You are on page 1of 127

WATER QUALITY CRITERIA

First National Meeting


on Water Quality Criteria
AMERICAN SOCIETY FOR
TESTING AND MATERIALS
Philadelphia, Pa., Sept. 21-22,1966

ASTM SPECIAL TECHNICAL PUBLICATION NO. 416

List price $5.50; 30 per cent discount to members

published by the
AMERICAN SOCIETY FOR TESTING AND MATERIALS
1916 Race Street, Philadelphia, Pa. 19103
EDgin.Ubrirf
T7)
37$
•M7
I9U

BY AMERICAN SOCIETY FOR TESTING AND MATERIALS 1967


Library of Congress Catalog Card Number: 67-14533

NOTE

The Society is not responsible, as a body,


for the statements and opinions
advanced in this publication.

Printed in Baltimore, Md.


April, 1967
t/J
£/S(S~
GsJG-iKj
A S7~^
7
'- t.lr-b-)
OT&'LO - Z / 0
A
°t> VOC

Foreword

The papers in this volume were presented at the first ASTM National
Meeting on Water Quality Criteria, held in Philadelphia, Pa., 21-22
Sept., 1966. The symposium was sponsored by ASTM Committee D-19
on Industrial Water, in cooperation with The Academy of Natural Sci-
ences of Philadelphia. Chairman of the Program Executive Committee
was Henry C. Bramer, Mellon Institute for Industrial Research. Session
chairmen were Dr. Bramer and James K. Rice, Cyrus Wm. Rice & Com-
pany.
The dinner address, "ASTM's Role in the Field of Water Pollution,"
which was presented by ASTM President James B. Rather, Jr., manager,
toxicology and pollution, Mobil Oil Corp., was published in the November
1966 issue of Materials Research & Standards (Vol 6, No. 11, pp. 580-
581).
Related
ASTM Publications

Industrial Water and Industrial Waste Water, STP 337


(1963), $3.00

Manual on Industrial Water, STP 148-H (1966), $16.75


Contents

Introduction 1
Municipal Water Supplies
Municipal Water from Eastern Rivers—M. E. Flentje 5
Municipal Water from Western Rivers—L. K. Cecil 13
Industrial Water Supplies
Water for Industrial Processes—E. P. Partridge 25
Industrial Water for Cooling and General Purposes—A. J. von Frank and
R. L. Fawcett 44
Agricultural Use
Quantitative Assessment of Irrigation Water Quality—Leon Bernstein... 51
Water for Supplemental Irrigation—Jesse Lunin 66
Recreational Use
Water for Aquatic Life—C. B. Wurtz 81
Water Quality Requirements of Estuarine Organisms—Max Katz and
C. E. Woelke 90
What Do We Really Know About Stream Quality Criteria and Standards?
—F. E. Clarke 100
Measurement of Water Quality with the Pacific Oyster Embryo Bioassay
—C. E. Woelke 112
STP416-EB/Apr. 1967

FIRST NATIONAL MEETING ON


WATER QUALITY CRITERIA—1966

/nfroducf/on
Federal Water Pollution Control legislation enacted in 1965 requires
that the States establish water quality criteria and provide for implementa-
tion of quality controls based upon such criteria. These criteria should be
based upon sound scientific fact if the public interest is to be served by
this legislative directive.
ASTM Committee D-19 on Industrial Water believes that this National
Meeting has been a worthwhile contribution of the professional com-
munity to the task faced by the States in complying with the Water Quality
Act of 1965.
It has been the purpose of this Meeting to evaluate the present state of
scientific knowledge of quality requirements for water to be used in agri-
culture, for municipal and industrial supplies, and for recreation.
The Meeting has been successful in its purpose, that is, in pointing out
both the adequacies and inadequacies of knowledge in the field. In many
instances quality criteria can presently be established on firm scientific
bases; in other cases criteria can only be based upon informed judgment.
Needs for sound, scientific data have been pointed out in the papers pre-
sented at the Meeting. Of particular interest is the emphasis on the needs
for adequately accurate, precise, and standardized methods of analysis.
No criterion or water quality standard is meaningful if methods of analysis
are inadequate for measurements at the appropriate levels of concentra-
tion.
This meeting is the first of a series to be devoted to the examination of
water quality criteria and methods of achieving these goals.
Henry C. Bramer
Mellon Institute, Pittsburgh, Pa.

Copyright© 1967 by ASTM International www.astm.org


This page intentionally left blank
Municipal Water Supplies
This page intentionally left blank
STP416-EB/Apr. 1967

Municipal Water from Western Rivers

REFERENCE: L. K. Cecil, "Municipal Water from Western Rivers,"


Water Quality Criteria, ASTM STP 416, Am. Soc. Testing Mats., 1967,
p. 5.
ABSTRACT: Western rivers are modified by construction of reservoirs
to serve irrigation, flood control, and power production. Municipal sup-
plies from these sources must be taken as is and treated to meet the
Varied requirements of the users. Return irrigation water, natural brine
pollution, and contamination from pesticides, herbicides, and fertilizing
nutrients from agricultural land runoff introduce purification problems
that require adoption of new technics such as activated carbon filtration
and ionic irradiation. More sophisticated analytical tools are needed to
help the purification plant operator control treating systems.
KEY WORDS: rivers, water pollution, water quality, municipal water
quality, brine, irrigation runoff, turbidity, nutrients, activated carbon,
algae, evaporation, herbicides, pesticides

Western rivers are considered to include the Mississippi and its western,
but not eastern, tributaries, and all other rivers to the Pacific Coast. About
the only thing these rivers have in common is water. Some of the Pacific
Northwest rivers have so little turbidity that the water needs no clarifica-
tion for municipal use, so even turbidity cannot be taken as a common de-
nominator. It seems fair to suggest that man-made pollution may become
a common denominator, considering such practices as airplane spraying
of remote forest areas with pest control chemicals.
Opening of vast areas of the semi-arid West to agricultural production
required the construction of dams to even out the flow so water for irri-
gation would be assured. These dams were made large enough for flood
control and power production. Thus, many Western rivers must be con-
sidered as rivers interspersed with very large lakes. The influence of these
lakes as affecting suitability of the river water for municipal use is con-
siderable, and fortunately it is usually beneficial.
Quality Requirements
A municipality must take the river water as is and modify it to serve the
requirements of its customers. Clarification to the extent that no visible
1
Consultant, Tucson, Ariz. Personal member ASTM.
5

Copyright© 1967 by ASTM International www.astm.org


6 WATER QUALITY CRITERIA

sediment is found in the drinking glass and bath tub, and sterilization so
that no water-borne organisms cause disease are the basic requirements.
Palatability comes next, and then perhaps hardness. These criteria serve
the ordinary domestic requirements of drinking, cooking, and bathing.
The agricultural requirements of the home are astonishingly complex.
Different kinds of lawn grass and flowers, some not suited to the area, will
be planted and nurtured by the householder, usually the feminine one, in
a competitive battle to have a different and better display than the neigh-
bors. There is not much the municipality can do about the chloride and
sulfate content of the water. If the pressure for soft water results in selec-
tion of zeolite softening, the conversion of calcium bicarbonate to sodium
bicarbonate can have disastrous results on the lawn and flowers in a long
dry spell. This applies also to golf course greens. Engineering considera-
tions may dictate the selection of zeolite softening, but public relations
may force the selection of lime softening instead.
An important but seldom recognized source of trouble in swimming
pool operation is the bicarbonate content of the make-up water. The higher
the carbon dioxide content, the greater will be the growth of algae, other
things being equal. In most of the West, home swimming pools have
ceased to be a status symbol; they are a necessity. With low alkalinities of
30 to 50 ppm, algal growth will be sparse, but if the alkalinity is 150 ppm
or higher algal growth will be lush. Home pools have intermittent use
and intermittent purification. With the low turnover rates in the pool fil-
tration system it is difficult to clean up a pool with a good load of algae
by starting up the recirculation pump Friday evening to get ready for the
Saturday pool party.
Corrosion of ferrous surfaces is both an irritant and expense to the
householder. The usual problem is the hot water system. The reasons are
complex, and sometimes the cure is beyond the capabilities of the munici-
pal treatment plant. The municipality also has corrosion problems in its
extensive ferrous distribution system. Perhaps the trouble comes not from
the quality of water in the river, but the manner in which it is modified in
the treatment plant.

Extent of Municipal Responsibility for Upgrading River Waters


It is debatable how far a municipality should go hi preparing water for
specialized uses. Most industries have specific water quality requirements
that are higher than those for general domestic use, including domestic
agricultural use. Such industries do not object to furnishing their own
treatment plants to upgrade water quality as needed. They do object if
the municipality treats the water so the "industry system will not produce
satisfactory results. A typical case is carbonated beverage plants treating
with lime to reduce the total alkalinity of the bottling water to or below
50 ppm to maintain proper acid taste in the beverage. If the city starts
CECIL ON MUNICIPAL WATER FROM WESTERN RIVERS 7

using polyphosphate to minimize main corrosion, the phosphate inhibits


the lime reaction, resulting in readings of 75 to 100 ppm, to the detriment
of taste in the bottled drink. Another instance would be maintenance of
a high chlorine residual that would damage organic ion exchange systems.
Industries that are local are competing with other local industries with
the same treating costs. Industries that ship their products over a wide
area compete with industries whose water treating costs vary with the
source water. If water treatment costs are an important part of production
costs, an industry located on a river with bad water is at a disadvantage.
The industry management is likely to blame the municipality, unfairly,
for this situation, and threaten to abandon a planned expansion, or even
to move the plant to another location. If the industry is large enough to be
an important factor in the community this pressure can be very strong.
Also, community boosters out looking for new industries may get an ad-
verse answer if their water supply does not measure up to requirements.
Characteristics of Western Rivers
The three most important characteristics of Western rivers are very
high turbidities, plant nutrients, and high mineral salt content.
Some rivers, such as the Rio Grande and the muddy Missouri, carry
unbelievably high turbidities. On my first visit on a bright summer day in
1932 to the modern treatment plant at Laredo, Tex., I was surprised to
see the plant superintendent in shorts and sneakers dumping lime and
alum into various spots in the coagulation basin, with two workmen hus-
tling to keep him supplied. To my comment that the chemical feeders must
be broken down the superintendent replied that they were running wide
open, and if I thought this looked bad I should have been there when the
turbidity was 50,000 ppm instead of the mere 30,000 ppm now. Bad as
these high turbidities are, the design and operation of treatment plants
with presedimentation basins with continuous solids removal, and good
coagulation and settling basins, permits the delivery of well clarified water
to the filters. There are no unsolved problems here to justify detailed con-
sideration in this paper.
Feth's [I]2 review of nitrogen compounds in natural waters shows
clearly that there are many natural sources for introduction of nitrogen
compounds into rivers. Agricultural, domestic, and industrial waste
sources supplementing the natural sources ensure a sufficient supply of
nitrogen to sustain good algal growths when turbidity is low enough to
permit sunlight penetration. In the highly mineralized soils of the West
sufficient phosphate is always present. Weissner [2] points out the need
for inorganic nutrients in algal growth, particularly iron, manganese,
molybdenum, cobalt, and copper, all of which are present in most Western
2
The italic numbers in brackets refer to the list of references appended to this
paper.
8 WATER QUALITY CRITERIA

rivers as decomposition rock products and leaching from mineral de-


posits. The sustained plant and algal growth as the water moves down-
stream ensures renewal of nutrients as plants die and release nutrients for
further growth. Appreciable numbers of algae do interfere with clarifica-
tion and filtration, but routine treating methods are adequate. Odorous
organic compounds released from decomposition of dead algae, particu-
larly those accentuated by partial chlorination, are no longer a problem
since water plant operators have had the courage to use superchlorina-
tion. Gorham [3] reports that toxins released by certain algae are readily
inactivated by adsorption or aluminum sulf ate flocculation, and that there
seems to be little danger to public water supplies by the sudden release of
toxins during the natural or induced decomposition of blooms. Thus, the
plant nutrient composition of rivers may produce a nuisance in algae re-
moval, an increase in chlorine demand, and no doubt an increase in the
residual soluble organic matter of the treated water.

Eliminating Natural Brine Pollution


One of the most serious problems of Western rivers is the mineral salt
content. There are many very thick salt beds in the Great Plains area, and
seepage from these deposits produce the salt springs that so pollute the
rivers they are no longer satisfactory for municipal use. The Arkansas
River, a beautiful, brawling, clear stream in the Colorado mountains picks
up so much salt in Western Kansas it is no longer fit for municipal use by
the time it reaches the municipalities that want to use it. At one time
enormous quantities of oil field brine were dumped into the Arkansas and
its tributaries, but this pollution has been stopped. Hager and De Geer
[4] report on the U.S. Public Health Service investigation of sources of
natural brine pollution of both the Arkansas and Red rivers and estimate
the quality improvement that would result from control. De Geer [5] re-
ports on studies made by the U.S. Army Corps of Engineers to eliminate
or minimize pollution from these brine springs. Two methods are consid-
ered. One consists of building dams to contain the brine streams in
reservoirs large enough to permit evaporation. The other is to apply
sufficient hydraulic head to stop the brine flow. Success of this depends
on local geologic formations. In one case it failed because the brine
broke through in several other areas, but in another case the formations
were tight and the flow was completely controlled. The study is not fin-
ished, but it seems probable the necessary work will be done. This will so
improve the water quality of these streams the water will be available for
irrigation, industrial, and municipal use in these rich valleys. Cities like
Wichita, Tulsa, Muskogee, and Fort Smith on the Arkansas, and Wichita
Falls, Fort Worth, Dallas, and Shreveport on the Red will have a new,
very large source for their expanding water needs.
CECIL ON MUNICIPAL WATER FROM WESTERN RIVERS 9

Water Quality Degradation by Irrigation Return Flow and Reservoir


Evaporation
The major water use for most Western rivers is irrigation. Much of the
water is consumed by evapotranspiration, but the salts remain behind and
are concentrated in the water moving downward in the soil. Some of this
water percolates back into the river, degrading the quality of water in the
river. In very long rivers this can be a serious pollution problem. Pillsbury
and Blaney [6] consider the problem from the standpoint of keeping the
river water suitable for irrigation. Their conclusions, which are not opti-
mistic, can also be applied to urban use.
Sylvester and Seabloom [7] report on the quality and significance of
irrigation return flow in a study of the Yakima River Valley of Washing-
ton. Their report considers the effect of evaporation and transpiration,
leaching, ion exchange, flocculation and precipitation, erosion, filtration,
heat transfer, and biochemical action. They give detailed analyses of all
the important mineral constituents of the water under various conditions,
such as irrigation season and nonirrigation season, and at various loca-
tions to show progressive degradation as the water moves down the valley.
Hill [8] makes a projection on future quantity and quality of Colorado
River water based on completion of various projects already authorized,
and also on completion of projects not authorized, but in the planning
stage. Particular attention is given to the downstream effect of additional
uses by the Upper Basin states, thus decreasing the volume of water flow-
ing into the reservoirs serving the Lower Basin states. The volume of
water lost by evaporation in the river and several reservoirs will not
change, but the dilution factor will, so the dissolved solids will increase.
He estimates that the concentration of salts in Lake Havasu will increase
from the 1965 level of 655 ppm to 926 ppm by 1990. Lake Havasu is
the source of supply for the Metropolitan Water District serving much of
Los Angeles and other municipalities in Southern California. A 1962
analysis of M.W.D. water as delivered shows a total dissolved solid con-
tent of 719 ppm.
Effects of Pesticides and Herbicides
The law of Malthus (often refuted but never repealed) is again rearing
its ugly head. Now that we are committed to feeding the burgeoning popu-
lation of the world there will be a great increase in farmed acreage and
application of chemicals that help increased crop production. Both are
bad news for the rivers: more salts in the return flow from irrigated fields,
but no greater volume of river flow for dilution. Worst of all may be the
lavish use of nonbiologically degradable organic chemicals for insect and
weed control, and the overuse of fertilizers. Even where the organic com-
pounds are partially degraded, the residual fractions may have objectional
effects on humans.
10 WATER QUALITY CRITERIA

Ettinger [9] in 1960 reviewed the available knowledge on organic com-


pounds in river waters. Some of these are known to be toxic. He points
out that where the carbon filter chloroform extract shows a concentration
greater than 200 ppb, the water nearly always has an objectional taste and
odor. A concentration below 200 ppb is not necessarily safe. In 1962 the
Subcommittee on Organic Chemicals of AWWA Committee 8930P [10]
published a tentative method for Carbon Chloroform Extract (CCE) in
water. Again 200 ppb is referred to, and this figure has become a generally
accepted standard.
The problem of residual pesticides and herbicides in river waters may
become one of the most important problems for the researcher and the
water purification plant operator. There is already an extensive literature
on many phases of the problem, and it is sure to grow. Durham [11] doubts
that the individual intake of pesticides from municipal supplies will be
harmful. Faust and Aly [12] and Warnick et al [13] report on the pollution
of water by organic pesticides. Ryckman et al [14], Faust and Hunter [75],
and Sigworth [16] report on analytical methods for detection of herbicides,
pesticides, and other trace organic matter in water. Booth et al [17] evalu-
ate the effect of sampling conditions on the carbon adsorption method,
and Greenberg et al [18] make a general evaluation of the carbon ad-
sorption method. Ettinger [19] in 1965 updates his earlier discussion on
the detection of trace organic contaminants in water.
The evidence of varying amounts of pesticides and herbicides in river
water is overwhelming. Therefore, the treatment plant must be capable of
reducing these compounds to an unknown safe level. The elegant article
by Robeck et al [20] is must reading for anyone interested. This study
shows that there is a variable, usually small, removal in the ordinary
coagulation, filtration, and chlorination procedures. Heavy oxidation may
even be harmful, as in the case of conversion of parathion to the more
toxic paraoxon. Activated carbon, either powdered or granular, is effec-
tive in proper dosage. The requirements for powdered carbon vary widely
with different pollutants and different concentrations of these pollutants.
Analytical procedures cannot be relied upon to keep up with these varia-
tions, so maximum dosage is required to assure good removals. Granular
carbon beds would seem to be better engineering design, if the beds are
made deep enough. One very important point shown by this study is that
pesticides did not penetrate ahead of other organics in the granular carbon
beds.
Dostal et al [21] report on design studies for carbon beds at Nitro,
W. Va. Hager and Flentje [22] and Cooper and Hager [23] present a
wealth of detail on the mechanism of removal of organic matter by granu-
lar carbon and cost evaluations for both capital investment and operation.
In 1960 the U.S. Public Health Service held a Conference on Physio-
logical Aspects of Water Quality. The Proceedings [24] are well worth
CECIL ON MUNICIPAL WATER FROM WESTERN RIVERS 11

reading six years later. As part of the purpose of the conference is the
statement, "It was recognized that almost no information is presently
available concerning the physiological effects of consuming, over long
periods of time, water that contains either minute amounts of potentially
toxic chemicals or excessive amounts of the common minerals."
The water quality standards that are going to be set by the States,
backed up by the Federal Water Pollution Control Administration, may
have some startlingly strict individual standards because of lack of knowl-
edge of the safe level, so they will be set low.
Before leaving the problem of herbicides, it should be pointed out that
they are very effective in extremely low concentrations. It is entirely
possible the average water purification plant could pass enough of these
compounds to ruin some of the prize plantings of the suburban housewife.
In biochemical and medical circles there is a fast growing interest in
molecular biology studies. We will be learning more about the effect on
humans of many of these toxic organic compounds and their degrada-
tion compounds. Various types of chromatographic procedures have given
much information on the exact composition of these trace compounds in
water, but much more detail is needed. It is a challenge to an organization
dealing with analytical procedures to be ready with the information needed
by the molecular biology researcher.
May I make a prediction that soon ionic irradiation using radioactive
isotopes will be used to fragment compounds like the refractory chlori-
nated organics into harmless units? The introduction of an ionic irradia-
tion unit into the water purification plant is sure to call for new analytical
tools.

References
[1] J. H. Feth, "Nitrogen Compounds in Natural Water. A Review," Water Re-
sources Research, Vol 2, No. 1, 1966, pp. 41-58.
[2] W. Weissner, Physiology and Biochemistry of Algae, Academic Press, New
York, 1962, pp. 267-286.
[3] P. R. Gorham, "Toxic Algae as a Public Health Hazard," Journal, Am. Water
Works Assn., Vol 56, 1964, pp. 1481-1488.
[4] W. R. Hager and M. W. De Geer, "Investigation of Natural Brine and Its
Control in the Arkansas and Red River Basins." Journal, Am. Water Works
Assn., Vol 57, 1965, pp. 383-390.
[5] M. W. De Geer, "Brine Control Projects in the Arkansas and Red River
Basins," Journal, Am. Water Works Assn., Vol 57, 1965, pp. 707-714.
[6] A. F. Pillsbury and H. F. Blaney, "Salinity Problems and Management in
River Systems," Proceedings, Am. Soc. Civil Engrs., Journal, Irrigation
Drainage Div., Vol 92, No. IR1, 1966, pp. 77-90.
[7] R. O. Sylvester and R. W. Seabloom, "Quality and Significance of Irrigation
Return Flow," Proceedings, Am. Soc. Civil Engrs., Journal, Irrigation Drain-
age Div., Vol 89, No. IR3, 1963, pp. 1-27.
[8] R. A. Hill, "Future Quantity and Quality of Colorado River Water," Pro-
ceedings, Am. Soc. Civil Engrs., Journal, Irrigation Drainage Div., Vol 91,
No. IR1, 1965, pp. 17-30; Discussion, Vol 91, No. IR4, pp. 78-86.
[9] M. B. Ettinger, "Proposed Toxicity Screening Procedure for Use in Protect-
12 WATER QUALITY CRITERIA

ing Drinking Water Quality," Journal, Am. Water Works Assn., Vol 52, 1960,
pp. 689-694.
[10] 'Tentative Methods for Carbon Chloroform Extract (CCE) in Water," Sub-
committee on Organic Chemicals of AWWA Committee 893 0-P, Journal,
Am. Water Works Assn., Vol 54, 1962, pp. 223-227.
[11] W. F. Durham, "Physiologic Effects of Pesticide Use," Journal, Am. Water
Works Assn., Vol 57, 1965, pp. 1311-1318.
[12] S. D. Faust and O. M. Aly, "Water Pollution by Organic Pesticides," Journal,
Am. Water Works Assn., Vol 56, 1964, pp. 267-279.
[73] S. L. Warnick, R. F. Gaufin, and A. R. Gaufin, "Concentration and Effects of
Pesticides in Aquatic Environments," Journal, Am. Water Works Assn., Vol
58, 1966, pp. 601-608.
[14] D. W. Ryckman, N. C. Burbank, Jr., and E. Edgerley, Jr., "New Techniques
for the Evaluation of Organic Pollutants," Journal, Am. Water Works Assn.,
Vol 56, 1964, pp. 975-983.
[75] S. D. Faust and N. E. Hunter, "Chemical Methods for the Detection of
Aquatic Herbicides," Journal, Am. Water Works Assn., Vol 57, 1965, pp.
1028-1037.
[16] E. A. Sigworth, "Identification and Removal of Herbicides and Pesticides,"
Journal, Am. Water Works Assn., Vol 57, 1965, pp. 1016-1027.
[17] R. L. Booth, J. N. English, and G. N. McDermott, "Evaluation of Sampling
Conditions in the Carbon Adsorption Method," Journal, Am. Water Works
Assn., Vol 57, 1965, pp. 215-220.
[75] A. E. Greenberg, C. Z. Maehler, and J. Cornelius, "Evaluation of the Carbon
Adsorption Method," Journal, Am. Water Works Assn., Vol 57, 1965, pp.
791-799.
[79] M. B. Ettinger, "Developments in Detection of Trace Organic Contaminants,"
Journal, Am. Water Works Assn., Vol 57, 1965, pp. 453-457.
[20] G. G. Robeck, K. A. Dostal, J. M. Cohen, and J. F. Kreissl, "Effectiveness
of Water Treatment Processes in Pesticide Removal," Journal, Am. Water
Works Assn., Vol 57, 1965, pp. 181-199.
[27] K. A. Dostal, R. C. Pierson, D. G. Hager, and G. G. Robeck, "Carbon Bed
Design Criteria Study at Nitro, W. Va.," Journal, Am. Water Works Assn.,
Vol 57, 1965, pp. 663-674.
[22] D. G. Hager and M. E. Flentje, "Removal of Organic Contaminants by Granu-
lar-Carbon Filtration," Journal, Am. Water Works Assn., Vol 57, 1965, pp.
1440-1450.
[23] J. C. Cooper and D. G. Hager, "Water Reclamation with Granular Acti-
vated Carbon," Chemical Engineering Progress, Vol 62, No. 10, 1966, pp.
85-90.
[24] H. A. Faber and L. J. Bryson, "Conference on Physiological Aspects of Water
Quality," Proceedings, U.S. Public Health Service, Washington, D. C., 1960.
STP416-EB/Apr. 1967

M. E. Flentje1

Municipal Water from Eastern Rivers

REFERENCE: M. E. Flentje, "Municipal Water from Eastern Rivers,"


Water Quality Criteria, ASTM STP 416, Am. Soc. Testing Mats., 1967,
p. 13.
ABSTRACT: In the eastern United States, 47 per cent of the population
derive their water supply from surface sources (reservoirs and rivers)
Generally, the supplies are taken from river tributaries rather than from
the more polluted main stems. The paper suggests, in meeting the grow-
ing need for water it may become necessary to use more polluted sources
through the utilization of the full capabilities of modern water purification
plants. Examples are cited where good operation and relatively simple
plant additions have increased plant capability for handling high coliform
bacteria loadings, algae in high concentrations, tastes and odors, and
organic matter removal. Suggested criteria limits for such plants are
given.
KEY WORDS: water supply, rivers, water treatment, water quality,
pollution, algae, coliform loadings, taste, odors

The area covered in this paper is that generally east of the Mississippi
River and comprises 26 states. The area's population is approximately
120,000,000, and, while the exact data are not readily available regarding
the direct use of rivers for public water supply, an indication of this use
from such sources exists in the fact that 47 per cent of the area population
use water from surface sources [I].2 In only 4 of the 26 states—Con-
necticut, Florida, Indiana, and Mississippi—does a majority of the popu-
lation use ground water for its public supply. When numbers of water
systems or treatment facilities are considered, every state has a greater
number of facilities distributing well water than surface water. The U.S.
ratio of ground water supplies to surface sources is 4.3 to 1; in this Eastern
States area this ratio varies from 2:1 in New England, New York, and
the Ohio-Tennessee Valley, to 4:1 in the Atlantic and Eastern Gulf States
and 5:1 in the area draining to the Western Great Lakes.
Fifty-eight of the Nation's 100 largest cities [2] are located in this
Eastern section. These obtain their public water supply from:
1
Consultant, Cherry Hill, N. J.
2
The italic numbers of brackets refer to the list of references appended to this
paper.
13

Copyright© 1967 by ASTM International www.astm.org


14 WATER QUALITY CRITERIA

Water supply sources—700 largest cities.


1. Impoundments on streams 27
2. Directly from streams 8
3. Directly from streams with short-term (l-20d) raw water storage 8
4. Wells 9
5. Large natural lakes 6
Total 58

Conventional water purification treatment for these largest cities' di-


rect-river-water treatment plants includes prechlorination, coagulation
with alum and lime when needed, sedimentation, rapid sand filtration,
postchlorination, corrosion-pH adjustment when needed, taste and odor
control when necessary and, in approximately 50 per cent of the cases,
ammoniation.
Practically all of the water supply streams in this section of the United
States would be classed as polluted. The degree of pollution varies from
that in sections of Pennsylvania and New England where chlorination
alone is sufficient preparation for public, domestic use to such heavily
polluted streams as the Potomac, Delaware, Ohio, Hudson, and the Kana-
wha in West Virginia—one of the most highly polluted streams in the
United States. Treatment problems vary widely from simple turbidity
removal, iron and manganese removal in Illinois, New England, and
Florida, water softening in Ohio and Florida, high conform concentrations
in numerous locations, and, universally, taste and odor control. The growth
of algae in these streams is a growing problem.
It is of interest that a quick survey of these public river water supplies
indicate that relatively few communities take water out of the main stems
of the principal rivers. Faust [3] has pointed out that only three cities
(Lowell and Lawrence, Mass., and Biddeford-Saco, Me.) in New England
use water from the main rivers. Also only four systems take water out of
the Hudson (Rensselaer, Highland, Port Ewen, and Poughkeepsie) [4].
Instead of the main stream, intakes or impoundments on less polluted
tributaries seems to be the rule.

General Discussion
The term "water pollution" is obviously not a specific one capable of
precise definition. The ability to define the term depends a great deal on
the terminal use of the water in question.
To the average operator of an eastern municipal water supply treat-
ment plant, pollution is a general term applied as a matter of course to his
particular supply source—if it were not polluted why would he and the
plant he is supervising be there? Seldom will this average operator either
make tests for or know values of raw water, dissolved oxygen (DO), bio-
chemical oxygen demand (BOD), or chemical oxygen demand (COD).
FLENTJE ON MUNICIPAL WATER FROM EASTERN RIVERS 15

His knowledge of chlorine demand will often be derived from the chlorine
dosages in use and the necessity to satisfy this demand to obtain chlorine
residual values insuring disinfection rather than from any specific test.
As a practical matter, the particular coliform loading to be met from
day-to-day does not enter into his immediate treatment control considera-
tions—he has come to rely on clarification, filtration, and generally multi-
ple chlorination to insure elimination of these bacteria and their associ-
ated pathogens. Pollution parameters of immediate, day-to-day concern
to him are: turbidity, sometimes color, sometimes microscopic organisms,
pH and alkalinity, tastes, and odors. Others could be added in specific
instances—chlorides, for example. Generally, also, the limiting pollution
criteria values for water uses, in nearly all cases, will be more stringent
than those for water destined for public water supply use.
This situation exists, in this writer's opinion, because conventional
water treatment processes are more dependable and capable of handling
greater pollution loads than they are generally given credit for, and the
average plant operator soon develops a feeling for this capability and
dependability.
It is hardly necessary to repeat that the ability to handle future pollu-
tion loads will require full use of every facility connected with the problem.
In the eastern United States this will mean making greater use of the water
in the main stems of the major rivers as well as taking advantage of the
full capability of treatment facilities. To do this means fuller knowledge
is required of the capability of water treatment processes.
Capability of Municipal Water Treatment and Purification Facilities to
Produce Satisfactory Water
That conventional water treatment plants do deliver bacterially safe
water is self-evident in the vital statistics of the eastern United States as
well as for the rest of the nation. Generally, the water delivered consumers
is palatable, although noticeable chlorinous odors are fairly frequent.
The filtered water is usually extremely low in turbidity and color.
In general, the author's experience has been that no undue problems of
a general nature exist in the eastern United States that have prevented the
production of public, domestic water supplies meeting Public Health
Service Drinking Water Standards requirements even though some rather
highly polluted streams are in use. In some few cases, "auxiliary proc-
essing" has been found necessary to do this. In view of the scarcity of data
concerning the capability of conventional water treatment processes and
the fact that new raw water quality standards can hardly be set without
knowing this capability, and because of the future need to make greater use
of the water in the main stems of our larger rivers, this paper offers some
discussion of the problems of full utilization of public water treatment
facilities.
16 WATER QUALITY CRITERIA

For purposes of discussion, "conventional" public water treatment is


herein considered to include prechlorination, chemical coagulation with
pH correction if required, 3 to 4 hr sedimentation, rapid sand filtration,
postchlorination (with or without ammonia), pH correction for corrosion
prevention, and powdered activated carbon for taste and odor removal.
Thus, presedimentation would be considered a secondary treatment proc-
ess as would secondary granular carbon filters for taste and odor control
or a four- to five-day holding basin for algae control.
Turbidity
Generally, turbidity is not a limiting water treatment parameter in
eastern U.S. supplies. Formerly, many of the water treatment plants along
the Missouri River, for example, considered it necessary to have presedi-
mentation basins in which no chemical coagulant was used. It is my
recollection that such basins were considered necessary when turbidities
were somewhat frequently found above approximately 1000 to 1500
Jackson units (JU). Turbidity values in this range are practically never
encountered in eastern United States.
Microscopic Organisms
The general decrease in the turbidity of river waters due to construction
of dams together with the increase in nutrient matter has created an algae
problem particularly in water treatment plants taking water either directly
out of streams without [5] any short term or with short-term (1 to 3 days)
raw-water storage. Algae are readily killable with copper sulfate in doses
up to approximately 0.5 ppm provided three or more days holding time
is available for kill to be complete.
Possibly, the experience of two Wisconsin plants handling heavily
laden algal waters can be used in this connection. In Menasha and
Oshkosh construction of presedimentation basins of approximately three
days detention with continuous copper sulfate dosing has restored plant
capacity. Menasha plankton concentrates in the range of 6000 units were
reduced 90 per cent by such treatment.
At Celina, Ohio, a 20-plant test with a similar three-day detention
holding and copper sulfate dosing basin in addition to removing algae
gave reductions of 65 to 75 per cent in threshold odor values, 70 per cent
turbidity removal, and reduced the chlorine demand from 2.5 to 2.0
ppm to 0, levels at which a conventional plant could again produce ac-
ceptable water.
Removal of Coliform Bacteria
Most raw-water quality standards limit the coliform bacteria concen-
tration to a monthly average most probable number (MPN) of 5000 per
100 ml or to a value between 5000 and 20,000 per 100 ml if prechlorina-
FLENTJE ON MUNICIPAL WATER FROM EASTERN RIVERS 17

tion or other treatment is provided. That the conventional water treatment


plant can handle considerably higher coliform loadings than this seems
evident [6-9, 18]. The hesitancy to increase this limiting concentration
may be laid to a present lack of knowledge of the importance of water-
borne virus organisms and the fact that human operator judgment still
plays too important a role in the final degree of purification achieved.
Operator training, required operator certification, and computerized
automation are available methods for improving this situation.
Among cities regularly treating highly polluted water, Philadelphia
successfully handles coliform loadings considerably above 5000 per 100
ml, MPN. In 1963 the monthly average MPN values varied between
48,000 and 543,000 at the Queen Lane (Schuylkill River source) and
between 32,000 and 171,000 at Torresdale (Delaware River) [7]. The
annual average MPN for these plants is given as 173,000 and 81,000,
respectively [7]. Each of these plants has an auxiliary process (as defined
earlier) consisting of short-term chlorinated raw water storage. Detention
in this added safety measure is approximately one to two days. Evidence
of the success of treatment has been annual coliform most probable
number values of distribution system samples (10,000+ collected an-
nually) wherein only 2.7 per cent of that allowable by U.S. Public Health
Service Drinking Water Standards (1962).
At Beaver Falls, Pa., it is reported [10] that the Water Authority found
it desirable to install a second coagulation and sedimentation step to in-
sure the production of safe, palatable water when the coliform concentra-
tion, tastes and odors, iron and manganese reached levels of 31,000
coliforms/MPN/100 ml (maximum monthly average 171,000), threshold
odors values of 20 to 200, iron 0.87 to 5 mg/liter and manganese
0.39 to 0.80 mg/liter, phenol concentrate 0 to 35 mg/liter. A treated
water is produced in which no coliform bacteria were detected, free of
turbidity, practically free of color, and with iron below limiting concen-
tration and manganese only at times slightly over 0.05 ppm.
In neither Beaver Falls nor Philadelphia are direct chemical costs
excessively high. At Beaver Falls these annual costs averaged between
$11.14 and $12.27/mg and at the Torresdale (Philadelphia) plant on the
Delaware River [11] average monthly costs were between $6.79 and
$11.06/mg in 1964 and between $7.67 and $12.92/mg in 1963.

Acidity and Alkalinity


Western Pennsylvania and West Virginia streams often contain acid
mine water drainage. This type of pollution may be of help to water
treatment when present in very small amounts, since the total acidity of
these wastes is due mainly to acid salts of iron and aluminum which are
available as coagulants to obtain treatment cost savings. All too often,
however, the acid content is high enough to affect treatment and water
18 WATER QUALITY CRITERIA

quality. As an indication of the size of this problem, in the Appalachia


Area alone 50,000 square miles of land are underlain with coal deposits,
and acid mine wastes result in serious water-quality damage in such
streams as the West Branch Susquehanna River, Kiskiminetas and Cassel-
man River basins in Pennsylvania and West Virginia and Raccoon Creek
in Ohio [19].
The Monongahela River in Pittsburgh has pH values of 5.4 (near 0
alkalinity) and below approximately 25 per cent of the time. Total acidity
(to phenolphthalein) may go 30 to 40 mg/liter as calcium carbonate
(CaCO3). When treated with lime the resulting total hardness added to
that present in the raw water results in total hardness of 200 to 250 ppm.
Several public water supplies have been abandoned in this area because
of excessive pollution with acid mine water wastes. The Vandergrift
Water Co. Apollo station on the Beaver Run tributary of the Kiskiminetas
River was one of these and was abandoned in approximately 1940 when
coping with existing contamination became burdensome from both fin-
ished water quality and treatment expense standpoints. Accurate figures
are difficult to obtain at this late date, but approximate raw water values
were: free water acidity—30 mg/liter and total acidity—100 mg/liter
as CaCO3.
Another public supply completely abandoned because of similar pol-
lution was the treatment plant of the West Perm Water Co., McDonald,
Pa. Free acid concentrations of 100 ppm with total iron values near 40
ppm with a high aluminum content were found entirely impossible to
handle in a conventional type purification plant even if it had been possi-
ble to produce an acceptable water otherwise.

Tastes and Odors


Odor in public water supplies is probably the most easily recognized
parameter of pollution. Slightly over 10 per cent of the Nation's water
treatment plants employ taste and odor control measures, and this treated
water is served to 35 per cent of the population using treated, purified
public water supplies. These general percentages hold in the Eastern
States where slightly less than 30 per cent receive water treated with odor
removal procedures.
Generally, it is possible to control taste and odor levels in finished
water through the use of activated carbon. Powdered activated carbon
dosages ranging from 20 Ib/mg to over 1000 Ib/mg have been reported.
In some instances very troublesome tastes and odors have imposed sec-
ondary treatment on water plants (Hopewell, Va., Beaver Falls, Pa.,
Nitro, W. Va.), and these provide some information at what degree this
type of pollution made necessary departures from conventional treatment.
Characterization of the taste and odor problem is hindered by lack of a
truly quantitative analytical method [73,14].
FLENTJE ON MUNICIPAL WATER FROM EASTERN RIVERS 19

At Hopewell, Va., after years of operation with virtually no taste and


odor problem whatever, a problem of considerable magnitude developed
in 1958 [15]. Thereafter in almost every year, for a three to six months
period, odor difficulties have been encountered that were beyond the
removal capabilities of conventional taste and odor control methods.
Acceptable water quality was not achieved until a supplementary filtra-
tion-adsorption step (granular activated carbon filters) was added to the
domestic supply section of the Hopewell plant. Raw water odor values
ranged from in the troublesome periods below 100 to 400 threshold odor
(TO) which could be reduced by 85 per cent but not sufficiently to
prevent consumer complaints. No difficulty was encountered in meeting
this requirement with the granular activated carbon filters. The adsorp-
tion-facility addition to the Hopewell water treatment plant increased
the filtration equipment by 33 per cent and the capital investment by
an estimated 40 per cent.
At Beaver Falls, as mentioned earlier, objectionable odor concentra-
tion was among the contributing factors making necessary a duplication
of flocculation and sedimentation facilities [10]. Here TO values of 41
to 202 were encountered.
At Nitro, W. Va., a long record exists of taste and odor control
difficulties with water drawn from the Kanawha River [16, 17]. This
river is probably one of the most highly polluted streams in the United
States. In a typical period from January through October, 1963, average
finished water TO values varied from 8 to 25 with maximum daily values
reaching 60 in spite of double aeration (the second stage at about 35
psi) and average powdered activated carbon dosages of 21 ppm and high
chlorine applications. The carbon chloroform extract (CCE) concentra-
tions varied from 960 to 2000 ppb.
A full-scale plant test with granular activated carbon filtration as a
final purification step was carried out in late 1963. The entire plant effluent
was put through two 30-in. deep carbon beds at filtration rates around
3.7 gal/min/ft.2 During the test period raw water TO values varied from
300 to 1300 which were reduced by prior-to-carbon filtration treatment
to 40 to 400 TO. The organic CCE load applied to the carbon beds varied
from 960 to 2000 ppb and was reduced to 130 to 200 ppb [76]. Until ex-
haustion from a taste and odor standpoint, finished water TO values
below 3 to 5 were obtained, and reduction of the CCE extract continued
beyond that.
As the outcome of these tests, 12 filters at Nitro are now in operation
as dual purpose (filtration and adsorption) beds with granular activated
carbon the filter media. This dual use has been practiced elsewhere [17].
The Nitro installation includes a regeneration furnace. It has been esti-
mated that equipment costs for such an installation will be about $4400
per 1 million gallons per day mgd capacity and carbon dosage cost at 3
cents/lb or less [17].
20 WATER QUALITY CRITERIA

In Peoria, 111., disagreeable tastes and odors and high algae counts
required doubling the clarification period in the public water treatment
plant taking water from the Illinois River. In this instance, raw water TO
values were not excessively high but consistently somewhat above TO 25.
In Terre Haute, Ind., extremely high algae concentrations and odor
values in the 50 to 60 TO range could not be reduced to acceptable levels
with conventional treatment and necessitated the development of a sup-
plementary well supply source.
Summary
An attempt has been made to show:
1. That eastern U.S. river waters vary widely in degree of pollution.
It appears that generally, public water supplies using such sources use the
lesser polluted tributaries rather than the main stems of these rivers.
2. Some highly polluted water supply sources are being successfully
treated and demonstrate the ability of conventional water treatment meth-
ods to produce good water from such sources.
3. The capability of such conventional treatment plants can be in-
creased.
4. Such "auxiliary treatment" additions provide an indication of the
limits of certain pollution loadings conventional plants (as defined) can
handle.
In the author s personal opinion these limits may be:

Pollution Parameter Suggested Limiting Value at Which Auxiliary


Treatment Becomes Necessary

1 Turbidity (750 to 1000) JU avg


? Plankton . . 6000 to 7500 units avg
3 Coliform bacteria 100,000 to 150,000 MPN/100 ml
monthly avg
4 Acidity (coal mine wastes) free 35 to 40 mg/liter
total 100 mg/liter
(to pH 8.3)
Odor 50 to 400 TO
6^ Organic matter" 250 ppb
" Plant experience indicates values above this value can be reduced with carbon
filtration.

References
[1] "Statistical Summary of Municipal Water Facilities in the U.S., Jan. 1, 1963,"
PHS Publication 1039, Public Health Service, 1965.
[2] C. N. Durfor and E. Becker, "Public Water Supplies of the 100 Largest Cities
in U.S., 1962," USGS Water Supply Paper, 1812, Reprint 1965.
[3] R. J. Faust, "Northeast Water Crisis and Its Solution," Journal, Am. Water
Works Assn., Vol 58, January, 1966, p. 3.
[4] "Pollution of the Hudson River and Its Tributaries," PHS Publication, Public
Health Service, September, 1965.
[5] J. A. Borchardt and C. R. O'Melia, "Filtration of Algal Suspension," Journal,
Am. Water Works Assn., December, 1961, p. 1493.
FLENTJE ON MUNICIPAL WATER FROM EASTERN RIVERS 21

[6] G. Walton, "Effectiveness of Water Treatment Processes as Measured by


Coliform Reduction, PHS Publication 898, Public Health Service, 1961.
[7] Annual Report, Philadelphia City Water Dept., Laboratory Control and
Water Analyses Section, 1963.
[8] S. J. Baxter, "Water Quality Requirements for Public Water Supplies," Pro-
ceedings, Nat. Symposium on Quality Standards for Natural Water, University
of Michigan, Ann Arbor, Mich., 1966.
[9] C. Biemond, "Relation of River Pollution to Public Water Supply in Europe
and the U.S.," Journal, Am. Water Works Assn., July, 1963, p. 917.
[10] "USPHS Report on Quality of Interstate Waters—Mahoning River Ohio-
Penna.," 1965.
[11] "Preliminary Report and Findings, Delaware Estuary Comprehensive Study,"
PHS Publication, Public Health Service, July, 1966.
[12] "Statistical Summary of Municipal Water Facilities in the U.S., Jan. 1, 1963,"
PHS Publication 1039, Public Health Service, 1965.
[13] R. A. Baker, "Taste and Odor in Water," Final Report to Manufacturing
Chemists Association by the Franklin Institute, Philadelphia, Pa.
[14] Rosen Baker, et al, "Tastes and Odors—Joint Discussion," Journal, Am.
Water Works Assn., Vol 58, June, 1966, p. 695.
[15] E. F. Eld and M. E. Flentje, "Quality Improvements Resulting from Industrial
Needs at Hopewell," Journal, Am. Water Works Assn., Vol 53, March, 1961,
p. 283.
[16] K. A. Dostal, R. C. Pierson, D. G. Hager, and G. G. Robeck, "Carbon Bed
Design Criteria Study," Journal, Am. Water Works Assn., Vol 57, May, 1965,
p. 663.
[17] D. G. Hager and M. E. Flentje, "Removal of Organic Contamination by
Granular-Carbon Filtration," Journal, Am. Water Works Assn., Vol 57, No-
vember, 1965, p. 1440.
[18] R. L. Woodward, "Relation of Raw-Water Quality to Treatment Plant De-
sign," Journal, Am. Water Works Assn., Vol 56, April, 1964, p. 432.
[19] J. E. Biessecker and J. R. George, "Stream Quality in Appalachia as Related
to Coal Mine Drainage," USGS Circular 1526, 1965.
This page intentionally left blank
Industrial Water Supplies
This page intentionally left blank
STP416-EB/Apr. 1967

E. P. Partridge1

Water for Industrial Processes

REFERENCE: E. P. Partridge, "Water for Industrial Processes," Water


Quality Criteria, ASTM STP 416, Am. Soc. Testing Mats., 1967, p. 25.

ABSTRACT: The quality of water suitable for use in different industrial


processes ranges widely. As a result, only broad criteria have been or
should be established by state and interstate agencies for water for in-
dustrial operations. Industry itself has created relatively few criteria for
the appearance properties, taste and odor, electrical conductivity (as a
measure of dissolved salts), dissolved oxygen, and pH required for process
operations where water of a particular quality is needed. The criteria de-
termining the quality of a specific source of water will tend to become
those for municipal water supply and recreation. Possible use of all or
part of a river for transportation of wastes to a treatment plant should
not, however, be arbitrarily eliminated from consideration.

KEY WORDS: water quality, water criteria, appearance, taste, odor, elec-
trical conductivity, dissolved oxygen, pH, water pollution, water renova-
tion, water reuse

This symposium is unique in one respect. Probably never before has


any group in our highly industrial society ever said in a kindly tone to
industry, "what kind of water would you like to have?" instead of shouting
the accusing challenge, "what have you done to our water?"
But asking industry what kind of water it wishes to have supplied to
it for process use is like asking each person in this room what he or she
would prefer for dinner. Moreover, regardless of personal tastes, each of
us will be limited to what is on the menu of the restaurant we choose—
unless we go out into its kitchen and prepare our own individual diet.
Of necessity, this latter independent self-help approach has usually
been employed by industry. In locating any new plant, it has had to select
its "restaurant" in terms of the water available from local sources, but
then has rolled up its sleeves to make some of that water more suitable
for one or for several process steps.
The responsibility of industry for the preparation of the quality of
water required in its operations has been reiterated recently by the Water
1
Vice president, director of Hall Laboratories Div., Calgon Corp., Pittsburgh,
Pa. Personal member ASTM.
25

Copyright© 1967 by ASTM International www.astm.org


TABLE 1—Preferred limits for several criteria of water for use in industrial processes.
The values in this table are derived from summaries in the comprehensive review by McKee and Wolf [5]. They should be used
only after study of these summaries and of the original references cited in them.
p'H Conductivity. Dissolved
max max Odor, max taste and
odor, max
min max
micr
°™hos'
max
Solids
> Oxygen,
mg/litermax ml/llter

Aluminum (hydrate wash) low low


Baking 10 10 low
Boiler feed
0 to 150 psi 80 8.0 3000 <2
150 to 250 psi 40 8.4 2500 <0.2
250 to 400 psi 5 9.0 1500 0
400 to 1000 psi 2 9.6 50 0
>1000 psi <0.5 <0.05
Brewing 10 10 low 6.5 7.0 ... 1500
Carbonated beverages 2 10 low
Carbon black low
Confectionery low 7.0 100
Dairy none none 500
Electroplating and metal finishing, rinse . . . low
Fermentation low low
Food canning and freezing 10 low 7.5 850
Food equipment washing 1 20 none
Food processing, general 10 10 low
Ice manufacturing 170 to 1300 ...
Laundering : 6.6 6.8
Malt preparation . low
Oil well flooding 7.6
Photographic process low low
Pulp and paper
Groundwood paper 50 30 500
Soda and sulfate pulp 25 5 250
Kraft paper
bleached 40 25 300
unbleached 100 100 500
Fine paper 10 5 200
Sugar manufacture low
Tanning operations 20 ioo 6.0 8.0
Television tube manufacture 1
Textile manufacture 0.3 0
25 70
Drinking water 5 15 3 500
Irrigation water
Class I 1000 700
Class II 3000 2100
Class III >3000 >2100
Propagation of fish 6.3 9.0 500 >5.0
2000
28 WATER QUALITY CRITERIA

Improvement Commission of the State of Alabama. On April 6, 1966,


the Commission adopted water quality objectives for five classes of water,
defined according to their best usage. Class D waters are defined in part as
being "usable after special treatment by the user as may be needed under
each particular circumstance for industrial purposes, including cooling
and process water supplies."
The quality of water adequate for use in specific process operations
throughout industry ranges from untreated surface water carrying pollu-
tion from natural, municipal, or industrial sources to completely demin-
eralized water. A similar broad spread is found within many individual
plants with varied process steps. To write general criteria for industrial
process water representing other than the lowest common denominator
accordingly seems useless. The "minimum conditions applicable to all
waters at all places at all times" of the Ohio River Valley Water Sanitation
Commission (ORSANCO) provide an example of such general criteria
seasoned by 17 years of experience [I].2
These minimum conditions state that the water shall be free to the
indicated degree from pollution attributable to municipal, industrial, or
other waste discharges with respect to four categories:
1. Substances that will settle to form putrescent or otherwise objection-
able sludge deposits.
2. Floating debris, oil, scum, and other floating materials in amounts
sufficient to be unsightly or deleterious.
3. Materials producing color, odor, or other conditions in such degree
as to create a nuisance.
4. Substances in concentrations or combinations which are toxic or
harmful to human, animal, or aquatic life.
Beyond such minimum requirements lie the criteria established by
industry for specific operations.

Criteria Developed by Industry


Few industrial associations have paid any attention to the quality of
water to be used in a process operation. The Technical Association of the
Pulp and Paper Industry is the notable exception. A monograph in 1942
summarized a 1939 survey of practice in the industry with respect to
industrial water for the manufacture of pulp, paper, and paperboard [2].
Tentative standards subsequently were written for the chemical composi-
tion of process water to be used in manufacturing certain types of pulp
and paper [3], Those within the scope of this symposium are listed in
Table 1.
How the pulp and paper industry, like all industrial operations, actually
has managed to utilize the water available to it may be glimpsed from a
2
The italic numbers in brackets refer to the list of references appended to this
paper.
PARTRIDGE ON WATER FOR INDUSTRIAL PROCESSES

FIG. 1—Characteristics of water supplies actually used by a number of pulp


and paper mills [4\.

survey in 1951 by the Water Resources Division of the U.S. Geological


Survey [4]. Pertinent though sometimes partial analyses of untreated
water supplies were obtained from the various mills and from Federal and
State agencies. The spread of the available values for color, turbidity, pH,
and dissolved solids is shown in Fig. 1.
Anyone seeking published information on water quality criteria should
start with the monumental report edited by McKee and Wolf [5]. The
second edition, published hi 1963, contains a bibliography of 3827 refer-
ences through the year 1961. These have not only been listed, but where
data were thought to be sufficiently definitive they have been summarized.
In this report, 15 of the 466 pages of text are devoted to industrial
water supply, with 162 references. One of these is the classic, often
TABLE 2—State criteria for pH of water and dissolved oxygen in water for industrial use.
Lowest Class Listed pH Status
ot&tc or Xntcrst&tG Agency lor Industrial .Dissolved Oxygeiij nig/Iitcr nun P—proposed V«ai*
Year
Process Use min max A—adopted

Alabama [8] D (Note 4) 2.0 (Note 10) A 1966


Alaska [23] E 6.0 9.5 3.0 A
Connecticut
(see NEIWPCC)
Illinois [24]
Grand Calumet River 6.5 9.0 3.0 (May-September) P 1966
1.0
Illinois River 6.0 8.5 5.0 day avg
3.0
Ohio River 5.0 9.0 2.0 day avg
1.0
Wabash River 5.0 9.0 2.0 day avg
1.0
Mississippi River
Zone 1 6.5 8.5 5.0
Zone 2 6.5 9.0 4.0
Maryland [25] B 3.8 10.5 4.0 mo. avg A 1949
3.0 day min
Massachusetts
(see NEIWPCC)
Michigan [26] (Note 1) 6.7 8.8 6.0 day avg P 1966
(Note 2) 6.5 8.8 4.0
Minnesota [18] Industrial A 6.5 8.5 P 1965
Industrial B 6.0 9.0
Industrial C 6.0 9.5
Missouri [10] (Note 3) 6.5 8.5 5.0 P 1966
Montana [12] E 6.0 9.5 3.0 A 1958
Nebraska [27] E (Note 5) (Note 5) A 1964
New Hampshire [7] C 5.0 8.5 5.0 A 1963
New Jersey [25] FW-4 6.5 8.5 (Note 11) A 1964
New York [20] D 6.0 9.5 3.0 A 1950
E 5.0 (Note 12)
North Carolina [21] D (Note 6) 3.0 A 1963
E (Note 7) (Note 13)
• . v- ,--~-x
Ohio [29] ... 6.3 9.0 (Note 14) p
p
Tennessee [17] ... (Note 8) (Note 15) A 1952
Utah [9] D 6.5 9.0 (Note 16) A 1965
West Virginia [30] ... (Note 9) 3.0 (Note 17) A 1965
ICPRB [ll]a ... 6.5 8.5 5.0 monthly avg A 1958
4.0
NEIWPCC [31]b C no free acids 5.0 A 1959
or alkalies
ORSANCO [l] . c
... 5.0 9.0 2.0 day avg p 1966
1.0
0
ICPRB—Interstate Commission on the Potomac River Basin.
6
NEIWPCC—New England Interstate Water Pollution Control Commission.
c
ORSANCO—Ohio River Valley Water Sanitation Commission.
NOTES:
1—Recommended for interstate watercourses at the Michigan-Indiana state boundary.
2—Recommended for interstate watercourses at the Michigan-Ohio state boundary.
3—Tentative for interstate waters between Missouri, Oklahoma, and Arkansas.
4—Shall not deviate more than 1.5 units from normal for the waters in the area.
5—No bacteria, organisms, chemicals, or high temperature wastes alone or in combination in sufficient quantity which will render the
receiving waters unsuitable for industrial use.
6—Normal for waters in the area, generally 6.0 to 8.5, except swamp waters may have a low of 4.3.
7—Not lower than 4.3 in waters used for navigation. Otherwise it shall be maintained at such levels as not to adversely affect the
quality of waters in any other class to which these waters may be tributary.
8—No substances added which will make the waters acid or alkaline to such an extent as to interfere with the reasonable and neces-
sary uses of the waters.
9—Acid mine drainage shall be treated so that the receiving stream shall not have a pH of less than 5.5.
10—As measured at five feet in waters ten feet or greater in depth and at mid-depth in waters less than ten feet in depth.
11—Not less than 50 per cent saturation.
12—Sufficient to prevent odor nuisances due to anaerobic decomposition unless other effective means are used to control odors.
13—Sufficient to prevent the development of an offensive condition.
14—Not significant of itself for this use, however, other quality parameters may place limits on this parameter.
15—Not depleted to an extent that will interfere unduly with the reasonable and necessary uses of the waters.
16—Biochemical oxygen demand in mg/liter not to exceed: 25 monthly arithmetical average; 25 in more than 20 per cent of samples
in month; 50 in more than 5 per cent of samples in month.
17—At the point of maximum oxygen depletion.
32 WATER QUALITY CRITERIA

quoted, but now inevitably somewhat dated progress report by the Com-
mittee on Water Quality Tolerances for Industrial Uses of the New
England Water Works Assn., published in 1940 [6].
Table 1 brings together items that the staff supervised by McKee and
Wolf found in the published literature specifying what quality of water
was desired in what process operations, as far as the criteria emphasized
hi this symposium are concerned.
Quoting from the literature has its built-in hazard of perpetuating a
typographical error or an obsolete practice. An example of the latter is
the indication in Table 1 that water containing up to 2 ml/liter (1.4
ppm) of dissolved oxygen may be fed to boilers operating at up to 150
psi. Mechanical deaeration, usually supplemented by the use of an oxy-
gen-scavenging chemical, is standard today even for such low-pressure
industrial boilers.

How Low a Quality for Industrial Process Use?


The water in a lake, a stream, or a particular reach of a stream may be
suitable for drinking without any treatment other than disinfection with a
minimum amount of chlorine. If so, it will undoubtedly earn a rating of
"A" from the state agency responsible for classifying it. Designating its
"best" use as water for drinking does not mean, however, that it should
not be used also in any industrial process for which it is available in ade-
quate supply and is actually the preferable source of supply.
Just because it is pure enough to drink does not, however, make it
universally superior for process use. When the value of the water in the
process is compared with the cost of getting it ready, the sparkling well
water available from the municipal water mains may actually be less
desirable than an apparently lower-quality water from the nearby river.
Industrial processing or cooling normally is considered to be the best
use to which a water classified as "C" or "D" can be applied. The descrip-
tion of such a water differs somewhat from state to state, but the quality
obviously is little, if any, higher than that represented by the previously
quoted criteria of ORSANCO for "all waters at all places at all times."
State agencies thus agree with industry that a plant can condition even
rather raunchy water for most purposes. This attitude may well have de-
veloped at least in part from the expectation, all too realistic in the past,
that the plant would foul up the water, anyway, so why start with too high
a quality?
Inquiry to each state and to several interstate commissions has pro-
vided an interesting composite picture of this quality of water which should
not be used for a more refined purpose than providing industrial process
water. Its characteristics with respect to appearance properties, taste and
odor, dissolved solids as measured by electrical conductivity, dissolved
PARTRIDGE ON WATER FOR INDUSTRIAL PROCESSES 33

oxygen, and pH are included in the following sections and summarized in


Table 2.
Table 2 was prepared from information obtained in the early spring of
1966. It represents a snapshot of a situation changing at an accelerated
rate as state agencies prepared to meet the challenge of the Federal Water
Quality Act of 1965.
Appearance Properties
From the viewpoint of the citizen who wants to go fishing or swimming
or water skiing, or even just desires to sit and look at a stream, the appear-
ance involves a much broader range of observations than just color or
turbidity. Floating solids, an oil slick, islands of foam, or shoals of un-
sightly sludge are even more likely to arouse the viewer's emotions.
Some one of these manifestations may occasionally be just as objection-
able to some operator of some specific industrial process. But how do you
reduce them to criteria expressed in meaningful numbers? The answer to
date is that you do not. Instead you write a general statement. Here are
some examples of recent efforts to grapple with this problem, with key
phrases emphasized:
"Class C waters shall be of the third highest quality and shall be free
from slick, odors and surface-floating solids of unreasonable kind or
quantity . . " [7].
"... Class D Waters.... Substantially free of floating solids, oil, bot-
tom deposits and any other contaminants attributable to sewage, industrial
wastes or other wastes that will render the water unsuitable for agriculture,
industrial cooling and process water supply purposes and fish survival"
[«].
"No person shall discharge into Class D waters any wastes ... which
result in any slicks, floating solids, suspended solids, or sludge deposits in
said waters which are readily visible, or which result in an appreciable
change in color of said waters,. . ." [9].
"There shall be no visible oil in the stream.... There shall be no man
made deposits of solids either organic or inorganic in nature on the stream
bed" [10].
"There shall be no floating solids, oil, settleable solids, or sludge de-
posits attributable to sewage, industrial wastes, or other wastes" [11].
While the proper task group struggles to quantify unreasonable and
substantially and readily and appreciable and even visible and attributable,
many decisions will have to be made on many reaches of many rivers.
Here is an opportunity for the users of water in a basin to adopt construc-
tive head-knocking among themselves to determine what must be done to
keep reusable this vital resource they share in common.
From the viewpoint of the writer of standards, turbidity and color are
relatively easy to reduce to numbers. These numbers are rather unim-
34 WATER QUALITY CRITERIA

portant to many industries, but vital to a few. If one is significant in an


industrial process, usually the other is also watched.
Clear, colorless water is considered essential in photographic processes,
the manufacture of fine paper, and various steps in the production of
synthetic fibers and in the processing of natural as well as synthetic fibers
into textiles [5]. Many of the other process operations in which the use of
clear, colorless water is emphasized actually are related to the preparation
of food and drink. Here there is an interesting situation. The limits of ten
units for color and turbidity of water for baking, brewing, food canning
and freezing, and food processing in general reported in Table 1 from the
summaries for these segments of industry by McKee and Wolf [5] seem
unnecessarily tight. Surely the other components in each of the systems
would be likely to contribute immeasurably more turbidity and color
to the finished product than would the quantity of water used, even if it
were of rather undistinguished appearance. Perhaps we see here a hold-
over from the era in which the quality of water was judged primarily by
how it looked to the eye. And the eye of the visitor to the food plant may
still be a most influential factor in determining the appearance of the
water used in it.
The eye of the customer likewise controls the decision in the carbonated
beverage industry. What enterprising company would enter the market
for club soda with a product that looked a bit muddy?
Agencies concerned with control of pollution not infrequently spell
out the special status of food processing by mentioning it as a "best use"
of the highest class of water together with drinking and culinary use, while
water of two or three classes lower is considered adequate for general
industrial process use.
Taste and Odor
The industries which provide us with food and drink are perhaps even
more concerned with taste and odor of water than they are with color
and turbidity. In plants manufacturing most nonfood products, the flavor
and smell of water is of little concern, except as a signal of possible trouble
from slime accumulation, corrosion of equipment, or dissatisfaction of
employees.
Because we evaluate only the intensity of taste and odor, and do this
with an instrument more sensitive than it is reproducible, the criterion for
water acceptable to the food industries is merely the vague word "low."
This word appears in Table 1 for water used hi baking, brewing, produc-
tion of carbonated beverages, confectionery, fermentation, food canning
and freezing, food processing in general, and malt preparation. A higher
standard of freedom from taste and odor is implied by the word "none"
in the case of the dairy industry and in the washing of food equipment.
Who will volunteer for a new task group to write specifications for the
taste and odor of water to be used in washing taffy-pulling machines?
PARTRIDGE ON WATER FOR INDUSTRIAL PROCESSES 35

FIG. 2—Variation in electrical conductivity of the Ohio River and several


major tributaries at selected monitoring points during 1964 [13].

Obviously, the agencies charged with controlling pollution have felt


that an industrial plant which needed "sweet" water should struggle with
its special problem. The few which specifically mention taste and odor in
relation to water for industrial use present criteria such as:
"There shall be no ... taste or odor producing substances, either alone
36 WATER QUALITY CRITERIA

or in combinations, in sufficient amounts to make the -water unsafe or un-


suitable as a source of industrial water supply, or for boating, shore recrea-
tion, passage of all species of fish or propagation of the hardier species of
fish" [11].
"Class E. Taste or odor producing substances. There shall be none
attributable to sewage, industrial wastes, or other wastes which will ad-
versely affect the marketability of agricultural or industrial products"
[12].
Electrical Conductivity (Dissolved Solids)
State or interstate agencies use the electrical conductivity of water as
a convenient indicator of dissolved solids. Figure 2 shows that conductivity
values varied from point to point on the Ohio River in 1964 and also that
at each sampling point they varied with time over rather wide limits [13].
The concentration of dissolved solids and hence the conductivity at any
point would obviously depend both upon the rate at which wastes were
being dumped into the river system above the point of sampling and the
rate of flow of water in both the main stem and its upstream tributaries.
Industry has relied on this same property to monitor the purity of water
produced by evaporators or demineralizers for make-up to power boilers
and to control the blowdown of dissolved substances from the concen-
trating water in boilers and in recirculating cooling systems.
Water of extremely high purity and therefore extremely low electrical
conductivity is needed in connection with nuclear reactors, as feed for
once-through boilers at high pressures below as well as above the critical
point, and as process water in some special operations, such as the manu-
facture of television tubes and the coating of steel strip with tin. The re-
peated emphasis on these demands for water containing only parts per
billion of dissolved substances has overshadowed the general significance
to industry of how much solute this universal solvent may bring with it
into a plant
Basically, the higher the electrical conductivity of water, the better
medium it is for encouraging corrosion of the metallic materials used to
confine and convey it [14].
If water of high purity is required for some process step in a plant,
variability in the content of dissolved salts of the raw water must be con-
sidered hi engineering the means by which the high-purity water is to be
produced. An example is a large demineralizing plant designed to handle
300 mg/liter of exchangeable cations. During initial operations the actual
concentration of the river water ran as high as 700 mg/liter, with an
average of 400 mg/liter, materially reducing the amount of demineralized
water which could be produced [15].
Dissolved solids in the water used for the final chromic acid rinse in a
phosphatizing system proved to be the critical factor in the blistering of
PARTRIDGE ON WATER FOR INDUSTRIAL PROCESSES 37

the finish on new cars assembled at a plant in California. Serious trouble


began shortly after the water supply to the plant was shifted from wells
providing about 500 ppm of dissolved solids to a new well characterized
by about 800 ppm.
That a much smaller increase of 10 ppm of dissolved solids can cause
a significant increase in the cost of using water for feed to industrial power
plants was estimated in a study made a decade ago by the Steel Industry
Action Committee of ORSANCO [16].
None of the states which provided some numerical criteria for this
paper included any reference specifically to electrical conductivity or to
dissolved solids, for which it is a convenient means of approximate meas-
urement. One [17] mentions:
"Hardness or mineral compounds. There shall be no substances added
to the waters which will increase the hardness or the mineral content of
the waters to such an extent as to appreciably impair the usefulness of the
waters as a source of water supply or interfere with other reasonable or
necessary uses of the waters."
Another [18] states that its Industrial Class B waters should be gen-
erally comparable to its Domestic Class D waters, which, after treatment,
are to meet the 1962 standards for drinking water of the Public Health
Service. This implies a limit of 500 ppm of dissolved solids.
ORSANCO has proposed the following criteria with respect to water
for industrial supply in the Ohio River Basin:
"Dissolved Solids: Not to exceed 750 mg/1 as a monthly average value,
nor exceed 1000 mg/1 at any time. For Ohio River Water, values of spe-
cific conductance of 1200 and 1600 micromhos/cm (at 25 degrees C)
may be considered equivalent to dissolved-solids concentrations of 750
and 1000 mg/1."
Dissolved Oxygen
Dissolved oxygen, as such, is usually not a significant criterion for water
for process use. Substantially zero oxygen in the ground water withdrawn
from a well is actually a boon, as far as reducing the tendency toward
localized corrosion by pitting in steel piping and equipment. If inhabitants
of a stream were not killed by a reduction in oxygen content, causing
troubles to industry ranging from blinded screens at intakes to corrosion
by sulfides to impairment of the quality of products, industry in general
would be pleased to have as little oxygen in surface water supplies as in
most ground water. But when the oxygen content of a surface supply is
used up by chemical or biochemical reactions, the indirect effects may be
as intolerable to the industrial plant as they are to the conservationist or
the housewife.
The criteria for the minimum dissolved oxygen in water for industrial
process use really are those for avoiding anaerobic decomposition of or-
a

ganic material in the water, as indicated in several of die notes in Table


2. But the river which supplies water for industry must also, almost with-
out exception, allow fish at least to survive. The numbers hi Table 2 reflect
concern more frequently for the fish than for industry.
PH
Occasionally, water to be used in an industrial process must be ad-
justed to stay within a narrow range of pH values. In the manufacture
of viscose rayon, for example, it is reported that the pH should be from
7.8 to 8.3. Sometimes there is a minimum value; in the making of hard
candy, the pH should not be lower than 7.0 to avoid inversion of sucrose,
which results in a sticky product. But industry is rarely concerned as long
as the raw water available to it for process use does not contain so much
free acid or alkali that it must be specially treated to bring it back into the
range from 6.0 to 8.5. A few values quoted by McKee and Wolf appear in
Table 1.
Criteria supplied by a number of states for water whose best use is for
industrial processing or cooling are included in Table 2. The narrowest
range is 6.5 to 8.5, the broadest 3.8 to 10.5.
Will Water Qualify Criteria for Industrial Process Use Be Significant?
Actually, criteria for water to be used in process operations are not
likely to be the controlling factor in most real situations. Industrial ac-
tivity will continue to involve people, who will continue to try to live
nearby and will want good water to drink. The criteria for potable water
supply and for recreation presumably will continue to be tighter than
those for industrial water supply with respect to appearance, taste and
odor, dissolved solids, and pH. The growing concern for aquatic life will
establish allowable temperatures lower than and allowable levels of
dissolved oxygen generally higher than would satisfy industry.
The report of water quality criteria for the Calumet Area-Lower Lake
Michigan, issued in January, 1966, provides a recent, possibly prophetic
example of the procedure by which standards may be established for a
body of water [19]:
1. For each of six water areas, some or all of the following seven cate-
gories of water use were chosen for the development of criteria by the
technical committee: (a) municipal, (b) industrial process and cooling,
(c) recreation with whole or limited body contact, (d) fish and wildlife,
(e) commercial shipping, (f) esthetics, and (g) waste water assimilation.
Use for irrigation was so minor it was left out of consideration.
2. Twenty-six constituents were considered for the development of
criteria for each area.
3. Criteria were selected for each constituent for each water use in each
area.
PARTRIDGE ON WATER FOR INDUSTRIAL PROCESSES 39

4. The most stringent criteria for any of the water uses were selected
as the governing values for the area.
The committee felt that it was "establishing a precedent in recommend-
ing criteria which, if attainable, will ensure the highest quality water that
is reasonably feasible."
In the area covered by the report, the open water of Lake Michigan
and the inner harbor basin were considered suitable for municipal supply
as well as for uses satisfied with lower quality. In contrast the Grand
Calumet River was judged to be useful only for industrial processing and

TABLE 3—Comparison of criteria developed for different uses of water in different


parts of the Calumet Area-Lower Lake Michigan Study [19].
Open Water Inner Harbor Grand Calumet
Basins River

Use municipal municipal industrial


and lower and lower and lower
quality quality quality
uses uses uses
Color (true) annual avg 5 5 25
Single daily value or avg 15 15 50
Turbidity (Note 1) (Note 1) (Note 2)
Dissolved solids, mg/liter, annual
avg 165 190
Single daily value or avg 200 230 500
Dissolved O2 , % saturation, an-
nual avg 90 80
Single value 80 65
Mg/liter (tentative) May-Sep-
tember 3.0
Single daily value or avg . . . . 1.0
pH, annual median 8.1 to 8.4 8.0 to 8.5 6.5 to 9.0
Daily median 7.7 to 9.0 7.5 to 9.0
NOTES:
1—No turbidity of other than natural origin that will cause substantial visible
contrast with the natural appearance of the water.
2—No criterion stated.

cooling, esthetics, and waste water assimilation. What are the differences
between the criteria selected for these three waters?
Table 3 includes only the values for the characteristics of water with
which this symposium is concerned, but it illustrates with numbers the
obvious point: criteria for water for industrial use will actually be signifi-
cant only where the water involved is not considered for any use with
more stringent criteria.
But what about the successive reaches of a major river such as the
Ohio, where water is withdrawn and returned by a long sequence of in-
dustrial plants and municipalities? It seems probable that the criteria for
drinking water will ultimately determine the quality to be maintained in
many rivers in each reach above the intake to a municipal water plant.
40 WATER QUALITY CRITERIA

The River as a Sewer


But should every river be cleaned up to the ultimate of 100 per cent
elimination of pollution, as suggested more and more frequently? Con-
sider again the Grand Calumet. This river first picks up industrial dis-
charges, but at the Illinois-Indiana line is essentially effluent from the
sewage treatment plant of Hammond. In any system of classification it
could not be considered today for a "best use" higher than industrial
processing and cooling, and might well be hi the final category of naviga-
tion and the transportation of sewage, industrial wastes, or other wastes.
This lowest class is usually not considered to be a source of water for
industrial processes, or for much of anything else. Here is a composite
picture assembled from several sets of criteria:
"The waters of this classification shall be considered as being devoted
primarily to the transportation of sewage or industrial wastes, or both,
without nuisance" [7].
"The quality of the water shall be such as to be suitable for esthetic
enjoyment of scenery and to avoid any interference with navigation or
damaging effects on property" [18}.
"Dissolved oxygen: sufficient to prevent the development of an offen-
sive condition" [8].
"Hydrogen sulfide: not to exceed a trace" [18].
"Odor producing substances: none in sufficient amounts to cause a
public nuisance as defined by the Penal Law" [20].
pH not lower than 4.3 in waters used for navigation [21]; not lower
than 5.0 [20]; 5.5 to 10.0 [18].
"No contaminant, irregardless of the above specifications, may be dis-
charged in such quantity so as to adversely affect the quality of waters in
any other class to which these waters may be tributary" [8].
The concept of such a class of water dedicated to providing transporta-
tion of wastes is explicitly challenged in the Guidelines for Establishing
Water Quality Standards for Interstate Waters issued by the Federal
Water Pollution Control Administration in May, 1966. After quoting
seven passages from the Water Quality Act of 1965, twelve additional
policy guidelines are stated, the second of which reads:
"No standards of water quality will be approved which provide for the
use of any stream or portion thereof for the sole or principal purpose of
transporting wastes."
There should be continuing debate of this position. In this debate,
German practice in the Ruhr will inevitably be invoked. Five rivers tra-
verse this highly concentrated industrial region on their way to the Rhine.
Of these the Emscher, a poor third in size, is used "for the principal pur-
pose of transporting wastes."
Near the mouth of the Emscher, the entire flow up to about 1000 ft3/sec
is now treated mechanically. Soon this will be supplemented by secondary
PARTRIDGE ON WATER FOR INDUSTRIAL PROCESSES 41

biological treatment to reduce materially the organic load now discharged


into the Rhine. The Ruhr area will then become essentially a closed water
supply and waste water system [22].
Treatment in a single plant of all the wastes from the entire watershed
of the Emscher River, plus some wastes pumped over from the basin of
the Ruhr River, will cost significantly less than separate treatment in
many individual plants.
Extending this approach to our own present situation and trying to
anticipate the exponential increase in manufacturing capability to provide
more and more goods to more and more people, it is not difficult to imagine
some of our own rivers passing through a renovation plant at the down-
stream end of any reach where people and industry have concentrated
to an unusual degree. By more than a loose analogy, these plants would
serve as the kidneys of society, keeping its vital fluid in condition for
continuous reuse. Certainly, this alternative should be one of those evalu-
ated wherever it may seem to have merit, rather than being summarily
excluded from consideration.
To what extent specific rivers or portions of rivers may be allowed to
conform to the only slightly more demanding criteria for industrial use
is a major question. It is a question that may be decided by political re-
sponse at the legislative level to an overenthused but economically un-
informed electorate. Brandishing federal credit cards without realizing
that they must eventually work to pay all the bills, enough good citizens
might decide we should all go for broke on complete elimination of all
pollution even before we have established an astronautical station on the
moon.
We are each moved to some degree by crusades. If we must have a
contemporary crusade, it does seem a bit more practical to start to clean
up the earth than to duty up the moon. But cleaning up the earth requires
more than the flash enthusiasm of a crusade; it is an endless commitment.
For generations our descendants will still be working steadily at improv-
ing control of pollution in each river basin, step by step.
In the continuing effort to meet more and more stringent criteria for
our surface waters, more and more industrial plants will be built or modi-
fied to "bottle up" then- process water, renovating and reusing it repeatedly
in a process cycle or cascading it through successive operations to disposal
other than by return to the supply. Not always will it be practical to go as
far as did the Kaiser Steel Corp. at Fontana, Calif., where the final, well-
worn fluid is evaporated while quenching coke and slag [32]. But, gen-
erally, as quality in the river is upgraded, quality at the end of the multiple
uses in a plant may well drop toward criteria such as the classic "too thick
to pump, too thin to plough."
Somewhere between the limiting concepts of the river serving as a
sewer and the river returned to pristine purity lies most of the reality of
42 WATER QUALITY CRITERIA

the future. While the quality of water required for a process use rarely
will define the criteria to be set for a surface water supply, industry ob-
viously needs to help in creating these criteria for every watershed on
which a manufacturing plant is located.
References
[1] ORSANCO River-Quality Criteria and Minimum Requirements, Ohio River
Valley Water Sanitation Commission, Cincinnati, 1966.
[2] "Industrial Water for Pulp, Paper and Paperboard Manufacture," TAPPl
Monograph Series, Technical Association of the Pulp and Paper Industry, No.
1, 1942.
[3] "Water Technology in the Pulp and Paper Industry," TAPPl Monograph
Series, Technical Association of the Pulp and Paper Industry, No. 18, 1957,
p. 162.
[4] O. D. Mussey, "Water Requirements of the Pulp and Paper Industry," Geo-
logical Survey Water-Supply Paper 1330-A, Government Printing Office, 1961.
[5] J. E. McKee and H. W. Wolf, Water Quality Criteria, State of California
Water Quality Control Board, Publication No. 3-A, 2nd edition, Sacramento,
1963.
[6] E. W. Moore et al, "Progress Report of the Committee on Quality Tolerances
of Water for Industrial Use," Journal, New England Water Works Assn., Vol
54, 1940, pp. 261-272.
[7] Laws Relating to the Water Pollution Commission, State of New Hampshire,
Concord, November, 1965, p. 32.
[8] Water Quality Objectives, State of Alabama, Water Improvement Commis-
sion, Montgomery, April, 1966.
[9] Code of Waste Disposal Regulations, Part II, Standards of Quality for Waters
of the State, Utah State Department of Health, Salt Lake City, May, 1965.
[10] Water Quality Criteria for Interstate Waters between Missouri, Oklahoma,
and Arkansas (Tentative), State of Missouri, Water Pollution Board, Jefferson
City, April, 1966.
[11] Water Quality Criteria for the Potomac River in the Washington Metropolitan
Area, Interstate Commission on the Potomac River Basin, Washington, Janu-
ary, 1958.
[12] Stream Criteria for Waste Discharges, State of Montana, Water Pollution
Council, Helena, reviewed September, 1958.
[13] 17th Annual Report, Ohio River Valley Water Sanitation Commission, Cincin-
nati, 1965, p. 19.
[14] U. R. Evans, The Corrosion and Oxidation of Metals: Scientific Principles
and Practical Applications, Arnold, London, 1960, p. 871.
[15] E. G. Paulson and R. E. Mickel, 'The Study of Operating Performance of
a Large Demineralizer on Ohio River Water," Proceedings, International
Water Conference, Vol 23, 1962, pp. 81-90.
[16] R. D. Hoak, "Industrial Water Quality Requirements," Public Works, Vol 87,
No. 11, 1956, p. 164.
[17] General Policies for the Control of Stream Pollution in Tennessee, Tennessee
Stream Pollution Control Board, Nashville, May, 1952.
[18] Water Pollution Control Commission, Proposed Procedures for Classification
of Waters of the State and Adoption of Standards of Quality and Purity, State
of Minnesota, St. Paul, November, 1965.
[19] Report of Water Quality Criteria, Calumet Area-Lower Lake Michigan,
Illinois-Indiana Interstate Conference, Technical Committee on Water Quality,
4th edition, January, 1966.
[20] Rules and Classifications and Standards of Quality and Purity for Waters of
New York State, State of New York, Water Resources Commission, Albany,
October, 1950, p. 10.
[21] Classifications and Water Quality Standards Applicable to Surface Waters of
PARTRIDGE ON WATER FOR INDUSTRIAL PROCESSES 43
43

North Carolina, State of North Carolina, Stream Sanitation Committee,


Raleigh, November, 1963, p. 12.
[22] A. V. Kneese, The Economics of Regional Water Quality Management,
Johns Hopkins Press, Baltimore, 1964. See Chapter 7, "Water Management
in the Ruhr," pp. 160-187.
[23] Water Quality Objectives and Minimum Treatment Requirements, State of
Alaska, Department of Health and Welfare, Juneau (not dated).
[24] Memorandum to Participants in Public Hearings—Water Quality for Inter-
state Waters or Portions Within the State, State of Illinois, Sanitary Water
Board, Springfield, June, 1966.
[25] Criteria for the Classification of Maryland Streams, State of Maryland, De-
partment of Water Resources, Annapolis, February, 1949.
[26] Recommended Water Quality Criteria for Interstate Watercourses at the
Michigan-Indiana State Boundary, State of Michigan, Water Resources Com-
mission, Lansing, 1966; also Recommended Water Quality Criteria for Inter-
state Watercourses at the Michigan-Ohio State Boundary.
[27] Water Quality Standards Applicable to Nebraska Waters, State of Nebraska,
Water Pollution Control Council, Lincoln, January, 1964.
[25] Regulations Establishing Certain Classifications to Be Assigned to the Waters
of This State and Standards of Quality to Be Maintained in Waters so Classi-
fied, State of New Jersey, Department of Health, Trenton, September, 1964.
[29] Suggested Water Quality Criteria for Various Uses, State of Ohio, Department
of Health, Columbus (not dated).
[30] Requirements Governing the Discharge of Sewage, Industrial Wastes and
Other Wastes into the Waters of the State, State of West Virginia, Water Re-
sources Board, Charleston, April, 1965.
[31] Classification and Standards of Quality for Interstate Waters, New England
Interstate Water Pollution Control Commission, Boston, October, 1959.
[32] H. I. Riegel, "Industry's Water Problems and Their Solution," Proceedings,
Association Iron and Steel Engrs., 1957, pp. 887-892, discussion, pp. 892-893.
STP416-EB/Apr. 1967

A. J. von Frank1 and R. L. Fawcett2

Industrial Water for Cooling and


General Purposes

REFERENCE: A. J. von Frank and R. L. Fawcett, "Industrial Water for


Cooling and General Purposes," Water Quality Criteria, ASTM STP 416,
Am. Soc. Testing Mats., 1967, p. 44.
ABSTRACT: Quality criteria for industrial cooling waters have received
little industrial attention in the context of use for source standards. The
reasons given for this passivity are: (a) a history of industrial adaptation to
a wide variety of water supplies without stress in most cases; (&) an aware-
ness that other beneficial uses, both public and private, involve quality
limits more stringent than those reasonably required for cooling water
supply; and (c) the practical problems in isolating meaningful threshold
concentrations in those multivariant technico-economic systems. The com-
mon parameters of biological growths, carbon dioxide, chlorides, dissolved
oxygen, dissolved solids, iron, manganese, and pH are discussed briefly
against this background.
KEY WORDS: water, industrial water, cooling water, criteria, standards,
water quality

The types of water use, under the title of this paper, include all indirect
cooling through carbon steel and other alloy metals used in construction
as well as recirculation through wood or concrete cooling towers. Apart
from the vital considerations of quantity and temperature, the only ele-
ments that seem significant in this context are those affecting operating
performance and economic life of equipment in such service. One would
not expect to find rigorous criteria for industrial cooling water when one
views its normal sources of supply, that is, sea water, brackish water, in-
land fresh waters with a wide range of natural and unnatural contaminants,
well water, city water, sewage plant effluents, etc.
Water used for cooling accounts for about 90 per cent of the 100 billion
gallons withdrawn and returned daily by manufacturing and thermal-
electric power industries combined. Despite this massive use, quality
1
Director, Air and Water Pollution Control, Allied Chemical Corp., New York,
N. Y.
2
Manager, Industrial Wastes Development Section, Plastics Div., Allied Chemi-
cal Corp., Philadelphia, Pa.
44

Copyright© 1967 by ASTM International www.astm.org


VON FRANK AND FAWCETT ON INDUSTRIAL WATER FOR COOLING45
45

criteria for industrial cooling waters have received relatively little in-
dustrial attention. There is a scarcity of organized information in the
literature on criteria for this major industrial use, particularly in forms
usable for translation into stream standards.
There are only rare and isolated instances where an industrial plant
has considered or proposed any sort of limits to control the general con-
tamination of its water supply [I]3 with silt, corrosive pollutants and slime-
forming organisms being the factors objected to on those infrequent oc-
casions.
The authors are not overlooking the fact that industry had invested
some $400 million [2] through 1960 for water quality treatment facilities
(exclusive of cooling towers). Essentially all, however, is for the 10 per
cent of its water use represented by direct process requirements. Most of
these facilities meet special purpose process needs unattainable with any
natural water quality on the required reliable basis.

Reasons for Passivity on Cooling Water Criteria


It may be useful to examine the ostensible reasons for industrial pas-
sivity on the subject of criteria for its largest water use. There are some
interesting aspects there that do not get beyond the terms of reference of
this symposium.
(a) Within limits, industry tends to be independent of the quality of its
cooling water (again excluding temperature) since it evidently has the
tools and capability to adapt itself to what is available, without stress.
This capability, of course, differs radically from some other beneficial uses
of the public waters such as bathing, fish and shell-fish propagation, etc.
which are sensitive to and completely dependent on the quality of the
water in situ. For industrial cooling, adjustment to the quality of the supply
involves principally the selection of appropriate materials of construction
for heat transfer equipment and its design, or modest and usually inex-
pensive treatments on a continuous or intermittent basis, or recourse to
alternate supplies such as ground water.
(6) Industry's seeming passivity on criteria for its major water use also
reflects its awareness that cooling water quality requirements are fre-
quently inconsistent with objectives for public waters or natural con-
stituents in such waters. Ideal industrial cooling waters would be sterile,
essentially free of dissolved oxygen and hardness salts as the more obvious
examples. When other criteria are involved such as appearance properties,
suspended solids, etc., the requirements of other beneficial uses, public and
private, place more stringent limits on the quality of the supply. Therefore,
when a single composited set of standards is adopted for a reach of a
3
The italic numbers in brackets refer to list of references appended to this paper.
46
46 WATER QUALITY CRITERIA

stream, the more evident criteria for industrial cooling water drop out of
consideration. Most states today are in the process of preparing such lists.
Those applicable to industrial cooling water available to date do not ex-
tend beyond temperature and pH.
(c) The concept of critical threshold levels—which had its origins in
physiological effects, that is, taste and odor, fish toxicity, human health
effects, etc.—has but limited equivalence when one deals with criteria
development for industrial cooling water supplies. For instance, the as-
signment of a threshold level for dissolved oxygen in the corrosion of
carbon steel has no practical meaning without specifying the whole en-
vironment acting or acted upon in the corrosive assault. The rate of at-
tack is accelerated as the carbonic acid, temperature, hydrogen ion con-
centration, and various ionized salts are increased, for instance, and is
retarded by the reverse conditions according to some multivariant rela-
tionship. The rate of attack will be suppressed by hardness scaling and by
chromates and some phosphates. Other variables include the quality and
precise composition of the carbon or stainless steel, or aluminum bronze
as the case may be, the temperature of the cooling water and the material
being cooled, whether the operation is continuous or intermittent, work-
manship in equipment fabrication, electrical grounding, maloperation,
etc.
In the face of these competing effects on the service life of the equip-
ment, the specification of one river water component, in isolation, has not
seemed to be a profitable outlet for research talent or economic investiga-
tion. This does not mean that adverse effects are disregarded. The major
industrial consideration in the selection of such equipment is whether it
will be serviceable over its depreciable life. And the generally successful
decision—which is an economic one and largely empirical—is based on
shared field experience that a given metal will be suitable hi contact with
a given water supply. This type of selective process inherently fails to de-
velop an adequate body of information on water quality criteria of a type
translatable to stream standards.
The authors have indicated three reasons for lack of conforming in-
formation on standards for industrial cooling water: (a) a technical inde-
pendence of quality, (b) the more stringent requirements of other's bene-
ficial uses, and (c) the absence of the go-no go threshold level concept.
The impact of these reasons is not necessarily unconstructive whether
or not one agrees with the accuracy of the assessment. The purpose hi
assembling a catalog of criteria is to provide for others a basis for legally
founded stream standards. For a proper foundation in law, such standards
must be demonstrable economically and socially, and they should be
integrable into the cost benefit equation that tests the validity of a com-
posited stream standards number. In other words, the criterion for criteria
development should be whether they can be usefully developed in that
VON FRANK AND FAWCETT ON INDUSTRIAL WATER FOR COOLING 47

context. Our remarks here today suggest that generalized standards for in-
dustrial cooling water based on supportable criteria are unlikely to emerge
rapidly, if at all.
Effects of Quality on Cooling Water Use
Nonetheless, it may be worthwhile to examine more specifically the
properties of cooling and general service water that affect its use.
Biological Growths—Both algal growths and bacterial slimes hi once-
through and recirculating systems can foul heat exchange systems. Chlori-
nation to a positive residual is an effective control for both and is normally
inexpensive except when the water is so badly polluted the chlorine de-
mand is high.
Color and Turbidity—Neither color nor turbidity due to colloidal or-
ganics is particularly significant.
Carbon Dioxide—This is significant in carbon steel corrosion but is
intimately bound in the complex carbonate equilibria. Even at pH 6 the
ratio of carbonic acid to bicarbonate is 2.86. Staniar [3] reports as "ex-
tremely corrosive" a water of pH 5.9 containing 27 ppm of carbon dioxide
(total solids 48 ppm). The presence of free carbon dioxide in water in-
creases the solvent action on calcium carbonate in cement and concrete
and is significant at 20 ppm [4]. Good industrial practice where water
supplies may be high in carbon dioxide is simple aeration for mechanical
reduction of carbon dioxide (as well as hydrogen sulfide) and the chemical
oxidation of such impurities as ferrous iron or manganous manganese.
Wood delignification is ascribed to sodium carbonate in excess to 200
to 300 ppm. However, its presence in harmful amounts is usually indi-
cated by a high pH reading, that is, 9 to 11, a level not normally encoun-
tered in public waters.
Chlorides—Ordinary carbon steels appear more resistant to corrosion
from chlorides than stainless steels. Common levels of chlorides in inland
"fresh" waters range up to 150 ppm. Field experience at various chloride
levels is not consistent up to several hundred ppm of chloride, possibly
due to lack of complete description of the environment. Larson and
King [5] conclude that the rates of corrosion of iron and steel in water
are a function of the specific mineral quality as well as the alkalinity and
pH values. They report that, with 400 mg/liter of alkalinity (as CaCo3)
and pH 7, the corrosion rate will be zero at 100 mg/liter of salt, but when
the salt concentration is 400 mg/liter, the corrosion rate will be about 100
mg/cm2 per day.
Dissolved Oxygen—For industrial cooling service, dissolved oxygen
may be classed as a pollutant. Zero dissolved oxygen would be desirable
as a means of inhibiting corrosion.
Dissolved Solids and Hardness Salts—Water hardness, caused by cal-
cium and magnesium ions, contribute to scale formation which impedes
4842 WATER QUALITY CRITERIA WATER QUALITY CRITERIA

heat transfer. Deposition is related to both concentration, and tempera-


ture rise across the heat exchanger which is controllable in the design.
Acceptable levels of concentration can be maintained in recirculation
systems by adjusting the amount of make-up and blowdown water.
Iron and Manganese—These are two elements that may cause prob-
lems in water distribution systems because they precipitate from solution.
In fire protection systems, where there is a long residence time, iron and
manganese will precipitate out and may in time interfere with proper
operation of sprinklers, valves, and related equipment. The American
Water Assn. [6] suggests a limit of 0.2 mg/liter iron plus manganese as
the maximum permissible concentration in cooling water.
pH—Attack on concrete is negligible at pH values greater than 7.0
and is critical at pH's less than 6.0 according to many references. Its re-
lationship to carbon dioxide has been mentioned earlier in the context of
steel corrosion.

Conclusion
A review of activity to date under the standards development section
of the Federal Water Pollution Control Act of 1965 would seem to reflect
little industrial interest in industrial cooling water specifications, un-
doubtedly for the reasons given earlier. Aside from temperature, which
is not involved in our discussion here, references to any standards for in-
dustrial cooling water are sparse and in most cases are qualitative not
quantitative.

References
[1] L. C. Burroughs, "Impacts of Water Pollution on Industry," Proceedings, Nat.
Conference on Water Pollution, U.S. Department of Health, Education, and
Welfare, December, 1960.
[2] "Water in Industry," Nat. Association of Manufacturers-U.S. Chamber of
Commerce Report, January, 1965, p. 5.
[3] D. C. Carmichael, "Plant Engineering Handbook," W. Staniar, editor, 2nd edi-
tion, McGraw-Hill Book Co., New York, 1959, pp. 11-30.
[4] J. E. McKee and H. W. Wolf, "Water Quality Criteria," Concrete Deterioration
Due to Carbonic Acid, R. D. Terzaghi, editor, Publication 3-A, California State
Water Quality Control Board, p. 97.
[5] T. E. Larson and R. M. King, "Corrosion By Water At Low Flow Velocity,"
Journal, Am. Water Works Assn., Vol 46, No. 1, 1954.
[6] "Water Quality and Treatment," 2nd edition, Am. Water Works Assn., New
York, 1950.
Agricultural Use
This page intentionally left blank
STP416-EB/Apr. 1967

Leon Bernstein1

Quantitative Assessment of Irrigation


Water Quality*

REFERENCE: Leon Bernstein, "Quantitative Assessment of Irrigation Water


Quality," Water Quality Criteria, ASTM STP 416, Am. Soc. Testing Mats.,
1967, p. 51.
ABSTRACT: New formulas are proposed for quantitatively assessing irriga-
tion water quality. Except for extremely unsuitable waters, irrigation water
quality can be assessed only in the context of the conditions under which the
water is to be used. These conditions include the infiltration rate of the soil
(/), the evapotranspiration rate (E), the irrigation frequency and duration
(tc and ti, respectively) the net downward drainage rate below the root zone
(O) and the maximum permissible salinity and chloride concentration for the
crop to be grown (ECd and Cl*, respectively). When drainage is non-limiting,
the potential leaching fraction (LF) is given by: LF = 1 — (Ete/Iti) and
when drainage is limiting by: LF = O/(E + O). The maximum permissible
EC and Cl concentration of irrigation waters (Ed and C/,-, respectively) are
calculated from: Ed = LF X ECd and C/t = LF X Cld • Potential leaching
fraction calculations will also be useful for estimating CaCO3 precipitation,
and the degree to which specific toxic materials will be concentrated as water
is evapotranspired.
KEY WORDS: water quality, irrigation water, salinity, irrigation, soil water,
evapotranspiration, soil infiltration, drainage, salt water, plants (organisms)

Many variations have been developed for classifying irrigation waters


according to quality. These range from general schemes designed to indi-
cate the suitability of irrigation waters under average use conditions to
specific water-quality ratings for a particular crop in a given area. Al-
though the user of the general scheme [1,2]2 is properly cautioned to take
into account specific conditions in evaluating the suitability of a given
water class, the cautions are often disregarded and are, in any case, diffi-

* Contribution from the U.S. Salinity Laboratory, Soil and Water Conservation
Research Div., Agricultural Research Service, U.S. Department of Agriculture,
Riverside,
1
Calif., in cooperation with the 17 Western States and Hawaii.
Plant physiologist, U.S. Salinity Laboratory, Soil and Water Conservation Re-
search Div., Agricultural Research Service, U.S. Department of Agriculture, River-
side, Calif.
8
The italic numbers in brackets refer to the list of references appended to this
paper.
51

Copyright© 1967 by ASTM International www.astm.org


52 WATER QUALITY4842 WATER QUALITY CRITERIA WATER QUALITY CRITERIA CRITERIA

cult to apply because of their qualitative nature. As a result, workers in


individual states and localities have felt it necessary to propose their own
water class limits for their particular conditions. However, even within
a state or irrigation district, conditions are usually highly variable, and
the modified classifications are also open to question and exception. Ob-
viously, the suitability of an irrigation water needs to be defined quantita-
tively in terms of specific conditions for its use, including soil properties,
irrigation management, climate, and crops.
An example will serve to indicate how water-quality standards depend
on specific use conditions. A water with a total salt content of 3000 mg/
liter would be unsuitable for a sensitive crop but could be used for a
highly tolerant one if other conditions, such as soil permeability and drain-
TABLE 1—Maximum allowable salinities, as electrical conductivity (EC) of the
soil solutions, for crops representing the full range of salt tolerance among crop
plants. 1 mmho/cm EC is equivalent, on the average, to approximately 640 mg of
salt per liter in the soil solution.
Maximum Permissible Maximum Permissible
EC at Indicated0
Crops EC in Soil Solution, Sensitive Stages,
mmho/cm at 25 C mmho/cm at 25 C

Bermudagrass, tall wheatgrass 18


Barley 18 10 (S)
Sugar beets 16 8 (G)
Cotton 16
Tomato, broccoli, spinach, alfalfa 8
Rice 8 4 (S, F)
Potato, corn, flax , 6
Bean clovers (Tri 'folium species) 4
Most pome and stone fruits, citrus . . 4
Strawberries, blackberries, boysenberries . 2 to 4
0
G = germination, S = seedling, F = flowering.

age, permit. Classifying this water as highly saline, or unacceptable for ir-
rigation, may be correct for the average soil, climate, and crop conditions
but is incorrect when exceptionally favorable conditions for the use of
saline waters prevail. In irrigation agriculture, a water has no inherent
quality, except for some extreme cases, independent of the conditions un-
der which it is to be used, and water quality can, therefore, be evaluated
only in the context of a specified set of conditions. This applies particularly
to two water-quality factors, total salt and chloride concentrations. Other
factors such as SAR, bicarbonate, and boron concentrations, although
influenced to some extent by crop tolerances to sodium and boron and by
soil characteristics, can be more readily evaluated in terms of existing
water-classification schemes. Also, adequate data for taking into account
the effect of leaching percentage on precipitation of carbonates are only
now being developed. Major emphasis will, therefore, be placed on the
total salt and chloride concentrations of irrigation waters as quality factors
BERNSTEIN ON QUANTITATIVE ASSESSMENT OF IRRIGATION WATER 53

determining the suitability of irrigation waters under various conditions of


use.
Conditions Affecting Water Quality Determination

Salt Tolerance of Crops


Obviously, the salt tolerance of the crop is a major factor determining
the suitability of an irrigation water. Since there is approximately a ten-
fold range in salt tolerance (Table 1), one might expect, roughly, a com-
parable range in permissible total salt contents of irrigation waters, other
factors being equal. Thus, with a range in tolerance to soil-water salinity of
about 2 to 20 millimho (mmho)/cm, the range in permissible salt content
of the irrigation water would be 0.4 to 4 mmho/cm if one assumes a five-
fold concentration of the irrigation water by evapotranspiration. But this
range would be shifted downward or upward, if the degree of concentra-
tion of the irrigation water by evapotranspiration was greater or less than
fivefold. The feasible degree of concentration of the irrigation water in
passing through the root zone of the crop will be determined by soil,
climate, management, and crop factors as speckled below.

Soil Permeability and Infiltration Rate


Control of soil salinity depends on the ability to introduce into the
root zone an amount of water in excess of evapotranspirational losses.
The excess water will carry salt below the root zone if drainage is adequate.
Therefore, infiltration rates must be adequate to permit the penetration
of a depth of water sufficient to meet evapotranspiration and leaching
requirements during an irrigation period that is not excessive for the crop
in question. Prolonged irrigation can damage crops because of water-
logging and poor aeration. Thus, the second set of factors that must be
taken into account is the infiltration capacity of the soil in relation to the
evaporative and transpirational losses of water. The frequency and dura-
tion of irrigation are variable but subject to such limits as the quantity of
water retained in the root zone, the availability of water from the supply-
ing source, and the tolerance of the crop to prolonged irrigation.

Drainage
If a calculated rate of leaching is to be obtained, natural or artificial
drainage must be adequate to convey the drainage water away from the
root zone, so that rising water tables and salt accumulation are avoided.
Thus, drainage becomes a third condition for assessing permissible salinity
in irrigation waters. Drainage, together with soil properties (infiltration
rate), climate (evapotranspiration rate), irrigation management (irrigation
frequency and duration of irrigation), and salt tolerance of the crops are
54 WATER QUALITY CRITERIA

the minimum conditions that must be specified to determine the suitabil-


ity of an irrigation water.
Leaching Potential in Relation to Soil, Climate, and Management Factors
The leaching fraction 9lf0
(LF} is defined as

LF=% J-'i
(1)
910

or the relative depth of drainage water, Dd, to infiltrated irrigation


water, Dt. (All water into or out of the soil will be considered as the
equivalent depth of water spread uniformly over the soil.) But
Dd = Di- De (2)
where De is the equivalent depth of water evapotranspired. Also,
Di = Iti (3)
and
D. = Ete (4)
where / and E are the average infiltration rates and evapotranspiration
rates in millimeters over the infiltration time, t*, and the irrigation cycle,
te, in days of 24 hr, respectively. Substituting Dd from Eq 2 in Eq 1,
we have

LF -*=.* or IF-l-l or
(5)
Substituting the expressions in Eqs 3 and 4 for D, and De in Eq 5, we
have

LF= 1-f* Hi
(6)
Equation 6 relates LF to evapotranspiration rate, infiltration rate, infil-
tration period, and irrigation cycle. Since E and / are expressed in
millimeters per day, and tc aand ti in days, the term Etc/Iti reduces to a
ratio.
If the drainage rate is limiting, that is, if
Ot < (It)
e 970
where O is the average net drainage rate in millimeters per day past the
lower boundary of the root zone during the irrigation cycle tc (that is,
Dd = Otc}, then the leaching fraction is determined by the drainage
limitation, since the permissible water infiltration, Itt, cannot exceed
Ete, the amount evapotranspired, plus Ote, the amount drained, with-
out a rising water table. Equation 1 for this limiting case becomes

LF =
Et „
+ °*'
Ot or LF=—2—
c cE + O (8)
980
BERNSTEIN ON QUANTITATIVE ASSESSMENT OF IRRIGATIO

FIG. l(a)—Relation of irrigation cycle-infiltration time ratio, t c /ti to infiltration


rate-evapotranspiration rate ratio, I/E for given leaching fractions, LF. Drainage
not limiting, (b)—Relation of evapotranspiration rate, E, to drainage rate, O, for
given leaching fractions, LF. Drainage limiting.
56 WATER QUALITY CRITERIA

Equation 6 defines the attainable leaching fraction when drainage rate is


not limiting, and Eq 8 when drainage rate is limiting.
Figure 1 presents graphic solutions of the two equations. In Eq 6,
four independent variables are involved in the term Ete/Iti. This can
conveniently be broken down into two ratios, E/I, the ratio of evapo-
transpiration rate to infiltration rate, and tc/tt, the ratio of the irrigation
cycle to infiltration time. The quantity I/E, rather than E/I, is plotted,
to avoid fractional values. If all four variables are evaluated, LF is
determined, but if only I/E is established, one can determine the ratio,
te/ti, for a desired LF (see following). Practical values for te and f,- can
then be chosen, to yield the necessary ratio. The primary upper limit for
te is, of course, determined by the quantity of water available in the root
zone.3 If the maximum permissible salinity of the irrigation water is to
be determined, te/ti must be minimized, that is, the minimum practical
tc and the maximum practical ti must be used. If, however, the suit-
TABLE 2—Maximum permissible EC in irrigation water for given LF's and crop
tolerances to EC of soil water, mmho/cm at 25 C.
Maximum Permissible 0.10 LF 0. 20 LF 0.30 LF 0.40 LF
EC in Soil Water

2 0.2 0.4 0.6 0.8


4 0.4 0.8 1.2 1.6
8 0.8 1.6 2.4 3.2
16 1.6 3.2 4.8 6.4

ability of the irrigation water for any larger value of te/ti is of interest,
the larger /c/ft- value is used.
Since only E and 0, the evapotranspiration rate and drainage rate,
determine LF when drainage rate is limiting, the representation of the
relationship in Fig. 16 for the limited drainage case is straightforward.
It is, of course, essential, in the first place, to determine whether drain-
age is or is not limiting according to Eq 7.
Once the LF is calculated by Eq 6 or 8, the suitability of an irrigation
water for the particular set of conditions is assessed by use of Eq 9.
FC
LF = ±£i or Ed = LF X ECd (9)
.cCd
[Equation 9 is derived from Eq 1 with the assumption of conservation
of salt in the soil-water system, that is, no net change in salt burden by
precipitation of salts, solution of soil minerals, and salt uptake by crops.
8
The upper limit of tc is determined by the size of the soil-water reservoir, that
is, irrigation is required when the reservoir is depleted to the point at which growth
will begin to be adversely affected to an unacceptable degree. One may irrigate before
this limit is reached to achieve necessary leaching, especially if it is not feasible to
achieve the leaching by extending the infiltration period.
BERNSTEIN ON QUANTITATIVE ASSESSMENT OF IRRIGATION WATER 57

With this assumption, Dd/Di = Ed/ECd . Quantities Ed and ECd are


electrical conductivities of irrigation water and drainage water, respec-
tively.]
Maximum tolerable ECds are given in Table 1 for a number of crops,
and limits for other crops are also available [J]. Leaching requirement
studies currently in progress may result in some change in permissible
soil-water salinities, and the ECd values may also be modified, depending
on the levels of Ed. A more saline water will probably require a lower
maximum ECd to compensate for the higher Ed . However, these refine-
ments must be deferred until adequate data are available.
In Table 2, maximum permissible EC values for irrigation waters
calculated by Eq 9 are tabulated for various LF's and for maximum
permissible EC of the soil water or drainage water below the root zone.
This table is intended to be illustrative and does not include the full
range of LF's. Leaching fractions greater than 0.30 or 0.40 are rarely
practical but may be achieved on some well-drained sandy soils; leaching
fractions of less than 0.10 are feasible with waters of exceptionally low
salt content or with highly salt-tolerant crops. Entering the table with the
calculated LF [Eq 6 or 8], one selects the line representing the maximum
permissible salinity for the crop in question, to obtain the maximum EC
of the irrigation water for the specified conditions. If the available water
has a lower EC than that found in the table, it is satisfactory for the use
specified; if it has a higher EC, it is unsatisfactory. In the latter case, the
use conditions, such as tc/ti, or crop may be altered to determine the
conditions under which the given water can be satisfactorily used. The
ratio I/E may also be increased by shifting to a season when the E
value is lower.
Even for a specified situation, precise values for all parameters may
not be available, or may be variable. If the values used are conservative
(minimum value for / or maximum value for E, and so forth), the final
judgment will be safe. If average values for these parameters are used,
then the judgment of the suitability of the water for average conditions
will be correct, but it may not be for less favorable situations (that is,
parts of a field having a lower /, or a smaller ti or O, etc.).
The preceding argument refers solely to total salinity of the water
expressed as EC. This is generally recognized as one of the predominant
quality-factors for irrigation water because of the relationship of EC to
osmotic pressure and the major effect of osmotic pressure of the root
medium in determining crop yields under saline conditions [3]. However,
some plants are specifically sensitive to certain ions, and this sensitivity
in particular cases may even overshadow osmotic pressure in determin-
ing plant growth and even survival. Woody plants, including most fruit
trees, berry and vine crops, and ornamental shrubs are specifically sensi-
tive to sodium and chloride ions and develop characteristic leaf-burn
58 WATER QUALITY CRITERIA

symptoms when leaves accumulate about 0.25 and 0.50 per cent on a
dry-weight basis of these elements, respectively. Maximum permissible
chloride concentrations in the soil solution are given for fruit crops in
Table 3. Because of large variations in the rate of chloride accumulation
by different rootstocks and varieties, tolerance for a given crop varies.

TABLE 3—Maximum permissible chloride contents in soil solution for various


fruit-crop varieties and rootstocks.
ijimit of Tolerance
Crop Rootstock or Variety to Chloride in Soil
cSolution, meq/liter

Rootstocks
Citrus Rangpur lime, Cleopatra mandarin 50
Rough lemon, tangelo, sour orange 30
Sweet orange, citrange 20
Stone fruit Marianna 50
Lovell, Shalil 20
Yunnan 14
Avocado .. West Indian 16
Mexican 10
Varieties
Grape Thompson seedless, Perlette 50
Cardinal, black rose 20
Berries Boysenberry 20
Olallie blackberry 20
Indian summer raspberry 10
Strawberry Lassen 16
Shasta 10

TABLE 4—Maximum permissible chloride concentrations in irrigation waters for


given LF's and crop tolerances to chloride in the soil water, meq/liter.
Maximum Permissible 0.10 LF 0.20 LF 0.30 LF 0.40 LF
Cl in Soil Water
10 .... 1 2 3 4
20 2 4 6 g
30 3 6 9 12
40 4 8 12 16
50 5 10 15 20

In Table 4, permissible chloride concentrations of irrigation waters


are tabulated according to crop tolerance to chloride and calculated
LF's. In dealing with a chloride-sensitive crop, one should check both
total salinity, EC, and chloride concentration of the irrigation water. A
given water may have a low enough total salt content but may develop a
higher chloride concentration at a given LF than a particular fruit crop
can tolerate. Remarks regarding the leaching fraction range in Table 2
also apply to Table 4 which is parallel in all respects except that chloride
instead of EC is the evaluated water property.
BERNSTEIN ON QUANTITATIVE ASSESSMENT OF IRRIGATION WATER 59

Leaching Fraction, Leaching Requirement, and Calculation of Total Irri-


gation Water Application
The leaching fraction as defined in this treatment is the predicted
fraction of the infiltrated water that will pass below the root zone calcu-
lated according to Eq 6 or 8. Leaching requirement, or "LR," on the other
hand, is defined as

TR-ECi
LR
~ECa
or the fraction of the applied water that must pass through the root zone
in order not to exceed a specified salinity in the soil water, ECd, deter-
mined by the salt tolerance of the crop. The two concepts are matched
when we compare the leaching fraction determined by specific use condi-
tions with leaching requirement determined by irrigation water salinity
and maximum permissible soil salinity. When the leaching requirement
is less than the leaching fraction, salinity control within the prescribed
limits is feasible. If the leaching requirement exceeds the leaching fraction,
then the necessary salinity control cannot be maintained, and conditions
must be modified to increase the leaching fraction by changing irrigation
practices or soil properties or drainage, or tc decrease leaching require-
ment by substituting crops that are more salt-tolerant.
Irrigators must generally apply more water than is necessary to meet
actual water deficits because of nonuniformity of application and infiltra-
tion. Such excess should be computed on the basis of total requirements,
including leaching, LR, as well as consumptive use, CU. The depth of ir-
rigation water, D*, required to meet the leaching requirement is

D - CU
Di-T=Th

If an additional 10 or 20 per cent of water must be applied to assure that


the desired depth of water penetrates all parts of the field, then the total
water application should be 1.1 A or 1.2 Dt. Applying 1.1 CU or 1.2
CU insures only that the CU is met but does not provide for LR. The drain-
age rate must be adequate for the extra drainage required.
Other Irrigation Water Properties Involved in Assessing Irrigation Water
Quality
Sodium-Adsorption-Ratio (SAR)
Major attention in this paper is devoted to total salt, EC, and chloride
concentration because new procedures for assessing these qualities with
regard to conditions of use are being proposed. However, other water
properties are fully as significant in determining the suitability of an ir-
60 WATER QUALITY CRITERIA

rigation water, even though the conditions of use do not generally need
to be specified as rigorously in assessing the suitability of waters. One of
these properties is sodium-adsorption-ratio, defined as
Na+
SAR =
/Ca++ + Mg-H-V/2
\ 2 J
where ionic symbols denote concentrations in milliequivalents per liter.
The exchangeable-sodium percentage, ESP, of soils is related to SAR by

esp
100 (- 0.0126 + 0.01475 SAR) ,}
1 + (- 0.0126 + 0.01475 SAR) L J
The ESP of soil is the fraction of negatively charged adsorption sites in
soils occupied by sodium ions. If more than 10 to 20 per cent of the
cation-exchange-capacity of soils is taken up by sodium, the physical con-
dition of the soil deteriorates. The effect usually becomes serious at lower
ESP's, the finer the soil texture, hence the range of 10 to 20 per cent ESP
associated with soil structure deterioration. Soils high in exchangeable
sodium are dispersed and have low permeability to water and air and are
unsuitable for crop production. Waters with high SAR's cause high
ESP's to develop as sodium in the irrigation water displaces calcium and
magnesium until the equilibrium condition indicated by the ESP-SAR re-
lation is achieved.
Sodium-sensitive crops (woody plants) may accumulate harmful levels
of sodium in the leaves when the ESP of the soil is as low as 5 per cent.
For fruit crops, the tolerable limit for SAR of the water is about 4, instead
of 8 to 18, as it is for nonsensitive crops when the soil effects predominate.
A few crops, such as beans, suffer nutritional effects because of increased
exchangeable sodium and associated decreases in exchangeable calcium
and magnesium. Such effects are noted on nonsaline, sodic soils (high in
exchangeable sodium but low in salinity). Exchangeable sodium of about
10 per cent can cause such nutritional disturbance in these crops [6].
SAR varies with the square root of the total salt content of the water
when cationic proportions are constant. A fourfold increase in total salt
content causes a doubling in SAR. The lower the LF, the greater will be
the SAR of the soil water at the bottom of the root zone. The effect of
leaching fraction on ESP is not of great concern because the empirical re-
lation of ESP to SAR takes this concentration factor partly into account
[4]. In summary, waters of high SAR (8 to 18 and above) are hazardous
because the high exchangeable sodium percentages they induce will criti-
cally impair soil structure. Waters with SAR's of about eight or above
may also impair the nutrition of bean crops, and waters having SAR's of
four to eight and above will cause damaging sodium accumulations in
some fruit crops and other woody plants.
BERNSTEIN ON QUANTITATIVE ASSESSMENT OF IRRIGATION WATER 61

Bicarbonate
The bicarbonate ion in concentrations of 10 to 20 milliequivalents
(meq)/liter or more causes iron chlorosis in plants, but this is rarely of
concern in the field where such high bicarbonate concentrations do not
develop because of reaction of bicarbonate with calcium and precipitation
as calcium carbonate. However, this precipitation reaction reduces the
concentration of divalent cations and increases the SAR or sodium hazard
of the water. Available data [7]4 indicate that the degree of precipitation
of calcium as CaCOs depends on the leaching fraction as well as on the
Ca+ + and the HCO3~ concentrations of the irrigation water. As much as
2/3 of the bicarbonate in the water may be involved in the precipitation of
calcium when the leaching fraction is low or about 10 per cent [7].4 With
increase in leaching fraction, the per cent of bicarbonate that precipitates
decreases. Eventually, it will be possible to relate degree of bicarbonate
precipitation quantitatively to leaching fraction and to the pertinent water
components. Total salt content in the water will then be corrected for the
precipitable calcium carbonate, and this adjusted salinity figure, instead
of the total salinity, will be used in judging suitability of a water according
to Table 2. The calculated leaching fraction will be useful in estimating
the amount of precipitable calcium carbonate, since leaching fraction is
one of the major factors in determining calcium carbonate precipitation.
SAR will also be adjusted in accordance with the quantity of the divalent
cations involved in precipitation of lime.
In addition to precipitation of lime, calcium sulf ate may also precipitate
from waters containing high concentrations of calcium and sulfate ions.
Leaching fraction is an important factor hi this reaction, also, since pre-
cipitation of calcium sulfate depends on reducing the volume of the soil
water by evapotranspiration. Detailed evaluation of salt precipitation
phenomena is beyond the scope of this paper.
Because salt conservation was assumed in calculating the data in
Tables 2 and 4, other deviations from this basic assumption must also be
examined. Not only may some salts precipitate in the soil as a result of
concentration of the soil solution, but soil minerals may dissolve to in-
crease the salt burden of the soil water. This effect may, in some instances,
largely offset the effect of salt precipitation [5].5 Also, plants absorb some
salt, although by far the major part of the salt contained in irrigation water
appears in the drainage water. The salt uptake by plants may be an ap-
preciable factor in waters of very low salinity. For example, with only 1
meq/liter of chloride in the water, and a water requirement of 500, fruit-
crop leaves that accumulate 10 meq of chloride per 100 g of dry leaf
absorb l/5 of the chloride originally present hi the water that they take up.
4
5
G.
Ogata and C. A. Bower, unpublished data, U.S. Salinity Laboratory.
Unpublished data, U.S. Salinity Laboratory.
62 WATER QUALITY CRITERIA

However, since the leaves return to the soil, this absorbed chloride is re-
leased and contributes to the chloride absorbed by the subsequent flush
of leaf growth. There is thus no net removal of chloride by the crop. (The
fruit itself accumulates very little chloride, compared to the leaves and
other plant parts.) Salt uptake by the crop can generally be neglected in
calculating the permissible salt contents in irrigation waters.

Boron and other Trace Elements


Boron, although essential for plant growth at a concentration of about
0.5 mg/liter, becomes toxic when concentrations exceed 1 to 4 mg/liter,
depending on boron tolerance of individual crops and and varieties [9].
Some varieties or rootstocks of fruit crops are more boron-resistant than
others because of lower boron uptake [10]. Because boron is weakly ad-
sorbed in soil, marginal levels of boron in the irrigation water may not be
immediately toxic, but prolonged use of waters exceeding the specified
levels cannot be tolerated.
Other elements such as lithium and selenium may be present in local
water supplies in sufficient concentrations to be damaging to crops. Irriga-
tion water quality standards should take into account such toxic trace ele-
ments. Leaching fraction should theoretically affect the tolerable level of
potentially toxic trace elements, since the lower the leaching fraction, the
greater the concentration of the elements in the reduced volume of soil
water. The tolerance limits proposed for boron and other trace elements
do take into account the leaching fraction to some degree because they are
based, in part, on field observations of plant response to waters of known
toxic element concentration. A more precise statement of the relation of
permissible toxic element content to leaching fraction will require more
precise observations than are currently available.

Other Management Practices That Affect Suitability of Irrigation Waters

Planting Techniques and Sensitive Growth Stages


If all use factors permit, waters with EC's of 2 to 6 mmho/cm and
higher may be used for irrigation (Table 2). However, this will leave a
residue of salinity in the plow layer that may affect the establishment of
subsequent crops. Soils with an electrical conductivity of the saturation
extract, ECe, of 1 to 2 mmho/cm in the plow layer may not permit germi-
nation of furrow-irrigated row crops because salinity is concentrated five-
to ten-fold in the seed row of single-row crops as a result of salt transport
into the center of the bed [11]. With double-row plantings on wider beds,
the concentration factor is reduced to about one. (Salinity about 5 cm in
from the bed shoulders after irrigation approximates average salinity in the
plow layer before irrigation.) With sloping seed beds, salt is transported
beyond the seed row, and salinity may be effectively reduced by furrow
BERNSTEIN ON QUANTITATIVE ASSESSMENT OF IRRIGATION WATER 63

irrigation [11]. Since the use of more saline waters will generally leave a
greater residual soil salinity in the plow layer, the successful establishment
of furrow-irrigated row crops will benefit from the use of planting beds that
minimize salt concentration in the seed row.
Some crops are particularly sensitive during early stages of growth
(Table 1). Care must be taken to avoid excessive salinity during these criti-
cal stages. Maximum EC of the soil water will usually not develop at early
growth stages and can usually be controlled by heavier irrigation and in-
creased leaching. For rice, draining the field just prior to flowering and
reflooding with less saline water accomplishes the necessary reduction in
salinity during flowering and seed-setting.

Irrigation Methods
With varieties or rootstocks that restrict chloride uptake (Table 3) and
suitable use conditions, waters with 5 meq/liter or more of chloride can
be used to irrigate fruit crops (Table 4). However, if sprinklers that wet
the foliage are used, foliar absorption of chloride and leaf injury can result.
As little as 3 meq/liter of chloride or sodium in the irrigation water may
result in harmful accumulations of the respective ions in the leaves of
citrus, stone-fruit trees, and almonds [12]. Some fruit crops such as avo-
cado and strawberry absorb salt very slowly through the leaves, so that
foliar absorption of chloride or sodium is not a hazard for all fruit crops.
Foliar absorption during the evening is only half as rapid as absorption
during the day. Since chloride or sodium in water wetting the foliage of
some crops becomes critical at about % 0 the critical concentration of
these ions in the soil water, foliar absorption of salt can severely restrict the
use of even good quality waters for sprinkler irrigation of some fruit crops.
Low-head sprinklers that do not wet the foliage can provide the benefits
of sprinkler irrigation while avoiding the hazards of foliar absorption of
salt.
64 WATER QUALITY CRITERIA

APPENDIX
Specimen LF Calculations

Case 1A, All Parameters Specified. Adequate Drainage


An irrigator is able to specify values for all parameters as follows: E, evapo-
transpiration rate, = 8 mm/day; /, infiltration rate, =120 mm/day; ?» = one
day (24-hr irrigation period); tc = 12 days (irrigation every 12 days); and O,
drainage rate, = 5 mm/day.
c
Ot= 60 mm, and //, - Etc = 120 - 96 = 24 mm, so that Otc > (Iti -
Etc), and drainage is not limiting.
From Eq 6, LF = 1 - [(8 X 12)/(120 X 1)] = 0.2.
The available irrigation water has an EC of 0.5 mmho/cm, and citrus is to
be grown. From Eq 9, Ed = 0.2 x 4 = 0.8. Since the EC of the water (0.5)
is less than the permissible value of 0.8, the water can be safely used under the
conditions specified. Actually, since the leaching requirement is 0.5/4, or 0.125,
the LF can be reduced by increasing tc or decreasing ti, that is, by irrigating
every 13 days instead of every 12 days for 24 hr, or irrigating for 22 hr instead
of 24 hr every 12 days. (Exact calculations, of course, are not intended to ex-
clude a need for additional application to compensate for nonuniformity.)
Since citrus is being grown, the irrigator should also consider the chloride
concentration. The chloride concentration of the irrigation water is 3 meq/liter.
The rootstock is sour orange, with a maximum permissible chloride concentra-
tion in the soil-water of 30 meq/liter. Since Ch = LF X Cld, the maximum
permissible chloride in the irrigation water for a LF of 0.20 is 6 meq/liter, and
even for the LF of 0.125, it is 3.75 meq/liter, or more than the actual chloride
concentration in the irrigation water of 3 meq/liter. The chloride concentration
of the water, therefore, is not excessive for the crop under the specified condi-
tions.

Case IB, All Parameters Specified. Inadequate Drainage


All conditions as in Case 1A, but O is 0.5 mm/day. Otc = 6 mm and
< (Iti — Etc) which is 24 mm. Even if drainage is minimized to give the leaching
requirement of 0.125, the required drainage rate from Eq 8 equals 1.14 mm/day,
or more than twice the actual value of 0.5. Drainage is therefore limiting, and
according to Eq 8, LF = 0.5/(8 + 0.5) = 0.059. The maximum permissible
Ed is 0.06 X 4, or 0.24 mmho/cm. The available water with an EC of 0.5
mmho/cm has a higher EC than the maximum permissible calculated. The
chloride concentration of the water, 3 meq/liter, also exceeds the maximum
permissible of 1.8 meq/liter, that is, 0.06 X 30. This water is unsatisfactory for
the crop specified unless drainage is improved to equal or exceed 1.14 mm/day,
thereby permitting a LF of 0.125.

Case 2, Only E and I Specified


E and / values specified at 8 and 120 mm/day, respectively, but tc and ti are
not specified. Crop is citrus on sour orange root. Water has an EC of 0.4 mmho/
cm. Leaching requirement is therefore 0.10. I/E = 120/8 = 15, and from Fig.
la, tc/ti is equal to 13.5 for I/E of 15 and LF of 0.10. (The same result is also
obtained by solving Eq 6 for LF of 0.1 and E/I = 1/15.) The irrigation fre-
quency should, therefore, not exceed 13.5 times the infiltration period, that is,
the irrigator can irrigate every 13 days for 24 hr and meet the leaching require-
BERNSTEIN ON QUANTITATIVE ASSESSMENT OF IRRIGATION WATER 65

ment under specified conditions.3 Otc must also exceed (Etc — Iti), as in the
previous examples.

References
[1] U.S. Salinity Laboratory Staff, "Diagnosis and Improvement of Saline and
Alkali Soils," Handbook 60, U.S. Department of Agriculture, 1954.
[2] U.S. Salinity Laboratory Staff, "Diagnosis and Improvement of Saline and
Alkali Soils," Handbook 60, U.S. Department of Agriculture, 1954, p. 79.
[3] U.S. Salinity Laboratory Staff, "Diagnosis and Improvement of Saline and
Alkali Soils," Handbook 60, U.S. Department of Agriculture, 1954, p. 59.
[4] U.S. Salinity Laboratory Staff, "Diagnosis and Improvement of Saline and
Alkali Soils," Handbook 60, U.S. Department of Agriculture, 1954, p. 26.
[5] Leon Bernstein, "Salt Tolerance of Plants," Information Bulletin 283, U. S. De-
partment of Agriculture, 1964.
[6] Leon Bernstein and G. A. Pearson, "Influence of Exchangeable Sodium on the
Yield and Chemical Composition of Plants. I. Green Beans, Garden Beets,
Clover, and Alfalfa," Soil Science, Vol 82, 1956, pp. 247-258; also G. A.
Pearson, "Tolerance of Crops to Exchangeable Sodium," Information Bulletin
216, U.S. Department of Agriculture, 1960.
[7] C. A. Bower and L. V. Wilcox, "Precipitation and Solution of Calcium Car-
bonate in Irrigation Operations," Proceedings, Soil Science Society Am., Vol 29,
1965, pp. 93-94.
[8] C. A. Bower, "Salinity Control in Irrigation Agriculture and Its Effect Upon
Stream and Ground Water Quality," Proceedings, Symposium on Agricultural
Waste Waters, University of California at Davis, April, 1966. Report No. 10,
Water Resources Center, L. D. Doneen, editor, pp. 57-60.
[9] L. V. Wilcox, "Boron Injury to Plants," Information Bulletin 211, U.S. De-
partment of Agriculture, 1960.
[10] H. E. Hayward and L. Bernstein, "Plant-Growth Relationships on Salt-Affected
Soils," The Botanical Review, Vol 24, 1958, pp. 584-635.
[11] Leon Bernstein, A. J. MacKenzie, and B. A. Krantz, 'The Interaction of
Salinity and Planting Practice on the Germination of Irrigated Row Crops,"
Proceedings, Soil Science Society Am., Vol 19, 1955, pp. 240-243; also Leon
Bernstein and Milton Fireman, "Laboratory Studies on Salt Distribution in
Furrow-Irrigated Soil with Special Reference to the Pre-Emergence Period,"
Soil Science, Vol 83, 1957, pp. 249-263.
[12] C. F. Ehlig and L. Bernstein, "Foliar Absorption of Sodium and Chloride as a
Factor in Sprinkler Irrigation," Proceedings, Am. Society Horticultural Science,
Vol 74, 1959, pp. 661-670; also C. F. Ehlig, "Salt Tolerance of Strawberries
under Sprinkler Irrigation," Proceedings, Am. Society Horticultural Science,
Vol 77,1961, pp. 376-379.
STP416-EB/Apr. 1967

Jesse Lunin1

Water for Supplemental Irrigation*

REFERENCE: Jesse Lunin, "Water for Supplemental Irrigation," Water


Quality Criteria, ASTM STP 416, Am. Soc. Testing Mats., 1967, p. 66.
ABSTRACT: There is no single set of water quality criteria for supple-
mental irrigation water because of the interactive effects of the many vari-
ables involved. These include the chemical and physical characteristics of
the soil, the physiological characteristics of the plant, and climatological
conditions. Detrimental effects of poor water quality on plant growth re-
sult from salinity, presence of toxic elements, and indirect effects on soil
properties. Problems vary depending on whether brackish water (contami-
nated by sea water) is used, or effluents from industrial plants. Criteria will
differ depending upon whether irrigation is used for crop production or
land disposal of effluents.
KEY WORDS: water, water quality, irrigation, salinity, toxicity, effluents,
criteria, soils, plant growth, farm crops

Recent droughts and expanding industrialization and urbanization in


the Northeastern United States have focused attention on the need for
water conservation and more efficient use of existing water supplies. In
the humid East, agricultural research is developing management practices
that will make the most efficient use of rainfall and, at the same time, mini-
mize water pollution to protect water supplies needed by other users.
Supplemental irrigation is often necessary in the humid East to insure
maximum production efficiency from fertilizer and soil management prac-
tices. While ample supplies of good-quality water are currently available
in many areas, increasing demands for water may result in the need to use
waters of inferior quality for supplemental irrigation. In coastal areas,
water resources may be contaminated by sea water. Sediment is a major
water pollutant in many areas [I].2 More recently, urban and industrial
areas have contributed to the pollution of both surface and ground water
resources. These pollutants may affect agricultural usage of these waters.

* Contribution from the Soil and Water Conservation Research Division, Agricul-
tural Research Service, U.S. Department of Agriculture.
1
Chief, Northeast Branch, Soil and Water Conservation Research Division,
Agricultural Research Service, U.S. Department of Agriculture, Beltsville, Md.
2
The italic numbers in brackets refer to the list of references appended to this
paper.
66

Copyright© 1967 by ASTM International www.astm.org


LUNIN ON WATER FOR SUPPLEMENTAL IRRIGATION 67

Current agricultural research is directed towards developing criteria for


the safe use of waters of inferior quality.
Cleary [2] has defined the term "criteria" as being "yardsticks of refer-
ence for appraising the suitability of water for various uses; they are guides
to decision making." Bernstein3 has pointed out, however, that no single
set of water quality criteria can be established for irrigation water because
of the many other variables which must be taken into consideration. These
include such factors as the chemical and physical properties of the soil,
the physiological characteristics of the plant, and climatological conditions.
This latter factor includes rainfall distribution and evapotranspirative con-
ditions. For this reason, criteria developed for full-field irrigation in the
arid West will differ in some respects from those developed for supple-
mental irrigation in the humid East. It is the objective of this discussion to
relate water quality requirements for supplemental irrigation to these
environmental factors.

Supplemental Versus Full-Field Irrigation


Basic principles involved whereby pollutants affect plant growth are
essentially the same for both supplemental and full-field irrigation, and
fall into three general categories: (1) high contents of dissolved solids in
the water tend to increase the osmotic pressure of the soil solution, thereby
rendering water less available for plant growth; (2) irrigation water may
contain certain elements which at certain concentrations are toxic to
plants; and (3) water may contain certain elements which, when added to
the soil, may impair soil quality, either directly or indirectly. These pol-
lutants may occur as dissolved solids or suspended colloidal material;
they may be organic or inorganic; they may be derived from natural
sources or result from urban and industrial sources.
Full-field irrigation in the arid West consumes large volumes of water.
Although some sprinkler irrigation is used, surface irrigation is used in
large areas. Large concentrations of salts may accumulate in these arid
areas unless an excess of irrigation water is applied for leaching. In the
humid East, where irrigation merely supplements rainfall, winter rains
are usually adequate to leach from the soil any salts which may have ac-
cumulated during the previous growing season.
Losses of water through evapotranspiration processes cause salt ac-
cumulation in the soil. Climatological conditions in the arid West result
in greater rates of evapotranspiration than those normally found in the
humid East. Hence, hazards from salt accumulation are further minimized
in the East.
Chemical and physical characteristics of soils also vary between these
two regions. Many of the soils in the East tend to be acid, whereas those
3
See p. 51.
68 WATER QUALITY CRITERIA

in the West are neutral to alkaline and may contain free lime. This is an
important factor, since the interaction between the soil and the irrigation
water determines the composition of the resulting soil solution and its
effect on crop growth.
Differences in water quality resulting from natural contaminants of
water resources is another source of variation. In the arid West, total salt
concentration does not vary too widely, but ionic composition may vary
considerably depending upon the source [3]. By contrast, brackish waters
sampled along the Eastern seaboard have a wider range in total concentra-
tions but have been found to contain calcium, magnesium, sodium, and
potassium in the same ionic proportions are those found in sea water [4].
The effects of urban and industrial pollutants in water will depend largely
upon the specific nature of pollutants involved.
The plant factor is an important one since there is considerable variation
in tolerance to overall salinity and to specific ions in water. Differences in
cropping practices represent, therefore, another source of variation be-
tween the humid East and the arid West.
All of the above factors are interrelated and present a rather complex
picture. A study of water quality in irrigation must take into consideration
the interactive effects of these various factors. Examples of some of the
interrelationships will be presented later.

Evaluation of Water Quality for Supplemental Irrigation


Many methods are available for the assessment of water quality, but
few relate directly to water used for irrigation. Characteristics such as
general appearance, taste, and odor cannot be related directly to plant
growth. These factors have only aesthetic value. The pH of water is of
some significance. In waters that have not been polluted by urban and in-
dustrial wastes, pH values are seldom low enough to be a problem, since
the soil itself is a buffered system. Very low pH values may be indicative
of pollution. Values as low as pH 2.80 have been reported [5] and were
a result of drainage from coal mines. In these waters, one might also
suspect the presence of toxic levels of certain metals. High pH values
could be indicative of the presence of other ions toxic to plants, such as,
bicarbonates.
Total suspended solids in water may be significant in determining water
quality for supplemental irrigation, since sediment is one of the largest
single pollutants in surface waters [7]. In addition, other suspended ma-
terials resulting from urban and industrial pollution must be considered.
These may be organic or inorganic. Deposition of fine sediments on the
soil surface may result in severe crusting that could deter seedling emer-
gence. Fine sediments may result in the plugging of soil pores, thereby re-
ducing soil aeration and the infiltration of water. Sprinkler irrigation with
LUNIN ON WATER FOR SUPPLEMENTAL IRRIGATION 69

water containing sediment may affect the quality of crops where the
marketable portion consists primarily of the leaves, such as tobacco and
leafy vegetable crops. There is also a possibility that suspended materials
may plug up sprinkler nozzles.
Dissolved oxygen has been mentioned as a water quality criterion. In
agriculture, we are concerned not with the oxygen content of the water
per se, but rather with the effect that irrigation with a given water will
have on the availability of oxygen in the soil for plant growth. This is of
little concern where unpolluted waters are used. In attempting to use
certain urban or industrial effluents for irrigation, the presence of readily
reducible organic compounds may result in the reduction of available
oxygen in the soil. Measurement of biological oxygen demand (BOD) or
chemical oxygen demand (COD) values would be indicative of this factor.
Unfortunately, very little information is available to evaluate the change
in BOD value of water as it is affected by sprinkler irrigation and infiltra-
tion into the soil, and the resultant effect on oxygen availability in the soil.
Total dissolved solids is one of the most significant single determinations
of water quality for agricultural use. Since these dissolved solids usually
occur as inorganic salts, the determination of the electrical conductivity
(EC) of water is a most useful measurement and may be related to the
osmotic pressure of the solution. The presence of certain dissolved organic
compounds in some polluted waters may also contribute to the osmotic
pressure of a solution; yet, this would not be reflected in the electrical con-
ductivity measurement.
Specific ion determinations are very important. In both the arid West
and humid East, the presence of sodium in water could result in the ad-
sorption of this element by the soil, causing poor physical conditions. The
sodium adsorption ratio (SAR);4 proposed by the U.S. Salinity Labora-
tory [6], is the ratio of sodium to divalent cations in the water. It is an
indication of the degree to which sodium will be adsorbed by the soil from
a given water when brought into equilibrium with it. The presence of
boron and bicarbonate in western waters is of greater significance than in
eastern waters. Yet these, along with other elements, may be a deterring
factor in the use of some eastern waters where they are subject to urban
and industrial pollution. The potential pollutants from industrial wastes
have recently been listed by Wilcox [7].
One final criterion relates to the microbial population of the water.
This is normally of little consequence in using water for irrigation where
no pathogens are involved. Where sewage effluents are used, or allowed to
contaminate agricultural water supplies, this may be a significant factor
where food crops are produced.
4
SAR = NaV([Ca+* + Mg++]/2)1/2 (ion concentrations in me./liter).
70 w
WATER QUALITY CRITERIA

Soil-Plant-Water Interactions
As previously stated, it is difficult to establish specific water quality
criteria for irrigation purposes since these will vary for different soils,
crops, and environmental conditions. The following discussion is designed
to point out some of these interactions and their relation to supplemental
irrigation in the humid East.
Soil water availability to plants is a function of the total soil water po-
tential. This value is the sum of the matric suction (physical attraction of

FIG. 1—Effect of drying on osmotic pressure of the soil solution.

the soil for water) and the solute suction (osmotic pressure of the soil water,
OP). Under nonsaline conditions, the latter factor has little significance.
Where brackish, or saline, water is used for irrigation, the solute suction is
sxtremely important since the salt content of the soil solution increases
rapidly as the soil dries. The effect of soil drying on the osmotic pressure
of the soil solution is shown in Fig. 1, which represents the addition of one
inch of several dilutions of synthetic sea water to a fixed volume of soil.
Since the matric suction of the soil increases logarithmically upon drying,
the combined effects can produce critical conditions with regard to soil
water availability. Where saline water is used for supplemental irrigation,
it is possible to use water having higher salt levels than would be permissi-
LUNIN ON WATER FOR SUPPLEMENTAL IRRIGATION 71

ble under full-field irrigation. As a result, there is a greater possibility of


damage.
One of the major differences between salinity problems in full-field irri-
gation and supplemental irrigation is the fact that the soil is usually non-
saline at spring planting in the humid East. Use of saline water for supple-
mental irrigation may result in a temporary buildup of salts in the soil,
but the leaching action of rainfall prevents a permanent accumulation
under average conditions. The distribution of salts in the soil profile is
also likely to be much more variable throughout crop growth where supple-
mental irrigation is used because of the effect of intervening rainfall.
Plants vary in their tolerance to salinity, but established tolerance values

FIG. 2—Effect of growth stage at time of salinization on yield of beets.

[6] are generally applicable to both arid and humid areas. Plant response is
also dependent upon the plant growth stage at which salinization occurs.
This is a significant factor where saline water is used only for supplemental
irrigation. Different portions of the plant may also react differently to
saline conditions. An example of this is shown in Fig. 2 [8], where, in a
greenhouse study, salinization of beets with dilute sea water for the A
treatment occurred in the seedling stage and that for B and C at two and
four weeks later. Yields are shown as a function of the electrical conduc-
tivity of a soil saturation extract, ECe, at harvest. Results indicated that
growth stage at time of salinization had little effect on yield of tops but
greatly affected the yield of roots. These results could be due either to the
fact that the plant was more sensitive to salinity in the early growth stage,
or to the fact that the plant was exposed to salinity for a longer period of
722 WATER QUAL'TY CRITERIA

time, or to a combination of both factors. It is apparent that where saline


water is used for supplemental irrigation, growth stage of the crop at
which salinized and duration of salinization in the soil must be taken into
consideration.
The SAR value of water is an important criterion since it is indicative of
the effect a given water will have on soil structure as a result of the amount
of sodium adsorbed. Most water of inferior quality currently being used

FIG. 3—Estimation of SAR values of a synthetic sea water from electrical con-
ductivity determinations.

along the Eastern seaboard results from contamination by sea water [4].
This being the case, ion distribution in the water as a function of concen-
tration would be essentially equivalent to dilutions of sea water, and the
SAR value can be estimated from the EC value by the graph shown in
Fig. 3. Where other pollutants are involved, however, the water should
be analyzed and the SAR value calculated.
The clay and some organic fractions of the soil are negatively charged
and function as an ion-exchange material. Any given soil, therefore, has a
characteristic cation exchange capacity. Acid, or base-unsaturated, soils
contain varying amounts of exchangeable hydrogen and aluminum in addi-
LUNIN ON WATER FOR SUPPLEMENTAL IRRIGATION 7
73

tion to exchangeable bases such as calcium, magnesium, potassium, and


sodium. Soils containing no exchangeable hydrogen or aluminum are base
saturated. The level of soil base saturation, as indicated by soil pH, has a
significant effect on the reaction between saline water with a given soil,
and the resultant influence on soil characteristics and plant growth.
The U.S. Salinity Laboratory [6] has established a relationship between
the SAR value and the resultant exchangeable sodium percentage (ESP)5
of a soil brought into equilibrium with that water. This relationship holds
well for western soils that are predominantly base saturated. Acid, or base-
unsaturated, soils frequently found in the humid East react differently.
The amount of sodium adsorbed by these soils is a function of the level of
base saturation, as well as the composition of the water applied as supple-
mental irrigation. This is illustrated in Table 1 [9], where an acidified

TABLE 1—Exchangeable sodium percentage in a Norfolk fine sandy loam as


influenced by the interaction of initial level of base saturation and treatment with
various saline solutions.
Soil, % base SAR of Equilibrating Solution"
Soil pH saturation a—21.2 b— 15.0 c—10.6 d—7.5

2.9 2.2 6.5 4.7 3.1 2.7


3.6 18.6 7.2 5.0 3.4 2.8
4.4 . . 35.7 8.4 5.7 4.0 3.3
5.2 50.7 9.2 6.4 4.5 3.9
6.0 66.7 10.1 6.8 5.1 4.3
6.7 82.7 10.7 7.9 5.2 4.4
"Solution compositions (me./liter): (a) Na—60, Ca—16; (6) Na—30, (c)
Na—15, Ca-^; (d) Na—7.5, Ca—2.

Norfolk soil was limed to give six levels of base saturation, and then
equilibrated with four dilutions of a solution containing sodium chloride
(NaCl) and calcium chloride (CaCl2) in the same relative proportions
found in sea water. The ESP decreases with decreasing SAR, and decreas-
ing level of base saturation in the soil. The interactive effect of ESP and
base saturation may be reflected in crop yields as shown in Fig. 4 [10].
This represents yield of bean pods grown in a greenhouse study on a
Bladen soil in which four levels of exchangeable sodium were established
at each of three levels of base saturation. It is apparent that base saturation
levels affect both the amount of sodium adsorbed by a soil from a given
solution, and the resultant yield of crops grown on that soil.
The interactive effects of salinity and soil acidity are significant even
where adsorbed sodium is not a problem. Addition of a neutral or acidic
saline solution to an acid soil tends to decrease the pH of that soil and
alter the composition of resulting soil solution in equilibrium with it. A
secondary effect of acidity created is the possibility of rendering soluble
s
ESP = adsorbed sodium expressed as a percentage of the total cation exchange
capacity.
74 WATER
w QUALITY CRITERIA

iron, aluminum, and manganese in concentrations toxic to plant growth.


This will vary with the specific chemical characteristics of the soil and
may occur where brackish waters are used for supplemental irrigation in
humid regions. This was demonstrated in a recent study [11], in which
an acid Sassafras fine sandy loam was limed to give three pH levels and
used in a greenhouse study to determine the effect of four levels of salinity
on the growth of beans. Salinization was accomplished by irrigating with
different dilutions of synthetic sea water during the growth of the crop. The

FIG. 4—Interactive effect of exchangeable sodium, ESP, and base saturation on


yield of bean pods.

results (Table 2), show both the detrimental effect of high salinity levels
on the growth of beans and the beneficial effect of liming. Liming increased
the yield of bean pods at all salinity levels, indicating that the low yields
of the unlimed beans were due to factors other than salinity alone.
Salinization resulted in a substantial decrease in soil pH which was ac-
companied by an increase in soluble manganese. This was also reflected
by a large increase in the manganese content of the plant. It is apparent
that the large decrease in yield on the unlimed soil was due to manganese
toxicity as well as to Salinization.
Soil Salinization by supplemental irrigation also alters the cation compo-
sition of the soil solution. This effect is further complicated by the varia-
LUNIN ON WATER FOR SUPPLEMENTAL IRRIGATION 75

tions in level of base saturation in acid soils. This was demonstrated in a


greenhouse study [11] in which an acid Portsmouth silty clay loam was
limed to give four levels of base saturation: A—24 per cent, B—42 per
cent, C—60 per cent, and D—77 per cent. Beans were grown and irri-
gated with saline water to give five salinity levels. The resultant effect on
the calcium concentration is shown in Fig. 5. Comparable results were
obtained with magnesium, but there was little or no effect of base satura-
tion on the sodium and potassium content. Leaching studies with dilute
sea water have shown that continued use of this water could result in a
gradual depletion of calcium in the soil profile.

TABLE 2—Relation of yield of beans to soil salinity, pH, and NH^O Ac-soluble Mn.
Soil Plant „
Soil pH EC ° Y iMrl r»f P/vle o
Mn6 ' ppm Mn ' /v

4.70 0.41 32.7 15 0.049


4.48 .. 3 02 11 2 18 0.120
4.30 ... 5.69 19 22 0.141
4.27 11.45 0 27 0.136
4.98 0.40 50.6 10 0.023
4.76 2.83 33.7 6 0.059
4.64 5.67 12.6 17 0.095
4.64 11.97 0 21 0.073
5.69 0.48 48 8 3 0.017
5.51 . 3.16 34 2 6 0.033
5.36 5.81 20 6 5 0.048
5.31 11.69 1.0 9 0.066
0
Electrical conductivity of saturation extract.
5
Extractable with ammonium acetate (pH 4, 8).

Selection of Water Quality Criteria


The previous discussion has dealt primarily with salinity. Irrigation
waters for full-field irrigation have been classified both as to salinity and
alkalinity hazards, using electrical conductivity and the SAR value as cri-
teria [6]. This classification is not necessarily applicable where supple-
mental irrigation is used because of factors previously discussed. In humid
regions, because of the leaching action of intervening rains, higher salt
contents can be tolerated in the water. Exchangeable sodium is not likely
to be a problem because of the general composition usually found in
brackish waters and the acidity of the soil. To aid farmers in evaluating the
effect of brackish water irrigation on soil salinity, the following equation
was proposed [12]:

T?fi ry-r I n\lL\^i


•£<-«(/) = -fcC^-) -+-
=—
w)

where:
76 WATER QUALITY CRITERIA

Salinity level

FIG. 5—Calcium content of soil solution as affected by levels of base saturation


and salinization.

ECe(f) = electrical conductivity of the soil saturation extract after


irrigation is completed,
ECea) = electrical conductivity of the soil saturation extract before
irrigation,
ECiw = electrical conductivity of the irrigation water, and
n = number of irrigations.
The maximum permissible level of soil salinity, ECe(f), is determined by
the salt tolerance of the specific crop to be grown. This equation can be
used to enable the farmer to predict the number of irrigations permissible
for a crop with water having a known level of salinity, assuming there
is no intervening rainfall. This has been a fairly good guide in practice to
date.
Little has been said thus far regarding specific ion toxicities resulting
from natural sources. No evidence has been found to date to indicate
that boron or bicarbonate would be a problem in the humid East. If
it were, it is probable that toxic limits used in the West [6] would be
applicable. We are concerned with the indirect effects resulting from
soluble aluminum, iron, and manganese which are rendered soluble in
acid soils by saline irrigation. An example of this was previously cited.
Toxic levels of these elements for certain crops have been determined,
LUNIN ON WATER FOR SUPPLEMENTAL IRRIGATION77
77

but it is not possible to use this as a criterion for water quality because
it is an indirect effect and is dependent upon soil characteristics. It must
be taken into consideration as a potential hazard, however.
Rapidly expanding industrialization and urbanization in the East has
resulted in an increasing possibility of pollution of water resources. The
number of potential pollutants is large and, as pointed out by Wilcox
[7], the amount of quantitative information on these pollutants relative
to water quality requirements for irrigation is scarce. Certain of these
pollutants might be detrimental to plant growth only by virtue of their
contribution to salinity. Existing criteria would cover this. Certain ele-
ments, such as the heavy metals, are detrimental because of their specific
toxicity to plants. Since the degree of toxicity may vary with different
soils and crops, additional research is necessary to establish effective
criteria.
Reference has been made to water pollution by industrial effluents.
Frequently, periods of drought in the East coincide with periods of low
streamflow. Water for supplemental irrigation during these periods is
scarce. The use of these effluents as a potential source of water for
supplemental irrigation of crops should be considered. Water quality
criteria previously discussed would be applicable here, but the chemical
composition of these industrial effluents will vary considerably from that
caused by natural contaminants. For this reason, a more complete char-
acterization will be necessary to determine the extent to which they can
be used. Preliminary studies on grass, corn, and vegetables [13, 14, 15]
indicate that this is feasible but that more information is needed regarding
effects of various constituents on crops and soils.
Another aspect of waste water utilization relates to sprinkler irriga-
tion for land disposal. Although this is not directly concerned with
supplemental irrigation for crop production, the determination of water
quality criteria is essential in order to relate this practice to the effect on
soils and plants. Water quality criteria will, of necessity, be different
because of the large quantities of water involved and the fact that crop
production is not necessarily a factor. Furthermore, we are concerned
with the potential pollution of ground water supplies. It is beyond the
scope of this discussion to present any great detail on this subject, but
the importance of relating water quality criteria to this practice should
be stressed. Considerable information is available regarding the en-
gineering aspects of waste disposal, but little is available regarding the
agronomic aspects.

Conclusions
Adequate water quality criteria for supplemental irrigation are availa-
ble regarding salinity factors, even though these criteria must be related
to specific crop, soil, and environmental conditions. Although some infor-
78
78 WATER QUALITY CRITERIA

mation is available on toxic levels of some specific elements, the increasing


hazard of industrial pollution of water resources make it imperative that
more quantitative information be obtained regarding the effect of different
trace elements on soils and crops. Here, again, soil and plant character-
istics must be considered. There is a great need for extensive research to
develop irrigation criteria for water subject to pollution from various urban
and industrial sources.
reference
[1] C. B. Brown, "Effect of Land Use and Treatment on Pollution," Proceedings,
Nat. Conference on Water Pollution, U.S. Department of Health, Education,
and Welfare, Public Health Service, December, 1960, pp. 209-219.
[2] E. J. deary, "Development of Water Quality Standards," Highlights, Water
Pollution Control Federation, Vol 3, No. 413,1966.
[3] L. V. Wilcox, "Classification and Use of Irrigation Waters," U.S. Department
of Agriculture Circular No. 969, 1955.
[4] M. H. Gallatin, J. Lunin, and A. R. Batchelder, "Brackish Water Sources for
Irrigation Along the Eastern Seaboard of the United States," USD A Production
Research Report No. 61, 1962.
[5] J. E. Biesecker and J. R. George, "Stream Quality in Appalachia as Related to
Coal Mine Drainage, 1965," U.S. Geologic Survey Circular No. 526, 1966.
[6\ "Diagnosis and Improvement of Saline and Alkali Soils," U.S. Department of
Agriculture Handbook No. 60, U.S. Salinity Laboratory Staff, 1954.
[7] L. V. Wilcox, "Effect of Industrial Wastes on Water for Irrigation Use," Water
and Industrial Waste Water, ASTM STP 273, Am. Soc. Testing Mats., 1959,
pp. 58-64.
[8] J. Lunin, M. H. Gallatin, and A. R. Batchelder, "Saline Irrigation of Several
Vegetable Crops at Various Growth Stages. I. Effect on Yields," Agronomy
Journal, Vol 55,1963, pp. 107-110.
[9] J. Lunin and A. R. Batchelder, "Cation Exchange in Acid Soils Upon Treat-
ment with Saline Solutions," Transactions, International Congress of Soil Sci-
ence, Vol 1, No. 6, 1960, pp. 507-515.
[10] J. Lunin, M. H. Gallatin, and A. R. Batchelder, "Interactive Effects of Base
Saturation and Exchangeable Sodium on the Growth and Cation Composition
of Beans," Soil Science, Vol 97, 1964, pp. 25-33.
[11] J. Lunin, M. H. Gallatin, and A. R. Batchelder, "Effect of Saline Water on the
Growth and Chemical Composition of Beans: II. Influence of Soil Acidity,"
Proceedings, Soil Science Society of Am., Vol 25, January-February, 1961, pp.
372-376.
[12] J. Lunin and M. H. Gallatin, "Brackish Water for Irrigation in Humid Regions,"
USDA ARS-41-29, January, 1960.
[13] "A Survey of Direct Utilization of Waste Waters," Publication No. 12, Cali-
fornia State Water Pollution Control Board, 1955.
[14] S. C. Crawford, "Spray Irrigation of Certain Sulfate Pulp Mill Wastes, Sewage
and Industrial Wastes, Vol 30,1958, pp. 209-219.
[75] L. L. McCormick, "Effects of Paper Mill Wastes on Cattle, Crops and Soils,"
Louisiana State University Bulletin No. 259, 1959.
Recreational Use
This page intentionally left blank
STP416-EB/Apr. 1967

C. B. Wurtz^

Water Use for Aquatic Life

REFERENCE: C. B. Wurtz, "Water Use for Aquatic Life," Water Quality


Criteria, ASTM STP 416, Am. Soc. Testing Mats., 1967, p. 81.
ABSTRACT: Aquatic life in general is quite tolerant of deteriorated
water. Aquatic organisms vary in their responses to environmental altera-
tion from extreme sensitivity to equally extreme tolerance. Compared to
the quality of water required for domestic and industrial use, the aquatic
life of the nation is not very demanding. To protect all water-use interests
pollution must be defined, criteria for natural waters developed, and
standards for specific discharges fixed. It is probable that no single, equita-
ble set of criteria for aquatic life can be established over a geographic area
larger than the basin of a first- or second-order tributary of our major
rivers. Water quality change brings about a concomitant biological
change, but extreme conditions must exist and be persistent before aquatic
life is eliminated. Environmental control is essential if optimal conditions
for aquatic life are to be developed and maintained. Waste disposal, prop-
erly controlled, need not conflict with the use of water for aquatic life.
KEY WORDS: criteria, standards, water pollution, organic waste, indus-
trial waste, dissolved oxygen, fish, eutrophication, sewage, stream classi-
fication

Practically all publicity about water pollution refers to debilitating


effects on aquatic life, particularly game fish. However, aquatic life in
general is quite tolerant of deteriorated water quality. Aquatic organisms
vary in their responses to environmental alteration from extreme sensi-
tivity to equally extreme tolerance. It is impossible to establish criteria for
aquatic life without taking this variability into consideration.
Aquatic organisms are widely tolerant of variation in water-quality
characteristics until maximum or minimum thresholds are crossed and
their natural range of tolerance is violated. Many other water uses are
far more demanding of cleaner water. For example, Dunbar and Henry
[I]2 tabulated several recommended stream quality standards with stream
classification systems. Five types of stream quality were presented. The
first was the highest quality; the fifth was the poorest quality. The fifth,
1
Assistant professor, Biology, La Salle College, Philadelphia, Pa.
2
The italic numbers in brackets refer to the list of references appended to this
paper.
81

Copyright© 1967 by ASTM International www.astm.org


ater
82 WATER QUALITY CRITERIA

considered unsuitable for water supplies, was suitable for recreation and
fish life. Compared to domestic and many industrial water quality needs,
the aquatic life of the nation is not very demanding.
Criteria and Standards
To protect all water-use interests pollution must be defined. Any defini-
tion developed must be broad enough to specifically include items injurious
to the best interests of society and specifically exclude items beneficial to
the best interests of society. For this reason the definition of pollution must
be based on a sound consideration of limiting values for desirable and
undesirable water quality characteristics. In short, criteria must be de-
veloped. This has been done, of course, in many states and under the
aegis of many different agencies.
As I understand and use the terms, criteria differ from standards by
defining all the water quality characteristics of a body of water and by
being subject to review and modification as knowledge increases or condi-
tions change. Standards, in turn, would constitute specific limiting values
for particular environmental factors. This differs from the usual dictionary
definition, where the terms are considered synonymous, but does reflect a
common use of the terms in pollution literature.
Industrial Discharges
Criteria do not necessarily imply ideal conditions for every specific
interest. Ideally it would be desirable to exclude all toxicants from water,
but this simply cannot be done. There are many naturally occurring toxi-
cants present in water. Because such materials were present long before
man learned to walk erect it is apparent that aquatic organisms have long
since developed a modus vivendi with such conditions. Heilbrunn [2] ob-
served: "One of the most striking and, from the standpoint of the continu-
ance of life on this planet, one of the most important aspects of acclimatiza-
tion is the capacity of living systems to adjust themselves to the presence
of poisonous substances."
Limiting values have been designed for many potential toxicants, and
these permit at least some minimal amount to be discharged. Such stand-
ards usually permit discharges below the minimal amount that will, in
fact, be injurious to aquatic life as determined by experience and bioas-
say studies. Such discharges do not affect the use of the water by aquatic
life.
Unfortunately the quantitative determination of the amount of any
given substance in water is never exact. This must be taken into considera-
tion when water quality criteria are established. For example, heavy metal
ions, in general, are highly toxic to aquatic life (although most are essential
nutrients in micro quantities). Each metal differs in its degree of toxicity.
Doudoroff and Katz [3] critically reviewed the literature on metal toxicity.
WURTZ ON WATER USE FOR AQUATIC LIFE 83

Cumulative experience as summarized in their article indicates that the


most toxic metal is silver, followed successively by mercury, copper, lead,
cadmium, gold, aluminum, zinc, nickel, and chromium. Other metals
appear to be much less toxic. To the extent that killing amounts of metal
are added to water in laboratory experiments, and only to that extent, can
we stipulate that a certain concentration is toxic. When aquatic life is de-
stroyed in a stream in association with a discharge of metal ions, we can-
not accurately measure the concentration that caused the kill. A report of
the Analytical Reference Service of the U.S. Public Health Service clearly
demonstrates this. Inherent weaknesses in analytical methods produce
widely divergent results. In a prepared mixed sample containing eight
metals, zinc was present in a concentration of 0.9 mg/liter. Forty-eight
agencies contributing data to that report performed 136 analyses in tripli-
cate using various analytical methods. The reported zinc concentration in
the sample ranged from 0.09 to 2.30 mg/liter (mean 0.94 mg/liter).
Zinc occurs naturally in water and apparently is not highly toxic to
humans. Data from drinking waters compiled from 37 locations in the
United States showed a mean value of 0.136 ppm of zinc. Wurtz [4] found
92 species of bottom organisms in the Portage River, N. B. Geochemical
studies in this region showed bottom sediments containing 50 to 300-ppm
zinc along with 25 to 75-ppm copper and up to 400-ppm lead. In the
water zinc was reported as 0.005 ppm. Mullican et al [5] found that 65
ppm of zinc in an industrial waste acted as a biological depressant in the
Nolichucky River, Tenn. At this concentration a resident bottom fauna
was reduced from 2934 individuals (representing 30 genera) per square
foot to 46 individuals (representing 22 genera) per square foot. In natural
waters not subject to mine drainage or industrial wastes zinc concentra-
tions may range to 0.2 ppm.
In view of the preceding comments on zinc, what water quality criteria
can be established that will include this metal, and how is the concentration
to be measured? This is a question that must be resolved for every organic
and inorganic substance discharged from any source. In addition to the
formulation of criteria that include toxicity, other considerations must be
given close attention. These include additive effects, synergistic reactions,
antagonistic reactions, temperature increments that affect biological rates,
all other factors that affect biological rates, the individuality of every
stream and lake, the variable responses of the whole spectrum of aquatic
species, and so forth, virtually ad infinitum. It is probable that no single,
equitable set of criteria for aquatic life can be established over a geographic
area larger than the basin of a first- or second-order tributary of our major
rivers.
Recent federal legislation indicates that river-basin coherence will in-
deed be the approach to water pollution control. The Federal Water Re-
sources Planning Act enables the Department of the Interior to establish
844 WATER QUALITY CRITERIA

river basin commissions as well as a Water Resources Council. The latter


will establish policy on a regional or basin basis.

Heated Discharges
The need for the regional approach is obvious. All environmental fac-
tors vary seasonally and geographically. The most pronounced example
of this is water temperature. We all recognize that, over the years, our
surface waters have warmed. The public is quick to point at industrial
thermal discharges as the culprit. However, it is unlikely that the cumu-
lative effect from industrial discharges has equaled the cumulative effect
of opening a continent by deforestation and agriculture.
A recent newspaper item drives home the concept of cumulative temper-
ature effects. It would appear the Columbia River has been warming
through the retention of water by dams, and, with further river develop-
ment, fishery authorities estimate that the river's temperature will reach
29.5 C by 1980. This is considered to be about four to five degrees above
the maximum temperature tolerated by salmon. This could prove to be
true if development rates hold constant and no control is exerted.
Water temperature does have a profound influence on aquatic life.
Innumerable examples of major biological change due to a persistent
difference of a few degrees can be cited. However, this is change, not
destruction. Temperature-wise many streams are borderline trout streams,
and an average annual temperature increase of one or two degrees centi-
grade can alter such streams so that they will not support trout. This has
undoubtedly happened many times in the historic development of our
nation. Brook trout in Pennsylvania are probably not found in 50 per cent
of the streams where they were resident 100 years ago. What trout fisher-
men fail to recognize, however, is that there probably now exists more
trout water in Pennsylvania than there was 100 years ago. This is ac-
counted for by the introduction of the western rainbow trout and the Euro-
pean brown trout. Both these species tolerate higher temperatures than
the brook trout. A few years ago a strain of rainbow trout was developed
by a private hatchery that was resistant to water temperatures as high as
29 C. Selective breeding of fish is common today and may someday be
comparable to that of domestic animals.

Organic Discharges
In recent months the pollution of Lake Erie has been much in the news.
Costs of renovation have been estimated as high as ten billion dollars.
Unquestionably Lake Erie has been receiving a massive influx of pollution.
Much of this loading includes materials that represents biological nutri-
ents. These fertilizers have increased the rate of eutrophication of the lake.
The geologic history of all lakes leads ultimately through eutrophication
to extinction. Man's activities simply alters the rate. Most articles dealing
WURTZ ON WATER USE FOR AQUATIC LIFE 85

with the Lake Erie problem overlook a basic phenomenon: during the
five-year period from 1925 to 1930 the average annual temperature of the
lake was increased from 10.5 C by about one degree according to Beeton
[6]. Previous to 1928 the bottom fauna was dominated by mayfly nymphs,
but now the dominant bottom organisms are oligochaete worms and midge
larvae, well-known pollution-tolerant forms. Their abundance today ap-
pears to be the damning evidence for pollution of the lake. It is important
to note that these changed environmental conditions do not deny the use
of the water to aquatic life. Water quality change brings a concomitant
biological change, but extreme conditions must exist and be persistent be-
fore aquatic Me is completely eliminated.
In November, 1965, the Michigan State Chamber of Commerce issued
a special report on Lake Erie fish. This report states that the ecology of
Lake Erie has changed drastically, due in part to pollution and in part to
the natural aging process. For the past two or three years the overall
catch of walleyes has been low, but the least productive walleye years of
record were 1913 to 1920. From that low the catch rose to five million
pounds per year by 1957. Currently the productivity curve for this species
appears to be falling. This reflects characteristic biological fluctuation.
This Michigan report includes the statement: "Apparently Lake Erie has
been chosen as the 'demonstration project,' to demonstrate to the rest
of the Nation the newfound strength and capabilities of the Federal
Government. At this moment they have everything in their favor, for
these waters are not as dead as some would have us believe."
Eutrophication, an aging process, is readily recognized by the increased
development of biological growths. Successive growth stages represent
steps along a continuum extending from oligotrophy to extinction. From
the viewpoint of water use the most objectionable feature of eutrophica-
tion is the possible development of anaerobic conditions and nuisance
growth of algae. However, increased biological production also includes
increased production of bottom organisms and fish. If an increase in
aquatic life as expressed by fish production is desired, eutrophication must
be accepted. The ideal situation for fish production would be development
of the optimal carrying capacity of the water and then holding environ-
mental conditions constant. Unfortunately, nothing in nature is stable,
and achieving such an ideal may be impractical. Nevertheless environ-
mental control is essential if optimal conditions are to be developed and
maintained.
The occurrence of anaerobic conditions in a stream or lake subject to
overloading with organic wastes represents the most undesirable condi-
tion for aquatic life. This is also the most offensive condition to the pub-
lic. Historically the control of anaerobic conditions developed early in
the United States. By 1890 the chief engineer of the Massachusetts State
Board of Health concluded that a stream flow of 2.5 ft3/sec per thousand
866 WATER QUALITY CRITERIA

sewered population was insufficient to prevent offensive conditions, but


that a flow of 7 ft3/sec would be adequate. This represents the concept of
pollution control by taking advantage of the assimilative capacity of the
receiving stream. Organic wastes are biologically reduced in a receiving
stream and the amount a stream can handle represents its assimilative
capacity. This is disposal by dilution. In the absence of excessive over-
loading that will produce anaerobic conditions this method of nontoxic
organic waste disposal is not injurious to aquatic life. Necessary levels of
dissolved oxygen must, of course, be maintained. In point of fact, decom-
posable organic wastes enhance biological productivity through fertiliza-
tion.

Assimilative Capacity
Disposal by dilution is still practiced; it always will be. The nation's
watercourses must carry the rejectamenta of man, but this need not be
injurious to aquatic life. Today waste treatment plants are designed to
reduce the burden on the receiving water. However, no treatment plant
is 100 per cent efficient. In the United States the overall efficiency of
municipal sewage plants is expected to reach 70 per cent by 1980. The
difference between the percentage efficiency and 100 per cent represents
the waste load being disposed of by dilution. Whether we wish it or not,
the assimilative or self-purification capacity of a receiving body of water
must be calculated into all waste disposal designs. Whether all interests
agree or not, waste disposal is a valid water use. It is also an inevitable
water use since there is no alternate solution to ultimate waste disposal.
The universality of waste disposal via the nation's waterways is recog-
nized by every public and private agency associated with pollution. The
very fact that, under the Water Quality Act of 1965, criteria are to be es-
tablished reflects cognizance of this. Otherwise enforcement agencies could
simply demand that no discharge except water at ambient temperatures
be permitted. Water quality criteria and standards are not new. Many
states now have limiting values for common water quality characteristics.
Such values are restrictive enough to protect local aquatic life, although
the particular kinds of organisms present may be altered. In general, such
an alteration is only meaningful to fishermen who selectively angle for
particular fish species.

Conflicts with Fish


In a recent interview the Executive Director of the Izaac Walton
League of America recalled a demonstration in which a carp, swimming
in an industrial effluent, was used to demonstrate the degree of purity of
the water. The Executive Director remarked, "... was a carp and a carp
can live in water that is practically solid." I would like to point out that
this fish was indeed reflecting the existence of water quality suitable for
WURTZ ON WATER USE FOR AQUATIC LIFE 87

aquatic life. In completely unpolluted waters there are more acres and
miles of carp water than there are for any so-called game fish species.
The interviewee also said, "We believe that, ideally, all water should
be fit for direct human use without special treatment." Presumably he
meant human consumption. Such a condition has never existed in North
America—not even before 1492. Ideals are fine, but they must be real-
istic if they are to be realized. The ideal suggested in this interview can
never be realized. I, for one, for example, will hold to the view that
chlorination is desirable, although I know chlorine is a toxicant for
aquatic organisms.
Like most anglers, I have found a satisfying solution to the conflict
between recreational fishing and water quality alteration due to com-
merce and public use. I simply go to a body of water where the species
I want lives. Actually I prefer this because it gets me away from the
things I probably should be doing. I suspect that the majority of the
members of the Izaac Walton League prefer to fish at a locale some
distance from their home.
Stream Classification
Whether we formally recognize it or not, we do classify our surface
waters and particularly our streams. There are streams we consider trout
streams, others we call wild rivers, still others we call warm-water
streams, etc. The need is to formalize this classification and establish
criteria for each class within each river basin. Several states classify their
waters now, and apparently it is a workable arrangement. Any of these
classificatory arrangements may be used as a model. I am most familiar
with one proposed for Pennsylvania. In this model several stream classes,
each meeting different criteria, were proposed. These included:
1. Trout streams—those streams which occur at elevations above 132
m in eastern Pennsylvania, and above 264 m in western Pennsylvania;
which have velocities of 1 ft/sec or more; which have maximum, natural,
warm-weather temperatures below 24 C; and which are able to support
a resident, breeding population of trout.
2. Intermediate streams—those streams which maintain some flow at
all times; which have scouring velocities during the spring months at
least; which have stable bottoms; which have maximum, natural, warm-
weather temperatures below 29 C; and which are able to support a
resident, breeding population of smallmouth bass or walleye pike or both.
3. Warm-water streams—those streams which are subject to long
periods of quiescence or very low velocities; which are chiefly stretches of
deposition of suspended load and bedload materials; which have maxi-
mum, natural, warm-weather temperatures that may exceed 29 C; and
which are able to support a resident, breeding population of large-mouth
bass.
8
88 WATER QUALITY CRITERIA

4. Acid streams—those streams subject to a mineral acid load great


enough to lower the hydrogen ion concentration of the water to a pH
value of 4.0 or less, and which will not support fish life for this reason.
In view of recent legislation in Pennsylvania (HB 585, effective Jan. 1,
1966) this classification would probably not be acceptable in that state.
It is unlikely, however, that the 2906 miles of Pennsylvanian streams cur-
rently recognized as acid by the Pennsylvania Fish Commission will be
relieved of this burden within the forseeable future. Approximately half
the acid in the state's streams stems from abandoned mines. The cost of
sealing these must be borne by the state. (Pennsylvania Senate Bill 1024,
now signed into law by the governor, appropriates a half million dollars
for an active program of correcting stream pollution from abandoned
deep and strip mines.) One provision of HB 585 grants tune extensions
for completing treatment plans. To date, no economical treatment tech-
nique that can be considered broadly applicable is known.
5. Industrial streams—those streams or stream stretches where, dur-
ing low-flow periods at least, 100 per cent of the stream's volume is used
by industry; where the water's characteristics are pronouncedly altered
from upstream water; and which will not support a resident, breeding
population of any game fish species.
Transitional areas will occur between any two classes that abut each
other. In addition, nonindustrial classes may alternate for some distance
depending upon the physiography of the stream channel. Industrial waters
may, of course, be superimposed upon any stream. Such an imposition
would be expected to be permitted on a best-use basis, and recovery to a
higher quality condition with downstream flow would be anticipated.
The classification model for Pennsylvania also included two further
classes. These were clean tidal waters and industrial tidal waters. Obvi-
ously these two categories were limited to the Delaware River. They are
not further developed here.
Suggested Criteria
In the report proposing this model stream classification system certain
criteria were presented as representing minimal desiderata. These cri-
teria included:
1. Wastes discharged should not be allowed to lower the dissolved
oxygen content of the receiving water below 5 ppm in trout streams, 4
ppm in intermediate streams, 3 ppm in warm-water streams, or so deplete
the oxygen content of industrial waters that septic conditions could re-
sult.
2. Toxic materials should not be discharged at concentrations in
excess of one tenth their median tolerance limit as derived by a 96-hr
bioassay.
3. No stream stretch, regardless of class, should be subjected to the
introduction of noticeable floating solids.
WURTZ ON WATER USE FOR AQUATIC LIFE 889

4. No stream stretch, regardless of class, should be subjected to sus-


pended solids loads that would increase turbidity more than 50 per cent
or that would exceed the competence of the stream for transporting the
load.
5. No stream stretch, regardless of class, should be subjected to bed-
load solids that would abrade the stream channel or otherwise alter the
bottom configurations of the channel to the extent that the biological
structure of the stream would be altered.
6. No nonindustrial water should be subjected to oil discharges in an
amount that will produce slick, and no industrial water should be sub-
jected to oil discharges in an amount greater than 100 gal per square mile
of water surface, which is the quantity that first shows traces of color
bands in the slick.
7. Heated discharges in nonindustrial waters should not cause temper-
ature increases above the natural, warm-weather temperature maximum
for the stream type concerned, and heated discharges in industrial waters
should not cause temperatures that will induce septicity or destroy the
biological self-purification capacity of the stream.
The classification system proposed for Pennsylvania was not designed
to permit brutalization of those waters designated as industrial, nor was
it intended to discourage improvement of receiving water quality through
refinements in waste treatment. The proposal simply recognized a de
facto situation. Formal recognition of such existing conditions will help
define our water problems and bring us much closer to equitable solutions
where conflicting interests are found.

References
[7] D. D. Dunbar and J. G. F. Henry, "Pollution Control Measures for Stormwaters
and Combined Sewer Overflows," Journal, Water Pollution Control Federation,
Vol 38, 1966, pp. 9-26.
[2] L. V. Heilbrunn, An Outline of General Physiology, W. B. Saunders Co., Phila-
delphia, 1943.
[3] Peter Doudoroff and Max Katz, "Critical Review of Literature on the Toxicity
of Industrial Wastes and Their Components to Fish. II. The Metals, as Salts,"
Sewage and Industrial Wastes, Vol 27, 1953, pp. 802-839.
[4] C. B. Wurtz, "Zinc,Effects on Fresh-Water Mollusks," The Nautilus, Vol 76,
1962, pp. 53-61.
[5] H. N. Mullican, R. M. Sinclair, and B. G. Isom, "Aquatic Biota of the
Nolichucky River in the State of Tennessee," Tennessee Stream Pollution Con-
trol Board, Tennessee Department of Public Health, Nashville, 1960.
[6] A. M. Beeton, "Environmental Changes in Lake Erie," Transactions, Am.
Fisheries Soc., Vol 90, 1960, pp. 153-159.
STP416-EB/Apr. 1967

Max Katz1 and C. E. Woelke*

Water Quality Requirements of


Estuarine Organisms

REFERENCE: Max Katz and C. E. Woelke, "Water Quality Require-


ments of Estuarine Organisms," Water Quality Criteria, ASTM STP 416,
Am. Soc. Testing Mats., 1967, p. 90.
ABSTRACT: The disposal of industrial and domestic wastes into the
estuarine and marine environment is feasible and practical. Properly
designed disposal systems can distribute wastes so that the other beneficial
uses of the aquatic environment are not harmed. A prerequisite to the
planning of the disposal system is a knowledge of the life history and eco-
logical requirement of the important aquatic species which utilize these
environments. The estuarine environment is an extremely variable one,
and the organisms that utilize this habitat have specialized life histories
that have enabled them to adapt to these conditions. Some fish, the
salmonids, migrate throught the estuary at certain limited times of the
year, and water quality must not impede free migration. Shrimp reproduce
in the open ocean and feed in the estuaries. Some fish species use the
estuaries only for a spawning and nursery area for their young, and only
utilize the estuary at certain times of the year. Oysters remain in the
estuary throughout the year; the shelled adults can close their shells and
survive a toxicity episode, but their free-swimming larvae are very sensi-
tive, and a population can be destroyed readily by toxic effluents. A good
deal of material on the ecological requirements of the important marine
species is available and should be applied by those responsible for waste
discharges in the estuarine and marine environment in order to protect
the important biological communities.
KEY WORDS: water, water pollution, waste disposal, estuaries, aquatic
organisms, fishes

The accelerating population increase in the United States and the


rest of the world has forced people concerned with waste disposal to look
toward the estuaries and open oceans as sites for the disposal of the ever-
increasing loads of domestic and industrial wastes. In many areas the
streams, lakes, and rivers are now handling wastes to their full capacity,
and in too many cases these waters are already well beyond their ca-
pacity to handle the loads that are imposed upon them, let alone sustain
1
College of Fisheries, University of Washington, Seattle, Wash.
2
Washington State Department of Fisheries, Olympia, Wash.
90

Copyright© 1967 by ASTM International www.astm.org


KATZ AND WOELKE ON REQUIREMENTS OF ESTUARINE ORGANISMS 991

desirable fish populations or have any other recreational use. To accom-


modate the industrial expansion necessary to take care of our demands
for goods, it is necessary either to develop more efficient or entirely new
waste treatment methods which, in many cases, will prove to be ex-
tremely expensive; or move to areas in which large amounts of wastes can
be discharged into a system which could dilute and remove the waste
load without the necessity of elaborate treatment facilities. For some
time it was believed that discharging wastes into an estuary or into the
open ocean would prove to be the ideal inexpensive solution to all waste
disposal problems. All that would be necessary is a pipeline that would
be covered by water at mean low tide, and the outgoing tide would ob-
ligingly take the waste out beyond the three-mile limit. By the time the
waste would reach the coast of China, Europe, or South America, it
would be so dispersed and diluted that no one could detect that a waste
had ever been discharged. This method of waste disposal would have
many advantages in addition to low cost. One would not have to worry
about the fish, the state pollution control agencies, or the Federal Water
Pollution Control Administration, and treatment cost would be minimal.
Disposal of wastes in a proper manner into the estuaries and into the
oceans is in fact an ideal method. Proper planning can develop a situation
where on the outgoing tide the wastes can be flushed out to the open sea
to take advantage of the almost infinite dilution available.
Unfortunately, as many disposal engineers have discovered to their
chagrin, for every outgoing tide there is an incoming tide, and even when
the outfall is located close to the mouth of the estuary, a substantial
amount of the discharge is often carried by the incoming tide back to
the plant site, and sometimes the wastes will end up above the mouth of
the outfall before being carried out of the estuary by another change of
the tide. This problem, the flushing of estuaries, is best left to the physical
oceanographers who are busy developing formulas to describe the hydro-
dynamics of estuarine flushing.
The disposal of wastes in the estuarine environment presents some
other problems of a biological nature. There are several important and
fundamental differences between the estuarine environment and the
freshwater systems wherein most of the biological studies regarding the
effects of wastes upon aquatic life have been made. In the freshwater
environment, the biological resource that most often concerns the waste
disposal man is the fish population. This concern is sometimes simplified
because in many cases these waterways have carried heavy waste loads
for so many years that no one can remember what species of fish were
prevalent, and the state fisheries agencies and the sportsmen's groups
no longer regard them as fish producing areas. The species of interest in
freshwater are usually game fish. There are not many areas in the United
States, with the exception of the Great Lakes, that still have significant
922 WATER QUALITY CRITERIA

commercial fisheries. Most conservation agencies have no idea of the


economic value of the game species that could utilize a particular area
if water quality was improved to the level in which fish life could be
sustained and could support a sports fishery. It should be mentioned
that economists have become interested in the very important recreational
and tourist industry that has been stimulated by our prosperity and
shorter working hours. The presence of good fish populations often
forms the basis of an extremely important recreational industry, the
value of which cannot be disregarded.
One of the areas that has stimulated the interest of some keen econo-
mists has been the development of techniques to evaluate the economics
of sport fisheries. It is almost unbelievable what an individual sportsman
will spend in terms of time and money to kill a fish, bird, or animal.
What this interest means from one practical standpoint is that one can-
not dismiss any sports fishery resource as a "few lousy fish" as some-
times has been the accepted practice in certain circles in the past. Game
fish are very expensive fish, and fisheries agencies can now show in the
terms of dollars that game fish are important economic resources in
their own right—even if relatively few pounds are captured.
The biological resources of the typical estuary are far greater and far
more important from any standpoint that you may choose, than those
of a typical freshwater stream unless it is, perhaps, a Pacific salmon
stream. As we mentioned before, the principal biological resource of
our freshwaters are fish, but in a typical estuary one is concerned with
species of fish and shellfish of interest to both the sportsman and the
commercial fisherman. Many estuaries maintain populations of crabs,
oysters, and shrimp, as well as fish. These are important commercial
species with substantial capital investments and large numbers of men
dedicated to their harvest, processing, and marketing. There is no need
to resort to sophisticated economic theory to evaluate the contribution to
our economy of these aquatic resources. The U.S. Bureau of Commercial
Fisheries and many state conservation agencies have long maintained
records regarding the production of fish and shellfish. The industries
that harvest these resources are not primitive—they are efficient and
well organized. A good deal of scientific research has been and is being
carried on in university, state, and federally supported laboratories, on
various aspects of the biology of these species. To cite a few examples—
the Bureau of Commercial Fisheries has maintained a highly respected
and productive shellfish laboratory at Milford, Conn., that has been in
existence since 1930. The Public Health Service has shellfish sanitation
laboratories on the West Coast, the Gulf, and the East Coast. The Water
Pollution Control Administration of the U.S. Department of the In-
terior plans to develop a marine water quality laboratory at Narra-
gansett, R. I., in which the water quality requirements of these inverte-
KATZ AND WOELKE ON REQUIREMENTS OF ESTUARINE ORGANISMS 993

brates will be studied. Other marine water quality laboratories are in


the planning stages.
We know a good deal about the life history and ecological require-
ments of the commercial species of interest in our estuaries and in the
open ocean. Unfortunately, for those with waste disposal problems, what
we do know frequently conflicts with the interests of one who wants to
dispose of an effluent at the lowest possible cost. It must be understood
that these organisms, whether they are fish, crustaceans, or molluscs,
have developed by the same slow and orderly processes of evolution that
has controlled human development, in order to adapt to and compete
successfully in the estuarine habitat. These organisms must have certain
rather well-defined areas in which they can breed and carry on their
activities. They cannot simply move out to the open ocean for the con-
venience of a waste disposal man to carry out their life activities. They
cannot physically adapt to the open ocean any more than we can. There
are certain temperature extremes beyond which they cannot survive and
reproduce. They must have particular types of food at different stages
of their life history. There are certain limits of salinity within which
they can survive or grow. In general, the aquatic forms of the estuaries
are quite specialized and fastidious in their requirements—far more
fastidious in many respects than humans—and have only a few mecha-
nisms that will help them adapt to changes in their environment. If the
certain ecological requirements for oysters, clams, shrimps, or striped
bass are not maintained, the population cannot survive. If the habitat
of these important forms is changed significantly, that is, significantly
in regard to the requirements of the animals, then the organism perishes.
Let us remember the organism (shellfish, oyster, or fish) does not destroy
a habitat—only man destroys habitats. If we wish to use the estuaries to
receive our wastes, we must treat these wastes so that the habitats of
these valuable organisms are not destroyed. We have the ability to
devise these techniques; the animals, unfortunately, do not.
The estuary is a complex environment, and the organisms which live
in the estuaries have developed complex life processes to adapt to the
vagaries of an environment of which the outstanding characteristic is
continual change. On the West Coast the tide flows out twice a day and
the mudflats which are very productive of biological life are exposed.
The clams and oysters that reside on these flats, and do quite well there,
are exposed to the hot sun in the summer and to below-freezing tem-
peratures in the winter. Twice a day the mudflats are covered with water,
and the oysters and clams, along with the hundreds of other organisms in
the community, feed quite happily. During parts of the year the salinity
of the estuary is quite low as the freshets of the winter and spring seasons
are discharged, and during the warm, dry summers the salinity may
approach that of the open ocean. Even during the daily tidal cycle the
944 WATER QUALITY CRITERIA

salinity may change rapidly as the saline tidal prism moves up the bay or
estuary under the freshwater layer. The organisms of the estuary can
adapt to wide changes in temperature and salinity. Certainly some of
these organisms tolerate or perhaps even require certain wide fluctua-
tions in temperature and salinity for optimum growth, but they cannot
tolerate anything and everything. They cannot tolerate the sudden loss
of their food supply which has been killed off by a toxic waste any more
than we could survive if we had to subsist for weeks on a diet of water
and ah*.
Not only do the physical properties in an estuary change markedly,
but the animal species of interest hi an estuary change constantly. The
normal populations of animals change with the seasons as the mobile
species move into and out of the estuary in response to their needs. Let
us examine briefly some of the important species that we find in the
estuaries and see how their life histories complicate the problems of
those responsible for the maintenance of water quality and the biological
resources of an estuary.
In the Pacific Northwest, the most important fish that uses the estu-
aries are the various species of Pacific salmon. The salmon are transients
and migrate through the estuaries as adults on their spawning migration,
or as young fish on their way to the ocean feeding grounds. The various
species of our Pacific salmon lay their eggs in the headwaters of the
streams that flow into the Pacific Ocean. These fish range from the
rivers south of San Francisco Bay through Washington, Oregon, British
Columbia, and Alaska. These important and beautiful species of fish
have had and are having their trials and tribulations. In many cases,
some of the most important races of these fish have been reduced to
pitiful remnants by dams, loss of watersheds, and pollution in the rivers
or estuaries. Important runs of salmon still go through San Francisco
Bay into the Sacramento River which, if our informant is correct, now
sustains the most important salmon runs in the continental United States
(the splendid runs of spring chinook which used to go up the San Jao-
quin, however, were recently eliminated by an engineering masterpiece).
Salmon still migrate through the harbors of Portland, Ore., Grays
Harbor, Wash., and the various harbors in the Puget Sound region of
Washington. These harbors are the repositories of considerable amounts
of domestic and industrial wastes—some treated and others untreated.
The dissolved oxygen during the periods in which the adult fish gather
in these estuaries frequently dips close to the level at which mortalities
can be expected to occur. Fish kills due to toxic industrial wastes are
fortunately becoming less common in these areas.
The adult salmon start moving upstream in September. They move
into the upper stretches of the river where they deposit their eggs. After
hatching in the early spring months and spending varying amounts of
KATZ AND WOELKE ON REQUIREMENTS OF ESTUARINE ORGANISMS 995

time in the fresh water (depending on the requirements of the various


species), the young fish migrate downstream in the late spring into and
through the estuaries with their surrounding industrial complexes and
move out into the ocean where they spend from two to five years feed-
ing. Usually the spring downstream migration does not present serious
problems because the rivers are high with melting snow water. Adequate
dilution, cool water temperatures, and plenty of dissolved oxygen are
characteristic of the rivers at these times, and a minimum of trouble is
anticipated or observed. In certain areas, however, the young fish mill
about for long periods in the estuaries where they are vulnerable to
industrial wastes. In the areas with which we are most familiar, pulp
mill effluent is the most -common industrial waste of concern. At times,
kills of these young fish have been observed. Many other kills are not
observed because dead fish sink to the bottom, and the fish are often
carried away by tidal currents.
After spending two to five years in the marine environment the salmon
return to the mouths of the rivers in which they were born. Again, de-
pending upon the species, they return from July until December. The
most important species, at least in the continental United States, is the
chinook salmon which, in the Seattle area, returns to the estuaries late
in July. They mill about in the harbor and the estuary of the Duwamish
river which flows through the industrial section of Seattle, until about
the first part of October. These fish are prized sport fish, and thousands
of anglers converge on them. During the late summer months, in com-
mon with most of the rivers of the United States, the flow of the
Duwamish is at its lowest, water temperatures are high, the dissolved oxy-
gen content of the water is low, and thousands of not entirely disin-
terested salmon lovers are out in the estuary fishing. While everyone else
is enjoying a pleasant summer, the regulatory agencies and the waste
treatment people sit with their fingers crossed, hoping that it does not get
too warm, that no slug of toxicant reduces the efficiency of the sewage
treatment plant, or that somebody does not turn the wrong valve allow-
ing a batch of plating wastes to escape to the river. Everyone sits on the
edge of his chair until the first freshet of the autumn allows the salmon
to ascend out of the critical areas.
Another very important species complex that is utterly dependent
upon the water quality in the estuaries are the shrimp of the Gulf of
Mexico. These are the shrimp of our shrimp cocktails and our breaded
shrimp. The adult shrimp scatters its eggs offshore, and the young shrimp,
after hatching, move into the nursery areas which are the sheltered la-
goons and estuaries that lie behind the barrier islands of the Gulf Coast
states. In these protected waters, which are high in nutrients, the young
shrimp grow rapidly, and by late summer many migrate back out to the
deeper waters of the Gulf. The young of some species are found in the
966 WATER QUALITY CRITERIA

lagoons at all times of the year. These lagoons are very productive, but
the young shrimp are restricted to an area subject to changes in water
quality. An area of particular concern is Galveston Bay, one of the most
important shrimp nursery areas on the entire Gulf Coast, yet one which
is subject to the wastes emanating from the giant petrochemical and
industrial complex adjacent to the Houston ship channel. Major kills
of fish and immature shrimp have resulted from the masses of toxic and
oxygen depleting waters which are flushed out of the channel by the
torrential rains which are characteristic of this area. From certain stand-
points (those of the polluting industries) shrimp kills do not pose serious
problems. The organisms at this stage of then* development are quite
small, and, even if an entire population is eliminated, the kills are diffi-
cult to detect unless one is carefully monitoring an area. In the past, a
resource like a shrimp fishery could be totally destroyed in an estuary,
and no one would be aware of the fact until the shrimp fisherman fail
to bring in catches. In recent years, however, the Bureau of Commercial
Fisheries, in the course of their studies of the biology of the shrimp, are
keeping detailed records of the abundance of young shrimp in some
important estuaries and are concerned with the effects of water quality
upon these species.
The estuaries are the nursery grounds for many species of marine
fish and invertebrates. Water quality changes in an estuary which do not
hold large populations of resident adult fish, Crustacea, and molluscs,
can destroy large and valuable coastal populations which must utilize
these areas for spawning grounds or nursery areas. Severe alteration of
water quality during a very short period of the year can destroy impor-
tant fish populations. Species of fish that are particularly susceptible to
pollution in estuaries are herring, striped bass, and species of flatfish.
Yet there is frequently no readily observable evidence of deleterious
conditions; the fish are simply no longer to be found.
Oysters, which we dearly love, are particularly sensitive to changes
in water quality in estuaries. In contrast to the salmon mentioned previ-
iously, and the shrimp which must live in the estuaries during part of
their life history, oysters spend their entire lives in the estuary. There
they are spawned, develop, live, feed, and reproduce, and they are
extremely sensitive to alterations of water quality. If the water quality
is altered sufficiently, the adult oyster will protect itself by closing its
shells. It can survive in this manner for long periods of time, but it can-
not feed. The inability of the oyster to feed at this time is purely aca-
demic, for the microscopic organisms upon which it feeds are often
destroyed. Frequent alterations in water quality or constant low levels
of some pollutants results in starved oysters of poor quality. These
oysters cannot be sold, and their reproduction is frequently inhibited.
We may have an estuary in which water quality is good most of the
KATZ AND WOELKE ON REQUIREMENTS OF ESTUARINE ORGANISMS 997

year, the oysters grow fat, and they produce eggs and sperm. Yet de-
terioration of water quality during the short period in which the oyster
larvae are in their free-swimming stage, before they set and develop
their protective shell, can destroy an entire generation. If oyster beds
are situated in an estuary where, for example, industries adhere to prac-
tices which reduce water quality at a particular time of the year, an en-
tire oyster population could be destroyed, and yet, not a dead oyster will
be visible. There are simply no new young oysters. The biologist responsi-
ble for the oysters in the area often cannot determine the cause of the
reproductive failures. Kills of this type are illustrated in part by a series
of fish kills that we once investigated. In this situation there was only
one industry in the watershed, and the kill always occurred below this
industry. The major complication was that the kill always occurred when
the plant was closed down; obviously, the plant could not be involved.
It finally was discovered that the cleanup operations being carried on in
the plant during the shutdown period were responsible. The cleanup
crews were flushing accumulated wastes into the plant outfalls and were
using, in addition, highly toxic substances as cleaning materials. The
plant management in all sincerity swore that their operations were not
responsible—it turned out to be the fault of the cleanup crew.
A summer shutdown which would involve a thorough plant scrub-
down and the removal of a year's accumulation of wastes could result
in acute toxic conditions for a brief period in an estuary. The adult
oysters might close their shells and weather the toxicity episode without
any trouble. The free-swimming unprotected oyster larvae, on the other
hand, could be totally eliminated.
In the Pacific Northwest, the long-standing controversy between the
forest-based industries and the oyster growers has stimulated a large
amount of research on oysters and their reactions to toxic wastes. For
many years the collection of information necessary to resolve the dispute
was hampered by the ability of the adult oysters to close their shells
as a response to an unsatisfactory environment. Although interesting
and useful information was gradually obtained, the test procedures in-
volved were time-consuming and costly, and the data obtained were so
complicated by seasonal and other variables that the analysis could only
be handled by sophisticated statistical techniques—to the dismay of the
biologists and the attorneys.
One of us, Charles Woelke, has developed a technique in which by
regulating water temperatures in the laboratory he can induce oysters
to mature and spawn whenever needed. Adult oysters are spawned and
the eggs fertilized. Toxicity bioassays are conducted with the developing
embryos, and good data regarding the toxicity of a particular waste are
obtained within 48 hr. By the use of a float plane, water samples are
collected in the areas under study and brought into the laboratory where
988 WATER QUALITY CRITERIA

freshly fertilized embryos are ready for use in testing the water sample.
This procedure has worked beautifully. Definite answers have been
made available to the regulatory agencies and to the industries concerned,
and the basis for appropriate action has been obtained.
This has been a lengthy introduction to the main topic of the discus-
sion, the establishment of water quality standards for the marine life
of the estuaries. We regret that we are not prepared to do so except in
general terms. Without long and careful study, we would hesitate to
assign values to be maintained for a body of water whose salinity varies
within a very short period from almost fresh water to a salinity approach-
ing that of the ocean, for organisms which are adapted to life on moist
ground for part of the day, and for the other part of the day are sub-
merged in water. We hesitate, without additional careful preparation,
to recommend water standards which would allow the optimum growth,
optimum reproduction, and the optimum commercial and recreational
values that should be obtained from the harvest of the various species of
fish, Crustacea, and molluscs. We would hesitate to suggest standards
for estuaries in our northern states in which water temperatures seldom
get over 50 F in the summer and expect them to apply to the almost
tropical water conditions that are obtained in Florida and the Gulf
Coast states. We can assure you, however, if realistic water quality
standards are established—and we mean by realistic standards, stand-
ards that are designed to protect the biological resources of our
estuaries and oceans—these standards will be high, and they will not be
easy to meet without careful waste disposal procedures. We are not in-
terested in seeing standards established that will allow a few token indi-
viduals of a population perpetuate themselves from year to year. We are
interested in oyster beds from which oyster men can make a living, and
we are interested in oysters that you will enjoy eating. We do not want
to see beds of oysters in such poor condition that an oysterman cannot
afford to harvest them because they cannot be marketed.
We have had enough experience in this field to know that these or-
ganisms require water of the highest possible quality to maintain them-
selves. As careful research is completed on the physiological require-
ments of our estuarine biological resources, it will be proved that there
can be no compromise in quality. Some studies available may indicate
that good survival of an organism will be observed at 50 per cent of
the saturation value of dissolved oxygen in sea water. But it will be shown
for the organism to grow, to fatten, to reproduce, and to swim rapidly
enough to escape its enemies the water must be very close to 100 per
cent dissolved oxygen saturation. We are sure that these are bold state-
ments for a fish biologist to voice to an audience whose success is based
upon efficient and profitable production and who, for the most part,
must regard waste treatment as an expensive nonessential.
KATZ AND WOELKE ON REQUIREMENTS OF ESTUARINE ORGANISMS 99
99

You have only two alternatives—relocate your waste discharges


where they will be rapidly dispersed in the deep ocean waters—or treat
your wastes to a degree where they will not affect the water quality.
Before concluding, we would like to say that it has been a considera-
ble source of annoyance to us while attending some of the recent con-
ferences on water quality standards, to hear some of our colleagues
apologizing for the lack of information regarding the state of knowledge
on the water quality requirements of marine forms. We wish to differ
with them. There is a tremendous amount of information available on
the physiological requirements of many of the organisms of interest to
us. Over one hundred years of research have been conducted on the
physiology of aquatic organisms, and much useful data have accumulated
in the literature. It is true that this information is widely scattered in the
many scientific journals that are being published all over the world.
Much of this information is not in English. Much of this research has
been performed by academically oriented biologists who never heard of
water pollution at the time they did their work. Much of this information
is not in a form that an engineer, chemist, or even a water quality biolo-
gist can plug easily into his favorite equation. The information is there,
it is on paper; but we have to collect it, and we have to transpose it into
forms that we are accustomed to using.
It is true that we do not have the answers to many of the problems
that face us. We cannot give you the information regarding the toxicity
to crabs or oysters of an exotic secret rocket fuel, whose formula you
cannot divulge, and which is still in the pilot-plant stage of development.
But we can find out the answers if given sufficient tune and adequate
funds. There are many biologists that have the training, the ability, and
the dedication to the resource that they are studying to do the necessary
research. We must admit, however, that aquatic biologists are not magi-
cians and may not always come up with the complete answer on schedule.
After all, our fellow biologists, the medical scientists, are still struggling
with the common cold—despite the armies of workers, the rivers of
funds, and a 5000-year period of research and development.
The biologists will not produce a miracle on demand, but we can get
most of the answers required that will allow industry to fill the needs of
our population and still preserve the natural resources that make life
worth living. We stoutly maintain that we must and can have both in-
dustry and fish!
STP416-EB/Apr. 1967

F. E. Clarke1

What Do We Really Know About Stream


Quality Criteria and Standards?*

REFERENCE: F. E. Clarke, "What Do We Really Know About Stream


Quality Criteria and Standards?," Water Quality Criteria, ASTM STP
416, Am. Soc. Testing Mats, 1967, p. 100.
ABSTRACT: Generally speaking, quality standards for streams are aimed
at insuring maximum practicable utility of water resources for all in-
tended users. Thus, the basic criterion in setting stream quality standards
has been to insure absence of those quality factors which are dangerous
or otherwise objectionable to the users. Lack of accurate information on
effects of many pollutants hinders standardization, as does inability to
perform accurate analysis for certain pertinent water components. Even if
one could analyze accurately and be sure of the quality he desires in a
river, the complex interactions of flow, water composition, seasonal
variations and aquatic processes would make it quite difficult to set a
single quantitative standard for any quality parameter. In some cases,
efforts to reduce pollution are opposed by almost immovable forces of
nature and thus are impracticable, if not almost impossible. The proper
approach to setting stream quality standards under these circumstances
is to accept arbitrary limits or ranges for the most pertinent solutes, sus-
pended materials, and aesthetic factors, while performing the data collec-
tion and research necessary to place standardization on a firmer basis.
KEY WORDS: criteria, standards, municipal water, industrial water, de-
tergents, algae, phosphates, oxygen, biological oxygen demand, pesticides,
sediments, lakes, water, water quality, water pollution

Criterion and standard are words that have been used somewhat in-
terchangeably, but they have rather different meanings. The fact that we
are uncertain about criteria makes it difficult to set practicable standards
for stream quality. One can find a variety of definitions for the two
words. The Webster definitions that appear best suited to the use in this
case are:
Criterion—"A standard of reference—a basis for discrimination."
Standard—"That value established by an authority as a rule for meas-
uring quality or quantity."
1
F. E. Clarke, associate chief hydrologist, U.S. Geological Survey, Washington,
D. C.
* Approved for publication by director, U.S. Geological Survey.
100

Copyright© 1967 by ASTM International www.astm.org


CLARKE ON STREAM QUALITY CRITERIA AND STANDARDS 101

In differentiating between the two terms, it would be better to avoid


the word standard in defining criterion, and to say a basis of reference,
because criteria are far more abstract and elusive than standards. The
latter can be established by administrative decision, stated in rather
quantitative terms and monitored by relatively objective measurements.
The purpose of setting stream quality standards has been stated many
times and in a variety of ways. In general, standards have been aimed
at insuring maximum practicable use of water resources with minimum
probable damage, inconvenience, and cost to intended users. This does
not necessarily mean that all users will derive equal benefit from stand-
ards, except in an indirect way, nor should it. Kneese2 pointed out in a
recent symposium on environmental quality the common misconception
that municipal and industrial operations are highly sensitive to the quali-
ties of waters entering their intakes. This, of course, is not the case. A
well-designed municipal water treatment plant is capable of handling
almost any contaminant, discounting serious unpredicted spills of pol-
lutants and trace quantities of toxins beyond the capability of its treat-
ment capacity. Industries generally are equipped with special flexible
in-plant treatments, so that normal fluctuations in quality of the raw
water have no serious impact on overall cost of operation or quality of
process. For example, a steam electric plant must use essentially pure
water in feeding its high-pressure steam generators, and it is prepared
to make this feedwater from almost any quality of raw water available.
The cost of treatment, either municipal or industrial, will vary, to be
sure, depending on raw water composition; but actual treatment cost is
such a relatively small part of overall expense of operation that raw
water quality usually is not a major consideration.
It also is clearly evident that health authorities have done a remarkable
job in eliminating water-borne infections, regardless of the raw water
source. Thus, human health also is a relatively minor consideration in
setting stream quality standards, except for the possibility of unknown
long-term effects of synthetic organics, trace quantities of unusual in-v
organics, and radioactive components.
The most pressing need for proper stream quality standards is made
evident by the following words taken from the President's creed on
individual rights and duties in conserving our natural resources:
"... The right to surroundings reasonably free from man-made ugli-
ness—and the duty not to blight.
"... The right of easy access to places of beauty and tranquility where
every family can find recreation and refreshment—and the duty to pre-
serve such places clean and unspoiled."
These words emphasize the important and relatively new concept
2
Allen Kneese, director of water and environmental studies, Resources for the
Future, Inc., Washington, D. C.
102 WATER QUALITY CRITERIA

that stream quality standards must encompass: the quality of the entire
stream environment—not simply the water it contains. It is a credit to
our economic prosperity and relatively high standard of living that we
recognize the need for better criteria and standards primarily for insuring
aesthetic values of our waterways, including healthy biological balance
and recreational potentials. Unfortunately, limiting our prime objective
to this area of concern does not simplify the task of setting stream quality
standards. Instead, it becomes a more difficult task, involving evaluation
of rather elusive benefits, and area in which we have little experience or
competence.

FIG. 1—Biological equilibrium.

Quality Criteria—The First Unknown


It would be a relatively simple matter (but still a difficult one) to
establish stream quality standards according to the definition and for
the purpose stated above, if one could first establish sound criteria on
which to base the standards. Nature is far from cooperative in such a
task. Consider, for example, the problem of selecting criteria and stand-
ards to minimize obnoxious blooms of blue-green algae. There is much
evidence that sewage nutrients, particularly detergent phosphate, are
related to nuisance blooms of this kind. Therefore, one might use as his
criterion of corrective action reduction of phosphate nutrient below the
critical limit at which algae bloom, and set his water quality standard
accordingly. But even decreasing phosphate to the limit of analytical
detection may not do the job in every aquatic environment subject to
the influences of both nature and man. As pointed out by Stumm and
Morgan,3 what really determines algae blooms is biological stoichiometry,
or balance, between autotrophic and heterotrophic organisms in the par-
ticular environment. The balance is illustrated in Fig. 1, a modification
of the Stumm-Morgan diagram. This shows a one-way flow of energy
3
Werner Stumm and J. J. Morgan, "Stream Pollution by Algal Nutrients,"
Transactions, Twelfth Annual Conference on Sanitary Engineering, University of
Kansas, Lawrence, Kans., 1962, pp. 16-26.
CLARKE ON STREAM QUALITY CRITERIA AND STANDARDS 103

and continuous recycling of nutrients in a semiclosed system involving


minimum input and loss of nutrients. As long as there is equilibrium,
troublesome autotrophic blooms of blue-green algae should not occur.
However, the heterotrophs must have organic carbon from sources other
than carbon dioxide to build the 106-C:16-H:1-P structure, which is
characteristic of both types of organisms. A municipal sewage treatment
which discharges effluent relatively rich in nitrogen and phosphorus and
relatively lean in organic carbon will favor the autotrophs and inhibit the
heterotrophs which feed upon them. Blooms of autotrophs therefore
are likely to occur in most domestic sewage effluents, because they origi-
nally are relatively low in organic carbon and are subjected to primary
and secondary treatments which are more effective in removing organic
than in removing phosphorus and nitrogen compounds. The degree of
treatment, whether 90 or 95 per cent removal, will make little difference
under these conditions. Only if one resorts to chemical removal of phos-
phate, as by precipitation with aluminum, calcium, or iron compounds,
can he insure proper biological balance in a stream receiving such waste-
water. Raw sewage actually may cause less bloom than treated sewage,
but, of course, it would increase biological oxygen demand (BOD) and
cause oxygen depletion problems. The proper criterion in this case is
maintenance of biological balance. A standard based on BOD or dis-
solved oxygen would not do the job. Failure to recognize this point could
easily result in continued construction of bigger and better conventional
treatment plants, with little or no improvement in the problem. Lack of
understanding of other complex quality control processes obviously will
be a stumbling block to our efforts to develop practicable standards.
While nobody would say it is a simple matter to establish criteria under
the conditions cited above, some criteria are even more elusive. Fear of
unknown long-term effects of synthetic organics already has been men-
tioned. Berkner4 recently mentioned in the Engineering Science Centen-
nial Convocation at Lafayette College the possibility that pesticide resi-
dues eventually may affect certain biota of the continental shelf, which
biota are believed to be primarily responsible for generation of at-
mospheric oxygen. Because the atmosphere is a rather poorly buffered
system, it is possible that such an effect might significantly reduce oxygen
content of the earth's atmosphere in a relatively short period of time and
thereby markedly change our environment. Who is to say how much
pesticide it would take to do this, if at all, and whether or not the critical
concentrations are detectable by present sampling and analytical meth-
ods? In such areas of uncertainty, it is easy to conclude that one should
avoid every trace component when its effects are unknown or when it
does not appear to be a product of the natural environment. This cri-

*L. V. Berkner, board chairman of the Graduate Research Center of the


Southwest.
104 WATER QUALITY CRITERIA

terion simplifies the task of setting stream quality standards, whether or


not they can be attained, but is not foolproof. Our complex environment
is capable of generating almost any kind of compound, and many of the
things we call "unnatural" simply have not yet been discovered in nature.
Man evolved in this environment of many "unknowns" and body proc-
esses may depend on certain trace components without our knowing it.
It already is well known, for example, that some arsenic is essential to
human development and that a zero limit of this element in one's diet
is detrimental. The standard for copper in drinking water has been
changed from time to time as we learned more about its effects. These
observations, plus the knowledge that too much of almost anything is
toxic, leads to the logical conclusion that standards should be aimed at
maintaining that concentration range which represents neither too much
nor too little.
Our ability to reach agreement on criteria for aesthetic factors and
appearance properties, such as color, taste, odor, and turbidity, is little
better than our ability to handle the problems already discussed. A per-
son reared in limestone country is likely to find soft water tasteless and
unpleasant, while others consume sulfide-bearing well waters of the
eastern seaboard with little concern for odor that would be considered
intolerable elsewhere. One who has visited the Ganga (Ganges) River
in India is keenly aware of the marked bearing background and culture
can have on criteria of river quality.
Appearance may give a completely erroneous impression of water
quality regardless of the background of the observer. A stream in Penn-
sylvania which is crystal clear, well aerated, and populated with green
plants may be contaminated with mineral acid from mine wastes to the
extent that it is uninhabitable for fish and unsuitable for many other
purposes. By contrast, a muddy stream in the Southwest may be a rather
healthy stream.

Analytical Uncertainties
Inability to establish suitable criteria and standards is due in some
cases primarily to analytical difficulties. Again, this can be illustrated by
reference to the phosphate compounds. It is possible, by empirical meth-
ods and pilot studies, to confirm phosphorus as a major contributing
factor in algal blooms. It is a much more difficult task, however, to de-
termine the critical concentration of phosphate in the aquatic system
where blooms are occurring. As a matter of fact, it is almost impossible
to determine the concentration of soluble inorganic phosphate in such
a system. A typical aquatic environment probably includes simple and
complex phosphates in soluble organic, soluble inorganic, and insoluble
(particulate) form, but even the most experienced analysts have been
unable to separate the forms satisfactorily. Attempts to remove particu-
CLARKE ON STREAM QUALITY CRITERIA AND STANDARDS 105

late matter by filtration often show that the finer the filter medium, the
lower the apparent insoluble phosphate content, until finally there is
virtually none at all. It is very difficult to determine how much phos-
phorus originally was inside the organism (for example, algae), and how
much was in the surrounding water, and equally difficult to determine
the particular form of phosphorus in either case.
In the laboratory one determines orthophosphates and complex phos-
phates by standard analytical procedures. He can determine the molecu-
lar structure of particulate compounds with the conventional tools of the
trade. These methods cannot be used in the microenvironment of bloom-
ing algae, nor can a sample be collected and preserved with complete
certainty that even the total phosphate will remain the same as it was in
the stream. There is considerable evidence that most soluble phosphate
determinations made during the last few decades in natural waters con-
taining aquatic organisms are in error by factors of ten or more. Evi-
dence compiled by biological assay methods suggests that as little as
1.0 jug per liter of excess soluble phosphate may be all that is required
for a well balanced aquatic system. It would be a difficult task to insure
sewage effluent with this low concentration by conventional treatments.
These facts are not very encouraging either to the person who is trying
to set water quality standards or to the analyst who must determine
if the standards are being met.

Hydrologic Complications
Normal hydrologic variations are as troublesome as aquatic processes
and analytical uncertainties. It is reasonably safe to say that surface
waters never are in chemical equilibrium with their environments and that
they seldom attain steady states. In setting water quality standards, one
must be aware not only that any period of record contains marked
variations in flow and quality, but also that this period may not be at all
representative of long-term behavior. Even if the period of record is
representative, there are such complications as simultaneous increase in
sediment load and decrease in dissolved solid load under certain condi-
tions of precipitation and flow. During periods of low flow, generally
in late summer and early autumn, high concentrations of troublesome
pollutants coincide with high oxidation rates to scavenge dissolved oxy-
gen and make the water lethal to fish and generally unpalatable and un-
sightly to humans. In the winter, high flow rates and relatively sluggish
organic degradation prevent dissolved oxygen troubles but result in
relatively short travel times which deliver undesirable pollutants to mu-
nicipal intakes before natural recovery processes have been able to
dispose of them. At both seasonal extremes, most streams have a dis-
tinctly spotty quality pattern representing areas of contamination and
recovery. In short, a river is a very different system in summer than in
106 WATER QUALITY CRITERIA

winter, and not a very uniform one in either season. One therefore must
decide whether he will settle for standards which vary with the seasons,
with or without quality zoning of the stream, or whether he will rest his case
on the criterion of the best quality attainable under worst conditions.
A statistician undoubtedly would argue for setting standards only in
terms of probability and for synthesizing hydraulic data to avoid serious
errors from short-term or spotty sampling. This involves a risk factor—
a concession that standards will not be met 100 per cent of the time.

Possible Approach to Better Knowledge


Our ability to establish better criteria and standards never will im-
prove unless we devise some plan for determining the interactions be-
tween the variety of water quality controls, both natural and man-made.
While the catalogs of data already collected by a variety of investigators
include analytical errors of the type discussed and exclude certain data
and river reaches that should have been covered, it should be possible
to make better use of them. To say this another way, we should be able
to make a better estimate of the standards which are compatible with the
behavior of water resources under the stresses of nature and man's
activities. One approach would be to explore the usefulness of relatively
simple correlations between water quality data and streamflow records,
like that illustrated in Fig. 2 from a recent paper by Durum and Blakey.5
Figure 3, compiled from U.S. Geological Survey data, indicates a rela-
tively uniform relationship between dissolved solid loads and streamflow
in a number of rivers ranging from the east coast to the west coast of the
United States. Generalizing these observations, it can be said that dis-
solved solid load increases uniformly with increasing river flow (or
runoff) up to about 10 in. of precipitation in the areas considered.
Granted, these are average figures for single sampling points, and
greater variation would be expected for individual sampling points along
any of the streams.
Figure 4 is a diagram of a hypothetical situation showing what might
be expected in an ideal drainage basin where total load, both of sus-
pended and dissolved solids, increases uniformly downstream. Points A,
B, and C represent successive downstream sampling stations in this
stream system, which is subject only to normal seasonal variations in
flow and quality. The dotted lines represent percentages of time during
which indicated conditions might be expected to occur at each sampling
point. The parallel positive slopes of the solid lines for the three stations
represent increases in total load of solid matter (not the concentration)
as discharge (flow) increases for each of the stations. There also is an
5
W. H. Durum and J. F. Blakey, 'Time-weighted Analysis as an Indicator of
Natural Stream Quality," presented at Association of State and Interstate Water
Pollution Control Administrators, Dayton, Ohio, 1965.
CLARKE ON STREAM QUALITY CRITERIA AND STANDARDS 107

increase from station-to-station in the downstream direction. Such a pat-


tern would be expected in areas receiving normally distributed rainfall
and involving relatively uniform and stable terrains, with relatively uni-
form erosion of sediment and leaching of solubles.
A similar diagram (Fig. 5) with negatively sloping solid lines would
mean decreasing load with increasing discharge at each station and

FIG. 2—Streamflow conditions for selected years, Maumee River at Waterville,


Ohio.

increasing total load from station-to-station downstream. Such a pattern


could result from dilution of relatively concentrated ground water with
less concentrated precipitation and runoff. The character of the terrain
again would control the relationship, and uniform slopes would be ex-
pected only in the case of uniform environment relatively free from
human influence.
A variety of slopes could result from nonuniform river basin condi-
tions, including serious pollution. For example, flushout of mine acid
108 WATER QUALITY CRITERIA

WATER DISCHARGE -
FIG. 4—Ideal drainage basin, load increasing with discharge downstream.

during periods of high water table might explain a steep slope at a


sampling station. The slope at a succeeding station might be different
because of combining a relatively neutral low solid water from one tribu-
tary with polluted water from another affecting the station. Load-dis-
charge slopes below such a confluence would show the effect of relatively
uniform effluent resulting from the mixing.
CLARKE ON STREAM QUALITY CRITERIA AND STANDARDS 109

The person who tries diagrams of this kind on actual stream data,
station-by-station, is likely to obtain less ideal plots than are shown here.
The conditions that apply to average dissolved solids loads may not be
applicable to individual contaminants and specific samples. Nevertheless,
quality-flow relationships may be sufficiently uniform to allow interpola-
tion between sampling points. At least, they will be helpful in showing
where new sampling points are needed.
The same general approach might be used in conjunction with mathe-
matical models and computer techniques for integrating flow charac-
teristics, seasonal variations, and water quality data, at least for rela-

FIG. 5—Ideal drainage basin, load decreasing with flow at sampling stations.

tively conservative water components, such as sediment and persistent


dissolved solids. It is not beyond the realm of reason to assume that even
the more complex patterns of nonconservatives, such as dissolved oxy-
gen, temperature, and degradable organics, eventually will yield to com-
puter techniques as our knowledge of stream processes and analytical
methods improves. Only when we reach this point will we be able to
establish standards which are reasonably reliable for any point in space
and time.
Exactly Where Do We Stand?
If one answers frankly the title question, "What do we really know
about stream quality criteria and standards?" he must admit that there
is need for much more knowledge than we now have if the most practi-
1 10 WATER QUALITY CRITERIA

cable standards are to be set. For example, if one were to plot dissolved
oxygen data on a map for almost any river basin, serious gaps would be
apparent. Needless to say, it is difficult to establish sound quantitative
standards on such bases. There is no immediate answer to this problem
but to make administrative decisions and arbitrary standards until we
have more information. The Ohio River Valley Sanitation Commission
already has done this, using the records at its disposal. The resulting
standards, in general, represent the minimum quality expected at any
place at a particular time. In other words, the criterion used in estab-
lishing these preliminary standards might be considered the philosophy
that we must accept the best possible conditions while working to improve
our water resources. This is a logical basis for setting realistic standards.
Such standards insure maintenance of the qualities attainable under
present conditions and provide insurance against further degradation.
Perhaps more stringent standards will be possible in the future, despite
predicted economic development and increasing use of water resources.

Can We Live with the Standards We Set?


The uninitiated may think that almost any standard of stream quality
can be maintained once we have agreed upon criteria and set our sights
accordingly. This is true only if quality standards are realistic ones. For
example, geomorphic processes will dictate in large measure the total
sediment loads of streams regardless of soil conservation practices and
reduction of erosion in urban developments. We must accept this fact
and learn to live with suspended matter unless we are prepared either to
inhibit powerful natural processes or to purify streams on a massive
scale. A meandering river, whether controlled by the principle of least
work or some other basic law of science, will continue to explore its
flood plain, cutting one bank while building another, and spilling the
excess sediment downstream.
Certain rivers in primitive parts of Alaska are quite red (with ferric
hydroxide), and on-site analyses show their dissolved oxygen contents to
be zero, this component having been consumed by converting leached
ferrous ion to the ferric form. Surrounding peat deposits are the source
of the iron, so this is what one might call "natural" pollution on a gi-
gantic scale. It would be difficult to counteract a natural phenomena of
this kind, just as it would be difficult to prevent the early "dying" of
many lakes located in the same area.
One should not expect the impossible in setting water quality stand-
ards, and certainly should not expect the raw water from every river to
be usable in untreated form for every purpose. It is logical to expect
that we can attain the basic objective, namely, to eliminate eventually
many unsightly and troublesome pollutants within the realm of practica-
ble control. This philosophy is apparent in the guidelines recommended
CLARKE ON STREAM QUALITY CRITERIA AND STANDARDS 1 11

by the Secretary of the Interior to the States for implementing the Water
Quality Act of 1965. No one will debate the fact that many of our rivers
can be improved considerably by methods and at costs within our reach.
Excess heat, troublesome nutrients, and unnecessary silt are examples
of pollutants than can be controlled to a significant extent. In some
cases, even "natural" pollutants can be minimized, as by diversion of
salt seeps and other measures which do not oppose powerful forces of
nature. In all of these efforts it is important to remember that stream
quality standards involve the entire environment of the stream, not
simply the water. Water quality, as such, can be handled reasonably
well after withdrawal. This is not true of many aesthetic quality factors.

Summary
It can be concluded that:
We know far too little about certain natural water quality controls to
be sure that our criteria are appropriate or that we can set practicable
standards.
Our inability to analyze properly for a variety of components makes it
difficult not only to select criteria and set stream quality standards, but
also to be sure that the standards are being met.
We must learn new ways of integrating water quality data, hydraulic
characteristics, and other factors before the most practicable stream
quality standards can be set, and be content with arbitrary standards in
the meantime. Far more basic data and better interpretive processes will
be needed to facilitate this task.
Our goal, then, should be to overcome recognized deficiencies as
rapidly as possible and to upgrade stream quality standards continuously
on the basis of the information available. This is not a task for any
single agency concerned with water resources. It will require the effort
of everyone who has an interest in the subject. Prudence dictates that
we not wait until all necessary information is available before setting
interim standards to assist in the task of overcoming obvious pollution
problems.
STP416-EB/Apr. 1967

C. E. Woelke1

Measurement of Water Quality with the


Pacific Oyster Embryo Bioassay

REFERENCE: C. E. Woelke, "Measurement of Water Quality with the


Pacific Oyster Embryo Bioassay," Water Quality Criteria, ASTM STP
416, Am. Soc. Testing Mats, 1967, p. 112.
ABSTRACT: Meaningful water quality standards or criteria must be ex-
pressed in terms of consumer needs. If these needs are based on chemical
parameters, the criteria or standards should be based on chemical meas-
urements. If consumer needs are based on biological factors, standards
or criteria should be based on biological measurements. Methods have
been developed in which Pacific oyster embryo bioassays measure water
quality in terms of response by a type of animal (consumer) found in
many of our bays and estuaries. The method has been successfully em-
ployed in the laboratory to measure relative toxicity of pulp and paper
wastes. Both polluted and unpolluted water from bays and estuaries are
routinely bioassayed with oyster embryos. Based on these bioassays,
areas of acceptable and unacceptable water quality have been delineated
relative to oysters. It is recommended that this method and similar bioas-
say techniques be adopted as part of the measurements employed in de-
fining water quality standards and criteria.
KEY WORDS: water, water pollution, water quality, estuaries, bioassay,
oysters, toxicity, molluscs, industrial wastes, pulp mills, paper mills, re-
fineries

The previous papers have indicated some of the complexities in-


volved in evaluating water quality in an environment subjected to wide
short-term physical and chemical changes. While not specifically spelling
it out, they have implied that the plethora of physical and chemical
measurements in routine use may still not detect unsatisfactory biological
conditions.
Biological assays are actually the most logical, and frequently the only
available, method for defining water quality. Doudoroff et al [I]2 de-
scribed a standardized bioassay procedure for fish, and Woelke [2] pro-
posed the use of bivalve larva for bioassays of waters in which oyster
1
Washington State Department of Fisheries, Olympia, Wash.
2
The italic numbers in brackets refer to the list of references appended to this
paper.
112

Copyright© 1967 by ASTM International www.astm.org


WOELKE ON PACIFIC OYSTER EMBRYO BIOASSAY 113

and clam populations are present. Bioassays with bivalve larva of simu-
lated wastes prepared in the laboratory have been reported by a number
of investigators [3-10]. Dimick and Breese [11] propose the bay mussel
as a standard bivalve for marine water bioassays because these molluscs
are found in nearly all estuarine areas in the world.
In this paper I shall give a brief description of a rapid, inexpensive,
and dependable bioassay technique I developed while working for the
Shellfish Research Unit of the Washington Department of Fisheries.
For the remainder of my presentation I shall deal with my subject
in four steps. First, state the assumptions and justifications; second,
describe the method itself; third, present data obtained in a series of
toxicity bioassays conducted on 29 samples of pulp mill and oil refinery
wastes; and fourth, describe how this technique has been extended to
actual estuarine pollution problems.
In developing this technique I assumed a general acceptance of the
need for and merit of biological assays in the field of water quality re-
search. I further assumed an acceptance of the concept that development
of a bioassay procedure with a commercially valuable marine organism,
which could be applied at any tune of the year with the same ease and
reproducibility currently attributed to the biological oxygen demand and
coliform mean probable number tests, would be a useful tool for evaluat-
ing water quality. It is my contention that the first 48 hr in the develop-
ment of fertilized eggs of the Pacific oyster, Crassostrea gigas, provides
a biological system whose response can be utilized to satisfy these as-
sumptions. During this 48-hr period the fertilized eggs normally develop
into free-swimming, fully shelled veliger larvae. My final assumption is
that failure to develop to fully shelled (normal) larvae in 48 hr will
break the life cycle of the Pacific oyster. I consider failure of the eggs
to develop, or the proportion (per cent) of larvae developing in an ab-
normal manner to constitute a measure of the biological response to a
particular stimulus.

Method
The basic steps I have followed in carrying out bioassays with fertil-
ized Pacific oyster eggs are relatively simple and straight-forward. Adult
oysters (spawners) are thermally conditioned at 20 C in flowing sea-
water until they can be readily spawned. This usually requires about
four to six weeks. To insure the availability of spawners during all
months of the year, several groups of oysters at various stages of sexual
maturity are kept on hand at all times. Several hours before a bioassay
is to be conducted, 10 to 20 mature oysters are placed into Pyrex dishes
filled with filtered ultraviolet-light-treated water. These dishes are placed
in a water bath and the temperature is raised to 28 to 30 C. About 30
min before the time the spawning is desired, a sperm suspension from
TABLE 1—Levels of response of Pacific oyster embryonic development to waste samples bioassayed.
No Effect0 20% Abnormal 50% Abnorma 100 % Abnormal Controls
Waste Numb %
FBI Dilution FBI Dilution FBI Dilution FBI Dilution Abnormal

39 5 1:10,000 15.0 1:2100 19.5 1:1650 33 1:1000 0.40


40 0 <1:100,000 0.0 1:525 0.0 1:390 0 1:100 0.40
41 0 1:100 3.2 1:61 7.2 1:46 15 1:20 0.41
42 1 1:2000 5.2 1:145 8.2 1:112 89 1:20 0.41
43 6 1:10,000 13.0 1:4540 21.0 1:3000 68 1:1000 0.41
44 18 1:20,000 18.5 1:19,000 23.5 1:16,500 43 1:10,000 2.20
45 4 1:1000 13.5 1:72 19.5 1:50 50 1:20 2.20
46 5 1:100 14.0 1:22 18.0 1:17 30 1:10 2.20
47 <18 < 1 : 1000 29.0 1:620 41.0 1:450 97 1:200 2.20
48 4 1:2000 8.4 1:920 13.0 1:600 36 1:200 2.20
62 2 1:20 >2.0 >1:10 >2.0 >1:10 >2 >l:10 0.05
63 3 1:100,000 17.5 1:11,500 28.0 1:7200 104 1:2000 0.05
64 1 1:1000 9.2 1:170 11.5 1:145 75 1:20 0.05
65 >0 >1:10 >0.0 >1:10 >0.0 >1:10 >o >l:10 0.05
66 2 1:100,000 14.0 1:3400 24.0 1:1950 301 1:200 0.05
67 1 1:2000 <1.0 1:690 <1.0 1:500 1 1:200 0.98
68 0 1:10,000 5.7 1:68 7.8 1:48 17 1:20 0.98
69 0 1:200 1.5 1:65 2.1 1:47 5 1:20 0.98
70 0 1:100,000 1.4 1:7800 2.4 1:5400 10 1:2000 0.98
71 0 1:100,000 1.3 1:7900 2.1 1:5450 7 1:2000 0.98
72 0 1:100,000 9.1 1:550 11.5 1:475 22 1:200 0.98
73 1:1000 1:700 1:430 1:100 1.79
74 0 1:20 >o!o >1:10 >0.0 >1:10 >0 >1:10 3.42
75 1 1:100 1.0 1:18 1.0 1:15 1 1:10 3.42
76 0 1:1000 1.2 1:59 20.0 1:39 6 1:10 0.58
77 0 1:2000 1.2 1:560 2.0 1:370 4 1:100 0.58
78 4 1:200 9.4 1:62 12.8 1:46 30 1:20 0.58
79 15 1:100 24.0 1:59 32.0 1:43 65 1:20 0.58
80 0 1:100 >4.0 >1:10 >4.0 >1:10 >4 >1:10 2.65
0
Same as controls or nearly so.
WOELKE ON PACIFIC OYSTER EMBRYO BIOASSAY 115

a sexually mature, sacrificed male oyster is added to the water. The


combination of increased temperature and sperm will induce one or more
of the female oysters to spawn. Eggs from a single female are selected
for use in the bioassay, and the number of eggs per unit volume are
determined by sampling the sperm-egg suspension. Glass or plastic
TABLE 2—Wastes ranked and grouped from most toxic to least toxic, based on
estimated dilution level which produced 20 per cent abnormal larvae.
Toxicity Waste Description of Waste Dilution Ratio FBI
Group Number

I... . 44 alcohol plant 1:19,000 18.5


63 diffuser line 1:11,500 17.5
II 71 surge tank diffuser line 1:7900 1.3
70 red stock washer 1:7800 1.4
III . 43 main sewer 1:4540 13.0
66 diffuser 1:3400 14.0
39 composite sewer 1:2100 15.0
IV . 48 bleach plant and lignin products 1:920 8.4
73 kraft mill 1:700
67 bleach plant sewer (chlorinator) 1:690 <1.0
47 pulp wash 1:620 29.0
77 bleach wash 1:560 1.2
72 main mill sewer 1:550 9.1
40 bleach plant waste 1:525 0.0
V . 64 main sewer 1:170 9.2
42 screen room 1:145 5.2
VI 45 board mill 1:72 13.5
68 composite 1:68 5.7
69 caustic extractor 1:65 1.5
78 ground wood screenings 1:62 9.4
41 barker waste 1:61 3.2
76 main mill paper machine 1:59 1.2
79 ground wood (refiner) 1:59 24.0
VII . 46 barker 1:22 14.0
75 oil refinery 1:18 1.0
62 white water >1:10 >2.0
65 paper machine >1:10 120
74 oil refinery >1:10 12
80 paper mill >1.10 >4.0

beakers containing the water to be bioassayed are each inoculated with


a sufficient amount of the egg suspension to give 20,000 to 30,000
fertilized eggs per liter. Approximately 10 per cent of the cultures in a
given bioassay are controls. The culture containers are placed in a 20 C
water bath for 48 hr. At the end of this time the cultures are poured
through a 37 p. sieve to collect the oyster larvae. Samples containing
about 150 to 250 larvae taken from each culture are preserved and later
examined under a microscope. The number of normal and abnormal
1 16 WATER QUALITY CRITERIA

TABLE 3—Gallons of dilution water needed per day to reduce the waste streams
toxicities to 0 and 20 per cent response levels.
Bioassay Dilution Present Waste Million G lo s Per Da
? ^. ° f y
Sample No. Volume, million Required for
allons er da
0 response 20% response « P y 0 response 20% response

44 >20 000 19 000 2.74 >54 800 52 060


63 100 000 11 500 13.15 1 315 000 151 225
71 100 000 7 900
70 100 000 7 800 4.50 450 000 35 100
43 10 000 4 540 29.09 290 900 130 680
66 100 000 3 400 5.67 567 000 192 780
39 10 000 2 100 4.60 46 000 9 660
48 2 000 920 13.20 26 400 12 144
73 1 000 700
67 .... 2 000 690 8.00 16 000 5 520
47 1 000 620 17.30 17 300 10 726
77 2 000 560 2.92 5 840 1 635
72 100 000 550 7.80 780 000 42 900
40 > 100 000 525 0.70 >70 000 367
64 1 000 170 40.32 40 320 6 854
42 2 000 145 9.00 18 000 1 305
45. 1 000 72 1.13 1 130 81
68 10 000 68
69 200 65 5.00 1 000 325
78 . . . . . 200 62 0.91 182 56
41 100 61 0.70 70 43
76 1 000 59 5.90 5 900 348
79 100 59 1.34 134 79
46 100 22 1.43 143 31
75 100 18
62 20 <10 3.49 69.8 <35
65 <10 <10 8.40 <84 <84
74 20 <10
80 100 <10

Totals 187.29 3 706 273 654 038

larvae are counted in each sample. The response which forms the basis
of the bioassay is the per cent abnormal larvae. The effect of any varia-
ble tested is described in terms of the per cent oyster larvae which de-
velop abnormally.

Results
In a study comparing the toxicity of 29 composite wastes serial dilu-
tions of one part waste sample to 10, 20, 100, 200, 1000, 2000, 10,000,
20,000, 100,000, and 200,000 parts of fresh seawater were prepared.
Each dilution was divided between three one-liter beakers for replication
of the bioassay. At least nine beakers were filled with pure seawater to
be used as controls in each of the nine separate bioassays made during
the study. The mean per cent abnormal larvae from each dilution of each
waste bioassayed was plotted on probability paper, and levels of no
WOELKE ON PACIFIC OYSTER EMBRYO BIOASSAY 1 17

FIG. 1—Flow diagram for oyster larva bioassay.

effect, 20, 50, and 100 per cent abnormals (response), were determined.
The Pearl-Benson Index (FBI), a chemical measure commonly used for
measuring pulp and paper wastes, was determined for each dilution of
each waste bioassayed. The results of this study are shown in Table 1.
In Table 2 these wastes are arranged in order from most to least toxic,
based on the amount of dilution needed to reduce their toxicity to the
20 per cent abnormal level. It should be noted that when ranked in this
manner, the FBI values for the waste dilutions designated do not follow
any particular pattern. This indicates that the FBI does not measure the
relative toxicity of the different wastes. To give a more practical de-
scription of these results, the gallons of dilution water needed to reduce
FIG. 2—Port Angeles, Wash., stations for oyster larva bioassay samples.
WOELKE ON PACIFIC OYSTER EMBRYO BIOASSAY 119

the waste streams bioassayed to the 0 and 20 per cent response levels
are summarized in Table 3.
In a further refinement of this bioassay procedure, water samples are
collected by airplane from various estuarine environments and are flown
to the laboratory where they are bioassayed within a few hours of the
time collected. In this type of bioassay, a control water sample is carried
from the laboratory on the airplane and is subject to the same handling
stresses as the samples to be bioassayed. The water is collected in one-
gallon polyethylene containers. As many as 50 water samples have been
collected for a single bioassay. These samples are divided between four
one-liter beakers. Three samples are inoculated with freshly spawned
oyster embryos, and the fourth is used for chemical analysis. At the end
of the 48-hr bioassay, further chemical measures are often made on the
waters in which the embryos actually developed. The flow diagram hi
Fig. 1 outlines the procedure followed in this type of bioassay. Average
per cent abnormal larvae and average FBI values from seven consecutive
monthly bioassays in one area of Washington state are summarized in
Fig. 2. High larva response levels in the Port Angeles area which decrease
with increasing distance from Port Angeles are readily apparent. At
present, water quality is being monitored annually with oyster larvae at
more than 130 stations, which include 95 per cent of the oyster growing
areas of Washington state and over 80 per cent of the estuarine areas
of the state.
Conclusions
I feel that the results achieved with the oyster embryo bioassay justi-
fies considering it for general use in:
1. Evaluating existing estuarine water quality.
2. Monitoring estuarine water quality.
3. Determining toxicity of new potentially toxic materials.
4. Measuring relative toxicity of wastes or potential wastes and esti-
mating their probable effect on molluscan populations.
5. Aiding in determining the degree and type of treatment a particular
waste might require.
6. Evaluating the effectiveness of waste treatment facilities discharg-
ing into estuarine waters.
7. Establishing estuarine water quality standards.
Among the advantages of this method are its speed (relative to many
other types of bioassays), its simplicity, its low cost, the fact that it is
based on a commercially valuable species, the availability of test organ-
isms on a 12-month basis, and the clear-cut response of the oyster em-
bryos. While not mentioned previously, biological problems such as age,
size, sex, and prior exposure of the animals to stress, which tend to
confound the results of bioassays with many other organisms, are not
120 WATER QUALITY CRITERIA

present with this method, since all embryos have the same parents and
are exactly the same age and size.
I would be less than honest if I did not caution that while the results
of oyster embryo bioassays can be used with a fair degree of confidence
when defining water quality for molluscs, the extension of the results to
crabs, shrimp, swimming fish, diatoms, or the plankton forms on which
these animals feed may be a hazardous and ill-advised procedure. At
present it appears that where water quality does not interfere with em-
bryonic development of oyster larvae, other animal forms will thrive.
In spite of this it is my firm belief that similar bioassay procedures must
be developed with species representing the other major groups of organ-
isms found in the estuarine environment, particularly those of social or
economic importance to man.
I feel strongly that water quality criteria must include bioassay meas-
urements of water of the type outlined here, in addition to chemical or
physical measurements.

References
[1] P. Doudoroff et al, "Bioassay Methods for the Evaluation of Acute Toxicity
of Industrial Wastes to Fish," Sewage and Industrial Wastes, Vol 23, No. 11,
1951, pp. 1380-1397.
[2] C. E. Woelke, "Bioassay The Bivalve Larvae Tool," Proceedings, 10th Pacific
Northwest Symposium Water Pollution Research, U.S. Department HEWPHS,
Portland, Ore., 1961.
[3] H. C. Davis, "Effects of Some Pesticides on Eggs and Larvae of Oysters
(Crassostrea virginica) and Clams (Venus mercenaria)," U.S. Fish and Wildlife
Service, Commercial Fisheries Review, Vol 23, No. 12, 1961.
[4} H. C. Davis and P. Chanley, "The Effects of Some Dissolved Substances on
Bivalve Larvae," Proceedings, Nat. Shellfisheries Assn., Vol 46, pp. 59-74.
[5] H. Hidu, "Effects of Synthetic Surfactants on the Larvae of Clams (M.
mercenaria) and Oysters (C. virginica)," Journal, Water Pollution Control
Federation, Vol 37, No. 2, pp. 262-270.
[6] V. L. Loosanoff, "Some Effects of Pesticides on Marine Arthropods and
Molluscs," U.S. Public Health Service, Transactions, 2nd seminar on Biological
Problems in Water Pollution, April 20-24, 1959.
[7] K. Okubo and T. Okubo, "Study on the Bioassay Method for the Evaluation
of Water Pollution. II. Use of Fertilized Eggs of Sea Urchins and Bivalves,"
Bulletin, Tokai Regional Fisheries Research Laboratory No. 32, 1962.
[8] C. E. Woelke, 'The Effects of Spent Sulphite Waste Liquor on the Develop-
ment of Eggs and Larvae of Three Marine Molluscs and Three of Their Food
Organisms," Washington State Department Fisheries, Research Bulletin No. 6,
1960.
[9] C. E. Woelke, "Effects of Sulfite Waste Liquor on the Normal Development
of Pacific Oysters (Crassostrea gigas) Larvae," Washington State Department
Fisheries, Research Bulletin No. 6, 1960.
[10] C. E. Woelke, "Bioassays of Pulp Mill Wastes with Oysters," U.S. Public
Health Service, Transactions, 3rd seminar on Biological Problems in Water
Pollution, 1962.
[11] R. E. Dimick and W. P. Breese, "Bay Mussel Embryo Bioassay," Proceedings,
12th Pacific Northwest Industrial Waste Conference, University of Washing-
ton, College of Engineering, 1965, pp. 165-175.

You might also like