National Academies Press: OpenBook

Ground Water Recharge Using Waters of Impaired Quality (1994)

Chapter: 2 Source Waters and Their Treatment

« Previous: 1 An Introduction to Artificial Recharge
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

2
Source Waters and Their Treatment

Water for human use comes from various sources—generally lakes, rivers, or underground aquifers. This report, which examines the potential of artificial recharge, uses the term ''source water" to mean the recharge source—the water supplied to a surface infiltration or injection well recharge system. Potential source waters of impaired quality for artificial recharge include treated municipal wastewater, stormwater runoff, and irrigation return flows. The quality of source waters may be improved by the use of various pretreatment and disinfection processes. This chapter evaluates the quality of municipal wastewater and the quality improvement gained from primary, secondary, and advanced wastewater treatment. The quality of urban stormwater runoff and its possible treatment methods are also discussed, as is the quality of irrigation return flow and the problems inherent in treating it. Industrial wastewater and industrial stormwater runoff are not considered in depth; although industrial wastewater might at times be suitable for ground water recharge, its potential use would be extremely site specific and a general evaluation is not useful.

The quality of the source water considered for ground water recharge has a direct bearing on operational aspects of recharge facilities and also on the use to be made of the recovered water. The source water characteristics that affect the operational aspects of recharge facilities include suspended solids (SS), dissolved gases, nutrients, biochemical oxygen demand (BOD), microorganisms, and the sodium adsorption ratio (which affects soil permeability). The constituents that have the greatest possible adverse effects when the recharge is intended to support potable use include organic and metallic toxicants, nitrogen compounds, and pathogens.

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

MUNICIPAL WASTEWATER

Characteristics

The quantity and quality of wastewater delivered from varies among communities, depending on the number of commercial and industrial establishments in the area, the per capita in-house water use (which may vary from 400 1/day or more in industrialized countries to 40 1/day or less in developing or water-short countries), and the condition of the sewer system. Raw municipal wastewater may include contributions from domestic and industrial sources, infiltration and inflow from the collection system, and, in the case of combined sewer systems, urban stormwater runoff. The typical composition of untreated municipal wastewater appears in Table 2.1.

The occurrence and concentration of pathogenic microorganisms in raw wastewater depend on a number of factors, and it is not possible to predict with any degree of assurance what the general characteristics of a particular wastewater will be with respect to infectious agents. Important variables include the sources contributing to the wastewater, the original purpose of the water use, the general health of the contributing population, the existence of "disease carriers" in the population, and the ability of infectious agents to survive outside their hosts under a variety of environmental conditions. Table 2.2 lists infectious agents potentially present in untreated municipal wastewater. Table 2.3 illustrates the variety and order of magnitude of the concentration of microorganisms in untreated municipal wastewater.

Viruses are not normally excreted for prolonged periods by healthy individuals, and the occurrence of viruses in municipal wastewater fluctuates widely. Viral concentrations are generally highest during the summer and early autumn months. Viruses shed from an infected individual commonly range from 1,000 to 100,000 infective or plaque forming units (pfu's) per gram (g) of feces, but may be as high as 1,000,000 pfu/g of feces (Feachem et al., 1983). Viruses as a group are generally more resistant to environmental stresses than many of the bacteria, although some viruses persist for only a short time in municipal wastewater. Vital levels in the United States have been reported to be as high as 700 pfu/100 ml, but are typically less than 100 pfu/100 ml (American Society of Civil Engineers, 1970; Melnick et al., 1978).

Dissolved inorganic solids (total dissolved solids or salts, TDS) are not altered substantially in most wastewater treatment processes. In some cases, they may increase as a result of evaporation in lagoons or storage reservoirs. Therefore, unless wastewater treatment processes specifically intended to remove mineral constituents are employed, the composition of dissolved minerals in treated wastewater used for ground water recharge can be expected to be similar to the composition in the raw wastewater. The concentration of dissolved minerals in untreated wastewater is determined by the concentration in the domestic water

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

TABLE 2.1 Typical Contaminants in Untreated Municipal Wastewater

 

 

 

Concentration

 

Unit

Weak

Medium

Strong

Solids, total

mg/l

350

720

1,200

Dissolved, total

mg/l

250

500

850

Fixed

mg/l

145

300

525

Volatile

mg/l

105

200

350

Suspended solids

mg/l

100

220

350

Fixed

mg/l

20

55

75

Volatile

mg/l

80

165

275

Setteable solids

mg/l

5

10

20

Biochemical oxygen demanda

mg/l

110

220

275

Total organic carbon (TOC)

mg/l

80

160

290

Chemical oxygen demand (COD)

mg/l

250

500

1,000

Nigrogen (total as N)

mg/l

20

40

85

organic

mg/l

8

15

45

Free ammonia

mg/l

12

25

50

Nitrites

mg/l

0

0

0

Nitrates

mg/l

0

0

0

Phosphorus (total as P)

mg/l

4

8

15

Organic

mg/l

1

3

5

Inorganic

mg/l

3

5

10

Chloridesb

mg/l

30

50

100

Sulfateb

mg/l

20

30

50

Alkalinityc

mg/l

50

100

NA

Grease

mg/l

50

100

150

Total coliform

no./100 ml

106-107

107-108

107-109

Volatile organic compounds

µg/l

<100

100-400

> 400

Note: NA = not available.

a 5-day, 20ºC(BOD5, 20ºC).

b Values should be increased by amount present in domestic water supply.

c As calcium carbonate (CaCO3).

Source: Metcalf & Eddy, Inc., 1991.

supply plus mineral pickup resulting from domestic water use, which in the United States varies from 200 to 400 mg/l.

Wastewater treatment levels are generally classified as preliminary, primary, secondary, and advanced. The nature of each level of treatment is discussed in the following sections. (Note that the costs of treatment at least through secondary treatment is required before disposal to receiving waters and therefore would not be the responsibility of the recharge operation.)

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

TABLE 2.2 Infectious Agents Potentially Present in Untreated Municipal Wastewater

 

Disease

Protozoa

Entamoeba histolytica

Amebiasis (amebic dysentery)

Giardia lambliai

Giardiasis

Balantidium coli

Balantidiasis (dysentery)

Cryptosporidium

Crytosporidiosis, diarrhea, fever

Helminths

Ascaris lumbricoides (roundworm)

Ascariasis

Ancylostoma duodenale (hookworm)

Ancylostomiasis

Necator americanus (roundworm)

Necatoriasis

Ancylostoma (spp.) (hookworm)

Cutaneous larva migrans

Strongyloides stercoralis (threadworm)

Strongyloidiasis

Trichuris trichuria (whipworm)

Trichuriasis

Taenia (spp.) (tapeworm)

Teaniasis

Enterobius vermicularis (pinworm)

Enterobiasis

Echinococcus granulosus (spp.) (tapeworm)

Hydatidosis

Bacteria

Shigella (4 spp.)

Shigellosis (dysentery)

Salmonella ryphi

Typhoid fever

Salmonella (1,700 serotypes)

Salmonellosis

Vibro cholerae

Cholera

Escherichia coli (enterophathogenic)

Gastroenteritis

Yersinia enterocolitica

Yersiniosis

Leptospira (spp.)

Leptospirosis

Legionella pneumopilla

Legionnaire's disease

Campylobacter jejune

Gastroenteritis

Viruses

Enteroviruses (72 types) polio, echo, coxsackie, new enteroviruses

Gastroenteritis, heart anomalies meningitis, others

Hepatitis A virus

Infectious hepatitis

Adenovirus (47 types)

Respiratory disease, eye infections

Rotavirus (4 types)

Gastroenteritis

Parvovirus (3 types)

Gastroenteritis

Norwalk agent

Diarrhea, vomiting, fever

Reovirus (3 types)

Not clearly established

Astrovirus (5 types)

Gastroenteritis

Calicivirus (2 types)

Gastroenteritis

Coronavirus

Gastroenteritis

 

Source: Adapted from Sagik et al., 1978; Hurst et al., 1989.

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

TABLE 2.3 Microorganism Concentrations in Untreated Municipal Wastewater

 

Concentration (number per 100 ml)

Fecal Coliforms

104-109

Fecal streptococci

104-106

Shigella

1-103

Salmonella

102-104

Pseudomonas aeruginosa

103-104

Clostridium perfringens

103-105

Helminth ova

1-103

Giardia lamblia cysts

10-104

Cryptosporidium oocysts

10-103

Entamoeba histolytica cysts

1-101

Enteric viruses

102-104

 

Source: Adapted from various sources.

Primary Treatment

The first step in treatment, sometimes referred to as preliminary treatment, consists of the physical processes of screening, or comminution, and grit removal. Coarse screening is usually the first step and is used to remove large solids and trash that may interfere with later treatment processes. Comminution devices are sometimes used to cut up solids into a smaller size to improve downstream operations. Grit chambers are designed to remove material such as sand, gravel, cinders, eggshells, broken glass, seeds, coffee grounds, and large organic particles, such as food waste. Settling of most organic solids is prevented in the grit chamber because of the high flow velocity of wastewater through the chamber. Other preliminary treatment operations can include flocculation, odor control, chemical treatment, and pre-aeration.

Past this initial screening, primary treatment consists of physical processes to remove settleable organic and inorganic solids by sedimentation and floating materials by skimming. These also removes some of the organic nitrogen, organic phosphorus, and heavy metals. Additional phosphorus and heavy metal removal can be achieved through the addition of chemical coagulants and polymers. Primary treatment, together with preliminary treatment, typically removes 50 to 60 percent of the suspended solids and 30 to 40 percent of the organic matter. Primary treatment does not remove the soluble constituents of the wastewater.

Primary treatment has little effect on the removal of most biological species present in wastewater. However, some protozoa and parasite ova and cysts will settle out during primary treatment, and some particulate-associated microorgan-

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

ism may be removed with settable matter. Primary treatment does not reduce the level of viruses in municipal wastewater. While primary treatment by itself generally is not considered adequate for ground water recharge applications, primary effluent has been successfully used in soil-aquifer treatment systems at some spreading sites where the extracted water is to be used for nonpotable purposes (Carlson et al., 1982; Lance, Rice, and Gilbert, 1980; Rice and Bouwer, 1984). The higher organic content of primary effluent may enhance nitrogen removal by denitrification in the SAT system (Lance, Rice, and Gilbert, 1980) and may enhance removal of synthetic organic compounds by stimulating greater microbiological activity in the soil (McCarty, Rittman, and Bouwer, 1984). A disadvantage of using primary effluent is that infiltration basin hydraulic loading rates may be lower because of higher suspended solids and weaker biological activity on and in the soil of the infiltration system. Also, too much organic carbon in the recharge water can have adverse effects on processes that occur in the soil and aquifer systems. In most cases, wastewater receives at least secondary treatment and disinfection, and often tertiary treatment by filtration, prior to augmentation of nonpotable aquifers by surface spreading.

Secondary Treatment

Secondary treatment is intended to remove soluble and colloidal biodegradable organic matter and suspended solids (SS). In some cases, nitrogen and phosphorus also are removed. Treatment consists of an aerobic biological process whereby microorganisms oxidize organic matter in the wastewater. Several types of aerobic biological processes are used for secondary treatment, including activated sludge, trickling filters, rotating biological contactors (RBCs), and stabilization ponds. Generally, primary treatment precedes the biological process; however, some secondary processes are designed to operate without sedimentation, e.g., stabilization ponds and aerated lagoons. Typical microorganism removal efficiencies for activated sludge and trickling filter secondary treatment processes are given in Table 2.4. Concentration ranges for inorganic constituents and some other parameters in secondary-treated municipal wastewater are presented in Table 2.5. Information on the concentration of trace organics in activated sludge secondary effluent from the City of Phoenix's 23rd Avenue treatment plant is given in the "Phoenix, Arizona Projects" case study in Chapter 6.

The activated sludge process is considered to be a high-rate biological process because of the high concentrations of microorganisms used for the metabolization of organic matter. Trickling filters may be classified as either low-rate or high-rate based on their hydraulic and organic loading, mode of operation, and other factors. These processes accomplish biological oxidation in relatively small basins and use sedimentation tanks (secondary clarifiers) after

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

TABLE 2.4 Typical Percent Removal of microorganisms by Conventional Treatment Processes

 

Primary Treatment

Secondary Treatment

 

Activated Sludge

Trickling Filter

Fecal coliforms

< 10

0-99

85-99

Salmonella

0-15

70-99

85-99 +

Mycobacterium tubercolulosis

40-60

5-90

65-99

Shigella

15

80-90

85-99

Entamoeba histolytica

0-50

Limited

Limited

Helminth ova

50-98

Limited

60-75

Enteric viruses

Limited

75-99

0-85

 

Source: Crook, 1992.

the aerobic process to separate the microorganisms and other settleable solids from the treated wastewater.

In the activated sludge process, treatment is provided in an aeration tank in which the wastewater and microorganisms are in suspension and continually mixed through aeration. Trickling filters utilize media such as stones, plastic shapes, or wooden slats in which the microorganisms become attached. RBCs are similar to trickling filters in that the organisms are attached to support media, which in this case are partially submerged rotating discs in the wastewater stream.

These processes are capable of removing up to 95 percent of BOD, COD, and SS originally present in the wastewater and significant amounts of many (but not all) heavy metals and specific organic compounds (Water Pollution Control Federation, 1989). Trickling filters are not as effective as activated sludge processes in removing soluble organics because of less contact between the organic matter and microorganisms. Activated sludge treatment can reduce the soluble BOD fraction to 1 to 2 mg/l, while the trickling filter process typically reduces soluble BOD to 10 to 15 mg/l (U.S. Environmental Protection Agency, 1992). Biological treatment, including sedimentation, typically reduces the total BOD to 15 to 30 mg/l, COD to 40 to 70 mg/l, and TOC to 15 to 25 mg/ 1 (U.S. Environmental Protection Agency, 1992). Very few dissolved minerals are removed during conventional secondary treatment.

Pond systems require relatively large land areas and are most widely used in rural areas and in warm climates and where land is available at reasonable cost. They are often arranged in a series of anaerobic, facultative, and maturation ponds with an overall hydraulic detention time of 10 to 50 days, depending on the design temperature and effluent quality required (Mara and Cairncross, 1989). Most organic matter removal occurs in the anaerobic and facultative ponds.

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

TABLE 2.5 Constituent Concentrations and other Parameters for Secondary-Treated Municipal Wastewater

 

Concentrationa (mg/l)

Calcium

9-84

Potassium

9-108

Magnesium

12-176

Sodium

44-1320

Amonium

0-501

Chlorine

43-2450

Fluoride

0.2-3.8

Bicarbonate

76-563

Nitrate

0.4-30

Phosphate

1.2-46

Sulfate

14-490

Silicondioxide

10-76

Hardness (as calcium carbonate)

62-951951

pH (units)

6.3-8.4

Electrical conductivity

423-6570 µmho/cm

Total dissolved solids

210-4580

Arsenic

< 0.005-0.023

Boron

0.3-2.5

Cadmium

< 0.005-0.22

Chromium

< 0.001-0.1

Copper

0.006-0.053

Lead0.003-0.35

0.003-0.35

Molybdenum

0.001-0.018

Molybdenum

0.001-0.018

Mercury

0.001-0.018

Nichol

0.003-0.60

Zinc

0.004-0.35

Biochemical oxygen demand

1.5-30

Chemical oxygen demand

40-70

Total suspended solids

10-25

Total organic carbon

15-25

a Concentration expressed in milligrams per liter unless otherwise noted.

Source: Treweek, 1985; Crook, 1992.

Maturation ponds, which are largely aerobic, are designed primarily to remove pathogenic microorganisms following biological oxidation processes. Well-designed stabilization pond systems are capable of reducing the BOD to 15 to 30 mg/l, COD to 90 to 135 mg/l, and SS to 15 to 40 mg/l (Shuval et al., 1986). The organic matter remaining in the effluent consists of soluble, biodegradable organic matter present in the raw wastewater, but not removed, plus intermediate

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

products formed during the biological degradation of organic compounds and microbial cellular constituents (Metcalf & Eddy, Inc., 1991). The suspended solids are mainly organic in nature, consisting of biological solids produced during secondary treatment and other solids that escaped treatment and separation.

Stabilization ponds use algae to provide oxygen to the system. This system is considered a low-rate biological process. Mechanically aerated lagoon systems sometimes are used to provide secondary-level treatment. Stabilization ponds are capable of providing considerable nitrogen removal under certain conditions (e.g., high temperature and pH and long detention times) and are effective in removing microorganisms from wastewater. Well-designed and well-operated pond systems are capable of achieving a 6-log (99.9999 percent) reduction of bacteria, a 3-log (99.9 percent) reduction of helminths, and a 4-log (99.99 percent) reduction of viruses and cysts (Mara and Cairncross, 1989). Algae produced during pond treatment may present soil clogging problems during recharge.

Tertiary Treatment

The treatment of wastewater beyond the secondary or biological stage is sometimes called tertiary treatment. The term normally implies the removal of nutrients such as phosphorus and nitrogen, and a high percentage of suspended solids. However, the term tertiary treatment is now being replaced in most cases by the term advanced wastewater treatment—which refers to any physical, chemical, or biological treatment process used to accomplish a degree of treatment greater than that achieved by secondary treatment.

Advanced Wastewater Treatment

Advanced wastewater treatment processes are designed to remove suspended solids and dissolved substances, either organic or inorganic in nature. Advanced wastewater treatment processes generally are used when a high-quality reclaimed water is necessary, such as for direct injection into potable aquifers. Commonly used processes and their principal removal functions are given in Table 2.6. The major advanced wastewater processes associated with ground water recharge are coagulation-sedimentation, filtration, nitrification, denitrification, phoshorus removal, carbon adsorption, and reverse osmosis.

Coagulation-Sedimentation

Chemical coagulation with lime, alum, or ferric chloride followed by sedimentation removes suspended solids, heavy metals, trace substances, phosphorus, and turbidity. Vital inactivation under alkaline pH conditions can be accom-

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

TABLE 2.6 Constituent Removal by Advanced Wastewater Treatment Operations and Processes

Principal Removal Function

Description of Operation or Process

Type of Wastewater Treateda

Suspended solids removal

Filtration

EPT, EST

 

Microstrainers

EST

Ammonia oxidation

Biological nitrification

EPT, EBT, EST

Nitrogen removal

Biological nitrification/denitrification

EPT, EST

Nitrate removal

Separate-stage biological

EPT + nitrification denitrification

Biological phosphorus removal

Mainstream phosphorus removalb

RW, EPT

 

Sidestream phosphorus removal

RAS

Combined nitrogen and phosphorus removal by biological methods

Biological nitrification/denitrification and phosphorus removal

RW, EPT

Nitrogen removal by physical or chemical methods

Air stripping

EST

 

Breakpoint chlorination

EST + filtration

 

Ion exchange

EST + filtration

hosphorus removal by chemical addition

Chemical precipitation with metal salts

RW, EPT, EBT, EST

 

Chemical precipitation with lime

RW, EPT, EBT, EST

Toxic compounds and refractory organics removal

Carbon adsorption

EST + filtration

 

Powdered activated carbon

EPT

 

Chemical oxidation

EST + filtration

Dissolved inorganic

Chemical precipitation

RW, EPT, EBT, EST

 

Ion exchange

EST + filtration

 

Ultrafiltration

EST + filtration

 

Reverse osmosis

EST + filtration

 

Electrodialysis

EST + filtration + carbon adsorption

Volatile organic compounds

Volatilization and gas stripping

RW, EPT

a EPT-effluent from primary treatment; EBT-effluent from biological treatment (before clarification); EST-effluent from secondary treatment: RW-raw (untreated sewage); and RAS-return activated sludge.

b Removal process occurs in the main flowstream as opposed to sidestream treatment.

Source: Metcalf & Eddy, Inc., 1991.

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

plished using lime as a coagulant, but pH values of 11 to 12 are required before significant inactivation is obtained.

Filtration

Filtration is a common treatment process used to remove particulate matter prior to disinfection. Filtration involves the passing of wastewater through a bed of granular media, which retain the solids. Typical media include sand, anthracite, and garnet. Removal efficiencies can be improved through the addition of certain polymers and coagulants. Table 2.7 presents average constituent removal efficiencies for filtration. The concentrations of organic and inorganic constituents at three water reclamation plants operated by the Sanitation Districts of Los Angeles County are shown in Tables 2.8 and 2.9, respectively. All three of the treatment plants have conventional activated sludge treatment followed by filtration and disinfection.

Treatment of biologically treated secondary effluent by chemical coagulation, sedimentation, and filtration has been demonstrated to remove more than 99 percent of seeded poliovirus (Sanitation Districts of Los Angeles County, 1977). This treatment chain reduces the turbidity of the wastewater to very low levels, thereby enhancing the efficiency of the subsequent disinfection process. Chemical coagulation and sedimentation alone can remove up to 99 percent of viruses, although the presence of organic matter can significantly decrease the amount of viruses removed. Direct filtration, that is, chemical coagulation and filtration without sedimentation, has also been shown to remove up to 99 percent of seeded poliovirus (Sanitation Districts of Los Angeles County, 1977). In one study, sand and dual-media filtration of secondary effluent, without coagulant addition prior to filtration, did not significantly reduce enteric viral levels (Noss et al., 1989). The primary purpose of the filtration step is not to remove viruses, but to remove protozoa and helminth eggs and floc and other suspended matter that may contain adsorbed or enmeshed microorganisms, thereby making the disinfection process more effective.

Chemical coagulation and filtration followed by chlorine disinfection to very low total coliform levels can remove or inactivate 5 logs (99.999 percent) of seeded poliovirus through these processes alone and subsequent to conventional biological secondary treatment can produce effluent essentially free of measurable levels of bacterial and viral pathogens (Sanitation Districts of Los Angeles County, 1977; Sheikh et al., 1990).

All parasitic cysts may not be removed by direct filtration. In one study, Giardia cysts were present in concentrations ranging from 3 to 7 cysts/100 liters after direct filtration (Casson et al., 1990). Rose and Carnahan (1992) found both Giardia and Cryptosporidium cysts and oocysts in reclaimed water after direct filtration of activated sludge effluent in about 20 percent of the samples collected. Filtered, secondary effluent from the Reedy Creek Wastewater Treat-

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

TABLE 2.7 Treatment Levels Achievable with Various Combinations of Advanced Wastewater Treatment Unit Processes

 

Typical Effluent Qualityb

Treatment Processa

SS (mg/l)

BOD5 (mg/l)

COD (mg/l)

Total N (mg/l)

NH3-N (mg/l)

PO4 as P (mg/l)

Turbidity (NTU)

AS+F

4-6

<5-10

30-70

15-35

15-25

4-10

0.3-5

AS+F+CA

<3

<1

5-15

15-30

15-25

4-10

0.3-3

AS/N

(single stage)

10-25

5-15

20-45

20-30

1-5

6-10

5-15

AS/N-D

(separate stages)

10-25

5-15

20-35

5-10

1-2

6-10

5-15

MS addition to AS

10-20

10-20

30-70

15-30

15-25

<2

5-10

MS addition to AS +N/D+F

<5-10

<5-10

20-30

3-5

1-2

<1

0.3-3

BP

10-20

5-15

20-35

15-25

5-10

<2

5-10

BNP+F

<10

<5

20-30

<5

<2

<1

0.3-3

a AS = activated sludge; F = granular-medium filtration; CA = carbon adsorption-N = nitrification; D = denitrification; MS = metal salt addition; BP = biological phosphorus removal; and BNP = biological nitrogen and phosphorus removal.

b SS = suspended solids; BOD5 = biochemical oxygen demand over 5 days; COD = chemical oxygen demand; N = nitrogen; NH3 = ammonia; P04 = phosphate; P = phosphorus; and NTU = Nephelometric Turbidity Units.

Source: Adapted from Metcalf & Eddy, Inc., 1991.

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

TABLE 2.8 Organic Constituent Concentrations in Oxidized, Filtered, Disinfected Reclaimed Water at Three Water Reclamation Plants

 

Range of Concentration (µg/l)

Constituent

Whittier Narrows

San Jose Creek

Pomona

Methylene chloride

3.3-5.3

2.1-3.6

ND-3.9

Chloroform

3.5-6.9

3.8-7.4

3.8-5.8

Bromodichloromethane

ND-0.8

ND-1.3

ND-2.0

Dibromochloromethane

ND

ND-0.6

ND-1.2

Bromoform

ND

ND

ND

Carbon tetrachloride

ND

ND

N-0.4

1,1-Dichloroethane

ND

ND

ND

1,2-Dichloroethane

ND

ND

ND

1,1,1-Dichloroethane

0.5-2.3

ND-1.1

ND

1,1,2-Dichloroethane

ND

ND

ND

1,1-Dichloroethylene

ND

ND

ND

Trichloroethylene

ND

ND

ND

Tetrachloroethylene

ND-0.6

ND-33.6

ND-1.1

Benzene

ND

ND

ND

Toluene

ND-0.4

ND-1.2

ND

Chlorobenzene

ND

ND

ND

O-Dichlorobenzene

ND

ND-0.6

ND

M-Dichlorobenzene

ND

ND

ND

P-Dichlorobenzene

0.9-2.2

0.9-1.5

ND-1.1

Trans-1,2-dichloroethylene

ND

ND

ND

Bromomethane

ND

ND

ND

Chloroethane

ND

ND

ND

2-Chloroethylvinylether

ND

ND

ND

Chloromethane

ND

ND

ND

1,2-Dichloropropane

ND

ND

ND

Cis-l,3-dichloropropene

ND

ND

ND

Trans-1,3-dichloropropene

ND

ND

ND

1,1,2,2-Tetrachloropropene

ND

ND

ND

Vinyl chloride

ND

ND

ND

Bis(2-ethylhexyl)phthalate

ND-2

ND-1

ND-5

1,2,4-Trichlorobenzene

ND

ND

ND

2,4,6-Trichlorophenol

ND

ND

ND

2,4,5-Trichlorophenol

ND

ND

ND

2,3,4-Trichlorophenol

ND

ND

ND

2,3,6-Trichlorophenol

ND

ND

ND

3,4,5-Trichlorophenol

ND

ND-3.0

ND

Pentachlorophenol

ND

ND

ND

Phenanthrene

ND

ND

ND

Fluoranthrene

ND

ND

ND

Atrizine

ND

ND

ND

Simazine

ND

ND

ND

Phenylacetic acid

ND-3

ND-1

ND

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

 

Range of Concentration (µg/l)

Constituent

Whittier Narrows

San Jose Creek

Pomona

Phenol

ND-1

ND-2

ND-1

DDT

ND

ND

ND

BHC

ND

ND

ND

Lindane

0.03-0.06

0.03-0.05

0.03-005

Heptachlor

ND-0.01

ND

ND

Heptachlor epoxide

ND

ND

ND

Aidrin

ND

ND

ND

Dieldrin

ND

ND

ND

Endrin

ND

ND

ND

Toxaphene

ND

ND

ND

PCB (Arochlor 1242)

ND

ND

ND

PCB (Arochlor 1254)

ND

ND

ND

Note: ND = none detected.

Source: Bookman-Edmonston Engineering, Inc., 1992.

TABLE 2.9 Physical Properties and Inorganic Constituent Concentrations in Oxidized, Filtered, Disinfected Reclaimed Water at Three Water Reclamation Plants

 

Range of Concentration (mg/l unless otherwise noted)

Constituent

Whittier Narrows

San Jose Creek

Pomona

Total dissolved solids

490-541

603-680

539-603

Electrical conductivity (µmbo/cm)

930-1450

1030-1380

1070-1130

Calcium

40-56

33-75

46-61

Magnesium

13-25

17-22

13-15

Sodium

96-112

120-146

116-132

Potassium

10-13

13-15

11-13

Bicarbonate

229-301

246-342

195-295

Sulfate

79-135

105-190

92-144

Chloride

67-94

131-160

173-211

Hardness

154-208

202-263

173-211

Alkalinity

188-247

202-280

160-242

Fluroide

0.73-1.01

0.40-.063

0.25-0.45

Nitrate (as N)

1..3-3.4

1.95-5.01

1.5-3.9

Nitrite (as N)

0.15-1.11

0.52-1.11

0.43-1.85

Ammonia nitrogen (as N)

13.6-19.9

10.2-15.0

10.7-19.5

Organic nitrogen (as N)

1.4-2.7

1.0-2.2

1.3-3.4

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

 

Range of Concentration (mg/l unless otherwise noted)

Constituent

Whittier Narrows

San Jose Creek

Pomona

Chemical oxygen demand

28-42

29-64

28-42

Biochemical oxygen demand

4-8

4-9

2-6

Total organic carbon

10-15

9-13

10-15

Iron

ND-0.07

0.03-0.09

0.04-0.07

Manganese

ND-0.01

0.02-0.03

ND-0.02

Arsenic

0.001-0.002

0.002-0.004

0.002-0.006

Barium

0.03-0.05

ND-0.06

0.03-0.04

Boron

0.27-0.60

0.16-0.68

0.09-0.57

Cadmium

ND-0.004

ND-0.004

ND-0.003

Chromium (hexavalent)

ND

ND-0.02

ND

Chromium (total)

ND

ND

ND

Copper

ND

ND

ND

Cyanide

ND-0.02

ND-0.03

ND

Lead

ND-0.001

ND-0.002

ND

Mercury

ND-0.0002

ND-0.0006

ND

Nickel

ND-0.04

ND-0.06

ND-0.04

Selenium

0.004-0.010

0.001-0.002

ND-0.004

Silver

ND

ND-0.005

ND

Zinc

0.07-0.13

0.05-0.06

0.02-0.04

pH

6.9-7.1

6.8-7.1

7.0-7.3

Color (CU)

17

15-20

25-46

Temperature (ºF)

69-82

71-81

68-83

Turbidity (T = NTU)

1.1-2.1

0.8-2.1

1.1-2.0

Note: ND = none detected.

Source: Bookman-Edmonston Engineering, Inc., 1992.

ment Plant in Florida was found to contain both Giardia cysts and Cryptosporidium oocysts in 5 of 12 samples analyzed for the organisms. The positive samples had average parasite concentrations of 1 cyst or oocyst/100 liters (CH2M Hill, 1993). The viability of the parasitic cysts was not determined in any of these studies.

Several types of filtration systems are used at municipal wastewater treatment plants including conventional dual-or multi-media filters, mono-medium deep bed filters, automatic-backwashing shallow-bed travelling bridge filters, downflow continuously backwashing filters, and upflow continuously backwashing filters. They are usually designed and operated to achieve low suspended

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

solids or turbidity prior to disaffection. Most types of filters have been demonstrated to be capable of producing effluent meeting an average turbidity limit of 2 NTU (James M. Montgomery Consulting Engineers, Inc., 1979; Matsumoto and Tchobanoglous, 1981; Lang et al., 1986; and Weinschrott and Tchobanoglous, 1986).

Nitrification

Nitrification is the biological conversion of ammonia nitrogen sequentially to nitrite nitrogen and nitrate nitrogen. Nitrification does not remove significant amounts of nitrogen from the effluent; it merely converts it to another form. Nitrification can be done in many suspended and attached growth aerobic treatment processes when they are designed to foster the growth of nitrifying bacteria. In the traditional activated sludge process, it is accomplished by designing the process to provide a retention time for suspended solids that is long enough to prevent the slow-growing nitrifying bacteria from being washed out of the system. Nitrification also occurs in trickling filters that operate at low BOD/TKN (total kjeldahl nitrogen) ratios, either in combination with BOD removal or as a separate advanced process following any type of secondary treatment. A well-designed and well-operated nitrification process can produce an effluent containing 1 mg/l or less ammonia nitrogen. Ammonia nitrogen can also be removed from effluent by several chemical or physical treatment methods, such as air stripping, ion exchange, reverse osmosis, and breakpoint chlorination.

Denitrification

Denitrification removes nitrate nitrogen from the wastewater. As with ammonia removal, denitrification is usually best done biologically for most municipal applications. In biological denitrification, nitrate nitrogen is used by a variety of heterotrophic bacteria as the terminal electron acceptor in the absence of dissolved oxygen (anaerobic conditions). In the process, nitrate nitrogen is convened to nitrogen gas which escapes to the atmosphere. A carbonaceous food source is also required by the bacteria in these processes. Stoichiometrically, about 4 mg of organic carbon is required for every milligram of nitrate to be denitrified. Biological denitrification processes can achieve effluent nitrogen concentrations between 2 mg/l and 12 mg/l nitrate nitrogen. The effluent total nitrogen will be somewhat higher, depending on the concentrations of volatile suspended solids and soluble organic nitrogen present.

Phosphorus Removal

Phosphorus can be removed from wastewater by either chemical or biological methods, or a combination of the two. Chemical phosphorus removal is

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

accomplished by precipitating the phosphorus from solution by the addition of iron, aluminum, or calcium salts. Biological phosphorus removal relies on the aerobic culturing of bacteria that will store excess amounts of phosphorus when exposed to anaerobic conditions in the treatment process. Chemical phosphorus removal can attain effluent orthophosphate concentrations of less than 0.1 mg/l, while biological phosphorus removal will usually produce an effluent phosphorus concentration between 1.0 and 2.0 mg/l.

Carbon Adsorption

One of the most effective advanced wastewater treatment processes for removing biodegradable and refractory organic constituents is the use of granular activated carbon (GAC). GAC can reduce the levels of synthetic organic chemicals in wastewater by 75 to 85 percent. The basic mechanism of removal is by adsorption of the organic compounds onto the carbon. Carbon adsorption preceded by conventional secondary treatment and filtration can produce an effluent with a BOD of 0.1 to 5.0 mg/l, COD of 3 to 25 mg/l, and TOC of 1 to 6 mg/l.

The major organic fraction adsorbed by activated carbon is in the low-molecular-weight range, generally less than 10,000. High-molecular-weight fractions, consisting of mostly polar organic compounds and compounds having molecular weights greater than 50,000, are adsorbed poorly. They can be displaced by more strongly adsorbed organics. Further, organic compounds having low solubilities adsorb better than those having high solubilities in the wastewater. Other molecules having highly branched structures are removed much more slowly than those of identical molecular weight, but with configurations that permit the coiling and compactness that result in high rates of diffusion into the pores of the carbon.

Granular activated carbon will remove several metal ions, particularly cadmium, hexavalent chromium, silver, and selenium. It has also been used to remove unionized species, such as arsenic and antimony, from an acidic stream, and it reduces mercury to low levels, particularly at low pH values.

Reverse Osmosis

Reverse osmosis (RO) is used mainly as a wastewater treatment process to remove suspended and dissolved solids (including microorganisms), either organic or inorganic. Removal is accomplished by the passage of wastewater through a semipermeable membrane. Constituent removal is influenced by the size, shape, chemical characteristics, and concentration of the chemical species as well as the physical and chemical characteristics of the feed water and type of RO unit employed. Because of the nature of the RO process, feed water must be of a fairly high quality (low suspended solids content) to prevent membrane clogging and deterioration.

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

In a homologous series of organic compounds, solute removals (i.e., rejection by the membranes) increases as the number of carbon atoms or molecular weight increases. The removal of organic compounds by RO membranes increases as the degree of molecule branching increases and as the cross-sectional area increases. Organic molecules having molecular weights greater than 200 are rejected by the membrane on the basis of size alone. Depending on membrane pore size, RO can also remove viruses and virtually all larger microorganisms.

The effectiveness of RO in removing inorganic constituents at Orange County Water District's Water Factory 21 is illustrated in Table 2.10. Reverse osmosis can readily remove more than 90 percent of the gross organics and with proper pretreatment can reduce TOC levels to less than 1 mg/l. Table 2.11 provides information on the removal of organic priority pollutants by both GAC and RO at Water Factory 21.

DISINFECTION

The most important process for the destruction of microorganisms is disinfection. Although, the most common disinfectant is chlorine, ozone (O3) and ultraviolet (UV) radiation are other prominent disinfectants used at wastewater treatment plants. Other disinfectants, such as gamma radiation, bromine, iodine, and hydrogen peroxide, have been considered for the disinfection of wastewater but are not generally used bemuse of economical, technical, operational, or disinfection efficiency considerations. Membrane processes (e.g., microfiltration, ultrafiltration, and reverse osmosis) have been shown to be effective in removing microorganisms, including viruses, from municipal wastewater, but again are not commonly used. The strategy in the selection and use of disinfectants for source waters prior to recharge should recognize the possibility that the nature and quantities of the disinfection by-products (DBPs) that may be formed are different from those in conventional water treatment. For example, both chlorine and ozone react in wastewater with organic precursors, which are likely to be greater in number and concentration than in freshwater sources of drinking water, to form DBPs. Accordingly, treatment of water for potable purposes is being modified to minimize the use of oxidizing disinfectants. However, in the treatment of water for nonpotable purposes, the numbers and concentrations of DBPs are of less concern because long-term ingestion is not an issue.

Disinfectants

Chlorine

The efficiency of disinfection with chlorine depends on the water temperature, pH, degree of mixing, time of contact, presence of interfering substances,

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

TABLE 2.10 Removal of Inorganic Constituents by Reverse Osmosis at Water Factory 21, Orange County, California

Constituent

Units

Concentration Before Reverse Osmosis

Concentration After Reverse Osmosis

Total dissolved solids

mg/l

1230

72

Sodium

mg/l

270

24

Potassium

mg/l

20

2.1

Magnesium

mg/l

4.3

< 0.1

Calcium

mg/l

98

1.9

Iron

µ.l

40

40

Manganese

µg/l

12

1.5

Silver

µg/l

0.1

50

Aluminum

µg/l

82

59

Barium

µg/l

23

1.5

Beryllium

µg/l

< 1.0

< 1.0

Cadmium

µg/l

2

< 1.0

Cobalt

µg/l

1.3

< 1.0

Chromium

µg/l

< 1.0

< 1.0

Copper

µg/l

1.3

< 1.0

Mercury

µg/l

< 0.5

< 0.5

Nickel

µg/l

19

1

Lead

µg/l

< 1.0

< 1.0

Selenium

µg/l

< 5.0

< 5.0

Zinc

µg/l

< 50

< 50

Ammonia (as N)

mg/l

23

2.6

Organic nitrogen

mg/l

< 0.1

< 0.1

Total kjeldahl nitrogen

mg/l

23

2.6

Total alkalinity

mg/l

20

13

Total hardness

mg/l

262

5

Fluoride

mg/l

0.5

0.16

Chloride

mg/l

373

40

Nitrate (as N)

mg/l

0.4

0.1

Sulfate

mg/l

431

4.8

Boron

mg/l

0.6

0.5

Silica

mg/l

14

< 1.0

Chemical oxygen demand

mg/l

27

1

Total organic carbon

mg/l

10.3

0.8

MBASa

mg/l

0.24

0.06

Color

color units

13

< 3

Free chlorine

mg/l

< 0.1

< 0.1

Total chlorine

mg/l

< 0.3

< 0.3

Total coliform

CFU/100 ml

< 0.3

< 0.3

Fecal coliform

CFU/100 ml

< 1

< 1

a MBAS = methylene blue active substances, a measure of surfactant concentration.

Source: Wesner, 1991.

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

TABLE 2.11 Removal Organic Priority Pollutants at Water Factory 21, Orange County, California by Granular Activated Carbon (GAC) and Reverse Osmosis (RO)

 

Secondary Effluent Concentration (µg/l)

Concentration Before GACa (µg/l)

Concentration After GAC (µg/l)

Concentration After RO (µg/l)

Chloroform

3.5

7.9

10.2

1.8

Bromodichloromethane

0.46

3.9

4.4

0.7

Dibromochloromethane

0.71

2.1

2.8

0.7

Bromoform

0.46

0.9

0.8

0.3

1,1,1-Trichloroethane

4.8

0.17

0.18

0.05

Trichloroethylene

1.1

0.04

0.06

0.01

Tetrachloroethylene

3.6

0.14

0.71

0.06

Carbon tetrachloride

0.05

0.10

0.21

0.06

Chlorobenzene

0.13

0.13

0.07

0.07

1,3-Dichlorobenzene

0.25

0.03

0.01

0.01

1,4-Dichlorobenzene

1.9

0.14

0.04

0.02

1,2-Dichlorobenzene

0.74

0.09

0.05

0.03

1,2,4-Trichlorobenzene

0.31

0.06

0.01

0.06

Napthalene

0.11

0.13

0.03

0.06

Ethylbenzene

0.04

0.03

0.02

0.04

2,4-Dichlorophenol

0.16

0.01

<0.05

<0.05

2,4,6-Trichlorophenol

0.13

0.05

<0.05

<0.05

Pentachlorophenol

1.23

0.48

<0.05

<0.05

PCB (Arochlor 1242)

0.40

0.00

<0.05

<0.05

Lindane

0.11

0.06

<0.05

<0.05

DDT

0.01

0.05

<0.05

<0.05

Di-n-butyl phthalate

0.94

0.41

0.24

0.20

Diethyl phthalate

1.14

0.59

0.82

0.32

Bis(2-ethylhexyl)phthalate

11

1.2

0.63

1.2

Isophorone

0.30

0.05

<0.05

<0.05

Note: All values are geometric mean concentrations.

a Treatment before GAC includes biological secondary treatment, chemical (lime) clarification, air stripping, chlorination, and filtration.

Source: McCarty et al., 1982.

concentration and form of chlorinating species, and the nature and concentration of the organisms to be destroyed. In general, bacteria are less resistant to chlorine than are viruses, which in turn are less resistant than parasite ova and cysts.

The chlorine dosage required to disinfect a wastewater to any desired level is greatly influenced by the constituents present in the wastewater. Some of the interfering substances are organic constituents, which consume the disinfectant; particulate matter, which protects microorganisms from the action of the disinfectant; and ammonia, which reacts with chlorine to form chloramines, a much

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

less effective disinfectant species than free chlorine. In practice, the amount of chlorine added is determined empirically, based on desired residual and effluent quality. Chlorine, which in low concentrations is toxic to many aquatic organisms, is easily controlled in wastewater by dechlorination, typically with sulfur dioxide or thiosulfate. Chlorine has the disadvantage that it must be handled carefully, and the safety precautions required can be expensive.

Secondary effluent can be disinfected with chlorine to achieve very low levels of coliform bacteria, although complete destruction of pathogenic bacteria and viruses is unlikely to occur. Chlorination of secondary effluent that has received further treatment to remove suspended matter can produce wastewater that is essentially free of bacteria and viruses. Chlorine, at the normal concentrations used in wastewater treatment, may not destroy helminth eggs, Giardia lamblia, and Cryptosporidium species.

Ozone

Ozone is a powerful disinfecting agent and a powerful chemical oxidant in both inorganic and organic reactions. Due to the instability of ozone, it must be generated on site from air or oxygen carrier gas. Ozone destroys bacteria and viruses by means of rapid oxidation of the protein mass, and disinfection is achieved in a matter of minutes. Some disadvantages are that the use of ozone is relatively expensive and energy intensive, ozone systems are more complex to operate and maintain than chlorine systems, and ozone does not maintain a residual in water. Ozone is a highly effective disinfectant for advanced wastewater treatment plant effluent, and it removes color and contributes dissolved oxygen. It also breaks down recalcitrant organic compounds into more biodegradable compounds, which is advantageous for ground water recharge and soil-aquifer treatment.

Ultraviolet Radiation

Irradiation of wastewater with ultraviolet radiation for disinfection is potentially a desirable alternative to chemical disinfection, owing to its inactivating power for bacteria and viruses, affordable cost, and the absence of chemical disinfection by-products. Exposure of microorganisms to the appropriate amount of electromagnetic (EM) radiation disrupts the cells, genetic material and interferes with the reproduction process. Some bacteria have repair enzyme systems that are activated by similar EM energies, and thus disinfected waters may be repopulated by these particular bacteria after disinfection when exposed to light. UV disinfection for water and wastewater is the newest of the disinfection technologies and therefore valuable large scale field applications are still under study. However, the trend is toward more use of UV disinfection.

The effectiveness of UV radiation as a disinfectant where fecal coliform

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

limits are on the order of 200 per 100 ml has been well established, as is evidenced by its use at more than 120 small to medium wastewater treatment plants in the United States (U.S. Environmental Protection Agency, 1986a). An increasing amount of research is being conducted on the ability of UV radiation to achieve high levels of disinfection. In one pilot study, a UV dose of 60 mWs/ cm2, or greater, consistently disinfected unfiltered secondary effluent to a total coliform level of 23 per 100 ml, or less, and a UV dose of at least 97 mWs/cm2 consistently disinfected filtered secondary effluent to a total coliform level of 2.2 per 100 ml, or less (Snider et al., 1991). The study also indicated that filtration, which was effective in removing significant amounts of suspended solids and providing an effluent with a turbidity of less than 2 NTU, enhanced the performance of the UV disinfection. Because water after soil-aquifer treatment has a very low turbidity, UV radiation may be the method of choice in this case.

A recent study applied UV radiation (dose, 47 mWs/cm2) in a full scale pilot plant to a blend of 70 percent secondary municipal wastewater effluent and 30 percent surface water with 90 to 99 percent reductions in total coliform organisms, E. coli, fecal streptococci, Salmonella, and coliphages (Dizer et al., 1993). The inactivating effect was counteracted by binding of the coliphage to suspended particles, prompting the authors to recommend that wastewater be clarified prior to UV disinfection. A similar pilot study investigated both secondary and tertiary effluent and the effects of UV on reclaimed water characteristics, microbial regrowth potential in transport lines, and photoreactivation of bacteria after UV treatment (Chen et al., 1993). At a radiation dose of 100 mWs/cm2, total coliform concentrations were reduced below the 2.2 MPN/100 ml drinking water limit, and doses between 100 and 200 mWs/cm2 had no significant effect on other wastewater characteristics, including TOC, soluble COD, total COD, and chlorine demand. No photoreactivation was observed in samples kept dark after UV exposure, and the extent of reactivation in samples exposed to the light decreased as UV dose increased. Heterotrophic plate count increases in the transport lines after U irradiation were observed, although total coliforms were unchanged.

Studies of virus survival kinetics for UV doses up to 40 mWs/cm2 conclude that UV irradiation can effectively inactivate viruses of public health concern in drinking water (Battigelli et al., 1993). Protozoan cysts (Giardia, Cryptosporidium) and bacterial spores were the most resistant to U-V radiation, but the net public health risk, as noted in Chapter 4, is probably low because these organisms are large enough to be physically removed in most ground water systems. Other recent laboratory and pilot plant studies (CH2M Hill, 1992; Wilson, 1992; Dizer et al., 1993) indicate that UV radiation is very effective in inactivating enteric viruses and coliphages in water and tertiary-treated (i.e., filtered) wastewater. However, the effect of subsequent chlorination on dissolved organic carbon (DOC) surviving UV radiation is still unknown.

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

Disinfection By-Products

Disinfection by-products (DBP) are the chemical transformation products of the disinfection of water to remove pathogens. There are a number of uncertainties regarding disinfection by-product chemistry. These include uncertainties regarding the chemical nature of the reduced carbon macromolecules commonly present in global water systems, the reaction chemistry of these carbon molecules with various oxidant/disinfectants, the toxicological properties of the identified by-products and residual disinfectant chemicals, and the reactivity and stability of the by-products in the soil and underground environment. We do not have an equal understanding of the reaction by-product chemistry for the various disinfectants employed in water treatment. It is safe to say that chlorine has been the most widely used disinfectant, but that within the last 10 years, lower chlorine doses and alternative disinfectants such as monochloramine and ozone have become more popular because of the concerns over the toxicity of chlorine disinfection by-products (Bull and Kopfler, 1991). It is probably true that more is known of chlorine by-products (at least from traditional water sources) than is known for chloramine by-products and that chloramine is less reactive with native carbon. Some current research is focusing on ozone by-product identification and some new halogenated by-products (bromohydrins) are being identified (Cavanaugh et al., 1992 and Shukairy et. al., 1994) resulting presumably from ozone conversion of bromide to hypobromite.

Chlorine Disinfection

Chlorine gas reacts reversibly and rapidly with water-forming aqueous species (HOCl, OCl-, C13-, Cl2O) containing the oxidizing power of the original chlorine molecule. These species can participate in virtually every major class of reaction with organic molecules, and in the natural water environment complex natural product organic molecules (dissolved organic matter, DOM) are commonly available, largely in the form of humic substances, which may account for as much as 40 to 50 percent of the dissolved organic carbon in terrestrial streams. As mentioned earlier, the coincidence in natural waters of appreciable concentrations of reduced carbon and chlorine species added from the disinfection process results in a variety of chlorinated and nonchlorinated reaction products (John et al., 1992 and Christman et al., 1983).

Trihalomethanes (THMs) may or may not be the most concentrated principal chlorination by-product, depending on pH, water source, and chlorine dose. A wide variety of other chlorination by-products have been identified, including haloacids, unhalogenated carboxylic acids, haloacetonitriles, halogenated aldehydes, ketones, and phenols, as well as small, but toxicologically interesting group of halogenated hydroxyfuranones. There is no guarantee that the most important by-products from a health effects standpoint have been identified,

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

because the major, known by-products usually account for 10 percent or less of the total organic halogen at treatment plant dosage levels, and only about 50 percent at laboratory chlorination levels. Most researchers have reported that residues of chlorinated water samples are mutagenic in the Ames histidine reversion assay (Kronberg et al., 1993). Despite this general conclusion, only five mutagenic derivatives of a trichlorinated hydroxyfuranone (MX) have been identified unambiguously as chlorination by-products (MX, EMX, red-MX, ox-MX, ox-EMX), and at the concentration levels at which they are ordinarily produced, they account for only a minor fraction (less than 20 percent) of the overall mutagenicity of whole water residues. Research points to the generally inadequate characterization of the toxicological properties of DBPs and the fact that mutagenicity data, alone, are of little use in the quantitation of risk (Bull and Kopfler, 1991). It could be concluded that we have not identified the majority of mutagens in chlorinated drinking water, nor do we understand the overall importance of their human health risk. On the other hand, chlorine and other disinfectants have played an extremely important role in controlling waterborne infectious disease, and the hazards that could arise from the abolition of disinfection would far outweigh any benefits from reduced toxicological hazards (Bull and Kopfler, 1991).

Monochloramine Disinfection

It is generally believed that monochloramine is inherently less reactive with native organic carbon in water systems than is chlorine with similar DBP formation at lower concentration levels (Jensen et al., 1985). In water treatment industry practice, chloramine is formed via the sequential or simultaneous additions of chlorine and ammonia, and in cases where chlorine is added before ammonia, higher levels of DBPs are observed than for simultaneous addition, or for prior addition of ammonia. In any event, the level of DBP production is considerably less than for disinfection with chlorine. A laboratory study could not detect ether-extractable products following dosing a fulvic acid solution with monochloramine at pH 9 (Jensen et al., 1985). However, THM production is well documented from studies of actual water supplies and other traditional by-products have been measured (Amy et al., 1984). A unique DBP of toxicological significance has been reported, namely, cyanogen chloride, the concentration of which may be increased when ozone is used in conjunction with monochloramine (Krasner et al., 1989).

The similarity of the identified monochloramine DBPs to those from chlorine is probably due to the very slow hydrolysis of monochloramine in water. Monochloramine is not unreactive with organic molecules and has been shown to form nitriles and other products after reaction with aldehydes and ketones, to form chlorophenols from phenol when a reaction time of several days is permitted, and to participate in electrophilic addition to activated olefinic bonds. Al-

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

though it is uncertain whether all DBPs from monochloramine have been detected, it is more certain that the major by-products have been identified (Jensen et al., 1985).

Ozone Disinfection

Ozone is reactive toward reduced organic carbon molecules in water and attacks organic molecules directly as molecular ozone, O3, and via hydroxyl radicals produced from ozone decomposition. Molecular ozone is a more selective oxidant, attacking olefinic bonds (epoxide formation), phenols, and simple amines, whereas hydroxyl radicals will attack a wider variety of organic substrates, including aliphatic acids, ketones, and unactivated aromatic rings. High pH values and ultraviolet radiation promote ozone decomposition favoring the hydroxyl radical pathway, whereas low pH values and the presence of radical scavengers such as carbonate and bicarbonate ions promote the molecular ozone pathway. In water treatment practice, ozone is often used in conjunction with hydrogen peroxide (peroxone process) to increase the generation of hydroxyl radicals.

Ozone produces a variety of disinfection by-products that are generally less oxidized and less halogenated than chlorine by-products, including aldehydes, ketones, carboxylic acids, and unstable peroxides. The smaller molecular weight aldehydes are typically found in greater concentration (5 to 20 micrograms per liter (µg/l) than the larger alkanals (i.e., C6 to C10, approximately 0.1 to 2 µg/l). There have been some reports of aromatic aldehydes, although these may be products from postchlorination (Glaze et al., 1993).

Ozone (like chlorine) oxidizes bromide-ion-producing hypobromous acid (HOBr). Although hypobromous acid is a weaker acid than hypochlorous acid (HOCl), and since the conjugate acid form of the hypohalite ions is a better nucleophile, HOBr reacts more rapidly with organic substrates than does HOCl. Thus, although ozone contains no halogen, brominated disinfection by-products are known, and, in general, brominated derivatives are more worrisome than chlorinated derivatives toxicologically. This circumstance may be fortuitous, however, because consumption of hypobromite at least prevents or reduces its conversion to bromate by residual ozone. Although ozone oxidizes OBr-to BrO3- at a significant rate in dilute aqueous solution, this reaction may be inhibited if radical scavengers are present.

There has not been as much research effort in identifying bromination reaction by-products as there has been for chlorination by-products. It is known, however, that HOBr reacts readily with aquatic humic material, and the formation of bromoform, other brominated THMs, and bromoacetic acids, has been reported (Glaze, 1986 and Cavanaugh et al., 1992). More recently, other brominated DBPs such as bromopicrin, cyanogen bromide, and bromoacetones, have been detected. A new group of labile brominated organic DBPs, the bromo-

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

hydrins have been reported (Cavanaugh et al., 1992). The most abundant brominated alcohol has been assigned the structure of 3-bromo-2-methyl-2-butanol.

There is no direct toxicological concern for ozone because it is not expected to persist to the point of drinking water consumption.

Chlorine Dioxide Disinfection

Most of the reaction by-products of chlorine dioxide with aquatic humic materials appear to be monobasic and dibasic aliphatic acids, although a variety of polybasic aromatic acids have also been reported. Relatively few chlorinated products have been identified other than monochloromalonic, monochlorosuccinic, and dichloroacetic acids, all of which are observed with chlorination (Colclough et al., 1983). Both chiorite and chlorate have been shown to be present in waters disinfected with chlorine dioxide, and owing to the toxicological significance of chlorite (oxidative hemolytic anemia), measures should be taken to remove both chlorine dioxide and chlorite from finished waters prior to ingestion. The relative difficulty in preparing chlorine dioxide, the need for chlorite removal and the unknown consequences of dosing waters with the required reducing agents (themselves of unknown toxicological concern) would appear to obviate consideration of chlorine dioxide for disinfection in ground water recharge programs.

URBAN STORMWATER RUNOFF

Characteristics

Urban stormwater runoff can be a candidate for ground water recharge because of its close proximity to points of use and water supply infrastructure and because substantial water volumes are associated with urban runoff. Unfortunately, some stormwaters from urban areas may be badly polluted, requiring either careful selection of the water to be used for recharge or significant treatment before recharge, especially if the water is to be used for portable purpose. Stormwater runoff is also erratic in timing and quantity. Many studies have investigated urban stormwater runoff quality, with the Environmental Protection Agency's (EPA) Nationwide Urban Runoff Program (NURP) providing the largest and best-known database (U.S. Environmental Protection Agency, 1983). Unfortunately, the extensive analytical results reported by NURP and other studies have not included many of the pollutants that are likely to cause the greatest concern in ground water contamination or recharge studies.

Urban runoff comprises many different flow types. These may include dry weather base flows, urban stormwater runoff, combined sewer overflows (CSOs), and snowmelt. The relative magnitudes of these discharges vary considerably based on a number of factors. Season (especially cold versus warm weather) and

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

land use have been identified as important factors affecting base flow and stormwater runoff quality, respectively (Pitt and McLean, 1986). This section summarizes a number of observations of runoff quality for these different flow types and land uses, along with observations of source area flows contributing to these combined discharges. This information can be used to identify the best urban stormwater runoff candidates for ground water recharge and the ones to avoid.

Land development increases urban stormwater pollutant concentrations and runoff water volumes. Impervious surfaces, such as rooftops, driveways, and roads, reduce infiltration of rainfall and runoff into the ground and degrade runoff quality. The average runoff volume from subdivisions has been reported to be more than 10 times greater than that of typical predevelopment agricultural areas (Madison et al., 1979).

Hydraulic factors affecting runoff water volume (and therefore the amount of water available for recharge) include rainfall quantity and intensity, slope, soil permeability, land cover, impervious area, and depression storage. Research during NURP showed that the most important hydraulic factors affecting urban runoff volume were the quantity of rain and the extent of impervious surfaces directly connected to a stream or drainage system (U.S. Environmental Protection Agency, 1983). Directly connected impervious areas include paved streets, driveways, and parking areas draining to curb and gutter drainage systems, or roofs draining directly to a storm sewer pipe.

Table 2.12 presents a summary of the NURP stormwater data collected from about 1979 through 1982 (U.S. Environmental Protection Agency, 1983). BOD and nutrient concentrations in urban stormwater are relatively close in quality to those of typically treated municipal wastewater. However, as shown later, urban stormwater has relatively high concentrations of bacteria, along with high concentrations of many metallic and some organic toxicants. Land use and source areas (parking areas, rooftops, streets, landscaped areas, and so on) can all have important effects on urban stormwater runoff quality.

Bacterial Characteristics

Most descriptions of bacterial characteristics of urban runoff focus on fetal coliform analysis because of its historical use in water quality standards. Fecal coliform bacterial observations have long been used as an indicator of municipal wastewater contamination and therefore as an indicator of possible pathogenic microorganism contamination (Field and O'Shea, 1993), Fecal streptococcal analyses are also relatively common for urban runoff. Unfortunately, relatively few analyses of specific pathogenic microorganisms have been made for urban runoff.

Pathogenic bacteria have been found in urban runoff at many locations and are probably from several different sources (e.g., Field et al., 1976; Olivieri et al., 1977; Qureshi and Dutka, 1979; Environment Canada, 1980; Pitt, 1983; Pitt

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

TABLE 2.12 Median Stormwater Pollutant Concentrations for All Sites by Land Use (Nationwide Urban Runoff Program, NURP)

 

Residential

Mixed Land Use

Commercial

Open/Nonurban

 

Median

COV

Median

COV

Median

COV

Median

COV 

Biochemical oxygen demand (mg/l)

10

0.41

7.8

0.52

9.3

0.31

-

-

Chemical oxygen demand (mg/l)

73

0.55

65

0.58

57

0.39

40

0.78

Total suspended solids (mg/l)

101

0.96

67

1. 14

69

0.85

70

2.92

Total Kjeldahl nitrogen (µg/l)

1900

0.73

1288

0.50

1179

0.43

965

1.00

Nitrite nitrogen plus nitrate nitrogen (µg/l)

736

0.83

558

0.67

572

0.48

543

0.91

Total phosphorus (µg/l)

383

0.69

263

0.75

201

0.67

121

1.66

Soluble phosphorus (µg/l)

143

0.46

56

0.75

80

0.71

26

2.11

Total Lead (µg/l)

144

0.75

114

1.35

104

0.68

30

1.52

Total Copper (µg/l)

33

0.99

27

1.32

29

0.81

 

 

Total Zinc (µg/l)

135

0.84

154

0.78

226

1.07

195

0.66

Note: COV = coefficient of variation-standard deviation/mean.

Source: U.S. Environmental Protection Agency, 1983.

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

and McLean, 1986; Field and O'Shea, 1993). Table 2.13 summarizes the occurrence of various pathogenic bacterial types found in urban stormwater at Burlington, Ontario; Milwaukee; Baltimore; and Cincinnati. The observed ranges of concentrations and percentage isolations of these biotypes vary significantly from site to site and at the same location for different times. However, many potentially pathogenic bacterial types can be present in urban stormwater runoff. As an example, the occurrence of Salmonella biotypes generally is low, and their reported density is usually less than one organism per 100 ml while Pseudomonas aeruginosa organisms are frequently encountered at densities greater than one thousand organisms per 100 ml (Pitt and McLean, 1986).

Salmonella has been reported in some, but not all, urban stormwaters (Qureshi and Dutka, 1979; Olivieri et al., 1977). Typical concentrations were from 5 to 300 Salmonella organisms per 10 liters. The types of Salmonella found in southern Ontario were S. thompson and S. typhimurium var. copenhagen (Qureshi and Dutka, 1979). Almost all of the urban stormwater samples that had fecal coliform concentrations greater that 2,000 organisms per 100 ml had detectable Salmonella concentrations. However, about 27 percent of the samples having fecal coliform concentrations less than 200 organisms per 100 ml had detectable Salmonella. Other research, however, did not find significant correlations of Salmonella isolations with fecal coliforrn concentrations (Schillinger and Stuart, 1978).

Evidence has been found that Shigella is present in urban runoff and receiving waters and could present a significant health hazard (Olivieri et al., 1977). Shigella species causing bacillary dysentery are one of the primary human enteric disease producing bacterial agents present in water. The infective dose of Shigella necessary to cause dysentery may be quite low (it can be as low as 10 to 100 organisms, although the median infective dose is 10,000 and the infective dose will of course vary depending on the host's age, health, and other factors) (Feachem et al., 1983). Because of this low infective dose and the assumed presence of Shigella in urban waters, it may be a significant health hazard associated with urban runoff.

The most abundant of the potentially pathogenic bacteria that can be found in urban runoff and streams is Pseudomonas aeruginosa (Olivieri et al., 1977). This opportunistic pathogen is widely found in aquatic environments; it is associated with eye and ear infections and is resistant to antibiotics. This biotype has been detected frequently in urban runoff studies in concentrations that may cause potential infections. Typical populations of 1,000 to 10,000 organisms per 100 mL have been frequently reported in urban stormwater (Pitt and McLean, 1986). E. coli also is common in urban stormwater. Viruses may also be important pathogens in urban runoff, although they are usually present at low levels (Olivieri et al., 1977).

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

TABLE 2.13 Microorganisms Found in Urban Stormwater (organisms per 100 ml)

City, Province/State

Catchment/ Land-use

Staphylococcus aureus

Pseudomonas aeruginosa

Samonella

Streptococcus species

Enterovirus (PFU/10 liters)

Others

Reference

Burlington, Ontario

Aldershot Plaza

 

14-3,000

S. seftenberg & S. newport isolated

 

 

Total fungi: 2×104-2×106

Qureshi and Dutka, 1979

 

Malvern Road

 

1-740

100% negative

 

 

Total fungi: 9-400 Heterotroph count: 4×105-2×107

Qureshiand Dutka, 1979

Milwaukee, Wisconsin

Highway runoff

all < 1,000

all < 1,000

45% positive

 

 

 

Gutpa, et al. 1981

Baltimore, Maryland

Bush St.

1,200

20

0.03

560,000 (FS)

6.9

 

Olivieri, et al., 1977

 

Northwood

120

6

0.006

50,000 (FS)

170

 

 

Cincinnati, Ohio

Business district

 

 

 

79% positivea

 

 

Geldreich and Kenner, 1969

 

Residential area

 

 

 

80% positiveb

 

 

Geldreich and Kenner, 1969

 

Rural area

 

 

 

87% positivec

 

 

Geldreich and Kenner, 1969

a Streptrococcal bacteria types found; S. bovis/S. equinus (2%); Atypical S. faecalis (1%); S. faecalis liquifaciens (18%); and S. thompson; 4,500/100 ml.

b Streptococcal bacteria types found: S. bovis/S. equinus (0.5%); Atypical S. faecalis (1%); and S. faecalis liquifaciens (18%).

c Streptococcal bacteria types found: S. bocis/S. equinus (0.5%); Atypical S. facecalis (0.2%); and S. faecalis liquifaciens (12%).

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Important Toxicants

Urban stormwater research has quantified some inorganic and organic hazardous and toxic substances frequently found in urban runoff (U.S. Environmental Protection Agency, 1983; Pitt and McLean, 1986). The NURP data (Table 2.14), collected from mostly residential areas throughout the United States, did not indicate any significant regional differences in the substances detected or in their concentrations (U.S. Environmental Protection Agency, 1983). However, residential and industrial data from Toronto showed significant concentration and yield differences for these two distinct land uses and for dry weather and wet weather urban runoff flows (Pitt and McLean, 1986).

The concentrations of many of these toxic pollutants exceeded the EPA ambient water quality criteria for human health protection by large amounts. As an example, typical standards for polycyclic aromatic hydrocarbons (PAHs) in surface waters used as drinking water supplies are 2.8 nanograms per liter (ng/l) (U.S. Environmental Protection Agency, 1986b). As shown on Table 2.14, urban runoff concentrations of chrysene (600 to 10,000 ng/l), fluoranthene (300 to 21,000 ng/l), phenanthrene (300 to 10,000 ng/l) and pyrene (300 to 16,000 ng/l) (four of the most common PAHs found in urban runoff) have been reported to be from 100 to almost 10,000 times greater than this criterion. Even though most of the PAHs are associated with particulate solids, filterable concentrations of these PAHs are still likely to be many times greater than this criterion.

Table 2.15 lists toxic and hazardous substances that are generally found in more than 10 percent of industrial and residential urban runoff samples analyzed (Galvin and Moore, 1982; U.S. Environmental Protection Agency, 1983; Pitt and McLean, 1986). Available NURP data do not reveal that toxic urban runoff conditions differ significantly in pans of the United States (U.S. Environmental Protection Agency, 1983). The pesticides shown were found mostly in urban runoff from residential areas, while heavy metals and other hazardous materials were much more prevalent in industrial areas. Urban runoff dry weather base flows may also be contributors of hazardous and toxic pollutants. Lindane and dieldrin may be very important in residential dry weather flows, while polychlorinated biphenyls (PCBs) may be very important in industrial dry weather flows. Many of the heavy metals found in industrial urban runoff were high during both dry weather and wet weather conditions.

Contamination by Industrial Wastewater

The potential for toxicants and hazardous materials in industrial areas to contaminate urban stormwater runoff is a serious problem. Inappropriate discharges of industrial wastewaters to storm drainage systems are relatively common and can cause serious contamination. Usually, the wastewater streams from all steps in the manufacturing process are collected in one location prior to

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

TABLE 2.14 Summary of National Urban Runoff Program Priority Pollutant Analyses

 

Frequency of Detection (%)

Range of Detected Concentrations (µg/l)

Pesticide

 

 

 

α-BHC

20

0.0027 to 0.1

 

y-BHC (lindane)

15

0.007 to 0.1

 

Chlordane

17

0.01 to 10

 

α-Endosulfan

19

0.008 to 0.2

Metals and Inorganics

 

 

 

Antimony

13

2.6 to 23

 

Arsenic

52

1 to 51

 

Beryllium

12

1 to 49

 

Cadmium

48

0.1 to 14

 

Chromium

58

1 to 190

 

Copper

91

1 to 100

 

Cyanides

23

2 to 300

 

Lead

94

6 to 460

 

Mercury

10

0.6 to 1.2

 

Nickel

43

1 to 182

 

Selenium

11

2 to 77

 

Zinc

94

10 to 2400

PCBs and Related Compounds (detected in less than 1% of all samples)

 

 

Halogenated Aliphatics

 

 

 

Methylene chloride

11

5 to 15

Ethers (none detected in any of the samples)

 

 

Monocyclic Aromatics (detected in less than 6% of all samples)

 

 

Phenols and Cresols

 

 

 

Phenol

14

1 to 13

 

Pentachlorophenol

19

1 to 115

 

4-Nitro phenol

10

1 to 37

Phthalate Esters

 

 

 

bis(2-ethylhexyl) phthalate

22

4 to 62

Polycyclic Aromatic Hydrocarbons

 

 

 

Chrysene

10

0.6 to 10

 

Fluoranthene

16

0.3 to 21

 

Phenanthrene

12

0.3 to 10

 

Pyrene

15

0.3 to 16

Note: Analyses are based on 121 samples from 17 cities. Only those compounds found in greater than 10% of outfall samples are shown.

Source: U.S. Environmental Protection Agency, 1983.

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

TABLE 2.15 Hazardous and Toxic Substances Found in Urban Runoff

 

Residential Areas

Industrial Areas

Halogenated Aliphatics

 

 

 

1,2,-Dichloroethene

 

x

 

Methylene chloride

 

x

 

Tetrachloroethylene

 

x

Phthalate

 

 

 

Bis(2-ethylene)phthalate

x

 

 

Butylbenzyl phthalate

x

x

 

Diethyl phthalate

 

x

 

Di-N-butyl phthalate

x

x

Polycyclic Aromatic Hydrocarbons

 

 

 

Phenanthrene

 

x

 

Pyrene

 

x

Other Volatiles

 

 

 

Benzene

x

x

 

Chloroform

 

x

 

Ethylbenzene

 

x

 

N-Nitro-sodimethylamine

 

x

 

Toluene

 

x

Heavy Metals

 

 

 

Aluminum

x

x

 

Chromium

 

x

 

Copper

x

x

 

Lead

x

x

 

Zinc

x

x

Pesticides and Phenols

 

 

 

BHC

x

 

 

Chlordane

x

 

 

Dieldrin

x

 

 

Endosulfan sulfate

x

 

 

Endrin

x

 

 

Isophorone

x

 

 

Methoxychlor

x

 

 

PCB-arochlor 1254

 

x

 

PCB-arochlor 1260

 

x

 

Pentachlorophenol

x

x

 

Phenol

x

x

Note: Substances were found in at least 10 percent of the stormwater samples analyzed.

Sources: Galvin and Moore, 1982; U.S. Environmental Protection Agency, 1983; Pitt and McLean, 1986.

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

treatment and disposal. These wastewaters originate from many different areas of the plant and/or steps of the industrial process. Activities and areas of a plant that are likely to discharge contaminated wastewater into the storm drainage system include loading and unloading operations, outdoor storage or processing, cooling or process wastewater discharges, particle-generating processes, and illicit or inadvertent connections to the storm drainage system (Pitt et al., 1993). Industrial runoff usually is much more polluted by toxicants than runoff from other land uses. Many of these compounds may present serious problems in the operation of a ground water recharge facility and in ensuring the safety of the water for later reuse. In addition, the behavior of many of these compounds in the soil-land-aquifer system is not well known. Industrial stormwater should therefore not be considered an appropriate source water for ground water recharge.

Relative Contributions of Different Flow Periods

Tables 2.16 and 2.17 summarize residential/commercial and industrial urban runoff characteristics during both warm and cold weather in Toronto (Pitt and McLean, 1986). These tables show the relative importance of wet weather and dry weather flows coming from separate urban stormwater systems. Possibly 25 percent of all separate urban stormwater outfalls have water flowing in them during dry weather, and as many as 10 percent of them are grossly contaminated with raw municipal wastewater and industrial wastewaters (Pitt et al., 1993). EPA's Stormwater Permit program requires municipalities to conduct urban stormwater outfall surveys to identify, and then correct, inappropriate discharges into separate storm drainage. However, substantial outfall contamination probably will exist for many years. If urban stormwater is recharged before it enters the drainage system (such as by using French drains, infiltration trenches, grass swales, porous pavements or percolation ponds in upland areas), the effects of contamination problems in the drainage system on ground water recharge will be reduced substantially. If outfall waters are to be recharged in larger regional facilities, then these periods of dry weather flows will have to be considered.

Similar problems occur in areas having substantial snowfalls. Table 2.18 (Part A and B) summarizes Toronto snowmelt and cold weather base flow characteristics (Pitt and McLean, 1986). The bacteria densities during cold weather are substantially less than during warm weather, but are still relatively high (U.S. Environmental Protection Agency, 1983). However, chloride and dissolved solids concentrations are much higher during cold weather. Early spring urban stormwater events also contain high dissolved solids concentrations (R. Bannerman, personal communication, 1993). Upland infiltration devices do not work well during cold weather because of freezing soils. Outfall flows occur under ice into receiving waters (including detention ponds) and may enter regional ground water recharge devices if not specifically diverted.

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

TABLE 2.16 Median Concentrations Observed at Toronto Outfalls During Warm Weather

 

Warm Weather Baseflow

Warm Weather Stormwater

 

Residential

Industrial

Residential

Industrial

Stormwater volume (m3/ha/season)

 

 

950

1,500

Base flow volume (m3/ha/season)

1,700

2,100

 

 

Total residue

979

554

256

371

Filterable residue (TDS)

973

454

230

208

Particulate residue (SS)

< 5

43

22

117

Chlorides

281

78

34

17

Total phosphorus

0.09

0.73

0.28

0.75

Phosphates

< 0.06

0.12

0.02

0.16

Total Kjeldahl nitrogen

(organic nitrogen plus ammonia)

0.9

2.4

2.5

2.0

Ammonia nitrogen

< 0. 1

< 0.1

< 0.1

< 0.1

Chemical oxygen demand (COD)

22

108

55

106

Fecal coliform bacteria (no./100 ml)

33,000

7,000

40,000

49,000

Fecal streptoccal bacteria (no./100 ml)

2,300

8,800

20,000

39,000

Pseudomonas aeruginosa bacteria (#100 ml)

 

 

 

 

 

2,900

2,380

2,700

11,000

Arsenic

< 0.03

< 0.03

< 0.03

< 0.03

Cadmium

< 0.01

< 0.01

< 0.01

< 0.01

Chromium

< 0.06

0.42

<0.06

0.32

Copper

0.02

0.05

0.03

0.06

Lead

< 0.04

< 0.04

< 0.06

0.08

Selenium

< 0.03

< 0.03

< 0.03

< 0.03

Zinc

0.04

0.18

0.06

0.19

Phenolics (µg/l)

< 1.5

2.0

1.2

5.1

α-BHC (ng/l)

17

< 1

1

3.5

y-BHC (lindane) (ng/l)

5

< 2

< 1

< 1

Chlordane (ng/l)

4

< 2

< 2

< 2

Dieldrin (ng/l)

4

< 5

< 2

< 2

Pentachlorobiphenol (PCB) (ng/l)

< 20

< 20

< 20

33

Pentachlorophenol (PCP) (ng/l)

280

50

70

705

Note: Concentrations are given in milligrams per liter unless otherwise indicated.

Source: Pitt and McLean, 1986.

Pollutant Contributions from Different Urban Source Areas

Sheet flow quality data are available from studies conducted in California, Washington, Nevada, Wisconsin, Illinois, Ontario, Colorado, New Hampshire, New York, and Alabama since 1979. A relatively large amount of parking and roof runoff quality data has been obtained from all of these locations, but only a

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

TABLE 2.17 Median Concentrations Observed at Toronto Outfalls During Cold Weather

 

Cold Weather Baseflow

Cold Weather Snowmelt

 

Residential

Industrial

Residential

Industrial

Stormwater volume (m3/ha/season)

 

 

1,800

830

Base flow volume (m3/ha/season)

1,100

660

 

 

Total residue

2,230

1,080

1,580

1,340

Filterable residue (TDS)

2,210

1,020

1,530

1,240

Particulate residue (SS)

21

50

30

95

Chlorides

1,080

470

660

620

Total phosphorus

0.18

0.34

0.23

0.50

Phosphates

< 0.05

< 0.02

< 0.06

0.14

Total Kjeldahl nitrogen

(organic nitrogen plus ammonia)

1.4

2.0

1.7

2.5

Ammonia nitrogen

< 0.1

< 0.1

0.2

0.4

Chemical oxygen demand (COD)

48

68

40

94

Fecal coliform bacteria (no./100 ml)

9,800

400

2,320

300

Fecal streptoccal bacteria (no./100 ml)

1,400

2,400

1,900

2,500

Pseudoomonas aeruginosa bacteria (#100 ml)

85

55

20

30

Cadmium

< 0.01

< 0.01

< 0.01

0 01

Chromium

< 0.01

0.24

< 0.01

0.35

Copper

0.02

0.04

0.04

0.07

Lead

< 0.06

< 0.04

0.09

0.08

Zinc

0.07

0.15

0.12

0.31

Phenolics (µg/l)

2.0

7.3

2.5

15

α-BHC (ng/l)

NA

3

4

5

y-BHC (lindane) (ng/l)

NA

NA

2

 

Chlordane (ng/l)

NA

NA

11

2

Dieldrin (ng/l)

NA

NA

2

NA

Pentachlorobiphenol (PCB) (ng/l)

NA

NA

NA

40

Note: Concentrations are given in milligrams per liter unless otherwise indicated. N.A. = not analyzed.

Source: Pitt and McLean, 1986.

few of these studies evaluated a broad range of source areas or land uses. However, there is adequate information to identify which upland areas are preferable for use as sources for ground water recharge and which ones to avoid because of excessive contamination. The major urban source area categories that have been studied include the following:

  • roofs,

  • paved parking areas,

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

TABLE 2.18 (Part A) Toronto Cold Weather Snowmelt Source Area Sheet flow Quality (median observed concentrations, mg/l)

Source Area

Total Solids

Filterable Solids

Suspended Solids

Reactive Chlorides

Total Phosphorus

Phosphates

Total Kjeldahl Nitrogen

Ammonia

Chemical Oxygen Demand

Industrial

 

 

 

 

 

 

 

 

 

Pervious areas

 

 

 

 

 

 

 

 

 

Grass/open areas

390

282

77

100

0.33

0.10

1.4

<0.1

47

Unpaved storage/parking

2,925

1000

2105

113

1.1

0.46

5.3

0.2

160

Impervious areas

 

 

 

 

 

 

 

 

 

Sidewalks

1,050

200

847

48

0.45

0.20

1.6

<0.1

63

Paved parking, storage, etc.

1,690

349

392

260

0.55

0.18

3.8

<0.1

135

Road gutters

1,320

575

625

230

0.60

0.15

1.8

<0.1

230

Residential/Commercial

 

 

 

 

 

 

 

 

 

Pervious areas

 

 

 

 

 

 

 

 

 

Grass/open areas

94

78

40

4.0

0.29

0.20

1.2

0.4

26

Impervious areas

 

 

 

 

 

 

 

 

 

Sidewalks

390

29

281

6.4

0.63

0.38

2.6

2.6

98

Paved parking, driveways, etc.

918

274

380

81

0.64

0.08

2.5

<0.1

110

Paved roads

890

166

284

56

0.30

0.06

1.8

<0.1

140

Road gutters

530

190

152

25

0.54

0.28

2.3

<0.1

66

Roadside grass swales

380

155

50

37

0.59

0.17

1.8

0.1

40

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

TABLE 2.18 (Part B) Toronto Cold Weather Snowmelt Source Area Sheet Flow Quality (median observed concentrations, mg/l)

Source Area

Fecal Coliforms

Fecal streptococcaa

Pseudomonas aeruginosa

Cadmium

Chromium

Copper

Lead

Zinc

Phenolics µ/l

Industrial

 

 

 

 

 

 

 

 

 

Pervious Areas:

 

 

 

 

 

 

 

 

 

Grass/open areas

< 20

100

<20

<0.005

0.01

 

0.01

0.06

3.0

Unpaved storage/parking

< 100

100

< 20

0.011

0.07

0.13

0.26

0.51

9.0

Impervious Areas:

 

 

 

0.11

 

 

 

 

 

Sidewalks

< 50

< 50

< 20

< 0.005

0.02

0.05

0.09

0.47

3.7

Paved parking, storage, etc.

< 100

450

< 20

< 0.005

0.02

0.12

0.20

0.40

4.0

Road gutters

< 100

100

< 20

< 0.005

0.05

0.45

0.66

9.0

 

Residential/Commercial

 

 

 

 

 

 

 

 

 

Pervious Areas:

 

 

 

 

 

 

 

 

 

Grass/open areas

< 20

350

< 10

< 0.005

< 0.01

< 0.01

0.04

0.02

1.4

Impervious Areas:

Areas:

 

 

 

 

 

 

 

 

Sidewalks

75

600

< 20

< 0.005

< 0.01

0.02

0.15

0.16

1.4

Paved parking, storage, etc.

< 20

200

10

< 0.005

0.02

0.04

0.23

0.23

2.6

Paved roads

50

200

< 10

< 0.005

0.01

0.05

0.26

0.26

3.2

Road gutters

60

4,00

<10

< 0.005

0.01

0.02

0.12

0.09

1.8

Roadside grass swales

60

1,00

<10

< 0.005

< 0.01

0.01

0.05

0.08

1.6

a Fecal streplococca measured in number per 100 ml.

Source: Pitt and McLean 1986

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
  • paved storage areas,

  • unpaved parking and storage areas,

  • driveways,

  • streets,

  • landscaped areas,

  • undeveloped areas,

  • freeway paved lanes and shoulders, and

  • vehicle service areas.

Table 2.19 summarizes data describing urban area runoff pollutants from these source areas for different land uses and seasons. Snowmelt waters, especially from industrial source areas, are shown to be contaminated with many heavy metal pollutants, such as lead, zinc, and copper. Lead and zinc concentrations are generally the highest in sheet flows from paved parking areas and streets, with some high zinc concentrations also found in roof drainage samples. High bacterial populations have been found in sidewalk, road, and some bare ground sheet flow samples (collected from locations where dogs would most likely be ''walked"). Bacterial levels are much lower during the cold season, but are still higher than most criteria allow.

Some of the sheet flow contributions observed at these locations were not sufficient to explain the concurrent concentrations of the same constituents observed in runoff at the outfall. The low chromium surface sheet flow concentrations at the Toronto industrial area occurring, at the same time as higher outfall chromium concentrations, indicated a high likelihood for direct industrial wastewater connections to the storm drainage system, for example. Similarly, most of the fecal coliform populations observed in sheet flows were also significantly lower than those observed at the outfall.

Treatment Methods

Stormwater runoff has been treated for reuse successfully in several U.S. cities. Many processes affect fate and removal mechanisms of pollutants in treatment facilities (Callahan et al., 1979). Sedimentation is the most common control mechanism for particulate-bound pollutants which typically are the constituents of concern in stormwater runoff. Exceptions include salt, zinc, 1,3-dichlorobenzene, fluoranthene, and pyrene, which may be mostly associated with the filterable sample portions of stormwater. Particulate removal can occur in many control processes, including catch basins, vegetation filtration, swirl concentrators, screens, drainage systems, and detention ponds.

Biological or chemical degradation of the toxicants may occur, but is quite slow for many of the pollutants in anaerobic environments. Degradation of soluble pollutants during treatment may occur, especially near the surface in aerated waters. Volatilization is also a mechanism that may affect many organic

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

Table 2.19 Birmingham Source Area Sheet Flow Quality

 

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

 

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

toxicants. Increased turbulence and elevated oxygen levels promote these processes. Sorption of pollutants onto solids or metal precipitation increases the sedimentation potential of pollutants and also encourages more efficient bonding of the pollutants in soils, preventing their leaching to ground water. The following discussion summarizes observed pollutant removals obtained in stormwater control devices.

Sedimentation Treatment

Wet Detention Ponds and Artificial Wetlands. Detention ponds are probably the most common management practice for the control of stormwater runoff. If properly designed, constructed, and maintained, they can be very effective in controlling a wide range of pollutants. Artificial wetlands are being proposed as a method for stormwater control, especially in conjunction with wet detention ponds. However, performance data are extremely limited at this time.

There are many kinds of detention ponds, including dry ponds (which typically contain no water between storms), wet ponds (which contain standing water between storms), and combination ponds (which drain slowly after storms and may contain a small permanent pool). In a survey of cities in the United States and Canada, the American Public Works Association found more than 2,000 wet ponds (about half of which were publicly owned), more than 6,000 dry ponds, more than 3,000 parking lot multiuse detention areas, and more than 500 rooftop storage facilities (Smith, 1982).

Detention ponds have been required for some time in selected areas of the United States and are therefore more numerous in certain regions than in others. In Montgomery County, Maryland, for example, detention ponds were first required in 1971; more than 100 facilities were planned during that fast year, and about 50 were actually constructed. By 1978, more than 500 detention facilities had been constructed in the county (Williams, 1982). In DuPage County, Illinois, near Chicago, more than 900 stormwater detention facilities (some natural) receive urban runoff (McComas and Sefton, 1985).

The Nationwide Urban Runoff Program (NUPR) included full-scale monitoring of nine wet detention ponds (U.S. Environmental Protection Agency, 1983). About 150 storm events were comprehensively monitored at these ponds, and performances ranged from negative removals for the smallest up-sized pipe installation to more than 90 percent consistent removals of suspended solids at the largest wet ponds. The best ponds reported BOD and COD removals of about 70 percent, nutrient removals of about 60 to 70 percent, and heavy metal removals of about 60 to 95 percent.

Catchbasin and Sewer Cleaning. The mobility of catch basin sediments was investigated using particulate fluorescent tracers mixed with catch basin sediment (Pitt, 1979). The amount of sediment in catch basins (and on streets) and

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

the sewer system at any time was large in comparison with individual storm runoff yields, but was not very mobile. Cleaning the material from catch basins reduces the potential of very large discharges during rare scouring rains and enables additional material to be captured.

Further research in Bellevue, Washington, investigated the accumulation rate of sediment in storm sewers and the effects of sewer cleaning on runoff discharges (Pitt, 1984). The main source of the sediment in the catch basins and the sewer system was found to be the street surfaces. A few unusual locations were dominated by erosion sediment originating from steep hillsides adjacent to the storm sewer inlets. The catch basin and sewer sediment consisted of the largest particles that were washed from the streets. Smaller particles that had washed from the streets during rains proceeded into the receiving waters, leaving behind the larger particles.

Catch basin sump particulates can be removed to eliminate this potential source of urban runoff pollutants. Cleaning catch basins twice a year was found to allow the catch basins to partially capture particulates for most rains. This cleaning schedule reduces the total solids and lead urban runoff yields by between 10 and 25 percent, and COD, total Kjeldahl nitrogen, total phosphorus, and zinc by between 5 and 10 percent (Pin and Shawley, 1981; Pitt, 1984).

Fate of Pollutants in Sedimentation Facilities. The major fate mechanism in wet detention ponds, and in smaller sumps such as catch basins, is sedimentation. Unfortunately, sedimentation results in the accumulation of polluted sediments. These sediments can be anaerobic, with associated chemical and biochemical transformations. Resulting toxic chemical releases from heavily polluted sediments, plus the potential problems associated with the disposal of contaminated dredging spoils during maintenance, can present problems.

Other important fate mechanisms possible in wet detention ponds and wetlands, but probably not important in small sump devices, include volatilization and photolysis. Biodegradation, biotransformation, and bioaccumulation (into plants and animals) may also occur in ponds. Most wet detention ponds are completely flushed by moderate rains (probably every several weeks), depending on their design. Much of the runoff during moderate and large rains passes through the ponds over several hours during and immediately after rains. Sediments may reside in ponds for several to many years. Therefore, the time available for these other removal or transformation processes can vary greatly for different detention ponds. The residence time in small sedimentation devices is just a few minutes, and significant biological activity is not likely, except for some long-residing sediments that may become anaerobic in catch basin sumps.

Most sedimentation devices (especially ponds) are designed to provide effective sedimentation with sufficient storage for the long-term maintenance of accumulated sediment. The removal rates of many toxicants by other processes

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

can possibly be increased by increasing water mixing, oxygen content, and biological activity in ponds.

Pollutant Removal During Infiltration

Grass filter strips can be quite effective in removing particulate pollutants from overland flows. The filtering effects of grasses, along with increased infiltration and recharge, reduce the particulate sediment load from urban landscaped areas. Grass filters can help reduce the particulate pollutant yields to the storm drainage system. Specific situations may include directing roof runoff to grass areas instead of pavement, planting grass between eroding slopes and the storm drainage system, and planting grass between paved or unpaved parking or storage areas and the drainage system.

Grass swale drainages are a type of infiltration device and can be used in place of concrete curb and gutters in most land uses, except possibly strip commercial and high density residential areas. Grass swales allow the recharge of significant mounts of surface flows while also providing some pollutant trapping in vegetation and surface soil. Because they are basically an infiltration device, swales should not be used in industrial areas because of the threat of ground water contamination.

Several large-scale urban runoff monitoring programs have included test sites that were drained by grass swales. For instance, one study of an area with poorly drained soils showed significantly lower surface flows (up to 95 percent lower) compared to a curb and gutter drained area (Bannerman et al., 1979). In another study, a special swale was constructed to treat runoff from a commercial parking lot. Soluble and particulate heavy metal (copper, lead, zinc, and cadmium) concentrations were reduced by about 50 percent. COD, nitrate nitrogen, and ammonia nitrogen concentrations were reduced by about 25 percent, while no significant concentration reductions were found for organic nitrogen, phosphorus, and bacteria (U.S. Environmental Protection Agency, 1983).

IRRIGATION RETURN FLOW

Characteristics

Irrigation return flow is the drainage water (surface and subsurface) collected from irrigated farmland. Because of the many types of crops irrigated, the wide variety of agricultural chemicals applied, the varying quality of supply water, and the different physical and chemical characteristics of soils, it is difficult to characterize the physical and chemical quality of irrigation return flow in any general way. Moreover, there is a paucity of data on the quality of irrigation return flow, except for data on dissolved solids and nitrate concentrations. Recent studies have addressed the content of selected pesticide residues and trace

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

Stormwater Infiltration at Fresno, California

The nationwide Urban Runoff program (NURP) (U.S. Environmental Protection Agency, 1983) was conducted in the early 1980's to gather data on stormwater pollutant quality and quantity throughout the nation and to obtain information concerning the effectiveness of different stormwater management practices (harrison, 1984) One site selected for study was Fresno, California Fresno had 74 recharge basias, with a total infiltration area of over 1,000 acres. The basins provided local stormwater drainage and helped recharge the local aquifer (which is a sole source aquifer). Five basins were studied; three were mosity grass-lined, while two were mostly bare dirt-lined. Two of the basins had been in operation for over 20 years. Stomwater from four land uses (medium-density residential, high-density residential, commercial shopping center, and moderate industrial) was sampted. Pollutant analyses were conducted on rainfall, dry atmospheric deposition, street dirt, stormwater, recharge basin soils, vadose water, and ground water.

Industrial stormwater runoff was the most severely polluted and fluctuated greaty during the storms. Runoff quality from other land uses, as expected, had the worst quality during the beginning of storms and early in the rainy season. The study found that soils in the recharge basin provided a high degree of removal of most of the stormwater pollutants. A correlation was found between the degree of metal removal and the proportion of silt plus clay and organic matter in the surface soils. The only organic compound found in the ground water beneath the basins and in the basin soils. Lead and chlordane were the most commonly detected soil contaminants. Some of these pollutants also were found in the regional ground water. Other pollutants of potential concent in studies of ground water contamination by stormwater, but not monitored during the Fresno project, include bacterial and viral pathogens.

Although there was evidence of downward movement of some pollutants in the soils, the study concluded that there were no apparent adverse effects on ground water quality from infiltration stormwater, especially with cleaner water being used for infiltration during dry months. The stormwater runoff had better, mineral quality (as indicated by specific conductance) and lower nutrients than the regional ground water.

This study did not investigate recharge basins receiving mostly industrial waters because the researchers felt that more data would be needed before industrial water recharge effects could be known. There was some concern that lead might create problems in basins used for recreation, but it was felt that the exposure would be reasonably low. As a precaution, periodic replacement of turf or soil (with proper disposal) could be conducted in basins having recreational uses to reduce potential exposure and the possibility of leaching of contaminants from the soil in future years.

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

elements in irrigation return flow, but the data are incomplete and comprehensive summary assessments are not yet available. From the information available, the principal water quality variables of irrigation return flow with respect to its possible use as a source for ground water recharge are dissolved salts, fertilizer and/or pesticide residues, trace elements leached from soils, and suspended-solids load.

Drainwater from irrigated fields historically has been known to affect the quality of water in receiving streams, reservoirs, and wetlands by increasing concentrations of dissolved solids and major constituents (Engberg et al., 1991). In the last decade or so, attention has been directed toward trace constituents in irrigation drainage because of the linkage made between elevated concentrations of selenium in drainwater and the damage to the waterfowl and shorebird population at Kesterson Reservoir in California. In 1985 the Department of the Interior (DOI) initiated the National Irrigation Water Quality Program, a five-phase research program to identify water quality problems caused by irrigation drainwater, through reconnaissance and detailed studies, and to plan and implement remediation actions. The reconnaissance study phase involved sampling of water, bottom sediment, and biota at 26 areas in western states, before, during, and after the irrigation season (Engberg et al., 1991). Samples of each medium were analyzed for major constituents; trace elements including arsenic, barium, boron, cadmium, chromium, copper, lead, mercury, molybdenum, nickel, selenium, silver, uranium, vanadium, and zinc; and pesticide residues in some cases. These studies should provide a consistent database useful for characterizing the quality of irrigation drainwater, but currently only limited summary results are available in the literature.

Overall, research at the seven reconnaissance sites showed that selenium was the constituent most frequently detected at elevated concentrations in wetland ecological systems (Deason, 1989). Also, the concentrations of constituents were found to vary widely on a spatial basis and, therefore, irrigation induced contamination problems are likely to be very site specific. In addition to the selenium, boron, arsenic, uranium, and mercury found at elevated concentrations, one pesticide residue, DDE, was frequently found at elevated concentrations (Feltz et al., 1990). Closed drainage basins, especially terminal ponds, wetlands, and playas, had the highest concentrations of dissolved solids and specific constituents of concern.

The great variation in irrigation return flow quality can be seen by inspecting results reported from several widely scattered sites in the western United States. From a study of return flow from the Milk River Irrigation Project in northeastern Montana, Lambing et al. (1988) concluded that, "Results of the current study indicate that irrigation drainage sampled in 1986 had relatively small concentrations of most constituents. The only significant differences between the supply water (Milk River) and irrigation drainage were zinc (56 µg/l)

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

and uranium (13 µg/l) concentrations that were several times larger in one of the irrigation drains."

For another DOI study site, the Angostura Reclamation Unit in southwestern South Dakota, Greene et al. (1990) reported that "irrigation return flow had relatively small concentrations of trace elements. Overall, there appeared to be minor differences between concentrations of trace elements in water of the Cheyenne River upstream of irrigated land and in water downstream from all irrigation return flow."

For yet another DOI site, the Fallon agricultural area in west central Nevada, Hoffman et al. (1990) reported that irrigation drainwater had a specific conductance that ranged from 566 to 41,000 microsiemens per centimeter (µS/cm) at 25ºC with a median of 1,990 µS/cm. In contrast, the source water had a conductance that ranged from about 200 to 400 µS/cm with a median of about 250 µS/ cm. These data indicate an eight to tenfold increase in dissolved solids in irrigation drainwater compared to the source water. With respect to trace elements, they report that dissolved aluminum, barium, cadmium, chromium, copper, lead, lithium, molybdenum, nickel, silver, and vanadium were present in concentrations either below Nevada criteria for the protection of aquatic life or propagation of wildlife, or below the analytical reporting level. On the other hand, arsenic (range 1 to 190, median 44 µg/l), boron (range 190 to 28,000, median 2,200 µg/l), and uranium (up to 300 µg/l) were notably high in irrigation drainwater. Although some enrichment of nitrogen and phosphorus occurred during the irrigation process, drainwater concentrations of these constituents were typically less than 1 mg/l.

Other studies conducted in California and Colorado further illustrate the variability of irrigation return flow. Tanji (1981) reported on irrigation return flow quality for the Glenn-Colusa Irrigation District and the Panoche Drainage District in the northern and central part of the Central Valley of California, respectively. He showed an increase in total dissolved solids (TDS) of 2 to 10 times, suspended solids of 1.5 to 5 times, nitrate-nitrogen of 1.3 to 20 times, and boron of almost 40 times over the concentrations in the supply water. The constituent concentrations in the irrigation return flow were as high 2,050 mg/l for TDS, 348 mg/l for suspended solids; nitrate-N: 12.6 mg/l for nitrate-nitrogen; and 4.3 mg/l for boron.

Keys (1981) reported the irrigation return flow quality contrasted with the supply water quality for the Grand Valley area in west central Colorado. TDS concentrations remained about the same in the surface irrigation return flow as in the supply water (about 400 mg/l), but increased by about 10 times (to 4,100 mg/ l) in the subsurface irrigation return flow water. Nitrate-nitrogen increased from 0.3 mg/l in the source water to 4 mg/l in the combined surface and subsurface irrigation return flow. Phosphates did not change in the return flow, but remained the same as in the supply water (0.05 mg/l).

Both the California and the Colorado cases indicate a significant elevation

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

in constituent concentrations or physical property in the return flow compared to the supply water, especially for the subsurface drainage. However, the extent of enrichment through the irrigation process and the concentrations in the drainwater differ widely.

Analysis of San Luis drainwater in the San Joaquin Valley of California indicated the drainwater there to be high in dissolved solids (about 10,000 mg/l), sulfate (about 5,000 mg/l), sodium (about 2,200 mg/l), and chloride (about 1,500 mg/l) (Lee, 1990). Of the trace elements, boron (14.4 mg/l) and selenium (0.3 mg/l) are notably high. Nitrate/nitrite as N was also high, averaging 48 mg/l. The average and maximum concentrations of various constituents in the drainwater are given in Table 2.20.

Agricultural drainage water from the San Joaquin Valley contained an average of 20 mg/l of nitrate-nitrogen but some areas had concentrations of 100 to 200 mg/l of (Bouwer, 1987). Total concentrations of other forms of nitrogen rarely exceeded 1 mg/l. The average concentration of phosphorus was 0.09 mg/ and salt (TDS) was 3,625 mg/l. The average concentrations of nitrate-nitrogen were 19 mg/l, and TDS 3,600 mg/l for tile drainage systems in the San Joaquin Valley for the period from 1962 to 1969 (Schmidt and Sherman, 1987).

Research on pesticide concentrations in surface irrigation runoff water and in tile drain effluents following application of pesticides to large fields of cotton, sugarbeets, alfalfa, lettuce, onions, and cantaloupes in the Imperial Valley, California, found that concentrations of pesticides in surface runoff water were dependent on the characteristics of the pesticides, their methods and rates of applications, the time elapsed between application and the first irrigation, the number of irrigation cycles since the pesticide application, irrigation efficiency, and other soil management practices (Spencer et al., 1985). Seasonal totals of insecticide in surface runoff were less than 1 percent of the mount applied, whereas herbicides in surface runoff were usually 1 to 2 percent of the mounts applied. None of the pesticides were identified in tile drain effluents at concentrations above minimum detectable levels of 1 to 2 ng/l. The highest mean concentrations calculated for the various pesticides in irrigation runoff water are listed in Table 2.21.

Several studies investigated pesticide residues in surface and tile drainage water (Pierce and Wong, 1988). These studies indicated (1) atrazine residue mean concentrations of about 14 to 56 µg/l and maximum of 1,100 µg/l from tailwater pits in Kansas cornfields; (2) alachlor, cyanazine, propazine, and terbutryne present in water and sediment from the same tailwater pits; (3) terbacil and 2,4-D at concentrations of 10 and 110 µg/l, respectively, in drainage water from citrus groves; (4) residues of the triazine herbicides, atrazine, cyanazine, cipiazine, and metabuzine in tile drainwater from Quebec cornfields; and (5) atrazine residues in tile drainwater from Ontario cornfields.

Leaching of nitrate and, to a lesser extent, pesticides to ground water beneath irrigated fields has been observed in many states (Law, 1987; Sabol et al.,

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

TABLE 2.20 Drainage Water Analysis for the San Luis Drain, Mendota, California

Constituent

Units

Average

Maximum

Sodium

mg/l

2,230

2,820

Potassium

mg/l

6

12

Calcium

 

mg/l

554

714

Magnesium

mg/l

270

326

Alkalinity (as calcium carbonate)

mg/l

196

213

Sulfate

mg/l

4,730

6,500

Chloride

 

mg/l

1,480

2,000

Nitrate/nitrite (as N)

mg/l

48

60

Silica

 

mg/l

37

48

Total dissolved solids

mg/l

9,820

11,600

Suspended solids

mg/l

11

20

Total organic carbon

 

mg/l

10.2

16

Chemical oxygen demand

mg/l

32

80

Biochemical oxygen demand

mg/l

3.2

5.8

Temperaturea

 

ºC

19

29

pH

 

8.2

8.7

Boron

µg/l

14,400

18,000

Selenium

 

µg/l

325

420

Strontium

µg/l

6,400

7,200

Iron

µg/l

110

210

Aluminum

 

µg/l

< 1

< 1

Arsenic

µg/l

 

 

Cadmium

µg/l

< 1

20

Chromium (total)

 

µg/l

19

30

Copper

µg/l

4

5

Lead

µg/l

3

6

Manganese

 

µg/l

25

50

Mercury

µg/l

< 0.1

0.2

Molybdenum

µg/l

88

120

Nickel

 

µg/l

14

26

Silver

µg/l

<1

<1

Zinc

µg/l

33

240

a Temperature varied from 23 to 25ºC (summer) to 12 to 1 5ºC (winter).

Source: U.S. Bureau of Reclamation, 1985, after Lee, 1990.

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

TABLE 2.21 Highest Mean Concentration Calculated for Pesticides in Irrigation Runoff Water for Eight Fields in Imperial Valley, California

 

Highest Mean Concentration µg/l

Herbicides

 

Cycloate

2.5

DCPA

153

Dinitramine

15

EPTC

1,250

Prometryn

180

Trifluralin 

10.7

Insecticides

 

Organophosphates

 

Azinphosmethyl

ND

Chloropyrifos

22.4

Diazinon

8.6

Malathion

13.3

Methidathion

64

Mevinphos

1.26

Ethyl parathion

50

Methyl parathion

18

Sulprofos

1.8

Carbamates

 

Methomyl

119

Organochlorines

 

Endosulfan

71

Ethylan

2

Pyrethroids

 

Fenvalerate

5.1

Permethrin

3.4

Note: ND = not detected.

Source: Spencer et al., 1985.

1987; Sonnen, et al., 1987; Mossbarger and Yost, 1989; Ritter et al., 1989, 1991). Nitrate contamination. of ground water is especially prevalent in irrigated areas (Power and Schepers, 1989). Irrigation return flow collected by file drains, therefore, should be expected to contain elevated nitrate concentrations and perhaps some level of pesticide contamination, depending on the type of pesticide applied.

Irrigation return flow can also contain wide ranging concentrations of suspended sediment, as shown by Boucher (1984) for a study of irrigated land in

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

south central Washington. Discharge-weighted mean concentrations of suspended sediment for four drain outflows ranged from 7 (±3) to 1,390 (±100) mg/1 for the 1980 irrigation season and from 9 (±1) to 2,800 (±200) mg/1 for the 1981 irrigation season. Suspended solids concentrations in irrigation return flow for other areas are listed in analyses presented earlier and also show wide variations.

Irrigation return flow constitutes a large supply of water nationally, but especially in the semiarid West. Its physical and chemical quality, however, varies widely. Typically, irrigation return flow contains high concentrations of dissolved solids. It also may contain objectionable concentrations of nitrate, pesticide residues, and trace elements. The suspended solids load of irrigation return flow may also be high. Collectively, these quality characteristics generally indicate the need for extensive treatment of irrigation return flowbefore it might be used for recharge without causing quality problems in the receiving ground water.

Treatment Methods

Water quality parameters of greatest concern in irrigation return flow include suspended solids, TDS, nitrogen and phosphorus compounds, pesticide residues, and various trace metals. Treatment technologies to remove each of these pollutants are available.

Suspended solids can be removed by settling ponds, chemical clarification, filtration, and membrane processes. These methods are sometimes used in series to produce a highly clarified end product free of turbidity and very low in suspended solids concentration. For surface infiltration systems, only settling ponds may be required; however, if the water is to be recharged by wells directly into the aquifer, then chemical clarification, filtration, and membrane processes are needed to produce a suitable effluent. These processes are commonly used to treat surface waters intended for water supply and to treat municipal wastewater prior to disposal or ground water recharge.

Irrigation return flow most often has a TDS concentration that renders it undesirable for recharge without treatment to reduce the mineralization. Reverse osmosis can be used to reduce TDS levels regardless of their initial concentration. For instance, roughly 90 to 95 percent of the inorganics and 95 percent of dissolved organics can be removed by a well-maintained reverse osmosis facility (Treweek, 1985). In addition, particulate matter, bacteria, and viruses also are removed, although the latter are generally not of great concern in irrigation return flow. Although reverse osmosis is widely used in the treatment of brackish and saline waters for water supply, it probably would be economically prohibitive for irrigation return flows.

Nitrate, the dominant nitrogen species in irrigation return flow, can be reduced to acceptable levels by reverse osmosis, if overall water composition re-

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

quires such treatment. If it is unnecessary to treat the water for TDS reduction, then nitrate levels can be reduced instead by biological denitrification, a process that has been applied in the past to treat municipal wastewater effluent. Phosphorus compounds are readily removed by chemical clarification, as are most of the trace metals commonly found in irrigation return flow.

Granular activated carbon (GAC) adsorption can be used if needed to remove soluble pesticide residues remaining after other treatment processes are completed. GAC is a commonly used treatment method for removing dissolved organics in drinking water supplies and has also been used to treat municipal wastewater (Treweek, 1985).

SUMMARY

Wastewaters considered suitable source waters for ground water recharge include municipal wastewater effluent, stormwater runoff, and irrigation return flow. Table 2.22 is a summary of advantages and disadvantages pertaining to the use of wastewaters for recharge, Table 2.23 is a summary of the qualities of the three primary types of wastewater considered in this report. Of the three, treated municipal wastewater effluent is by far the most consistent, spatially and temporally, in both quantity and quality. An exception to this generalization is where raw municipal wastewater and stormwater are commingled in a combined sewerage system.

When compared to other potential impaired water sources, the quality of treated municipal wastewater has been characterized extensively for various levels of treatment because of regulations pertaining to the disposal of municipal wastewater effluent and because municipal wastewater has a history of use as a recharge water source. The body of information on quality of stormwater runoff and irrigation return flow is far less developed, especially when their greater variability is taken into account. Therefore, characterization of stormwater runoff and irrigation return flow quality must be drawn from a much less systematic and comprehensive database than is available for municipal wastewater.

Constituents of concern in municipal wastewater include organic compounds, nitrogen species, pathogenic organisms, and suspended solids. Treatment processes are readily available and have been used successfully to treat municipal wastewater effluent to levels acceptable for various recharge applications. However, even when treated to a very high degree, disinfection of the effluent with chlorine results in the formation of disinfection by-products with the residual organic compounds. These DBPs are of concern if the recovered ground water is to be used for potable purposes.

Urban stormwater runoff quality is affected by several factors, including rainfall quantity and intensity, the natural and anthropogenic characteristics of the drainage basin, time since the last runoff event, and, in northern areas, the time of year. Constituents of concern in urban stormwater runoff include trace

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

metals, organic compounds, pathogenic organisms, suspended solids, and in northern climates in the winter, dissolved solids and chloride enrichment by road deicing practices. Stormwater runoff typically is not treated. However, field experience suggests that many suitable treatment methods exist that may adequately treat most stormwaters before surface infiltration. Overall, stormwater from residential areas is generally best in quality, but its quantity may be extremely erratic and unpredictable due to natural rainfall variations. Recharge with stormwater often requires surface storage and flow regulation because recharge systems do not have the capacity to allow immediate infiltration all the runoff produced by a given precipitation event. Residential area stormwater runoff is best allowed to infiltrate through source area recharge devices, such as french drains, grass filter strips, and grass drainage swales.

Although not examined in depth in this report, industrial stormwater runoff is very irregular in quality, especially for toxicants. Because of this irregular quality and the great potential for severe contamination, industrial area stormwater runoff is not a good candidate for ground water recharge use. Urban snowmelt may also be a poor choice for recharge because of its high salt content. Dry weather flow in stormwater drainage systems may be associated with highly contaminated inappropriate discharges (such as raw municipal wastewater, industrial process water, and illegal dumping of hazardous materials) and should also be avoided. Therefore, to take advantage of urban stormwater runoff as a source of recharge water, care must be taken to isolate the acceptable residential area runoff from the more contaminated flows or to provide additional source area treatment for runoff from the critical areas.

Irrigation return flow exhibits the widest variation in quality of the three potential source waters. It varies from having basically the same quality as high-quality surface water to having a salinity of as much as 10,000 mg/1. The quality characteristics of irrigation return flow are not well studied, except for salinity and concentrations of nitrate. In humid areas, the salt content of irrigation return flow is not a problem, but in semiarid areas it can be enriched to 8 to 10 times that of the water applied. Nitrate concentrations can be as high as 100 to 200 mg/1. Suspended solids and trace element concentrations including selenium, uranium, boron, and arsenic are also of concern. Pesticide residues may also pose problems in irrigation return flow, but in general most of the pesticide residues are associated with particulates and are readily removed with suspended solids. Treatment of irrigation return flow is not generally done, but treatment processes are available to remove the constituents of concern to acceptable levels. The cost-effectiveness of doing so for saline waters is questionable.

In the past, surface and subsurface return flows from irrigated agricultural areas were simply ''disposed" of in streams, lakes, and the ocean without any environmental concerns. This attitude is changing and there is a trend toward increased management to minimize degradation of the environment, such as storage in evaporation ponds and ultimate disposal of salts as solid waste, treat-

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

TABLE 2.22 Advantages and Disadvantages for Using Various Wastewaters in Ground Water Recharge

 

Advantages

Disadvantages

Municipal Wastewater

Primary-treated municipal wastewater

• High TOC for possible improved denitrification

• Poor water quality; higher toxicants, nutrients, BOD, and suspended solids than other municipal wastewaters

 

• Relatively constant flow

 

• Located near major point of use

• High disinfection by-product formation potential

Secondary-treated municipal wastewater

• Most common

• Moderate to poor water quality

 

Relatively constant flows

 

 

High volume

 

 

Located near major point of use

 

Advanced treated municipal wastewater

• Best quality municipal wastewater

• High cost

 

• Low TOC for reduced disinfection by-product formation potential

 

 

• Relatively constant flows

 

 

• Located near major point of use

 

Agricultural Irrigation Return Flows

Irrigation return flow

 

• High pesticides and herbicides

 

 

• High nutrients and salts

 

 

• Irregular flows

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

 

Advantages

Disadvantages

Urban Stormwater

Residential area stormwater

• Likely best quality wastewater

• Irregular flows (highly intermittent)

 

• Most common stormwater

 

 

• Located near major point of use

 

Industrial area stormwater

 

• Highly irregular toxicant quality (likely, contamination from industrial processes and contact with grossly polluted soils)

Urban snowmelt water

 

• High salt content in areas using common de-icing procedures

Dry weather stormwater sewerage flows

 

• High pesticides and herbicides

 

 

• Likely contamination from inappropriate discharges

Combined Sewage

Combined sewage

 

• contains raw sewage with pathogen contamination

 

 

• higher likelihood of toxicants from stormwater from older industrial and commercial areas

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

TABLE 2.23 Comparison of Quality Parameters for Irrigation Return Flow, Urban Stormwater Runoff, and Treated Municipal Wastewater

 

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

ment, or deep-well injection into closed geologic formations. To minimize the cost of these techniques, irrigation efficiencies must be increased, and return flows must be reused as much as possible for irrigation (e.g., raising salt tolerant crops) to minimize the volume of irrigation return flow ultimately produced.

The availability of wastewater for recharge can vary widely throughout the country. Total urban stormwater runoff may provide about 100 acre-feet per square mile per year in areas having about 25 cm (10 inches) of rain per year, but can increase to about 1,000 acre-feet per square mile per year in more humid areas of the United States (having about 125 cm (50 inches) of rain per year). Obviously, some of this water would be diverted from recharge facilities because of poor quality (such as that from industrial areas), reducing the amount available. Municipal wastewater flow would also vary, depending on the population served in a community. Large cities may provide as much as 1,500 acre-feet per square mile per year of wastewater for recharge, but most small towns will generate only about one-tenth as much because of lower population densities. The amount of irrigation return flow available for recharge would vary greatly, depending on irrigation practice, return flow collection efficiency, crop requirements, and rainfall amounts. The amount of irrigation return flow to surface or ground water supplies in 1985 in the United States was estimated to be about 45 million acre-feet (Solley et al., 1988). This amount is 29 percent of the 154 million acre-feet withdrawn for irrigation in 1985. California and Idaho were by far the largest users of irrigation water, together accounting for 37 percent of the national total.

Finally, location of the wastewater source is also important. While urban stormwater runoff and municipal wastewater are usually located near the area of use, most irrigation return flows would be located further from populated areas.

REFERENCES

American Society of Civil Engineers. 1970. Engineering evaluation of virus hazard in water. J. Sanit. Eng. Div., 96(SA1):111.

Amy, G. L., P. A. Chadki, and P. H. King 1984. Chlorine Utilization during formation of THM in presence of ammonia and bromide. Environ. Sci. Technol. 18:781-786.


Bannerman, R., J. G. Konrad, and D. Becker. 1979. The IJC Menomonee River Watershed Study. EPA-905/4-79-029. U.S. Environmental Protection Agency. Chicago, Ill.

Battigelli, D. A., M. D. Sobsey, and D. C. Lobe. 1993. The inactivation of hepatitis A virus and other model viruses by UV irradiation. Water Sci. Technol. 27(3-4):339-342.

Bookman-Edmonston Engineering, Inc. 1992. Annual Report on Results of Water Quality Monitoring: Water Year 1990-91. Report prepared for the Water Replenishment District of Southern California by Bookman-Edmonston Engineering, Inc., Glendale, California.

Boucher, P. R. 1984. Sediment transport by irrigation return flows in four small drains within the DID-18 drainage of the Sulphur Creek basin, Yakima County, Washington, April 1979 to October 1981: U.S. Geol. Surv. Water Resour. Invest. Rep. 83-4167, 149 pp.

Bouwer, H. 1987. Effect of irrigated agriculture on groundwater. J. Irrig. Drain. Eng. 113(1):4-15.

Bull, R. J., and F. C. Kopfler. 1991. Health Effects of Disinfectants and Disinfection By-products. Denver: Am. Water Works Assoc. Res. Found.

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

Callahan, M. A., M. W. Slimak, N. W. Gabel, I. P. May, C. F. Fowler, J. R. Freed, P. Jennings, R. L. Durfee, F. C. Whitmore, B. Maestri, W. R. Mabey, B. R. Holt, and C. Gould. 1979. Water Related Environmental Fates of 129 Priority Pollutants, EPA-4-79-029a and b. Monitoring and Dam Support Division, U.S. Environmental Protection Agency, Washington, D.C.

Carlson, R. R., K. D. Lindstedt, E. R. Bennett, and R. B. Hartman. 1982. Rapid Infiltration Treatment of Primary and Secondary Effluent. J. Water Pollut. Contr. Fed. 54(3):270-280.

Casson, L. W., C. A. Sorber, J. L. Sykora, P. D. Gavaghan, M. A. Shapiro, and W. Jakubowski. 1990. Giardia in Wastewater: Effect of Treatment. Res. J. Water Pollut Contr. Fed. 62(5):670-675.

Cavanaugh, J., H. Weinberg, A. Gold, R. Sangaiah, D. Marburg, W. H. Glaze, T. Collete, S. Richardson, and A. Thruston. 1992. Ozone by-products: Identification of bromohydrins from the ozonation of natural waters with enhanced bromide levels. Environ. Sci. Technol. 26(8): 151-157.

CH2M Hill. 1992. UV Disinfection Pilot Study: Rapid Infiltration/Extraction (RIX) Demonstration Project Prepared for the Santa Aria Watershed Project Authority, City of San Bernardino, and City of Colton. Santa Ana, Calif.

CH2M Hill. 1993. Advanced Wastewater Reclamation Program Final Report. Prepared for the Reedy Creek Improvement District, Lake Buena Vista, Florida, by CH2M Hill, Gainesville, Florida.

Chen, C., R. Nut, J. F. Stahl, R. W. Horvath, and J. F. Kuo. 1993. UV inactivation of bacteria and viruses in tertiary effluents. In Proceeding Specialty Series: Training design and operation of effluent disinfection systems, May 23-25, Hanover, NJ. County Sanitation Districts of Los Angeles County, Whittier, Calif.

Colclough, C. A, J. D. Johnson, R. F. Christman, and D. S. Millington. 1983. Environmental Impact and Health Effects. Pp. 219-229 in Water Chlorination: Chemistry, Environmental Impact and Health Effects, Vol. 4, R. L. Jolly et el., eds. Ann Arbor, MI: Ann Arbor Science.

Crook, J. 1992. Water reclamation. Pp. 559-589 in Encyclopedia Physical Science and Technology, Vol 17. San Diego, Calif.: Academic.


Deason, J. P. 1989. Irrigation-induced contamination: How real a problem. J. Irrig. Drain. Eng. 115(1):9-20.

Dizer, H., W. Bartocha, H. Barrel, K. Seidel, J. M. Lopez-Pila, and A. Grohmann. 1993. Use of ultraviolet radiation for inactivation of bacteria and coliphages in pretreated wastewater. Water Res. 27(3):397-403.


Engberg, R. A., M. A. Sylvester, and H. R. Feltz. 1991. Effects of drainage on water, sediment, and biota. Pp. 801-807 in Proceedings of the 1991 National Conference Sponsored by the Irrigation and Drainage Division of the American Society of Civil Engineers, W. F. Ritter, ed. July 22-26, 1991, Honolulu, Hawaii.

Environment Canada. 1980. Rideau River Water Quality and Stormwater Monitor Study, 1979. MS Rep. No. OR-29 (Feb). Hull, Quebec.


Feachem, R. G., D. J. Bradley, H. Garelick, and D. D. Mara. 1983. Sanitation and Disease—Health Aspects of Excrete and Wastewater Management. Chichester, England: John Wiley, for the World Bank.

Feltz, H. R., R. A. Engberg, and M. A. Sylvester. 1990. Investigations of water quality, bottom sediment, and biota associated with irrigation drainage in the western United States. Pp. 119-130 in The Hydrological Basis for Water Resources Management: Proceedings of the Beijing Symposium, Oct. 1990. IAHS Publ. No. 197.

Field, R., and M. O'Shea. 1993. The detection and disinfection of pathogens in storm-generated flows. Water Sci. Technol. 28(3-5):311-315.

Field, R., V. P. Olivieri, E. M. Davis, J. E. Smith, and E. C. Till, Jr. 1976. Proceedings of Workshop on Microorganisms in Urban Stormwater. EPA-600/2-76-244. Office of Res. and Develop., U.S. Environmental Protection Agency. Cincinnati, Ohio.

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

Galvin, D. V., and R. K. Moore. 1982. Toxicants in Urban Runoff. Toxicant Control Planning Section, Municipality of Metropolitan Seattle. Contract P-16101. U.S. Environmental Protection Agency. Lacy, Wash.

Geldreich, E. E., and B. A. Kenner. 1969. Concepts of fecal streptococci in stream pollution, J. Water Pollut. Contr. Fed. 41(8):R336-R352.

Glaze, W. H. 1986. Reaction products of ozone: a review. Environ. Health Persp. 69:151-157.

Glaze, W. H., J. B. Andelman, R. J. Bull, R. B. Conolly, C. D. Hertz, R. C. Hood, and R. A. Pegram. 1993. Determining health risks associated with disinfectants and disinfection by-products: research needs. J. Am. Water Works Assoc. 83(Feb.):3.

Greene, E. A., C. L. Sewards, and E. W. Hansmann. 1990. Reconnaissance investigation of water quality, bottom sediment, and biota associated with irrigation drainage in the Angostura Reclamation Unit, southwestern South Dakota, 1988-89: U.S. Geol. Surv. Water Resour. Invest. Rep. 90-4152, 75 pp.

Gupta, M. K., R. W. Agnew, D. Gruber, and W. Kreutzberger. 1981. Constituents of Highway Runoff. Vol IV: Characteristics of highway runoff from operating highways. Fed. Highway Admin. Rep. No. FHWA/RD-81/045.


Hoffman, R. J., R. J. Hallock, T. G. Rowe, M. S. Lico, H. L. Burge, and S. P. Thompson. 1990. Reconnaissance investigation of water quality, bottom sediment, and biota associated with irrigation drainage in and near Stillwater Wildlife Management Area, Churchill County, Nevada, 1986-87. U.S. Geol. Surv. Water Resour. Invest. Rep. 89-4105, 150 pp.

Hurst, C. J., W. H. Benton, and R. E. Stetler. 1989. Detecting viruses in water. J. Am. Water Works Assoc. 81(9):71-80.


James M. Montgomery Consulting Engineers, Inc. 1979. Pilot Filtration Studies at the Tapia Water Reclamation Facility. Report prepared for the Las Virgenes Municipal Water District by James M. Montgomery Consulting Engineers, Inc., Pasadena, California.

Jensen, J. N., J. J. St. Aubin, R. F. Christman, and J. D. Johnson. 1985. Characterization of the reaction between monochloramine and isolated aquatic fulvic acid. Vol. 6, Chapter 73, pp. 939-949 in Water Chlorination: Chemistry, Environmental Impact and Health Effects, R. L. Jolly ed. Chelsea, MI: Lewis Publishers.


Keys, J. W. 1981. Grand Valley irrigation return flow case study. J. Irrig. Drain. Eng. 107(IR2):221-232.

Krasner, S. W., M. J. McGuire, J. G. Jacalangelo, N. L. Patania, K. M. Reagen, and E. M. Aieta. 1989. The occurrence of disinfection by-products in U.S. Drinking Water. J. Am. Water Works Assoc. 81:41-53.


Lambing, J. H., W. E. Johnes, and J. W. Sutphin. 1988. Reconnaissance investigations of water quality, bottom sediment, and biota associated with irrigation drainage in Bowdoin National Wildlife Refuge and adjacent areas of the Milk River basin, northeastern Montana, 1986-87. U.S. Geol. Surv. Water Resour. Invest. Rep. 87-4243, 71 pp.

Lance, J. C., R. C. Rice, and R. G. Gilbert. 1980. Renovation of sewage water by soil columns flooded with primary effluent. J. Water Pollut. Contr. Fed. 52(2):381-388.

Lang, R., B. Weinschrott, and G. Tchobanoglous. 1986. Evaluation of the Aqua-Aerobic Automatic Backwash Filter for Wastewater Reclamation in California. University of California at Davis, Department of Civil Engineering. Davis, California.

Law, J. P., Jr. 1987. Irrigation effects in Oklahoma and Texas. J. Irrig. Drain. Eng. 113(1):49-56.

Lee, E. W. 1990. Drainage water treatment and disposal options. Pp. 450-468 in Agricultural Salinity Assessment and Management, K. K. Tanji, ed. New York: American Society of Civil Engineers.


Madison, F., J. Arts, S. Berkowitz, E. Salmon, and B. Hagman. 1979. Washington County Project. EPA 905/9-80-003. U.S. Environmental Protection Agency. Chicago, Ill.

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

Mara, D., and S. Cairncross. 1989. Guidelines for the Safe Use of Wastewater and Excreta in Agriculture and Aquaculture: Measures for Public Health Protection. Geneva, Switzerland: World Health Organization

Matsumoto, M. R., and G. Tchobanoglous. 1981. Evaluation of the Pulsed-Bed Filter for Wastewater Reclamation in California. University of California at Davis, Department of Civil Engineering, Davis, California.

McCarty, P. L., B. E. Rittman, and E. J. Bouwer. 1984. Microbiological Processes Affecting Chemical Transformations in Groundwater. Pp. 89-116 in Groundwater Pollution Microbiology, G. Bitton and C.P. Gerba eds. New York: John Wiley.

McComas, S. R., and D. F. Sefton. 1985. Comparison of stormwater runoff impacts on sedimentation and sediment trace metals for two urban impoundments. In Lake and Reservoir Management: Practical Applications, proceedings of the Fourth Annual Conference and International Symposium. Oct. 1984. NcAfee, N.J.: North American Lake Management Society.

Melnick, J. C., C. P. Gerba, and C. Wallis. 1978. Viruses in water. Bull. WHO 56:499-508.

Metcalf & Eddy, Inc. 1991. Wastewater Engineering: Treatment, Disposal, and Reuse. New York: McGraw-Hill.

Mossbarger, W. A., Jr., and R. W. Yost. 1989. Effects of irrigated agriculture on ground-water quality in corn belt and lake states. J. Irrig. Drain. Eng. 115(5):773-790.

Noss, C. I., R. P. Camolan, and L. Stark. 1989. Virus Removal by Wastewater Effluent Filtration. Prepared for the Florida Department of Environmental Regulation. Tallahassee, Fla.


Olivieri, V. P., C. W. Kruse, K. Kawata, and J. E. Smith. 1977. Microorganisms in Urban Stormwater. EPA-600/2-77-087. U.S. Environmental Protection Agency, Washington, D.C.


Pierce, R. C., and M. P. Wong. 1988. Pesticides in agricultural waters: The role of water quality guidelines. J. Canadian Water Resour. Assoc. (Winnipeg) 13(3):33-49.

Pitt, R. 1979. Demonstration of Nonpoint Pollution Abatement Through Improved Street Cleaning Practices. EPA-600/2-79-161. U.S. Environmental Protection Agency. Cincinnati, Ohio.

Pitt., R. 1983. Urban Bacteria Sources and Control by Street Cleaning in the Lower Rideau River Watershed, Ottawa, Ontario. Rideau River Stormwater Management Study. Ontario Ministry of the Environment. Ottawa, Ont.

Pitt, R. 1984. Characterization, Sources, and Control of Urban Runoff by Street and Sewerage Cleaning. Contract No. R-80597012. Office of Research and Development, U.S. Environmental Protection Agency, Cincinnati, Ohio.

Pitt, R., and J. McLean. 1986. Toronto Area Watershed Management Strategy Study. Humber River Pilot Watershed Project. Ontario Ministry of the Environment. Toronto, Ont.

Pitt, R., and G. Shawley. 1981. A Demonstration of Non-Point Source Pollution Management on Castro Valley Creek. Alameda County Flood Control and Water Conservation District (Hay-ward, CA) for the Nationwide Urban Runoff Program. Water Planning Division, U.S. Environmental Protection Agency, Washington, D.C.

Pitt, R., M. Lalor, R. Field, D. D. Adrian, and D. Barbe. 1993. Investigation of Inappropriate Pollutant Entries into Storm Drainage Systems. A User's Guide. EPA/600/-R-92/238. U.S. Environmental Protection Agency. Cincinnati, Ohio.

Pitt, R, R. Field, M. Lalor, and M. Brown. 1994. Urban stormwater toxic pollutants assessment: Sources and treatability. Accepted for publication in Water Environ. Res. To be published in late 1994.

Power, J. F, and J. S. Schepers. 1989. Nitrate contamination of groundwater in North America. Agr. Ecosyst. Environ. 26:165-187.


Qureshi, A. A, and B. J. Dutka. 1979. Microbiological studies on the quality of urban stormwater runoff in southern Ontario, Canada. Water Res. 13:977-985.


Rice, R.C. and H. Bouwer. 1984. Soil-Aquifer Treatment Using Primary Effluent. J. Water Pollut. Contr. Fed. 56(1):84-88.

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

Ritter, W. F., F. J. Humenik, and R. W. Skaggs. 1989. Irrigated agriculture and water quality in east. J. Irrig. Drain. Eng. 115(5):807-820.

Ritter, W. F., R. W. Scarborough, and A. E. M. Chirnside. 1991. Nitrate leaching under irrigation on Coastal Plain soil. J. Irrig. Drain. Eng. 117(4):490-502.

Rose, J. B., and R. P. Carnahan. 1992. Pathogen Removal by Full-Scale Wastewater Treatment. Prepared for the Florida Department of Environmental Regulation, Tallahassee, Fla.

Sabol, G. W., H. Bouwer, and P. J. Wierenga. 1987. Irrigation effects in Arizona and New Mexico, J. Irrig. Drain. Eng. 113(1):710-748.

Sagik, B. P, B. E. Moore, and C. A. Sorber. 1978. Infectious disease potential of land application of wastewater. Pp. 35-46 in State of Knowledge in Land Treatment of Wastewater, Vol. 1. Proceedings of an International Symposium, U.S. Army Corps of Engineers, Cold Regions Research and Engineering Laboratory, Hanover, New Hampshire.

Sanitation Districts of Los Angeles County. 1977. Pomona Virus Study: Final Report . Prepared for the California State Water Resources Control Board. Sacramento, Calif.

Schillinger, J. E., and D. G. Smart. 1978. Quantification of Non-Point Water Pollutants from Logging, Cattle Grazing, Mining, and Subdivision Activities. NTIS No. PB 80-174063. Springfield, Va.: National Technical Information Service.

Schmidt, K. D., and I. Sherman. 1987. Effect of irrigation of groundwater quality in California. J. Irrig. Drain. Eng. 113(1):16-29.

Sheikh, B., R. P. Cort, W. R. Kirkpatrick, R. S. Jaques, and T. Asano. 1990. Monterey Wastewater Reclamation Study for Agriculture. Res. J. Water Pollut. Cont. Fed. 62(3):216-226.

Shukairy, H. M., R. J. Miller, and R. S. Summers. 1994. Bromide's effect on DBP formation, speciation, and control: part 1, ozonation. J. Am. Water Works Assoc. 86(6):72-87.

Shuval, H. I., A. Adin, B. Fattal, E. Rawitz, and P. Yekutiel. 1986. Wastewater Irrigation in Developing Countries—Health Effects and Technical Solutions. Techn. Paper No. 51. The World Bank. Washington, D.C.

Smith, W. G. 1982. Water quality enhancement through stormwater detention. In Proceedings of the Conference on Stormwater Detention Facilities, Planning, Design, Operation, and Maintenance, Henniker, New Hampshire , W. DeGroot, ed. New York: American Society of Civil Engineers.

Snider, D. E., J. L. Darby, and G. Tchobanoglous. 1991. Evaluation of Ultraviolet Disinfection for Wastewater Reuse Applications in California. Department of Civil Engineering. Davis, Calif.: University of California at Davis.

Solley, W. B., C. F. Merk, and R. R. Pierce. 1988. Estimated use of water in the United States in 1985. U.S. Geol. Surv. Circ. 1004, 82 pp.

Sonnen, M. B., J. L. Thomas, and J. C. Guitjens. 1987. Irrigation effects in six western states. J. Irrig. Drain. Eng. 113(1):57-68.

Spencer, W. F., M. M. Cliath, J. W. Blair, and R. A. LeMert. 1985. Transport of pesticides from irrigated fields in surface runoff and tile drain waters. U.S. Dept. Agr. Conserv,. Res. Rep. 31, 76 pp.


Tanji, K. K. 1981. California irrigation return flow case studies. J. Irrig. Drain. Eng. 107(IR2):209-237.

Treweek, G. P. 1985. Pretreatment processes for groundwater recharge in Artificial Recharge of Groundwater, T. Asano, ed. Boston, Mass.: Butterworth.


U.S. Environmental Protection Agency. 1983. Results of the Nationwide Urban Runoff Program. Water Planning Division, PB 84-185552, U.S. Environmental Protection Agency. Washington, D.C.

U.S. Environmental Protection Agency. 1986a. Design Manual—Municipal Wastewater Disinfection, Ultraviolent radiation, Chapter 7. EPA/625/1-86/021. Water Engineering Research Laboratory, U.S. Environmental Protection Agency. Cincinnati, Ohio.

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×

U.S. Environmental Protection Agency. 1986b. Quality Criteria for Water. EPA 440/5-86-001. U.S. Environmental Protection Agency. Washington, D.C.

U.S. Environmental Protection Agency. 1992. Guidelines for Water Reuse. EPA/625/R-92/DO4, Cemer for Environmental Research Information, U.S. Environmental Protection Agency. Cincinnati, Ohio.

Water Pollution Control Federation. 1989. Water Reuse (second edition)—Manual of Practice. Water Pollution Control Federation, Alexandria, Va.

Weinschrott, B., and G. Tchobanoglous. 1986. Evaluation of the Parkson DynaSand Filter for Wastewater Reclamation in California. Department of Civil Engineering, University of California at Davis, Davis, California.

Wesner, G. M. 1991. Annual Report: Orange County Water District Wastewater Reclamation and Recharge Project, Calendar Year 1990. Prepared for Orange Country Water District, Fountain Valley, California.

Williams, L. H. 1982. Effectiveness of stormwater detention. In proceedings of the Conference on Stormwater Detention Facilities, Planning, Design, Operation, and Maintenance, Henniker, New Hampshire, W. DeGroot, ed. New York: American Society of Civil Engineers.

Wilson, B., P. Roessler, E. Van Dellen, M. Abbaszadegan, and C. P. Gerba. 1992. Coliphage MS-2 as a UV water disinfection efficacy test surrogate for bacterial and viral pathogens. Poster presentation at the American Water Works Association 1992 Water Quality Technology Conference, Nov. 16-17.

Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 35
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 36
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 37
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 38
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 39
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 40
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 41
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 42
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 43
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 44
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 45
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 46
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 47
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 48
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 49
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 50
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 51
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 52
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 53
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 54
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 55
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 56
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 57
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 58
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 59
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 60
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 61
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 62
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 63
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 64
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 65
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 66
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 67
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 68
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 69
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 70
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 71
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 72
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 73
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 74
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 75
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 76
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 77
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 78
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 79
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 80
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 81
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 82
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 83
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 84
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 85
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 86
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 87
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 88
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 89
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 90
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 91
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 92
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 93
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 94
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 95
Suggested Citation:"2 Source Waters and Their Treatment." National Research Council. 1994. Ground Water Recharge Using Waters of Impaired Quality. Washington, DC: The National Academies Press. doi: 10.17226/4780.
×
Page 96
Next: 3 Soil and Aquifer Processes »
Ground Water Recharge Using Waters of Impaired Quality Get This Book
×
Buy Paperback | $80.00
MyNAP members save 10% online.
Login or Register to save!
Download Free PDF

As demand for water increases, water managers and planners will need to look widely for ways to improve water management and augment water supplies. This book concludes that artificial recharge can be one option in an integrated strategy to optimize total water resource management and that in some cases impaired-quality water can be used effectively as a source for artificial recharge of ground water aquifers. Source water quality characteristics, pretreatment and recharge technologies, transformations during transport through the soil and aquifer, public health issues, economic feasibility, and legal and institutional considerations are addressed. The book evaluates three main types of impaired quality water sources—treated municipal wastewater, stormwater runoff, and irrigation return flow—and describes which is the most consistent in terms of quality and quantity. Also included are descriptions of seven recharge projects.

  1. ×

    Welcome to OpenBook!

    You're looking at OpenBook, NAP.edu's online reading room since 1999. Based on feedback from you, our users, we've made some improvements that make it easier than ever to read thousands of publications on our website.

    Do you want to take a quick tour of the OpenBook's features?

    No Thanks Take a Tour »
  2. ×

    Show this book's table of contents, where you can jump to any chapter by name.

    « Back Next »
  3. ×

    ...or use these buttons to go back to the previous chapter or skip to the next one.

    « Back Next »
  4. ×

    Jump up to the previous page or down to the next one. Also, you can type in a page number and press Enter to go directly to that page in the book.

    « Back Next »
  5. ×

    Switch between the Original Pages, where you can read the report as it appeared in print, and Text Pages for the web version, where you can highlight and search the text.

    « Back Next »
  6. ×

    To search the entire text of this book, type in your search term here and press Enter.

    « Back Next »
  7. ×

    Share a link to this book page on your preferred social network or via email.

    « Back Next »
  8. ×

    View our suggested citation for this chapter.

    « Back Next »
  9. ×

    Ready to take your reading offline? Click here to buy this book in print or download it as a free PDF, if available.

    « Back Next »
Stay Connected!