Sie sind auf Seite 1von 20

QUALITY REQUIREMENTS FOR IRRIGATION WITH SEWAGE

WATER
By Herman Bouwer1 and Emanuel Idelovitch,2 Members, ASCE

ABSTRACT: Irrigation is an excellent use for sewage effluent because it is


mostly water with nutrients. For small flows, the effluent can be used on
Downloaded from ascelibrary.org by Selcuk Universitesi on 01/31/15. Copyright ASCE. For personal use only; all rights reserved.

special, well-supervised "sewage farms," where forage, fiber, or seed


crops are grown that can be irrigated with standard primary or second-
ary effluent. Large-scale use of the effluent requires special treatment so
that it meets the public health, agronomic, and aesthetic requirements
for unrestricted use (no adverse effects on crops, soils, humans, and
animals). Crops in the unrestricted-use category include those that are
consumed raw or brought raw into the kitchen. Most state or govern-
ment standards deal only with public health aspects, and prescribe the
treatment processes or the quality parameters that the effluent must
meet before it can be used to irrigate a certain category of crops.
However, agronomic aspects related to crops and soils must also be
taken into account. Quality parameters to be considered include bacte-
ria, viruses, and other pathogens; total salt content and sodium adsorp-
tion ratio of the water (soil as well as crop effects); nitrogen; phospho-
rus; chloride and chlorine; bicarbonate; heavy metals, boron, and other
trace elements; pH; and synthetic organics (including pesticides).

INTRODUCTION

The use of sewage effluents for irrigation of agricultural crops is an


attractive and popular wastewater reuse option for the following reasons:

1. Where crops need to be irrigated, water tends to be scarce, and


wastewater can supplement available freshwater resources.
2. Irrigated agriculture requires large amounts of water which are used
only once, since irrigation basically is a consumptive use; consequently,
irrigation water requirements represent a major portion of the total water
demand in dry areas.
3. Agriculture can beneficially use not only the water, but also, within
certain limitations, the additional resources found in wastewater, such as
organic matter, nitrogen, phosphorus, potassium, minor elements, and
other nutrients.
4. Irrigation is relatively flexible with respect to water-quality require-
ments: Some crops may be irrigated with low-quality water without major
risks, and some water quality problems can be overcome by suitable
agronomic practices.
'Dir., U.S. Water Conservation Lab., Agric. Res. Service, USDA, 4331 East
Broadway, Phoenix, AZ 85040.
2
Sr. Sanitary Engr., Water Supply Div., World Bank, 1818 8th St. NW,
Washington, DC 20433; formerly, Head, Wastewater Reuse and Water Quality
Div., TAHAL Consulting Engrs., Ltd., Tel Aviv 64-364, Israel.
Note. Discussion open until April 1,1988. To extend the closing date one month,
a written request must be filed with the ASCE Manager of Journals. The manuscript
for this paper was submitted for review and possible publication on September 30,
1985. This paper is part of the Journal of Irrigation and Drainage Engineering, Vol.
113, No. 4, November, 1987. ©ASCE, ISSN 0733-9437/87/0004-0516/$01.00. Paper
No. 21962.

516

J. Irrig. Drain Eng. 1987.113:516-535.


RESTRICTED VERSUS UNRESTRICTED IRRIGATION

A distinction should be made between two types of irrigation using


municipal effluent: restricted and unrestricted irrigation. The concept of
restricted irrigation refers to the use of a low-quality effluent only in
specific agricultural areas and for specific crops. Restrictions imposed in
connection with this reuse category are related not only to the type of
Downloaded from ascelibrary.org by Selcuk Universitesi on 01/31/15. Copyright ASCE. For personal use only; all rights reserved.

crops to be cultivated, but also to the type of soils to be irrigated, the


location of irrigated fields with respect to potable aquifers, irrigation
methods, crop-harvesting techniques, fertilizer application rates, distance
of irrigated fields to roads and houses, distance between pipelines carrying
potable and nonpotable water, etc.
The "restricted irrigation" concept has the advantages of simplicity and
low treatment cost, but it is generally applicable only to small amounts of
wastewater to be used in specific locations, where areas to be irrigated and
crops to be cultivated are well-defined and unlikely to change. Restricted
irrigation of large areas is technically feasible, but problems increase
dramatically as the size of the system increases. However, a few large
systems (notably at Melbourne, Australia) have also been constructed.
Farmers are, in general, unwilling to accept a low-quality effluent in equal
exchange for fresh water. Also, farmers and farm personnel will have to be
trained to properly handle the effluent (minimum contact, washing hands
before eating, etc.). Irrigation techniques should be selected that minimize
contact between irrigators and wastewater. Aerosols can result in trans-
port of microorganisms to a considerable distance from the field, especially
if the effluent is applied with sprinklers. Accidental ingestion of the
low-quality water and the accidental (or deliberate!) use of the water to
irrigate vegetables or other crops brought raw into the kitchen represent
additional health hazards.
The concept of unrestricted irrigation refers to the use of a high-quality
effluent for irrigation of all crops on all types of soil in any area during a
prolonged period of time, without adverse effects on crops, soils, animals,
people involved in the various stages of the agricultural production
process, and consumers. The water for unrestricted irrigation should be
suitable for livestock watering, too. High-quality effluents which are
suitable for unrestricted agricultural use should be acceptable to the
farmers in exchange for fresh water allocations; thus, they can be traded
for other sources of water supply. Contact and even accidental drinking of
such water should not pose major health risks.

WATER QUALITY STANDARDS FOR IRRIGATION WITH EFFLUENT

Water quality standards established for crop irrigation with fresh water
are, at present, also the best available criteria for effluent reuse (2, 18).
However, there are additional constituents in wastewater, which are
usually absent from or unimportant in fresh water. For such constituents,
specific reuse standards will have to be developed in the future. At present,
only preliminary guidelines can be established, based on available knowl-
edge. Further research is required in order to establish rational water
quality standards for irrigation with sewage effluent. The question of
synthetic organic compounds, for example, should be more fully explored.
517

J. Irrig. Drain Eng. 1987.113:516-535.


TABLE 1. Standards for Irrigation with Reclaimed Wastewater in Arizona (1)
Crop and Land-Use Category
Characteristic A B C D E F G H
(1) (2) (3) (4) (5) (6) (7) (8) 0)
PH 4.5-9 4.5-9 4.5-9 6.5-9 4.5-9 4.5-9 4.5-9 4.5-9
Fecal coliforms (CFU/100 ml):
Downloaded from ascelibrary.org by Selcuk Universitesi on 01/31/15. Copyright ASCE. For personal use only; all rights reserved.

Geometric mean 1,000 1,000 1,000 1,000 1,000 200 25 2.2


(5-sample minimum)
Single sample not to exceed 4,000 4,000 4,000 4,000 2,500 1,000 75 25
Turbidity (NTU) 5 1
Enteric virus (PFU/40 1) 125 1
Entamoeba hystolytica N.D.
Ascaris lumbricoides (roundworm eggs) N.D. N.D.
Common large tapeworm N.D. N.D.
Note: CFU = colony-forming unit; NTU•= nephelometer turbidity units; PFU = plaque-forming
units', N.D. = none detectable, using: correct samples and methods, and qualified personnel. The
crop and land-use categories are:
A. Orchards.
B. Fiber, seed, and forage crops.
C. Pastures.
D. Livestock watering.
E. Processed food crops.
F. Landscaped areas, restricted access.
G. Landscaped areas, open access.
H. Crops to be consumed raw.

In areas where irrigation with effluent is practiced, standards or recom-


mendations for effluent quality are usually available. In general, these
standards are prepared by the health authorities, and they take into
account only the public health aspects of effluent quality. They do not
include agronomic considerations related to crops or soils, or aesthetic
aspects. Some standards specify the treatment processes for the sewage
effluents. Others specify quality parameters and leave the treatment up to
the city or other entity preparing the effluent for irrigation use. Such
standards or guidelines for irrigation with effluent are available in several
states in the United States—e.g., Arizona, California, Nevada, Texas,
Utah (7)—as well as in Israel (13, 14).
In 1983, the Arizona Department of Health Services established a
revised set of reuse standards for irrigation with effluent (1). In accordance
with these standards, maximum limits were determined for five crop
categories, as well as for livestock watering (Table 1). The standards also
establish the monitoring frequency required for some water quality param-
eters (Table 2). Maximum limits for trace substances were also specified.
These limits are about the same as those for alkaline soils shown later in
this paper. Suggested treatments to achieve the standards in Table 1 are
secondary (biological) treatment and disinfection for categories A-F.
Filtration may be required for category G, and chemical coagulation or
soil-aquifer treatment (groundwater recharge and collection of renovated
water from aquifer with wells or drains) may be acceptable for category H.
In California, the only parameter specifically referred to in the standards
is the number of coliform organisms, but the quality of the effluent is
518

J. Irrig. Drain Eng. 1987.113:516-535.


TABLE 2. Minimum Monitoring Requirements for irrigation with Reclaimed Waste-
water in Arizona (1)
Crop and Land-Use Category

Characteristic A B C D E F G H
(D (2) (3) (4) (5) (6) (7) (8) (9)
pH 1/month 1/month 1/month 1/month 1/month 1/month 1/month 1/month
Downloaded from ascelibrary.org by Selcuk Universitesi on 01/31/15. Copyright ASCE. For personal use only; all rights reserved.

Fecal coliforms 1/month 1/month 1/month 1/week 1/week 1/week i/day 1/day
Turbidity continuous continuous
Note: The user of wastewater is not required to monitor routinely for viruses or other biological contaminants, or
for trace substances for which no sampling frequency is specified. However, should the Department of Health Services
find or have good reason to believe such contaminants are present in excess of allowable limits, corrective action
including monitoring will be required to eliminate or reduce the contaminants to acceptable levels.

defined in terms of the treatment processes required, e.g., oxidation,


coagulation, filtration, disinfection, etc. (9). For unrestricted irrigation, for
example, the effluent must be well oxidized (organic matter stabilized),
coagulated, clarified, filtered, and disinfected, so that the 7-day median
coliform concentration does not exceed 2.2/100 ml and the 30-day maxi-
mum coliform concentration does not exceed 23/100 ml. If the effluent will
only be used to irrigate fodder, fiber, or seed crops, or orchards or
vineyards, primary treatment is sufficient.
In 1979, the Israeli Ministry of Health (13) published standards for
irrigation with sewage effluent in draft form (Table 3). For food crops to be
consumed raw (category IV), sand filtration was recommended as polish-
ing treatment. In 1981, the Ministry of Health published only the standards
for category I crops as official law, indicating that for irrigation of other
crops, specific permission must be obtained (14). In giving such permis-
sion, the Ministry presumably is guided by the recommendations in Table
3. The crop categories in this table are:
TABLE 3. Standards for irrigation with Reclaimed Wastewater in Israel (13,14) (See
Text for Explanation of Crop Categories)

Crop Category3
Characteristic 1 II III IV°
(1) (2) (3) (4) (5)
Biochemical oxygen demand (mg/1) 60° 45c 35 15
Biochemical oxygen demand, filtered — — 20 10
Suspended solids content (mg/1) 50= 40° 30 15
Total coliforms, per 100 ml (80%) — — 250 12
Total coliforms, per 100 ml (50%) — — — 2.2
Chlorination contact time (min) . — — 60 120
Residual chlorine (mg/L) — — 0.15 0.5
Minimum distance from residences (m) 300 250 — —
Minimum distance from paved roads 30 25 — —
a
AU values refer to the 80-percentile, except for total coliforms in category IV, where
the 50-percentile is also specified.
b
Unrestricted irrigation; sand filtration of the effluent is mandatory.
°Not applicable to effluent from oxidation ponds with detention times of more than 15
days, where most BOD and suspended solids are of algal origin.

519

J. Irrig. Drain Eng. 1987.113:516-535.


Category I: industrial crops (cotton, sugarbeets, etc.), cereals, dry
fodder, seeds.
Category II: green fodder, olives, nuts, almonds, citrus.
Category III: fruits and vegetables for processing, vegetables to be
cooked, peeled fruits, woodlands, golf courses, soccer
fields.
Category IV: all crops without restriction, including crops to be con-
Downloaded from ascelibrary.org by Selcuk Universitesi on 01/31/15. Copyright ASCE. For personal use only; all rights reserved.

sumed raw, municipal parks, lawns.

In the case of deciduous fruits (category III), irrigation should be under the
trees only (surface or drip irrigation), and should be stopped at least 15
days before harvesting. No fruit should be picked from the ground.
Health department or other official standards for irrigation with sewage
effluent often are a compromise between what is theoretically desirable to
safeguard the public health and what is practically achievable. Regulations
that are too strict could price effluent reuse for irrigation out of the market
and, hence, hinder the use of a valuable water resource. Potentially, there
are definite health risks associated with the use of sewage effluent for
irrigation. In practice, however, these risks may not be too severe, as
evidenced by the dearth of documented cases relating disease outbreak to
irrigation with reasonably treated sewage effluent. This evidence does not
include, of course, such blatant violations of basic health rules as irrigation
of vegetables consumed raw with untreated or poorly treated sewage, as is
practiced in some countries.

WATER QUALITY FOR UNRESTRICTED IRRIGATION

The effluent to be supplied for unrestricted agricultural reuse (namely


unrestricted crop irrigation and livestock watering) has to be of such
quality that it: (1) Will not have adverse effects on crops, i.e., will not
cause yield reduction or product quality deterioration; (2) will not have
adverse effects on soils; (3) will not affect the health of people involved in
the various stages of agricultural production, i.e., the farmers, the irriga-
tion operators, the population living in the proximity of the irrigation site,
and the consumers of agricultural produce; and (4) will be acceptable to the
farmers, as well as to the general population. The major parameters of
importance in connection with unrestricted crop irrigation and livestock
watering, and their recommended limits, are discussed in the following.

Salinity
The salinity of irrigation water is a very important water quality factor
affecting plant growth. Salt concentrations in sewage effluent usually are
expressed in milligrams per liter or parts per million, which are essentially
identical for the range of concentrations found in sewage effluent. Soil-
water salinities commonly are expressed as the electrical conductivity
(EC) of the saturation extract of the soil. This extract is obtained by adding
small amounts of distilled water to a soil sample and mixing it until a
glistening paste is obtained (19). The paste is then filtered through a
vacuum filter to get the "saturation extract." The unit of EC for many
years was millimhos per centimeter at 25°C. Recently, the SI unit deciSie-
mens per meter, which is identical to millimhos per centimeter, has
520

J. Irrig. Drain Eng. 1987.113:516-535.


become the preferred expression. For most natural waters, 1 mmho/cm or
1 dS/m corresponds to a salt concentration of 640 mg/L.
Many studies have been made on the effect of soil salinity on crop yields.
Bernstein (4), for example, grew a large number of crops in artificially
salinized plots to determine the relation between yield and EC of the
saturation extract. The EC-values producing a yield that was half the
normal yield unaffected by salinity are shown in Table 4 for a number of
Downloaded from ascelibrary.org by Selcuk Universitesi on 01/31/15. Copyright ASCE. For personal use only; all rights reserved.

crops. Crops differ with respect to their tolerance to soil salinity. Where
salinity is a problem, salt-tolerant crops should be grown. Special man-
agement techniques, such as growing furrow-irrigated row crops on the
side of the ridges instead of on the top, and adequate excess irrigation for
leaching of the root zone, also should be employed.
Because all irrigation waters contain salt (sewage effluent usually has
200-400 mg/L more salt than the input water for the municipal water
supply), irrigation water must be applied to the soil in excess of the
evapotranspiration (ET) of the crop (ETis the sum of evaporation from the
soil and transpiration by the plants) to prevent salt accumulation in the root
zone. Several methods have been developed to calculate the minimum
amount of extra irrigation water or leaching (the "leaching requirement")
that is necessary to flush the salts brought into the soil with the irrigation
water out of the root zone (11). Some of the methods for calculating the
leaching requirement are based on the salt balance equation
CtDt = CtDd (1)
where C, = salt concentration of irrigation water; £>, = amount of irrigation
water that entered the soil (expressed as depth of water); Cd = salt
concentration of drainage of "deep-percolation" water leaving the bottom
of the root zone: and Dd = amount of deep-percolation water leaving the
bottom of the root zone. Dt and Dd can be taken per irrigation cycle,
season, year, or whatever time period is considered. The value of £>, is Det
+ Dd, where Det is the evapotranspiration or consumptive use of water by
the crop. The ratio DJDt is called the leaching fraction, since it indicates
how much of the irrigation water applied to the soil moves through the root
zone for leaching out salts. Eq. 1 shows that the leaching fraction is equal
to CtICd.
Several models have been proposed to relate the minimum leaching
fraction, or leaching requirement, to some readily available value of soil
salinity that is indicative of the crop's leaching requirement (11). An early
model was developed by Bernstein (4), who took Cd in Eq. 1 as the salt
concentration of the saturation extract of uniformly salinized experimental
plots that gave reductions of 50% in crop yield (Table 4). Experience has
shown that when these values are used to calculate leaching requirements
for irrigated fields, essentially normal crop yields are obtained (22). This is
so because, in the field, the salinity increases from a value associated with
C, at the top of the root zone where the roots are concentrated, to a value
associated with Cd at the bottom of the root zone. Thus, yields in the field
for a certain Cd are much higher than in plots where that Cd is uniformly
distributed throughout the entire root zone. Calculating Cd on the basis of
the CVvalues in Table 4 tends to overestimate the leaching requirement
(11).
In later years, Bernstein refined his model by reducing the calculated
521

J. Irrig. Drain Eng. 1987.113:516-535.


TABLE 4. Salt Tolerance of Crops: Electrical Conductivity (at 25°C) of Saturation
Extract of Sallnlzed Plots Which Produces 50% Reduction in Yield

Electrical
Conductivity
Crop (dS/m)
(1) (2)
Downloaded from ascelibrary.org by Selcuk Universitesi on 01/31/15. Copyright ASCE. For personal use only; all rights reserved.

(a) Field Crops


Barley 17.6
Sugarbeets 16
Cotton 16
Safflower 14
Wheat 14
Sorghum 12
Soybean 9
Sesbania 9
Paddy rice 8
Com 7
Broadbean 6.5
Flax 6.5
Beans 3.5
(b) Vegetable Crops
Beets 11.6
Spinach 8
Tomato 8
Broccoli 8
Cabbage 7
Potato 6
Corn 6
Sweet potato 6
Lettuce 5
Bell pepper 5
Onion 4
Carrot 4
Beans 3
(c) Forage Crops
Bermuda grass 18
Tall wheatgrass 18
Crested wheatgrass 18
Tall fescue 14.7
Barleyhay 13.5
Perennial rye 13
Harding grass 13
Beardless wildrye 11
Birdsfoot trefoil 10
Alfalfa 8
Orchard grass 8
Meadow foxtail 6.5
Clovers, alsike and red 4

522

J. Irrig. Drain Eng. 1987.113:516-535.


leaching requirement by 75% for crops of low to moderate salt tolerance,
and by 60% for crops of high salt tolerance. The 60% reduction improved
the fit with Hoffman's experimentally determined leaching requirements
(11), Later models use yield threshold values to estimate leaching require-
ments (11).
By way of example, Table 4 shows that for cotton Cd is 16 dS/m, or
about 10,240 mg/L. For the Salt River valley in Arizona, De, of cotton is
Downloaded from ascelibrary.org by Selcuk Universitesi on 01/31/15. Copyright ASCE. For personal use only; all rights reserved.

about 3.4 ft (104 cm) for the entire growing season. Assuming that C,- is 800
mg/L, Eq. 1 becomes (3.4 + Dd) 800 = Dd X 10,240, which gives a
deep-percolation requirement Dd of 0.3 ft/yr (9 cm/yr). Thus, D, will be 3.7
ft (113 cm), of which 3.4 ft (104 cm) is for evapotranspiration and 0.3 ft (9
cm) for leaching. Defining the field irrigation efficiency as DjDh this gives
an irrigation efficiency of 3.4/3.7 = 92%. In reality, field irrigation
efficiencies tend to be much lower. Because cotton is a salt-tolerant crop,
the leaching requirement could be reduced by 60%, yielding a Devalue of
0.12 ft (3.6 cm) and a maximum irrigation efficiency of 97%. These results
show that the normal "inefficiency" of irrigation generally is adequate for
leaching salts out of the root zone, except where crops have a low salt
tolerance, where irrigation water has a high salt content, or where very
efficient irrigation systems are used (level basins, sprinklers, drip systems).
Where Dd exceeds the minimum value, there is more deep percolation
water generated than the leaching requirement, and its salt concentration
accordingly will be less than the values indicated in Table 4. For this
reason, salt concentrations of deep percolation water from irrigated fields
in arid regions with reasonably good quality irrigation water typically are
in the range of 2,000-5,000 mg/L. High field irrigation efficiencies and the
associated small amounts of deep percolation water with a high salt
concentration are preferred, however, because:

1. Less irrigation water needs to be applied. This is important where


irrigation water is pumped, or is otherwise expensive or scarce.
2. Less fertilizer is leached from the root zone.
3. Crop yields may be improved.
4. The deep-percolation water is more concentrated, so more salt is
stored in the vadose zone en route to the groundwater. Less salt is
mobilized from soil minerals or from deeper formations, and more salt is
precipitated (mostly calcium carbonate and gypsum) in the vadose zone.
These effects result in a reduction of the total salt load on the underlying
groundwater.
5. Downward velocities of the deep-percolation water in the vadose
zone are lower, thus delaying adverse effects on groundwater, and
allowing more time for biological degradation of pesticides and other
organics.
6. The potential for developing high water tables in heavy or poorly
drained soils is reduced.

Since the irrigation efficiency Ei is De,/D,- = (A ~ Dd)ID, and the leaching


requirement Lr is DJDj, the relation between irrigation efficiency and
leaching requirement is
Et=l-Lr (2)
523

J. Irrig. Drain Eng. 1987.113:516-535.


Thus, the irrigation efficiency cannot exceed 1 - Lr if a salt balance is to
be maintained in the root zone. If Lr is relatively high and the efficiency of
the irrigation system is high, special allowance for leaching will have to be
made in the management of the irrigation system. For most graded-surface
or border-irrigation systems, the irrigation efficiency is low enough to more
than meet the leaching requirement. As long as the normal system
efficiency Et is more than 1 - Lr, however, special allowance for leaching
Downloaded from ascelibrary.org by Selcuk Universitesi on 01/31/15. Copyright ASCE. For personal use only; all rights reserved.

will have to be made in the management of the irrigation system.


The maximum allowable salt concentration of irrigation water is gov-
erned by the salt tolerance of the crop. Crops tend to be more sensitive to
salt in irrigation water at the early stages of growth than later in the
growing season. There are also limits on the allowable total salt content of
irrigation water because leaching requirements may become unpractically
high, requiring very frequent, large irrigations, and even then, crop yields
may become reduced because the average salt concentration in the root
zone may become too high. In view of these considerations, it is difficult to
prescribe maximum salt concentrations for irrigation water, and only
general classes of suitability of the water can be given. Irrigation water
standards developed by Ayers (2), for example, classify water with less
than 500 mg/L salt {Ec < 0.75 dS/m) as giving "no problems" when used
for irrigation. If the salt content is between 500 and 2,000 mg/L (0.75 and
3 dS/m), "slight to moderate" problems can be expected, and when it is
above 2,000 mg/L (3 dS/m), "severe" problems can be expected (2).
Sewage effluents typically will have a salt concentration of 300-2,000
mg/L, and most will be in the 400-800 mg/L range.
The higher the salt content of the irrigation water, the more irrigation
and crop management have to be aimed at salinity control and minimizing
salinity effects. In addition to adequate leaching and the growing of
furrow-irrigated crops on the side of the ridges rather than on the top,
farmers should select salt-tolerant crops, use good-quality water (if avail-
able) for germination and early development, and increase the plant
populations to compensate for the fact that plants irrigated with salty water
are generally smaller than when irrigated with good-quality water.

Sodium and Permeability Hazard


High concentrations of sodium in the irrigation water may adversely
affect the soil structure and reduce the soil hydraulic conductivity in
fine-textured soils.
The degree to which sodium will be absorbed by a soil is a function of the
proportion of sodium to the divalent cations (Ca and Mg), and is usually
expressed by the sodium adsorption ratio (SAR):
Na+
SAR = —== (3)
/(Ca + + + M g + + )

where the ion concentrations are expressed in meq/L.


The tendency for calcium carbonate to precipitate in soils is related to
the Langelier index of the irrigation water (18). An adjusted SAR value
usually is calculated, which takes into account the effects of precipitation
524

J. Irrig. Drain Eng. 1987.113:516-535.


TABLE 5. Classification of irrigation Water Using &4ifadj and C,

c,
Slight to
moderate Severe
SAR* No problems problems problems
(1) (2) (3) (4)
Downloaded from ascelibrary.org by Selcuk Universitesi on 01/31/15. Copyright ASCE. For personal use only; all rights reserved.

0-3 >580 130-580 <130


3-6 >830 160-830 <160
6-12 > 1,280 220-1,280 <220
12-20 > 1,980 580-1,980 <580
>20 >3,580 1,150-3,580 <1,150

and dissolution of calcium carbonate in the soil as related to the concen-


tration of COf + HCQf. The adjusted SAR value is calculated as
SARadi = SAR[9A - p(K2 - Kc) - p(Ca + Mg) - p(Alk)l (4)

where pK2 = negative logarithm of the second dissociation constant for


carbonic acid; pKc = solubility constant for calcite; and p = negative
logarithm of ion concentrations (meq/L). Values of p(K2 - Kc), p(Ca +
Mg), and P{Alk) in relation to Ca++ + Mg++ + Na+, Ca++ + Mg++, and
COf - + HCOf, respectively, have been tabulated (Ref. 6, p. 364, and
references therein).
The hydraulic conductivity of the soil is also affected by the salt
concentration of the soil solution (17). The higher the salt concentration of
the soil solution, the higher the soil hydraulic conductivity will be for a
given SAR. Ayers and Tanji (2) classified irrigation water as to its effect on
soil structure and hydraulic conductivity, taking both SARadi and the salt
concentration into account. They developed the classification shown in
Table 5 (C,-values (mg/L) as obtained by multiplying EC by 640).
Sodium also has adverse effects on the crops, such as leaf burn in
almonds, avocado, and stone fruits (Ref. 18, and references therein).
Ayers and Tanji (2) suggested that if the SAR^ of the irrigation water is
below 3, there are no sodium problems. If the SARa6i is between 3 and 9,
there are increasing problems, and if the SARadj is above 9, there are
severe problems. If water is also absorbed by the leaves, as with sprinkler
irrigation, there are no problems if the sodium concentration of the
irrigation water is below 70 mg/L, but increasing problems as it gets above
70 mg/L.

Nitrogen
Nitrogen is an intriguing constituent of municipal wastewater because of
its importance as a fertilizer in irrigated agriculture, its adverse effects
when too much is applied, and the various forms in which it can be found
in wastewater effluents. There is a tendency among some people involved
in water reuse to consider that nitrogen is beneficial to crops in any
concentration found in sewage effluents, and that, therefore, it should not
be removed from effluents used for crop irrigation. However, research
525

J. Irrig. Drain Eng. 1987.113:516-535.


TABLE 6. Nitrogen Application (lb/acre) by Irrigation with Effluent

Water Application
Nitrogen
(ft/yr)
concentration in
effluent (mg/L) 1 2 3 4 5
(1) (2) (3) (4) (5) (6)
Downloaded from ascelibrary.org by Selcuk Universitesi on 01/31/15. Copyright ASCE. For personal use only; all rights reserved.

5 14 28 42 56 70
10 27 54 81 108 135
15 41 81 122 162 203
20 54 108 162 216 270
25 68 135 203 270 338
30 81 162 243 324 405
35 95 189 284 378 473
40 108 216 324 432 540
45 122 243 365 486 608
50 135 270 405 540 675

Note: 1 lb/acre =1.12 kg/ha.

work has shown that yields, as well as product quality, may be adversely
affected by excess nitrogen application in the case of the following crops:
cotton, tomatoes for processing, sugar beets, sugar cane, potatoes, citrus,
avocados, peaches, apricots, apples, and grapes (3, 10, 15). The sugar
content of beets and the quality of the fruit (e.g., the color of apples and
oranges) were affected adversely by excess nitrogen concentrations. Over-
fertilization may cause excessive vegetative growth, lodging, and delay in
harvest for some crops (e.g., cotton), as well as toxicity to the consumers
via the nitrate accumulated in the forage. When ensiling high-nitrate
forage, nitrogen oxide gases can be formed, which are deadly to humans
and animals (3). Excess nitrate can also be converted to nitrite, which may
produce methemoglobin instead of hemoglobin in blood, and may cause
cyanosis, a fatal animal disease. Nitrogen application rates above 150-200
lb/acre (170-220 kg/ha) were found to be detrimental to some crops (3).
With each mg/L of nitrogen contained in the wastewater, about 2.7
lb/acre N are applied with each foot of irrigation water (10 kg/ha per meter
application). Table 6 shows the amounts of nitrogen applied by irrigation
with sewage effluent with N concentrations of 5-50 mg/L, and at water
application rates of 1-5 ft (0.3-1.5 m). Considering that the nitrogen
concentration in raw wastewater or secondary effluent is generally within
the range of 15-40 mg/L, and that the normal irrigation water application in
dry, warm areas like Arizona is 3-6 ft/yr (0.9-1.8 m/yr), the nitrogen
application with effluent would normally vary in the range of 120-650
lb/acre (130-730 kg/ha) per year. These higher values are much higher than
the amounts of nitrogen required by crops (Table 7).
Another problem related to the use of nitrogen in effluent as fertilizer is
that the water demand and the nitrogen demand are not parallel. For most
crops, nitrogen demand is highest during the period of active growth and
lowest during the initial growth stages and when harvest time approaches.
A comparison carried out in Israel between the nitrogen applied to various
crops in freshwater irrigation and the amount that would be applied in
526

J. Irrig. Drain Eng. 1987.113:516-535.


TABLE 7. Nutrient Uptake Rates for Various Crops (20)

Uptake (lb/acre • yr)a


Crop Nitrogen Phosphorus Potassium
0) (2) (3) (4)
(a) Forage Crops
Downloaded from ascelibrary.org by Selcuk Universitesi on 01/31/15. Copyright ASCE. For personal use only; all rights reserved.

Alfalfa" 200-480 20-30 155-200


Bromegrass 116-200 35-50 220
Coastal Bermuda grass 350-600 30-40 200
Kentucky bluegrass 180-240 40 180
Quackgrass 210-250 27-41 245
Reed canary grass 300-400 36-50 280
Ryegrass 180-250 55-75 240-290
Sweet cloverb 158 16 90
Tall fescue 135-290 26 267
(b) Field Crops
Barley 63 15 20
Corn 155-172 17-25 96
Cotton 66-100 12 34
Milomaize 81 14 64
Potatoes 205 20 220-288
Soybeans" 94-128 11-18 29-48
Wheat 50-81 15 18-42
(c) Forest Crops
Young deciduous 100
Young evergreen 60
Medium and mature deciduous 30-50 — —
Medium and mature evergreen 20-30
a
l lb/acre =1.12 kg/ha.
"Legumes will also take nitrogen from the atmosphere and will not withstand wet
conditions.

effluent irrigation indicated that the effluent should have N concentrations


of about 15-20 mg/L in order not to exceed the requirements of most crops
(15). Some of the nitrogen not used by the crop will be denitrified in the soil
and returned to the atmosphere as free nitrogen gas or nitrogen oxides. The
rest will be leached out. As a rule of thumb, one-half of the amount of
nitrogen applied to normally fertilized crops actually is taken up by the
crop, one-fourth is denitrified, and another one-fourth leaches out of the
soil (mostly as nitrate). As the nitrogen application rate increases above
the capacity of the crop to use N, the proportion of N leached increases.
Nitrate-N concentations in deep-percolation water from fields irrigated
with fresh water typically are in the 10-50 mg/L range (Ref. 6, and
references therein). Where nitrogen applications exceed crop nitrogen
requirements, higher nitrate concentrations in the deep percolation water
can be expected. Thus, sewage irrigation should be carefully managed, and
nitrate in underlying groundwater should be monitored to make sure that
undesirable nitrate pollution of the groundwater does not occur.
527

J. Irrig. Drain Eng. 1987.113:516-535.


Phosphorus
Phosphorus is another major nutrient found in wastewater, which has, in
general, a beneficial effect on crops. A comparison between phosphate
fertilization rates usually applied to various crops in Israel and phosphorus
concentrations in municipal effluent has shown that for citrus and cotton,
the amount of phosphorus that would be applied with effluent irrigation
exceeds the fertilization rate usually applied in freshwater irrigation (15).
Downloaded from ascelibrary.org by Selcuk Universitesi on 01/31/15. Copyright ASCE. For personal use only; all rights reserved.

At the normal phosphorus concentration in raw wastewater or in sewage


effluent from conventional wastewater treatment plants which do not
provide for phosphorus removal (about 10 mg/L), the amount of fertilizer
applied would be 80-100 lb/acre or 90-110 kg/ha (assuming a water
application of 3-4 ft or 0.9-1.2 m), which is more than that required by
crops (Table 7).
Phosphorus excess may have a negative effect on crops and soils. It can
cause crop yield reduction because of "nutrient imbalance"—the excess
available P may reduce the availability of micronutrients such as Cu, Fe
and Zn (20). This would occur in alkaline soils low in minor elements. High
P concentrations may also increase the precipitation of Ca from the
effluent, and consequently increase the sodium adsorption ratio of the soil
solution.
The available information on the effect of irrigation with phosphorus-
rich effluents is limited. Many soils, however, are successfully irrigated
with sewage effluent having "normal" P-concentrations of about 5-10
mg/L (mostly as P0 4 ). Soils generally have a high adsorptive capacity for
P. On sandy soils, excessive P-applications could result in leaching of P
and in P-contamination of underlying groundwater. If P-concentrations in
the effluent are reduced to 1-5 mg/L, the amount of P added with the
effluent would be similar to that added as fertilizer in freshwater irrigation
for most crops and soils.

Chloride and Chlorine


High chloride levels affect the growth of many plants, although some are
more sensitive to chloride than others. No problems would be expected if
the chloride concentration is below 140 mg/L, increasing problems when it
is in the 140-350 mg/L range, and severe problems when it is above 350
mg/L (2). These concentrations apply to surface or other irrigation systems
where water is absorbed by roots only. If water is sprayed on the field and
absorbed by the leaves also, the chloride concentrations of the irrigation
water should be below 100 mg/L to avoid problems.
Many sewage treatment plants chlorinate their effluent for disinfection,
so that there is a residual chlorine level. Because of the relatively high
concentrations of ammonium and organic matter in the effluent, residual
chlorine levels are usually quite low, and well below the 0.5-mg/L level
where problems (leaf burn in spray-irrigated crops, for example) may start
to occur.

Bicarbonate
Bicarbonate in irrigation water can leave a white residue on fruits and
leaves when the water is applied with sprinklers or other spray techniques.
While the bicarbonate deposits do not affect the yield or the flavor and
texture of fruits or other harvested products, they do affect the appearance
528

J. Irrig. Drain Eng. 1987.113:516-535.


and, hence, the salability of the products. Ayres and Tanji (2) indicate that
there are no problems when the bicarbonate concentration (expressed as
HCO3) in the irrigation water is less than 90 mg/L, increasing problems
when it is between 90 and 520 mg/L, and severe problems when it is above
520 mg/L.
Trace Elements
Downloaded from ascelibrary.org by Selcuk Universitesi on 01/31/15. Copyright ASCE. For personal use only; all rights reserved.

While some elements in small concentrations, such as B, Cu, Fe, and


Zn, are essential nutrients for plant growth, excessive concentrations of
most trace elements may have toxic effects on crops.
Maximum concentrations of trace elements allowable in irrigation water
have been established for water to be used continuously on all soils, as well
as for water to be used up to 20 yrs on fine-textured, neutral-to-alkaline
soils with a pH of 6.0-8.5 (Table 8). The former were determined for sandy
soils that have low capacities to remove trace elements from solution,
whereas the latter are for soils that have a high removal capacity (18). For
unrestricted irrigation, the concentration of trace elements should be
below the limits given for water to be used continuously on all soils, as
shown in Table 8. These limits also satisfy most of the water-quality
requirements for livestock watering.

TABLE 8. Recommended Maximum Limits (mg/L) for Trace Elements in Irrigation


Water (18)

Up to 20 yr irrigation of
Permanent irrigation fine-textured neutral to
Trace element of all soils alkaline soils (pH 6-8.5)
0) (2) (3)
Aluminum 5 20
Arsenic 0.1 2
Beryllium 0.1 0.5
Boron-sensitive crops" 0.75 2
Semitolerant crops 1
Tolerant crops 2
Cadmium 0.01 0.05
Chromium 0.1 1
Cobalt 0.05 5
Copper 0.2 5
Fluoride 1 15
Iron 5 20
Lead 5 10
Lithium: citrus 0.075 0.075
other crops 2.5 2.5
Manganese 0.2 10
Molybdenum 0.01 0.05b
Nickel 0.2 2
Selenium 0.02 0.02
Vanadium 0.1 1
Zinc 2 10
a
See Table 9 for boron sensitivity of crops.
b
For acid soils only.

529

J. Irrig. Drain Eng. 1987.113:516-535.


TABLE 9. Relative Tolerance of Plants to Boron, Listed in Decreasing Order of
Tolerance within Each Group (19)

Tolerant Semitolerant Sensitive


0) (2) (3)
Athel (Tamarix asphylla) Sunflower (native) Pecan
Downloaded from ascelibrary.org by Selcuk Universitesi on 01/31/15. Copyright ASCE. For personal use only; all rights reserved.

Asparagus Potato Black walnut


Palm (Phoenix canariensis) Acala cotton Persian (English) walnut
Date Palm (P. dactylifera) Pima cotton Jerusalem artichoke
Sugar beet Tomato Navy bean
Mangel Sweetpea American elm
Garden beet Radish Plum
Alfalfa Field pea Pear
Gladiolus Ragged Robin rose Apple
Broadbean Olive Grape (sultanina & Malaga)
Onion Barley Kadota fig
Turnip Wheat Persimmon
Cabbage Corn Cherry
Lettuce Milo Peach
Carrot Oat Apricot
Zinnia Thornless blackberry
Pumpkin Orange
Bell pepper Avocado
Sweet potato Grapefruit
Lima bean Lemon

Boron and molybdenum are two trace elements of special importance in


connection with wastewater reuse. Boron is toxic to some crops, espe-
cially citrus. The maximum allowable boron concentration in order to
avoid problems with any crop is 0.33 mg/L, according to some studies (19),
and 0.75 mg/L according to EPA limits for water to be used continuously
on all soils (18). A figure of 0.5 mg/L probably can be used as a safe
criterion for all crops. The sensitivity of various crops to boron is shown
in Table 9. Molybdenum involves a special hazard: It is not toxic to crops,
but it may be toxic to animals ingesting forage grown on soils with high Mo
concentrations (18, 20).
Suspended Solids
Deposition of suspended solids on the soil surface may clog the soil, thus
reducing water infiltration and soil aeration. Since most organic solids are
biodegradable, the organics on the surface of the soil also form an oxygen
sink, which could hinder oxygen movement from the atmosphere into the
root zone. Severe deposits of organic solids from sewage effluent on crops
and soils, as sometimes develop in rapid-infiltration basins or other land
treatment systems, can actually kill crops. In the case of sprinkler irrigation,
colloidal particles also can be deposited on leaves, where they may reduce
photosynthetic activity and adversely affect product appearance (in the
case of leafy vegetables, for example). High concentrations of suspended
solids in the irrigation water may interfere with the flow of the water in
pipes, sprinklers, drip emitters, and hydraulic structures.
Suspended solids, which are mostly of organic nature in sewage effluents,
also adversely affect the efficiency of chlorination, since bacteria and
530

J. Irrig. Drain Eng. 1987.113:516-535.


viruses can be protected by organic particles from effective contact with
chlorine. Suspended solids also detract from the aesthetics of using sewage
effluent for irrigation. This is particularly important in populated areas.
For these reasons, suspended solids should be removed as much as
possible before sewage effluent is used for irrigation. For unrestricted
irrigation, this may require sand filtration, or soil-aquifer treatment via
groundwater recharge to filter the effluent through natural soil, sand, and
Downloaded from ascelibrary.org by Selcuk Universitesi on 01/31/15. Copyright ASCE. For personal use only; all rights reserved.

gravel deposits in vadose zones and aquifers.

Acidity and Alkalinity


Bicarbonate is one form of alkalinity in wastewater. Since the soil is
usually a buffered system, the pH of irrigation water, which indicates its
acidity or alkalinity, probably will not significantly affect the pH of the soil
in the root zone. The main danger connected with water having low or high
pH values is not from the direct effects of acidity or alkalinity (because of
the high buffer capacity of most soils), but from the indirect association of
such water with high concentrations of undesirable elements: iron, man-
ganese, and aluminum in the case of acid waters and sodium; and carbonates
and bicarbonates in the case of alkaline waters (18).

Dissolved Organics
Theoretically, the presence of organic substances in high concentrations
in effluents used as irrigation water could deplete the available oxygen in
the plant root zone and, thus, adversely affect the plant growth. However,
irrigation with water having relatively high concentrations of organics has
been practiced in many areas around the world without major problems.
The biodegradable organics, which enter the sewage water with human and
kitchen wastes, are readily decomposed in the soil. In a rapid infiltration
system where 50-100 times as much effluent entered the soil as in typical
irrigation systems, the biodegradable organics (expressed as biochemical
oxygen demand, BOD), were essentially completely degraded (8). Even
for primary effluent, residual total organic carbon levels after passage
through loamy sand were the same as for secondary effluent at high
hydraulic loading rates (16).
In addition to the biodegradable organics of toilet and kitchen origin,
sewage effluent also contains a wide spectrum of synthetic organics which
are present in small concentrations, often at the microgram-per-liter level
(5). Some of these organics are decomposed in the soil under aerobic
conditions, some under anaerobic conditions, and some not at all. In a
rapid-infiltration system in Phoenix, Arizona, where about 300 ft (90 m) of
secondary effluent infiltrated the soil per year, samples of the effluent in the
infiltration basins and after it had moved about 50 ft (15 m) down the
vadose zone and 10 ft (3 m) down through the aquifer were analyzed by gas
chromatography and mass spectrometry. The soil materials were mostly
sands and gravels. The results showed that, of the volatile organics (trihalo-
methanes, chlorobenzenes, hexanes, nonanes, xylenes, etc.), 30-70% was
removed by direct volatilization from the infiltration basins. Passage
through the sands and gravels removed 50-99% of nonhalogenated hydro-
carbons. Halogenated hydrocarbons were more refractory, and decreased
to a lesser extent. These results indicate that sewage irrigation systems,
where application rates are only 1-2% of those in rapid-infiltration systems

531

J. Irrig. Drain Eng. 1987.113:516-535.


and where soils generally are heavier, on average will not pose serious
problems with trace organics and possible pollution of underlying ground-
water. However, if the sewage contains unusually high concentrations of
one or more synthetic organics (as may occur in certain industrial waste
discharges) that are not biodegradable and not readily adsorbed to the soil
matrix, problems may result.
Downloaded from ascelibrary.org by Selcuk Universitesi on 01/31/15. Copyright ASCE. For personal use only; all rights reserved.

Pesticides
Pesticides are used extensively in agriculture because of their important
benefits in insect, weed, and algae control. A great variety of products is
available as either insecticides or herbicides.
There are two types of dangers connected with pesticides in irrigation
water: (1) Their possible effects on the growth and quality of crops; and (2)
their possible effect on groundwater underlying irrigated fields. No data are
available on the effect of herbicides in sewage effluent on crops and, thus,
no limits have been established for allowable herbicide concentration in
irrigation water. Because the soil has only a limited capacity for reducing
the concentration of some pesticides such as chlorinated hydrocarbons, it
is considered advisable that pesticide contamination should not greatly
exceed the limits recommended for drinking water (20).

Polynuclear Aromatic Hydrocarbons (PAH)


These are high-molecular-weight, refractory compounds such as chry-
sene, fluorene, benzopyrene, and anthracene. Six PAHs are known to be
carcinogenic, and are included in the World Health Organization Interna-
tional Standards for Drinking Water (23). Little information is available on
the behavior of these compounds in soil, but it seems safe to assume that
they are not readily biodegraded. They may, however, be sorbed to a
considerable extent. To be on the safe side, their concentrations in irri-
gation water to be used without any restriction (including in areas where
leaching may occur to groundwater used for potable purposes) should not
be much higfier than those recommended for drinking water.

Surfactants
The presence of surfactants in irrigation water represents an aesthetic
hazard because of the foam occurring at concentrations higher than 0.5
mg/L. Foam may appear near hydraulic structures, as well as in the
irrigated fields. Moreover, the presence of detergents in high concentra-
tions in the irrigation water has been associated with changes in infiltration
rates. Biodegradable detergents should present no agronomic problems.
The common test for surfactants—measurement of methylene blue active
substances (MBAS)—is of limited significance, since it reflects only the
presence of anionic detergents. Cationic and nonionic detergents that
should also be of concern in agriculture do not react with methylene blue.
Little quantitative information is available on the effect of detergents on
crops and soils.
Phenols
This group of compounds is another group of organic substances which
is included in drinking-water standards because of its adverse effect on the
water taste, particularly after chlorination. Considering that adverse
effects of phenols on crops or soils have not been reported, and that they
532

J. Irrig. Drain Eng. 1987.113:516-535.


are biodegradable and, hence, removed in the soil-aquifer system, normal
concentrations of phenolic compounds in sewage water used for irrigation
should not be objectionable.

Pathogens
Bacteria and viruses in wastewater present a major health hazard in
agricultural reuse. The effluent to be used for unrestricted agricultural
Downloaded from ascelibrary.org by Selcuk Universitesi on 01/31/15. Copyright ASCE. For personal use only; all rights reserved.

reuse has to be of a very high bacteriological and virological quality, so


that food crops for raw consumption or brought raw into the kitchen can be
included in the effluent irrigation scheme without any health hazards. A
high bacteriological quality is also important in order to minimize the
danger of aerosols spreading organisms during sprinkler irrigation, and to
improve the public acceptance of the effluent.
The bacteriological quality is determined by the number of fecal coli-
forms, which are indicators of fecal contamination and may imply the
presence of pathogenic bacteria, such as salmonella, shigella, cholera, and
others.
The bacterial standard for irrigation water is controversial. Whereas
WHO standards for unrestricted irrigation are not more than 100 fecal
conforms per 100 ml, more recent standards in various parts of the world
are more stringent. In California and in Arizona, the bacterial standard for
irrigation of food crops to be eaten raw is not more than 2.2 fecal coliforms
per 100 ml (geometric mean) and not more than 23/100 ml (25/100 ml in
Arizona) per single sample (1,9). In Israel, for irrigation of food crops to be
eaten raw, the concentration of total coliform bacteria should be less than
12/100 ml in at least 80% of the samples, and less than 2.2/100 ml in at least
50% of the samples (13).
The virological quality of effluents used for irrigation is being given more
and more consideration, especially as new methods of virus detection and
enumeration become available. In Arizona, a maximum limit for enteric
viruses of 1 PFU (plaque-forming unit) per 40 L has been proposed for
irrigation of food crops to be eaten raw. The major problems associated
with detecting viruses in effluents are: (1) The need to concentrate the
viruses from large amounts of water; (2) the lack of a simple, universally
accepted, method of virus detection; (3) the existence of a great variety of
viruses; and (4) the lack of an indicator organism for viruses.

CONCLUSIONS

Irrigation is an excellent use of sewage effluent because it is mostly


water with nutrients. There are, however, other substances in the effluent
that could adversely affect the crop, the soil, the underlying groundwater,
the farm workers, and/or the consumers (human and animal) of the crops.
To determine the suitability of a given effluent for irrigation or the quality
requirements for irrigation of certain crops, the chemical and biological
composition of the effluent is compared with known quality standards for
irrigation. Restricted irrigation usually involves the use of conventional
primary or secondary effluent on special "sewage farms" where there is
close control over the water management and the type of crops grown
(usually forage, fiber, and seed crops not intended for human consump-
tion).

533

J. Irrig. Drain Eng. 1987.113:516-535.


Large-scale irrigation with sewage effluent by farmers without close
control or supervision from regulatory agencies requires that the effluent
be treated for unrestricted irrigation, which includes crops that are
consumed raw or brought raw into the kitchen. Bacteriological standards
for such use are formulated by state health or water resources depart-
ments, and they usually specify a mean fecal coliform concentration of less
than 2.2/100 ml, with maximum values not exceeding 25/100 ml. Arizona
Downloaded from ascelibrary.org by Selcuk Universitesi on 01/31/15. Copyright ASCE. For personal use only; all rights reserved.

also requires a virus count of less than 1 unit per 40 L, and absence of
parasitic worm eggs and of Amoeba hystolytica. For proper disinfection,
the suspended solids content of the effluent should be low. A low
suspended solids content also enhances the aesthetics of irrigation with
sewage effluent, which is especially important in populated areas. Agro-
nomic parameters to be considered include salinity, sodium adsorption
ratio, nitrogen, phosphorus, chloride, chlorine, bicarbonates, boron,
heavy metals and other trace elements, pH, and synthetic organics (in-
cluding pesticides).
Most effluents from predominantly residential areas meet the agronomic
requirements for unrestricted irrigation. Problems may arise where there
are significant industrial discharges in the municipal sewer system. In
warm, arid areas where irrigation applications are three to seven feet (1-2
m) per year, too much nitrogen may enter the soil with the applied effluent.
This situation may require nitrogen removal from the sewage in the
treatment system (including soil-aquifer treatment via groundwater re-
charge where applicable), or blending the effluent with normal irrigation
water, at least during the early and late stages of the growing season. Many
successful sewage irrigation systems exist throughout the world.
Treatment procedures to achieve the desired quality of sewage effluent
for irrigation include primary treatment, secondary (biological) treatment,
oxidation ditch, lagooning, coagulation, sedimentation, filtration, lime
precipitation, ammonia volitalization, denitrification, phosphate precipita-
tion, disinfection, and soil-aquifer treatment (SAT). The latter is achieved
via groundwater recharge, using rapid-infiltration basins to put partially
treated sewage effluent underground, and wells or drains to collect the
sewage water after it has filtered through the vadose zone and aquifer, and
has become "renovated" water (7). Primary or secondary sewage effluent
can be used for SAT systems. The renovated water from SAT systems
generally meets all quality requirements for unrestricted irrigation. Thus,
where local soil and hydrogeological conditions are favorable for ground-
water recharge with rapid-infiltration basins, SAT is an effective way to
treat the sewage effluent for unrestricted irrigation use. For a more detailed
discussion of the various treatment procedures and the optimum sequence
of treatment steps, reference is made to Idelovitch and Bouwer (12).

APPENDIX. REFERENCES

1. Arizona Department of Health Services (1983). "Regulations for the reuse of


wastewater." Water quality standards, Arizona Department of Health Ser-
vices, Jun.
2. Ayers, R. S., and Tanji, K. K. (1981). "Agronomic aspects of crop irrigation
with wastewater." Proc. Water Forum '81, 1, ASCE, New York, N.Y.,
579-586.
3. Baier, D. C , and Fryer, W. B. (1973). "Undesirable plant responses with

534

J. Irrig. Drain Eng. 1987.113:516-535.


sewage irrigation." / . Irrig. and Drain. Div,, 99 (IR2), 133-142.
4. Bernstein, L. (1964). "Salt tolerance of plants." U.S. Dept. ofAgr. Informa-
tion Bull. No. 283, Washington, D.C.
5. Bouwer, E. J., et al. (1984). "Organic contaminant behavior during rapid
infiltration of secondary wastewater at the Phoenix 23rd Avenue Project "
Water Res., 18(4), 463-472.
6. Bouwer, H. (1976). Groundwater hydrology. McGraw-Hill Book Company,
New York, N.Y.
Downloaded from ascelibrary.org by Selcuk Universitesi on 01/31/15. Copyright ASCE. For personal use only; all rights reserved.

7. Bouwer, H. (1982). "Wastewater reuse in arid areas. Water reuse, E. J.


Middlebrooks, ed., Ann Arbor Science Publishers, Inc., Ann Arbor, Mich.,
137-180.
8. Bouwer, H., Lance, J. C , and Riggs, M. S. (1974). "High-rate land
treatment. II. Water quality and economic aspects of the Flushing Meadows
Project." J. Water Pollut. Control. Fed., 46(5), 844-859.
9. California Department of Health. (1978). "Wastewater reclamation criteria."
Excerpt from the California Administrative Code, Title 22, Division 4—
Environmental Health, California Dept. of Health.
10. Childers, N. F., ed. (1966). "Temperate to tropical fruit nutrition." Horticul-
tural Publication, Rutgers State University, New Brunswick, N.J.
11. Hoffman, G. J. (1982). "Leaching requirements for managing salinity."
Proc. Specialty Conf. Adv. in Irrig. and Drain. ASCE, New York, N.Y.,
409-416.
12. Idelovitch, E., and Bouwer, H. (1984). "Wastewater treatment for irriga-
tion." Internal report, TAHAL Consulting Engrs., Ltd., Tel Aviv 64-364,
Israel.
13. Israeli Ministry of Health. (1979). "Recommendations for treatment of
wastewater to be used for crop irrigation," 2nd Draft, Israeli Ministry of
Health, Nov.
14. Israeli Ministry of Health. (1981). "Public Health Law No. 4263: Purification
of sewage water to be used for irrigation." Israel: Ministry of Health, Aug.
27.
15. Kary, S. (1966). "The use of effluent for unrestricted irrigation." Progress
Rpt. No. 1, Tahal Publication No. 01/86-66, Sep., TAHAL Consulting
Engineers, Ltd., Tel Aviv, Israel.
16. Lance, J. C , Rice, R. C , and Gilbert, R. G. (1980). "Renovation of sewage
water by soil columns flooded with primary effluent." J. Water Pollut.
Control Fed., 52(2), 381-388.
17. McNeal, B. L. (1968). "Prediction of the effect of mixed-salt solutions on
the soil hydraulic conductivity." Proc. Soil Sci. Soc. Am., 32, 190-193.
18. National Academy of Sciences and National Academy of Engineering.
(1973). "Water quality criteria 1972." Report of the Committee on Water
Quality Criteria, Ecological Research Series, EPA-R3-73-033, Mar.
19. United States Department of Agriculture. (1954). Diagnosis and improve-
ment of saline and alkaline soils." Agr. Handbook No. 60, U.S. Salinity
Laboratory Staff, USDA, Washington, D . C , Feb.
20. U.S. Environmental Protection Agency. (1977). Process design manual for
land treatment of municipal wastewater. EPA 625/1-77-008, Cincinnati,
Ohio, Oct.
21. U.S. Environmental Protection Agency, (1977). Manual of treatment
techniques for meeting the interim primary drinking water regulations.
EPA-600/8-77-005, Cincinnati, Ohio.
22. Van Schilfgaarde, J., et al. (1973). "Irrigation management for salt control."
Proc. Irrig. and Drain. Div. Specialty Conf. ASCE, New York, N.Y., Ap.,
647-672.
23. World Health Organization. (1972). "International standards for drinking
water." World Health Organization, Geneva, Switzerland.

535

J. Irrig. Drain Eng. 1987.113:516-535.

Das könnte Ihnen auch gefallen