance if the type of pump to be installed has not yet been determined. Blowing-yield during
development will tell the supervisor that a particular borehole has the potential for good
production. Low-yield boreholes (less than 0.5 litre/second) will require only a hand-pump for
extraction and do not need to be tested, and time should not be wasted doing this. Boreholes of
very low yield (less than 0.2 litre/second), but with high water levels (say, 10 meters or less) may not
be suitable for long-term water supply, but can give an indication that groundwater exists in the
area, which could be exploited by dug wells. Here, community participation should be encouraged.

Main types of pumping test

There are many different types of pumping test from which to choose:

1. Step test: Designed to establish the short-term relationship between yield and drawdown
for the borehole being tested. It consists of pumping the borehole in a series of steps, each

45
at a different discharge rate, usually with the rate increasing with each step. The final step
should approach the estimated maximum yield of the borehole.
2. Constant-rate test: Carried out by pumping at a constant rate for a much longer period of
time than the step test, and primarily designed to provide information on the hydraulic
characteristics of the aquifer. Information on the aquifer storage coefficient can be deduced
only if data are available from suitable observation boreholes.
3. Recovery test: Carried out by monitoring the recovery of water levels on cessation of
pumping at the end of a constant-rate test (and sometimes after a step test). It provides a
useful check on the aquifer characteristics derived from the other tests but is valid only if a
foot-valve is fitted to the rising main; otherwise water surges back into the borehole.

These tests can be carried out singly or in combination. A full test sequence usually starts with a step
test, the results of which help to determine the pumping rate for the constant-rate test, with a
recovery test completing the sequence. The test design can be adapted for use in small, medium or
large boreholes, the main differences being the pumping rates, the length of test and the
sophistication of the monitoring system.

- Preparations for test pumping

Before commencing any pumping test, there are certain basic preparations that should be made.
These include gathering information about the borehole or well that is about to be tested. The
outcome of the preparations may influence the choice of test and will certainly increase the value of
the results obtained from the test.

1. Basic monitoring equipment

The two parameters that must be measured in any pumping test are the water level in the pumped
borehole and the rate at which water is being abstracted (pumped or bailed). The basic equipment
necessary for monitoring these two parameters is as follows:

- Monitoring water levels

The hand-held water-level monitor, commonly known as a “dipper,” is the most practical, robust and
easily available method of monitoring water levels in boreholes and wells. The dipper probe is
lowered down the borehole, and when it reaches the water surface, an electrical circuit is completed
and a ‘bleep’ is heard. The water level is then read off a graduated tape, usually to a resolution of
the nearest centimeter. The water level is typically recorded in meters below a local measuring
datum, such as the lip of the borehole casing. Manual dipping is widely trusted as a reliable and
relatively trouble-free way of obtaining water-level data, but it is not without its problems, for
instance:

 The graduated dipper tape can suffer from stretch due to age, temperature or misuse,
introducing a systematic inaccuracy, especially if different dippers are used.
 If the water level is falling or rising quickly, as during the early stages of a pumping test,
it can be difficult to take manual readings fast enough, although this problem improves
with practice.

46
 It can be difficult to get a ‘clean’ water-level reading from the borehole, especially if
there is water cascading down the side of the borehole, or there is turbulence at the
water surface.
 The dipper can become stuck, entangled or wrapped around the pump, rising main,
electrical cables, or other items down the borehole. This can be avoided by the use of a
dip tube (an open-ended plastic tube installed in the borehole specifically for the dipper
to go down), which also solves the problem of cascading water and turbulence.

- Monitoring pumping rates

There are many methods of measuring pumping rates, of which the most common, or the ones most
likely to be of use to hydrogeologist, are as follows:

1. Bucket and stopwatch: The simplest method of measuring relatively low pumping rates is to
use a bucket and a stop -watch. Arrangements are made for the discharge from the pump to
flow freely into a bucket of known volume, and the time taken for the bucket to fill is
recorded. The flow rate is then calculated by dividing the volume of the bucket by the time
taken to fill it. For this method to be precise, it should take a minimum time of about 100
seconds to fill the bucket. If necessary, use a larger container of known volume, such as an
oil drum.
2. Flow meters: Where more sophisticated equipment is available, pumping rates can be
measured using flow meters, of which there are various types. This uses spring-loaded
pistons that are deflected by the flow of water, and the flow rate is read off the graduated
scales. It is important to double-check the flow rate by using another method, to operate the
gauge correctly, and to keep the equipment in good condition.

- Other equipment available

Before choosing the type of test to conduct, hydrogeologist should establish what equipment is
available, practicable or affordable. In addition to the water-level and flow-monitoring equipment
described above, potential equipment includes the following:

- Submersible pump.
- Generator
- Rising main (GI Pipes)
- Manually-operated valves
- Discharge pipes
- Water-quality monitoring equipment
- Surface water flow-gauging equipment
- Bailers

Whatever equipment is used, it should be maintained in good condition and used correctly
(according to the manufacturer’s instructions). It should also be designed so that it can be operated
safely, and if necessary, should be calibrated to give reliable and accurate data. Equipment should be
tested when it is in position, before a pumping test is begun, in order to ensure that it is all working
properly and to determine pump or valve settings that will give appropriate pumping rates.

47
Figure 3.5: Pumping test process and data collection, Somalia.

2. Information collection

When planning a pumping test, it is useful to gather together all the information that can be found
about the aquifer and the borehole itself. The results from the pumping test will be added to the
information, and will improve your understanding of the local groundwater system. Try to collect
information on the following:

- Are there any other boreholes in the area (especially in the same geological formation)?
- What are the typical water levels and yields, and what is the quality of the water, from those
boreholes?
- Are the boreholes being pumped at the moment? Ideally, other boreholes in the area should
not be pumped during your pumping test, or for at least 24 hours before the start of the test
(and they might serve as observation boreholes).
- What drawdown can be expected in the borehole about to be tested? At what depth should
the pump intake be set so that it remains well below the water level during the test?
- Are the rocks crystalline basement, volcanic, consolidated sediments or unconsolidated
sediments?
- Is the aquifer confined, unconfined or leaky?
- How deep is the borehole, and of what diameter? Has solid casing, screen or gravel pack
been installed?
- Installed equipment: If a pump is already installed in the borehole, what are its type and
capacity, and at what depth is the pump’s intake? Can the pumping rate be varied?
- Does the water level vary much from wet season to dry season?
- In the period before the test takes place, is the water level already falling or rising or is it
stable?
- What is the current water level? And does the water level respond to rainfall?
- Is the water safe for drinking, and does the water quality change over time?

There may not be much information available, in which case the planned pumping test will be the
starting point for your understanding of the local groundwater system. It is good discipline to write
down all the data collected so the work does not have to be duplicated in the future. A form for
recording basic borehole information can be found in Annexes.

48
Water hygiene

The links between water quality and public health are well known, and much has been written on
the importance of good sanitation and hygiene, the protection of groundwater resources and the
improvement of traditional water sources. It is essential that the principles of good water hygiene
not be forgotten during test pumping. Things to look out for include the following:

- Make sure that contaminated water cannot enter the borehole during test pumping,
especially when using temporary pumping equipment in an open borehole, and when there
is water spillage or run-off from rainfall during the test.
- Provide adequate sanitation facilities for the field staff and test-pumping crew, and insist
that they follow good hygiene practices, particularly hand-washing. Also, if one of the
workforces has symptoms such as persistent diarrhoea or prolonged unexplained fever,
recommend that he or she not work on the test.
- Ensure that all equipment that will come into contact with the groundwater or the wellhead
(pumps, pipes, valves, dippers, samplers, bailers, ropes, tools, etc.) has been cleaned
properly before use, especially if it has previously been in contact with contaminated water.
Don’t forget to flush contaminated water out of pump chambers, valves and rising mains.
- If mechanical equipment such as mobile generators, air compressors and drilling rigs is being
used, make sure that it is in good condition, with no leaks of hydraulic fluid, lubricating oil or
diesel or other fuels. Items such as drip-trays and absorbent mats should be available in case
of leaks or spills.
- Make adequate arrangements for the temporary storage of drums or containers of fuel, oil
or other hazardous substances, and enforce good practices for re-fuelling, so that there is no
danger of contaminating the water supply during the test pumping.
- Make sure that the borehole has been secured when you leave it, so that foreign objects,
animals or dirty water cannot enter.

Water-quality monitoring

Although the focus of most pumping tests is on monitoring water levels and pumping rates, water-
quality monitoring can be an important part of the test, and should be considered at the planning
stage. As mentioned above, there is a strong link between water quality and public health, and even
if a certain borehole can sustain a high yield, the water produced by it may be unsuitable for
drinking. Water-quality monitoring during a pumping test can help answer key questions such as:

- Is the water quality suitable for the intended use (particularly for drinking)?
- Is the water quality stable in the long term?
- Does the water quality change with the pumping rate?
- Is there a pumping rate above which the water quality suddenly deteriorates?
- Is any treatment necessary before the water can be used?
- Is the groundwater vulnerable to pollution, or to ingress of contaminated surface water?

When planning a pumping test, therefore, take into account the following practical issues:

49
- Some parameters, such as electrical conductivity, temperature, pH and turbidity, must be
measured at the well-head, or very soon after the water has come out of the ground.
Readings are usually taken with hand-held probes.
- Measuring some parameters involves collecting samples in bottles, for subsequent analysis
by a field testing kit or in a laboratory. Make sure that there is a suitable sampling point
included in the discharge arrangements, so that good clean samples can be obtained,
without splashing from the ground, for example.
- Unless there is a field testing kit available, samples for microbiological analysis need to be
kept cool and must reach a laboratory within a certain time limit. Is this practicable?
- Electrical conductivity can be correlated with total dis -solved solids, thus providing a useful
field indicator of water quality (especially salinity). Observe how the conductivity changes
during the test, particularly during a step test as the pumping rate progressively increases.
- Don’t forget simple clues such as the appearance, smell and color of the water. Make notes
of these during the test. Do they change?
- Some boreholes produce sand, which can damage pumping equipment and fill up storage
tanks. If this is suspected, collect a sample of the discharge water in a clear container. Set it
aside, allow the sand to settle, and then measure the depth of sand. Using the same
container, take more samples at intervals, and record how the sand content changes.
- With all water-quality sampling, record basic information such as the time and date of
sampling, and the name and location of the borehole.

I. Step Test

The step test (sometimes referred to as the step-drawdown test) is designed to establish the short-
term relationship between yield and drawdown for the borehole being tested. It consists of pumping
the borehole in a sequence of different pumping rates, for relatively short periods (the whole
sequence can usually be completed in a day). There are many different ways to perform a step test,
but the most common practice is as follows:

- Start with a low pumping rate, and increase the rate with each successive step, without
switching off the pump between steps.
- Aim for four or five steps in total, with the pumping rates roughly spread equally between
the minimum and maximum rates.
- All steps should be of the same length in time, with somewhere between 60 and 120
minutes per step being common.
- The pumping rate for the final step should be at or beyond the intended operational
pumping rate when the borehole is fully commissioned. Of course, this depends on whether
the pump being used for the step test is capable of that pumping rate.

Figure 3.6 illustrates a typical series of pumping rates (Q) and the behavior of the water level. It is
immediately clear why it is called a step test.

50
Figure 3.6: Outline of step test

As mentioned above, step tests are primarily designed to provide information about the borehole
performance characteristics (the yield-drawdown relationship). It is possible to use step-test results
to estimate aquifer transmissivity. However, the present guidelines focus on borehole performance.
Step-test results are not very good for predicting the behavior of a borehole under long-term
pumping, for which a constant-rate test should be used.

- Step-test procedure

Assuming that all the equipment is ready and people have been assigned their tasks, the procedure
for conducting a step test is as follows:

1. Choose a suitable local datum (such as the top of the casing) from which all water-level
readings will be taken, and measure the rest-water level. The water level must be at rest
before the start of the test, so the test should not be conducted on a day when the borehole
is being drilled or developed, or when the equipment is being tested.
2. Open the valve to the setting for the first step (determined by prior experiment, as
described above) and switch the pump on, starting the stopwatch at the same time. Do not
keep changing the valve setting to achieve a particular pumping rate (a round number in
litres per minute, for example). Rather, aim for an approximate rate and measure the actual
rate (see 4) below).
3. Measure the water level in the borehole every 30 seconds for the first 10 minutes, then
every minute until 30 minutes have elapsed, then every 5 minutes until the end of the step
(the length of each step having been decided during the test preparations). If you miss the
planned time for a water-level reading, write down the actual time the reading was taken.
Record all the readings on the standard step-test form (Annexes).
4. Measure the pumping rate soon after the start of the step, and then at intervals during the
step (every 15 minutes would be reasonable). If there is a noticeable change in the rate of
increase of drawdown or the pump sounds different, then measure the pumping rate at
those times as well. If the pumping rate changes significantly (say by more than 10%), then

51
adjust the valve setting to maintain as steady a pumping rate as possible throughout the
step. Be careful not to over-adjust and make the problem worse.
5. At the end of Step 1, open the valve further, to the set-ting for Step 2, note the time (or
restart the stopwatch) and repeat the procedures for measuring water levels and pumping
rates (see 3) and 4) above).
6. Repeat the procedure for subsequent steps, progressively increasing the pumping rate for
each step.
7. At the end of the final step (which will probably be Step 4 or 5), switch the pump off, note
the time (or restart the stopwatch), and measure the water-level recovery at the same
measurement intervals as for measuring the drawdown in each step. Continue for at least
the length of a step, and ideally for much longer, until the water level approaches the pre-
test level.

- Analysis and interpretation

There are many different ways to analyses step-test results, some of them very sophisticated, but
the present guidelines describe simple methods that concentrate on borehole performance.

Jacob’s Equation

The theory of groundwater hydraulics assumes that during pumping from a borehole, the flow
conditions in the aquifer are laminar. If this is the case, then drawdown in the borehole is directly
proportional to the pumping rate. However, turbulent flow may occur in the aquifer close to the
borehole if pumping takes place at a sufficiently high rate, and the final path of the water from the
aquifer, through the gravel pack and screen, into the borehole and the pump intake itself is nearly
always subject to turbulent flow conditions. This results in ‘well losses,’ meaning that additional
drawdown is required to get the water into the pump. If turbulent flow is present, Jacob suggested
that drawdown in a borehole can be expressed by the following equation:

s = BQ + CQ2

Where: s is the drawdown, Q is the pumping rate, and B and C are constants. If all the terms in
Equation above are divided by Q, it becomes:

s/Q = B + CQ

Which is the equation of a straight line (if s/Q is plotted against Q on linear graph paper). Note that
the term s/Q is called the specific drawdown, and the inverse (Q/s) is called the specific capacity. So,
for this analysis of the step test results, do the following:

1. Calculate the average pumping rate for each of the steps in the test (take all the
measurements of the pumping rate recorded during Step 1 and calculate the average; now
repeat the procedure for the other steps). If there were five steps in the test, you should end
up with five values for the pumping rate (Q1, Q2, Q3, Q4, and Q5).
2. Take the water-level readings from the very end of each step (in meters below datum) and
convert them into drawdowns, by subtracting the rest-water level. Again, for a test with five
steps, you should end up with five values of drawdown (s1, s2, s3, s4 and s5).

52
3. Calculate the specific drawdowns from the pairs of values (s1/Q1, s2/Q2, etc.). Now draw a
graph of s/Q against Q on linear graph paper (by plotting s1/Q1 against Q1, s2/Q2 against Q2,
etc.), as shown in Figure 3.4 below. Draw a best-fit line through the points (the solid blue
line in Figure 3.4): the intercept of the line on the y-axis represents the constant B, and the
gradient of the line represents the constant C.

The values for B and C can then be used in Equation above to calculate the expected drawdown for
the other pumping rates or, with a little rearrangement of the equation, the expected pumping rate
for a given drawdown. If the step test is repeated at a later date and the best-fit line (in Figure 3.7)
has shifted vertically (different B) but has the same gradient (C), that represents a change in aquifer
conditions. If B is the same but C has increased, then the borehole performance has deteriorated,
probably due to a factor such as clogging of the screen. Jacob’s equation is often used to calculate
borehole efficiency, but there is a lot of confusion about borehole efficiency, and the reality is much
more complex. When analyzing step-test results, it is much more practical to concentrate on
understanding the borehole performance characteristics, as will now be described.

Figure 3.7: Step-test analysis

II. Constant-rate Test

The constant-rate test is the most common type of pumping test performed, and its concept is very
simple: the borehole is pumped at a constant rate for an extended period (from several hours to
several days or even weeks) while the water levels and pumping rates are monitored. If the most
value is to be gained from constant-rate tests, water levels should be monitored in an observation
borehole as well as in the pumping borehole (or better still, several observation boreholes at
different distances from the pumping borehole). As this is rarely possible in most places, the present
guidelines concentrate on what to do with the data obtained from the pumping well alone.

Data from constant-rate tests can be analyzed to derive the transmissivity of the aquifer. The
storage coefficient of the aquifer can be calculated only if data from observation boreholes are
available, which is assumed not to be the case here.

53
- Constant-rate test procedure

Assuming that all the equipment is ready and people have been assigned their tasks, the procedure
for conducting a constant-rate test is as follows:

1. Choose a suitable local datum (such as the top of the casing) from which all water-level
readings will be taken, and measure the rest-water level. The water level must be at rest
before the start of the test, so the test should not be conducted on a day when the borehole
is being drilled or developed, or when the step test is taking place.
2. Open the valve to the appropriate setting and switch the pump on, starting the stopwatch at
the same time. Do not keep changing the valve setting to achieve a particular pumping rate
(a round number in litres per minute, for example). Rather, aim for an approximate rate and
measure the actual rate (see 4) below).
3. Measure the water level in the borehole every 30 seconds for the first 10 minutes, then
every minute until 30 minutes have elapsed, then every 5 minutes until 2 hours have
elapsed. After 2 hours, observe how quickly the water level is still falling, and decide an
appropriate frequency for water-level readings until the end of the test. If the water level is
falling very slowly, then a reading every 30 minutes or even every hour may be sufficient. If
the test is to continue for several days, review the measurement frequency depending on
the behaviour of the water level. If you miss the planned time for a water-level reading,
write down the actual time the reading was taken. Record all the readings on the standard
form (Annexes).
4. Measure the pumping rate soon after the start of the test, and then at intervals during the
test (every 15 minutes would be reasonable for the first few hours, then decide a suitable
frequency for the remainder of the test). If there is a noticeable change in the rate of
increase of drawdown, or if the pump sounds different, then measure the pumping rate at
those times as well. If the pumping rate changes significantly (say by more than 10%), then
adjust the valve setting to maintain as steady a pumping rate as possible throughout the
test, but be careful not to over-adjust and make the problem worse.
5. At the end of the test, switch the pump off, note the time (or restart the stopwatch), and
measure the water-level recovery at the same measurement intervals as for measuring the
drawdown. Continue until the water level has recovered to the pretest level, or at least
approaches that level. See the next chapter for a full explanation of the recovery period.

If there is a problem during the test, such as an interruption to the power supply or a pump failure,
then use your judgement, depending on when the problem occurs and how long it is likely to last.
For example, if something goes wrong in the first few minutes, wait for the water level to recover
and start again. If the failure occurs well into the test and can be solved quickly, just restart the
pump and carry on. If it is going to take a long time to solve, it may be better to allow full recovery of
the water level and start again. For long constant-rate tests, it is especially important to ensure that
there is an adequate fuel supply to last the planned duration of the test.

- Analysis and interpretation

The method of analysis presented here is called the Jacob (sometimes referred to as the Cooper-
Jacob) straight-line method, which is based on a simplification of the Theis method. The procedure is
as follows:

54
1. Prepare a graph on semi-log graph paper, with water levels on the (linear) y-axis, in meters
below datum, and time on the (logarithmic) x-axis (time since the start of pumping, in
minutes). See Figure 3.8. Note that draw-downs can be plotted on the y-axis instead of
water levels, if preferred this does not affect the analysis.

Figure 3.8 Constant-rate test analyses

2. Plot the water levels against time for the duration of the test. The data should plot roughly
as a straight line. Draw a best-fit line through the data, ignoring the early data and
concentrating on middle to late data.
3. From this line, measure a parameter known as Δs, which is the difference in water levels (in
meters) over one log cycle (best understood by looking at Figure 3.5).
4. Calculate the average pumping rate for the duration of the test, Q, in m3/day.
5. Insert the values of Q and Δs into the formula below to calculate the transmissivity T. Make
sure that the correct units have been used, in which case the units of T will be m 2/day.

T = 0.183 Q/Δs

When a fine line was fitted to the data points in 2) above, the early data were ignored because they
tend to be affected by the volume of water stored in the borehole itself, and the points would
probably not have fallen on the straight line. If there are other deviations from the straight line, first
look for explanations such as sudden changes in the pumping rate or heavy rainfall during the test.
Different types of deviation from the standard Jacob straight line are commonly observed, as shown
in Figure 3.9 and described below (the figure and the explanations are taken from MacDonald et al
[2005]).

- Gradual decrease in drawdown: This occurs because the aquifer is gaining water from
another source, either because the aquifer is leaky, or because the expanding cone of
depression has intercepted a source of recharge, such as surface water. This is an
encouraging sign for the borehole as a sustainable water source, and the transmissivity value
should be measured using the data before the leakage is observed.

55
- Gradual increase in drawdown: This indicates that the aquifer properties away from the
borehole are poorer than those closer to the borehole. This can be because the aquifer is
limited in extent (in other words, the expanding cone of depression has encountered a
hydraulic barrier), or because shallow parts of the aquifer are being dewatered. This is not
an encouraging sign, and indicates that less water is available than appeared at first. If the
test has been continued long enough (for the data to stabilize on a new straight line),
calculate the transmissivity from the late data.
- Sudden increase in drawdown: This can result from the dewatering of an important fracture
or the interception of a hydraulic barrier. Such behavior is of serious concern and indicates
that the borehole may dry up after heavy usage or during the dry season. All is not lost,
however, as the borehole may still be usable, at a lower pumping rate.

Figure 3.9: Deviations from straight line during constant-rate test (From: MacDonald et al [2005])

56
III. Recovery Test

The recovery test is not strictly a pumping test, because it involves monitoring the recovery of the
water level after the pump has been switched off. We have already come across it in the final stages
of the procedures for undertaking step tests and constant-rate tests. It has been given a chapter to
itself because recovery data are not always given the attention they deserve. Recovery tests are
valuable for several reasons:

- They provide a useful check on the aquifer characteristics derived from pumping tests, for
very little extra effort – just extending the monitoring period after the pump has been
switched off.
- The start of the test is relatively ‘clean.’ In practice, the start of a constant-rate test, for
example, rarely achieves a clean jump from no pumping to the chosen pumping rate.
Switching a pump off is usually much easier than starting a pump, and the jump from a
constant pumping rate to no pumping can be achieved fairly cleanly.
- Similarly, recovery smoothest out small changes in the pumping rate that occurred during
the pumping phase, and there is no problem with well losses from turbulent flow. This
results in more reliable estimates of aquifer properties when the recovery data are analyzed.
- The water levels in the borehole are easier to measure accurately in the absence of
turbulence caused by the pumping (especially in the early stages of the test, when water
levels are changing quickly). Some people find that it is easier to take readings quickly with a
dipper when the water level is rising than when it is falling.
- Recovery tests represent a good option for testing operational boreholes that have already
been pumping at a constant rate for extended periods. In these cases, the recovery test can
be performed when the pumps are first switched off, followed by a constant discharge test
when the pumps are switched back on again.

- Recovery-test procedure

The procedure for undertaking a recovery test is as follows:

1. Switch the pump off and start the stopwatch at the same time.
2. Measure the water level in the borehole in the same way as for the start of the pumping
test, that is, every 30 seconds for the first 10 minutes, then every minute until 30 minutes
have elapsed, then every 5 minutes until 2 hours have elapsed. After 2 hours, observe how
quickly the water level is still rising, and decide an appropriate frequency for water-level
readings until the end of the test. If the water level is rising very slowly, then a reading every
30 minutes or even every hour may be sufficient. If you miss the planned time for a water-
level reading, write down the actual time the reading was taken. Record all the readings on
the standard form (Annexes). Make sure the same datum is used for measuring water levels
as for the pumping phase.

- Analysis and interpretation

Methods of analysis for recovery tests are supposed to be used only if the pumping was at a
constant rate during the pumping phase, with the water level at or approaching equilibrium.
Recovery data following an extended constant-rate test are therefore preferable (as opposed to

57
after a step test). As before, a simple analytical method will be presented here; the reader is
referred to the reading list (Annexes) for more complex methods. The procedure for analyzing
recovery data is as follows:

1. Take all the water levels measured during the recovery phase (in meters below datum) and
convert them to residual drawdowns (s’) by subtracting the original rest-water level
measured just before the start of the pumping phase.
2. The time elapsed since the start of the recovery phase (in minutes) is denoted by t’. For the
entire residual draw -downs, calculate t, which is the time elapsed since the very start of the
pumping phase of the test. For example, if the pumping phase was 600 minutes long, for
recovery readings taken at times t’ of 1, 10 and 100 minutes, the respective times t would be
601, 610 and 700 minutes.
3. For all these pairs of times, divide t by t’.
4. Prepare a graph on semi-log graph paper, with residual drawdown s’ on the (linear) y-axis, in
meters, and t/t’ on the (logarithmic) x-axis. See Figure 3.10.
5. Plot s’ against t/t’ for the duration of the test, noting that time runs from right to left on this
graph. The data should plot roughly as a straight line. Draw a best-fit line through the data,
ignoring the early data (those on the right-hand side) and concentrating on middle to late
data. Under normal circumstances, the line should trend towards t/t’ = 1 when s’ = 0.

Figure 3.10: Recovery-test analysis

6. From this line, measure a parameter known as Δs’, which is the difference in residual
drawdowns (in meters) over one log cycle (best understood by looking at Figure 3.11).
7. Calculate the average pumping rate for the duration of the pumping phase of the test, Q, in
m3/day. This should already have been done during the analysis of the constant-rate test.
8. Insert the values of Q and Δs’ into the formula below to calculate the transmissivity T. Make
sure that the correct units have been used, in which case the units of T will be m 2/day.

T = 0.183 Q/Δs’

58
As in the analysis of data from the constant-rate test, the early data were ignored because they tend
to be affected by the volume of water stored in the borehole itself, and the points will probably not
fall on the straight line. Several types of deviation from the straight line are commonly observed, as
shown in Figure 3.7 and explained briefly below.

Figure 3.11: Deviations from straight line during recovery test (From: MacDonald et al [2005])

- Well-storage effects: The water level does not recover as quickly as it should do in theory,
because water is required to fill up the volume of the borehole itself.
- Leakage from other aquifers: The aquifer being tested is receiving water from other aquifers
or aquifer layers by vertical leakage.
- Cascading fracture: As the water level recovers, it eventually submerges a fracture from
which water was cascading (when the water level was below the fracture).
- Dewatered fracture: The rate of recovery is affected by the fact that a fracture was
dewatered during the pumping phase.
- Very low-yielding: The recovery is very slow, and likely to be dominated by the need to fill up
the volume of the borehole.

59
c. Checking verticality

After the testing equipment has been removed, a check of borehole verticality (‘plumbness’) and
alignment (‘straightness’) should be conducted. This is usually done by inserting and lowering into
the hole a perfectly straight, 12-metre-long steel rod or pipe, the external diameter of which should
be a maximum of 13 mm (about 0.5 inch) less than the inner diameter of the main or longest section
of casing (i.e. in which the pump will be housed, unless a hand-pump is to be installed).

d. Disinfection

Finally, assuming that it has passed the tests above, the borehole should be thoroughly disinfected
with a chlorine-rich solution, such as HTH (High-test Hypochlorite), leaving a concentration of
residual chlorine of 50 milligrams/litre for at least four hours. Table 8 gives the quantity or chlorine
compound to be added to 20 meters of water-filled casing for various diameters. The borehole may
then be re-sealed with the locked cap or welded plate.

Table 3.4: Quantity of chlorine compound to produce a 50 mg/l solution in 20 m of water-filled


casing

Casing diameter Volume per 20 m 65% HTH 25% Chloride of 5.25% Sodium
Inches m3 (dry weight)g lime (dry weight)*g hypochlorite (Jick)
(liquid measure)
4 0.16 12.98 37.18 0.20
6 0.36 37.18 74.10 0.39
8 0.65 55.80 129.84 0.66
10 1.01 74.10 204.59 1.11
12 1.46 111.48 297.7 1.57
16 2.59 204.59 520.66 2.49
*When powder is used, it should first be put in solution in a water container before being introduced
into the borewell

60
4. Borehole Monitoring and Deterioration
4.1 Borehole Monitoring
Continuous monitoring of borehole performance can be cost-effective, helping to detect any
problems before they become serious. Maintenance programs should consist of regular field visits,
water sampling (for chemical/microbial analyses), water level measurements, and routine
monitoring by simple step-drawdown tests. The data collected can be compared with those
obtained when the well was new or last monitored. A regular testing schedule consisting of a basic
step-drawdown test every year is sufficient, with maintenance carried out if there is any sign of
deterioration. Low-risk areas (in terms of borehole incrustation or corrosion) may require
maintenance work only every few years. It is prudent to erect a lockable fence around the borehole
to prevent tampering and accidental or malicious damage. Table 4.1 sets out the symptoms to be
noted in a monitoring program, along with causes and suggested remedial actions.

Table 4.1: Borehole monitoring: Symptoms, causes, and remedies

Monitored symptom Causes Remedial action


Regional fall of groundwater Regional factors, e.g. drought, Lower pump inlet
level large-scale abstraction,Deepen borehole
extensive deforestation Drill new (deeper) borehole
Localized fall of groundwater Over-pumping Check/compare earlier test
level Blocked screens or gravel pack pumping data
Reduce pumping rate
Rehabilitate: Inspect screens,
surge-develop to clean
screens and gravel pack
Change in water quality Chemical pollution Analyze water; if hazardous,
(chemical) Saline influx shut down borehole
Aquifer mixing production and reassess
situation
Change in water quality Pollution Analyze water. If hazardous,
(biological) Change in water chemistry shut down borehole
production. If temporary,
pump out water and disinfect
borehole
Unusual Water quality, e.g. carbonate Remove pump, inspect
corrosion/incrustation of (hard water), acidic water, iron borehole. Rehabilitate
borehole head works bacteria
equipment
Reduction of yield (pumping Pump faulty Remove and inspect pump
level unchanged) Piping blocked (incrustation) Inspect piping; replace if
necessary
Unusual noise or vibration Damaged/faulty pump Remove and inspect pump
(submersible pump) Inspect borehole

61
Local staff should be recruited and trained in the monitoring of boreholes, and in the repair of
pumps (especially hand-pumps), particularly in those areas where the failure of the water supply
would have the most serious consequences. This would apply most to boreholes supplying
settlements or institutional facilities in remote arid or semi-arid regions of the world and in places
where rapid borehole deterioration is a possibility.

Water quality monitoring

Chemical analysis of the bore water should indicate the potential for damage to borehole structures.
The physical condition of the abstraction system at a borehole may give an indication of developing
conditions within the borehole itself. If unusual and significant corrosion or incrustation is taking
place among borehole headwork structures, the same is likely to be happening inside the borehole.
Water quality monitoring is particularly important if boreholes are close to densely populated areas
or in coastal zones.

Pollution (chemical or biological) may be caused by the former; in coastal areas there is the
possibility of intrusion by salt water, from a fluctuating fresh water/sea water transition zone. In the
latter case, of course, simple tasting will confirm the problem, but regular conductivity or total
dissolved solids (TDS) analysis will provide predictive data.

Distilled water has a conductivity of 1 µS,

good fresh water <2,000 to 3,000 µS,


and saline water >6,000 µS
(S = Siemen, 1 Ω-1cm-1).

The equivalent TDS classification is:

fresh water 0 to 1,000 mg/l;


brackish water 1,000 to 10,000 mg/l;
and saline water 10,000 to >100,000 mg/l.

Borehole monitoring should include regular step-drawdown tests, which can be further analyzed to
determine the basic hydraulic parameters of aquifers. Drawdown in a borehole is essentially the sum
of losses due to movement of water from the aquifer into the borehole space. Mathematical analysis
of step-test data allows these losses to be determined, along with the relationship between
drawdown and discharge for the borehole under test. From these data, an indication of the
efficiency of the borehole (and hence of any reduction of efficiency over time) can be obtained.

4.2 Borehole Deterioration


The life expectancy of a production borehole will be limited if it was incorrectly designed or not
constructed for maximum efficiency, or if it has been over-pumped. Many production wells are
seldom monitored or maintained; they are neglected until a problem arises. But if a borehole is
properly designed, constructed with the correct materials, and given regular attention, it can
produce water for 50 years or more. Common causes of borehole deterioration or failure include the
following:

62
a. Water level drawdown

Production from a borehole or a well field can decline because of a drop in the water table, which
might be due to natural causes such as drought, but also to well deterioration and over-pumping
(excessive drawdown). A drop in the water level can result in submersible pumps shutting off
automatically.

b. Mechanical failure

Pumps eventually lose their effectiveness as parts become worn, corroded, or clogged, and borehole
screens become partly blocked by damaging organic and inorganic accumulations and scale deposits.
If pumps are not turned off before they begin sucking in air, they will be irreparably damaged.
Decline in or loss of production can be at least partly (if not mostly) remedied by a program of well
maintenance and rehabilitation.

c. Incrustation

Most ground waters are only mildly corrosive, if at all, so corrosion is not usually a problem if good
quality plastic and steel (such as stainless steel) casings and screens have been installed. The main
cause of deterioration is the build-up of incrustations around screen openings, which reduce
borehole efficiency.

As a borehole is pumped, pressure is reduced by the local drawdown, and water velocity and
turbulence around the borehole increase. In this agitated zone, carbon dioxide gas is released from
the water, which reduces the solubility of certain compounds in the water, such as calcium
carbonate. Incrustation is mainly the result of the precipitation of insoluble carbonates,
bicarbonates, hydroxides, or sulphates of calcium, magnesium, sodium, manganese, or iron.
However, these deposits are rarely composed of a single mineral.

Normally, the level of dissolved iron in groundwater is low, but slight changes in water chemistry,
such as acidification due to dissolved carbon dioxide or organic matter (humic acids) can result in
higher iron concentrations (up to tens of milligrams per litre). Iron will remain in its soluble (ferrous)
state unless there is a rise in the pH (alkalinity, equivalent to reduction of acidity) or Eh (redox
potential) of the water. Increased oxygenation of the turbulent zone can initiate iron precipitation by
oxidation from the ferrous (soluble Fe2+) to the ferric (insoluble Fe3+) form in the screen area. Serious
mineral deposition can occur at the top of screens, which become exposed to air owing to excessive
drawdown. Inorganic silts and clays often add to the problem, but organic deposits can also be
involved. Oxidation of ferrous to ferric iron at the borehole boundary can encourage the growth of
86 certain bacteria. Organic slime formation by species of iron bacteria is a result of the life cycle of
such organisms. They inhabit groundwater by metabolizing ammonia, methane, or carbon dioxide,
again changing iron into deposits of insoluble salts (mainly hydroxide), which worsens incrustation.

Iron biofouling is a complex process influenced by interactions between the aquifer environment
and the borehole structures. Microbial matter consists of filamentous cell colonies, mats, and slime
sheaths (which cells secrete for protection), often of a sludgy consistency, but able to harden with
age. Such incrustations impair hydraulic efficiency and specific capacity, clogging pipes, filter packs,
screens, and pumps. They can cement a gravel pack into something akin to concrete. They
encourage corrosion and reduce water quality, but remedial measures are likely to be less effective

63
once hardening has occurred. If an incrustation has aged and recrystallized it will be extremely
difficult to loosen and remove.

d. Corrosion

The most common corrosion process is electrochemical, in which iron (or another metal) is dissolved
and re-precipitated as a hydroxide deposit.

Corrosion in a water well most often occurs at localized physical imperfections on metal pipes and
screens; the process can be encouraged by high salinity, high temperatures, oxygen, carbon dioxide,
hydrogen sulphide, and organic acids (from peat or pollution).

Corrosion can perforate metal screens and casings, weakening the structure and allowing pollutants
(or even gravel pack material) to enter the borehole. As has been mentioned before, incrustation or
corrosion can be slowed down by installing screens with the greatest possible slot area, to reduce
pumping rates and inlet velocities, and by periodic cleaning or redevelopment of the borehole.

64
5. Borehole Rehabilitation
Rehabilitation is the action taken to repair a borehole whose productivity has declined or that has
failed through lack of monitoring and maintenance of the pump and/or well structure. This is often a
financial problem, or a logistical one – a function of remote location and, possibly, of conflict
preventing easy access. Surface pumps, such as wind-pumps or hand-pumps, often fail for purely
mechanical reasons – broken rods or corroded risers, for instance – and disused boreholes silt up or
have objects dropped into them. Unfortunately, if a borehole has become tightly blocked by hard
debris, such as stones and pieces of metal (a not uncommon occurrence), it is probably totally lost.
Existing boreholes are likely to be well sited in terms of usage, since they must originally have been
drilled for a purpose. Therefore, it is almost always advantageous to rehabilitate them.

It can be reckoned as a rule of thumb that a simple rehabilitation (no casing replacement) will cost
around 10% of the price of a new borehole.

5.1 When to Rehabilitate


All pre-existing boreholes within a project area should be inspected for the possibility of
rehabilitation, unless they are on privately owned land. The extra water might not be needed, but as
boreholes provide access to groundwater, they could be used as observation holes for monitoring
local water levels. Abandoned boreholes may act as pathways for the contamination of an aquifer,
or enable the mixing of ground waters of differing quality from separate aquifers. They might also
present a physical hazard to, say, local children, especially if they are of large diameter and open.
Redundant boreholes are potentially useful as groundwater monitoring points, even if they cannot
be rehabilitated for production; but holes that are beyond repair should be backfilled using clean,
inert, non-polluting materials such as gravel, sand, shingle, concrete, bentonite, rock, or cement
grout.

A borehole that has stood unprotected – by a top casing cap or a surface installation – for some time
will almost certainly have been lost because of, say, objects being dropped into it by children. If a
blockage can be reached from the surface it should be probed with a strong metal bar to get an idea
of its solidity. Loose fine material might be removable using compressed air (see below); if this can
be done, full rehabilitation might be a possibility. If the borehole was protected by a cover and is
apparently clear, it should be checked for depth by plumb-line dipping, for static water level by dip
meter, and for method of construction and internal condition by means of down hole camera.

Before carrying out rehabilitation, it is advisable to sample and analyses the local groundwater (if
possible) to ensure that it is not unduly chemically aggressive.

5.2 Rehabilitation Methods


The basic rehabilitation process should consist of the following principal stages in this order:

65
1. Collection of archives and information (from water authorities, drilling companies, aid
organizations, etc.) on the borehole design
2. Inspection by down hole camera
3. Breaking-up of clogging deposits and incrustations
4. Removal of silt and debris by surging and airlift clearance pumping
5. Borehole disinfection
6. Step-drawdown test

I. Inspection by down-hole camera

Prior to commissioning camera inspection, efforts should be made to locate borehole design and
construction details as that may save a lot of time. However, in Somalia archives of borehole design
might be hard to find.

Typically, rehabilitation might consist of an initial camera run before de-silting by conventional air
surging. A second camera inspection should then be carried out to check the efficacy of the de-silting
operation and to obtain a clearer picture of down-hole conditions. All camera runs should be logged
in detail and videotapes of the inspection retained for future reference.

A survey video enables full inspection of the inside of a borehole to be carried out, from top to
bottom, in ‘real time.’ Side views allow casing or screen condition to be observed at accurate,
recorded depths. With information of this quality, problems can be identified and complete
rehabilitation of a borehole planned. Construction details can be observed directly and compared
with the original log, if one is available. Objects or debris dropped into a hole can be inspected and
the possibility of removal assessed. Water cascades, and to a certain extent, water quality (chemical
precipitates, turbidity), can be viewed on a television monitor.

II. Breaking up of clogging deposits and incrustations

It is usually difficult – if not impossible – to remove old casings or screens to clean or replace them,
so other methods often need to be used. Screens can be cleaned using a rotating wire brush or
scratcher, but they may have been weakened by corrosion, so care should be taken not to worsen
their condition. Borehole restoration methods are similar to those used in development, except that
incrustations have to be broken up and removed.

a. Water jetting

If it is done systematically, water jetting at high pressures can be a particularly effective means of
de-clogging and cleaning the internal surfaces of boreholes. A jetting nozzle on the end of a length of
high-pressure air hose or pipe is required. Test trials have shown that nozzle exit pressures of 17,000
kPa (for a 1.5 to 2" nozzle, positioned about 1" from the screen) will be effective on most occasions.
In unlined boreholes, the jetting pressure limit is around 40,000 kPa. To avoid damage to plastic
screens, pressures greater than 20,000 kPA should be avoided, because very high pressure jetting
(greater than 30,000 kPa) can cut through plastic casing. Steel casing can withstand pressures of up
to at least 55,000 kPa, and the screens that best respond to jetting treatment are those with high
open areas and continuous slots, such as wire-wrap types like Johnson screens.

b. Acidization

66
For seriously affected boreholes, a combination of physical and chemical methods might be most
effective. Acidization can remove carbonate incrustations and ferric hydroxide deposits in their early
non-cemented stage. Hardened iron deposits would require physical breaking up by the methods
described above. A 30% sulphamic acid solution to the volume of the screened or open section to be
cleaned can be used for 15 to 24 hours, with the water in the borehole being periodically agitated by
air that is blown in.

c. Hydrofracturing

Old boreholes drilled into low-yield formations, such as Precambrian crystalline rocks, can be
stimulated by a process known as hydrofracturing. The technique can be applied only to open,
uncased sections such as might occur towards the bottom of a hole. First, inspection by down-hole
camera or down-hole geophysical log must be run to assess the suitability of the borehole to such
treatment. The section to be worked should already be fractured to a certain extent, and must be
isolated using some kind of packer. This might consist of a series of rubber seals that can be
expanded in the borehole by a hydraulic ram or by compressed air from the surface.

An injection pipe runs down the center of the packing system. High-pressure water is injected into
the borehole in order to create or enlarge the fractures. Sand can be added to the water to keep
open (‘prop’) newly developed fractures. Reports indicate that yield increases of 20 to 80% have
been achieved using hydrofracturing. Depending upon the nature of the formation, injection
pressures of 35 (soft) to 140 (hard) bar are used. After treatment, water and debris are air-lifted out
in the normal way.

III. Relining

A borehole seriously affected by corrosion – that is now pumping out sediments – can be restored
only by partial or total relining. The necessary course of action may be decided only after a borehole
camera survey, or a geophysical logging, has determined the extent of the damage or deterioration.
Down-hole logs might contain indications (water temperature, conductivity, flow, resistivity, or
casing collar logs) of holes in casing.

Any new casings or screens that are installed should be of corrosion-resistant materials to avoid a
repetition of the original problem. A new lining will be of smaller diameter, so the new pump will
have to be chosen with this in mind.

Corroded screens should not be relined if at all possible, because concentric screens create
turbulence and abrasion, and fragments of corroded metal could be sucked into the borehole during
pumping.

Although it can be extremely difficult, corroded screens should be removed and replaced by new
corrosion-resistant materials. Any attempt to do this would involve bringing a large drilling rig on site
and using its pulling power to remove the old casing string. Any lost gravel pack material can be
blown out. With new casings installed, the borehole can be developed in the usual way.

New casings and screens can be protected from corrosion under water by means of sacrificial
electrodes (cathodic protection). Sacrificial anodes of a metal that is higher in the electromotive
series (relative tendency to oxidation) than steel – such as magnesium and zinc – are attached and

67
corrode in preference to the protected metal of the casing. Such systems are used to protect ships,
underwater pipelines, and pump installations, but are rarely applied to borehole casings or screens.

Techniques for the removal of clogging deposits and incrustations include high pressure surging,
jetting, air-lift pumping, air-bursting, and chemically assisted dispersion. Periodic rehabilitation,
which ought to be carried out on a regular basis, can remove deposits before they harden with age.

In addition to air surging, a drilling rig can be used to redesign an uncased borehole or one from
which linings have been pulled out. A borehole can be reamed or deepened to intersect more of the
aquifer or to provide greater available drawdown.

Shallow dug wells can also be rehabilitated using a drilling rig. When the water level drops below the
bottom of a well, the well runs dry. If the structural integrity of the well (its sidewall and surface
structure) is sound, a rig can be brought in to drill a borehole through the base of the well and
further into the shallow aquifer (or even deeper, but it may be better to drill a completely new
borehole if this is desired). The borehole can, if necessary, be cased and screened by the methods
described above, and a hand-pump mounted on the dug well slab with risers extending down into
the borehole.

IV. Borehole sterilization

Boreholes affected by iron incrustation should be sterilized by chlorination between the clean-out
pumping and the step test, to destroy ubiquitous iron bacteria to and delay re-infection of the well.
Granular HTH can be dissolved and added so as to leave about 50 milligrams per litre of residual free
chlorine in the borehole water. Mixing the solution in the borehole can be done by blowing with the
airline used for the air-lift pumping. The concentration should be monitored using a water testing
kit. The borehole can then be pump tested

V. Step-drawdown testing

A step-drawdown test will indicate whether rehabilitation has been successful; it can also serve as a
new baseline against which future well performance can be measured. Test pumping of a
rehabilitated borehole will also help to re-establish normal groundwater flow and remove remaining
silt particles.

VI. Mechanical repair

Many boreholes lie disused because pumps have broken down or because of the lack of necessary
expertise or spares. In the case of hand-pumps, this can be relatively easy to fix: all that is needed is
a set of standard tools with which to remove the pump handle, chain, riser pipes, rods, and piston,
for inspection and repair or for replacement.

Pumps and risers on deeper boreholes might require a tripod and a vehicle with a winch for removal.
If at all possible, a borehole should be inspected by video camera once a pump has been removed
(see above).

68
Annexes

69
Annex 1: Formation type in Somalia:
Unconsolidated:
Water quality
Named Aquifers General Description Recharge
issues
Terrace deposits in major wadis (ephemeral
river beds - called toggas). Younger
Holocene/Recent deposits often overlie and
are in hydraulic continuity with older
Pleistocene deposits, which can result in very
thick aquifers of over 100 m.
Typically high productivity aquifers, with
medium to high permeability and high
infiltration capacity. Estimated transmissivity
values are commonly in the range 10-2 to 10-
3
m²/sec. In the Geed Deeble area (source for Generally low levels
the Hargeysa water supply), only one in ten of mineralisation,
tested boreholes showed a transmissivity of with TDS below
less than 10-3 m²/sec; the others ranged from 1000 mg/l, and of
2.86 to 5.18 x 10-3m²/sec. Calculated moderate to good
equivalent hydraulic conductivities were in drinking water
the range 1.4 x 10-4 m/sec to 7.7 x 10-5 m/sec. quality. Water from
Test yields of the production boreholes shallow dug wells
ranged from 12 to 20 l/s, with drawdowns and some springs
Alluvial terrace
typically less than 20 m (data provided by often has a High
deposits -
Hargeysa Water Utility). conductivity in the infiltration
Pleistocene to
Generally unconfined, but where covered or range 2000 to 4000 capacity
Holocene/Recent
associated with Quaternary volcanic basalts, microS/cm, but
they can be confined, sometimes with other samples of
considerable artesian pressure (e.g. in the shallow
Xunboweyle area). In unconfined aquifers the groundwater in the
water table is typically 2 to 3 m deep western part of
throughout the year, related to seasonal northern Somalia
flows along riverbeds. In deeper confined, have conductivity
artesian aquifers in older deposits, the values of less than
piezometric head does not fluctuate much 1500 microS/cm.
throughout the year.
Thickness varies from a few meters to over
100 m. At Geed Deeble (source for the
Hargeysa water supply), the tapped aquifer
depth is over 150 m. Boreholes are typically
between 10 m and 50 m deep.
In Somaliland in the north of Somalia,
dynamic (sustainable) groundwater reserves
in the major alluvial aquifers are estimated at
an average flow of ~30 m³/sec.
Alluvial Direct rainfall
sediments filling recharge,
Low to high productivity, depending on local
major valleys and and indirect
lithology, thickness and lateral extent.
plateaus - recharge
Pleistocene to from

70
Holocene/Recent infiltration of
river water

Key references for this aquifers are Faillace and Faillace 1986, FAO/SWALIM 2012, Water Supply
Survey Team of the PRC 1983, Petrucci, 2008 and German Agro-Action, 2005.
Volcanic
Water quality
Named Aquifers General Description Recharge
issues
These are a potential aquifer in some areas.
They contain groundwater only where In some
fractured and/or weathered, or in lenses of areas,
pyroclastic material between lava flows. They vertical
typically have low to moderate permeability, fractures
but are locally highly fractured, increasing resulting
Pleistocene permeability. However, they occur primarily from cooling
basaltic lava as elevated plateaus, and are often of the basalts
flows unsaturated. In some areas, such as Agabar may occur,
and Las Dhure, they are found in the lowlands and are likely
and may be saturated, and in this case are to form
likely to be unconfined. Boreholes drilled in primary
these areas have intersected water-bearing recharge
zones composed of sand/pyroclastic lenses routes.
and weathered basalt.

Key references for this aquifer are Faillace and Faillace 1986, FAO/SWALIM 2012 and German Agro-
Action 2005.
Sedimentary - Intergranular and Fracture Flow
Water quality
Named Aquifers General Description Recharge
issues
The Yessoma Formation is of Nubian
Recharge is
sandstone type and can form a high
estimated to
Upper productivity aquifer. The coarsest grained Groundwater of
be
Cretaceous part of the formation occurs between 140 m good quality is
approximately
Yessoma and 180 m depth. Calculated aquifer generally supplied
in the range
Formation transmissivity is around 2 x 10-3 m²/sec (220 by dug wells in the
of 3 to 5% of
(Nubian m²/day), with an average specific capacity of weathered part of
annual rainfall
sandstone) 7.5 m³/hour/m. Most boreholes penetrating the aquifer.
(Van der Plac
the formation can sustain a yield of more
2001).
than 30 m³/hour.
Jurassic sedimentary rocks in the south of
Somalia are likely to be dominated by
sandstone. Their groundwater potential is not
Jurassic
well known. Groundwater storage and flow
sandstones
may be by both intergranular and fracture
flow. Low to moderate yields may be
possible.

71
Sedimentary - Fracture Flow
Water quality
Named Aquifers General Description Recharge
issues
These form moderate productivity aquifers.
Fractures act as pathways for rapid
groundwater flow, but permeability and
Tertiary: Recharge is
groundwater storage are small.
Iskushuban estimated to
A borehole drilled into the Miocene
Formation be
Iskushuban Formation in Timirishe in the Bari
(Miocene); approximatel
area yielded 5 l/s for a drawdown of some 50
Mudug y in the range
m, with a calculated transmissivity of 4.5 x 10-
Formation 4 of 3 to 5% of
m²/sec.
(Oligocene/Mioc annual
Boreholes in the Oligocene/Miocene Mudug
ene); Daban rainfall (Van
Formation are drilled to 180 to 220 m deep,
Formation der Plac
and provide yields of 3 to 5 l/s for drawdowns
(Oligocene) 2001).
in the range 3 to 24 m. Transmissivity values
of 3.1 x 10-3 to 2.9 x 10-4 m²/sec were
calculated.
Cretaceous
undifferentiated:
sandstones, Little is known about the aquifer properties of
conglomerates, these rocks.
limestones and
evaporitic rocks

Key references for these aquifers are: Faillace and Faillace 1986, FAO/SWALIM 2012, Petrucci 2008,
German Agro-Action 2005, GKW 1977 and Van der Plac 2001.
Sedimentary - Karstic
Water quality
Named Aquifers General Description Recharge
issues
The Eocene limestone (Karkar and Auradu) Groundwater in the
and limestone/evaporite (Taalex) formations Karkar karst aquifer
are often karstic, and are among the most is slightly
significant aquifers in the north of Somalia, in mineralised, with
the Somaliland and Puntland regions. an SEC
The Karkar limestone represents the most (conductivity) value
promising fresh groundwater resource for typically between
further development in the Sool and Hawd 1500 and 1800
Eocene Karkar,
plateaus in the north of Somalia. It typically micromhos/cm.
Taalex and
forms a moderately productive aquifer. The Taalex aquifer
Auradu
The Auradu limestones can form a high usually yields
limestones
productivity aquifer, with good quality moderately to
groundwater, although more investigation is highly mineralised
needed. If groundwater is present, the groundwater,
overlying Taalex aquifer should be sealed off derived from
to prevent inflow of lower quality water. geogenic evaporitic
Many boreholes abstract from the aquifer, minerals. Ca or
particularly in the Puntland region, with an CaSO4 type
average transmissivity of 10-3m²/sec (860 groundwater is

72
m²/day). Other boreholes over 200 m deep dominant, with TDS
are drilled in limestones in the Garoowe area. usually greater than
Where these limestones are overlain by the 3800 mg/l. SEC
Karkar formation, they are often semi- (conductivity) levels
confined, with low sub-artesian pressure. are generally very
The depth to water table in unconfined parts high, from 890 to
of these aquifers is usually between 5 and 15 7270 microS/cm.
m throughout the year. Sulphate
Fresh groundwater reserves in the Auradu concentrations are
aquifer in the Somaliland and Puntland in the range 125
regions are estimated as equivalent to an mg/l up to 3100
average flow of 63.4 m³/sec. The estimated mg/l, with an
fresh groundwater reserve in the Karkar average of 1300
aquifer is lower at approximately 10 m³/sec. mg/l. Many
boreholes have
been abandoned
because of a high
salinity content.
Groundwater from
the Auradu
limestones is
typically of
sulphate-
bicarbonate type
with moderate to
high mineralisation,
and an SEC
(conductivity) value
generally lower
than 1000
micromhos/cm.
Sulphate is the
dominant element
in almost all
samples with a
range from 3 to 220
mg/l.
The Jurassic limestones in the north of the
Approximate
country have the greatest potential for
Groundwater in the estimates of
groundwater development in the country.
Jurassic aquifer is recharge are
There is usually pure limestone in the upper
generally of between 35%
part of the formation, with marly levels and
bicarbonate type of annual
calcareous sandstones in the lower part. The
with low levels of rainfall for
Jurassic upper parts in particular are usually
mineralisation, with the Karkar
limestones characterised by a high degree of fracturing
SEC (conductivity) aquifer to
and probably karstic cavities, and
commonly in the 50% of
groundwater circulation probably develops
range 600 rainfall for
mainly in this zone. The limestones can be
microS/cm to 1200 the Jurassic
highly permeable, with a transmissivity value
microS/cm. limestone
from one test borehole at Borama of 3.1 x 10-
3 aquifer.
m²/sec (270 m²/day).

73
The depth to water table in unconfined parts
of the aquifer is usually between 5 and 15 m
throughout the year.
Groundwater reserves in the Jurassic
limestone aquifer in the Awadal region are
estimated as equivalent to an average flow fo
18.9 m³/sec.

Key references for these aquifers are: Faillace and Faillace 1986, FAO/SWALIM 2012, Petrucci 2008,
German Agro-Action 2005, GKW 1977 and Van der Plac 2001.
Basement
Water quality
Named Aquifers General Description Recharge
issues
Groundwater has
low to moderate
mineralisation, with
conductivity often
between 300
mS/cm and 1400
mS/cm, up to a
maximum of 3570
Forms a low productivity aquifer or an
mS/cm in some
aquitard, depending on the development of
shallow wells. More
permeability by weathering/fracturing.
than 70% of
analyzed waters
have good
characteristics
according to WHO
standards for
drinking water in
arid regions.

74
Annex 2 a. Examples of borehole construction designs (not to scale)

75
Annex 2 b. Examples of borehole construction designs (not to scale)

76
Annex 2 c. Sample Well Design (un-consolidated formation)

77
Annex 2 d. Sample Well Design (semi-consolidated formation with risk of collapse)

78
Annex 2 e. Sample Well Design (consolidated formation – casing and screen in bedrock)

79
Annex 2 f. Sample Well Design (consolidated formation – open hole)

80
Annex 3: Suggested Format for Borehole Completion Record
1. General
2. Drilling Operation
3. Casing and Well Completion
4. Well Development and Pumping Test Summary
5. Water Quality Summary
6. Lithology
6a. Lithological Logging
6b. Characteristics to be evaluated and assessed during logging of drilling
samples
7. Pumping Test Details
7a. Step Drawdown Test
7b. Constant Rate Test
7c.Recovery Test
8. Water Quality Analysis Parameters

81
82
83
84
85
86
87
88
89
90
91
92
93
94
Annex 4:

- Top Free Satellite Imagery Sources

1. USGS Earth Explorer: https://earthexplorer.usgs.gov/


2. Landviewer: https://eos.com/landviewer/
3. Copernicus Open Access Hub: https://scihub.copernicus.eu/
4. Sentinel Hub: https://apps.sentinel-hub.com/eo-browser/
5. NASA Earthdata Search: https://search.earthdata.nasa.gov/
6. Remote Pixel: https://search.remotepixel.ca/
7. INPE Image Catalog: http://www.dgi.inpe.br/catalogo/

- Top Open source Software’s for Hydrogeologist

1. Geographical Information Systems:


- QGIS: https://www.qgis.org/en/site/
- SAGA GIS: http://www.saga-gis.org/en/index.html
2. River modeling
- HEC-RAS: http://www.hec.usace.army.mil/software/hec-ras/
- iRIC: https://i-ric.org/en/
3. Hydrologic modeling
- HEC-HMS: www.hec.usace.army.mil/software/hec-hms
- PRMS: https://www.usgs.gov/software/precipitation-runoff-modeling-system-prms
- SWAT: https://swat.tamu.edu/
4. Hydrogeological modeling
- MODFLOW: http://water.usgs.gov/ogw/modflow/
- MT3DMS: http://hydro.geo.ua.edu/mt3d/
- OpenFOAM: www.openfoam.org
5. Subsurface Data Management and Reporting
- gINT: https://www.datgel.com/gint-logs
- GAEA: http://www.gaeatech.com/proddetail.php?prod=3011

6. Environmental Data Collection


- ESdat pLog: http://esdat.net/pLog.aspx
- EnviroInsite: https://earthsoft.com/enviroinsite/

95

You might also like