You are on page 1of 44

Source: HYDRAULIC DESIGN HANDBOOK

CHAPTER 24

ARTIFICIAL RECHARGE
OF GROUNDWATER:
SYSTEMS, DESIGN,
AND MANAGEMENT
Herman Bouwer
USDA-ARS Water Conservation Laboratory
Phoenix, Arizona.

24.1 ABSTRACT
The type of system to be selected for artificial recharge of groundwater and how it should
be designed and managed for optimum performance depend entirely on local conditions
of soil, hydrogeology, topography, water availability (quality, continuous, or interrupted
supply), and climate. Unlike water or wastewater treatment plants, guidelines, standards,
and blueprints for artificial recharge systems cannot be given. Rather, the responsible
planner should have extensive knowledge of the various types of recharge systems that can
be used, how they work, the processes, cause and effect relations, what experiences have
been obtained elsewhere, and how they should be adapted to local conditions. Once the
type system has been selected, a protocol of site investigations should be followed for preliminary design. Key factors in the design and management of successful artificial
recharge systems are site and system selection, maintenance of adequate infiltration rates,
hydraulic continuity between the recharge system and the aquifer (no clay layers in the
vadose zone), and groundwater control for effective water recovery and prevention of
undue groundwater rises in the recharge area. Another important factor in the selection of
the type of recharge system is the pretreatment of the water required before recharge to
minimize physical, chemical, or biological clogging of the infiltrating surface (bottom in
basins and walls in trenches, shafts, and wells). Since clogging layers are easier to control and remove in surface infiltration systems, proper pretreatment is especially important for trenches, shafts, and wells. Since trenches and shafts are relatively inexpensive,
they can be replaced when their useful life has come to an end. On the other hand,
recharge wells, while much more expensive than trenches and shafts, are more amenable
to clogging prevention and control by frequent pumping and periodic redevelopment. All
these aspects have to be factored in, however, when selecting the type system that gives
recharge at minimum cost per unit volume of water put into the aquifer. While presedimentation in settling ponds may be adequate pretreatment for surface infiltration systems,
recharge with trenches, shafts, and wells may require coagulation and sand filtration to
remove suspended solids to acceptable levels. Where sewage effluent is used for groundwater recharge, conventional primary and secondary treatment may be sufficient for surface infiltration systems. However, recharge with effluent through wells may require
24.1
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

24.2

Chapter Twenty-Four

advanced treatment processes including nitrogen and phosphorus removal, sand filtration
for removal of suspended solids and parasitic organisms like helminths, Giardia, and
Cryptosporidium, disinfection for removal of viruses and bacteria, and reverse osmosis
(RO) for removal of organic compounds and other chemicals.
If the site investigations indicate no fatal flaws, a small pilot or test project should be
installed to gain a better understanding of how the system behaves and how the full-scale
system should be designed and managed for best performance. Even then, a phased construction should be used for the full-scale system, so that operating experience is obtained
and refinements can be made as the rest of the system is designed and constructed. This
is especially true for new areas where artificial recharge has not yet been practiced.
Where local recharge experience already is available, additional systems then are essentially expansions of existing systems and the normal protocol of preinvestigations and
pilot testing can be reduced. The objective of this chapter is to provide information on the
types of recharge systems, how they work, and how they should be designed and managed,
so that planners, designers, and managers can make best available decisions.

24.2 INTRODUCTION
Artificial recharge of groundwater is accomplished by applying water to the soil for infiltration and downward movement through the unsaturated or vadose zone to the groundwater. Such surface systems (Fig. 24.1A) require soils that are sufficiently permeable,
vadose zones that have no clay or other restricting layers, and aquifers that are unconfined.
Where permeable surface soils are not available but sufficiently permeable material is
found at relatively small depth (about 1 m, for example), artificial recharge can be
achieved with excavated basins that are sufficiently deep to reach the permeable material
(Fig. 24.1B). If the permeable material is too deep for removal of overlying material, but
is within trenchable depth (less than about 7 m, for example) seepage trenches can be used
(Fig. 24.1C). Such trenches are also suitable where soils are highly stratified with alternating layers of fine and coarse materials. Where permeable material is too deep for
trenches, large-diameter wells, pits, or shafts in the vadose zone can be used (Fig. 24.1D).
Such shafts can be drilled with bucket augers to a depth of about 50 m with a diameter of
about 1 m. Both trenches and shafts are backfilled with coarse sand or fine gravel to prevent caving. Where permeable surface soils are not available, vadose zones are not sufficiently permeable to transmit water to the underlying aquifer, or aquifers are confined,
artificial recharge can be achieved by applying water to recharge wells penetrating the
aquifer (Fig. 24.1E).

24.3 SYSTEMS
The first step in planning and designing an artificial recharge system is selecting the type
of system to be used, based on soil and hydrogeologic information. Often, the choice is
obvious or determined by other factors such as high cost of land in urban areas which preclude the use of surface systems or excavated basins. Another important factor in the
selection of the type of recharge system is the pretreatment of the water required before
recharge to minimize physical, chemical, or biological clogging of the infiltrating soil surface in basins or of the walls in trenches, shafts, and wells. Since clogging layers are easier to control and remove in surface systems, proper pretreatment is especially important
for trenches, shafts, and wells. Since trenches and shafts are relatively inexpensive, they

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

FIGURE 24.1 Recharge systems for increasingly deep permeable materials: surface basin (A), excavated basin (B),trench (C), shaft or vadose zone well (D),
and aquifer well (E).

Artifical Recharge of Groundwater: Systems, Design, and Management 24.3

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

24.4

Chapter Twenty-Four

can be replaced when their useful life has come to an end. On the other hand, recharge
wells, while much more expensive than trenches and shafts, are more amenable to clogging prevention and control by frequent pumping and periodic redevelopment. All these
aspects have to be factored in, however, when selecting the type of recharge system. While
presedimentation in settling ponds may be adequate pretreatment for surface infiltration
systems, recharge with trenches, shafts, and wells may require coagulation and sand filtration to remove suspended solids to low enough levels. Where sewage effluent is used
for groundwater recharge, conventional primary and secondary treatment may be sufficient for surface infiltration systems. However, effluent recharge with wells may require
advanced treatment processes (nitrogen and phosphorus removal, sand filtration, removal
of suspended solids and parasitic organisms like helminths, Giardia, and
Cryptosporidium, disinfection for removal of viruses and bacteria, and RO for removal of
organic compounds and other chemicals).
While conjunctive use of surface and groundwater often is the main objective of artificial recharge, recharge is also used for other purposes such as creating hydraulic barriers
against seawater intrusion, stopping land subsidence, stopping spread of contaminated
groundwater (plumes) in the aquifer from point sources of pollution, raising groundwater
levels for reduction of pumping costs, protecting streams, wetlands or other ecosystems,
protecting wooden building foundations that would decay when above groundwater, etc.
Dams for surface storage of water have a number of disadvantages, such as evaporation
losses (about 2 m/year in warm, dry climates), sediment accumulation, potential of structural failure, and increased malaria, schistosomiasis, and other diseases in the local population (Devine, 1995; Knoppers and van Hulst, 1995; Pearce, 1992). New dams often are
more and more difficult to build because of poor economics and public opposition. They
interfere with the river ecology and may flood sensitive areas (cultural, religious, archeological, environmental, and recreational, scenic, and so on.). People living on the reservoir
site have to relocate and good dam sites are becoming scarcer. Dams are not sustainable
because eventually they all silt up, and they are not effective for long-term storage of water
(years or decades), which may become increasingly necessary as increases of carbon dioxide and other greenhouse gases in the atmosphere may cause global climatic changes and
increase the probability of extremes in weather such as more prolonged droughts and periods of excessive rain. This increases the need for storing excess water in wet periods to meet
water demands in dry periods. Underground storage via artificial recharge then is preferred
because of essentially zero evaporation, economics, and other factors. The water sources for
artificial recharge include water from perennial or intermittent streams that may or may not
be regulated with dams, storm runoff (including from urban areas), aqueducts or other water
conveyance facilities, irrigation districts, drinking water treatment plants, and sewage treatment plants. Of all the water in the world, 97 percent is salt water in the oceans (Bouwer,
1978). Of the remaining fresh water, two-thirds is in the form of ice in arctic and mountainous regions. Of the remaining liquid fresh water, less than 2 percent is surface water in
streams and lakes, and much of that is fed by groundwater. More than 98 percent of the
worlds liquid fresh water thus is groundwater. Not only is groundwater the dominant water
resource, but the potential for underground storage of water also vastly exceeds that for surface storage.

24.3.1

Surface Systems

Surface infiltration systems for artificial recharge of underlying unconfined aquifers can
be constructed in streambeds (in-channel systems) or outside streambeds (off-channel systems). In-channel systems (Fig. 24.2) typically consist of weirs or dams across the

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

Artifical Recharge of Groundwater: Systems, Design, and Management 24.5

FIGURE 24.2 Schematic of in-channel infiltration systems consisting of low weirs in narrow steep channel (upper left), bigger dams in wider, milder sloping channel (upper right),
and T-levees in wide, flat channels (bottom).

streambed to back the water up and spread it over the entire width of the streambed to
increase the water depth and the wetted area and, hence, the infiltration of water. The
weirs or dams are small and closely spaced where channel slopes are steep, and larger and
more widely spaced where slopes are flat. The groundwater table for these systems must
be well below the streambeds so that the streams are losing. The weirs or dams are made
of sheet metal (low weirs), earth, concrete, or inflatable rubber (fabridam). The dams
should have adequate spillway capacity or structural washout sections to handle large
flows. Such flows can clean the channel from fine sediment that may have accumulated
above the dam. Otherwise, the channel periodically must be mechanically cleaned to
remove accumulated sediment. Such sediment can be a layercake of coarse and fine
material deposited during high and low flows, respectively. This sediment greatly restricts
infiltration of water and storage of water behind the dams when flood waters recede and,
hence, must be removed. In coastal plains or other areas of mild channel slopes, dikes or
berms about 1 m high are pushed up by bulldozers to make T- or L-dikes in the streambed
that force the water into circuitous paths that cover the entire streambed (Fig. 24.2) and,
hence, increase the infiltration and groundwater recharge. Periodic high flows can destroy
the dikes but also clean the bottom from fine sediment and other clogging material. For a
system in southern California, this happens at least once a year. When the high flows have
receded, bulldozers restore the dikes in a few days.
In-channel systems are most effective if the stream is dammed higher up in the watershed; for example, in the mountains. This removes sediment from the water and also regulates the flow by storing water during high-flow events and releasing it slowly for
recharge. Storage dams themselves are, of course, ineffective for groundwater recharge
because of the sediment that accumulates on the bottom of the reservoirs.
Off-channel systems consist of basins constructed by building berms from soil that is
normally excavated from the basin areas. They can be square or rectangular, or of irregular form to adapt to existing topographies or for environmental enhancement. The latter
may include nature and wildlife benefits, scenic enhancement (landscaping, and the like),
trails for walking and bicycle riding, recreational facilities (picnic tables, playgrounds),
and so on. Each basin should have its own water supply. Deeper basins may also need a
drainage facility so that water can be drained out for drying and maintenance where dry-

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

24.6

Chapter Twenty-Four

ing by natural infiltration would take too long. Each basin is then hydraulically independent and can be operated according to its best schedule of flooding, drying, and cleaning.
The inflow facility should be carefully designed to dissipate the energy of the inflowing
water to avoid soil erosion in the basin. In finer textured soils such erosion will cause the
water to become muddy which will lead to clogging as the fine particles settle out in the
rest of the basin. Basin banks also need to be protected against erosion by rainfall and
wave action.
Occasionally, agricultural irrigation-type systems are also used for recharge, such as
borders, furrows, and terraces. On very permeable soil in irregular topographies like
kames, eskers, and other glacial deposits, low-pressure perforated pipe (hole diameter
about 5 mm) can be used to apply the water so as to avoid construction of basins and
destruction of trees or other natural vegetation.
24.3.2 Trenches, Shafts, and Wells.
Trenches (Fig. 24.3) are constructed with backhoes or other trenching equipment.
Vadose zone shafts (Fig. 24.3) are drilled with bucket augers. Trenches and shafts are
backfilled with fine gravel or coarse sand. For trenches, water is supplied with a perforated pipe on top of the backfill and the trench can be covered to avoid exposure to
sunlight and public access. For shafts, water is supplied with a vertical perforated pipe
or well screen in the center of the backfill. Recharge wells in aquifers are drilled with
conventional well drilling techniques and are constructed much like pumped wells,
including sand or gravel envelopes in unconsolidated materials. Careful grouting is
necessary where large injection pressures are used as, for example, in fractured rock
aquifers. Injection water flows into the well with one or more pipes inside the well.
These pipes should terminate some distance below the water level in the well to avoid
free falling water that can cause air entrainment and subsequent air blocking of pores
in the aquifer. For the same reason, the pipe or pipes themselves also should have full
flow. This can be achieved by using several pipes with different diameters so that the
small flows can go through the smaller pipe to maintain full-pipe flow. Alternatively,
one pipe can be used if it has an exit valve at the bottom that can be adjusted to always

FIGURE 24.3 Vadose zone recharge well with sand or gravel fill and perforated supply pipe
(left) and recharge trench with sand or gravel fill, supply pipe, and cover (right).

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

Artifical Recharge of Groundwater: Systems, Design, and Management 24.7

create full-pipe flow. Special valves are available for this purpose. Recharge wells
should be equipped with permanent, dedicated pumps to allow periodic pumping of the
well for clogging control.

24.4 INFILTRATION
24.4.1 Infiltration Rates
Infiltration rates are expressed as volume of water moving into the soil or aquifer per unit
infiltration area and per unit time. The dimension of infiltration rate thus is length/time,
for example, cm/h or in/h, or m/day or ft/day. For infiltration basins, this rate can be visualized as the rate of fall of the water surface in the basin if there is no inflow to or outflow
from the basin. Infiltration rates for surface infiltration recharge systems typically range
from 0.3 to 3 m/day (110 ft/day), with most systems in the 0.51.5 m/day (25 ft/day)
range. For systems that are operated year-round, long-term infiltration rates that include
the time that basins are periodically dried and cleaned are called hydraulic loading rates.
These rates may vary from 30 m/year to 500 m/year, depending on soil type, water quality and climate. Since evaporation rates from free water surfaces and wet soils vary from
0.5 m/year or less in cool, humid climates to 2 m/year or more in hot dry climates, water
losses by evaporation normally are negligible.
Hydraulic loading rates must be known to determine how much water can be put
underground per year for a given recharge area. If, on the other hand, the recharge system must be designed to accept a certain flow rate throughout the year, as, for example,
where the entire flow from sewage treatment plants must go underground for recharge
and soilaquifer treatment (SAT) and there is no other disposal or temporary surface storage, the land area must be based on minimum infiltration rates. These usually occur during the winter when the water temperature is low and, hence, the viscosity of the water
is high which gives proportionally lower infiltration rates (Bouwer, 1978). Also, drying
is slower when the weather is cold so that basins need to stay dry longer to achieve infiltration recovery. Drying also can be delayed by more rainfall during the drying period
which also can produce crusting effects on the soil.
Predicting and managing infiltration rates are the most important aspects in planning,
designing, and managing of recharge systems, because they determine how much land
area is needed to put a certain flow of water underground or how much groundwater
recharge can be achieved with a certain amount of available land and how much water will
be lost by evaporation. For surface infiltration systems in uniform soils without surface
clogging, infiltration rates will be about equal to the vertical hydraulic conductivity of the
soil (Bouwer, 1978), which has the following approximate orders of magnitude:
Clay soils

<0.1 m/day

Loams

0.2 m/day

Sandy loams

0.3 m/day

Loamy sands

0.5 m/day

Fine sands

1 m /day

Medium sands

5 m /day

Coarse sands
>10 m/day
These values can be used for preliminary site evaluation and system selection.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

24.8

Chapter Twenty-Four

24.4.2 Cylinder Infiltrometers


After promising soils have been identified by studying soil maps and doing on-site soil
inspections, wet infiltration tests should be performed. These are typically done with
metal, cylinder infiltrometers about one foot in diameter. However, use of such small infiltrometers can seriously overestimate the large-area infiltration rates because of lateral
flow (divergence) around the cylinder due to capillary suction in the soil (Bouwer, 1986
and 1995; Bouwer et al. 1999). Double-ring or buffered infiltrometers are not the solution because the divergence also causes overestimation of infiltration in the center portion
of the cylinder. The obvious approach then is to use larger infiltration test areas like, for
example, 10  10 ft bermed areas, where divergence effects are less significant. These
tests are laborious and they can also require large volumes of water because it may take
more than a day to reach or approach final infiltration rates. Another approach would be
to use conventional single cylinders with significant water depth to speed up the infiltration process so that tests can be completed in a relatively short time (5 h, for example).
The resulting data are then corrected for water depth, limited depth of wetting, and divergence effects to get an estimate of the long-term infiltration rate for a large inundated area.
This rate should be about equal to the hydraulic conductivity K of the wetted zone.
The infiltrometers used for this procedure are single steel cylinders 24 in in diameter
and 12 in high with beveled edge (Figure 24.4). A piece of (2 in  4 in) lumber is placed
on top of the cylinder and the cylinder is driven straight down with a sledgehammer to a
depth of about 1 in to 2 in into the ground. The soil is packed against the inside and outside of the cylinder with a piece of 1 in2 in furring strip that is held at an angle on the
soil against the cylinder and tapped with a light hammer to get good soil-cylinder contact.
If the soil contains clay, the water used for the test should be of the same chemical composition as the water used in the recharge project to avoid complications resulting from
effects of water on status of clay (coagulated or dispersed; Bouwer, 1978). A plate or pie
tin is placed on the soil for erosion prevention. The cylinder is filled to the top, and clock
time is recorded. Water is allowed to drop about 2 in to 5 in, the drop is measured with a

FIGURE 24.4 Geometry and symbols for single-ring infiltrometer.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

Artifical Recharge of Groundwater: Systems, Design, and Management 24.9

yardstick, clock time is recorded, and the cylinder is filled back to the top. This is repeated for about 6 hours or when the accumulated infiltration has reached about 20 in,
whichever comes first. The last drop yn is measured and clock time is recorded to obtain
the time increment tn for yn. A shovel is used to dig outside the cylinder to determine
the distance x of lateral wetting or divergence (Fig. 24.4). The infiltration rate in inside
the cylinder during the last water level drop is calculated as yn/tn. The corresponding
downward flow rate or flux iw in the wetted area below a cylinder or radius r is then calculated as
inr2
(24.1)
iw  

(rx)2
The depth L of the wet front at the end of the test is calculated from the accumulated
drop yt of the water level in the cylinder as
ytr2
L

n(r+x)2

(24.2)

where n is the fillable porosity of the soil. The value of n is estimated from soil texture and
initial water content. For example, n may be about 0.3 for dry uniform soils, 0.2 for moderately moist soils, and 0.1 for relatively wet soils. Well-graded soils would have lower values
of n than uniform soils. The value of L can also be determined by digging down with a shovel right after the test and seeing how deep the soil has been wetted. This works best if the
soil initially is fairly dry, there is good contrast between wet and dry soil, and there are no
rocks. Applying Darcys law (see section 4.2.1) to the downward flow in the wetted zone
then yields
K (z + L  hwe)
iw  
(24.3)
L
where z is the average depth of water in the cylinder during the last water level drop. The
term hwe is the water-entry value of the soil and it is used to estimate the suction at the wet
front as it moves downward. The water-entry value is about half the air-entry value (due
to hysteresis) and may be estimated as follows (in inches of water):
Coarse sands
Medium sands
Fine sands
Loamy sandssandy loams
Loams
Structured clays (aggregated)
Dispersed clays

2
4
6
10
14
14
40

Since K is now the only unknown in Eq. (24.3), it can be solved for K as
iw L
K  )
(z + L  hwe

(24.4)

The calculated value of K can be used as an estimate of long-term infiltration rates in


large and shallow inundated areas, without clogging of the surface and without restricting
layers deeper down. Because of entrapped air, K of the wetted zone is less than K at saturation, for example, about 0.5 of K at saturation for sandy soils, and about 0.25 of K at
saturation for finer soils. If the K-values calculated with Eq. (24.4) look promising for an
infiltration system, the next step is to put in some test basins of about 1 acre for long-term

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

24.10

Chapter Twenty-Four

flooding to evaluate clogging effects and potential for infiltration reduction by restricting
layers deeper down. Good agreement has been obtained between predicted infiltration
rates (K in Eq. 24.4) and those of larger basins. Six infiltrometers were installed in a field
west of Phoenix, Arizona. They gave an average K of 40 cm/day. Two test basins of 0.75
acre each in the same field yielded final infiltration rates of 30 and 35 cm/day (Bouwer
et al, 1997), which is a good agreement. If the infiltrometer tests give infiltration rates
that are too low for surface infiltration systems, alternate systems such as excavated
basins, recharge trenches, recharge shafts or vadose zone wells, or aquifer wells can be
considered.
The above procedure is by no means exact. However, in view of spatial variability (vertical as well as horizontal) of soil properties, exact procedures and measuring water level
drops with vernier equipped hook gages are not necessary. The main idea is to account
somehow for divergence and limited depth of wetting, rather than applying a flat reduction percentage to go from cylinder infiltration rates to long-term large area infiltration
rates. Because of spatial variability, cylinder infiltration tests should be carried out at various locations within a given site. Finally, the resulting infiltration rates should never be
expressed in more than two significant figures!
Hypothetical Example: A cylinder 24 in in diameter and 12 in high was driven 1in into
a relatively dry sandy loam soil. The following time and drop-of-water level data were
recorded:
0800

Filled with water to top

0830

Water dropped 4 in, cylinder refilled to top

0900

Water dropped 3 in, cylinder refilled to top

1000

water dropped 5 in, cylinder refilled to top

1100

water dropped 4 in, end of test

Digging with a shovel showed a lateral wetting of 12 in outside the cylinder, n was
taken as 0.2 and hwe as  10 in. The value of in is 4/1  4 in/h which when substituted
into Eq. (24.1) yields iw  4 122/ (12 + 12)2  1 in/h. The value of yt is 16 in, which
when substituted into Eq. (24.2) yields L  16122/0.2 (12 + 12)2  20 in. Since the
average water depth in the cylinder during the last measured water level drop was 11 
4/2  9 in, K is calculated with Eq. (24.4) as 1  20/(9  20  10)  0.5 in/h. Thus, the
last measured cylinder infiltration rate of 4 in/h is reduced to 0.5 in/h to eliminate effects
of divergence, limited depth of wetting, and water depth in the cylinder to produce a value
that can be used to predict potential infiltration rates for extended inundation of large
areas.

24.4.3 Clogging
The main enemy of infiltration systems for artificial recharge is clogging of the infiltrating surface (bottoms of basins or other surface infiltration systems, and walls of trenches,
shafts, and wells). Clogging is caused by inorganic (clay, silt) and organic (algae, sludge)
suspended solids in the water that accumulate on the infiltrating surface, and by microorganisms that grow on the soil particles (biofilms) and produce polysaccharides and other
insoluble metabolites to form a soil-clogging biomass. Bacteria also can produce gases
(nitrogen, methane, carbon dioxide) that can block soil pores. Gas also can be formed in
aquifers when water from recharge wells contains entrained air or is cooler than the

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

Artifical Recharge of Groundwater: Systems, Design, and Management 24.11

groundwater itself. As the recharge water then warms up in the aquifer, dissolved air may
go out of solution to form pore blocking air pockets in the aquifer (called air binding).
Also, well recharge can lead to precipitation of iron and manganese oxides or hydroxides
as dissolved oxygen levels change, and to solution or precipitation of calcium carbonates
due to changes in pH and carbon dioxide levels.
Since clogging layers are much less permeable than the natural soil material, they
reduce infiltration rates and become the controlling factor or bottleneck in the infiltration process (Fig. 24.5). When the infiltration rate in surface systems has become
less than the hydraulic conductivity of the soil below the clogging layer, this soil
becomes unsaturated to a water content whereby the corresponding unsaturated
hydraulic conductivity becomes numerically equal to the infiltration rate (Bouwer,
1982). The resulting unsaturated downward flow then is entirely due to gravity with a
hydraulic gradient of 1. The thickness of clogging layers may range from 1 mm or less
(biofilms, thin clay, and silt layers or blankets) to several cm or more for thick sediment deposits.
Clogging is best controlled by prevention; that is, by removing the parameters that
cause clogging. For surface water, this typically means presedimentation to settle clay,
silt, and other suspended solids. This can be accomplished by dams in the river or aqueduct system (which would also regulate the flow) or by passing the water through dedicated presedimentation basins prior to recharge. Coagulants like alum and organic polymers can be used to accelerate settling. For well recharge, filtration may also be necessary.
Algae growth and other biological clogging in basins can be reduced by removing nutrients (nitrogen and phosphorus) from the water. This is also important where trenches,
shafts, or wells are used for recharge with sewage effluent or effluent contaminated water.
Organic carbon levels then must also be reduced, using activated carbon filtration and/or
possibly reverse osmosis or other membrane filtration. Disinfection with chlorine or other
disinfectants with residual effects reduces biological activity on and near the walls of the
trenches, shafts, or wells and, hence, clogging. Clogging rates increase with increasing
infiltration rates, because of the increased loading rates of suspended solids, nutrients, and
organic carbon on the surface. Because of this, increasing the injection pressures in

FIGURE 24.5 Schematic of infiltration basin with clogging layer, unsaturated flow to aquifer, and capillary fringe above water table.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

24.12

Chapter Twenty-Four

recharge wells that show signs of clogging can actually hasten the clogging process.
Regular pumping of recharge wells and periodic redevelopment of the wells can control
and delay clogging, but probably not forever. Increasing the water depth in recharge
basins or the injection pressure in recharge wells can compress the clogging layer which
reduces its permeability and, hence, the infiltration rates (see Sec. 24.4.4).
For surface infiltration systems, clogging is controlled by periodically drying the
basins or other infiltration facility, and letting the clogging layer dry, decompose, shrink,
crack, and curl up. This may be sufficient to restore infiltration rates to satisfactory values. If clogging materials continue to accumulate, they must be periodically removed at
the end of a drying period. This can be done mechanically with scrapers, front-end loaders, graders, or manually with rakes. After removal of the clogging material, the soil
should be disked or harrowed to break up any crusting that may have developed at or near
the surface. Disking or plowing clogging layers as such into the soil gives short-term
relief, but eventually fines and other clogging materials will accumulate in the topsoil and
the entire disk or plow layer must be removed. For good quality surface water with very
low suspended solids contents, drying and cleaning may be necessary once a year or
maybe every few years. Where soils are relatively fine textured or have many stones, clogging control becomes a major challenge (see Sec. 24.6). Where the water is extremely
muddy or where inadequately treated sewage effluent is used, drying and cleaning may be
needed after each flooding period which may then be as short as 1 or 2 days.

24.4.4 Effect of Water Depth on Infiltration


If there is no clogging layer on the bottom, the groundwater is relatively deep, and the
vadose zone is uniform, the downward flow in the vadose zone is governed by gravity and
water depth in recharge basins only has a minor effect on infiltration (Bouwer and Rice,
1989). For example, if the general groundwater table away from the infiltration system is
at a depth of 10 m, increasing the water depth in the basin from 0.1 to 1.1 m (about 10fold) would increase the infiltration rate only by about 10 percent. If, on the other hand,
infiltration is controlled by a clogging layer on the bottom and the underlying material is
unsaturated, the infiltration rate theoretically increases almost linearly with water depth,
so that doubling the water depth would double the infiltration rate (Bouwer, 1982). In reality, however, this seldom occurs because clogging layers often are loose deposits that are
quite compressible. Since increasing water depths produce increasing intergranular pressures in the clogging layer, this layer becomes compressed as the water depth in the basin
increases and, hence, becomes denser and less permeable (Bouwer and Rice, 1989). Thus,
in practice, increasing the water depth in an infiltration basin with a clogging layer will
cause a less than linear increase and maybe even a decrease (if the clogging layer is very
compressible) in infiltration rate. If the infiltration rate does not increase linearly with
water depth, increasing the water depth will then reduce the rate of turnover or the replacement rate of the water in the basin. This increases the time that water is exposed to sunlight which, for eutrophic water with relatively high concentrations of nitrogen, phosphorus, and organic carbon, means additional algae growth and especially the single-cell type
algae like Carteria klebsii that give the water a green appearance. On infiltration, these
algae form an algae filter cake on the bottom which increases the clogging process. Also,
photosynthezing algae absorb dissolved carbon dioxide from the water, which at high
algae concentrations increases the pH of the water to 9, 10, or maybe even 11. This in turn
causes precipitation of dissolved CaCO3, which further aggravates the clogging process
and, hence, decreases infiltration rates. This explains why increasing water depths in infiltration basins has actually produced decreases in infiltration rates, to the surprise and dismay of operators who wanted to increase infiltration rates!

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

Artifical Recharge of Groundwater: Systems, Design, and Management 24.13

For this reason, and also to allow quick draining of basins by natural infiltration when
a drying period is needed, shallow basins (water depth of about 20 cm) are preferred. If
plant growth becomes a problem, water depths may periodically have to be increased to
kill the plants before they emerge above the water level. Deep basins (10 m or more), like
old gravel pits, are not very good for recharge because clogging layers become compressed and infiltration rates are so low that the pits need to be pumped out to start a drying period so that clogging layers can be removed. This is expensive and slow, so that drying often is delayed and infiltration rates remain disappointingly low. An advantage of
deep basins is, however, that they can store more water in times of high runoff events. For
subsequent recharge, however, this water should then be conveyed to shallow basins
where clogging is less and easier controlled by drying and cleaning.
Where groundwater levels are so high that they are close to the bottom of the basin
or even above the bottom, the flow in the aquifer is mainly in the horizontal direction
so that infiltration rates are controlled by the slope of the water table away from the
basins. Under these conditions, increasing the water depth in the basins will increase
infiltration rates in direct proportion to the vertical distance between the water surface
in the basin and the groundwater level some distance away from the basin (Bouwer,
1990; Bouwer and Rice, 1989).
If the groundwater table is above the bottom of a recharge basin, the basin can never
be dried for infiltration recovery and removal of the clogging layer. Thus, sediment continues to accumulate in those basins and infiltration occurs mainly through the sides of the
basins which are then made and kept as steeply as possible to minimize sediment accumulation on the basin banks. Under those conditions, increasing the water depth in the
basin will also increase the infiltration or recharge rate.

24.4.5 Effect of Groundwater Depth on Infiltration Rate


Usually, bottoms and banks of infiltration basins are covered with a clogging layer that
controls and reduces infiltration rates so that the underlying soil material is unsaturated
(Fig. 24.5). The water content in the unsaturated zone then is at a value whereby the corresponding unsaturated hydraulic conductivity is numerically equal to the infiltration rate,
since the downward flow is due to gravity alone with a hydraulic gradient of 1 (Bouwer,
1982). The unsaturated zone breaks the hydraulic continuity between the basin and the
aquifer, so that infiltration rates are independent of the depth to groundwater, as long as
the water table is deep enough so that the top of the capillary fringe above the water table
is below the bottom of the basin. This capillary fringe may be about 1 ft (30 cm) thick for
medium sands, more for finer sands or soils, and less for coarse sands. Thus, a conservative conclusion is that as long as the groundwater table is more than 3 ft (about 1 m) below
the bottom of a basin where infiltration is controlled by a clogging layer on the bottom,
infiltration rates are unaffected by changes in groundwater levels. If the groundwater table
rises, infiltration rates will start to decrease when the capillary fringe reaches the bottom
of the basin, and continue to decrease linearly with depth of groundwater below the water
level in the basin, until they become zero when the groundwater table has risen to the
water surface in the basin.
Where the water for recharge is exceptionally clear and free from nutrients and organic carbon, temperatures are low, and soils are relatively coarse, infiltration can go on for
considerable time without development of a clogging layer on the bottom. In that case,
there is direct hydraulic continuity between the basin and the aquifer with the groundwater table joining the water surface in the basin (Fig. 24.6; Bouwer, 1969, 1978).
Groundwater levels are then characterized by the depth Dw of the groundwater table below

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

24.14

Chapter Twenty-Four

the water surface in the basin at sufficient distance from the basin to be relatively unaffected by the recharge flow system (Fig. 24.6). If Dw is relatively large, the flow below the
recharge basin is mainly downward and gravity controlled (Bouwer 1969, 1978) with a
hydraulic gradient of about 1 (Fig. 24.6). In that case, infiltration rates are essentially
unaffected by depth to groundwater. However, if groundwater levels rise and Dw decreases, the flow from the recharge basin becomes more and more lateral until it is controlled
by the slope of the water table away from the basin (Fig. 24.6; Bouwer 1969, 1978).
Modeling these flow systems on an electrical resistance network analog has shown that
the change from gravity controlled flow to slope-of-the-water-table-controlled flow
occurs when Dw is about twice the width W (or diameter) of the recharge system (Bouwer,
1990). This relation is shown in Fig. 24.7 where I is the infiltration rate per unit area of
water surface in the basin and K is the hydraulic conductivity in the wetted zone or aquifer.
Thus, as long as Dw  2K, infiltration rates decrease linearly with decreasing Dw and reach
zero when Dw  0 (Fig. 24.7). These relations apply to uniform, isotropic underground
formations. Anisotropic or stratified situations need to be considered on a case-by-case
basis.
Infiltration rates in clean basins (no clogging layers) thus are more sensitive to depth
to groundwater than in clogged basins. Clogged basins are the rule rather than the exception and groundwater mounds can rise much higher there than below clean basins before
reductions in infiltration rates occur. Sometimes maximum mound heights are dictated by
circumstances other than their effect on infiltration rates, such as presence of sanitary
landfills, underground sewers or other pipelines, basements (especially deep basements of
commercial buildings), cemeteries, and deep  rooted vegetation like old trees that may
die when groundwater levels rise too high.

24.4.6 Induced Recharge


One method for increasing recharge of groundwater from losing streams is to lower
groundwater levels near the stream by groundwater pumping. As described in the previ-

FIGURE 24.6 Recharge basin with high groundwater table and


lateral flow in aquifer controlled by slope of water table (top)
and with deep groundwater table with downward flow below basin
controlled by gravity (bottom).

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

DW W

FIGURE 24.7 Dimensionless plot of seepage (expressed as I/K) versus depth to groundwater (expressed as Dw/W) for clean stream channel or basin (no clogging layer on bottom).

1
I
K

Artifical Recharge of Groundwater: Systems, Design, and Management 24.15

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

24.16

Chapter Twenty-Four

ous section, this will only increase recharge or seepage losses from the stream if the
stream bottom is not covered by a clogging layer and the groundwater table is relatively
high so that the depth to groundwater below the water level in the stream is less than two
stream widths (Fig. 24.7). If the groundwater level is already more than two stream widths
below the water surface in the stream, recharge is already near maximum and further lowering of the groundwater will have little or no effect on recharge (Fig. 24.7). If the stream
wetted perimeter is covered by a clogging layer, lowering of groundwater levels will only
increase recharge if the groundwater table is above the stream bottom or only a small distance below the stream bottom so that the capillary fringe still touches the bottom. If the
groundwater level already is deep enough for the top of the capillary fringe to be below
the stream bottom, the soil material below the stream bottom is unsaturated. There is then
no direct hydraulic connection between the stream and the aquifer, so that further lowering of the groundwater level by pumping from wells will not increase the recharge or seepage losses from the stream.
Groundwater pumping can decrease streamflow in two ways. If the groundwater table
is above the water surface in the stream, the stream is gaining and groundwater moves into
the stream and supports the base flow in the stream. Any pumping of groundwater will
then diminish the flow of groundwater into the stream. If the groundwater table is below
the water surface in the stream, lowering the groundwater table will increase recharge or
seepage losses from the stream when the groundwater table is high, but not when it is relatively low, as explained above. These relationships can then be used in resolving situations where there are real or perceived conflicts between owners of surface water rights
and holders of groundwater pumping rights (Bouwer and Maddock, 1997).
Where groundwater levels are high enough to support growth of phreatophytes (plants
and trees that live off groundwater), a lowering of groundwater levels can reduce the water
uptake by phreatophytes (Bouwer, 1975) and, hence, increase the capture of groundwater
by the aquifer (Bouwer and Maddock, 1997). This means that groundwater pumping can
then lessen the reduction in flow of tributary groundwater to gaining streams, and reduce
water losses from losing streams.

24.5 GROUNDWATER MOUNDING


24.5.1 Perched Groundwater Mounds
When infiltrometer data indicate surface infiltration rates that are acceptable for artificial
recharge, the next step is to investigate the vadose zone to make sure that the infiltrated
water can move unimpeded to the underlying groundwater. Trenches or pits dug by backhoes allow inspection of the soil profile to a depth of about 7 m. If fine-textured or cemented (caliche) materials are found at a certain depth, their hydraulic conductivity can be
determined with infiltrometer tests or other methods (Bouwer, 1978) on the bottom of a
trench or pit excavated to that depth. Doing these measurements in a sloping trench bottom will give a profile of hydraulic conductivity. Soil borings are necessary to evaluate
deeper layers. Large-diameter (about 1 m) holes drilled with a bucket auger enable assessment of infiltration rates and hydraulic conductivities at great depth. After removing loose
material from the bottom, a few cm of water on the bottom are maintained, and the infiltration rate is measured. This can be managed from the top. Bucket augers can drill holes
to about 50 m.
If a restricting layer is detected, the height of the perching mound can be calculated by
applying Darcys equation to the vertically downward flow in the perching mound and the

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

Artifical Recharge of Groundwater: Systems, Design, and Management 24.17

flow through the restricting layer (Fig. 24.8). If the material below the restricting layer is
relatively coarse, the pressure head of the water at the bottom of the restricting layer can
be taken as zero. For finer materials below the restricting layer, the water-entry values listed in Sec. 24.4.2 be taken as a first estimate. The resulting equation for the perched
mound, assuming zero pressure head for the water at the bottom of the restricting layer, is
then (Bouwer et al., 1997).
i
  1
Kr
Lp  Lr 
(24.5)
1  i
Ks
where Lp  height of perching mound above restricting layer, Lr  thickness of restricting layer, i  infiltration rate, Kr  hydraulic conductivity of restricting layer, and Ks 
hydraulic conductivity of soil above restricting layer.
Often, i will be much smaller than Ks because surface soils are finer textured than
deeper soils, or there is a clogging layer on the surface soil that reduces infiltration.
Also, i often will be much larger than Kr. For these conditions, Eq. 24.5 can be simplified to
L
Lp  i r
Kr

(24.6)

Lp should be small enough so that the top of the perched mound is deep enough to
avoid reductions in infiltration rates, as discussed in the Sec. 24.4.5. Perched mounds
above noncontinuous perching layers (lenses or strips) are not as high as above continuous perching layers with the same Lr and Kr because there is then also lateral flow in the
perched mound. Thus, Eqs. 24.5 or 24.6 will then overestimate the height of the perched
mound. Where perched mounds would be too high, recharge may be achieved with long
narrow basins that allow lateral flow in the mound and reduce its height as the perched
mound spreads laterally on the restricting layer until all the recharge water will pass
through the restricting layer (Bouwer, 1978). Another possibility is to use infiltration
basins that are excavated through the restricting layer (Fig. 24.1B). If the restricting layer

FIGURE 24.8 Geometry and symbols for perched


mound above restricting layer in vadose zones.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

24.18

Chapter Twenty-Four

is too deep, recharge trenches (Fig. 24.1C) or recharge shafts (large diameter holes in the
vadose zone, also called dry wells or vadose zone wells) can be used (Fig. 24.1D).

24.5.2 Groundwater Mounds


Numerous papers have been published on the rise of groundwater mounds on the aquifer
in response to infiltration from a recharge system, and some also on the fall of the mound
after infiltration has stopped (Glover, 1964: Hantush, 1967; Marino, 1975a, b; Warner et
al., 1989). The usual assumption is a uniform aquifer of infinite extent with no other
recharges or discharges. One of the difficulties in getting meaningful results from the equations is getting a representative value of aquifer transmissivity. The most reliable transmissivity data come from calibrated models. Next come Theis-type pumping tests, step-drawdown and other pumped well tests, and slug tests, in decreasing order of sampling size.
Slug tests, while simple to carry out, always have the problem of how to get representative
areal values from essentially point measurements. Averages from various tests often seriously underestimate more regional values (Bouwer, 1996, and references therein).
Piezometers at two different depths in the center of a mound enable the determination of
both vertical and horizontal hydraulic conductivity of an aquifer already being recharged
with an infiltration system (Bouwer et al. 1974). In deep or thick unconfined aquifers,
streamlines of recharge flow systems are concentrated in the upper or active portion of
the aquifer, with much less flow and almost stagnant water in the deeper or passive portion of the aquifer. Use of transmissivities of the entire aquifer between the water table and
the impermeable lower boundary for mound calculations can then seriously underestimate
the rise of the mound. Older work (Bouwer, 1962) with resistance network analog modeling showed that for rectangular recharge areas, the thickness of the active, upper portion of
the aquifer is about equal to the width of the recharge area. This thickness should then be
multiplied by K of the upper aquifer to get an effective transmissivity for mounding predictions. Also, if Hantushs or another equation is used to calculate long-term mound formation, as for water banking in areas with deep groundwater levels, larger transmissivity
values should be used to reflect the increase in transmissivity as the groundwater level rises.
Otherwise, the Hantush equation will seriously overestimate the mound rise.
The best way to get representative transmissivity values for artificial recharge systems
is to have a large enough infiltration area that produces a groundwater mound, and to calculate the transmissivity from the rise of the groundwater mound using, for example,
Hantushs equation (Eq. 24.7). The fillable porosity to be used in the equations for mound
rise usually are larger than the specific yield of the aquifer, because vadose zones often
are relatively dry, especially in dry climates and if they consist of coarse materials like
sands and gravels. The fillable porosity should be taken as the difference between existing and saturated water contents of the material outside the wetted zone below the infiltration system. The Hantushs equation is
vt
hx,y,t  H  a {F[W/2  x)n, (L/2  y)n]
4f
+ F[W/2  x)n, (L/2  y)n]
+ F[W/2  x)n, (L/2  y)n]
+ F[W/2  x)n, (L/2  y)n]
(24.7)
where hx,y,t  height of water table above impermeable layer at x, y, and time (Fig. 24.9),
H  original height of water table above impermeable layer, va  arrival rate at water
table of water from infiltration basin, t  time since start of recharge, f  fillable porosi-

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

Artifical Recharge of Groundwater: Systems, Design, and Management 24.19

ty (1  f  0), L  length of recharge basin (in y direction), W  width of recharge basin


(in x direction), n  (4t T/f)-1/2, and F(, )  10 erf(  1/2)  erf(  1/2)d [which was
tabulated by Hantush (Table 24.1)]
where  (W/2 + x) n or (W/2  x)n and  (L/2 + y)n or (L/2  y)n.
Of most interest to operators and managers often, is the long-term effect of recharge
on groundwater; that is, where the groundwater mound will be 10, 20, or 50 years from
now, how much water can be stored or banked underground, will the whole area become
waterlogged, and how must the water be recovered from the aquifer to prevent waterlogging? Computer models can be used to simulate regional recharge inputs and pumped
well outputs for the aquifer. However, a quick idea about ultimate or quasi-equilibrium
mound heights can be obtained from a steady-state analysis where the mound is considered to be in equilibrium with a constant water table at some depth and at some distance
from the infiltration system. The constant faraway water table is obtained by groundwater pumping, discharge into surface water like rivers or lakes, or some other control.
Usually, recharge systems consist of a number of basins or other infiltration facilities.
Steady-state equations were developed for two general geometries of the entire recharge
area: (1) the basins form a long strip with a length of at least 5 times the width so that after
long times it still performs about the same as an infinitely long strip (Glover, 1964), and
(2) the basins are in a round or square or irregular area that can be handled as an equivalent circular area (Bouwer et al., 1999). For the long strip (Fig. 24.10) the groundwater
flow away from the strip was taken as linear horizontal flow (Dupuit-Forchheimer flow).
Below the infiltration area, the lateral flow was assumed to increase linearly with distance
from the center. The lateral flow was then assumed to be constant between the edge of the
recharge system at distance W/2 from the center and the constant control water table at
distance Ln from the edge (Fig. 24.10). This yielded the equation

FIGURE 24.9 Geometry and symbols for recharge system and groundwater mound.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

24.20

Chapter Twenty-Four

iW W + L
Hc  Hn  
 
n
2T 4

(24.8)

for the ultimate rise of the groundwater mound below the center of the recharge strip when
there is equilibrium between recharge and pumping from the aquifer (Bouwer et al.,
1997). The symbols in this equation are (Fig. 24.10):
Hc  height of groundwater mound in center of recharge area,
Hn  height of groundwater table at control area,
i

 average infiltration rate in recharge area (total recharge divided by total area),

W  width of recharge area,


Ln  distance between edge of recharge area and control area, and
T  transmissivity of aquifer.
For a round or square recharge area (Fig. 24.11), the groundwater flow will be radially away from the area. The equilibrium height of the mound below the center of the
recharge system above the constant groundwater table at distance Rn from the center
of the recharge system then can be calculated with radial flow theory (Bouwer et al.,
1999) as
iR2 1  2ln Rn
Hc  Hn  


4T
R

(24.9)

where R is the radius or equivalent radius of the recharge area, Rn is the distance from the
center of the recharge area to the water-table control area (Fig. 24.11) and the other symbols are as defined above.
Equations (24.8) and (24.9) thus can be used to predict the final mound height below
a recharge area for a given elevation of the control water table at distance Rn or Ln from
the recharge area. As indicated by the Hantushs equation, the value of T in Eq. (24.8) and
(24.9) must reflect the average transmissivity of the aquifer at the ultimate equilibrium
mound height. If the ultimate mound height is too high, Rn or Ln must be reduced by
groundwater pumping from wells closer to the recharge area or Hn must be reduced by
pumping more groundwater . Equations (24.8) and (24.9) can then indicate where groundwater should be recovered and to what depth groundwater levels should be pumped to prevent groundwater tables below the recharge areas from rising too high. Ultimate mound
heights can also be reduced by making the recharge area longer and narrower, or by reducing recharge rates as can be achieved by using less water for recharge or by spreading the
infiltration facilities out over a larger area.

24.6 CHALLENGING SOILS


Because artificial recharge of groundwater offers many advantages, municipalities and
water districts, or other entities may still wish to do artificial recharge even though permeable soils are not available and they may have to work with finer soils (sandy loams to
loams) with marginal hydraulic conductivity. Such systems may only have infiltration
rates of 0.050.1 m/day and hydraulic loading rates of 10 to 20 m/yr. At these rates, evaporation losses (0.52.5 m/year for year-round systems, depending on climate) can become
significant. To maximize infiltration rates, operators of such systems are tempted to disk

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

FIGURE 24.10 Geometry and symbols for groundwater mound below long infiltration area (strip).

Artifical Recharge of Groundwater: Systems, Design, and Management 24.21

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

Chapter Twenty-Four

FIGURE 24.11 Geometry and symbols for groundwater mounds below round infiltration area.

24.22

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

TABLE 24.1 Values of the Function F ( , ) in Eq. (24.7) for Different values of and

0.02
0.04 0.06
0.08
0.10
0.14
0.18 0.22
0.26
0.30

0.02
0.0041 0.0073 0.0101 0.0125 0.0146 0.0184 0.0216 0.0243 0.0267 0.0288
0.04
0.0073 0.0135 0.0188 0.0236 0.0278 0.0353 0.0416 0.0470 0.0518 0.0559
0.06
0.0101 0.0188 0.0266 0.0335 0.0398 0.0509 0.0602 0.0684 0.0754 0.0817
0.08
0.0125 0.0236 0.0335 0.0425 0.0508 0.0652 0.0776 0.0884 0.0978 0.1060
0.10
0.0146 0.0278 0.0398 0.0508 0.0608 0.0786 0.0939 0.1072 0.1188 0.1290
0.14
0.0184 0.0353 0.0509 0.0652 0.0786 0.1025 0.1232 0.1414 0.1573 0.1714
0.18
0.0216 0.0416 0.0602 0.0776 0.0939 0.1232 0.1490 0.1716 0.1916 0.2094
0.22
0.0243 0.0470 0.0684 0.0884 0.1072 0.1414 0.1716 0.1984 0.2222 0.2433
0.26
0.0267 0.0518 0.0754 0.0978 0.1188 0.1573 0.1916 0.2222 0.2494 0.2737
0.30
0.0288 0.0559 0.0817 0.1060 0.1290 0.1714 0.1094 0.2433 0.2737 0.3009
0.34
0.0306 0.0596 0.0871 0.1133 0.1391 0.1839 0.2251 0.2621 0.2954 0.3252
0.38
0.0322 0.0628 0.0920 0.1197 0.1461 0.1949 0.2391 0.2789 0.3147 0.3470
0.42
0.0377 0.0657 0.0963 0.1254 0.1532 0.2048 0.2515 0.2938 0.3320 0.3665
0.46
0.0349 0.0683 0.1001 0.1305 0.1595 0.2135 0.2626 0.3071 0.3474 0.3839
0.50
0.0361 0.0705 0.1035 0.1350 0.1650 0.2212 0.2724 0.3189 0.3612 0.3995
0.54
0.0371 0.0725 0.1065 0.1389 0.1700 0.2281 0.2812 0.3295 0.3735 0.4134
0.58
0.0380 0.0743 0.1091 0.1425 0.1744 0.2343 0.2890 0.3389 0.3844 0.4257
0.62
0.0387 0.0759 0.1115 0.1456 0.1783 0.2397 0.2959 0.3472 0.3941 0.4368
0.66
0.0394 0.0773 0.1136 0.1484 0.1718 0.2445 0.3020 0.3547 0.4027 0.4466
0.70
0.0401 0.0785 0.1154 0.1509 0.1849 0.2488 0.3075 0.3612 0.4104 0.4553
0.74
0.0406 0.0796 0.1117 0.1531 0.1876 0.2526 0.3123 0.3671 0.4172 0.4630
0.78
0.0411 0.0806 0.1185 0.1550 0.1900 0.2559 0.3166 0.3722 0.4232 0.4699
0.82
0.0415 0.0814 0.1198 0.1567 0.1921 0.2589 0.3203 0.3768 0.4286 0.4760
0.86
0.0419 0.0822 0.1209 0.1582 0.1940 0.2615 0.3237 0.3808 0.4333 0.4813
0.90
0.0422 0.0828 0.1219 0.1595 0.1957 0.2638 0.3266 0.3844 0.4374 0.4860
0.94
0.0425 0.0834 0.1228 0.1607 0.1971 0.2658 0.3292 0.3875 0.4411 0.4902
0.98
0.0428 0.0839 0.1236 0.1617 0.1984 0.2676 0.3314 0.3902 0.4442 0.4938
1.00
0.0429 0.0842 0.1239 0.1622 0.1990 0.2684 0.3324 0.3914 0.4457 0.4955
1.20
0.0437 0.0858 0.1263 0.1654 0.2030 0.2740 0.3396 0.4001 0.4558 0.5070
1.40
0.0441 0.0866 0.1275 0.1669 0.2049 0.2767 0.3431 0.4043 0.4608 0.5127
1.80
0.0444 0.0871 0.1283 0.1680 0.2062 0.2785 0.3454 0.4071 0.4641 0.5165
2.00
0.0444 0.0871 0.1284 0.1681 0.2064 0.2787 0.3457 0.4075 0.4645 0.5169
2.20
0.0444 0.0872 0.1284 0.1682 0.2065 0.2788 0.3458 0.4076 0.4646 0.5171
2.50
0.0444 0.0872 0.1284 0.1682 0.2065 0.2788 0.3458 0.4077 0.4647 0.5172
3.00
0.0444 0.0872 0.1284 0.1682 0.2065 0.2789 0.3458 0.4077 0.4647 0.5172
0.34
0.0306
0.0596
0.0871
0.1133
0.1381
0.1839
0.2251
0.2621
0.2954
0.3252
0.3520
0.3761
0.3976
0.4169
0.4341
0.4495
0.4633
0.4756
0.4865
0.4962
0.5048
0.5125
0.5192
0.5252
0.5305
0.5351
0.5392
0.5410
0.5540
0.5603
0.5645
0.5651
0.5653
0.5653
0.5654

0.38
0.0322
0.0628
0.0920
0.1197
0.1461
0.1941
0.2391
0.2789
0.3147
0.3470
0.3761
0.4022
0.4256
0.4466
0.4654
0.4823
0.4973
0.5108
0.5227
0.5334
0.5429
0.5513
0.5587
0.5653
0.5711
0.5762
0.5807
0.5827
0.5969
0.6039
0.6086
0.6092
0.6094
0.6095
0.6095

0.42
0.0337
0.0657
0.0963
0.1254
0.1532
0.2048
0.2515
0.2938
0.3320
0.3665
0.3976
0.4256
0.4508
0.4734
0.4937
0.5119
0.5281
0.5427
0.5556
0.5672
0.5774
0.5865
0.5946
0.6017
0.6080
0.6136
0.6184
0.6206
0.6362
0.6438
0.6489
0.6495
0.6497
0.6498
0.6499

0.46
0.0349
0.0683
0.1001
0.1305
0.1595
0.2135
0.2626
0.3071
0.3474
0.3839
0.4169
0.4466
0.4734
0.4975
0.5191
0.5385
0.5559
0.5715
0.5854
0.5977
0.6087
0.6185
0.6272
0.6348
0.6416
0.6476
0.6528
0.6552
0.6719
0.6801
0.6856
0.6863
0.6865
0.6867
0.6867

0.50
0.0361
0.0705
0.1035
0.1350
0.1650
0.2212
0.2724
0.3189
0.3612
0.3995
0.4341
0.4654
0.4937
0.5161
0.5420
0.5626
0.5810
0.5975
0.6122
0.6254
0.6371
0.6475
0.6567
0.6648
0.6721
0.6784
0.6840
0.6865
0.7044
0.7132
0.7190
0.7198
0.7200
0.7202
0.7202

0 .54
0.0371
0.0725
0.1065
0.1389
0.1700
0.2281
0.2812
0.3295
0.3735
0.4134
0.4495
0.4823
0.5119
0.5385
0.5626
0.5842
0.6036
0.6209
0.6364
0.6503
0.6627
0.6736
0.6834
0.6920
0.6996
0.7063
0.7123
0.7150
0.7339
0.7432
0.7494
0.7502
0.7505
0.7506
0.7506

0.58
0.0380
0.0743
0.1091
0.1425
0.1744
0.2343
0.2890
0.3389
0.3844
0.4257
0.4633
0.4973
0.5281
0.5559
0.5810
0.6036
0.6238
0.6420
0.6582
0.6728
0.6857
0.6972
0.7074
0.7165
0.7245
0.7316
0.7378
0.7406
0.7605
0.7704
0.7769
0.7778
0.7781
0.7782
0.7782

0.62
0.0387
0.0759
0.1115
0.1456
0.1783
0.2397
0.2959
0.3472
0.3941
0.4368
0.4756
0.5108
0.5427
0.5715
0.5975
0.6209
0.6420
0.6609
0.6778
0.6929
0.7064
0.7184
0.7291
0.7386
0.7469
0.7543
0.7608
0.7638
0.7846
0.7949
0.8018
0.8027
0.8030
0.8032
0.8032

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

Artifical Recharge of Groundwater: Systems, Design, and Management 24.23

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

1.00
0.0429
0.0842
0.1239
0.1622
0.1990
0.2684
0.3224
0.3914
0.4457
0.4955
0.5410
0.5827
0.6206
0.6552
0.6865
0.7150
0.7406
0.7638
0.7846
0.8034
0.8201
0.8351
0.8485
0.8604
0.8710
0.8803
0.8886
0.8924
0.9191
0.9324
0.9414
0.9426
0.9430
0.9432
0.9433

1.20
0.0437
0.0858
0.1263
0.1654
0.2030
0.2740
0.3396
0.4001
0.4558
0.5070
0.5540
0.5969
0.6362
0.6719
0.7044
0.7379
0.7605
0.7846
0.8064
0.8259
0.8434
0.8591
0.8731
0.8855
0.8966
0.9064
0.9151
0.9191
0.9472
0.9614
0.9709
0.9722
0.9726
0.9728
0.9729

1.40
0.0441
0.0866
0.1275
0.1669
0.2049
0.2777
0.3431
0.4043
0.4608
0.5127
0.5603
0.6039
0.6438
0.6801
0.7132
0.7432
0.7704
0.7949
0.8171
0.8370
0.8549
0.8710
0.8853
0.8980
0.9094
0.9195
0.9284
0.9324
0.9614
0.9759
0.9858
0.9871
0.9875
0.9878
0.9878

1.80
0.0444
0.0871
0.1283
0.1680
0.2062
0.2785
0.3454
0.4071
0.4641
0.5165
0.5645
0.6086
0.6489
0.6856
0.7190
0.7494
0.7799
0.8018
0.8243
0.8445
0.8627
0.8789
0.8935
0.9065
0.9180
0.9282
0.9373
0.9414
0.9709
0.9858
0.9959
0.9972
0.9977
0.9979
0.9980

2.00
0.0444
0.0871
0.1284
0.1681
0.2064
0.2787
0.3457
0.4075
0.4645
0.5159
0.5651
0.6092
0.6495
0.6863
0.7198
0.7502
0.7778
0.8027
0.8252
0.8454
0.8636
0.8799
0.8945
0.9075
0.9191
0.9294
0.9384
0.9426
0.9722
0.9871
0.9972
0.9985
0.9990
0.9992
0.9993

2.20
0.0444
0.0872
0.1284
0.1682
0.2065
0.2788
0.3454
0.4076
0.4646
0.5171
0.5653
0.6094
0.6497
0.6865
0.7200
0.7505
0.7781
0.8030
0.8255
0.8458
0.8640
0.8803
0.8949
0.9079
0.9195
0.9298
0.9389
0.9430
0.9726
0.9875
0.9972
0.9990
0.9995
0.9997
0.9998

2.50
0.0444
0.0882
0.1284
0.1682
0.2065
0.2788
0.3454
0.4077
0.4647
0.5172
0.5653
0.6095
0.6498
0.6867
0.7202
0.7506
0.7782
0.8032
0.8257
0.8460
0.8642
0.8805
0.8951
0.9081
0.9187
0.9300
0.9391
0.9432
0.9728
0.9878
0.9879
0.9992
0.9997
1.0000
1.0000

3.00
0.0444
0.0882
0.1284
0.1682
0.2065
0.2788
0.3454
0.4077
0.4647
0.5172
0.5654
0.6095
0.6499
0.6867
0.7202
0.7506
0.7782
0.8032
0.8257
0.8460
0.8642
0.8805
0.8951
0.9081
0.9187
0.9300
0.9391
0.9433
0.9729
0.9878
0.9980
0.9934
0.9998
1.0000
1.0000

24.24

Source: From Hantush (1967).

TABLE 24.1 Values of the Function F ( , ) in Eq. (24.7) for Different Values of and
0.62
0.66 0.70
0.74
0.78
0.82
0.86 0.90
0.94
0.98

0.02
0.0387 0.0394 0.0401 0.0406 0.0411 0.0415 0.0419 0.0422 0.0425 0.0428
0.04
0.0759 0.0773 0.0785 0.0796 0.0806 0.0814 0.0822 0.0828 0.0834 0.0839
0.06
0.1115 0.1136 0.1154 0.1171 0.1185 0.1198 0.1209 0.1219 0.1228 0.1236
0.08
0.1456 0.1484 0.1509 0.1531 0.1550 0.1567 0.1582 0.1595 0.1606 0.1617
0.10
0.1783 0.1818 0.1849 0.1876 0.1900 0.1921 0.1940 0.1957 0.1971 0.1984
0.14
0.2397 0.2445 0.2488 0.2526 0.2559 0.2589 0.2615 0.2638 0.2658 0.2676
0.18
0.2959 0.3020 0.3075 0.3123 0.3166 0.3203 0.3237 0.3266 0.3292 0.3314
0.22
0.3472 0.3547 0.3612 0.3671 0.3722 0.3768 0.3808 0.3844 0.3875 0.3902
0.26
0.3941 0.4027 0.4104 0.4172 0.4232 0.4286 0.4333 0.4374 0.4411 0.4442
0.30
0.4368 0.4466 0.4553 0.4630 0.5699 0.5760 0.4813 0.4860 0.4902 0.4938
0.34
0.4756 0.4865 0.4962 0.5048 0.5125 0.5192 0.5252 0.5305 0.5351 0.5392
0.38
0.5108 0.5227 0.5334 0.5429 0.5213 0.5587 0.5653 0.5711 0.5762 0.5807
0.42
0.5427 0.5556 0.5672 0.5774 0.5865 0.5946 0.6017 0.6080 0.6136 0.6184
0.46
0.5715 0.5854 0.5977 0.6087 0.6185 0.6272 0.6348 0.6416 0.6476 0.6528
0.50
0.5975 0.6122 0.6254 0.6371 0.6475 0.6567 0.6648 0.7721 0.6784 0.6840
0.54
0.6209 0.6364 0.6503 0.6627 0.6736 0.6834 0.6920 0.6996 0.7063 0.7123
0.58
0.6420 0.6482 0.6728 0.6857 0.6972 0.7074 0.7165 0.7245 0.7316 0.7378
0.62
0.6609 0.6778 0.6929 0.7064 0.7784 0.7291 0.7386 0.7469 0.7543 0.7608
0.66
0.6778 0.7953 0.7110 0.7250 0.7375 0.7486 0.7584 0.7671 0.7748 0.7816
0.70
0.6929 0.7110 0.7272 0.7417 0.7546 0.7660 0.7762 0.7852 0.7932 0.8002
0.74
0.7064 0.7250 0.7414 0.7566 0.7698 0.7816 0.7921 0.8014 0.8096 0.8168
0.78
0.7184 0.7375 0.7546 0.7698 0.7834 0.7956 0.8063 0.8159 0.8243 0.8317
0.82
0.7291 0.7486 0.7660 0.7816 0.7956 0.8080 0.8190 0.8288 0.8374 0.8450
0.86
0.7386 0.7584 0.7762 0.7921 0.8063 0.8190 0.8302 0.8402 0.8491 0.8569
0.90
0.7469 0.7671 0.7852 0.8014 0.8159 0.8288 0.8402 0.8504 0.8594 0.8674
0.94
0.7543 0.7748 0.7932 0.8096 0.8243 0.8374 0.8491 0.8594 0.8686 0.8767
0.98
0.7608 0.7816 0.8002 0.8168 0.8317 0.8450 0.8569 0.8674 0.8767 0.8849
1.00
0.7638 0.7846 0.8034 0.8201 0.8351 0.8485 0.8604 0.8710 0.8803 0.8886
1.20
0.7846 0.8064 0.8259 0.8434 0.8591 0.8731 0.8855 0.8966 0.9064 0.9151
1.40
0.7949 0.8171 0.8370 0.8549 0.8710 0.8853 0.8980 0.9094 0.9195 0.9284
1.80
0.8018 0.8243 0.8445 0.8627 0.8789 0.8935 0.9065 0.9180 0.9282 0.9373
2.00
0.8027 0.8252 0.8454 0.8636 0.8799 0.8945 0.9075 0.9191 0.9594 0.9384
2.20
0.8030 0.8255 0.8458 0.8640 0.8803 0.8949 0.9079 0.9195 0.9298 0.9389
2.50
0.8032 0.8257 0.8460 0.8642 0.8805 0.8951 0.9081 0.9197 0.9300 0.9391
3.00
0.8032 0.8257 0.8460 0.8642 0.8805 0.8951 0.9081 0.9197 0.9300 0.9391

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

Chapter Twenty-Four

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

Artifical Recharge of Groundwater: Systems, Design, and Management 24.25

or rip the soil at the end of each drying period. This repeated traffic can compact underlying soils and reduce their permeability to the point where infiltration rates become critically low. Deep ripping may then be necessary, after which the basin bottom should be
cleaned less frequently. Rather than disking which mixes clogging materials with surface
soil, clogging materials should be removed by shaving the basins with a front-end
loader at the end of a dry period, or raking them. This may have to be done about once a
year. After shaving or raking the basin, it may have to be disked or harrowed to a depth of
about 10 cm to break up deeper clogging layers.
Finer soils in recharge basins are more vulnerable to clogging than coarser soils
because pore sizes are smaller and easier to block by suspended solids. Biofilm growth on
the soil particles could significantly reduce effective pore diameters and, hence, infiltration rates. Also, clay and other small soil particles could migrate downward due to the
seepage force of the infiltrated water and accumulate a small distance (often only a few
mm or less) below the surface where it can form a micro clogging or restricting layer.
This process, called fine particle movement or wash-out wash-in, is an important factor in
soil crusting due to rainfall and has been well documented in the soils literature (Sumner
and Stewart, 1992). Because of this fine particle movement, the normal drying of basins
for infiltration recovery and periodic removal of the cracked and curled up clogging layer
may have to be followed by shallow disking or harrowing to break up any wash-in layers
that may have formed some small distance below the surface. This leaves a rough surface
and loose soils so that when the basins are filled again the surface soil may cave and
slough, causing soil to be stirred up and become suspended. This soil could then move
with the water and settle out to start a new clogging layer on the soil surface of the entire
basin. Also, as water rapidly infiltrates when it first enters loose soil, fine-particle migration may be enhanced and the fine particles could migrate further down and accumulate
on the underlying undisturbed soil, where they could form a minirestricting layer. This
layer may then be at a depth of 515 cm below soil surface, depending on how deep the
soil was disked or harrowed. This would require deeper disking, harrowing, or even ripping the next time the basin is dried and cleaned, etc.! To prevent this situation from developing, the basin after disking or harrowing may have to be smoothed and somewhat compacted, as achieved by rolling, so that the surface soil is not stirred up when the basin is
filled and fine particles cannot as readily move downward through the soil. Perhaps fine
particle migration can be minimized by making sure that there is sufficient dissolved calcium in the soil and water to keep the clay particles coagulated or aggregated. This could
be achieved by adding gypsum to the soil or the more soluble calcium chloride to the soil
or water, or by adding acid to the soil or water to dissolve calcium carbonate precipitates
in the soil. Research and testing are necessary to see how these amendments work. Other
infiltration problems that can develop on finer textured soils are hardsetting and structural or depositional crusts (Oster, 1997).
If infiltration basins are constructed in sand or gravel, erosion of the banks (wave action,
sloughing) or bottom (splashing at water inlet and overland flow when filling) is not all that
critical for infiltration because the suspended material quickly settles out and forms a sand
layer on top of the original sand. This layer probably still is permeable and may have a
minor if any effect on infiltration. However, if the same erosion happens to a basin in fine
soil, the suspended material settles out more slowly and there will be segregation of particles. The fine sand particles will settle out first, followed by silt particles which then will
be topped with a blanket of clay particles that can significantly reduce infiltration rates.
Because of this, infiltration basins in fine soils must be designed and managed to completely avoid all soil erosion. Berms or dikes must be compacted to maximum density and
must have mild slopes that are vegetated, covered by plastic sheets that are kept down by a
layer of rock or gravel, or otherwise protected against erosion by rain or wave action. Also,

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

24.26

Chapter Twenty-Four

when the basins are filled, water should be admitted very slowly to avoid erosion by splashing and overland flow. In the system of Fig. 24.12, this can be achieved by filling the basin
from the survival channel, which is on the contour on the uphill side of the berms and thus
can fill the basin over its entire length with minimal inflow per unit length of basin.
Survival channels are kept full of water during drying of the basins so that aquatic life can
survive. Shallow basins with water depths less than about 10in generally are preferable to
deep basins because the clogging layer then tends to be looser and, hence, more permeable
than under larger water depths where it may be compressed by increased intergranular pressures (Bouwer and Rice, 1989). Shallow basins also dry sooner by infiltration after turning off the inflow when decreasing infiltration rates indicate the need for infiltration
restoration by drying and possibly cleaning of the bottom. Deep basins, or basins with sloping bottoms, may take long times to dry completely after the inflow into the basins is
stopped. This reduces long-term hydraulic loading rates.
The role of vegetation in basins still is not clear. While it may aggravate insect problems, it also provides shade which could reduce algae growth in the water and on the bottom and, hence, clogging. Root activity may help keep fine soils open and, hence, more
permeable. Dead vegetation that accumulates on the soil (thatch, detritus) may protect
underlying soil against blockage of soil pores by suspended solids. However, vegetation
also retards drying and recovery of infiltration rates. This may cause problems in the winter which already is a critical period for infiltration because lower temperatures increase
the viscosity of the water which gives lower infiltration rates.
Some recharge systems must be able to accept a certain inflow, for example, sewage
effluent that cannot be discharged somewhere else. Where the total basin area is barely
sufficient to handle the inflow, operators of the system then tend to frequently disk or harrow the basins to break up clogging layers and to keep infiltration rates as high as possible. However, on sensitive (fine-textured) soils this can result in soil compaction problems
deeper down, which leads to reduced infiltration rates. To maintain system capacity,
basins then must be flooded longer and dried shorter, which reduces infiltration recovery
until eventually all basins must be filled to accept the inflow with no opportunity for drying. Infiltration rates then continue to decline due to increased clogging and eventually the
whole system fails and drastic measures must be taken.
If the soil contains a lot of large stones, disking or harrowing is not possible and the
soil must be ripped to restore infiltration rates. Ripping, however, causes stones to move
upward through the soil profile so that frequent ripping eventually produces a surface that
is almost completely covered by stones, like an armored streambed. Infiltration rates then
will be low, because infiltration cannot take place where the soil is covered by stones.
Infiltration then can occur only between the stones. Suspended solids and other clogging
materials accumulate on the soil between the stones and cannot be removed by drying and
scraping. The only solution is to remove the stones and expose as much of the soil surface
as possible.
Some soils change their structure on wetting. These include loose, windblown soils
like loess which settle when flooded (hydrocompaction; Bouwer, 1978). This reduces the
porosity and, hence, the hydraulic conductivity and infiltration rate of the soil. Other soils
are rather porous but the soil particles are cemented together by calcite at their contact
points. On flooding, the calcite dissolves in the water and the soils collapse. For such
collapsible soils, infiltration rates decline rapidly after the soil is first flooded and the
usual procedures of drying, cleaning, and disking are ineffective in restoring infiltration
rates to original values.
Because low-infiltration recharge systems require more land for infiltration basins,
they offer excellent opportunities for environmental and public benefits. The infiltration
systems could then be designed as attractive aquatic parks with nature, scenic, and recre-

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

Artifical Recharge of Groundwater: Systems, Design, and Management 24.27

FIGURE 24.12 Schematic of meandering supply stream (top)


with rock dams (dots) and outlets (small bars) to basins that are
formed by berms on the contours, and of crossection AA (bottom) with contour berms, infiltration basin upslope from berm,
and survival channels.

ational amenities. If the infiltration area is relatively flat and only mildly sloping, the supply channel could be constructed as a meandering stream (possibly with some wetland
sections) down the slope (Fig. 24.12). Shallow infiltration basins would then be obtained
with low berms on the contours (Fig. 24.12). The berms would be constructed from soil
on the uphill side, leaving a deeper channel that is kept filled with water during drying for
survival of aquatic life, including mosquito fish. Rock piles in the supply channel would
back the water up so that it can flow into the basins. Systems as in Fig. 24.12 require minimum earth moving and give minimum intrusion into the land area. They could be attractively landscaped and equipped with walking and biking trails, picnic areas, and other
facilities for public enjoyment.

24.7 PILOT BASINS AND SYSTEM DESIGN


24.7.1 Test Basins
If the proposed project survives all the preliminary investigations, it is still a big step to
go from small infiltrometers to multihectare recharge systems. Thus, it is advisable to
install a test basin of at least about 50  50 m to see what actual basin infiltration rates
will be, how the water will move down to the groundwater (piezometers above restrict-

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

24.28

Chapter Twenty-Four

ing layers will indicate perched water), what the response of the groundwater level to
recharge will be, and how the observed mounding agrees with predicted mounding using,
for example, Glovers or Hantushs equation. This is a test of hydraulic continuity
between the basin or other surface system and the aquifer. The basins should be operated for a long time (several months to a year or more if possible) to see if clogging occurs,
how it affects infiltration rates, and how it can be remediated by drying and cleaning.
Long-term operation of the pilot basin(s) will also show seasonal differences between the
infiltration rates. Knowing infiltration rates and required drying times during cold or
rainy periods is extremely important where the recharge system must be able to take a
certain flow such as, for example, the entire discharge from a sewage treatment plant.
The recharge area then must be sufficiently large to accommodate that flow during periods of minimum infiltration. The annual hydraulic loading rate is useful in determining
how much water can be put underground per year or other time period if the supply of
recharge water is not limiting.
In general, the more challenging the conditions are for artificial recharge, the more
effort should be made to adequately study the site conditions to make sure that there are
no fatal flaws and that artificial recharge indeed is feasible. This will cost money, but the
justificiation of additional expenses for adequate site investigations is that they will be a
lot less than the costs of installing a recharge system that turns out to be a failure. The
main tools for site investigations are infiltration measurement (infiltrometers and test
basins), testing the vadose zone for restricting layers (backhoe pits or trenches for sampling and visual inspection of soil profile), and borings to obtain samples of vadose zone
and aquifer material and well logs such as geologic, self-potential, resistivity, neutron, and
gamma logs, and testing the aquifer for adequate transmissivity (borings to obtain samples, pumping tests to measure transmissivity, and slug tests to measure hydraulic conductivity; Bouwer, 1978). Hydraulic conductivity values for the vadose zone can be measured with the double-tube or infiltration gradient methods (Bouwer, 1978). If pits or
trenches are dug, in situ hydraulic conductivities of the bottom material can be determined
with cylinder infiltrometers (using Eq. 24.4) or with the air-entry permeameter (Bouwer,
1978). Very stony soil materials defy insitu measurement of hydraulic conductivity with
cylinder techniques. Perhaps the reverse auger hole technique can be used (Bouwer,
1978), or the hydraulic conductivity can be estimated from the volume fraction of the rock
matrix itself and the hydraulic conductivity of the soil material between the rocks as determined in the laboratory on disturbed samples (Bouwer and Rice, 1984b).

24.7.2 Design and Management


Projects should be installed piecemeal, putting in only one or a few basins at first to get further experience in the operation and performance of the system, which can then be used to
fine tune the design and management of the rest of the project. The golden rule in groundwater recharge projects is: start small and simple, learn as you go, and expand as needed.
Individual infiltration basins typically vary in size from about 0.1 ha to about 30 ha.
Each project usually consists of a number of infiltration basins to allow flexibility in
operation and managing each basin (water depth, lengths of flooding and drying periods,
and frequency of cleaning) to maximize infiltration. This requires experienced operators
who know their basins and recognize that each basin has its own personality and
must be treated accordingly while keeping an awareness of the general principles presented in this chapter.
While mostly aimed at off-channel type infiltration basins, the above comments also
apply to in-channel systems. The main differences are that for these systems there often is

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

Artifical Recharge of Groundwater: Systems, Design, and Management 24.29

no control over flooding and drying periods, and mechanical cleaning of the spreading
system may not be possible. For T- or L-dikes, the water is moving throughout the entire
system which could keep suspended solids in suspension and reduce formation of clogging layers, especially in the upstream part of the system.
Example. A recharge system with a basin layout similar to that in Fig. 24.12 is proposed in a new area with no wells and no other groundwater recharges and discharges.
Infiltrometer tests indicated potential infiltration rates of 0.7 m/day, whereas tests with a
pilot basin of 30  30 m indicated an initial infiltration rate of 0.6 m/day which gradually decreased to 0.2 m/day in about 2 months due to clogging. The average infiltration rate
during the 60 days of flooding was 0.4 m/day. Assuming that 30 days of drying would be
necessary to restore infiltration rates and possibly remove the clogging layer after each 60day flooding period would then give a hydraulic loading of 365  60  0.4/(60  30) 
97 m/year. The design recharge rate is 50 million m3 year, which at hydraulic loading rate
of 97 m/year would require 51.5 ha of infiltration basins. These basins will be spread out
over an approximately square area of 83 ha similar to the system in Fig. 24.12, leaving
31.5 ha of dry areas between the basins. The average daily infiltration over the entire 83
ha thus will be 51.5  97/(83  365)  0.17 m/day. The 83 ha gross area and the average infiltration rate of 0.17 m/day over that area will be used for calculating the rise of
groundwater mounds in response to recharge.
The question is: How much water can be stored underground before the mound rises
too high and where should wells eventually be installed to establish a constant, control
groundwater level that will create an equilibrium condition between recharge and pumping to prevent further mound rises? The transmissivity is 930 m2/ day, the ground water
depth is 91 m, the maximum permissible rise of the mound is 82 m, and the fillable
porosity is 0.2. Using Hantushs equation (Eq. 24.7) the mound is calculated to rise to its
maximum level (82 m rise) in 30 years, which will then have stored 1500 million m3 of
water underground. To keep the mound at 9 m below ground surface, a controled water
table at some distance from the recharge area must be established at the original groundwater depth of 82 m below the top of the mound. Replacing the square recharge area by
an equivalent circular area gives a radius R of the circular area of 507 m. Substituting
these values into Eq. (24.9) then gives Rn  10 km. Thus, wells must be installed and
pumped in the vicinity of a circle about 10 km from the center of the recharge system or
closer to keep the groundwater at 10 km from the center at a depth of at least 91 m, so
that the top of the groundwater mound in the recharge area remains at least 9 m below
ground surface.

24.8 SUBSURFACE SYSTEMS


24.8.1 Vadose Zone Wells
Vadose zone wells, also called dry wells or recharge shafts, are boreholes in the vadose
zone, usually about 1050 m deep and about 12 m in diameter (Fig. 24.3). They are
commonly used for infiltration and disposal of storm runoff in areas of relatively low
rainfall that have no storm sewers or combined sewers. Dry wells normally are drilled
into permeable formations in the vadose zone that can accept the runoff water at sufficient rates. Where groundwater is deep (e.q. 100300 m or more) dry wells are much
cheaper than recharge wells, and hence it is tempting to use dry wells for groundwater

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

24.30

Chapter Twenty-Four

recharge instead of aquifer wells. Such vadose zone wells are similar to the recharge pits
or recharge shafts which have also been used for recharge of groundwater. To get adequate recharge, the vadose zone wells should penetrate permeable formations for an
adequate distance.
The main problem with using vadose zone wells for groundwater recharge is, of
course, clogging of the wall of the well and the impossibility of remediating that clogging
by pumping or redeveloping the vadose zone well, since it is in the vadose zone and
groundwater cannot flow into it. Thus, clogging must be prevented or minimized. First,
this can be achieved by protecting the water in the vadose zone well against slaking and
sloughing of clay layers in the vadose zone that could make the water in the well muddy,
causing clay to accumulate and form a clogging layer on the more permeable soil material where most of the infiltration takes place. This slaking can be minimized by filling the
well with sand and using a perforated pipe or screen in the center to apply the water for
recharge. Placing plastic sheets or geotextiles in the well against the zones with clay layers also may be effective. Second, the water must be treated before recharge to remove as
many clogging agents as possible, including suspended solids, assimilable organic carbon,
nutrients, and microorganisms, and it must be disinfected to maintain a residual chlorine
level (see also Sec. 24.8.3). If clogging still occurs (and long-term clogging is always a
possibility), it will then mostly be due to bacterial cells and organic metabolic products
like polymers (biofouling) on the wall of the well. Thus, while such clogging cannot be
remediated by pumping or cleaning or redevelopment, it may be possible that a very long
drying period could produce significant biodegradation of the clogging material to restore
the vadose zone well for another episode of recharge. Superdisinfection of the water
before recharge to push biological activity out into the vadose zone, as discussed in Sec.
21.93, Sec. 21.9.3, may also be effective in preventing clogging of vadose zone wells.
Since recharge with aquifer wells or vadose zone wells is much more expensive
than with surface infiltration systems, rigorous economic analyses are necessary to
come up with the best system. Factors to be considered include the cost of vadose zone
wells versus aquifer wells, their recharge capacities and the number of wells needed,
their useful lives, maintenance and/or replacement costs, and the cost of necessary pretreatment of the water. Contaminated vadose zones can preclude the use of vadose zone
wells.
Recharge rates for vadose zone wells in uniform soil materials can be calculated from
Zangars equation for reverse augerhole flow (Bouwer, 1978), which for a typical vadosezone well geometry with groundwater levels significantly below the bottom of the well
can be simplified to
2 KLw2
Q  
ln (2Lw/rw)  1

(24.10)

where Q is the recharge rate, K is the hydraulic conductivity of the soil material, Lw is the
water depth in the well, and rw is the radius of the well (Fig. 24.3). More research is needed on vadose zone recharge wells to develop an optimum design for well capacity, clogging control (including pretreatment and superdisinfection), useful life, and minimum
long-term cost of recharge per unit volume of water.

24.8.2 Seepage Trenches


Where permeable surface soils are not available but permeable strata occur within
trenchable depth (about 25 m, for example), drilled vadose zone wells may not be nec-

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

Artifical Recharge of Groundwater: Systems, Design, and Management 24.31

essary and seepage trenches may be more cost effective (Fig. 24.3). The trenches are
backfilled with coarse sand or fine gravel, water is applied to the surface of the backfill,
and the trench is covered to keep out sunlight, animals, and people (Hantke, 1983) or
to make the trenches invisible by giving them the same surface condition as the surrounding area. Several variations on this basic design are possible, such as a T-trench
with somewhat finer material in the wider T-layer on top to obtain better removal of
suspended solids, use of geotextiles on or in the backfill to filter the water, plastic
sheets against clay zones in the trench to prevent sloughing of the clay and mud from
entering the trench, and gravel walkways or landscaped covers on top of the trench. As
with vadose zone and aquifer wells the water for seepage trenches must have a very
low suspended solids content. The recharge rate for seepage trenches can be estimated
as about 20 percent of Q calculated with Eq. (24.10) for a vadose-zone well, for a
trench width and length section equal to the diameter of the well (i.e., 2rw) and a trench
water depth equal to the water depth in the well. Thus, if a 1 m diameter dry well 10
m deep and filled with water to the top can do 1000 m3/day, a 10mdeep trench 1 m
wide and full of water can do about 200 m3/day per m length of trench. As with surface infiltration systems, test vadose zone wells or trenches should be installed in new
areas where there is no previous experience with these systems, to see how they perform and how they should be designed and managed for optimum performance in the
full-scale system.

24.8.3 Aquifer Wells


Well recharge is done where permeable surface soils are not available, vadose zones
have restricting layers or contaminated areas, and/or aquifers are confined. As with
trenches and vadose zone wells, the main problem with aquifer wells is clogging around
the well, especially if the aquifer consists of relatively fine, unconsolidated material. To
predict the clogging potential of the recharge water, three main clogging parameters
have been identified (Peters and Castell-Exner, 1993): the membrane filtration index
(MFI), the assimilable organic carbon content (AOC), and the parallel filter index
(PFI). These parameters can also be used for vadose zone wells and trenches (Secs.
24.8.1 and 24.8.2).
The MFI describes the suspended solids content of the water determined as the slope
of the straight portion of the curve in a plot of time/volume versus volume in a membrane
filter test, using, for example, a 0.45 millipore filter. Thus, the dimension of the MFI is
time/volume2; for example, s/L2.
Assimilable organic carbon is determined microbiologically by plating out and incubating a water sample for growth of bacteria of the type Pseudomonas fluorescence,
counting the bacterial colonies, and expressing the results in terms of the carbon concentration of an acetate solution producing the same bacterial growth. AOC may be less than
1 percent of dissolved organic carbon (DOC). AOC levels in the recharge water should be
below 10 g/L to avoid serious clogging of the well if no chlorine is added to the water.
If a residual chlorine level is maintained prior to recharge, higher AOC levels probably can
be tolerated. Rather than using AOC, biodegradable organic carbon (BDOC) may be
preferable as a biological clogging parameter, especially for higher organic carbon concentrations. BDOC is easier to determine than AOC because it is based on degradation of
organic carbon in soil columns or in batch tests with soil slurries.
The PFI is determined by passing the recharge water through columns filled with the
appropriate aquifer material. The flow rates through the columns are maintained much
higher than the infiltration rates per unit area of aquifer around the well. Thus, clogging

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

24.32

Chapter Twenty-Four

occurs faster in the columns than in the well and the PFI serves as an early warning of
things to come for the recharge well so that preventive or remedial action can be taken
early.
Experience has shown that MFI, AOC, and PFI are useful parameters for comparing
relative clogging potentials of the water, but that they cannot be used to predict clogging
and declines in injection rates for actual recharge wells. Thus full-scale studies on
recharge test wells are still necessary to determine feasibility and design and management
criteria for operational recharge wells. Practical aspects such as a varying flow in the
water supply pipes to the recharge project and associated possibility of fluctuating suspended solids contents in the water can also play a major role in well clogging. The suspended solids fluctuations are then caused by formation of biofilms in the pipelines during periods of low flow and erosion of the biofilms during high flow. Treatment of the
water at the recharge site to remove suspended solids prior to well injection may then be
necessary.
Clogging of recharge wells can be controlled by adequate pretreatment of the water,
maintaining a residual chlorine level in the water, periodic pumping of the well (ranging from a few minutes every day to pumping once a month or once a year until the
brown, smelly water is gone and clean water comes out), and well redevelopment or
rehabilitation techniques (surging, jetting, an so on). Periodic pumping is the easiest
(and also often the cheapest!). To enable regular pumping, recharge wells should be
installed with a permanent pump, as with aquifer storage and recovery wells (see Sec.
24.8.4). Increasing injection pressures to overcome clogging effects generally is not
successful and can actually hasten the clogging process by compressing the clogging
layer in the same way as discussed in Sec. 24.4.4. Even if the clogging layer is not compressed by the higher injection pressures and injection rates are indeed increased, the
higher infiltration rates in the well can then increase pore clogging biomass production
by higher loading rates of nutrients and organic carbon. They can also increase physical clogging by higher loading rates of suspended solids. Increased injection rates by
increasing injection pressures then are relatively short lived as injection rates soon
decrease to lower values.

24.8.4 Aquifer Storage and Recovery (ASR) Wells


A new and rapidly spreading practice in artificial recharge is the use of aquifer storage
and recovery (ASR) wells (Pyne, 1995). These wells are combination recharge and pumping wells. They are used for recharge when surplus water is available, and pumped when
the water is needed. ASR wells typically are used for seasonal storage of finished drinking water with a residual chlorine level in areas where water demands are much greater in
summer than in winter, or vice versa, and where surface storage of water is not possible
or too expensive. The winter surplus is then stored underground with ASR wells, which
are pumped in summer (or vice versa) to augment the production from the water treatment
plant. The only treatment of the water pumped from the wells then is chlorination. The
combination of treatment plant capacities based on mean annual demand and ASR wells
to store surplus water and meet peak demands is cheaper than the use of treatment plants
and surface reservoirs with capacities based on peak demands without ASR wells. ASR
wells can also be used to store good quality raw water supplies when they are in surplus
and to pump them up when there is a need for that water. This is of special importance in
parts of Europe and other countries where people demand groundwater but where groundwater levels are depleted in the summer and must be replenished in the winter when there
is more streamflow.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

Artifical Recharge of Groundwater: Systems, Design, and Management 24.33

24.9 ROLE OF RECHARGE IN WATER REUSE


24.9.1 Quality Criteria
Planned water reuse will become increasingly important, not only in water short areas
where sewage effluent is an important water resource, but also where streams or other
surface water (including seawater at popular beaches) need to be protected (Bouwer,
1993). Sewage treatment for planned water reuse may then be cheaper than treatment
for discharge into surface water without causing unacceptable pollution. Planned water
reuse requires treatment of the effluent so that it meets the quality requirements for the
intended reuse. Due to treatment costs, economic feasibility, and aesthetics, the treated
sewage effluent most commonly will be used for nonpotable purposes such as agricultural and urban irrigation, golf courses, sports fields, recreational and decorative lakes,
power plant cooling, industrial processing, construction, dust control, fire protection,
toilet flushing (mostly in commercial buildings but also more and more in private
homes), and environmental purposes (wetlands, riparian habitats, perennial streams,
wildlife refuges). While unplanned or incidental use of sewage effluent for drinking or
public water supplies goes on all over the world, planned reuse for potable purposes is
still rare but can be expected to increase in the future (McEwen and Richardson, 1996;
National Research Council, 1994).
In areas with dry climates where rainfall is insufficient to meet the water requirements
of plants and crops, irrigation is a good use for sewage effluent. The necessary treatment
then is least for crops not consumed by humans (tree, fiber, or seed crops) or crops cooked
before human consumption (grain crops) and not brought raw into the kitchen (Bouwer
and Idelovitch, 1987; Shelef, 1990). This is called restricted irrigation. On the other
hand, unrestricted irrigation, which includes sprinkler irrigation of fruit and vegetable
crops consumed raw by people or brought raw into the kitchen, and urban irrigation of
parks, playgrounds, private yards, and so on, requires more intensive treatment. There are
now two main sets of public health water quality criteria for unrestricted irrigation with
municipal wastewater. One set is for developed-type countries which are technically and
financially capable of high technology treatment and where tourism and export of vegetables and other farm products demand essentially zero-risk water reuse. The other is more
for developing countries that cannot afford expensive treatment, where there is little or no
tourism, agricultural products irrigated with sewage effluent are for local consumption,
and stringent health standards would lead to no treatment at all and the use of raw wastewater for unrestricted irrigation, which, of course, is completely unacceptable.
The standard for developed countries is patterned after Californias Title 22 Effluent
Reuse Standards and calls for treatment of the wastewater to remove essentially all pathogenic organisms (no fecal coliforms, no viruses, no eggs of parasitic worms) and turbidity (to less than 2 nephelometric turbidity units). This can be achieved with conventional
primary and secondary treatment followed by coagulation (sometimes followed by sedimentation), granular media filtration, and chlorination or other disinfection. Where hydrogeological conditions are favorable for groundwater recharge with infiltration basins, the
movement of partially treated wastewater through soils and aquifers also can clean the
wastewater sufficiently so that it can be collected from the aquifer for unrestricted irrigation, as discussed in the next section. The guidelines for unrestricted irrigation in mostly
developing countries, as established by the World Health Organization (WHO) (1989) call
for a maximum fecal coliform concentration of 1000/100 ml and a maximum concentration of helminthic eggs of 1 per liter. This can be achieved by lagooning with sufficient
detention times (e.g., 1 month in warm regions). The lagoon effluent will then also have

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

24.34

Chapter Twenty-Four

greatly reduced concentrations of pathogenic bacteria and viruses, hopefully below minimum infectious doses.
The WHO guidelines are based on public health effects as manifested by documented disease outbreaks (epidemiology) due to sewage irrigation, and on feasilibility
of treatment systems. Case histories of disease outbreaks due to irrigation with poorly
treated wastewater showed that they were mainly caused by intestinal nematodes or
parasitic worms (helminthic eggs such as Ascaris and Trichuris species and hookworm,
where endemic). It was also concluded that presence of pathogenic organisms in the
wastewater does not necessarily mean disease outbreaks, especially if the organisms
are present in sufficiently low concentrations and/or there is local immunity. Because
treatment lagoons take up large land areas, the WHO guidelines are mainly for smaller towns. As towns grow, better treatment systems may be needed. Unlike the WHO
guidelines, the much more stringent California-type standards are based on avoiding
the presence of pathogens in the wastewater altogether, regardless of whether they are
capable of causing disease or not, and the essentially complete elimination of such
pathogens in the treatment process. This is the preferred approach where such treatment is feasible, where the public demands zero or minimum risk, and where municipalities, irrigation districts, and farmers need to protect themselves against lawsuits in
case of disease outbreaks where contaminated agricultural products are implied
(Shelef, 1990). Another factor to consider is whether the crops will be entirely consumed by local people with built-up immunities to certain diseases, or whether the
crops will also be consumed by outsiders (visitors to the region or people in other
regions to which the crops are exported). If the crops are also consumed by outsiders,
the more stringent standards should apply.
In addition to public health considerations, agronomic factors should also be considered and the wastewater should meet the normal quality requirements (salinity, sodium
adsorption ratio, nitrogen, toxic and trace elements, and so on) for irrigation water
(Bouwer and Idelovitch, 1987; Pettygrove and Asano, 1985). The California-type standards also apply to most other nonpotable urban uses of sewage effluent, like urban irrigation (parks, playgrounds, and the like), recreational lakes, toilet flushing, fire protection, car washing, and industrial uses. Since irrigation is mostly a consumptive use of
water (most of it evaporates back to the atmosphere while the rest moves downward
through the soil to underlying groundwater) and requires large quantities of water, it is an
excellent way of reusing water, especially in areas with dry climates where crops and
urban plantings need extra water. However, the effects of such irrigation on underlying
groundwater need to be better understood (Bouwer et al. 1999)

24.9.2 Artificial Recharge and Soil-Aquifer Treatment


Artificial recharge of groundwater has an important role to play in water reuse. Treated
sewage effluent is then infiltrated into the ground for recharge of aquifers. As the effluent
water moves through the soil and the aquifer, it undergoes significant quality improvements
through physical, chemical, and biological processes in the underground environment.
Collectively, these processes and the water quality improvement obtained are called soilaquifer treatment (SAT) or geopurification. Recharge systems for SAT can be designed as
infiltration-recovery systems, where all effluent water is recovered as such from the aquifer
(systems A, B, and C in Fig. 24.13), or after blending with native groundwater (system D).
SAT typically removes essentially all suspended solids, biochemical oxygen demand
(BOD), and pathogens (viruses, bacteria, protozoa, and helminthic eggs). Concentrations
of synthetic organic carbon, phosphorous, and heavy metals are greatly reduced. Nitrogen

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

Artifical Recharge of Groundwater: Systems, Design, and Management 24.35

can be left in the water if needed as fertilizer for irrigation use. The SAT system is then
operated to establish aerobic conditions in the soil that produce nitrification. If nitrogen levels are to be reduced, the SAT system can be managed for nitrification-denitrification by
including anaerobic phases. Nitrogen also can be removed in the pretreatment process. If
surface infiltration systems are used, SAT will slightly increase the salt concentration of the
water due to evaporation from the basins. For example, if the hydraulic loading rate is 100
m/year and the evaporation 2 m/year, the salt increase will be 2 percent.
The performance of SAT systems is site-dependent and controlled by wastewater
quality, soils, hydrogeology, and climate. Thus, in new areas without local SAT experience, pilot or experimental systems should always precede full-scale, operational systems so that the feasibility of SAT can be evaluated and the full-scale system can be
designed and managed for optimum performance. Examples of such experimental and
demonstration projects are the Flushing Meadows Project (Bouwer et al., 1980) and the
23rd Avenue Projects (Bouwer and Rice, 1984a) in the Salt River floodplain west of
Phoenix, Arizona. Both projects used secondary effluent (activated sludge) from
Phoenix. Average quality parameters for the secondary effluent as it infiltrated the soil in
the basins and of the water after SAT are shown in Table 24.2. The metal concentrations
of the effluent in this table are recent values (Pat Wokulich, Water and Wastewater
Department, City of Phoenix, personal. communication, 1990). Recent metal concentrations in the water after SAT were not determined. However, analyses of samples about
20 years ago indicated much higher metal concentrations in the secondary effluent and
removal percentages in the SAT system of 84 percent for zinc, 87 percent for copper, 12

FIGURE 24.13 Schematic of recharge and recovery SAT systems: (A) with
natural drainage of renovated water into stream, lake, or low area, (B) collection
of renovated water by subsurface drain, (C) infiltration areas in two parallel rows
and line of wells midway between, and (D), infiltration areas in center surrounded by a circle of wells.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

24.36

Chapter Twenty-Four

percent for cadmium, and 16 percent for lead (Bouwer et al., 1980). Since the present
metal concentrations in the wastewater, as shown in Table 24.2, are much lower, higher
removal percentages can be expected. The vadose zone and aquifer in the 23rd Avenue
Project consisted mainly of sand and gravel layers and the groundwater table was at a
depth of about 17 m. There were four parallel infiltration basins totaling 16 ha. The well
for pumping wastewater after SAT was located in the center of the basin area and was
perforated from 30 to 55 m. The four basins of the project were operated on a schedule
of 2 weeks flooding followed by 2 weeks drying to obtain nitrification and denitrification in the soil and to allow recovery of infiltration rates between flooding periods.
Infiltration rates during flooding were about 0.5 m/day. Since the basins were dry about
half the time, hydraulic loading rates were about 100 m/year.
The quality parameters of the water after SAT, listed in Table 24.2, indicate that the
water meets the agronomic requirements for crop irrigation and the health standards for
California Title 22 effluent (Bouwer, 1985; Pettygrove and Asano, 1985). Hence, the
water is suitable for unrestricted irrigation (including sprinkler irrigation of fruits and vegetables consumed raw or brought raw into the kitchen, and of parks and playgrounds) and
for unrestricted aquatic recreation (including swimming and fishing).
Where sewage effluent or other low-quality water is used and the recharge systems are
designed and operated as recharge and recovery systems, the optimum combination of

TABLE 24.2 Quality Parameters from Phoenix, Arizona, SAT system for mildly chlorinated
secondary effluent (activated sludge) as it entered the infiltration basins (left column) and
after SAT and pumping it from a well in the center of the infiltration basin area (right column).

Total dissolved solids

Secondary
Effluent
(mg/L)

Recovery Well
Samples
(mg/L)

750

790

Suspended solids

11

Ammonium nitrogen

16

0.1

Nitrate nitrogen

0.5

5.3

Organic nitrogen

1.5

0.1

Phosphate phosphorus

5.5

0.4

Fluoride

1.2

0.7

Boron

0.6

0.6

Biochemical oxygen demand

12

12

1.9

Total organic carbon


Zinc
Copper
Cadmium
Lead

0.036
0.008
0.0001
0.002

Fecal coliforms, per 100 ml

3500

0.3

Viruses, PFU/100 1

2118

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

Artifical Recharge of Groundwater: Systems, Design, and Management 24.37

pretreatment, soil aquifer treatment (SAT) and postreatment must be selected. The usual
pretreatment in the United States then is primary and secondary treatment and chlorination because that is what is typically necessary to comply with local treatment and discharge regulations. However, primary treatment may be sufficient and even advantageous
(Carlson et al., 1982; Leach et al., 1980; Rice and Bouwer, 1984), not only because it
saves money, but also because the higher total organic carbon (TOC) level of primary
effluent may actually remove more recalcitrant or refractory TOC in the SAT systems by
secondary utilization and cometabolism (McCarty et al., 1984) and more nitrogen by denitrification (Lance et al., 1980). Primary effluent may produce more clogging of the basin
bottoms because of the higher suspended solids content, and also because of the higher
dissolved organics content which can aggravate biological clogging. Where the resulting
increased need for cleaning the basins is a problem, the primary effluent could be filtered
first through a rapid sand or similar filter, using a flocculent if desirable.
Disinfection will be most effective after SAT when the suspended solids concentrations or turbidity are very low and most pathogens have already been removed by SAT.
Ultraviolet irradiation may then be the best disinfection technique to minimize formation
of disinfection by-products (DBPs). Some mild chlorination prior to infiltration may be
desirable so that the rest of the pathogens can be removed by SAT and postdisinfection is
not necessary if the water is used for nonpotable purposes. In general, pretreatment of the
sewage effluent should be matched to the benefits from SAT so that posttreatment other
than disinfection will not be necessary to meet the quality requirements of the intended
use of the water. If the system is designed and managed as a complete recharge and recovery system, posttreatment is a viable option that can reduce pretreatment and maximize
SAT benefits. However, there may be conflicts between optimum pretreatment for SAT
and the pretreatment necessary to comply with local regulations. Also, popular demand
may call for best available demonstrated control technology (BADCT)-type pretreatment
procedures at least in the United States.
Nitrogen removal in SAT systems is achieved by biological denitrification of nitrate,
which occurs under anaerobic conditions and requires organic carbon as an energy source
for the denitrifiers (about 2.5 mg/L BOD or 0.81.3 mg/L biodegradable organic carbon
for each mg/l nitrate nitrogen to be denitrified). For conventional secondary effluent,
nitrogen concentrations in the United States with an in-house water use of about 350 L per
person per day will be on the order of 15 to 20 mg/L, mostly as ammonium. Short, frequent flooding and drying of the basins (a few days each, for example) will then create
mostly aerobic conditions in the soil and almost all the ammonium in the effluent will be
converted to nitrate in the water moving down to the groundwater, with little nitrogen
removal in the SAT system. Where in-house water use is less, effluent nitrogen concentrations can be expected to be proportionally higher.
To get denitrification in the recharge system, flooding periods must be long enough to
get anaerobic conditions in the soil so that the ammonium is adsorbed to the clay particles
or organic matter. This process requires soils like loamy sands or sandy loams, or even
finer, with some clay so that the soils have some cation exchange capacity. Before the clay
particles are saturated with ammonium, flooding is stopped and the basin is dried.
Atmospheric oxygen will then enter the soil to create aerobic conditions for nitrification
of adsorbed ammonium. Part of the resulting nitrate will then be denitrified in microanaerobic zones still present in the aerobic zone. The remaining nitrate will be leached out
when flooding is resumed, causing a nitrate spike in the water moving down to the
groundwater. In the United States, these spikes can exceed 100 mg/L NO3-N but they are
attenuated with distance of travel and mixing in the aquifer, so that they are not found in
water from recovery wells outside the recharge area. Typical schedules for enhancing denitrification may be 10 days flooding alternated with 10 days drying in the summer and 20

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

24.38

Chapter Twenty-Four

days drying in the winter. Such schedules have yielded 50 to 70 percent removal of nitrogen (Bouwer and Rice, 1984a). Most of the nitrogen transformations take place in the
upper 1 m of the soil profile. Recently, a new process for nitrogen removal has been discovered where by ammonium and nitrate react biochemically to produce free nitrogen gas
while leaving some residual nitrate nitrogen at a concentration of about 10% of the total
ammonium and nitrate nitrogen input concentration. The process is autotrophic (requires
no organic carbon) and anaerobic so that it can occur in aquifers using carbonate as the
carbon source. It is called the anammox process for anaerobic ammonium oxidation (Van
de Graaf et., 1995)
If secondary effluent received additional aeration in the treatment plant or in an aerated lagoon, part of the ammonium is already converted to nitrate and BOD is reduced.
Denitrification may still occur in the SAT system when flooding periods are long enough
to create anaerobic conditions in the soil, but nitrogen removal may be less than for the
previous case because of insufficient BOD. In such cases, it may be preferable to nitrify
and denitrify in the treatment plant, which creates an effluent with sufficiently low nitrogen concentrations (mostly as nitrate) and low BOD levels. Extended flooding periods to
create anaerobic conditions in the soil may then further reduce the nitrate nitrogen, but
not by much (perhaps by a few mg/L at most, depending on the BOD level of the effluent and synthesis of organic carbon in the SAT system itself). More research is needed
to see how much organic carbon is formed in the SAT system by biological processes
(algae, microorganisms, autotrophic bacteria like nitrifiers, and so on). Also, more
research is needed on sustainability of SAT and associated pretreatment and/or posttreatment requirements.
A very important aspect of water reuse via artificial recharge of groundwater is that it
also greatly enhances the aesthetics and public acceptance of water reuse, because the water
has had SAT and it comes out of wells rather than out of the end of advanced wastewater
treatment plants where water is recycled after in-plant treatment only and put directly back
in the water distribution system (direct or pipe-to-pipe recycling).

24.9.3 Well Recharge with Sewage Effluent


Where sewage effluent is used for recharge with wells or other underground systems
(trenches or vadose zone wells), it must undergo extensive pretreatment, including
advanced wastewater treatment (AWT) processes. First, this is necessary to achieve
essentially drinking water quality standards before recharge because aquifers generally
are considered too coarse for significant aquifer treatment. The only treatment that
can be expected in aquifers is some additional TOC removal, removal of microorganisms, improvement in taste and odor, and similar aging and polishing effects.
Second nitrogen removal via the anamox process, AWT is necessary to minimize clogging in the well by removing the clogging parameters. Where the AWT includes RO,
proper disposal of the reject water or brine can be expensive. Also, water after RO will
have a very low TDS concentration which makes it corrosive, aggressive, and hungry.
The interaction between this water and the receiving aquifer must then be well understood to make sure that the hungry water does not mobilize chemicals from minerals
and other solid phases of the aquifer that are not wanted in the final product water.
Blending of the water after RO with water of a higher TDS content before recharge may
then be necessary to stabilize it and avoid unpleasant surprises when it is pumped up
again.
Perhaps the expensive RO or other membrane filtration of sewage effluent or similar
low quality water can be avoided by giving the effluent a relative high dose of chlorine or

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

Artifical Recharge of Groundwater: Systems, Design, and Management 24.39

other disinfectant prior to recharge. This may push biomass development farther into the
aquifer. The biomass would then be created at some distance from the well where the disinfectant has dissipated, so that biomass is formed over a larger area than right around the
well. It is hoped that the biomass would then be in the form of biofilms on the solid aquifer
matrix rather than as the pore plugging biomass formed by biological activity right around
the well. This biofilm zone could then actually perform as an in situ bioreactor that could
remove organic compounds (including disinfection by-products), nitrate, and other contaminants from the water. Chlorination before recharge has always been considered problematic because of the potential for formation of disinfection byproducts (DBPs).
However, recent work by Singer et al. (1993) has shown that DBPs like trihalomethanes
and haloacetic acids (THMs and HAAs) can be biodegraded in the aquifer, so that superchlorination or other disinfection (with chloramines or peroxide, for example) may not be
objectionable. Research is necessary to see what disinfection levels are needed in relation
to TOC of the water, how far biomass development has to be pushed out into the aquifer
to get a good bioreactor, and how the bioreactor will perform in terms of quality improvement of the water and reduction in local hydraulic conductivity so that water can still
move through it. However, if this super disinfection can be made to work it will be an
important contribution to using sewage effluent or other low quality water for recharge
with aquifer wells, vadose zone wells, or seepage trenches, because it will eliminate the
need for expensive RO or other membrane filtration with associated brine disposal as pretreatment to remove organic carbon. Conventional primary and secondary treatment followed by nitrification-denitrification, sand filtration, and possibly activated carbon filtration may then be sufficient pretreatment.

24.9.4 Constructed Aquifers


Constructed aquifers are basically sand filters for sewage or other water of impaired quality. They are constructed by excavating a pit, lining it with plastic, putting a gravel layer
with perforated drain pipe on the bottom, and filling it with about 2 m of sand or other
permeable soil (Fig. 24.14). They are flooded with sewage effluent for short periods (a few
days at most) and dried long enough for the clogging layer to dry, crack, and curl up so
that it can be raked off. The outflow from the drain is used for discharge into surface water
or for irrigation or other nonpotable use. While the drainage water does not meet
Californias Title 22 requirements for unrestricted irrigation, it can meet the less stringent
guidelines of the WHO for unrestricted irrigation (Bouwer, 1993). Because of their simplicity, constructed aquifers are increasingly used in the Third World where sewage often
is extremely strong and pretreatment usually is of the primary type (mostly septic tanks or
overloaded lagoons). Since the suspended solids content of these effluents is still high, the
clogging layer on the sand filter develops rapidly and may have to be removed after every
drying period. This is normally done manually by raking. Flooding periods may then only
be 1 or 2 days, followed by drying periods of 510 days, depending on climate. The systems thus are labor intensive. However, this is usually not a problem with the rural poor
where unemployment often is high and where simple, inexpensive, and labor-intensive
systems are preferable because they are dependable and they create jobs! Constructed
aquifers are basically poor mans intermittent sand filters (no back flushing). The advantage is that they give more protection against infectious diseases than lagooning alone.
They also are used for small villages, hotels, resorts, and similar small developments,
especially if the lagoons are overloaded. More than 100 systems have been installed in
France as of 1991 (Agences de l Eau et du Ministere de l Environnement, 1994) and
there is a research and demonstration project in Morocco, south of Agadir.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

24.40

Chapter Twenty-Four

FIGURE 24.14 Constructed aquifer or intermittent sand filter lined with plastic and with
perforated drain in gravel layer on bottom.

24.9.5 Potable Reuse after Soil Aquifer Treatment


The main concern in potable use of sewage water after recharge, SAT, and blending with
native groundwater is with the trace organic compounds that are still present in the water,
albeit in very small concentrations [much less than 1 g/L or parts per billion (ppb), for
example] that may well be below threshold values where they become toxic. Other constituents like nitrate, pathogens, and heavy metals are no problem because they are
removed in the pretreatment and SAT, and the water most likely will be disinfected again
before potable use.
Most natural organic carbon compounds (carbohydrates, proteins, fats, and so forth) in
sewage effluent are biodegradable and, hence, are decomposed in the soil. In addition,
however, there is a broad spectrum of synthetic organic compounds that enter the water as
chemicals in industrial wastewater, hospital wastes, storm runoff, and domestic wastewater because of use of household chemicals and pharmaceuticals. After treatment in the
water reclamation plant, the concentrations of these chemicals often are on the order of 1
part per billion (Bouwer and Rice, 1984a). A large number of these chemicals also are
biodegradable, especially the nonhalogenated compounds (E. J. Bouwer et al., 1984) and
also certain disinfection byproducts or DBPs (Singer et al., 1993). Some resistant organic compounds (also known as nonbiodegradable, stable, refractory, or recalcitrant compounds) may accumulate in the adsorbed state in the soil and aquifer and eventually may
move with the water when the adsorption sites are saturated with these chemicals. This
causes the transport of these chemicals in the underground environment to be retarded relative to the velocity of the water itself (Bouwer, 1991). Retardation factors may vary from
as low as 2 or less to as much as 100 or more, depending on adsorption of the chemicals.
Some of the resistant compounds eventually may also be biodegraded in the adsorbed or
mobile state through various bacterial processes, including acclimatization, secondary utilization, and cometabolism (McCarty et al., 1984).
Maximum concentration limits for organic and other chemicals in drinking water normally are based on rodent bioassays. The resulting values, however, are more and more
considered to be very much on the conservative side. In addition, their validity is increasingly questioned because the linear response theory used to extrapolate responses from
unrealistically high doses in the tests to low doses ingested by people does not recognize
threshold values (Ames and Swirsky Gold, 1990). Hence, many compounds now classified as carcinogenic actually may not cause cancer at the concentrations normally ingested. Also, rodent bioassays fail to recognize genetic effects. Some chemicals have produced cancers in mice but not in rats (Calabrese, 1987; Lave et al., 1988). So if rodent
bioassay results cannot be transferred from one rodent to another, how can they be transferred to people? Also, rodent bioassays do not take into account physiological effects
such as different metabolic pathways for high doses than for low doses. This has caused
toxic metabolic endproducts to be produced at high doses but not at low doses, showing
again that it is the dose that makes the poison. Thus, very small concentrations of syn-

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

Artifical Recharge of Groundwater: Systems, Design, and Management 24.41

thetic organic compounds in reclaimed water after SAT and mixing with native groundwater by macrodispersion (Bouwer, 1991) and other processes may not be a cause for
alarm in the potable use of such water, just as they are not considered a problem in conventionally treated drinking water obtained from rivers or other surface waters that
received sewage effluent. Concerns are rising over pharmaceutically active chemicals
(PACHs) in municipal wastewater, which along with pesticides and other chemicals may
have caused deformed frogs in sewage contaminated surface water and feminization of
males resulting in reduced reproduction and population, possibly because of hormone
effects (endocrine disruptors) (Bouwer et al., 1999). Epidemiological studies in areas
where people drink water from aquifers recharged with sewage effluents have failed to
provide evidence of adverse effects on public health (Nellor et al., 1984; Sloss et al.,
1996). Hence, the endorsement of indirect potable reuse of municipal wastewater by the
National Research Council (1994) and the American Water Works Association (McEwen
and Richardson, 1996). The key issue here is indirect reuse. Direct recycling with pipeto-pipe connections between the sewage treatment plant and the water supply system is
not permitted. Rather, the water after treatment must go through surface water (streams,
lakes, reservoirs) or through groundwater (aquifers), or both, before it can be distributed
for potable use. Going through surface water has several disadvantages like evaporation
losses, vulnerability to secondary contamination by animals and human activities, and
growth of algae which gives the water a bad taste and creates metabolic products (THMprecursors) which on chlorination can form trihalomethanes. These disadvantages do not
occur with groundwater recharge, which gives the additional benefit of SAT, storage, and
enhanced aesthetics and public acceptance of potable water reuse. Planned water reuse
basically compresses the hydrologic cycle from a noncontrolled global scale to a controlled local scale.

24.9.6 Integrated Water Management


Artificial recharge and water reuse are important aspects of integrated water management
where water resources management problems are solved by considering all aspects, using
a holistic or integrated approach. Integrated water management includes not just quantity
and quality aspects, and supply management, but also demand management, water conservation, reuse and recycling, artificial recharge of groundwater, conjunctive use of surface water and groundwater, economics, transfers of water rights and water marketing
with willing buyers and sellers while protecting third-party interests, environmental and
ecological aspects, sociocultural aspects, sustainability, regional solutions (often on a
watershed basis), community involvement (public meetings with all stakeholders, and so
on), more storage of water, multiple-purpose projects, desalination (ocean water, brackish
groundwater), and weather modification.
Integrated water management is readily achieved in countries where the water belongs
to the state (i.e., the people). Where water rights are in private hands, integrated water
management and equitable distribution of water resources hopefully can be achieved by
water marketing, where the water would essentially go to the highest bidder. This would
stimulate the most economical use of water resources. However, some government intervention may be needed to ensure equitable distribution of water and protection of
thirdparty interests. If necessary, water from private owners could be obtained through
the process of eminent domain, as is now done for land. An example of reforming water
law to achieve integrated water management and equitable distribution of water resources
is the White Paper developed in South Africa (Hayward, 1997).

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

24.42

Chapter Twenty-Four

24.10 REFERENCES
Agences de lEau et du Ministere de lEnvironnement, Epuration des Eaux Uses Urbaines par
Infiltration-Percolation: Etat de lArt et Etudes de Cas, Laboratoire de Hydrologie et
Modelisation, Universit Montpellier II, Montpellier-Cedex, France, 1994.
Ames, B. N., and L. Swirsky Gold, Too Many Rodent Carcinogens: Mitogenesis Increases
Mutagenesis, Science 249:940971. 1990.
Bouwer, H. , Analyzing groundwater mounds by resistance network, Journal Irrigation and
Drainage Division, Am. Soc. Civil Enginners 88 (IR 3): 15-36, 1962.
Bouwer, H., Theory of Seepage from Open Channels, in V. T. Chow, ed., Advances in
Hydroscience, Academic Press, New York, p. 121170.1969.
Bouwer, H., Predicting Reduction in Water Losses from Open Channels by Phreatophyte Control,
Water Resources Research, 11:96101, 1975.
Bouwer, H. , Groundwater Hydrology, McGraw-Hill, New York, New York, USA, 1978.
Bouwer, H., Design Considerations for Earth Linings for Seepage Control, Ground Water,
20(5):531537, 1982.
Bouwer, H. , Groundwater Recharge as a Treatment of Sewage Effluent for Unrestricted Irrigation
in A. Arar and M. D. Pescod, eds., Proceedings. FAO Regional Seminar on the Treatment and
Use of Sewage Effluent for Irrigation, Nicosia, Butterworth, London. pp. 116128.1985.
Bouwer, H. , Intake rate: Cylinder in filtrometer. In Methods of Soil Analyses, Part 1, Physical
and Mineralogical Methods, A. Klute, ed. Agronomy Monograph, 2nd. ed. P. 825-844, 1986.
Bouwer, H., Effect of Water Depth and Groundwater Table on Infiltration from Recharge Basins,
In S. C. Harris, ed., Proceedings 1990 National Conference Irrigation and Drainage Division
American Society of Civil Engineers, Durango, CO, July 1113, pp. 337384.1990.
Bouwer, H., Simple Derivation of the Retardation Equation and Application to Preferential Flow
and Macrodispersion, Ground Water Journal, 29(1):4146, 1991.
Bouwer, H., From Sewage Farm to Zero Discharge, European Water Pollution Control,
3(1):916, 1993.
Bouwer, H., Estimating the Ability of the Vadose Zone to Transmit Liquids, In L. G. Wilson, L.
G. Everett, and S. J. Cullen, eds., Handbook of Vadose Zone Characterization & Monitoring,
Lewis Publishers, Boca Raton, FL, pp. 177188.1995.
Bouwer, H. Discussion of Bouwer and Rice Slug Test Review Articles, Ground Water, 34:171,
1996.
Bouwer, H. and E. Idelovitch, Quality Requirements for Irrigation with Sewage Effluent, Journal
of Irrigation Drainage Division, American Society of Civil Engineers.113(4):516535, 1987.
Bouwer, H., and T. Maddock III, Making Sense of the Interaction Between Groundwater and
Streamflow: Lessons for Watermasters and Adjudicators. RIVERS. 6(1):1931, 1997).
Bouwer, H., and R. C. Rice, Renovation of Wastewater at the 23rd Avenue Rapid-Infiltration
Project, Phoenix, Arizona, Journal Water Pollution Control Federation 56(1):7683. 1984a.
Bouwer, H., and R. C. Rice, Hydraulic Properties of Stony Vadose Zones, Ground Water,
22(6):696705, 1984b.
Bouwer, H., and R. C. Rice, Effect of Water Depth in Groundwater Recharge Basins on
Infiltration Rate, Journal of Irrigation and Drainage Division, American Society of Civil
Engineers, 115(4):556568, 1989.
Bouwer, H. , R. C. Rice, and E. D. Escarcega, High-rate Land Treatment: I. Infiltration and
Hydraulic Aspects of the Flushing Meadows Project, Journal Water Pollution Control
Federation 46 (5): 835-843, 1974.
Bouwer, H., R. C. Rice, J. C. Lance, and R. G Gilbert, Rapid Infiltration Researchthe Flushing
Meadows Project, Arizona, Journal of Water Pollution Control Federation, 42(10):24572470,
1980.
Bouwer, E. J., P. L. McCarty, H. Bouwer, and R. C. Rice, Organic Contaminant Behavior During
Rapid Infiltration of Secondary Wastewater at the Phoenix 23rd Avenue Project, Water Research,
18:463472, 1984.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

Artifical Recharge of Groundwater: Systems, Design, and Management 24.43


Bouwer, H. , P. Fox, P. Westerhoff, and J. E. Drewes, Integrating Water Managment and Reuse:
Causes for Concern? Water Quality International, January - February 1999, pp. 19-22.
Bouwer, H., J. T. Back, and J. M. Oliver, Predicting Infiltration and Ground Water Mounding for
Artificial Recharge, Journal of Hydrologic Engineering, Am. Soc. Civil Engrs. In press, 1999.
Calabrese, E. M, Animal Extrapolationa Look Inside the Toxicologists Black Box, Journal of
Environmental Science And Technology, 21:612623 1987.
Carlson, R. D., K. D. Lindstedt, E. R. Bennett, and R. B. Hartman, Rapid Infiltration Treatment of
Primary and Secondary Effluents, Journal of Water Pollution Control Fededration, 54:270280.
1982.
Devine, R. S. The Trouble with Dams, Atlantic Monthly, August 1995, p. 6773. 1995.
Glover, Robert E. Ground Water Movement, U. S. Bureu of Reclamation, Engr. Monogr. 31, 67
pp, 1964.
Hantke, H., Der Sickerschlitzgraben, Brunnenbau, Bau von Wasserwerken, und Rohrleitungsbau
(BBR), 34(6):207208, 1983.
Hantush, Mahdi S. Growth and Decay of Ground Water mounds in Response to Uniform
Percolation, Water Resources Research 3: 227-234, 1967.
Hayward, K. South Africa Sets Out Policy on Water Management, Water Quality International,
p. 4 May/June 1997.
Knoppers, R., and W. van Hulst., De Keerzijde van de Dam, Jan van Arkel, Utrecht, The
Netherlands, 1995.
Lance, J. C., R. C. Rice, and R. G. Gilbert, Renovation of Sewage Water by Soil Columns
Flooded with Primary Effluent, Journal Water Pollution Control Federation, 52(2):381388,
1980.
Lave, L. B., F. K. Ennever, H. S. Rosenkranz, and G. S. Omenn, Information Value of the Rodent
Bioassay, Nature, 336 (Dec.):631633, 1988.
Leach, L. E., C. G. Enfield, and C. C. Harlin, Jr., Summary of Long-Term Rapid Infiltration System
Studies, Report No. EPA-600/2080-165, U.S. Environmental Protection Agency, Ada, OK, 1980.
Marino, Miguel A. Artificial Ground Water Recharge, I. Circular Recharging Area, Journal
Hydrology 25: 201-208, 1975a.
Marino, Miguel A. Artificial Ground Water Recharge, II. Rectangular Recharging Area, Journal
Hydrology 26: 29-37, 1975b.
McCarty, P. L., B. E. Rittman, and E. J. Bouwer, Microbiological Processes Affecting Chemical
Transformations in Groundwater, in G. Bitton and C. P. Gerba, eds.,Groundwater Pollution
Microbiology, John Wiley & Sons, New York, pp. 89116.1984.
McEwen, B., and T. Richardson, Indirect Potable Reuse: Committee Report Proceedings 1996
Water Reuse Conference, American Water Works Association and Water Environment Federation,
San Diego, CA, pp. 486503 1996.
National Research Council, Ground Water Recharge Using Waters of Impaired Quality, National
Academy Press, Washington, DC, 1994.
Nellor, M. H., R. B. Baird, and J. R. Smith, Summary of Health Effects Study: Final Report,
County Sanitation Districts of Los Angeles County, Whittier, CA, 1984.
Oster, J. D., Future Challenges of Irrigated Agriculture Using Poor Quality Water, Arabian
Journal for Science and Engineering, 22(IC):175198, 1997.
Pearce, F., The Dammed, The Bodley Head, London, 1992.
Peters, J. H., and C. Castell-Exner, eds., Proceedings Dutch-German Workshop on Artificial
Recharge of Groundwater, September, 1993, Castricum, The Netherlands, KIWA, Nieuwegein
3430 BB, The Netherlands, 1993.
Pettygrove, G. S., and T. Asano, eds. Irrigation with Reclaimed Municipal WastewaterA
Guidance Manual, Lewis Publishers, Chelsea, MI, 1985.
Pyne, R. D. G. Groundwater Recharge and Wells: A Guide to Aquifer Storage Recovery, Lewis
Publishers, Boca Raton, FL, 1995.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

ARTIFICIAL RECHARGE OF GROUNDWATER: SYSTEMS, DESIGN, AND MANAGEMENT

24.44

Chapter Twenty-Four

Rice, R. C, and H. Bouwer, Soil-Aquifer Treatment Using Primary Effluent. Journal of Water
Pollution Control Federation, 56(1):848, 1984.
Shelef, G. The Role of Wastewater Reuse in Water Resources Management in Israel, in
Proceedings 15th Biennial Conference, International Association on Water Pollution Research
Control, Kyoto, Japan, Water Scienc and Technology, 23:20812089. 1990.
Singer, P. C., R. D. G. Pyne, M. AVS, C. T. Miller, and C. Mojoonnier, Examining the Impact of
Aquifer Storage and Recovery on DBPs, Journal American Water Works Association, 8594.
1993.
Sloss, E. M., S. A. Geschwind, D. F. McCaffrey, and B. R. Ritz, Groundwater Recharge with
Reclaimed Water: An Epidemiologic Assessment in Los Angeles County, 1987-1991, RAND,
Corp. Santa Monica, CA, 1996.
Sumner, M. E, and B. A. Stewart, eds., Soil Crusting: Chemical and Physical Processes, Lewis
Publishers, Boca Raton, FL, 1992.
Van de Graff, A. A. , A. Mulder, P. de Bruyn, M. S. M. Jetten, L. A. Robertson, and J. G. Kuenen.
Anaerobic Oxidation of Ammonium is a Biologically Meditated Process, applied and
Enviromental Microbiology 61 (4): 1246-1251, 1995.
Warner, J. W., D. Molden, M. Chehata, and D. K. Sunada. Mathematical Analysis of Artificial
Recharge from Basins, Water Resources Bulletin, 25:401411, 1989.
World Health Organization, Health Guidelines for the Use of Wastewater in Agriculture and
Aquaculture, Technical Bulletin Series 77, WHO, Geneva, 1989.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

You might also like