Ecological Risk A^yssm^n^ Final Report Endpoints for Ecological Toxicity ssb TU Submitted To: Exposure Assessment Group Office of Research and Development U.S. Environmental Protection Agency Submitted By: Technical Resources, Inc. 3202 Monroe St., Rockville, MD 20852 Work performed under contract #68-02-4199 ------- Ecological Risk Assessment ENDPOINTS FOR ECOLOGICAL TOXICITY BY Abe Mittelman, Project Manager Joanne Settel Kathleen Plourd Rolland S. Fulton III Gene Sun Shakuntala Chaube Patrick Sheehan* TECHNICAL RESOURCES, INC. 3202 MONROE STREET, SUITE 200 ROCKVILLE, MARYLAND 20852 Contract No. 68-024199 September 30, 1988 Project Officer John Segna Exposure Assessment Group Office of Research and Development U.S. Environmental Protection Agency Washington, D.C. 20460 *Aqua Terra Technologies ------- TABLE OF CONTENTS PAGE Introduction iv Chapter 1 Organism-Level Ecotoxicological Endpoints 1 1.0 Introduction 1 1.1 Acute Mortality 2 1.2 Biochemical Alterations 7 1.3 Osmoregulatory Activity 9 1.4 Respiratory Activity 12 1.5 Behavioral Alterations 12 1.5.1 Avoidance reactions 13 1.5.2 Predator-prey interactions 14 1.5.3 Reproductive behavior 15 1.5.4 Locomotor activity 16 1.5.5 Feeding activity 17 1.6 Reproductive Toxicity 21 1.7 Musculoskeletal Effects 24 1.8 Growth and Development 24 1.8.1 Developmental toxicity 27 1.8.2 Growth 31 1.9 Genotoxicity 32 1.10 Carcinogenicity 34 1.11 Modifying Factors 35 1.11.1 Microbial toxicity 37 1.11.2 Multifactorial interactions 41 1.12 Conclusion 43 Chapter 2 Population-Level Endpoints 43 2.0 Introduction 43 2.1 Population Dynamics 44 2.1.1 Birth rate 44 2.1.2 Death rate 45 2.1.3 Population growth models 45 2.1.4 Other population endpoints 51 ------- 2.2 Life-history Strategy 52 2.2.1 Parental care and fecundity 54 2.2.2 Parental care and growth/development 54 2.2.3 Growth and age at maturity 55 2.2.4 Reproduction and life expectancy 55 2.2.5 Longevity and body size 56 2.2.6 Compensatory mechanisms of life-history strategy 56 2.2.7 Life histories and response to contaminants 57 2.3 Conclusion 58 2.4 Methods for Selecting Appropriate Populations to Monitor 59 2.4.1 Total population 60 2.4.2 Species dominance 62 2.4.3 Indicator species 63 2.4.4 Keystone species 65 2.4.5 Representative and important-species approach 69 2.5 Conclusion 70 2.6 Species Interactions 71 2.6.1 Predator-prey interactions 72 2.6.2 Interspecies competition 74 2.7 Conclusion 76 Chapter 3 Ecosystem-Level Endpoints 77 3.0 Introduction 77 3.1 Ecosystem Function 78 3.2 Energy Flow 79 3.2.1 Primary productivity 79 3.2.2 Respiration 83 3.2.3 Photosynthesis/respiration ratio 84 3.2.4 Methods of measurement of primary production 85 3.3 Biogeochemical Cycling 94 3.3.1 Nutrient analysis 95 3.3.2 Chlorophyll content 97 ii ------- 97 100 101 103 108 109 111 112 114 116 116 118 121 125 127 128 129 130 135 136 138 139 139 140 141 142 142 143 145 146 148 149 3.3.3 Leaching 3.3.4 Determination of a nutrient budget 3.3.5 Nutrient cycling: the nitrogen cycle 3.3.6 Methods for analysis 3.3.7 Decomposition 3.3.8 Methods of measurement 3.4 Conclusion 3.5 Ecosystem or Community Structure 3.6 Abundance and Biomass 3.7 Species Lists 3.8 Biological Indices (Pollution Indices) 3.9 Species Richness 3.10 Diversity 3.11 Comparative Indices (Similarity Indices) 3.12 Multivariate Analysis 3.13 Trophic Organization 3.14 Spatial Structure 3.15 Guilds 3.16 Conclusion 3.17 Stability 3.17.1 Resilience 3.17.2 Amplitude 3.17.3 Elasticity 3.17.4 Inertia 3.18 Conclusion Summary Introduction Organism-Level Endpoints Population-Level Endpoints Ecosystem-Function Endpoints Ecosystem-Structure Endpoints Conclusions iii ------- INTRODUCTION The performance of an ecological risk assessment requires a delineation of endpoints suitable for measurement. Ecotoxicological endpoints are here defined as physical or biological parameters, characteristic of the ecosystem, which are measurably and predictably affected adversely by contaminants. Such endpoints potentially include individual organism-, population-, and ecosystem-level parameters. While organism-level effects have often been used in isolation to assess environmental impact, these are not generally sufficient to describe ecosystem response. The complex nature of interactions between the abiotic and biotic components of an ecosystem necessitate a more integrative approach to ecological risk assessment. Both population- and ecosystem-level effects need to be considered, along with organism-level effects, in order to adequately describe ecosystem response. This document surveys the range of endpoints that have been or potentially may be used in ecological risk assessment at the levels of the individual organism, the population, and the ecosystem. For each endpoint discussed, strengths and weaknesses as measures of ecological damage are presented, and examples of its use in measuring pollutant effects are presented. The report focuses on the use of endpoints in the detection of existing effects, particularly at the population and ecosystem levels of organization, but many of the endpoints described are also appropriate for pre-release prediction of contaminant effects through their measurement in laboratory single-species bioassays, microcosm, mesocosm, or experimental ecosystem studies. Test methods currently used in hazard assessment are discussed in more detail in TRI (1988c). iv ------- A variety of endpoints have been used to assess the ecological hazards associated with anthropogenic stress on ecosystems. The use of specific response measures in ecological risk assessment depends, in part, upon the type of ecosystem impacted, the type of stress, the proposed hypothesis, and the required level of detection. This is to say that, although a general set of possible response endpoints can provide data on effects at various levels of biological organization, the selection of appropriate endpoints for each study must be situation specific. To evaluate the applicability and utility of response endpoints for ecological risk assessments, the following criteria should be applied: o Is there an appropriate theoretical base for the response endpoint? o Do these endpoints provide ecologically meaningful information? o Is the endpoint easy to measure? o Is the endpoint sensitive to stress, and is the response concentration- dependent? o Can the endpoint be applied to analysis of a variety of situations and different ecosystems? o Is there a history of successful use of the response endpoint in impact assessments? o Does it have socio-economic-legal significance? This document does not provide specific guidelines for selection of ecological endpoints. However, further guidance is presented in another document, "Ecological Endpoint Selection Criteria" (TRI, 1988a). v ------- Response endpoints at the individual organism level can be useful predictors of stress effects at the population level, if the connection between individual performance and population fitness is clear. Chemical injury to individuals resulting in premature death and reduced reproductive success and recruitment may be reflected in lowered abundance and altered distribution of exposed populations. The connection between individual and population responses will depend on the processes regulating population dynamics. In turn, the relevance of population response data for predicting community- and ecosystem-level impacts is related to the importance of the population in structuring the community and in controlling functional processes. These statements point to the importance of gathering ecologically meaningful information and, if possible, response endpoint data at the appropriate biological level to maximize confidence in extrapolations. vi ------- CHAPTER 1 ORGANISM-LEVEL ECOTOXICOLOGICAL ENDPOINTS 1.0 INTRODUCTION Physiological parameters are important biological endpoints in ecological risk assessment. The biological success of a population is based, in part, on the physiological status of its individual members. Physiological endpoints include mortality, biochemical alterations, osmoregulatory effects, respiratory effects, behavioral effects, reproductive effects, musculoskeletal effects, effects on growth and development, genotoxic effects, and carcinogenic effects. A major advantage of using physiological endpoints is that they often respond rapidly to pollutant stress. This early detection may allow corrective action to be applied before irreparable damage has occurred. However, it is often not possible to relate physiological endpoints to the population- or ecosystem-level effects of pollutants, because these levels are strongly influenced by interactions among species and between species and the physical environment (Levin et al., 1984). 1.1 ACUTE MORTALITY Acute mortality, or toxicity, is the level of toxic chemical that produces lethality in a specific proportion of test organisms in a given short period of time (usually 48 or 96 hours). The most frequently used measure is the median lethal dose or concentration that kills 50 percent of the organisms being tested (LDjq or LC50). These tests are simple to perform, relatively inexpensive, and practical to conduct, and an enormous background literature exists on them. Such tests 1 ------- have a bottom line appeal in that they can be presented as a single figure. There are, however, many drawbacks associated with them. Test outcomes depend on the physical and chemical environment (e.g., water hardness, water movement, temperature, pH, photoperiod), the organisms used (species, degree of acclimation, age, sex, genetic stock), synergistic and antagonistic interactions between toxicants, and variations in procedural methods. These drawbacks are shared by other physiological toxicological endpoints. Acute lethality tests are often considered insufficiently sensitive to provide information on sublethal or chronic effects; these occur at lower concentrations and may be of considerable ecological significance. Sublethal impairment may affect an organism's ability to cope with other naturally occurring stresses (such as environmental variability or interactions with other organisms), and thus its ability to survive in the natural environment. Acute toxicity testing represents the quickest and least sophisticated way of obtaining a preliminary evaluation of a hazardous substance (Monk, 1983). 1.2 BIOCHEMICAL ALTERATIONS Biochemical responses to pollutants have received considerable attention because they provide information on the mechanisms of action of toxic pollutants. Biochemical responses to many different chemicals have been studied in fish. In a study by Verma et al. (1981), Saccobranchus fossils were exposed to sublethal concentrations of copper sulfate, the detergent swascofix E45, and chlordane and metasystox for 40 days. These substances caused significant inhibition of the enzyme acetylcholinesterase (AChE). Associated behavioral responses were violent 2 ------- movements followed by loss of equilibrium, possibly due to accumulation of acetylcholine (ACh) at nerve endings, thus disrupting transmission of nerve impulses. Enzymes involved in energy metabolism were also affected by the toxicants. There was general inhibition of succinic dehydrogenase (SDH) and pyruvic dehydrogenase (PDH), and stimulation of lactic dehydrogenase (LDH) activity, indicating depression of aerobic metabolism and development of anaerobic conditions at the tissue level. Birds have also been used to monitor biochemical changes resulting from pollutant exposures. In one study, the pesticide fenitrothion suppressed brain cholinesterase (ChE) activity in several species of song birds (Busby et al., 1981). In another study, Japanese quail exposed to the pollutant parathion exhibited ChE inhibition. It was found that in the quail, inhibition in excess of 20 percent indicated stress while inhibition of more than 50 percent usually caused death (Ludke et al., 1975). ChE inhibition is useful as an ecological endpoint because of its specificity and clear association with pollutant-induced effects. While it can result in disruption of neural action, it is not significantly affected by neural stress, and it is associated with mortality. Mortality from ChE inhibition is preceded by various symptoms, including anorexia, lethargy, antagonistic behavior, muscular incoordination, and convulsions. Effects of sublethal inhibition of ChE can include weight loss, impaired growth and reproduction, and increased susceptibility to predation. A drawback associated with the use of ChE inhibition as an ecotoxicological endpoint is that it is difficult to correlate with chemical exposures in the field. This problem is, however, common to most 3 ------- ecotoxicological endpoints and arises from difficulties in determining field exposures. Brain neurotoxic esterase (NTE) activity has been used as a biochemical measurement of delayed neurotoxicity in birds. In chickens, delayed neurotoxicity is generally manifested as ataxia within 2 weeks of dosage, with eventual paralysis of the legs. An excellent correlation has been established for organophosphate pesticides, which cause delayed neurotoxicity and inhibition of NTE activity both in vivo and in vitro (Lotti and Johnson, 1978). Evidence for the effects of pesticides on NTE is provided by a study in which the organophosphate pesticide EPN (phenylphosphonothioic acid-0-ethyl-0-4- nitrophenyl ester) was fed to mallards (Anas platyrhynchos) for 90 days. The treated birds experienced inhibition of brain NTE activity by 16, 69, 73, and 74 percent at concentrations of 10, 30, 90, and 270 ppm, respectively. Brain NTE inhibition of 65 percent or greater was associated with severe ataxia or paralysis. Treatment-related demyelination and degeneration of axons of the spinal cord were also observed (Hoffman et al., 1984). Other enzymes have also been used as endpoints to monitor pollutant stress. One of these, delta-aminolevulinic acid dehydratase (ALA-D), is an essential enzyme in the biosynthetic pathway of heme synthesis. Its inhibition may result in significant reduction in heme synthesis and in neurological consequences. However, the association of the ALA-D index with overall fitness of exposed populations is not known. Depression of ALA-D activity has been observed in 4 ------- response to Pb exposure in fish (Jackim, 1973) and birds (Hutton, 1980; Hoffman et al., 1981). The mixed function oxygenase (MFO) enzyme system, catalyzed by the heme protein cytochrome P-450, may also be a useful endpoint. It is responsible for biotransformation of xenobiotics in vertebrates and invertebrates and has been used to indicate pollution by aromatic hydrocarbons or other pollutants. For example, elevation of MFO activity was detected in livers of brook trout (Salvelinus fontinalis) that had been fed a single dose of PCB (Aroclor 1254) (Addison et al., 1981). A number of studies have observed induction of hepatic cytochrome P-450 activity by petroleum hydrocarbons (Stegeman, 1980). Though changes in MFO activity appear to provide a useful endpoint for monitoring effects of pollutant stress, results must be interpreted cautiously because many factors can influence MFO activity, including diet, temperature, season, age, sex, and species. The extent of MFO induction by different chemicals varies widely; indeed some cause inhibition of MFO activity. While MFO induction may be used as an indicator of certain pollutants, the relationship between it and environmental damage remains unclear (Stegeman, 1980). The physiological basis of stress response is endocrinological. Under stress, the adenohypophysis is stimulated to produce more ACTH, which in turn results in elevated levels of corticosteroids secreted by the adrenal cortex. Thus, corticosteroid hormone levels can provide monitorable endpoints for pollutant- induced stress. For example, corticosteroid changes were observed in Sockeye salmon (Oncorhynchus nerka) exposed to low levels of Cu in freshwater aquaria. Treated fish showed a rapid, dose-related increase in Cortisol, cortisone, and total 5 ------- corticosteroid levels (corticosteroid response). Fish exposed to 10"^ molar of Cu died between 8 and 24 hours (Donaldson and Dye, 1975). Several other studies have noted changes in corticosteroid levels in response to various pollutants, including heavy metals, pesticides, coal dust, pulp mill effluent, and landfill leachate (Donaldson, 1981). It has been suggested that disruption of the endocrine balance (increased corticosteroids) is the cause of depressed growth in oil-dosed birds (Sheehan 1984a). However, corticosteroid changes are not specific responses to pollutants; levels can change in response to many environmental stresses, including handling, anesthesia, temperature or salinity change, hypoxia, confinement, crowding, or disease. Another difficulty is that organisms frequently acclimate to sublethal stresses, resulting in a transitory corticosteroid response (Donaldson, 1981; Schreck, 1981). The relationship between corticosteroid stress responses and longer-term effects on growth and reproduction is unclear. Lysosomal membrane stability is another biochemical endpoint for pollutant effects. Membrane stability is important in maintaining cellular integrity and in preventing autolytic cell damage by free hydrolases. In one study, pollutant effects on lysosomes were monitored in the hydroid Campanularia flexuosa. An increase in the levels of free glucosaminidase activity in endodermal cells was observed after exposure to threshold concentrations of Cu, Cd, and Hg. This change in glucosaminidase activity was hypothesized to result from decreased stability of the lysosomal membrane (Moore and Stebbing, 1976). Such lysosomal membrane stability has been shown to be correlated with growth indices both in the laboratory and in the field (Sheehan, 1984a). 6 ------- Biochemical responses to toxic pollutants can be quite complex. A wide spectrum of biochemical changes were observed in the eggs of brook trout (S. fontinalis) exposed to PCB (Aroclor 1254) before and after hatching. These changes decreased levels of hydroxyproline, vitamin C, collagen, and phosphorous but increased Ca concentrations in the spine. The decreased growth rate, lowered levels of hydroxyproline, and vitamin concentrations in sac fry exposed to PCB, and later increased mortality in these fry, suggest that PCB exposure induced competition between developmental and detoxification processes for the use of vitamin C (Mauck et al., 1978). In summary, biochemical responses are quite sensitive to short-term and sub- lethal pollutant stresses and are often easily associated with the mechanisms of toxicity. However, extrapolation to longer-term organismal responses, and to population- or ecosystem-level responses, is not generally given with present knowledge. 1.3 OSMOREGULATORY ACTIVITY Osmoregulatory activity may be a good ecological endpoint because it can provide an important measure of physiological stress in aquatic organisms. Its function in aquatic species is prevention of loss of salts from the organism and maintenance of the salt water balance over a range of salinities. Environmental pollutants, by inhibiting or interfering with these adaptive mechanisms, might be expected to reduce the ability of aquatic organisms to tolerate stressful salinity. 7 ------- Exposure of juvenile coho salmon (Oncorhynchus kisutch) to Cu in fresh water reduced their tolerance to increased salinity. Effects of Cu on gill ATPase (inhibition) resulting in impaired survival in sea water occurred within 24 to 72 hours. This decreased gill-ATPase activity was probably one of the factors leading to loss in osmoregulatory ability and death (Lorz and McPherson, 1976). Impairment of osmoregulatory capabilities was evident in a study of the effects of sublethal concentrations of inorganic Hg and PCBs on chloride ion levels in the blood of estuarine shrimp (Palaemonetes pugio). The study showed significant effects on the ability of shrimp to adjust to rapid fluctuations in environmental salinity. Thus a rapid change in salinity during exposure to sublethal concentrations of the PCB could be lethal to the shrimp (Anderson et al., 1974). Studies of osmotic regulation in a variety of aquatic organisms have shown that this is a sensitive measure of the effects of a number of different metals. Jones (1975), for example, showed that Cd significantly lowered the blood osmotic concentration of the isopod Idotea neglecta in seawater, while Cd, Zn, and Hg significantly altered the osmoregulatory ability of another isopod, Jaera albifrons, in water of dilute salinity. Increases in mortality are associated with osmoregulatory changes at low salinities. Reduced osmotic regulation has also been observed in a number of other estuarine species exposed to chlorinated hydrocarbons (PCBs, DDT). Osmoregulatory changes have also been observed in killifish (Fundulus heteroclitus) and eels exposed to sublethal concentrations of DDT and Aroclor 8 ------- 1221. Blood of exposed fish exhibited elevated Na+ and K+ levels and higher osmolarity compared to controls (Anderson et al., 1974). In another study, the ability of sodium-depleted killifish (F. heteroclitus) to take up sodium was completely inhibited after a 24-hour exposure to sublethal levels of inorganic Hg (Renfro et al., 1974). The ability of organisms to acclimate to osmotic stress makes the interpretation of the ecological significance of short-term osmoregulatory studies difficult. The relationship between osmoregulatory changes and effects on mortality is rarely well established. Although impairment of osmoregulatory capabilities by pollutants has been repeatedly demonstrated in laboratory experiments, it is questionable whether this is a suitable endpoint for environmental monitoring studies. Osmotically impaired organisms have never been found in contaminated estuarine environments, perhaps due to the rapid death of such organisms following exposure to an osmotic stress (Bayne et al., 1980). 1.4 RESPIRATORY ACTIVITY Several investigators have used the respiratory activities of fish and aquatic macroinvertebrates as indicators of aquatic organism response to environmental stress. Physiological parameters that have been used to monitor respiration include ventilatory frequency and the cough response. Ventilatory frequency is defined as the buccal and opercular opening and closing frequency. The cough response has been frequently used as a short-term predictive parameter. It is defined as a regularly recurring break in the ventilation rhythm. These 9 ------- parameters may be affected by variables other than pollutants, such as dissolved oxygen, temperature, and predator influence. Thus, the interpretation of the ecological significance of these respiratory responses can be problematic. Respiratory responses provide rapid and inexpensive measures of response to pollutants. However, it is unclear whether they accurately predict long-term, chronic effects of pollutants. Among the most useful long-term chronic tests are those employing full-life-cycle and critical-life-stage exposures for deriving a maximum acceptable toxicant concentration (MATC). While the MATC facilitates projection of "safe" concentrations of chemicals in natural waters, determination of a MATC is expensive and time-consuming compared to many physiological endpoints, although not in comparison to population- or ecosystem-level endpoints. In one study, ventricular frequency was used to determine the diurnal respiratory response of bluegills (Lepomis macrochirus) to surfactants. The results were compared to previously existing full-life-cycle chronic toxicity data on other species of fish. A good correlation was found between the chronic MATC for fathead minnows and concentrations of surfactants that produced statistically significant changes in the diurnal ventilatory frequencies in exposed bluegills (Maki, 1979). These results support the notion that monitoring of ventilation frequency may have predictive utility as a tool for early estimation of long-term chronic effects. In another study, elevated ventilatory rates and reduced food conversion efficiency were observed in pinfish (Lagodon rhomboides) exposed to oil and bleached kraft mill effluent (Stoner and Livingston, 1978). These changes indicate 10 ------- that the pollutant caused an elevated metabolic demand -- a situation that could weaken the organism and leave it susceptible to additional stress. Anderson et al. (1974) concluded that changes in oxygen consumption provided a useful measure of pollutant effects in aquatic organisms. They noted that polychlorinated biphenyls significantly reduced oxygen consumption in killifish (Fundulus similus). In addition, the respiratory rate following exposure to PCBs was only about 20 percent of the rate measured in the same fish prior to exposure. In contrast, grass shrimp (P. pugio) exposed to PCBs showed an increase in oxygen uptake. These authors also reported an age- and species- specific response of crustaceans to petroleum hydrocarbons. Mussels (Mytilus edulis) taken from areas polluted with heavy metals and organic compounds displayed increased oxygen consumption rates. The cough response has been used in several respiratory studies to measure pollutant-induced stress. In one such study conducted by Bull and Mclnerney (1974), exposure of juvenile coho salmon (O. kisutch) to sublethal concentrations of the organophosphate insecticide fenitrothion caused an increase in the cough response. Similarly, an increase in cough frequency in yearling brook trout was observed within 2 to 24 hours after exposure of the fish to Cu. The mean cough frequencies tended to increase with Cu concentration (Drummond et al., 1973). The Maki (1979) study is suggestive of the utility of short-term respiratory responses as predictors of long-term pollutant effects. However, further workv needs to be done before respiratory responses can be used as indices of the effects of pollutants on survival, growth, and reproduction. 11 ------- 1.5 BEHAVIORAL ALTERATIONS Behavioral changes provide potentially useful endpoints for measuring pollutant effects, because they are elicited at very low pollutant concentrations and often affect population survival and success. Changes in behavior affect reproduction, migration, nesting, shelter construction, avoidance activities, and vulnerability of prey. Difficulties with behavioral measures arise, however, because they are not easily quantifiable. Practical assessment of pollution-related behavioral changes depends on qualitative definition of normal behavior patterns. 1.5.1 Avoidance Reactions Perception and avoidance of contamination is paramount for the survival of species exposed to polluted ecosystems. In several studies, the Atlantic salmon (Salmo solar) has been observed to avoid water contaminated with Cu or Zn. In 4 successive years of pollution, downstream returns of salmon were 22, 14, 10, and 15 percent of the upstream migrations, in contrast to the 1 to 3 percent downstream returns during the 6 years before pollution. The estimated threshold for the avoidance reaction was about 0.35 to 0.43 toxic units (1.0 toxic unit being equal to the LD50). A unit of 0.8 blocked all upstream movement (Sprague et al., 1965, cited in Sheehan, 1984a; Saunders and Sprague, 1967). Differential avoidance of insecticides was observed in two mosquito-fish (Gambusia affinis) populations differing greatly in insecticide tolerance (resistant and susceptible populations). The study showed that both populations avoided 12 ------- endrin, toxaphene, and parathion, but only the susceptible population avoided DDT. In addition, the concentrations at which the avoidance response occurred differed markedly in the two populations. With the exception of DDT, the resistant population showed a much greater avoidance of damaging concentrations of the other three pesticides than the susceptible population (Kynard, 1974). In another avoidance study, rainbow trout fry (Salmo gairdneri) were found to avoid a selected group of toxicants such as copper sulfate and dalapon but not others such as diquat (Folmar, 1976). Avoidance reactions have also been observed in a number of different invertebrate species. In one study, the distribution of burrowed marine bivalves (Macoma balthica) was assessed in a tank containing sediment contaminated with various concentrations of metals, including Cu, Pb, Zn, Cr, Hg, and Cd. At the highest pollutant concentrations, a statistically significant avoidance response was observed (McGreer, 1979). The midge larva (Chironomus tentans) was also found to display a linear avoidance reaction to metal-contaminated (Cd, Zn) sediment of Lake Palestine in northern Indiana (Wentsel et al., 1977). Other factors such as texture, organic matter content of the sediment, and death of chironomids did not influence the avoidance behavior. 1.5.2 Predator-Prey Interactions Sublethal concentrations of pesticides have been reported to interfere with prey escape and other antipredator behaviors. In a multiprey system, predators may be expected to consume a higher than normal proportion of the more 13 ------- affected or impaired species. Thus, changes in predator-prey interactions can provide an important indication of pollutant effects on a population. Examples of effects of pollutants on predator-prey interactions are discussed in Chapter 2. 1.5.3 Reproductive Behavior A number of different types of reproductive behavior have been used as ecotoxicological endpoints. Behavioral changes are often mediated through pollutant-induced changes in the endocrine system. These hormonal changes may ultimately induce behavioral anomalies, such as abnormal courtship and nest construction, decreased incubation attentiveness, nest desertion, impaired nest defense, and decreased parental care. Changes in parental behavior resulting from pollutant stress have been observed in birds. Herring gulls (Larus argentatus) of Lake Ontario exposed to a variety of pollutants, including DDE, PCBs, mirex, photomirex, and other pesticides, were found to be inattentive and showed decreased nest defense compared to birds from an uncontaminated site. Nests were left unattended for long periods, thus exposing the eggs to predators and to less than optimal temperature, which resulted in a high incidence of embryonic mortality (Fox et al., 1978). Similarly, laughing gulls (Larus atricilla) exposed to a single sublethal dose of the organophosphate insecticide parathion exhibited reduced parental care activity and increased rate of nest desertion. Hatchability and defense behavior, however, were not affected (King et al., 1984). 14 ------- Exposure of mallard hens (A. plalyrhynchos) to mercury has resulted in abnormal maternal behavior. The exposed hens laidx their eggs outside their nesting boxes and did not defend them against predators (Heinz, 1979). Likewise, hens fed the pesticide Abate prior to initiation of laying until ducklings were 21 days of age, showed abnormal incubation behavior (Franson et al., 1983). Courtship behavior has also been shown to be affected by toxicants. For example, in one study, the mean number of seconds of total courtship activity time displayed by male Ringed Turtle Doves (Streptopedia risoria) was reduced by exposure to DDE in the diet (Haegele and Hudson, 1977). The degree of change was found to be related to the level of DDE exposure. No change in courtship behavior was observed at 10 ppm, while birds on the 50 ppm DDE contaminated diet had reduction in activity time. 1.5.4 Locomotor Activity Locomotor activity is one of a number of behavioral endpoints that has been used to monitor pollutant stress on a population. Exposure to even low concentrations of environmental pollutants are known to elicit adverse effects on locomotor activity. Activities associated with changed behavior, however, vary widely between subjects and over time and are not easily quantifiable. Therefore, to properly measure a change in behavioral response, normal behavior patterns must first be defined so that a quantal or gradual change can be demonstrated. A number of studies document pollutant-induced changes in the locomotor activity of fish. Davey et al. (1972) observed locomotor changes in goldfish 15 ------- (Carassius aurathus) exposed to DDT. In unexposed fish, a significant correlation was established between the magnitudes of consecutive turns in opposite directions, which initially decreased as the time between the turns increased and then ceased abruptly. As a retention process is involved in this process, it is assumed that this behavior is controlled by the central nervous system. It was observed that DDT-exposed fish displayed a significant reduction in the correlations between turns, implying an impairment of the retention process. Returning the fish to clean water for up to 130 to 139 days did not restore normal behavior. Similarly, juvenile coho salmon (O. kisutch) exposed to the organophosphate insecticide fenitrothion displayed changes in behavior patterns, indicating physiological impairment (Bull and Mclnerney, 1974). All locomotor and some comfort behaviors ceased, and many fish, unable to hold position, were swept downstream. In another study, effects on the locomotor activity of yearling brook trout (S. fontinalis) exposed to Cu were observed. While locomotor activity varied widely, all groups were from 4 to 6 times more active than the controls. Increased activity lasted for 6 to 8 hours at all concentrations of Cu and then returned to pretreatment values (Drummond et al., 1973). 1.5.5 Feeding Activity Measures of changes in feeding response are potentially useful endpoints for ecological risk assessment because they provide an early indication of 16 ------- physiological stresses that could result in growth retardation. Unfortunately, the feeding response is difficult to measure, and the baseline data for this activity is sparse. One study in which feeding response proved to be an effective measure of pollutant effects was conducted by Drummond et al. (1973). The investigators monitored the feeding habit of the brook trout (S. fontinalis) following exposure to Cu. Feeding was markedly depressed or ceased after 2 hours at Cu concentrations of 9 ug/1 or higher and did not return to normal within 24 hours. Feeding was also found to be more sensitive than other behavioral parameters as a measure of the effects of sublethal concentrations of the organophosphate insecticide fenitrothion on juvenile coho salmon (O. kisutch). The chemical was found to depress the feeding response of the fish at concentrations of 0.1 ppm (Bull and Mclnerney, 1974). 1.6 REPRODUCTIVE TOXICITY Reproductive success, crucial for species perpetuation, primarily depends on the normal functioning of the neuroendocrine system. This, in turn, is closely associated with normal maturation of the reproductive organs and their functioning during various phases of the reproductive cycle (e.g., gamete formation and its maturation, fertilization, courting and mating behavior, and rearing of the young (see the section on Behavior Alterations). Environmental pollutants that adversely influence these reproductive parameters also influence population dynamics. Birge et al. (1980, cited in Sheehan, 1984a) suggested that 17 ------- 10 percent or greater increase in mortality in the developmental stages (embryonic, larval) would significantly alter the population dynamics in natural populations. Reduced hatch success, one of the most commonly used indices of reproductive success, was observed in a study of pollutant effects on barnacles (Wu and Levings 1980). The investigators noted reduction in egg production and survival in barnacles (Balanus glandula) exposed to bleached kraft pulp mill effluent near a pulp mill outfall. Similarly, a significant reduction in brood size and number of survivors was observed among cladocerans (Moina macrocopa) exposed to sublethal concentrations of Cd, with 50 percent reproductive impairment occurring at a concentration of 0.78 ug/1 of Cd (Hatakeyama and Yasuno, 1981). Fish eggs exposed to cyanide, to organochlorines such as Lindane, or to sublethal concentrations of heavy metals such as Hg, Fe, and Cr showed decreased fertilization rates (Billard, 1978, cited in Sheehan, 1984a). Schofield (1976) reviewed studies of declining freshwater fish populations in acidified waters in Scandinavia and eastern North America. He reported rapid extinction rates of fish populations inhabiting acidified lakes (pH less than 4.5), often resulting from chronic reproductive failure. Reduced hatch success in birds also provides a sensitive endpoint for toxic effects. Blus et al. (1972) found that dieldrin impaired reproduction in brown pelicans (Pelecanus occodentalis). Hatching success was also reduced in nests of 18 ------- the red-breasted mergansers (Mergus serraior) contaminated with toxaphene (Heinz et al., 1983). In a study of three generations of mallard ducks (A. platyrhynchos) fed 0.5 ppm of methylmercury, exposed birds laid fewer hatchable eggs than controls (Heinz, 1979). Similarly, screech owls exposed to endrin in their feed experienced reductions in the number of eggs laid, number of eggs hatched, and number of chicks fledged (Fleming et al., 1982). Reduced hatch success in birds has been correlated with eggshell thickness. Eggshell thinning thus provides a potentially useful endpoint for measuring reproductive toxicity. Cooke (1973) reviewed experimental avian data and concluded that environmental pollution can result in thin, easily cracked eggshells. In addition, there are interspecific differences in response among birds, with falcon eggs being more susceptible to thinning than eggs of gallinaceous birds. Studies of the mechanisms involved in reduction of shell thickness by organochlorines suggest general impairment of calcium metabolism, a reduction in available carbonate in the shell gland lumen, a deleterious effect on the thyroid and adrenal glands, and an alteration in organic matter being incorporated into the developing shell. The extent to which each of these mechanisms is likely to contribute to shell thinning will depend on the species, condition of the bird, and environmental conditions. These effects are illustrated in a study of red-breasted mergansers (M. senator) nesting on islands in northwestern Lake Michigan. The birds experienced a small degree of eggshell thinning, which was attributed to DDE (Heinz et al., 1983; Peakall, 1983). In another study conducted on mallard ducks 19 ------- {A. platyrhynchos), shells of eggs laid by Hg-fed hens were significantly thinner and of poorer quality, than the shells of those laid by controls (Heinz, 1979). Egg shell thinning has been used as an indicator of population trends. In the brown pelican, thinning of 15 to 20 percent has been associated with declining populations on several widely separated islands (Blus et al., 1972). Pollutant effects on the reproductive responses of animals are often mediated through changes in the endocrine system. Rattner et al. (1984) recently reviewed endocrine-induced effects on the reproductive response of birds to environmental pollutants. They noted endocrine-related delays in breeding and reproduction in several species exposed to DDT and its metabolites. For example, in one study, ring doves (S. risoria) chronically fed DDE showed delayed egg laying and mating behavior with subtle changes in the levels of the luteinizing hormone (LH) in females. Endocrine changes were also observed in several studies of Japanese quail. One study showed that ingestion of kepone affected reproduction in Japanese quail (Coturnix c. japonica) directly by interfering with normal egg production of the oviducts and indirectly by altering hormone secretion of the hypothalamo- hypophyseal-gonadal system, thus impairing ovarian follicular development. Ingestion of PCBs affected egg laying capacity in female quails and reduced testicular seminiferous elements in males (Rattner et al., 1984). Reproductive effects were also correlated with endocrine changes in field studies conducted on herring gulls (L. argentatus). Herring gulls exposed to DDE 20 ------- and PCBs exhibited abnormal incubation behavior and hyperactivity. Histological examination and blood hormone levels indicated these effects were due to thyroid dysfunction There is very little information on the effect of environmental pollutants on amphibian and reptile reproduction. Martin (1983) reviewed the experimental data and reported that quinacrine exposure produced testicular lesions in lizards, Cd suppressed spermatogenesis in sexually active frogs, and Pb and Cu disturbed the germination of frog spawn. A major problem in using reproductive success as an ecotoxicological endpoint is that it can be influenced by many factors, including nutritional state, weather, disease, and predation. Thus, the cause for reproductive failure may be difficult to identify in field populations. In particular, there are few studies of effects of toxic pollutants on reproductive success in wild mammalian populations. 1.7 MUSCULOSKELETAL EFFECTS Among the numerous pathological signs that are associated with pollution stress, skeletal abnormalities are most amenable to quantification. Some of the most readily detectable and widely used abnormalities include vertebral and spinal deformities in fish and shell erosion in crustaceans. The vertebral and spinal cord deformities may be manifested as dorsoventral flexures (lordosis), lateral flexures (scoliosis), or backward spinal curvature (kyphosis). These deformities may have a variety of behavioral effects, including impaired swimming performance, lowered feeding rates, impaired ability to avoid predators, decreased 21 ------- ability for territorial defense, reduced ability to compete for a sexual partner, and general physiological weakness (Bengtsson, 1979). Pollutant-induced stresses have been responsible for skeletal abnormalities in fish. For example, skeletal abnormalities have been observed in minnows exposed to Zn and Cd. Zn and Cd may affect the neuromuscular functioning of the fish, resulting in overloading of the vertebrae and skeletal fracture. Interference with calcium metabolism by heavy metal or chlorinated hydrocarbon contamination may also contribute to fish skeletal deformities (Bengtsson et al., 1985). Holcombe et al. (1976) observed a 20 percent increase in skeletal deformities in the second generation of brook trout (S. fontinalis) exposed to Pb over three generations. Spinal skeletal defects such as lordosis (16 percent), scoliosis (32 percent), kyphosis (25 percent), and extreme rigidity and coiling of the vertebral column (17 percent) have been described in the embryos and larvae of a variety of fish exposed to pesticides, such as atrazine, chlorobenzene, and trisodium nitrilotriacetic acid (Birge et al., 1979). Increased incidences of skeletal abnormalities have also been described in amphibian tadpoles (frogs, toads) exposed to DDT, oxamyl and dieldrin (Cooke, 1972, 1981; Martin, 1983) and in sea urchin (Paracentrotus livodus) embryos exposed to Cu, Zn, Se, and Cd during the pre- and post-hatching stages of development (Pagano et al., 1986). Other kinds of physical abnormalities may serve as endpoints for monitoring pollutant effects on a population. In one study, a high incidence of fin and gill 22 ------- raker abnormalities appeared in sea-spawning whitefish (C. lavaretus) exposed to heavy metal pollution (Bengtsson et al., 1985). In another study, dover sole (Microstomas pacificus), collected from the vicinity of a major municipal waste-water discharge site in southern California, exhibited fin erosion and tumor-like growth on the skin (Mearns and Sherwood, 1974). Histological examination showed the fin disease to be external (Wellings et al., 1976) and apparently caused by contact with contaminated sediments around the waste-water outfall. One problem with the use of physical deformities as an ecotoxicological endpoint is that they are not specific responses to pollutants. There may be many causes of such abnormalities, including genetic factors, disease, and many environmental variables, as well as interactions among these factors. Another problem is that some of the commonest abnormalities (fin rot, shell erosion) are difficult to quantify, while more quantifiable deformities rarely occur. The potential effects of physical abnormalities on fitness of the affected organisms remain mostly hypothetical. The very fact that deformed animals are captured shows that the deformities have not prevented growth and survival in the affected environment. Little is known how deformities affect such important population parameters as growth rates, lifespan, reproduction, competitive ability, feeding rates, and predator avoidance in the natural environment. 23 ------- 1.8 GROWTH AND DEVELOPMENT 1.8.1 Developmental Toxicity The developmental period of an organism's life cycle is the period extending from fertilization of the egg through maturity. Pollutants can interfere with development by disrupting early cell division, growth, and migration (embryonic growth) or by interfering later with critical growth phases during organogenesis (e.g., larval, fetal) causing embryotoxicity or lethality, altered growth (e.g., retardation), functional deficiencies (e.g., metabolic, behavioral), structural abnormalities (e.g., terata), and diminished maturation of offspring. A number of studies have examined pollutant effects on development and survival of the organism through the embryonic and larval stages, because of the high sensitivity of these developmental stages to pollutants, and because of the short duration of early development. A variety of studies demonstrate developmental effects resulting from pollutant exposures. In one such study, fertilization success was significantly reduced in sea urchins (e.g., Echinus esculentus) following exposure of sperm to Cd or Zn (Pagano et al., 1986). Amphibian embryonic and larval stages are often very sensitive to pollutants. For example, low pH caused embryonic death in salamanders, while inorganic Cu greatly reduced growth and prevented metamorphosis in frog tadpoles (Martin, 1983). 24 ------- In another study conducted with birds, injecting DDT and its metabolites into gull eggs (Larus californicus) induced abnormal development of ovarian tissue and oviduct in male gull embryos. This feminization process probably resulted from estrogenic action (Rattner et al., 1984). In addition to skeletal abnormalities discussed earlier, pollutant-induced stresses have caused soft-tissue abnormalities. Soft tissue anomalies such as optic malformations were observed in Atlantic silverside (Menidia menidia) embryos exposed to insecticides, rainbow trout (S. gairdneri) embryos treated with benzo(a)pyrene (B(a)P), and embryos of killifish (Fundulus heteroclitas) exposed to magnesium chloride or methylmercury. Cardiac anomalies have been reported for several species of fish exposed to pesticides (carbaryl, parathion, malathion, mercury compounds) (Weis and Weis, 1987). Cell and tissue pathology have often been used to demonstrate incidence of internal injuries or anomalies caused by pollutants. Hose et al. (1984) exposed rainbow trout (5. gairdneri) alevins (from fertilization through hatching) to B(a)P and observed nuclear pycnosis and karyorrhexis in the brain, retina, and muscles; microphthalmia associated with patent optic fissure; abnormal mitosis in hepatic, neural, and muscular tissues; and skeletal malformations in the skull and vertebral column. Possible ecological effects of such morphological abnormalities would be decreased feeding and growth, and inability to escape predation. The teratogenic responses described in these studies suggest that monitoring of abnormal egg, embryos, and larvae may provide a good estimate of the severity of environmental stresses. 25 ------- Behavioral abnormalities may also be elicited in offspring of parents exposed to environmental contaminants. Thus, exposure of three generations of mallard ducks (A. platyrhynchos) to 0.5 ppm of Hg resulted in a significant inhibition of the ducklings' response to simulated maternal calls. The ducklings were hyper- responsive to frightening stimuli in avoidance tests, but their locomotor activity was normal (Heinz, 1979). A similar behavioral effect was observed in ducklings from mallard hens chronically fed 3 ppm of DDE (Heinz, 1976). Studies have demonstrated that the embryo-larval stages of aquatic organisms are particularly susceptible to pollutant stress. Accumulation of organochlorine pesticides (e.g., mirex, DDT) in several species of fish eggs, for example, resulted in high embryonic and larval mortality with the embryos being more susceptible than the larvae (Birge et al., 1979; Livingston, 1977). The rainbow trout was the most sensitive species tested. Reduced survival rates were also observed in embryos, larvae, and juveniles of fathead minnows (Pimephales promelas) exposed to 4-methyl-2-pentanone, 1,4-dimethoxybenzene, benzophenone, and 3,4- dichlorotoluene (Call et al., 1985). In a recent review of early-life-stage toxicity tests on fish, McKim (1985) found that, in 83 percent of 72 comparisons, estimates of MATC obtained from early-life-stage tests were identical to MATCs established by longer, more involved, and more costly complete- or partial-life-cycle toxicity tests. Early- life-stage tests are thus considered to be particularly useful estimators of long- term chronic toxicity. Limitations of utilizing early-life-stage tests are that the duration of exposure may not be sufficient to observe cumulatively toxic effects, 26 ------- and they provide no mechanism for identifying alternative modes of action resulting from toxicity to other life stages (Macek et al., 1978). 1.8.2 Growth Growth provides an integrated index for measuring the physiological status of an organism that has not as yet attained its maximum biomass. It is a widely used endpoint for monitoring the effects of pollutants on both aquatic and terrestrial organisms. The use of changes in growth rate as a measure of pollutant stress in aquatic organisms is illustrated in a study conducted by McKim and Benoit (1971). These investigators monitored the growth rate of juvenile brook trout and found it to be extremely sensitive to pollutant stress. Juvenile fish exposed to Cu for 14 months showed a significant reduction in growth rate. This reduction was inversely correlated with Cu concentrations. By comparison, adult fish were unaffected. The study indicated that long-term growth rates are an effective measure of chronic stress. Other studies have reported growth rate retardation in microinvertebrates at high but sublethal pollution levels. Borgmann et al. (1980) studied the effects of Cd, Cu, Hg, Pb, and As on the growth and survival of copepods. Experiments were conducted seasonally, using naturally occurring water and food rather than defined media, in order to determine the extent to which varying environmental conditions affect toxicity. Growth rates of the copepod population were affected 27 ------- at sublethal metal concentrations. A seasonal cycle in toxicity was observed with all metals except As. Numerous studies indicate that growth rate provides a sensitive endpoint for monitoring pollutant effects on plants. Pollutants may influence forest growth and development via multiple pathways and mechanisms and over varying time scales. This influence may be exerted via physiological disruption of the plant- water balance (osmoregulatory activity), change in forest nutrition, effects on growth and reproduction, or altered resistance to secondary stresses (McLaughlin, 1985). For example, ozone exposure disrupts the stomata-controlled leaf-water balance (Heath, 1975). The physiologic disruption of the plant-water balance may result in diminished capacity of the plant to take up water from the soil or control its water loss to the atmosphere through its foliage. In one plant study conducted by Miller et al. (1977), the vegetative growth of corn shoots was suppressed by both Pb and Cd. Further, the concentration of Cd in corn shoots was increased by the addition of Pb to the soil, while the presence of Cd in the soil reduced the uptake of Pb. Changes in rates of seedling germination and growth have provided sensitive measures of pollutant effects. For example, reductions in growth rate, stem elongation, leaf area, plant and root weight, fruit and seed set, and floral productivity occurred in several plants grown in air with ozone levels ranging from 8 to 10 ppm (Feder, 1973). Reductions in growth and germination also occurred in plants exposed to heavy metal contaminated soil (Walley et al., 1974). This last study is one of a number showing heritable variation in tolerance to 28 ------- heavy metals, thus indicating the potential for evolution of tolerance to toxic pollutants. Pollutant effects on seedlings were also observed by Constantinidou and Kozlowski (1979) in a study of 4-month-old Ulmus americana seedlings exposed to sulfur-dioxide (SO2) or ozone or both. Injury to the leaves was observed within 48 hours after all treatments. Expansion of new leaves was inhibited by the mixture and SO2; expansion of young leaves was inhibited by all three treatments. Leaf emergence was significantly reduced by SO2 and the mixture at the end of the first week; stem dry weight and root dry weight after 5 weeks were reduced by ozone, and/or SO2. Quiescent seedlings responded to all fumigation treatments with severe defoliation. These experiments emphasize that the inhibitory effects of pollutants on plant growth vary markedly for different organs and tissues and at different stages of plant growth. A variety of methods are available for evaluating and predicting forest growth response to air pollution. One such method, developed in a 1982 project entitled FORAST (Forest Response to Anthropogenic Stress), involves surveying and characterizing forest damage based on long-term growth changes and effects on wood density (McLaughlin and Braker, 1985). These dendroecological techniques document growth responses of individual trees over their entire life- spans. Techniques include measuring the width of annual rings and cross-dating to rings measured in other trees of the same species. Confounding elements may be eliminated by statistical analysis (e.g., effects of tree age, competitive status, climate). Additionally, wood density can be measured annually and correlation of growth decline with causal factors (e.g., emissions, meteorological data) 29 ------- can be combined with elemental analysis of tree rings to correlate pollutant accumulation with measured growth changes. Elemental analysis of tree rings provides another possible mechanism for determining pollutant effects on forest trees. For example, increases in Fe, Ti, Cd, Cu, and Mn in trees correspond to increased regional combustion of fossil fuels (McLaughlin and Braker, 1985). The sophisticated technology that is now available (e.g., PIXE-photon - induced X-ray emission spectrophotometer, ICP, inductively coupled plasma emission spectroscopy, laser spectrophotometer, and portable computer-based systems) makes it possible to collect and analyze physiological data in the field, in situ, on foliage-stressed and nonstressed trees. Using forest growth as an ecotoxicological endpoint presents a number of problems. Multiple surveys are necessary to determine changes in rates of growth and wood density. Many potential confounding factors, such as soil characteristics, geologic factors, elevation, aspect, wind, temperature, soil moisture capacity, disease, grazing, and competition, cannot be controlled, so their effects must be estimated statistically. There might be indirect as well as direct effects of pollutants, such as pollutant-induced changes in nutrient availability and altered resistance to secondary stresses. 30 ------- 1.9 GENOTOXICITY Mutation is a permanent change in genotype other than one brought about by genetic recombination. Genetic agents can produce changes in cellular deoxyribonucleic acid—a group of changes that are implicated as the initial events in carcinogenesis. Analytic methods have been developed that make it possible to measure genotoxic endpoints in the laboratory. The two most frequently used methods for detecting genotoxic effects are cytogenetic analysis and sister chromatid exchange (SCE). The two most popular in vivo test models are the central mudminnow (Umbra limi) and the eastern mudminnow (Umbra pygmaea). Both species are available in large numbers and are easily captured; they possess a low number (22) of large, easily observed chromosomes; they survive well in a laboratory atmosphere; and they possess the microsomal system necessary for activating promutagens to genotoxic intermediates (Bishop and Valentine, 1982). In one study, exposure (intraperitoneal injection) of the central mudminnow {U. limi) to two direct-acting mutagens—methyl-methanesulfonate (MMS) and N- methyl-N'-nitro-N-nitrosoguanidine (MMNG)--and to two indirect-acting mutagens --cyclophosphamide (CP) and dimethylnitrosamine (DMN)--resulted in a significant dose-dependent increase in the frequency of SCE (Bishop and Valentine, 1982). In another study, the eastern mudminnow (U. pygmaea) exposed to polluted Rhine water containing organic acids, esters, aldehydes, and phenolic compounds showed a time-related increase in SCEs in gills and testicular tissues (Alink et al., 1980). The SCE method of analysis was also successfully used to examine the genotoxicity of mitomycin C (MMC) in the marine polycheate worm (Neanthes 31 ------- arenaceodentata). A dose of 5 x 10"^ mol/1 of MMC increased the rate of SCE from a baseline frequency of 0.14/chromosome to 0.5/chromosome (Pesch and Pesch, 1980). Other studies showed that the larvae of the mussel, M. edulis were 1.5 times more sensitive to the genotoxic effects of MMC than the polycheate worm (Harrison and Jones, 1982). One problem in using SCE as an ecotoxicological endpoint is that information on SCE is available for only a few species. In addition, the mechanism for SCE induction is not well understood, and the biological significance of SCE is unknown. Some researchers, in fact, question whether SCE is correlated with a mutational event (Bishop and Valentine, 1982). 1.10 CARCINOGENICITY The results of some studies have demonstrated that selected fish and invertebrates could be used as indicators of carcinogenic pollutants in the environment through monitoring carcinogenic effects, such as the development of benign or malignant neoplastic growth of cells in tissues. Chemical induction of cancer has been demonstrated in numerous environmental epidemiologic studies (Couch and Harshbarger, 1985). Approximately 300 species of fish and 15 species of bivalve molluscs have varied spontaneous and experimentally induced tumors. The etiology of bivalve neoplasms is problematic despite some studies that implicate one or a combination of viral, chemical, and genetic factors. On the other hand, there is clear evidence of induction in fish. Many of the fish neoplasms are idiopathic. Others, such as the lymphomas of 32 ------- northern pike, appear to be caused by a retrovirus. Chemical and genetic impairment have been implicated in neoplasms originating from numerous cell types in certain platyfish/swordfish hybrids. Liver and skin cancers in various bottom fish have been associated with point source concentrations of environmental chemicals. These cancers have been produced in the laboratory by extracts of the bottom sediments of polluted waterways. Studies have shown that fish liver metabolizes carcinogenic pollutants into reactive intermediates and that fish experimentally exposed to known carcinogens have frequently developed liver neoplasms (Couch and Harshbarger, 1985). Although the most frequent tumor type reported in fish is the hepato- carcinoma, other types such as fibrosarcomas of the skin, tumors of the gill, and other organs have also been reported (Edwards and Overstreet, 1976; McCain et al., 1977; Meyers and Hendricks, 1982; Sindermann, 1980). The development of neoplasms has been strongly correlated with pollutant exposures in other vertebrate classes. For example, neotenic tiger salamanders (Ambystoma tigrinum) inhabiting a small isolated lagoon (Lubbock County, Texas), which was heavily polluted with secondarily treated domestic sewage, developed neoplastic and related skin lesions. The neoplasms were of epidermal, fibrotic, and melanocytic origin. No neoplasms were discovered among the larvae sampled from 16 proximal nonsewage lagoons (Rose and Harshbarger, 1977). 33 ------- It is important to develop methods for introducing and monitoring sensitive fish/invertebrate species at or near suspected pollutant sources in an aquatic ecosystem. The current histopathologic techniques for positive identification of fish cancer should be backed up by identification of biochemical endpoints. These endpoints would include specific enzyme inductions in selected tissues, early detection of carcinogenic changes, and histopathological bioassays of fish liver and bile extracts and invertebrate digestive glands. Fish bioassays, especially those using embryos, have been advocated to pre- screen chemicals and mixtures for carcinogenicity with considerable advantage in cost and time over similar tests for carcinogenicity in rodents. Sensitivity is excellent, and interpretation is rarely complicated by spontaneous tumors (Couch and Harshbarger, 1985). However, synergistic effects and viral interactions complicate analysis, especially in the natural habitat. Also, there have been few studies of pollutant-induced carcinogenicity in avian and mammalian populations. 1.11 MODIFYING FACTORS In an ecosystem, each organism lives within a number of physical, biological, and chemical constraints because of exposure to many factors that may act independently or together. Under these constrained circumstances of existence, any additional environmental stress (such as toxic pollutants) may exacerbate or ameliorate pre-existing conditions. These multifactorial interactions and stresses commonly occur among inhabitants of all ecosystems. 34 ------- When there are multiple stresses, their effects may be additive or interactive. If the effects are additive, the combined effect of two or more stresses would be equal to the sum of the independent effects. If the effects are interactive, they may be classed as either synergistic or antagonistic. In a synergistic interaction, the combined effect of the multiple stresses is greater than the sum of their independent effects. In an antagonistic interaction, the combined effect of the multiple stresses is less than the sum of their independent effects. 1.11.1 Microbial Toxicity Multifactorial interactions, including pollutant toxicity and the stress response of exposed microbial communities, have been extensively reviewed by Babich and Stotzky (1980). These authors have shown that abiotic, physicochemical factors influence pollutant toxicity to microbes and viruses either by potentiation or attenuation and have emphasized the characteristics of the specific environment into which the pollutants are deposited. Both cationic and/or anionic elements of the environment can influence the toxicity of heavy metals to microbes and viruses. For example, Mg reduced the toxicity of Co, Cd, and Ni to bacteria (Escherichia coli) and filamentous green algae (Hormidium rivulare); in a similar reaction, Ca reduced the toxicity of Zn or Cd to H. rivulare and of Cd to Aspergillus niger and E. coli. In a synergistic reaction, the combined toxicity of Cd and Zn to Klebsiella aerogenes was greater than the sum of their individual toxicities. 35 ------- The inorganic anionic composition of the environment is known to influence the chemical form of metals. For example, adding phosphate reduced the toxicity of Pb to H. rivulare and of Cu to Aerobacter aerogenous. The reduction probably resulted from the formation of sparingly soluble salts of both elements, i.e., Pb3(P04)2 and Cu(P04)2. The pH of the environment into which the pollutant is deposited is important because it may influence a chemical's form, solubility, and toxicity. For example, as the pH is increased, Cd+2 is sequentially hydroxylated to other species, i.e., Cd(OH)^, Cd(OH)2, Cd(OH>3, and Cd(OH)42~. Each species is differentially toxic to the microbiota and possesses differential affinities to viral particles, microbial cells, and other particulates such as clay minerals, hydrous metal oxides, and humic acids. The oxidation-reduction potential (Ejj) of an environment may influence toxicities to microbes because it may influence the form of some inorganic anions (e.g., S2" vs. SC>42~), and thus influence the solubility, mobility, and toxicity of heavy metals deposited into that environment. For example, Cu+ was more toxic to E. colt than Cu2+, while Fe2+ showed a greater potential for mutagenicity in E. coli than Fe^+. The effect of temperature on the physiological and biochemical state of the microbiota is best illustrated by the inhibition of the growth of the alga Cyclotella meneghiniana by Cr, where increased inhibition accompanied an increase in temperature from 5 to 25°C. 36 ------- In bacteria (Bacterium B III39), increases in hydrostatic pressure in the environment resulted in increased toxicity of Ni to microbes. In addition, the presence of organic particulates reduced the toxicity of Cd to a variety of organisms (Bacillus megalerium, Aspergillus fischeri, Penicillium). These effects resulted from the ability of particulates to remove heavy metal cations from solution. Microbial response to heavy metal pollutants in the environment may also be influenced by the presence of other chemicals. This was demonstrated in experiments in which bacterial (E. coli, Pseudomonas sp.) levels of respiration and nitrogen fixation were altered in the presence of Cd to a variable extent depending upon the species of Cd and the presence of the citrate ion (Lighthart, 1980). 1.11.2 Multifactorial Interactions A strong positive correlation has been found between levels of selenium (Se) and Hg in marine mammals and fish. The antagonistic interrelationship between Se and Hg has been established in many studies (Beijer and Jernelov, 1978). Synergism and antagonism have also been noted in several plant species exposed to ozone and SO2. Tingey et al. (1973) observed significantly more foliar injuries from mixed ozone/S02 treatment than from additive effects of single gas treatment in tobacco and alfalfa. In other species such as cabbage, broccoli, and tomato the foliar injuries from mixed gas exposure were additive or less than additive. 37 ------- Plant injury from ozone, other oxidants, and air pollutants (SO2) may be differentially affected by using agricultural chemicals, especially fungicides and nematocides. Tobacco plants were protected against ozone-induced injury by benomyl (methyl -1 -butyl-carbamyl-2-benzimidazolecarbamate), dodine (dodecylguanidine acetate), and maneb (manganous ethylenebisdithiocarbamate). Benomyl was most effective, reducing leaf injury by 59 percent; maneb decreased the number of brown spot lesions per leaf by 66 percent. Smaller decreases were observed with benomyl and dodine (Reinert and Spurr, 1972). Ozone also enhanced Cd-induced injury in cress shoots (Lepidium sativum) by stimulating early development of chlorosis and necrosis. The combined elements induced these injuries at lower concentrations than the separate elements (Czuba and Ormrod, 1984). Assessment of nutritional stresses on populations of cladocerans (Daphnia magna) maintained on a vitamin-enriched algal diet showed that these animals were less sensitive to chronic Cu stresses than those fed a trout-granular diet (Winner et al., 1977). This conclusion was based on the mean brood size, survival, and the instantaneous rate of population growth. Mortality of blue crabs (Callinectes sapidus) in a DDT-contaminated marsh was increased when environmental temperatures were low (Koenig et al., 1976). Examination of the hepatopancreas and swimmeret muscles from dead or dying crabs revealed high concentrations of residual DDT (39.0 and 1.43 ppm, respectively). These results suggested reduced temperatures and body burdens of 38 ------- DDT interact to produce acutely toxic effects. Several mechanisms of action were proposed. Survival was reduced and development was delayed in larval blue crabs (C. sapidus) following exposure to 50 ppb of Cd at 25°C. Twelve combinations of Cd and salinity at 25°C were examined in this study (Rosenberg and Costlow, 1976). In similar tests with the mud-crab (Rhithropanpeus harrisii) larvae, where 63 combinations of Cd, salinity, constant temperatures, and cycling temperatures were used, combination exposures of 50 or 150 ppb Cd and 10, 20, or 30°/oo salinity significantly reduced survival (from hatched to first adult stage) of the larvae at the 20 and 35°C temperature extremes. Results from cycling temperature experiments (20 to 25° and 25 to 30°C) showed a stimulatory effect on survival of larvae compared to constant temperature, both in presence and absence of chronic Cd exposure. Acidification may influence metal-organism interactions in at least two ways: a decrease in pH increases solubility of heavy metals, making them more available to be taken up by organisms; and a pH decrease may affect biological sensitivity to metals at the level of the cell surface (Campbell and Stokes, 1985). Elevated levels of aluminum exacerbate toxic effects of reduced pH on crustacean zooplankton (Havas and Hutchinson, 1982). Examples of antagonism also occur; aluminum can increase survival of some cladocerans and fish at some pH levels. Proposed mechanisms for this are that trace metals may compete with H+ ion for exchange sites on the gill surface, or they may reduce membrane permeability (Havas, 1985). 39 ------- Heavy metals can reduce tolerance to other environmental stresses, such as drought stress. For example, cadmium (20 ug/g soil) significantly reduced the dry weight yield of two native plant species (Andropogon scoparius, Monarda fistulosa) under drought stress. The effects of the stress and Cd treatment appeared to be additive (Miles and Parker, 1980). There are many examples of interactions between pollutants and biotic environmental stresses. Studies on the effect of air pollutants on the plant-insect relationship have demonstrated increased susceptibility of greenhouse-grown soybeans to the Mexican bean beetle (Phaseolus vulgaris) following fumigation with SO2 (Hughes et al., 1982). Adult females showed feeding preference for fumigated leaves, produced eggs with higher viability, and showed increases in both number of eggs per brood and in number of broods produced. A possible mechanism of action involves modification of amino acid or sugar levels in the soybean plants. Other examples of interactions between pollutants and predator- prey relationships are covered in Chapter 2. Toxic pollutants can also increase susceptibility to disease. In the marine fish populations, the incidence of various nonspecific diseases (e.g., ulcers, lymphocytosis, and fin rot in fish and shell erosion in crustaceans) are good qualitative indicators of environmental stress. In one study, Hetrick et al. (1979) reported increased susceptibility of rainbow trout to infectious hematopoietic necrosis virus following exposure to less than 0.01 ppb of Cu. Mortality was twice as high in Cu-exposed fish as in the controls. Suppression of the immune response is probably the mechanism involved in causing increased susceptibility. 40 ------- In another study, eels exposed to water contaminated with 30 to 60 ppb of Cu died with vibriosis (Vibrio anguillarum) infection, whereas eels kept in non- contaminated water remained healthy (Rodsaether et al., 1977). These interactions illustrate the many types of interactions that can take place between environmental pollutants and abiotic and biotic components of the ecosystem. This raises serious questions about the utility of single-species or single-factor laboratory toxicity testing in developing ecotoxicological endpoints for ecological risk assessment. 1.12 CONCLUSION Ten major categories of physiological ecotoxicologic responses (endpoints) that may be used to evaluate adverse effects of an environmental pollutant on populations have been identified. They include acute mortality, biochemical alterations, osmoregulatory effects, respiratory effects, behavioral effects, reproductive effects, musculoskeletal abnormalities, and effects on growth and development, genotoxicity, and carcinogenicity. The ecotoxicological endpoints most directly related to individual success are acute mortality, growth and development, and reproduction. A frequently recommended approach is to begin with acute lethality tests to establish a crude estimate of toxicity, followed by testing for chronic sublethal effects on growth and reproduction (Macek et al., 1978; Monk, 1983). The primary advantage of using growth and reproduction as endpoints is that they integrate all other physiological processes necessary for individual success. The major disadvantage 41 ------- of these endpoints is the time and expense required to conduct full-life-cycle chronic toxicity tests. Other physiological and biochemical responses provide more rapid and inexpensive measures of pollutant effects. These are potentially useful for early detection of pollutant effects, thus allowing corrective action before irreparable damage occurs. In addition, these endpoints are often useful in determining the mechanisms of ecotoxicological effects. However, the relationship of these short- term physiological responses to individual growth and reproductive success is rarely well established. Among the few exceptions to this include the demonstration that, in fish, early-life-stage toxic effects are highly correlated with MATCs derived from life-cycle chronic toxicity tests. Other short-term physiological responses cannot be recommended as ecotoxicological endpoints until relationships between them and growth and reproductive success are determined. Within an ecosystem, organisms are continually exposed to many environmental stresses that may act independently or together. Any additional, adverse environmental condition could further exacerbate pre-existing stresses. It has been demonstrated that toxic pollutants may interact with other environmental stresses. These interactions may be either synergistic or antagonistic and may involve both abiotic and biotic factors in the environment. The numerous examples of interactions among environmental factors make it clear that laboratory single-species, single-factor tests will not be adequate to estimate all the ecological effects of contaminants. This points to the necessity of further relating ecotoxicologic effects on individual organisms to population- and ecosystem-level effects of pollutants. 42 ------- CHAPTER 2 POPULATION-LEVEL ENDPOINTS 2.0. INTRODUCTION Population-level endpoints form an important component of an ecological risk assessment. These endpoints focus on the relationship between populations and the physical, chemical, and biological factors in their environment. The first step in a population-level assessment requires selection of the most appropriate endpoints. Pollutants may affect both birth and death rates of populations. These effects may cause changes in other measurable endpoints, such as abundance, age structure, distribution, genetic makeup, and life history patterns. Other potential endpoints are measures of such species interactions as predator- prey relationships and interspecies competition. Selection of the most appropriate population(s) to monitor is the next step in a population-level assessment. Population-level endpoints provide essential information about the structural dynamics of an ecosystem. In addition, they can often provide useful indicators of whole ecosystem changes in response to pollutant stress. 2.1 POPULATION DYNAMICS The concept of population dynamics was first used by Elton in 1933 to describe the area of study "concerned with rates of increase, fluctuations in numbers, and the relation of problems of numbers to the environmental factors 43 ------- which influence the population." This subject deals with the influence of environmental factors on the rates of birth, death, immigration, and emigration (Odum, 1971). Population size is determined by the sum of gains from birth and immigration minus the losses due to death and emigration (Moriarty, 1983). 2.1.1 Birth Rate The birth rate of a population has long been used as a means of determining its health. The population birth rate is determined by combining age-specific fecundity rates with the age structure of the population (Warren, 1971). For some species, there is a clear correlation of intraspecific competition (density- dependent factor) with age-specific birth rates and age at maturity (Frank et al., 1957, cited in Warren, 1971). Fisheries investigations generally do not measure the birth rate because newly hatched/born individuals are often too small or too dispersed to catch. Instead, fisheries studies measure recruitment, the young large enough to be caught with fishing gear (Royce, 1972). Rate of birth (or recruitment) is a good measure of contaminant stress, as a change can affect not only the success of individual organisms, it can also be detrimental to population abundance, age structure, and gene pool. Birge et al. (1980, cited in Sheehan, 1984a) suggested that a 10-percent-or-greater increase in mortality in the developmental stages (embryonic, larval) would significantly alter population dynamics in natural populations. 44 ------- 2.1.2 Death Rate The death rate is calculated as the number of individuals dying per unit of population over time. As in the case of birth rates, the death rate is generally described by age class. Death rates are inversely related to survivorship. A survivorship curve is plotted as the number of individuals surviving to successive age classes (Andrewartha and Birch, 1954; Warren, 1971). The death rate is not easily measured in either terrestrial or aquatic animals. Dead and dying animals are rarely seen in nature. As a result of this, death is generally estimated indirectly from other data, such as historical information on the number of survivors that are members of the same age class throughout the life cycle (Royce, 1972). 2.1.3 Population Growth Models Evaluation of populations has rested on the development of mathematical descriptions of "populations in terms of abundance, mortality, and reproduction, defining relationships between life history and population growth, explaining fluctuations in abundance, and identifying regulatory mechanisms" (Barnthouse et al., 1986). Many population models have been designed to manage exploited populations (fisheries in particular) or to control nuisance species such as Hydrilla. The simplest population growth models describe exponential growth. However, no population can grow in this fashion forever; in reality population sizes are limited. There is little consensus in the scientific community as to the 45 ------- mechanisms regulating population size. The argument centers around the role of density-dependent factors versus density-independent factors. Density-dependent factors, e.g., intra- and interspecific competition, tend to stabilize the population growth; density-independent factors, e.g., weather, act upon a population without regard to its equilibrium size and are considered nonstabilizing influences. Murray (1979) states that the prevailing view is that the size of populations is generally regulated by density-dependent factors; however, in some populations "density-dependent factors are inadequate, nonfunctional, or nonexistent; thus, periods of exponential growth are followed by population crashes." However, Andrewartha and Birch (1954) dispute a distinction between density-dependent and density-independent factors. They argue that all factors can act in a density- dependent fashion, but most populations are primarily influenced by nonstabilizing, density-independent processes. A third view is that population growth is affected by both density-dependent and density-independent factors (Ricklefs, 1973; Warren, 1971). Most population modeling deals only with density-dependent limitations on growth. Population growth rates are generally described by the logistic equations as modified for competition by Lotka and Volterra. Most basic population models are based on the logistic equation (Murray, 1979). The basic logistic equation is represented by: dN = rN (1-N) where N=number of individuals dt K K=carrying capacity t=time r=per capita rate of increase which may be restated verbally as: 46 ------- the population's = the population's actual growth rate potential growth x rate the proportion by which the population is below the environment's carrying capacity (Moriarty, 1983) Carrying capacity (K) is defined as the number of individuals that the resources of a habitat can support. It can be thought of as the total resources available divided by the minimum maintenance requirement of each individual (Ricklefs, 1973). Thus, carrying capacity is determined by characteristics of both the environment and of the population (McNaughton and Wolf, 1979). The per capita rate of increase (r) = b - d, where b = instantaneous birth rate and d = instantaneous death rate. The logistic equation describes a population growing in a sigmoid fashion, rising to an equilibrium population size at the carrying capacity. This model assumes a stable age structure in the population with overlapping generations (Moriarty, 1983; Murray, 1979). Population growth rates are assumed to be strictly density-dependent in the logistic model, with birth and death rates linearly related to population size. Using this model, population growth rates decline as population size increases, reaching zero as the carrying capacity is attained. The model can be modified to include coefficients for competition, predation, and other two-species interactions. Population growth models used in fisheries often take a somewhat different form. In the logistic equation, population growth rate is represented by the number of individuals over a unit of time. Fishery models utilize changes in population biomass over a unit of time. There are a variety of fishing yield models; however, they all are based on the general relationship between the unit 47 ------- stock and the effects of additions and losses. In general, recruitment and growth add to the unit stock while natural mortality and fishing mortality decrease it. This is in complete agreement with the earlier descriptions of changes in population numbers as described by Moriarty (1983) and Barnthouse et al. (1986). However, the actual mathematical representation of this relationship is different for fish populations. For example, this logistic model is often used in fishery studies (Royce, 1972): dP = kiP(L-P) where dt P=total biomass of fish in stock t=time L=limiting biomass to the stock ki=growth rate constant Another approach to predicting population growth is the Leslie matrix model. This model predicts future population size and age structure of a population, given both the structure at the present time and a matrix whose elements represent age-specific fecundity and mortality (Usher, 1972). It can include density-dependent factors as well as density-independent factors (Murray, 1979) and can be modified to include some biological interactions, such as competition. A common application of this model is in determining survivorship and recruitment rates to estimate the maximum allowable harvest of fisheries, forests, or wildlife (Boyce, 1977). In the basic Leslie matrix model, birth and death rates remain constant at the initially measured values. As a result, the model describes exponential population growth. However, the model can be modified by incorporating logistic-type functions to make fecundity and mortality density- dependent (Usher, 1972). 48 ------- There are other mathematical models that represent both terrestrial and aquatic species. Different models are used depending on the type of population being studied and the amount of data available on parameters, such as age structure, size composition, and mortality rates. A brief classification of other analytical methods of evaluating population interactions is given in Table 2.1. These are covered in greater detail in a separate document, "Ecological Model Selection Criteria" (TRI, 1988b). The logistic equation is an adequate description of population growth of organisms with simple life cycles in constant laboratory environments, but populations with more complex life cycles and those in natural environments seldom follow logistic growth. In particular, the stable carrying capacity predicted by the logistic equation is not achieved in natural populations; in reality, population numbers fluctuate (Krebs, 1985). As a result, efforts to model toxicant effects on population growth have had limited success. Barnthouse et al. (1986) concluded that population theory cannot now provide models that accurately predict long-term consequences of toxic pollutant release. The greatest success in using population growth models in studying pollutant effects has been in laboratory studies of microbial growth. For example, Christensen and Nyholm (1984) fit logistic, Weibull, and probit growth models to growth rates of the alga Selenastrum capricornutum exposed to potassium dichromate and copper, and of the alga Scenedesmus suspicatus affected by 3,5- dichlorophenol and potassium dichromate. The concentrations of toxicant which gave 10 percent and 90 percent growth rate reductions (EC10 and EC90) were 49 ------- Table 2.1 Model Classification System Model Fully Described Quantified type defined experimen- tally Partially simpli- fied and quanti- fied Economic Objectives criteria quanti- fied STATISTICAL X Information from data DIORISTIC Distinguish component planning research and teaching holistic description COMPONENT Complete quanti- fication of system. Transfer functions experimen- tally determined STRATEGIC X X Analysis or simu- lation of complex systems MANAGEMENT Southwood (1978) Decisions on management of complex systems 50 ------- lowest for the Weibull model. Thus, the Weibull model was most sensitive in detecting the EC 10, which is often considered a threshold level in environmental risk assessment. 2.1.4 Other Population Endpoints Although population growth models, at their current state of development, do not appear to be useful for determining effects of pollutants on natural populations, certain components of the models and other population characteristics can be useful endpoints. As discussed earlier, birth and death rates are potential measures of pollutant effects. Other population characteristics that are functions of age-specific birth and mortality rates, such as population size, growth rates, distribution, age structure, or genetic composition, are also potential ecotoxicological endpoints. Alterations in species abundance are easily observed and are clearly related to the health of the population. A decline in population number is often an early indicator of population stress. Distribution alterations may also be very apparent. For example, populations may be most susceptible to pollutant stress near the limits of their distribution and in environments in which they are only marginally suited. Therefore, changes in spatial distribution may be an indicator of pollutant stress (Sheehan, personal communication). Effects of pollutants on per capita growth rates (r) in laboratory populations have been determined from age-specific birth and mortality rates. For example, Gentile et al., (1982) studied effects of mercury and nickel on population growth 51 ------- rates of the marine mysid, Mysidopsis bahia. Metal concentrations at which per capita growth rates were reduced to zero were closely correlated with other estimates of chronic toxicity. The age structure of a population, which is determined by mortality and natality rates in various age groups, can also be used as a population-level endpoint. For example, acidification of lake waters often causes failures of reproduction in fish populations, resulting in a shift in age and size structure of the population to one consisting of older and larger fish. Despite the absence of recruitment, large populations of long-lived fish may persist for several years (Schofield, 1976; Schindler et al., 1985). So monitoring of age structure can provide an early indicator of pollutant effects on populations. Finally, pollutants may have effects on the genetic composition of populations. Elimination of all but resistant genotypes may reduce genetic variability within a population, which may impair the ability of the population to respond to other naturally occurring stresses. As a possible example of this, heavy-metal-tolerant genotypes of plants are frequently competitively inferior to nontolerant genotypes on uncontaminated soils (Antonovics et al., 1971). 2.2 LIFE-HISTORY STRATEGY Detailed knowledge of the life-history strategy of a species is essential to the evaluation of the effects of chronic pollution on that population. Life-history traits, including fecundity, growth and development, age at maturity, parental care, and longevity of an organism, have some genetic basis and hence are subject 52 ------- to natural selection. Natural selection adjusts these traits to maximize the fitness of each organism (Ricklefs, 1973). The two most important factors in determination of a life-history strategy are age-specific birth and survivorship (or death) rates within a population. As described earlier, these factors are the driving variables in many population models. K- versus r-selection is a life-history theory describing evolutionary responses to environmental variability and patterns of mortality. The terms r- selection and K-selection come from the exponential and logistic representations of population growth; r-selected species are characterized by high population growth rates. They are dominant in early-succession ecosystems and in systems following disturbance (environments in which density-independent mortality is high) because they have the ability to disperse rapidly to unoccupied habitats and increase population size rapidly. This is typical of invertebrates and many fishes. K-selected species, on the other hand, are characterized by low mortality and high competitive ability. They are favored in late-successional-stage ecosystems where mortality is density dependent, abundance is near carrying capacity, and abundance is largely determined by biological interactions (competition, predation). Large mammals and trees are typical K-selected species. Knowledge of a population's life-history strategy is extremely important when selecting appropriate indicators of stress in ecological risk assessment. All of the life-history parameters are interrelated; a change in any one of these 53 ------- factors may cause a complementary change in another. Relationships among life- history characteristics are discussed below. 2.2.1 Parental Care and Fecundity Intense parental care increases the survival rate of offspring (Ricklefs, 1973). The more protection and feeding that the young receive, the greater their survivorship. Any disruption of the ability of a parent to care for its young may have an adverse effect on the survival of the young. Examples of disruptions of parental behavior by pollutants are discussed in Chapter 1. The intensity of parental care is inversely related to fecundity (Ricklefs, 1973). Lack (1954) proposed that the ability of parents to feed the young limits brood size and that the average brood size maximizes the number of surviving offspring. One experimental study has shown that offspring growth decreases as brood size is increased. However, contrary to Lack's hypothesis, offspring survivorship did not decrease as brood size was increased. The number of surviving offspring per brood was maximized for larger broods than the average naturally occurring brood size (Nur, 1984). 2.2.2 Parental Care and Growth/Development Species that provide a high degree of parental care generally have broods that are altricial (initially underdeveloped) (Ricklefs, 1973). Conversely, species that provide little parental care generally have broods that are precocial (capable of providing for themselves). A high degree of parental care permits the young 54 ------- to expend most of their energy on rapid growth and development. Any pollutant that reduces the ability to care for altricial young can reduce their survival. 2.2.3 Growth and Age at Maturity Many species, including certain invertebrates, fish, reptiles, amphibians, and plants, continue to grow throughout life. In such species the fecundity rate is directly proportional to the size of the organism (Ricklefs, 1973). Reproduction requires energy and nutrients that otherwise could be allocated to growth. Fecundity early in life reduces organism growth, and this can reduce future fecundity. Survivorship patterns may determine the life-history strategy favored by natural selection (McNaughton and Wolf, 1979). In organisms with higher survivorship, natural selection may favor expenditure of energy on growth to increase total fecundity. In organisms with low survivorship, natural selection will favor early reproduction and high expenditure of energy on reproduction. These are the reproductive strategies expected in K-selected and r-selected species, respectively. 2.2.4 Reproduction and Life Expectancy Larger litter size appears to reduce the life expectancy of females (and nurturing males). Murray (1979) cites eight different studies of both vertebrates and invertebrates in which virgin females were found to live longer than reproducing females. He also cites several studies in which population mortality increases during the reproductive season. 55 ------- 2.2.5 Longevity and Body Size Body size is strongly correlated with animal longevity. Larger mammals tend toward longer life and smaller individual broods; they also exhibit extended interbirth periods (Eisenberg, 1980). Diet also has an effect on longevity for large animals; large herbivores tend to live longer than large carnivores. As body size increases, mortality rate decreases due to a lack of predators large enough to capture such prey, an increased ability to find food, and the increased buffering against microclimatic changes that results from heavier body weight (Ricklefs, 1973). 2.2.6 Compensatory Mechanisms of Life-History Strategy Stress to a population does not always cause a change in population numbers. Populations have means of compensating for change just as ecosystems do. At the ecosystem level, compensation may be in the form of replacement species; at the population level, compensation occurs through changes in life-history strategy. Such compensatory mechanisms may be very important in the success of a population subjected to pollutants or other stresses. They should be considered when assembling an ecological risk assessment. Jensen and Marshall (1983) noted population-level compensation mechanisms in a population of cladocerans exposed to metal contaminants. Population birth rates of Daphnia galeata mendolae exposed to cadmium increased to compensate for increased death rates. This relationship has also been observed in other species, and life-history theory predicts that females are under selective pressure 56 ------- to produce more offspring at reduced population levels (Fowler, 1981). Decreased availability of food or parental care for the young can cause phenotypic changes, including asynchronous hatching (in birds), cannibalism of siblings, and selective starvation of the weakest in the litter (McNaughton and Wolf, 1979). Knowledge of a population's life-history strategy is vital to risk assessment of the ecosystem (Sheehan, 1984a). As life-history traits are interrelated, any pollutant stress affecting one trait may affect other aspects of life-history. For example, a chemical that reduces reproduction may result in a longer life expectancy for particular individuals; however, in the long run, it can lead to extinction. Resources that would normally be devoted to producing and caring for young can be used by the organism to enhance growth and survival. Thus, compensatory mechanisms may ameliorate contaminant effects. However, if a pollutant stress exceeds the compensatory capacities of the organisms, the population will decline or disappear. In addition, energy expended on compensatory responses may reduce a population's ability to respond to other naturally occurring environmental stresses. 2.2.7 Life Histories and Response to Contaminants There seem to be no universally applicable generalizations concerning sensitivities of different taxa to pollutants (Sheehan, 1984a). However, life- history strategies of organisms may be good predictors of their response to pollutants. Pollution effects can be loosely separated into two categories: intermittent acute toxic effects and chronic effects (alternatively termed disturbance and stress, Gray, 1979). During recovery from intermittent toxic 57 ------- events, the community would be expected to be dominated by r-selected species. This is because their entire life-history strategy is adapted to colonization and growth in disturbed habitats: high dispersal, rapid growth and reproduction, and short generation times. As recovery proceeds, these species are gradually out- competed and replaced by K-selected species. An example is the polychaete Capilella capitata, which develops large, transient populations following pollution or other disturbances (Gray, 1979). Organisms with stress-tolerant life histories would be expected to be least affected by chronic pollution. Stress-tolerant organisms share some of the life- history characteristics of K-selected species: low dispersal, low reproduction, and long life spans. However, unlike K-selected species, stress-tolerant organisms allocate more resources to maintenance of productive tissue in very stressful environments. These characteristics would predispose organisms with stress- tolerant life-histories to success during chronic pollution concentrations that may be sublethal to them but lethal to r- or K-selected organisms (Grime, 1977; Gray, 1979). 2.3 CONCLUSION Moriarty (1983) aptly states that "no population remains constant in number forever." Population sizes increase and decrease in response to many factors, such as weather, availability of food, and competition. Pollution is only one factor that can influence population size. Even when an effect is directly traceable to a particular pollutant, populations may have mechanisms to compensate for the stress. Stress-tolerant life histories may be pre-adapted to 58 ------- chronic pollutant stress, while r-selected organisms appear to be pre-adapted for recolonization during recovery from intermittent acute toxic disturbances. However, life histories of organisms are adapted to a suite of physical, chemical, and biological characteristics of the environment. Compensatory responses to a pollutant may reduce a population's ability to cope with other environmental regulating factors. While population-level responses have been used for decades to determine the effects of pollution, it is important to realize that an in-depth understanding of the physical and biological processes regulating populations is necessary before effects of contaminants can be predicted. For the most part, current population models cannot effectively deal with all of the physical and biological processes affecting populations. As an endpoint, population dynamics are extremely complex and absolutely species-specific. There are no formal methodologies for using changes in population dynamics as a risk assessment tool; however, knowledge of the dynamics of a population is necessary to any assessment of stress to a population. 2.4 METHODS FOR SELECTING APPROPRIATE POPULATIONS TO MONITOR Evaluation of individual species forms an important part of any ecosystem study. Individual species effects can provide readily measurable indicators of ecosystem health. They pinpoint some of the factors responsible for changes in population density. However, individual species do not exist in ecological vacuums; interspecific interactions, such as competition or predation, have strong 59 ------- effects on population dynamics. These interspecies interactions may be very sensitive to effects of pollutants. An important initial step in assessing population-level ecotoxicological endpoints is the selection of the most appropriate populations for study. The methods for assessing population-level changes range from individual evaluation of all species to the evaluation of certain representative species of the community. There are benefits to using each of these methods. They are discussed in detail below. 2.4.1 Total Population The total-population approach involves evaluation of the total number of organisms present in the area of concern. Measurements are made using one of three major field sampling techniques. The first is the nearest neighbor method. In this method, a geographic point is randomly selected and then is searched in ever widening circles until an organism is found. The searching continues until a second member of the same species is found. The density of individuals per unit area is determined by using simple mathematical expressions based on the distance between the nearest neighbors (Southwood, 1978). This method and several of its variations have been used with some success in making rough estimates of plant and insect populations present in very small areas (such as the organisms present on one plant). This method is most useful for immobile, sedentary, or territorial organisms; it is less useful for mobile animals (Tanner, 1978). 60 ------- More accurate measures of determining a total population include marking techniques or taking absolute samples of a unit area. Techniques for marking animals, particularly mark and recapture, have been used on large or slow animals that can be readily marked, such as snails and many vertebrates. An advantage of mark/recapture techniques is that they can provide estimates of birth and death rates in addition to population size. The total-population approach is quite comprehensive and is useful for population studies. This approach samples quadrats that are representative of the system of interest. Sedentary organisms, such as plants or barnacles, may be censused nondestructively. But in many other cases, samples must be taken of the organisms. present in air, soil and litter, water, or sediments (if appropriate). This is a complicated and labor-intensive process requiring biota sampling equipment, such as suction traps for air samples, corers for soil samples, and nets and traps for water and land samples (Southwood, 1978). The organisms in the soil cores must be extracted, and the organisms from all of the samples must be identified and counted. Single measurements of abundance obtained by one of these methods do not provide much interpretable information, but repeated measures over time can be used to estimate important population parameters, such as rates of recruitment, mortality, and population growth. However, estimation of population parameters does not necessarily identify the causes for many of the changes in populations, which may include interactions with other members of the community or with the physical environment. 61 ------- 2.4.2 Species Dominance The concept of species dominance relates to the importance of an individual species relative to others in the ecosystem. The dominant species is variously defined as the one with the greatest abundance, the greatest productivity, the largest individuals, the most occupied space, or the greatest impact on community dynamics (Clapham, 1973; Sheehan, 1984b). Shifts in dominance may be used as measures of pollutant effects on community structure. For example, in a bay stressed by pulp and paper effluent, yellow perch replaced white sucker as the dominant fish species in the area near the mill discharge. The shift was attributed to pollutant-induced decreases in the abundance of benthic invertebrates, the primary food source of the white sucker (Kelso, 1977). Dominance shifts were also noted in a study of macroinvertebrate communities of streams stressed by heavy metal pollution (Winner et al., 1975 cited in Sheehan, 1984b). In this particular case, reductions in the numbers of sensitive taxa leads to the dominance of metal-tolerant chironomid species. A number of abundance-based dominance indices have been devised to provide more rigorous methods of assessing this measure. These, however, have not been widely used in pollution studies (Sheehan, 1984b). Clapham (1973) suggests that there are a number of problems associated with the practical use of the dominance concept. For instance, the most commonly used measure of dominance is species abundance. However, a large-bodied species may be more influential than a far more abundant smaller species. Similar problems are associated with other measures of dominance. Dominance may 62 ------- reflect lack of competition, i.e., pollutant-sensitive organisms may die or migrate away from the study area. Invertebrates are very opportunistic in this way—with no competition, there are dramatic increases in numbers (Folmar, personal communication). Another consideration is that dominance within a single community may be best assessed using different measures at different trophic levels. Thus, a dominant plant species might be best defined based on ground cover, while a dominant carnivore species might be best defined based on biomass or net productivity. Although dominance implies a position of advantage in community interactions, it does not necessarily correlate with tolerance to chemical stress. Abundance (dominance) under pollution stress often depends on the opportunities and life-history strategy of the species (see section on Life Histories and Response to Contaminants). Nevertheless, in some ecosystems dominance can supply a clear simplifying concept that is useful for assessing pollutant impact (Clapham, 1973). 2.4.3 Indicator Species The indicator-species method of evaluating ecosystem stress has been used for most of this century. The method is based on the notion that the continued presence of certain species indicates acceptable environmental conditions, whereas their absence would indicate the lack of appropriate environmental conditions (Sheehan, 1984b). The indicator method is useful as an evaluative tool, and it can be employed to determine if damage to a system has already occurred. Sheehan (1984b) points out problems in the general definition of indicator species. For example, the absence of a species in a system does not always occur 63 ------- as a result of poor environmental conditions. Absence can also result from a lack of dispersal (e.g., changes in wind patterns resulting in altered seed distribution in the study area), sampling of an inappropriate seasonal lifestage (particularly for invertebrates), or biotic influences such as introduction of a new predator. It is clear that merely observing the presence or absence of a species does not prove a cause-and-effect relationship between the disturbance being investigated and the occurrence of the species. The indicator concept has been modified somewhat to recognize the ability of indicators to reflect the subtle as well as the gross effects of pollution (Sheehan, 1984b). Effects that occur prior to the total disappearance of a population from a system must be measured. To accomplish this, the species selected must evoke characteristic toxicological responses. One of the problems associated with the indicator method involves the selection of the indicator. Because there are no consistently reliable algorithms for selecting appropriate species, the determination of an indicator is subjective (Levin et al., 1984). The methodology for determining appropriate indicators is based on two different techniques. The first is subjective selection of the organism (Levin et al., 1984; Limburg et al., 1984), involving identification of species that possess characteristics such as: o Amenability to laboratory handling and testing; o Characteristic toxicological responses; o Economic, recreational, and/or ecological importance; o Universal distribution; 64 ------- o Susceptibility to the toxicants; and o Availability and abundance. The second basis for selecting species is mathematical. Statistical analysis can be performed based on the log-normal distribution of individuals among species. A method has been proposed that selects species of moderate abundance (16 to 63 individuals per species) to be appropriate as indicators (Gray and Pearson, 1982; Pearson et al., 1983). A problem with this method arises because it does not account for practical matters, such as exhibiting a toxicological response to a contaminant or being easy to handle in a laboratory. In conclusion, the indicator method can be a useful endpoint for ecological risk assessment in both aquatic and terrestrial systems; however, it must be used with caution. The selection of the particular species to monitor is the key. The indicator species must be chosen by scientists with a strong expertise concerning the particular environment being studied. 2.4.4 Keystone Species Keystone species provide a valuable endpoint in ecological risk assessment. They can be a means of bridging the gap between single-species toxicity testing and community level evaluations. Single-species tests are quick and easy, but their use can be limited if the species is a laboratory strain that may not represent the sensitivity of species present in the actual environment under examination. Community level evaluations are often complicated, costly, and time- consuming. However, evaluations of keystone species offer the advantages of a 65 ------- single-species approach, which yields insights as to whether a stress will affect community level endpoints. "A few species by their size, form, abundance, or activity may exclude or promote other macro species, may provide new niches of smaller organisms, or may modify the physical environment in which they live." (Lewis, 1978, cited in Bowmer et al., 1986). This is the currently held definition for keystone species. It is a more general definition than the original, which focused on predatory control of species densities. According to Paine (1966), keystone species are "predators that keep the population densities of the prey below levels where resources become limiting." This prevents a high level of competition for resources whereupon only a few species would become dominant and species diversity would decline. By definition, keystone species modify the community and/or the environment. Therefore, any disturbances to the keystone species would lead to a disturbance to the community and/or environment. The process of keystone species selection is more objective than selection of indicators. Furthermore, there has been experimental demonstration that keystone species strongly influence abundances of other species in the community. The keystone relationship is illustrated in a study of the starfish Pisaster ochraceus and mussel Mytilus californianus conducted by Paine (1974). Paine observed a natural population of mussels in a well-defined band. The principal predator for the mussel larvae was the starfish Pisaster ochraceus. Most of these predators were removed manually from the area on a monthly or twice- monthly basis for 5 years; monitoring of the area continued for 6 years. Under the conditions of a predator-free environment, the mussels excluded many other 66 ------- species that were in competition for the limited resource of space. Paine concluded that removal of the starfish resulted in "the local elimination of at least twenty-five macroscopic species that otherwise would have existed on or immediately associated with the primary space." Hixon and Brostoff (1983) discuss a keystone relationship that is different from the previous examples. In the Paine study, predation by the keystone species was found to maintain a high diversity of prey by preventing competitive exclusion among the prey. Hixon and Brostoff observed a relationship whereby the predator maintained a high diversity of prey by reducing the overall intensity of predation. The predator was the territorial damselfish. Damselfish are one of several herbivorous fish that feed on algae in tropical reefs. In the study, plates for growing algae were placed in three different areas: one set was placed outside the damselfish territories; one set was placed within some of the damselfish territories; and the final set was placed in fish-exclusion cages within the damselfish territories. Plates were sampled over the course of 1 year. Initially, colonization of the plates was comparable in the territories and the cages, although colonization was somewhat slower outside the territories. Throughout the experiment, biomass and diversity remained lowest where there were no damselfish to moderate the amount of grazing. Within the territories, diversity was highest on the caged plates at the 6-month mark. By the 1-year mark, diversity was greatest on the plates inside the territories, but biomass was higher on the caged plates. This study shows that damselfish can maintain a high diversity of prey by excluding other herbivorous fish from their territories, thus reducing the overall intensity of predation. This study supports the "intermediate disturbance hypothesis," which holds that diversity is maximized by an 67 ------- intermediate level of predation or other disturbance and declines at both low and high predation intensities (Connell, 1978). Many criticisms of this method are identical to criticisms of species-level approaches. However, keystone species do have a demonstrable influence on community structure. It is important to identify the types of ecosystems in which key species exert influence. In monitoring keystone species, it is important to note that systems have mechanisms, such as functional redundancy, that can minimize general structural or functional changes that could take place (Levin et al., 1984). For example, predators may switch their food choice from a toxicant- sensitive prey species to a more resistant one without changing the food web structure. Pollutant-sensitive species may be replaced by competitors without changing system biomass or productivity (Sheehan, 1984c). If compensatory mechanisms are not present, keystone species removal may have wide-ranging indirect effects on community structure that would not be evident from species- level approaches. For example, removal of sea otters from Alaskan islands led to increases in grazing sea urchins, decimation of kelp beds, and exposure of shorelines to greater wave damage (Estes and Palmisano, 1974). This example illustrates that interactions among species and between species and their environment must be understood before effects of contaminants can be predicted. Thus, these species-level effects may not adequately serve as indicators of whole ecosystem-level effects and should be used in combination with other measures in an ecosystem risk assessment. 68 ------- 2.4.5 Representative and Important Species Approach A multiple-species approach based on an examination of the Representative and Important Species (RIS) concept was presented by Limburg et al. (1984). This is a regulatory method that is used in the Clean Water Act and the National Environmental Policy Act. It combines several of the species-level effect measures (like keystone organisms or indicator species) with practicality. Limburg et al. (1984) recommended general criteria for selecting a RIS. The species should have some of the following characteristics: o Representative of the ecological community most exposed to an impact; o Direct commercial, recreational, or aesthetic value; o Critical to maintaining the integrity of an ecological community whether by structure (e.g., coral reefs) or function (e.g., primary producers, keystone predators); o Characteristically predominant in a habitat either by virtue of numbers or biomass (e.g., Spartina in a salt marsh); o Particularly sensitive to the perturbation of concern; usually the protection of the most sensitive species provides a "safety window" for the rest of the organisms in the area; and o Able to sequester chemical substances with known or suspected toxic effects, either because of physiology or life history. Several states have modified these criteria for particular locations. For example, Maryland Regulation 10.50.01.13 defines the following criteria for developing RIS lists for estuaries and other coastal zones: 69 ------- o Consider only those species normally present in the local salinity regime; o Determine the spatial and temporal distribution of resident and migratory species with respect to their various life stages; and o Select at least one fin fish, mollusc, and arthropod, and one other species for intensive study, using as selection criteria species abundance, commercial or recreational importance, and sensitivity per life stage to facility operations. In fresh water studies, one insect species shall also be selected. Additional R1S may be selected from a list of important taxa, which includes three species of waterfowl, one mammal, eight molluscs, one crab, three insects, and twenty-eight species of fish. This approach to evaluating ecosystem health is a valuable one. It can be made site-specific, and it solves some of the main criticisms of laboratory toxicological testing. Examination of a set of species provides some information on interspecies sensitivity. The RIS species selected are members of the natural system, and looking at species like keystone predators predicts some of the effects on ecological processes and structure (Levin et al., 1984). One objection to the uses to which the RIS concept has been applied is that little attention has been paid to microorganisms, (such as bacteria, fungi, microscopic algae, or microinvertebrates), which may be extremely important in ecosystem structure and function. 2.5 CONCLUSION Species-level methods are valuable indicators of ecosystem effects, but they do have drawbacks. They do not accurately account for changes in community or ecosystem structure that are caused by interactions between species or with other environmental factors. Also, changes in population levels are not necessarily correlated with adverse effects on the system due to buffers (replacement species) 70 ------- that minimize disruption of system function. Substitutions can occur without causing noticeable modification in ecosystem processes (Levin et al., 1984). Single-species measures provide methods of monitoring changes that are easier to define and measure than measurements at the community level. When used in conjunction with community-level endpoints, populations can provide necessary information on ecosystem health. For example, keystone species modify the community, and thus can provide insights into changes at the community level. Similarly the representative and important species approach provides a more integrative indicator of ecosystem stress. It selects species for examination based on practical, economic, and scientific considerations. Used properly, this approach (with particular emphasis on keystone species) is extremely valuable for ecological risk assessment. 2.6 SPECIES INTERACTIONS Species interactions are defined as relationships between two or more species (Odum, 1971). Because communities are composed of groups of interacting populations, impairment of the ability of a population to function normally is likely to lead to adverse impacts on the structure and function of the community. Two types of species interactions have been used extensively as endpoints for determining stress to an ecosystem; these are alterations to predator-prey interactions and competition (Sheehan, 1984a). 71 ------- 2.6.1 Predator-Prey Interactions Predator-prey interactions have been studied extensively, particularly with regard to pesticide effects. Complex behavioral patterns may be utilized in avoidance of predators. For example, schooling, surface swimming, formation of a circular aggregation, and hiding under natural vegetation are used by fathead minnows to avoid their principal predator, the largemouth bass (Sullivan and Atchison, 1978). Any alteration in prey vulnerability to predation may strongly affect population dynamics of both prey and predator. The usefulness of predator-prey interactions as a measure of ecosystem stress was illustrated in a study conducted by Tagatz (1976). He observed1 the effect of a pesticide (Mirex) on predation by pinfish on grass shrimp. Sublethal concentrations of Mirex significantly increased the susceptibility of grass shrimp to pinfish predation. This type of behavior alteration was also observed by Kania and O'Hara (1974). Sublethal mercury concentrations of 0.1, 0.05, and 0.01 ppm reduced the ability of mosquitofish to avoid predation by largemouth bass. In this situation, the effect on the mosquitofish had a potentially more serious effect on the ecosystem, as large amounts of mercury were ingested from contaminated prey by largemouth bass. It is likely that bass prey selectively upon the most heavily contaminated mosquitofish, as they are most likely to have reduced capacity to avoid capture. This would lead to increased bioaccumulation and biomagnification of mercury up the food chain. Sylvester (1972) examined the effects of thermal stress on predator avoidance in sockeye salmon. In a laboratory study, tanks containing coho salmon 72 ------- (predator) and sockeye salmon fry (prey) were monitored. Elevated temperatures significantly increased predation rates on sockeye salmon fry. Industrial processes using cooling water that is discharged into rivers and streams may thus have a high impact on predator-prey relations. Toxic stress can also have an effect on the selection of prey species. Farr (1978) exposed three aquatic species to sublethal doses of methyl parathion under laboratory conditions. The predator was gulf killifish and the prey were grass shrimp and sheepshead minnow. Only the grass shrimp exhibited any toxic effects from the pesticide. As a result, methyl parathion increased the preference of killifish for grass shrimp. Increased predation on grass shrimp could affect the ecosystem-level processes of decomposition and nutrient recycling. Grass shrimp are important in the conversion of organic detritus to forms more readily usable by other organisms. Replacement species may functionally compensate for the reduction of grass shrimp, but many of these species are also susceptible to methyl parathion poisoning. Another way that pollutants can alter community structure is through predator removal. Hurlbert (1975) noted that pesticide-induced density increases of prey populations can occur due to either removal of a predator or increased food supply. He cites several cases where rotenone and other pesticides eradicated fish populations. After removal of fish, reduced predation allowed many types of invertebrates and frogs to substantially increase in number. Similarly, pesticide-induced depletions of crustacean zooplankton have led to algal blooms and increases in abundance of herbivorous rotifers (both prey and competitors with crustaceans). 73 ------- Most predator-prey studies are performed using aquatic organisms, largely due to the ease of laboratory handling of fish, the intensity of predator-prey interactions in aquatic systems, and the preponderance of data on the behavioral effects of contaminants such as pesticides on fish. There are several methods of performing these types of studies. Microcosms are a logical choice if adequate time and space are available. A common laboratory technique used to detect the effects of stress on aquatic predator-prey interactions is described by Goodyear (1972). He recommends the use of two controls: one with' the toxicant present but without the predator, the other without the toxicant but with the presence of the predator. In general, the tank should functionally represent the natural habitat (e.g., some predator species require hiding areas from which to pounce, and prey generally require a refuge). Goodyear further notes that the distractive presence of investigators can affect the behavior of certain species; therefore, knowledge of the test species is critical prior to beginning an investigation. 2.6.2 Interspecies Competition A less commonly used measure of stress to the ecosystem is alteration of competition. Interspecific competition is defined as two or more species making simultaneous demands on one resource that is present in limited amounts. Interspecific competition may result in either the emergence of one dominant population or coexistence of multiple species (McNaughton and Wolf, 1979; Moriarty, 1983). Competition studies have not been used as extensively as predation studies (Sheehan, 1984a); however, they can serve as an endpoint. For example, toxic 74 ------- stress can alter competition intensity in a variety of ways. As discussed in the preceding section, Hurlbert (1975) noted that pesticide-induced mortality of predator populations can result in increased abundance of prey. This reduction of the predator population could actually have an adverse effect on competitively inferior prey species, as release from predation may lead to competitive exclusion. Even if local extinction of one or more populations does not occur, release from predation will likely alter community function and structure (Moriarity, 1983; Sheehan, 1984a). A possible example of community effects of reductions of a dominant competitor has been illustrated by Schmidt (1986). Coyotes are territorial and are the dominant member of a guild of large canids that prey on domestic livestock. Population reduction aimed at coyotes had serious effects on the structure of the guild. As the coyote population declined, other large carnivore populations increased. There is insufficient information about whether the shift in guild structure affected prey populations; however, it does appear that the reduction of coyotes has resulted in an increase in the competitively inferior smaller canids. Pollutant stress could alter interspecific competition by increasing the amount of a limiting resource. For example, phosphates discharged into aquatic systems often result in algal blooms (Reynolds and Walsby, 1975; Barica and Mur, 1979). Under these conditions, blue-green algae often replace other species, apparently by competitive elimination (Smith, 1983). 75 ------- 2.7 CONCLUSION These examples show how pollutant stress can affect interactions among populations. Examination of population interactions is an integral part of determining the effects of toxicants on ecosystems. Communities are groups of interacting populations, and serious changes in the functional capabilities of populations are likely to lead to adverse impacts on the community structure and function. These effects of pollutants on behavioral interactions among organisms cannot be determined from single species toxicity tests. The limitations of single species endpoints suggest that assessment of effects of pollutants must incorporate population, community, and ecosystem level ecotoxicological endpoints. 76 ------- CHAPTER 3 ECOSYSTEM-LEVEL ENDPOINTS 3.0 INTRODUCTION An ecosystem is a highly complex structure composed of a diversity of interacting biotic and abiotic components (Odum, 1971). The stress responses of the system reflect, in part, the integrated responses of these components. In an ecosystem risk assessment, effects on ecosystem components can be evaluated using organism- or population-level measures. Such measures may provide information on dominant, keystone, or indicator species; these may then serve as indicator signals of ecosystem-level stress. Organism-level measures can also be used to evaluate the status of economically or aesthetically important species (Sheehan, 1984a; Levin et al., 1984). Organism- or population-level effects, however, are not always indicative of ecosystem-level changes, because the system often responds in ways that are independent of the responses of specific components (Sheehan, 1984c). For example, individual populations can fluctuate or die off, yet the system, through a variety of internal buffering mechanisms-- predator flexibility, replacement species—can survive. Thus, monitoring of only organism- or population-level changes may not provide an adequate picture of the responses of the system. Ecosystem-level endpoints, such as primary production and nutrient cycling, provide important indicators of ecosystem stress. Although these endpoints have not received much attention as measures of pollution-induced stress, the literature is now expanding concerning the use of endpoint changes in both ecosystem 77 ------- function and structure. The ecosystem-level work done to date has not been as comprehensive as that done at the individual organism level. The research does, however, identify tested, measurable, ecosystem-level endpoints such as primary production, leaching of nutrients from soil, species richness, and similarity that provide quantitative data for risk assessment. 3.1 ECOSYSTEM FUNCTION Functional components of a system include the processes involved in the movement and transformation of chemicals and energy. These processes in turn provide the basic support for the system's structural components. As a result, the maintenance of population and community structure is dependent on the functional integrity of the ecosystem (Sheehan, 1984c). Societal goals for environmental protection (which are embodied in the statutes mandating ecological risk assessment) focus on protection of population and ecosystem structure, such as maintenance of species diversity or protection of endangered species. The dependence of these structural components on ecosystem functional processes make functional endpoints important elements of ecological risk assessment. The processes involved in chemical and energy flow are revealing endpoints for an ecosystem risk assessment. Measurable processes include primary production, photosynthetic rates, respiration, decomposition, nutrient levels, and nutrient leaching from soil. These processes are often highly sensitive to low levels of ecosystem stress and are thus valuable indicators of ecosystem stress. 78 ------- Despite their importance, functional endpoints have not been widely used in ecosystem pollution studies. There is, therefore, a lack of extensive baseline data establishing levels at which changes in ecosystem function will adversely affect ecosystem structure. More work using functional endpoints to monitor pollutant effects on ecosystems must be done before these endpoints can provide fully predictive measurements. In the following section, the current state of knowledge concerning these functional endpoints and methods of their measurement will be reviewed. 3.2 ENERGY FLOW 3.2.1 Primary Productivity Production, or the "process of energy input and storage in an ecosystem" (McNaughton and Wolf, 1979), is one of the most tested endpoints for evaluating ecosystem stress in both aquatic and terrestrial systems. Studies usually focus on either gross or net primary production. Gross primary production has been variously defined as the "energy fixed in photosynthesis" (Krebs, 1985) or the "total assimilation of organic matter by the plant community" (Odum, 1971). Net primary production is defined as gross primary production less the energy lost through respiration (McNaughton and Wolf, 1979). Net primary production is thus the total amount of photosynthetic input that could be available to other trophic levels (Beadle and Long, 1985). Photosynthesis is the mechanism through which solar energy is transformed into biomass (Sheehan, 1984c). It is, therefore, the ultimate source of all 79 ------- production in an ecosystem, and the various measures of photosynthesis—CO2 uptake, O2 production, 14C uptake, chlorophyll levels—can serve as indicators of primary production. As the energy base for the ecosystem, primary productivity is a fundamental factor in ecosystem health. It is critical to both autotrophic maintenance and heterotrophic development, and it determines the amount of living tissue that an ecosystem can support (Woodwell, 1970). Measurements of change in primary productivity provide important information about ecosystem health. One of the measures of the rate of photosynthesis is often used as an endpoint of ecosystem toxicity. For example, SO2 and a number of other phytotoxic pollutants act to suppress photosynthesis rates. This suppression has been investigated in lichens and a variety of other producers. In one such study conducted by Bennett and Hill (1974), alterations in the rate of photosynthesis were measured as changes in CO2 uptake. These investigators observed a reversible suppression in photosynthesis in crop plants chronically exposed to pollutants such as CI2, O3, SO2, and NO2. In another study, Carlson (1979) observed the effects of SO2 and O3 on the rate of photosynthesis. Measured as CO2 uptake, photosynthesis was significantly reduced in exposed maple and ash leaves. Other studies using photosynthesis as an endpoint are described in Sheehan (1984c). Although primary productivity is clearly an important measure of ecosystem function, its usefulness as an endpoint for ecological risk assessment is dependent on the nature of both the pollutant and ecosystem being studied. Effects of 80 ------- different pollutants on primary productivity in different ecosystems are highly variable. SO2 and acid rain, for example, consistently reduce productivity in terrestrial systems, and yet they may either decrease or increase productivity in different aquatic systems. Similarly, low levels of heavy metals have been shown to produce rapid, readily measurable changes in net primary productivity in aquatic systems, while intermediate levels of metal pollution produce potential, but difficult to measure, changes in productivity of forest ecosystems (Sheehan, 1984c). Primary productivity therefore serves as an appropriate measure of ecosystem stress in some situations but not others. When productivity is used as a measure, changes should be evaluated in terms of the status of the system being monitored. Natural systems normally undergo fluctuations in productivity as a result of changes in grazing and in available resources such as CO2, nitrogen, phosphorous, and light. If resources flow into and out of the system at the same rate, the system is considered to be in a steady state. A system may naturally shift from one steady state, and therefore one level of productivity, to another, depending on resource availability (Odum, 1971). Thus, any measure of pollutant effects on a system's productivity can only be fully assessed with prior knowledge of the steady-state status of the system. In the absence of information on steady-state productivity levels, it may be possible to estimate them by using appropriate models or by comparison to similar, unimpacted systems. If ecosystems are not in a steady state, however, even prior knowledge of system productivity will not be sufficient to assess pollutant impact (Hurlbert, 1984; Stewart-Oaten et al., 1986). 81 ------- Productivity changes must also be assessed in terms of the seasonal and successional stages of an ecosystem. Ecosystems in different stages of seasonal development demonstrate different sensitivities to pollutants. For example, terrestrial systems tend to be more sensitive to pollutant-induced stress during the season when plants are in the germinal stages of growth (Sheehan, 1984c). Similarly, pollutant-induced changes in nutrient flux are most likely to affect plant productivity in late-succession terrestrial systems, where nutrient cycles are stabilized (Vitousek and Reiners, 1975, cited in Sheehan, 1984c). Another factor that must be evaluated is the transience of productivity changes. Long-term changes of large magnitude are clearly suggestive of ecosystem damage. Transient changes, however, may be suggestive of ecosystem stability, when in fact other kinds of functional and structural changes are taking place. Studies on the effects of hydrocarbons in lake systems have shown that while algal production drops initially in response to pollutants, the algae show rapid recovery and growth, returning to normal productivity levels, even as nutrient cycles and other parts of the system are breaking down (Hobbie, personal communication). Evidence from eutrophication studies suggests that recovery of algal production may involve a change in species composition, which can then affect other components of the ecosystem. Eutrophication commonly causes increases in primary production that are accompanied by a shift in algal species composition to dominance by blue-green algae (Reynolds and Walsby, 1975; Barica and Mur, 1979). Because blue-green algae are often toxic or of poor nutritional value to algal 82 ------- consumers (Fulton and Paerl, 1987), the transfer of primary production to higher trophic levels may be reduced following eutrophication. Pesticides, such as DDT, also induce transient changes in productivity. Pesticides cause an initial increase in crop productivity due to a reduction in pest species numbers; after a number of years, however, productivity drops, due to pesticide effects on other components of the system (Sheehan, 1984c). Clearly, changes in productivity must be evaluated with knowledge of changes that could be taking place in other parts of the system. 3.2.2 Respiration Respiration, defined as "any energy-yielding biotic oxidation" (Odum, 1971), provides a measure of the rate at which organic matter is oxidized. As such, it serves as an indicator of community metabolism (Cooper and Copeland, 1973). For example, soil respiration can be used as an indicator of the rate of decomposition of organic matter (Garten et al., 1985). In addition, respiration can be used with measures of net production to determine gross production. Methods for measuring respiration include determinations of nighttime C>2 uptake and CO2 release. A number of studies illustrate the use of respiration as an independent endpoint. In one study, conducted by Baddeley et al. (1973), respiration rates, measured as O2 uptake, decreased when lichens were exposed to SO2. Cooper and Copeland (1973) also observed a depression in respiration, measured as CO2 production, in estuarine ecosystems stressed with industrial effluent. 83 ------- 3.2.3 Photosynthesis/Respiration Ratio The photosynthesis/respiration (P/R) ratio provides an integrative measure of ecosystem metabolism. It was originally proposed by Odum (1971) as a method for classifying ecosystems and has since developed into a useful method of measuring ecosystem stress. Odum suggested that, in autotrophic systems where energy is obtained principally from the sun (McNaughton and Wolf, 1979), the rate of photosynthesis tends to exceed the rate of community respiration. In such a system, biomass tends to increase and therefore the P/R ratio is greater than 1. In heterotrophic systems, energy is obtained primarily from preformed sources of organic energy (McNaughton and Wolf, 1979) and the rate of respiration tends to exceed the rate of photosynthesis, yielding a P/R ratio of less than 1. During the process of succession, the P/R ratios of both autotrophic and heterotrophic systems approach 1, and energy cost is in balance with energy fixation (Odum, 1971). Studies have shown that the P/R ratio can be a sensitive indicator of ecosystem stress. In mature systems, the presence of toxicants and other stressors will cause deviations from the predicted P/R ratio of 1 (Giddings and Eddlemon, 1978). Such changes have been observed in a number of micro- ecosystems. In a study conducted by Maki and Johnson (1976), changes in measurement of O2 levels were used to determine primary production and community respiration in 3-trifluoromethyl-r-nitrophenol (TFM) poisoned model stream communities. Exposure to TFM was found to cause a significant depression in gross primary production and P/R ratios. The P/R ratios were found to be very sensitive indicators of TFM influence on the community. 84 ------- Similarly, Giddings and Eddlemon (1978) found that P/R ratios, determined from changes in 02 concentrations, declined in arsenic-stressed pond microcosms. The ratio decline was correlated with the arsenic concentrations of the individual systems. This study indicated that the P/R ratio could be useful when monitoring changes in ecosystems at varying distances from the pollutant source. Although it provides a useful endpoint, the P/R ratio should be used with some caution. In situations where a toxic chemical reduces both primary production and respiration, the P/R ratio may exhibit little change, even though both the total energy base for the ecosystem and biomass have been reduced (Sheehan, personal communication). 3.2.4 Methods of Measurement of Primary Production A variety of methods have been used to measure primary production. These methods all monitor some aspect of the energy-transforming photosynthetic process, and the general photosynthetic equation. 6C02 + 12H20 673 kilocalories C6H1206 + 602 + 6H20 can be used with these methods to calculate production. Measures of gross production additionally take into account energy lost through respiration. In making net and gross productivity determinations, it has generally been assumed that net production can be determined during the light period of the diel cycle, while respiration rates for calculating gross production can be determined 85 ------- al., 1963). Thus, using nighttime respiration rates to estimate daytime, i.e., total, plant community respiration, will probably lead to an underestimate of gross primary production. Because actual quantitative measures of daytime respiratory rates are lacking, however, daytime respiration is normally assumed to be equal to nighttime respiration (Cooper and Copeland, 1973). *4C UPTAKE: uptake is the most commonly used measure of primary production in aquatic systems. It is currently considered the most sensitive method of measuring aquatic primary production in ponds, lakes, and oceans. This method was introduced by Steeman Nielson in 1952, and it provides an estimate of the amount of carbon taken up by a system's plants and converted through photosynthesis to organic compounds (Peterson, 1980). A simple light-dark bottle technique is used to measure ^CC>2 uptake in aquatic systems. ^CC>2 may also be used to measure production in terrestrial systems by introducing it into a transparent chamber containing representative plants. The plants are then harvested, and whole-plant levels are determined (Krebs, 1985). Problems with methodology have been detailed by a number of investigators. Peterson (1980) suggests that uptake normally provides no direct estimate of respiration, as dark bottle uptake measures both active dark uptake of CO2 and abiotic formation of labelled particulate carbon. Thus, this method does not measure either gross or net production. Other problems associated with this technique include: 0 Potential contamination of samples with metals or other toxic materials from containers used during collection and incubation. 0 Potential loss of during inoculation, incubation, and sample 86 ------- o Potential contamination of samples with metals or other toxic materials from containers used during collection and incubation. o Potential loss of '^C during inoculation, incubation, and sample preparation. o Uptake and release of 14C by bacteria and zooplankton in natural plankton communities. o Underestimation of primary production when compared to other methods in both oligotrophic and certain highly productive ecosystems (Peterson, 1980; Bemer et al., 1986; Fitzwater et al., 1982; Sakamoto et al., 1984). The method has been carefully analyzed and refined over the years and remains the best, most sensitive method currently available for measuring primary production in lakes, ponds, and oceans (Hobbie, personal communication). It is evident, however, that this method requires careful monitoring of sampling and counting techniques. O; measurements: Another way to use primary production as an endpoint in an ecological risk assessment is through measurement of O2 production. This is a widely used method of measuring primary production in marine and freshwater environments. Light-dark bottle techniques, which were pioneered by Gaarder and Gran (1927), involve a setup similar to that already described for measuring uptake. Oxygen production in incubated samples is measured using an oxygen meter, or by chemical analysis. Data obtained from light bottles provide a measure of net community production (Odum, 1971). Data obtained from dark bottles represent the normal respiratory consumption of oxygen and thus can be added to data from light bottles to provide a measure of gross production (Kormondy, 1969), which cannot be measured using (Peterson, 1980). 87 ------- Oxygen production has been used in a variety of ecosystem pollution studies as an indication of changes in primary production. In one such study conducted by Patil et al. (1985), the effects of DDT contamination in a pond ecosystem were monitored. Primary production and respiration were measured, using the light and dark bottle method, as changes in Oj. It was found that DDT caused decreases in both gross production and net primary production, as well as increases in respiration. Oxygen production provides a good measure of aquatic primary production because it is simple to monitor and it produces data on both production and respiration (Peterson, 1980). While 14C is currently considered the most sensitive measure, recent studies suggest that oxygen may provide a more accurate measure of primary productivity than 14C in certain highly productive ecosystems, e.g., hypertrophic ponds (Berner et al., 1986). Oxygen measurements are also useful for estimating production in flowing water systems (Hobbie, personal communication). COj assimilation: Monitoring of CO2 assimilation using an infrared gas analyzer is a useful method of measuring photosynthesis, respiration, and primary production in terrestrial ecosystems. Measurements can be made with great speed and precision to + 1 ppm, a level of accuracy not achievable for field measures of O2 (Woodwell and Botkin, 1970). The ease of measurement has made this method one of the most widely used measures of terrestrial production. CO2 assimilation is determined by enclosing all or part of an ecosystem in an airtight or monitored, open-flow, light-transmitting chamber, and measuring 88 ------- CC>2 levels with an infrared gas analyzer. Daytime uptake of CO2 serves as an indicator of photosynthetic uptake minus respiratory release of CO2 and provides a measure of net community production, while nighttime production of CO2 is a measure of respiration. These can be added together to provide a measure of gross primary production (Brinson, et al., 1981; Carlson, 1979; Kormondy, 1969). CO2 has also been measured in open systems using an aerodynamic method developed by Huber (Odum, 1971). This method involves measurement of the vertical gradient of CO2 from the ground up. Vertical gradients result from uptake of CO2 by vegetation and release of CO2 through soil and litter respiration. Total community respiration can be estimated from nighttime gradients. In using this method, however, account must be taken of incidental factors such as mass air movements and soil CO2 evolution that could alter CO2 concentrations. Aerodynamic measurement methods have been used on crops, grasslands, and even forest communities (Odum, 1971). Refinements in monitoring techniques may make this an extremely useful way of measuring CO2 in the future. CO2 assimilation can be used to monitor primary production in a variety of terrestrial systems following toxicant insult. For example, Bennett and Hill (1974) measured CO2 uptake rates of barley and oat canopies, following 2-hour fumigations with environmental pollutants. Similarly, Carlson (1979) used the rate of CO2 exchange to measure photosynthetic rates in maple, ash, and oak branches exposed in an experimental chamber to sulphur dioxide and ozone. 89 ------- Brinson et al. (1981) cite several problems associated with measurement of primary production by CO2 assimilation, including: the potential for changes in temperature and air flow to alter rates of CO2 assimilation; the high cost of using sophisticated monitoring equipment to obtain long-term measures in the field; and the potential for error in large, complex systems where parts must be used to estimate whole ecosystem function. The latter problem, however, becomes insignificant in systems such as nonforested wetlands, where the whole system can be enclosed. dH method: Because the pH of water is a function of dissolved CO2, pH changes in aquatic ecosystems can be used as an index of production. To use pH as a measure, it is necessary to create a calibration curve that incorporates the buffering capacity of the system. Methods of preparing a calibration curve are described in Beyers (1963). pH measurements are useful in that they do not involve any disturbance to the community, which could potentially skew any conclusions regarding environmental hazards. This method has been used primarily to determine gross community production in laboratory microecosystems (Odum, 1971). Another advantage of using the pH method is that, in contrast to O2 and CO2, the H+ ions and OH- ions that largely govern the pH in natural waters are not readily lost to the atmosphere. pH measurements, additionally, are useful for pollutant studies because they are not generally influenced by the presence of foreign substances (Cooper and Copeland, 1973). 90 ------- Gorden et al. (1969) point out several problems associated with pH-derived measures of community production. They note that excretion of organic acids and ammonium by bacteria and algae can result in pH changes. Furthermore, heterotrophic CO2 uptake may affect measures of both photosynthetic uptake and respiratory production. Cooper and Copeland (1973), however, suggest that the error caused by heterotrophic uptake is likely to be minor. Another potential problem is that acidity produced by nitrification can reduce pH (Fenchel and Blackburn, 1979). Chlorophyll: A determination of chlorophyll levels or chlorophyll fluorescence is primarily used as a measure of plant biomass, but is sometimes used as an index of primary production or photosynthesis. Odum (1971) suggests that chlorophyll concentrations can be used to estimate gross production when both the assimilation ratio (rate of O2 production per unit biomass of chlorophyll) and the available light levels are known. The assumption of a constant assimilation ratio may be false when production is limited by factors other than light, such as nutrient availability, toxicity, or other stresses. Measurement of chlorophyll levels requires extraction of chlorophyll from phytoplankton using an organic solvent and determination of concentrations using a spectrophotometer (Kormondy, 1969). Chlorophyll extraction has proved to be a relatively speedy and inexpensive alternative to '^C and O2 methods for measuring production in large bodies of water. In a study conducted by Ryther and Yentsch (1957, as cited in Odum, 1971), production rates of marine phyto- plankton proved to be similar when the chlorophyll and O2 light-dark bottle methods were compared. 91 ------- Biomass change: Biomass can be defined as the total weight of living organisms or the total stored energy content of an ecosystem (McNaughton and Wolf, 1979). Estimates of biomass change over time is useful as an endpoint for ecological risk assessment. Historically one of the oldest methods of measuring production, the harvest method of biomass measurement is the most commonly used. The harvest method, in which plant material is removed and weighed at periodic intervals, can be employed in areas where herbivores are insignificant and the ecosystem does not approach a steady-state condition. Typically, farmers use this method when measuring production of cultivated crops. It is also useful in early successional systems, where small numbers of annual plants predominate and little herbivore consumption takes place before plants are fully grown. In these situations, a determination of peak standing crop can represent net production. Although some investigators consider biomass production a measure of primary production, technically it is a measure of net community production, as it does not account for biomass loss through herbivory and plant respiration (Odum, 1971). In diverse environments where all species do not mature simultaneously, peak standing crop does not account for plant mortality prior to the time of measurement or plant growth following the time of measurement, nor does it account for different species attaining peak standing crop at different times. To deal with these problems, Wiegert and Evans (1964) developed a method of determining net production by combined measurements of the annual disappearance of dead material from both green and dead standing crops. Lomnicki et al. (1968) 92 ------- simplified the Wiegert and Evans method by using the collection of green and dead material only. It was subsequently shown that, in a grassland community, this modified method produces results comparable to those of the original method (Lomnicki et al., 1968). In forests, biomass change can be used to determine production by trees, shrubs, and ground flora. Biomass change is determined by summing the mass of a number of components, including flowers, fruit, buds, and leaves. Their mass can be estimated using traps to collect litter fall, measuring branch mass by destructive sampling (i.e., selective tree-falling), and measuring root biomass through root sampling. Repeated measures can then be used to determine biomass change (Newbould, 1967). The measurement of biomass offers a relatively simple approach to the determination of primary production. There are, however, a number of problems associated with this form of measurement that hinder its utility for use in an ecological risk assessment. These problems (summarized in Beadle and Long, 1985) include: o A tendency by investigators to not consider inorganic (ash) content when determining biomass; o Difficulties in identifying living material in senescent leaves and in separating out live and dead roots; o Difficulty in obtaining representative samples in heterogeneous plant communities; o Errors in the use of regression estimates to determine forest biomass based on small samples; o Difficulties in measuring mass losses due to grazing; and 93 ------- o Failure to account for below-ground biomass, including both roots and root exudates. This factor may pose a particular problem in wetland communities with large amounts of underground biomass. Despite accuracy problems, the simplicity of the biomass measure has made it a standard technique for measuring production, particularly in terrestrial ecosystems. Biomass is also used as a measure of community structure. This use is described in a later section. 3.3 BIOGEOCHEMICAL CYCLING Biogeochemical cycles are defined as "the patterns of flow of chemical elements through biological organisms and their geological (abiotic) environment" (Kormondy, 1969). Elements are moved through the action of biochemical, chemical, and physical forces. During the process, the chemicals undergo a series of physical and chemical transformations making them more or less available to the system (Atlas and Bartha, 1981). The cyclic movement of chemical elements through an ecosystem is fundamental to the health of the system. Organisms are dependent on the constant availability of some 20 elements, which are required for all life processes. The major elemental components of living organisms include C, H, O, N, P, and S; minor and trace elements include Mg, K, Na, Fe, Mn, Ca, Al, B, Co, Cr, Cu, Mo, Ni, Se, Si, V, and Zn. These elements are cycled to varying extents through the biosphere. Toxicant effects on the availability of these elements in an ecosystem ultimately translate into effects on many other aspects of ecosystem 94 ------- structure and function. Various aspects of nutrient cycles thus provide important endpoints for use in an ecological risk assessment. Nutrient cycling can be monitored using a variety of measures. Information can be obtained on nutrient concentrations in soil and living organisms, and on levels of nutrient leaching in soil. Serious changes in any of these measures can provide good evidence of ecosystem stress. Measures of excessive nutrient export in a mature, normally nutrient-conservative terrestrial system, for example, are highly suggestive of a serious breakdown in nutrient cycling processes (Sheehan, 1984c). 3.3.1 Nutrient Analysis Nutrient analyses must be performed with an understanding of the numerous processes that affect nutrient concentrations in different parts of a system. In aquatic systems, the molecular forms of carbon, nitrogen, phosphorus, and sulfur are continuously being altered by physical and biological processes. These processes include continental erosion and dissolution of gases in atmospheric water droplets. The nutrient concentrations in aquatic systems are further affected by factors such as inputs from watersheds and from groundwater, chemical composition of rivers, stratification and vertical mixing of the water column, photosynthesis in the near-surface euphotic zone, settlement of organic matter to deep waters, and the microbial decomposition of organic matter to its component nutrients. 95 ------- In terrestrial systems, nutrient concentrations may also be affected by processes such as volcanic eruption, glacier melting, aerial transport of soil and seeds, precipitation, erosion and runoff, germination and growth of seeds into plants, photosynthesis, nitrogen fixation, chemical and physical decomposition of dead plants and organic matter by soil microorganisms, and microbial storage of nutrients. Nutrients can be analyzed using standardized laboratory techniques for element analysis. For example, ammonia-nitrogen can be analyzed by nesslerization, phenate, titration, or ammonia-selective electrode methods; nitrate- nitrogen can be analyzed by cadmium reduction or chromotropic acid procedures; nitrite-nitrogen can be analyzed by the formation of a reddish-purple azo dye; and sulfate can be analyzed by gravimetric or turbidimetric methods (American Public Health Assoc., 1985). These standard methods are well defined, inexpensive, and readily performed in an established laboratory. They may, however, lack sensitivity for detection of trace elements. Atomic absorption spectroscopy is another commonly used method for nutrient analysis. This method is highly sensitive and has the ability to measure elements in the ppb range. The substance to be analyzed is converted to an atomic vapor and absorbance at a selected wavelength is measured and compared with that of a reference substance. The usefulness of the method for ecological risk assessment, however, is limited by its cost and technical complexity. 96 ------- An autoanalyzer system is a fast and accurate method for measuring levels of nutrients such as nitrate, phosphate, and silicate. Like atomic absorption spectroscopy, however, autoanalysis tends to be expensive (Warfar et al., 1983). Nutrient levels clearly provide useful endpoints for monitoring nutrient cycles. They are readily measurable, using well standardized and often highly sensitive methodologies. Nutrient levels alone, however, are sometimes inadequate measures of pollutant-induced changes in nutrient cycles. During productive periods in aquatic systems, for example, macronutrient levels are often low and recycling processes are well developed. Disruptions to nutrient cycles in this case will only be obvious as changes in nutrient flux rates, rates of transformation, and associated biological activities (Sheehan, personal communication). 3.3.2 Chlorophyll Content Analysis of chlorophyll content is a rapid method for estimating the biomass of photosynthetic algae in aquatic systems. Because algal productivity, and thus biomass, increases in nutrient-rich systems, chlorophyll measurements provide indicators of nutrient conditions in aquatic ecosystems. Methods for analysis of chlorophyll are described in the section on primary productivity. 3.3.3 Leaching Leaching is the process through which percolating water removes soluble substances from the soil. The analysis of soil leachate provides a well-tested, 97 ------- sensitive method of monitoring changes in nutrient cycling because it measures the rate at which nutrients leave the system. Measurements of the rate of leaching of a chemical indicate the length of time that a chemical is retained in the topsoil, where it is most subject to degradation, dissipation, or plant uptake (Hamaker, 1975). In determining the rate of nutrient disappearance, it is important to consider not only leaching, but also the rate of degradation. There are few studies in the literature, however, that deal with the problem of simultaneous degradation and leaching (King and McCarty, 1968). The rate of the nutrient loss through leaching is a natural candidate for an endpoint for an ecosystem risk assessment. Changes in the rate of leaching can reflect the breakdown of any of a number of nutrient cycling processes. In addition, a leachate study can detect small changes in nutrient content that are not measurable using other methods. It can thus provide a sensitive, early measure of detrimental changes to systems. It is a particularly useful endpoint in situations where it is difficult to predict the exact mode of action of a new pollutant (O'Neill et al., 1977). Soil leaching analysis is generally performed using a lysimeter. Lysimeters were first used by De LaHire in 1703 and Ebermayer in 1897. Since that time, they have been modified and improved, and are employed to study water movements in agricultural and forest systems (Cole et al., 1961). In pollution impact studies, the lysimeter may be used to assess nutrient leaching, as well as the movement of toxic chemicals through the soil into the groundwater. 98 ------- Leachate studies can be performed in the laboratory by adding water, at seasonal rainfall averages, to soil samples and analyzing the leachate for metals. Although this procedure may oversimplify the leaching process, neglecting many factors that affect leaching rates in the natural habitat, it can be useful for making comparisons of leaching rates from different soil types. Numerous studies illustrate the usefulness of soil leachate measures in studying pollutant effects on ecosystems. In one study, Nieboer et al. (1980, cited in Sheehan, 1984c) found that K.+ efflux from lichens correlated with SO2 levels and the acidity of precipitation. Overrein (1972, cited in Sheehan, 1984c) found a similar increase in leaching of calcium from forest soils exposed to acid rain. In 1961, Cole et al. noted that a modified lysimeter proved to be a sensitive tool when used to monitor the effects of fertilizer containing potassium (added as KC1) and nitrogen (added as (NH^SC^), on ion movement in the soil. Data from the lysimeter indicated that both potassium and nitrogen acted as mass ions, stimulating increased leaching of Ca and other elements. In a soil leachate study conducted by O'Neill et al. (1977), it was found that arsenic and lead caused significant increases in losses of calcium and nitrate from soil microcosms, even though there were no measurable changes in population or community parameters. These results led O'Neill et al. to conclude that soil nutrient leaching can provide a sensitive, early-warning indicator of environmental stress. 99 ------- 3.3.4 Determination of a Nutrient Budget Nutrient budget determination is a "measurement of the total input minus total output of nutrients in a system." The nutrient budget provides a good measure of the dynamics of ecosystem function. It includes information on many nutrients, thus providing a comprehensive overview of ecosystem function. It can further provide information on a variety of channels of nutrient movement including sources of inflow (precipitation, dust, and weathering), recycling, and outflow (runoff, erosion, leaching, and volatilization). This information is critical in assessing the retention or loss of nutrients that can potentially be stored in the biomass of the ecosystem for long periods of time (Kormondy, 1969). A complete nutrient budget study includes determinations of standing stocks of nutrients, rates of movement of essential nutrients between different ecosystem compartments, and inputs and outflows of nutrients through the system (Westman, 1985). As part of this study, measurements are made of meteorologic inputs and geologic outputs. Meteorologic input to the ecosystem is measured through chemical analysis of precipitation and calculated for each element in terms of mass flux per unit area. Geological output can be measured through analysis of chemicals dissolved in water. This is combined with data on the flow rate and volume of drainage water to obtain an estimate of nutrient loss as grams of element lost per hectare watershed (Bormann and Likens, 1970). Likens et al. (1970) were able to obtain a substantial amount of information on nutrient cycling when they used the nutrient-budget method to study the effects of forest cutting and herbicide treatment in the Hubbard Brook watershed- 100 ------- ecosystem. Comparison of the streamwater nutrient concentrations in the deforested and undisturbed areas of the watershed showed that budgetary net losses from deforested watershed were up to 20 times higher than in the adjacent, undisturbed watershed. Thus, it appeared that alterations of the nutrient cycles resulted in an increased loss of dissolved nutrients from the deforested ecosystem. Although changes in the nutrient budget would seem to be excellent endpoints for ecosystem risk assessment, in general the costs and time required for such a study are prohibitive. A study of the effects of a contaminant on nutrient budgets requires continuous long-term measurements of both the undisturbed system and stressed ecosystem. In addition, measurements of nutrient budgets are hampered by the existence of nearly unmeasurable pathways of nutrient losses such as deep seepage, groundwater circulation, and wind (Bormann and Likens, 1967), rendering nutrient budget determination infeasible in many systems. Alternative approaches for nutrient budget estimation may include microcosm experiments or restricting attention to certain transformations in nutrient cycles. 3.3.5 Nutrient Cycling: The Nitrogen Cycle Various phases of nutrient cycles provide readily measurable endpoints for ecosystem risk assessment. Portions of the nitrogen cycle, in particular, have frequently been used as endpoints of toxicity. Measurements of alterations in the processing of nitrogen can serve as sensitive indicators of environmental change in ecosystems (Sheehan, 1984c; Bollag and Barabasz, 1979). As a critical constituent of living biomass, and a frequently limiting element in terrestrial and 101 ------- aquatic ecosystems, nitrogen is an appropriate endpoint for measurement (Westman, 1985; Cook, 1984). The nitrogen cycle includes four major steps: nitrogen fixation, mineralization, nitrification, and denitrification (Kormondy, 1969). Though high concentrations of nitrogen are present in the atmosphere, most organisms are unable to use atmospheric nitrogen. Nitrogen becomes available to organisms after it has been fixed into nitrate or other inorganic nitrogen compounds that can be utilized by other organisms (McNaughton and Wolf, 1979). In aquatic systems, nitrogen fixation is carried out by blue-green algae and bacteria. In terrestrial systems bacteria are the primary nitrogen fixers. Nitrogen fixation takes place through reduction of dinitrogen (N2), and the resulting ammonia may be assimilated by algae and larger aquatic or terrestrial plants. The reduced inorganic nitrogen is assimilated by various organisms into proteins and nucleic acids and is eventually released as metabolic waste or as protoplasm in dead organisms. Many heterotrophic bacteria, actinomycetes, and fungi convert the organic nitrogen and release it as ammonia. The process is referred to as mineralization or ammonification. Ammonia and ammonium salts are converted by nitrifying bacteria to nitrite, which in turn is converted to nitrate in a pH-dependent process known as nitrification. Both the nitrate and ammonia may be rapidly taken up by plants or microbes. Under certain conditions, particularly anaerobic ones, nitrate is reduced by denitrifying bacteria to nitrite, ammonia, and dinitrogen. Denitrification occurs 102 ------- primarily under anaerobic conditions in the presence of large amounts of decaying organic matter. Dinitrogen gas produced during denitrification is released back into the atmosphere (Kormondy, 1969). 3.3.6 Methods for Analysis The nitrogen cycle can be monitored through four general processes: soil nitrogen availability, soil denitrification, nitrification, and nitrogen fixation. Soil nitrogen availability: There is no generally accepted method for measuring available nitrogen in soil. Problems arise because quantities of available nitrogen are difficult to determine, in part because 97 to 99 percent of soil nitrogen is present in organic forms that are not directly available to plants. Only after mineralization occurs can nitrogen be used. The amount of nitrogen that would be mineralized is difficult to predict, as mineralization depends on numerous environmental factors. In addition, once nitrogen has been mineralized, it is subject to losses through leaching, denitrification, and microbial conversion to organic forms (Goh and Haynes, 1986). Two techniques commonly used for measuring soil nitrogen availability are the incubation and the chemical indices methods. A number of different incubation procedures have been proposed. These include short-term (1-6 week) aerobic incubations and anaerobic incubations. Proposed procedures have been reviewed by Keeney (1982), Stanford (1982), Sahrawat (1983), and Goh and Haynes (1986). 103 ------- Incubation methods are usually unsuitable for routine soil testing because they are both space and time consuming. In addition, it is likely that they provide only relative estimates of available soil nitrogen. Other problems associated with the use of incubation methods are discussed by Keeney (1982) and Goh and Haynes (1986). They include water loss, improper aeration, and improper sample pretreatment. These can be alleviated through careful handling of samples. Nonetheless, measurements made using incubation methods are highly correlated with nitrogen uptake by plants and have been used in a number of studies. In one such study, anaerobic incubation was used by Shumway and Atkinson (1977, in Keeney, 1980) to measure NH4-N production. The results indicated that NH4-N production was correlated with increased diameter growths in Douglas fir trees. Chemical indices have been proposed as speedy, precise, and convenient alternatives to biological incubation procedures. Chemical indices involve measurements of ammonium or total nitrogen from treated soil samples. Chemical index methods have been reviewed by Keeney (1982), Stanford (1982), Sahrawat (1983), and Goh and Haynes (1986). Soil denitrification: Soil denitrification is generally analyzed by measuring the products and conditions of anaerobic incubation of soil samples. Denitrification has been measured under field conditions by disappearance of applied (as 15^ and ^N20), and by measuring N2O fluxes in the presence of an N2O reduction inhibitor, acetylene (Rolston et al., 1982; Colbourn et al., 1984; Haynes and Sherlock, 1986). 104 ------- Soil denitrification was measured in a study assessing the effects of pollutant metals performed by Bollag and Barabasz (1979). The study, which used laboratory incubations, involved measurements of denitrification in liquid growth medium, autoclaved soil, and native soil. The study showed that, depending on the medium used, cadmium, copper, and zinc all inhibited denitrification to varying extents. Nitrification: Nitrification is the process by which ammonium is oxidized via nitrite to nitrate. Inhibition of nitrification by toxic pollutants can lead to accumulation of toxic levels of ammonia and nitrites (Liang and Tabatabai, 1978). Various methods have been used to measure nitrification rates. One commonly used method involves the incubation of samples in the presence of chlorate, which inhibits nitrite oxidation, the second step in nitrification. Under these conditions, the rate of nitrite accumulation provides an estimate of nitrifier activity (Belser and Mays, 1982). However, this method must be interpreted cautiously, as chlorate can also inhibit oxidation of ammonia in some cases (Hynes and Knowles, 1983). Other inhibitors used in measurement of rates of nitrification include nitrapyrin and allylthiourea (Hall, 1984). In another method, a tracer amount of '^N-nitrate is added to the sample and nitrification is measured as the rate of dilution of '^N-nitrate during incubation. Nitrification rates have also been estimated as changes in nitrate and ammonia concentrations during long-term incubations (Koike and Hattori, 1978). Nitrogen fixation: Nitrogen fixation in terrestrial systems can be studied through the analysis of symbiotic legumes. Fixation processes are sensitive to 105 ------- nonlethal doses of pollutants, which impair the survival of organisms dependent on nitrogen fixation. Methods of analyzing legume fixation performance include measures of ethylene formation and nodulation performance, as well as nitrogen mass balance and isotope analyses. The ethylene formation (acetylene reduction) method is based on the fact that the nitrogenase enzyme involved in nitrogen fixation also reduces acetylene to ethylene. The amount of ethylene produced during incubation can be easily measured by gas chromatography. The ethylene production rate is simple, sensitive, and inexpensive to measure, and is thus frequently used as a measure of nitrogen fixation. The nodulation performance method involves a quantification of nitrogen- fixing root nodules on leguminous plants. Simple nodule quantification, however, is not an adequate measure of nitrogen fixation, as not all nodules are active. Further analysis must be performed to test for nitrogenase activity in individual nodules. It is possible to quantitatively estimate rates and amounts of nitrogen fixed by nitrogen mass balance techniques. Kjeldahl digestion of the sample will provide a measure of total organic nitrogen (American Public Health Assoc., 1985). To use mass balance as a measure of nitrogen fixation, it would be necessary to account for other sources of available nitrogen, and for losses of nitrogen subsequent to fixation (Silvester, 1983). 106 ------- Nitrogen fixation has also been measured as the uptake of N1^ determined with a mass spectrometer. Although this method is quite sensitive, it is also time-consuming and expensive. This and other methods for analysis of nitrogen fixation are discussed in detail in Bergersen (1980) and Silvester (1983). Numerous studies have demonstrated pollutant-induced inhibition of nitrogen fixation. Home and Goldman (1974) observed the suppression of nitrogen fixation by blue-green algae with the addition of low levels of copper. Similarly, Francies et al. (1980, cited in Sheehan, 1984c) observed that nitrogen fixation by free- living bacteria was substantially reduced at pH levels below 6.4. Alexander (1980, cited in Sheehan, 1984c) additionally noted that the root nodulation process is highly sensitive to soil acidity. Various phases of the nitrogen cycle, and nutrient cycles in general, provide sensitive endpoints for ecological risk assessment. As nutrient cycles are critical functional processes for ecosystem survival, measurements indicating disruptions of these cycles can have far reaching implications concerning the health of an ecosystem. In addition, many nutrient-related measures such as rates of nutrient soil-leaching and nitrogen fixation are highly sensitive to low levels of pollutant stress. These endpoints are also readily measurable using a variety of techniques. The specific choice of a nutrient measurement endpoint is best made based on knowledge of the limiting factors in a particular ecosystem. 107 ------- 3.3.7 Decomposition Decomposition is the process through which organic material is degraded and organically bound nutrients are released into the ecosystem (Levin et al., 1984). Through its vital role in the immobilization and release of nutrients, decomposition links primary productivity and nutrient cycling. The process of degradation involves the movement of energy and carbon through a series of decomposer trophic levels. Microorganisms initiate the process by assimilating the proteins and carbohydrates from detritus (dead organisms and excreta). Bacteria, actinomycetes, and fungi further degradation, while cellulose-, hemicellulose-, and chitin-digesting organisms complete the process. Detritivorous animals strongly affect rates of decomposition by mechanical breakdown of detritus, and ingestion of detritus and decomposing microorganisms. Measures of decomposition thus provide good universal endpoints for monitoring pollution stress at the ecosystem level (Sheehan, 1984c). The rate of decomposition is affected by both the quantity and quality of available substrate, as well as the physical, chemical, and biological status of the ecosystem. Pollutants may alter decomposition by causing changes in the chemical status of the environment. Toxic organic pollutants (oil, industrial organic effluents, pesticides) have also been observed to affect the other controlling variables. 108 ------- 3.3.8 Methods of Measurement Methods commonly used for monitoring changes in the rates of decomposition include measures of litter bag decomposition rates and populations of litter decomposers, as well as indices of microbial activity. Litter bag analysis: Litter bag analysis is the most commonly used method for studying decomposition. The method makes use of litter bags of a predetermined pore size, which are filled with known weights of dried substrate samples and placed under a suitable amount of soil surface litter or suspended in the water of an aquatic system. Decomposition rates are measured as changes in dry-weight biomass over time (Gorden, 1972). Populations of litter decomposers: Populations of litter decomposers such as bacteria and macroinvertebrates can be monitored to assess potential effects of pollutants on decomposition. Macroinvertebrates can be measured using estimates of species abundance and biomass. Microfungal population densities can be estimated by using dilution plate count methods. Macroinvertebrate decomposer analysis can also be achieved by using species lists or by counting their numbers. Measures of decomposer populations are often combined with litter bag analysis to provide a more complete picture of changes in the decomposition process. Indices of microbial activity: The evolution of carbon dioxide from soil has often been used as a measure of microbial activity in soil. Unfortunately, the many types of organisms present in soil make it an invalid measure of microbial activity. It may instead be considered a measure of community respiration from 109 ------- the soil. The measurement of soil respiration by CO2 flux can be used as an index of litter breakdown. Estimates of CO2 flux can be made on litter samples using KOH to absorb respired CO2 (Gorden, 1972). Most naturally occurring compounds are degradable by microorganisms, and the degradation of organic compounds will stimulate microbial biomass, which may be estimated by a number of indirect methods. These methods include measurement of ATP levels and monitoring of soil enzyme activity (Visser et al., 1984). Although a large number of studies have monitored the effects of pollutants on decomposition, most of these involve laboratory, microcosm, and short-term, small-scale field investigations. Long-term effects on larger-scale natural ecosystems are thus not well understood. One field study of decomposition effects was conducted by Freedman and Hutchinson (1980b). The effects of smelter pollution were examined using analysis of forest litter decomposition. Measurements were made of litter standing crop and rates of litter input, and populations of litter-decomposers. Additional measures were also made of soil CO2 flux and soil enzyme activity. It was found that there was a reduction in the rate of litter decomposition in forested areas close to the smelter. This reduction was correlated with decreases in soil CO2 flux and acid phosphatase activity. In addition, reductions were noted in populations of soil microfungi and microarthropods at the contaminated sites (Freedman and Hutchinson, 1980b). 110 ------- In another decomposition study, Forbes and Magnuson (1980) monitored decomposition and microbial colonization of leaves in a stream stressed by coal ash effluent. ATP content proved to be a sensitive measure of pollutant-induced effects on leaf decomposition. After 27 and 96 days of exposure, the ATP content of leaves placed in the effluent-exposed stream was found to be significantly lower than that of leaves placed in a reference stream. In addition, there was a lack of normal macroinvertebrate colonization in the exposed leaf packs. This was correlated with reduced colonization and decomposition by fungi. The ash effluent appeared to indirectly affect macroinvertebrates through its interference with leaf decomposition and thus with food availability. Studies to date indicate that decomposition provides a good, readily measurable endpoint for an ecological risk assessment. Reliable, sensitive measures are available for monitoring changes in both the decomposition process and in decomposing organisms. More long-term field work must be done, however, to fully assess the usefulness of decomposition measures for monitoring long-term stress effects on an ecosystem. 3.4 CONCLUSION The functional characteristics of an ecosystem provide important endpoints for determining pollutant effects on that ecosystem. Primary production, nutrient cycling, and decomposition are critical measures of energy and nutrient flow through the system. As these processes form the foundation for ecosystem structure, they can serve as critical indicators of ecosystem survival. In addition, 111 ------- some measures, such as nutrient leaching from soil, are extremely sensitive to low levels of stress and may be particularly useful as sentinels of ecosystem decline. Decisions about which characteristics to use for a particular risk assessment should be made on a case-specific basis, depending on the nature of both the pollutants that are affecting a system and the system itself. Thus, while primary productivity might be an appropriate measure to use in a heavy-metal-polluted aquatic system, it might not be a very sensitive measure in a similarly polluted terrestrial system. In general, several kinds of functional measures are required to fully assess the responses of a system. In addition to their importance as endpoints, functional measures can also provide insights into the causes of changes in ecosystem structure. The cause of a change in species biomass, for example, might be revealed through functional measures as alterations in nutrient levels and/or primary productivity. When used in combination with measures of ecosystem structure and species level effects, functional measures can thus provide information on the dynamics involved in an ecosystem's response to stress. 3.5 ECOSYSTEM OR COMMUNITY STRUCTURE The structure of an ecosystem is defined by the communities of organisms that comprise it. Structural features include the abundance, biomass, diversity, and spatial distribution of populations, as well as the taxonomic, functional, and trophic organization of the community. While structural and functional aspects of an ecosystem are clearly linked to each other, they may be affected by stress in 112 ------- totally independent ways. For example, pollutant-induced changes in community diversity may not result in concurrent changes in ecosystem productivity. Similarly, changes in productivity can occur without any substantial change in diversity. Aspects of ecosystem response to stress may thus be missed, if only one endpoint is examined. Measures of both ecosystem structure and function must therefore be included in an ecosystem risk assessment (Sheehan, 1984b). Any analysis of stress-induced change in community structure has certain inherent limitations. Baseline data that would provide a description of the pre- exposure structure of a community is rarely available. Such baseline data is of limited value for communities that are not at equilibrium. An alternative procedure is to compare the exposed community with a similar unexposed site. This makes a questionable assumption that the two sites differ only in the presence of the pollutant (Hurlbert, 1984; Stewart-Oaten et al., 1986). Studies of structure are also limited by the ability of the investigator to identify all species and species interactions in the community. At best, a representative sample of the community must be used to deduce the state of interactions of the whole (Herricks and Cairns, 1982). Indicators of community structure range from simple, descriptive measures of abundance and biomass to more analytical, multivariate analyses of ecological similarity. The type of indicator used depends on the nature of the community to be assessed and the availability of time and resources. 113 ------- 3.6 ABUNDANCE AND BIOMASS Biomass is defined as the total weight (Clapman, 1973) and abundance as the absolute numbers of living organisms (McNaughton and Wolf, 1979) in a community. Both abundance and biomass provide simple, gross measures of community structure, and may be specified in terms of trophic or taxonomic units, thus providing comparative measures of changes in trophic structure. Abundance or biomass are often expressed in terms of population per unit of space (density). When absolute abundance measures are not feasible, often measures of relative abundance are obtained, such as number encountered per unit time, or proportion of individuals captured that are of a particular type. While these measures are useful in terms of their simplicity, they lack sensitivity and provide little information about the overall character of the system. They also display both seasonal and temporal variations. In addition, both measures carry an inherent bias. Measures of abundance tend to exaggerate the importance of small, abundant species, while measures of biomass tend to overemphasize the importance of large, nonabundant species. These measures are best used in conjunction with other measures of community structure (Sheehan, 1984b). Chlorophyll content can be used for estimating biomass of photosynthetic organisms. In aquatic systems chlorophyll estimates are a rapid and useful way to estimate phytoplankton biomass. 114 ------- A number of pollution studies have demonstrated variable effects of pollutants on ecosystem abundance and biomass. In a study of acidified Norwegian lakes, conducted by Leivestad et al. (1976), it was found that both the biomass and the density of benthic invertebrates declined with decreasing pH. Hendrey et al. (1976) also studied acidified lakes, however, and observed an increase in periphyton biomass accompanying a decrease in abundance of zooplankton and fish. In addition, Stokes (1986) reviewed studies of aquatic communities of phytoplankton, periphyton, and macrophytes and found that while acidification caused little change in biomass and productivity, it caused consistent reductions in species richness and composition. Sensitivity problems associated with abundance measures are illustrated in a study conducted by Winner et al. (1975, in Sheehan, 1984b). In this study, several measures of structure were used to examine a community of benthic macroinvertebrates dispersed along a pollutant gradient in a copper-contaminated stream. Although abundance did show an inverse relationship to copper concentrations, large variations in abundance obscured differences at median concentration levels. A clear, graded response was observable, however, using measures of species richness (see description below) (Sheehan, 1984b). Problems associated with accurate determinations of biomass are detailed in the section on Primary Productivity (Section 3.2.1). 115 ------- 3.7 SPECIES LISTS The listing of species is a straightforward indicator of community structure that is commonly used in studies of ecosystem stress. Species lists from samples at varying distances from a pollutant source can be compared to provide an indication of pollutant effects (Sheehan, 1984b) Species lists may also comprise the first step in determining other measures, such as species richness, and species dominance, which can in turn be used as components of biological diversity and similarity indices (Herricks and Cairns, 1982). In ecosystem-level studies of effects of pollutants on experimental lakes, among the most sensitive responses were species composition of phytoplankton and disappearances of sensitive aquatic species (Schindler, 1987). 3.8 BIOLOGICAL INDICES (POLLUTION INDICES) Biological indices are measures developed from observations of responses of groups of indicator species (see Indicator Species, Section 2.3.3) to pollutant stress in aquatic systems (Hellawell, 1977). They provide numerical rankings of species and species assemblages. These indices were first developed in the early 1900's, primarily as a method of evaluating the effects of municipal sewage or organic wastes on aquatic systems (Sheehan, 1984b). Numerous biotic indices have been proposed over the years. Some of the simpler indices compare numerical differences between groups of species or individual species, which are classed as being tolerant or intolerant to pollution. Many of these simple indices look only at species numbers and make no allowance 116 ------- for species abundance. For example, in a pollution index developed by Beck (1954, in Hellawell, 1977), the index is: I = 2C\-C2, where Cj represents the number of intolerant and C2 represents the number of tolerant macroinvertebrate species. Some of the more complex biotic indices integrate measures of key species abundance, their known pollution tolerance, and their reliability as indicators. In one such index, developed by Chandler (1970, in Hellawell, 1977), scores are determined for both species abundance and pollutant tolerance. The resultant values are then combined to produce a total biotic score. One of the oldest indicator systems is the saprobic approach (Kolkwitz and Marsson, 1902, 1908, 1909, cited in Sladecek, 1963). This system was designed for use in running water. It is based on the assumption that the presence of particular organisms living in running water bodies permits estimation of both the level of contamination in the water and the trend of the general conditions (for example, deterioration or self-purification). Water quality is predicted based on the number of specific microorganisms present in 1 ml of test water as well as chemical properties of the water such as DO, H2S content, and BOD (Sladecek, 1963). The data on the quantity of microorganisms and the chemical parameters are fit into rigid categories that describe water quality. Sladecek (1963) lists criticisms of the saprobic system including force fitting of data, consideration of only processes related to bacterial decomposition, and no validation of the system. Overall, saprobity is not considered a very valuable method for most situations. 117 ------- The usefulness of biotic indices in evaluating ecosystem stress was illustrated in a study conducted by Solbe (1977). In this investigation, two biotic indices, two diversity indices (see section on Diversity Indices) and a similarity index (see section on Similarity Indices) were used to evaluate the invertebrate communities in a zinc-polluted stream. It was found that both the Trent Biotic Index and the Chandler Biotic Score served as good indicators of pollutant effects, with the Chandler Biotic Score being most sensitive to low levels of pollution. MargalePs diversity indices and a cluster analysis that measured similarity, however, proved to be less sensitive than the biotic indices. While biotic indices provide a semi-quantitative measure for evaluating ecosystem stress, their usefulness for studying the effects of toxic chemicals has not been well tested (Sheehan, 1984b). In addition, many of the indices require subjective determinations of organism tolerance or indicator values (Herrick and Cairns, 1982). Scores produced, furthermore, may not clearly distinguish between different combinations of evaluated factors. Thus, the same score may be obtained both for a few individuals of a sensitive, pollution-intolerant species and a ubiquitous, pollution-indifferent species (Hellawell, 1977). As Solbe's study indicates, however, biotic indices can provide sensitive indicators for certain well- defined systems. 3.9 SPECIES RICHNESS Species richness can be defined as the number of species present in a system (McNaughton and Wolf, 1979). As it would be virtually impossible to determine the total number of species in a natural community, richness has been quantified 118 ------- as the number of species per fixed number of individuals, or the number of species per unit area (Peet, 1974). Species richness is widely used as an endpoint in measurements of polluted ecosystems. It is also a component of measures of species diversity (Levin et al., 1984). When determined by making direct counts of species numbers in a sample, it provides a simple, practical, objective measure of community structure. It can also be used to compare species diversity between communities, when the relationships of species importance are similar (Peet, 1974). A problem in the measurement of species richness is its inherent dependence on sample size. A series of simple richness indices have been developed that assume a consistent relationship between species numbers and sample size. If this assumption is not satisfied, these richness indices will be biased in an unknown manner. One such index, developed by Margalef (1951, cited in Peet, 1974), is described by the logarithmic relationship: Rj = (S-l)/Log N, where S is the number of species and N is the number of individuals in a sample. Another well-known example of a richness index that assumes a known relationship between species numbers and sample size is Preston's analysis. Preston's relationship calculates the expected number of species in a total sample, based on the proposition that a log-normal distribution best describes species abundance data (Peet, 1974). This log-normal distribution arises from the fact that a multiplicity of factors causes populations to increase in a geometric, rather than arithmetic, pattern. In a large sample, individuals in a population thus tend to be distributed in log-normal pattern (Gray, 1979). Patrick (1949, in Levin et 119 ------- al., 1984) was able to use this index to demonstrate a loss of diversity in diatom assemblages of stream communities polluted with sewage and organic outputs (Levin et al., 1984). Studies have shown that species richness is a good measure of pollutant stress. It is sensitive to gradients of pollutant concentration and thus provides a "concentration-response relationship" useful for monitoring stress in ecosystems (Sheehan, 1984b). For example, in a retrospective study conducted by Stokes (1986), the effects of acidification on phytoplankton, periphyton, and macrophytes in a number of aquatic systems were compared. The results showed that species richness declined consistently with declining pH. In contrast, biomass and productivity did not change substantially. Varied effectiveness of species richness measures have been observed in terrestrial studies. Freedman and Hutchinson (1980a) found that species richness provided a sensitive monitor of distance from a pollutant source in a forest community. In this study, richness of the ground flora vegetation was found to correlate well with distance from a metal smelter. In measures of the overstory or tree canopy, however, biomass-related measures were more sensitive than measures of either richness or diversity. The lack of correlation in the overstory results was attributed to the fact that only a small number of species were present in the initial community. While species richness can provide a useful measure of pollutant stress and community-level effects along a gradient, it does not provide a complete picture of ecosystem dynamics. As the Stokes (1986) study indicates, changes in species 120 ------- richness do not necessarily complement changes in biomass or productivity. It is evident that in a system with functional redundancy, one species might readily compensate for the loss of another (Sheehan, 1984c; Peet, 1974). Thus, richness alone may not be a complete measure but is a useful component of a complex of measures of ecosystem function. 3.10 DIVERSITY Although the concept of diversity has been broadly explored in the literature, there is still no universally accepted definition of the term (Sheehan, 1984b). A generally held definition, however, involves the dual-component concept introduced by Simpson (1949, cited in Peet, 1974). According to this definition, diversity incorporates both species richness and equitability (the relative evenness of abundance of species distribution) (Peet, 1974). The concept thus provides a measure of the "frequency of species occurrence in a community" (Herrick and Cairns, 1982). Diversity indices are widely used measures of community structure. The theory behind these indices is that diversity increases as a community ages and higher diversity is associated with community stability (Odum, 1971). Caswell (1976, in Gray, 1979), however, has shown that diversity decreases in the final stages of succession. Additionally, the relationship between diversity and stability has come under serious question. Nonetheless, there is evidence to suggest that pollution can produce a decline in ecosystem diversity (Sheehan, 1984b). 121 ------- Diversity calculations can additionally provide a useful method of summarizing large amounts of data (Herrick and Cairns, 1982), and they permit comparative measures, evaluating the degree to which species abundances in stressed environments deviate from predicted values (Hellawell, 1977). Because of these advantages, large numbers of diversity indices have been developed over the years. Some of these emphasize the richness component of diversity, some emphasize the equitability component, and others incorporate both components (heterogeneity). A variety of equitability-based indices have also been developed. These generally scale a heterogeneity measure to the maximum possible value for a fixed species number and sample size (Peet, 1974). Equitability indices are based on the assumption that if species numbers are equal, communities with the most equitable distribution of individuals among species can be considered to be the most diverse (Sheehan, 1984b). Some of the heterogeneity indices are derived from information theory. Heterogeneity, in these cases, is based on the degree of uncertainty associated with the species of an individual randomly selected from a population. As the number of species and evenness of the sample increases, so does the uncertainty of selecting an individual of a particular species. Thus, increased uncertainty is correlated with increased diversity. The Shannon-Weaver (Wiener) formulation is one of the most commonly used information-based indices. Pielou (1966) suggests that this method is useful for estimating diversity in a large collection of organisms in which all the species 122 ------- cannot be identified. She further demonstrates how different formulas for estimating Shannon's measure of information can be used to estimate diversity in different kinds of communities. The Shannon-Weaver and other diversity indices have produced mixed results in pollution studies. Bechtel and Copeland (1970) found that fish species and biomass diversity, determined using the Shannon-Weaver formula, provided useful indicators of environmental pollution in Galveston Bay. Similarly, in a second study by the same authors, the diversity of phytoplankton, zooplankton, and nekton provided a reliable indicator of pollutant levels (Copeland and Bechtel, 1971, cited in Sheehan, 1984b). On the other hand, in the previously mentioned study by Freedman and Hutchinson (1980a), the Shannon-Weaver formula did not prove to be a powerful indicator of pollutant-induced stress. It was less sensitive than measures of either biomass or total cover to changes in a forest community along a gradient of smelter pollution. Similarly, in a study of a deciduous forest exposed to air pollution, McClenahen (1978) found that diversity, as measured by the Shannon- Weaver index, was inadequate to describe changes in the pollutant-exposed shrub layer. In this situation, richness and evenness varied inversely and thus obscured any changes that might have been evident from measures of diversity. In another study conducted by Eisele and Hartund (1976, cited in Sheehan, 1984b), it was found that, in a methoxyclor-contaminated stream community, invertebrate populations declined while diversity remained unaffected. 123 ------- Because of studies like those described above, numerous authors have offered alternative methods for measuring structure. Gray (1979) has suggested, for example, that measures such as the Shannon-Wiener (Weaver) index and the rarefaction method of Sanders can be difficult to interpret. Additionally, they are not very sensitive to pollutant-induced changes in community structure. Gray suggests that a log-normal distribution analysis, an analysis that measures the departure from the log-normal distribution of individuals among species, is a much more sensitive measure of ecosystem stress. Sheehan (1984b), however, notes that this technique has not been widely used in ecotoxicology studies. Standard diversity indices have been criticized in the literature for their inherent biases. Shannon's formula and similar indices tend to overemphasize the importance of rare species at low population densities (Sheehan, 1984b). On the other hand, Simpson's and related indices tend to be most sensitive to changes in the most common species (Peet, 1974). Another problem associated with use of diversity indices is the difficulty of interpreting their results (Hellawell, 1977). Green (1975, cited in Read et al., 1978) noted that diversity indices could measure changes in community structures, but that they provide little information on the nature of the change. This point is further emphasized by Herricks and Cairns (1982), who suggest that information about the community structure is lost when large quantities of data are summarized in a single index. Green (1979) argues against the use of diversity indices. He indicates that the concept of diversity is vague, combining two components that may vary independently. He further notes that diversity indices are "often uncritically 124 ------- applied, without regard to the assumptions implicit in the various diversity formula and the biases in their estimation." Problems with sensitivity, bias, and information content of diversity indices may render them inadequate measures for assessment of pollutant effects in many ecosystems. Sheehan (personal communication) has suggested that, while these indices might be useful as measures of gross changes in pollutant-stressed systems, they are not particularly sensitive to low level effects. 3.11 COMPARATIVE INDICES (SIMILARITY INDICES) Comparative indices are designed to evaluate two or more parameters of community structure over space and time. These methods were initially designed by plant ecologists to evaluate plant communities. They are useful for monitoring the effects of pollutants at varying distances from a source or for detecting ecosystem recovery or deterioration over time (Hellawell, 1977). They also provide a means of comparing stressed and unstressed communities (Brock, 1977, cited in Sheehan, 1984b). Comparative indices identify the similarities or differences between samples. Some of these indices compare species lists or species abundance data. Those which make use of species lists include the Jaccard index, a measure that is particularly useful for distinguishing between similar samples. The Dice and Ochiai coefficients, on the other hand, are species list indices that are most sensitive for very dissimilar samples. 125 ------- Other similarity indices make use of quantitative abundance tabulations or distance measures between communities. These measures, which include percent similarity, coefficient of community, the Czekanowski index, and the Bray-Curtis coefficient, may require modification to account for overestimates of the importance of abundant species (Herricks and Cairns, 1982). Species ranking provides another method of comparing communities. In these methods, species are ranked by their relative importance in a community. The rankings in different communities are then compared (Hellawell, 1977). Ranking indices include Spearman's rank correlation coefficient and Kendall's coefficient. These measures have been criticized for a number of different reasons. Problems include methodological difficulties involved in evaluating species ties and species absence (Herricks and Carins, 1982) and similarities of rankings that can arise in structurally different communities (Hellawell, 1977). Finally, correlation coefficients have been used to identify relationships between communities based on species distributions or species presence or absence. These measures include the point correlation coefficient and the product moment correlation. Studies have shown that similarity indices provide sensitive measures of low- level stress in polluted ecosystems. These measures appear to be more sensitive than those obtained by diversity indices (Sheehan, 1984b). For example, in a study conducted by Marshall and Mellinger (1980, cited in Sheehan, 1984b) changes in a Lake Michigan zooplankton community in response to cadmium pollution were assessed with percentage similarity and coefficients of community 126 ------- indices. Using these measures, significant changes in community structure could be identified at cadmium levels that produced minimal effects on diversity indices. The coefficient of community was also found to be a useful measure by McClenahen (1978) in his study of air pollutant effects on a deciduous forest. As previously mentioned, McClenahen found that changes in community structure at low pollutant levels were difficult to identify using the Shannon diversity index. He noted, however, that the coefficient of community could identify significant correlations between community structure and air pollutant exposure along a decreasing pollutant gradient. 3.12 MULTIVARIATE ANALYSIS A more complex approach to similarity measures used in ecological risk assessment is multivariate analysis. Multivariate methods permit associations to be made between samples or groups of samples based on many different variables (Herricks and Cairns, 1982). Green (1979) describes such techniques as being powerful measures that are sensitive to subtle patterns of difference and provide a visual representation of relationships between samples. Multivariate measures of similarity can be broadly grouped into two catagories, cluster analysis and ordination. Clustering methods involve grouping samples with similar characteristics into clusters and fitting the clusters into a two-dimensional display of complex resemblance patterns. The display may then be tested using an indicator of accuracy. 127 ------- Ordination techniques involve the organization of samples along one or more axes or gradients, using statistical comparisons of variables such as species abundances or similarity coefficients,. Strong discontinuities between communities are represented by a steeply sloped gradient (Odum, 1971). These techniques have been found to be useful in determination of the causes of associations between communities (Herrick and Cairns, 1982). 3.13 TROPHIC ORGANIZATION Trophic structure may serve as a framework within which other measures of community structure or function can be monitored. Determinations may be made of productivity, dominance, biomass, and other endpoints at different trophic levels. Energy from the sun is stored in chemical bonds during photosynthesis and subsequently passes through a series of organisms in an ecosystem. This movement of energy through the system is characterized as a food chain. Food energy passes through the food chain in steps, as one organism feeds on another and is eaten itself. Feeding steps or levels are defined as trophic levels (McNaughton and Wolf, 1979; Odum, 1971). Organisms in a particular system are functionally grouped into different trophic levels. Green plants are classified as producers, herbivores as primary consumers, carnivores as secondary consumers, and secondary carnivores as tertiary consumers (Odum, 1971). Trophic relationships may also be described in terms of feeding mechanisms. Cummins (1974) categorized aquatic insects as 128 ------- shredders, collectors, scrapers, and predators. Functional feeding groups have also been termed guilds (Johnson, 1981). Changes in trophic organization may be measured in terms of changes in species numbers, species abundance, or species diversity at different trophic levels. They may also be quantified in terms of shifts in trophic dominance. Trophic shifts were observed in a study by Hall and Likens (1980, cited in Sheehan, 1984c). They found that in acidified streams, the ratio of consumers to producers declined. Another example of trophic shifts in aquatic systems are blooms of phytoplankton primary producers following insecticide-induced removal of crustacean herbivores (Hurlbert, 1975). 3.14 SPATIAL STRUCTURE The spatial structure of an ecosystem involves both the vertical and horizontal patterns of species organization within the community. This spatial organization is readily quantifiable for terrestrial producers. It is difficult to define, however, for mobile terrestrial consumers or in aquatic communities where many species are mobile. Spatial structure has been used as a measure of pollution-induced changes in ecosystem structure. Sheehan (1984b) describes a number of terrestrial studies in which simplification of structure was observed to result from environmental stress. Changes in plant distribution were also observed by Dawson and Nash (1980) in a study of the effects of copper smelter effluent on a forest ecosystem. Reductions in the abundance of annuals and grasses were observed in the area near the 129 ------- smelter. Large shrubs, however, remained abundant. The pattern of distribution was attributed to the presence of a thin layer of copper-contaminated, acidified soil in the area near the smelter, which killed plants such as annuals that have shallow root systems. A small number of studies thus suggest that spatial structure offers a potentially useful endpoint for measures of toxicant effects on the structure of terrestrial systems. 3.15 GUILDS The guild concept, though still in the early stages of research and development, provides a potentially useful organizing structure for the performance of a risk assessment. A guild was originally defined by Root in 1967 as a "group of species that exploit the same class of environmental resources in a similar way" and was designed as a method of grouping species based on niche overlap rather than taxonomy (Root, 1967). In general, this concept has been used to group animals based on feeding or nesting strategies. Thus, animals might be grouped into the same guild if they all feed on a common set of flowering plants, forage on the ground for insects, or nest in tree holes. A community can then be conceived of as containing a complex of potentially interacting guilds (Krebs, 1985) which can be organized into a matrix. A feeding and nesting guild matrix for an avian community in a terrestrial habitat is illustrated in Figure 3-1. Since its original development, the usefulness of the guild concept as a general research tool and as a component of the risk assessment process has been 130 ------- extensively analyzed in the literature. It has been suggested that the concept simplifies the analysis of community structure by reducing the number of components in a community to a manageable number. In addition, use of the guild concept might diminish the economic costs of a risk assessment by reducing the amount of long-term input required to monitor the status of a stressed system (Landres, 1983; Severinghaus, 1981). The study of a guild containing a small group of ecologically related species may thus provide a practical alternative to the single-species or community analysis approaches to risk assessment (Petersen, 1986). One of the main problems associated with the use of the guild concept arises from the lack of an objective mechanism of delineating guilds in a particular system. Guilds are often defined by an investigator, based on a subjective classification of resource use (Karr, 1980, cited in Landres, 1983). They may also be defined somewhat more objectively using statistical techniques such as cluster analysis, principal components analysis, and canonical correlation (Landres, 1983). For example, statistical methods were employed by Poysa (1983) to study guilds in a waterfowl community. Both cluster analysis and principal component analysis were used to identify guilds based on feeding habitat and feeding method. These analyses required investigator-determined classifications of feeding habitats and feeding methods, thus introducing a subjective component to even these forms of analysis. 131 ------- Figure 3-1. Guilds matrix with primary feeding and nesting zones. (Adapted from Verner, 1984) ai z o N o z 2 UJ HI u. > tr < 2 E Q. vaaw Snags ACWO COHA $8HA Am. ANHU BTM Tree canopies «.n WCKI YRWA WBK SOON PHAI HIM NOOA OCJU NUWO YSSA W»NU Tree boles and limbs WBSH HOWR 90 HfTH CATM Btwn RCKI WW Shrubs rov UM TUVU AOAO AMXI PTTHA A**0 CAQU SHCO SCOW MOOO MTO COMA LI 00 con. OHOW Ground WCMf ACV WTO star LfOW ftCJA HOfl © Jr Ground Shrubs Tree boles and limbs Tree canopies Snags 1 » 9 Jfi « M ¦D s Ł PRIMARY NESTING ZONE 132 ------- Because of the subjectivity of guild determination, there is a general lack of consistency in the ways in which different investigators define guilds. Yerner (1984) found this problem to be particularly acute in the literature on avian guilds. In some instances, different investigators identified different sets of guilds among avian species in similar habitats. It appears that the ability to classify organisms into guilds is, in part, based on the experience and understanding of the ecologist performing the analysis (Gauch and Whittaker, 1981, cited in Landres, 1983). Thus, the usefulness of the guild concept may vary with investigator competence. Recently, a number of investigators have attempted to define an approach to the use of guilds that would be useful in the performance of a risk assessment. One such approach involves the use of "guild indicators" and is based on the idea that all members of a guild will react similarly to environmental change. The responses of any one species of the guild can thus be used as an indicator of the responses of the others (Severinghaus, 1981). While this method might provide an economic approach for conducting a risk assessment, it has been criticized by a number of investigators. Landres (1983) notes that while all members of a guild may exploit a common resource, they may not necessarily be similar in all aspects of their life history and thus may not be equally affected by a change in that resource. Some species, for example, may be readily able to shift to a new resource while others cannot (Landres, 1983; Verner, 1984). It would thus be wrong to assume that a single species can automatically serve as an indicator for the rest of the guild. 133 ------- The whole guild approach is another method of using guilds as a tool in risk assessment. This approach, as described by Verner (1984), uses the entire guild as an indicator of habitat quality. Guild categories are based on "zones of habitat that are likely to respond in similar ways to various sorts of perturbations." The guild then provides a basis for obtaining information on individual guild species and on the ability of the habitat to support them. Verner used this approach to categorize bird populations in a pine-oak woodland. His guilds were organized into a grid based on primary feeding and nesting zones (see Figure 3-1). Intraguild comparisons were made of differences in species richness on grazed and ungrazed plots of land. It could then be determined how well an ungrazed plot would support a guild such as ground-feeding birds. Szaro (1986) took a third approach to using guilds by placing groups of bird species in a ponderosa pine forest into response guilds. These guilds were made up of bird species that demonstrated similar changes in densities, in particular habitat-zones, over time. Typical response guilds included: o Species that were absent in one year on most or all study plots and showed no preference for any forested site; o Species that had their highest densities on the medium-cut and light-cut plots; and o Species that had their highest densities on untreated, light-cut, and medium-cut plots. Szaro found that there was a strong correlation between species density and response guild density. Szaro further noted that this correlation was much weaker for functional guilds (habitat-zone guilds) as defined by Verner (1984). He 134 ------- concludes that response guilds revealed relationships that would not be seen using other types of guild matrices. In conclusion, changes in guilds are potentially useful endpoints for ecosystem risk assessments. They facilitate analyses of species responses to functional or structural changes in specific habitat zones of an ecosystem. Furthermore, they reduce the numbers of community components and thus simplify community structure analyses. Unfortunately, while there are numerous studies that make use of guilds to analyze ecosystems, little work has been done to apply the guild concept to systems exposed to pollutant stress. In addition, there is still no consistently agreed-upon approach for defining guilds. Further research and testing will be required to determine if an approach such as response guilds would be useful in an ecosystem risk assessment. 3.16 CONCLUSION Measures of changes in ecosystem structure can provide important components of an ecosystem risk assessment. Measures such as abundance, biomass, , and species lists provide relatively simple, gross measures of ecosystem stress. Species richness appears to be a good, sensitive measure of pollutant- induced stress, and can provide a partial picture of changes in community structure. Similarity indices and multivariate analyses further provide important measures of differences in pollutant-effects between systems or along gradients. They also seem to be sensitive measures of pollutant effects. 135 ------- Other methods may prove to be less sensitive or less desirable for use as endpoints of toxicity. Diversity indices, for example, produce mixed results in ecosystem studies and often prove to be less sensitive than richness or similarity indices. Biotic indices are often confounded by the need for subjective determinations of organism tolerance or value as indicators. In general, diversity and biotic indices are not strongly recommended as measures for use in ecosystem risk assessments. It is important to note that measures of ecosystem structure, such as species richness or diversity, are influenced by numerous poorly understood factors, including regional diversity, geographical dispersal, competition, predation, adaptation, and environmental variation (e.g., Ricklefs, 1987). All of these factors may vary among local sites. Therefore, in comparing contaminated with uncontaminated sites, it is difficult to conclude with certainty that any differences noted in ecosystem structure are due solely to the contaminant. Whole ecosystem experimentation may be the only way to determine the effects of a particular chemical on an ecosystem (Levin et al., 1984). 3.17 STABILITY One of the fundamental characteristics of an ecosystem is its degree of stability. The concept of ecosystem stability is associated with a wide spectrum of ecosystem traits such as inertia, elasticity, and amplitude. These traits generally refer to the tendency of a system to be upset by, and to recover from, the effects of perturbations. Elements of stability are clearly important factors to consider in any evaluation of an ecosystem's response to pollutant-induced stress. 136 ------- A large body of literature exists on the various aspects of ecosystem stability and recovery. While the theoretical aspects of these concepts have been debated for years, their practical use in the monitoring of pollutant-induced ecosystem stress is still in its early stages (Sheehan, 1984b). Because stability- related concepts have been detailed in another document (Ecological Risk Assessment Issues, Volume II. Ecosystem Stability/Recovery Report, TRI, 1988d), they will be briefly summarized here. Stability can be defined as the ability of a system to return to an equilibrium state after a temporary disturbance. The equilibrium state applies to aspects of both ecosystem structure and ecosystem function (McNaughton and Wolf, 1979). The stability definition implies that the less a system fluctuates, the more stable it is (Holling, 1973). Many systems, however, naturally experience repeated cyclic fluctuations. These systems would still be considered stable if they displayed a tendency to return to their original patterns of fluctuation (Sheehan, 1984b). Considerable debate has focused on the relationship between stability and diversity. The original hypothesis stated that species diversity stabilizes ecosystem functional properties (McNaughton, 1977). While some investigators provide evidence to support this concept (Van Voris et al., 1980), mathematical models suggest that under certain conditions increased diversity has a destabilizing effect (May, 1973, cited in McNaughton, 1977). Further theoretical and empirical research indicates that there is not a simple relationship between stability and diversity. Stability may not decrease as diversity increases if there is a decrease in connectance or strength of food-web interactions (Pimm, 1984; 137 ------- Moore and Hunt, 1988). An alternative view of the diversity-stability relationship is that diversity is a function of environmental stability. If, as theory predicts, complex systems tend to be dynamically fragile, they may be expected to occur in more predictable habitats (May, 1986). Some empirical evidence suggests that diversity is maximized at intermediate levels of environmental disturbance (Connell, 1978). Methods of determining stability involve examination of changes in both structural and functional characteristics representative of total ecosystem responses. In a study by Van Voris et al. (1980), loss of calcium in leachate was used as an index of ecosystem stability. Other studies have focused on fluctuations in diversity and population density. Sheehan (1984b) suggests that measures of similarity, species richness, total biomass, and primary productivity would also be suitable for describing ecosystem fluctuations. A large variety of terms have emerged from the concept of ecosystem stability. These include: inertia, elasticity, amplitude, resilience, hysteresis, malleability, constancy, and persistance (Sheehan, 1984b). A few of these are briefly summarized below. 3.17.1 Resilience Resilience can be defined as the ability of a system to absorb perturbations and return to a stable configuration (Holling, 1973). The resiliency of a particular ecosystem is dependent on a number of different factors. In general, high resiliency is found in moderately stable systems, located in temperate areas, 138 ------- which have high productivity and large niche sizes, and which experience large environmental fluctuations. On the other hand, resilience tends to be low in communities that are highly stable, such as tropical rain forests, or highly unstable, such as the tundra. These systems are either too specialized to adapt to large perturbations or too low in productivity to readily compensate for the stress-induced changes (Clapham, 1973). 3.17.2 Amplitude Amplitude is one component of resilience. The amplitude of a system refers to the amount of perturbation that the system can absorb and still retain the ability to recover (Westman, 1978, cited in Sheehan, 1984b). It describes the recovery threshold for the system. Amplitude can be ascertained by monitoring recovery along a pollutant gradient or comparing pollutant effects at varying levels of exposure. In one such amplitude study conducted by Baker (1971, cited in Sheehan, 1984b), it was found that salt marsh grass loses its ability to recover from crude oil applications after 12 successive oilings. 3.17.3 Elasticity Another component of resilience is elasticity. Elasticity is defined as the speed of recovery to a stable state and has been measured using an index of elasticity (Stauffer and Hocutt, 1980). Factors that influence this index were listed by Cairns and Dickson (1977). They include: 139 ------- o The availability of nearby epicenters of replacement organisms for the damaged system; o The mobility of organisms in the system; o The management capabilities for rapid control of the damaged system; o The condition of the system after perturbation; and o The extent to which residual toxicants remain in the system. One of the problems involved in estimating elasticity following pollutant stress is the fact that recovery may not be a linear process, and thus may be difficult to predict early on. In addition, different features of a system may recover at different rates, making it difficult to produce one comprehensive prediction of time-to-recovery (Sheehan, 1984b). 3.17.4 Inertia Inertia (alternatively termed resistance) represents the ability of a system to resist stress-induced changes (Cairns and Dickson, 1977; Sheehan, 1984b). It appears to be related, like resilience, to the degree of specialization and the amount of exposure to environmental fluctuations that are normally experienced by a system (Boesch, 1974, cited in Sheehan, 1984). According to Sheehan (1984b), while inertia may be quantified as the amount of pollutant per unit area per unit time that will cause a specified amount of ecosystem damage, resilience would be quantified as the maximum level of pollutant exposure that would still permit ecosystem recovery. An index for the prediction of the inertia of streams has been developed by Cairns and Dickson (1977) and has been used to classify stream fish communities by Stauffer and Hocutt (1980). The index of inertia of a stream may be calculated using the following parameters: o The tendency of the system to experience variable environmental conditions; o The functional and structural redundancy of the system (in terms of species numbers and trophic interactions); 140 ------- o The dependability of the stream flow, its flushing capacity, etc.; o The detoxification capacity of the water; o The proximity of the system to major ecological transition areas; and o The water quality management facilities of the area (Cairns and Dickson, 1977; Stauffer and Hocutt, 1980). 3.18 CONCLUSION Various aspects of stability have been briefly reviewed. One measure of stability, inertia, may be potentially useful for predicting effects of pollutants on ecosystems. Measures of stability in general, however, require long-term monitoring of numerous parameters. Thus, stability measures are not useful for rapid determinations of effects. In addition, these measures have not been widely used in the field and require more testing to establish their applicability. 141 ------- CHAPTER 4 SUMMARY 4.0 INTRODUCTION A variety of ecotoxicological endpoints have been proposed to assess the effects of pollutants on ecological systems. Potential endpoints occur at the level of the individual organism, the population, and the ecosystem. In general, endpoints at lower levels of organization (organism or suborganism levels) have been more widely used because they are simpler, more rapidly and inexpensively assessed, and are most useful in determining the mechanisms of toxicological effects. Endpoints at the population or ecosystem levels of organization are more complex and difficult to interpret, but are probably ecologically more realistic because they incorporate the complexity of interactions among organisms, and between organisms and their abiotic environment. A major unresolved question is to what extent endpoints at lower levels of organization can be used to predict pollutant impacts at higher levels of organization. It is important to note that the ecosystem- and population-level ecotoxicological endpoints, and to a lesser extent the organismal-level endpoints, are often generalized indicators of stress or disturbance, rather than specific responses to toxic pollutants. There are many sources of stress or disturbance in natural, unpolluted ecosystems (e.g., Sousa, 1984), so evidence of disturbance in an ecosystem is not necessarily an indication of significant effects of pollution. Ideally, measurements would be made before and after pollutant discharge to provide direct evidence of pollutant effects (Stewart-Oaten et al., 1986). 142 ------- 4.1 ORGANISM-LEVEL ENDPOINTS Physiological endpoints most closely related to individual fitness are acute mortality, growth and development, and reproductive success. Acute lethality testing is widely used to provide minimal estimates of toxicity. However, such testing is not sufficiently sensitive to assess sublethal or chronic effects, which occur at lower toxicant concentrations and may be of considerable ecological importance. Biochemical response endpoints may provide information on the mechanism of toxic action. Since biochemical processes are in general particularly sensitive to pollutants, biochemical response endpoints may provide early warning of potential impacts on the individual. However, most biochemical processes respond to conditions other than pollutant stress and the response of these endpoints may be adjusted with acclimation of the individual to the stress. Correlations between biochemical response endpoints and individual success need to be established to enhance the value of these sensitive endpoints as predictors of higher level impacts. Osmoregulatory activity is an appropriate endpoint for assessing impacts on certain freshwater and estuarine fish and invertebrates. Again, the ability of the individuals to acclimate to osmoregulatory stress must be considered in interpreting osmoregulatory response data. 143 ------- Musculoskeletal endpoints have been used to monitor stress responses in fish. Correlations between abnormalities and the ecological success of deformed fish need to be established. Respiratory activity has been used as a response endpoint for a number of species. However, it is difficult to generalize about the patterns of respiratory response to stress. Respiration rates may be elevated or inhibited by pollutants and exposed individuals may adjust ventilation rates with acclimation time. Behavioral alterations are appropriate endpoints for impact assessments if they act either to protect the individual from harm, as in avoidance behavior, or make the individual more vulnerable to the stress, as in the loss of antipredator behavior. Although behavioral responses are not easy to demonstrate in the laboratory or in the field, these endpoints, if demonstrated, may be easily extrapolated to predict potential population-level effects. Genotoxicity and carcinogenicity are endpoints that provide early warning of stress. Data must be gathered on the natural incidence of mutations and tumors to aid in interpreting the importance of chemically induced mutation and tumor incidence rates. Endpoints measuring growth, development, and reproductive success of individuals are of most obvious utility in predicting population-level impacts. Because these endpoints are directly related to population success, their use in impact studies where single-species test data is extrapolated to predict population- level impacts is recommended. These endpoints have been less frequently used 144 ------- because of the time and expense required to conduct full-life-cycle chronic toxicity tests. However, the more frequently used short-term physiological and biochemical endpoints cannot be recommended until their relationships to organismal growth and reproductive success are determined. A number of studies have documented interactions between effects of pollutants and abiotic and biotic factors in the environment. These examples illustrate the inadequacy of laboratory single-species, single-factor testing to estimate all ecological effects of contaminants, and point to the necessity of relating ecotoxicologic effects on individual organisms to population- and ecosystem-level effects of pollutants. 4.2 POPULATION-LEVEL ENDPOINTS At the population level, stress response may be monitored in terms of changes in the abundance, distribution, age structure, or gene makeup of exposed populations. The first three endpoints can be quite clearly related to the overall success of the exposed population. Changes in the gene pool may be related to future adaptability of the population to similar types of stress. Also at question is the selection of an appropriate population or populations to be monitored in an impact assessment. Quite clearly, monitoring the effects on commercially or aesthetically valuable species is important to predict impacts on those species. More valuable to predicting higher level impacts is population- response data on representative and ecologically important species within exposed communities. Included within this category are keystone species, which strongly 145 ------- influence the structure of the communities or the functioning of the ecosystem. If there is interest in extrapolating population response to predict ecosystem-level impacts, emphasis should be placed on gathering data on populations from major functional groups, including primary producers, primary and tertiary consumers, and decomposers. A problem in the use of population-level endpoints as indicators of pollution is that the numerous factors regulating population structure are, as yet, poorly understood. This makes it difficult to discriminate pollutant effects from naturally occurring processes. As population structure is influenced by interactions among population members, with other populations, and with the abiotic environment, it becomes necessary to examine the effects of pollutants at the ecosystem level. 4.3 ECOSYSTEM FUNCTION ENDPOINTS The analysis of functional response endpoints can provide data on energy flow and nutrient cycles. The functional capability of the ecosystem is, in fact, the ultimate criterion of ecological success. The effective use of endpoints in describing impacts is dependent upon a theoretical and practical knowledge of ecosystems for proper interpretation, and on the collection of sufficient baseline data to establish normal process rates. A longer history of measuring functional response variables will be necessary to establish threshold values for unacceptable reductions in functional capability. 146 ------- Primary productivity provides the energy for the base of the food web. This process has been shown to be sensitive to a variety of pollutants and other forms of stress. Reductions in primary productivity that are of substantial magnitude and of long duration are unquestionably detrimental to energy processing in exposed ecosystems. Disruptions in material cycles will be critical if the effects on cycling processes indirectly inhibit ecosystem production. Material cycles can be upset by pollutant inhibition of the decomposition process, interference with the functional links in specific nutrient cycles, or disruption of nutrient conservation mechanisms. The effects on decomposition can be measured in terrestrial and aquatic ecosystems and changes in decomposition rate and the completeness of mineralization can be related to the level of pollution stress. At present, little data are available on the long-term impacts of reduced decomposition on ecosystem production. Specific nutrient cycling processes are key to the production efficiency of ecosystems. Identification of the critical cycles in specific ecosystems will be necessary for the selection of appropriate monitoring points. Nutrient conservation is exceedingly important in terrestrial ecosystems. Evidence of excessive leaching of essential nutrients is a sign of stress. Leaching loss of nutrients has been correlated with reduced nutrient availability in the plant-root zone and reduced plant growth in nutrient-deficient soil. 147 ------- A problem in the use of ecosystem-function endpoints is their relative insensitivity to ecosystem structure. Shifts in species composition to more pollutant-resistant species may or may not result in changes in such functional processes as productivity or nutrient cycling. Thus, an assessment of pollutant effects at the ecosystem level should include both structure and function endpoints. Because the factors controlling ecosystem structure and function are numerous and poorly understood, it is difficult to distinguish ecosystem-level effects of pollutants from naturally occurring processes. Many of the ecosystem- level endpoints depend on the questionable assumption that unpolluted ecosystems are at a stable, undisturbed state. 4.4 ECOSYSTEM STRUCTURE ENDPOINTS Measures of ecosystem structure can provide important data for ecosystem risk assessments. Structural changes in stressed ecological communities may be visualized as an information network reflecting environmental conditions, but not demonstrating the external mechanisms or internal interactions that brought about the reorganization in species composition or dominance patterns. Structural endpoints such as abundance and biomass of communities provide relatively simple, gross measurements of ecosystem stress. Species richness has been shown to be sensitive to the level of stress and can provide a partial picture of changes in community composition accompanying stress. 148 ------- Combined numerical indices such as similarity and ordination measures may be used to track changes in community structure with changes in pollutant concentrations. Although diversity indices have been widely used in hazard assessment studies, these integrated measures are often insensitive to stress and provide data that are difficult to interpret. The use of numerical indices exclusive of the biological data from which they are calculated should be discouraged. 4.5 CONCLUSIONS A multilevel ecological risk assessment, which makes use of a combination of organism, population, and ecosystem-level endpoints, provides the most effective approach to examining ecosystem stress. Multilevel testing would both enhance the sensitivity of a risk assessment and broaden its focus to more complex levels of ecological organization. In contrast, the traditional approach of using only single-species testing is generally inadequate to account for the pollutant-induced effects on the complex organization of the ecosystem. Supplementation of organism-level endpoints with carefully selected population- and ecosystem-level endpoints may greatly improve the accuracy of ecological risk assessments. It is difficult to precisely specify the most appropriate endpoints for use in an ecological risk assessment because the choice depends partially on both the ecosystem being stressed and the nature of the pollutant stressor. A variety of organism-, population-, and ecosystem-level endpoints are available to choose from. Among the factors that should be considered in the choice of endpoints are sensitivity to low pollutant levels, predictability of long-term stress to 149 ------- individual species or to ecosystems, and ease of measurement. Endpoints that appear to be of generally high value in assessing the ecological risk of pollutants include primary productivity, nutrient cycling, trophic structure, and growth and reproduction, particularly for keystone species. Further guidance in choosing endpoints is given in a related document, "Ecological Endpoint Selection Criteria" (TRI, 1988a). There is still, however, an inadequacy of field data documenting the usefulness of organism-, population-, and ecosystem-level endpoints in predicting ecological risk. Future research in this area would facilitate the development of the multilevel risk assessment approach. 150 ------- REFERENCES Addison, R.F., M.E. Zinck, and D.E. Willis. 1981. Time- and dose-response of hepatic mixed function oxidase activity in brook trout Salvelinus fontinalis on polychlorinated biphenyl residues: Implications for biological effects monitoring. Environ. Pollut. Ser. A. 25: 211-218. Alink, G.M., E.M.H. Frederix-Wolters, M.A. van der Gaag, J.F.J, van de Kerkhoff, and C.L.M. Poels. 1980. Induction of sister-chromatid exchange in fish exposed to Rhine water. Mutat. Res. 78: 369-374. American Public Health Association, Inc. 1985. Standard Methods for the Examination of Water and Wastewater Including Bottom Sediments and Sludges. Sixteenth Edition. 1268 pp. Anderson, J.W., J.M. Neff, and S.R. Petrocelli. 1974. Sublethal effects of oil, heavy metal and PCBs on marine organisms, pp. 83-121 in Khan, M.A.Q. and J.P. Bederka, eds., Survival in Toxic Environments. Academic Press, New York. Andrewartha, H.G., and L.C. Birch. 1954. The Distribution and Abundance of Animals. Univ. of Chicago Press. Antonovics, J., A.D. Bradshaw, and R.G. Turner. 1971. Heavy metal tolerance in plants. Adv. Ecol. Res. 7: 1-85. Atlas, R.M., and R. Bartha. 1981. Microbial ecology: Fundamentals and Applications. Addison-Wesley Publishing Company, Reading, MA. Babich, H., and G. Stotzky. 1980. Physicochemical factors that affect the toxicity of heavy metals in microbes in aquatic habitats, pp. 181-203 in Colwell, R.R., and J. Foster, eds., Aquatic Microbial Ecology. Proceedings of the Conference of the Amer. Soc. Microbiology. Maryland Sea Grant Publ. Univ. of Maryland, College Park, MD. Baddeley, M.S., B.W. Ferry, and E.I. Finegan. 1973. Sulphur dioxide and respiration in lichens, pp. 299-313 in Ferry, B.W., M.S. Baddeley, and D.L. Hawksworth, eds., Air Pollution and Lichens. Univ. of Toronto. Barica, J., and L.R. Mur, eds. 1979. Hypertrophic Ecosystems. Developments in Hydrobiology 2. Junk, The Hague. Barnthouse, L.W., R.V. O'Neill, S.M. Bartell, and G.W. Suter. 1986. Population and ecosystem theory in ecological risk assessment, pp. 82-96 in Poston, T.M. and R. Purdy, eds., Aquatic Toxicology and Environmental Fate, 9th Symposium. STP 921. Am. Soc. Testing Materials, Philadelphia. Bayne, B.L., J. Anderson, D. Engel, E. Gilfillan, D. Hoss, R. Loyd, and F. Thurberg. 1980. Physiological techniques of measuring the biological effects of pollution in the sea, in Mclntyre, A.D. and J.B. Pierce, eds., Biological Effects of Marine Pollution and the Problems of Monitoring., Rapp, P.-V. Reun, Cons. Int. Explor. Mer. 179: 88-89. 151 ------- Beadle, C.L., and S.P. Long. 1985. Photosynthesis - is it limiting to biomass production? Biomass. 8: 119-168. Bechtel, T.J., and B.J. Copeland. 1970. Fish species diversity indices as indicators of pollution in Galveston Bay, Texas. Contrib. Mar. Sci. 15: 103-132. Beijer, K., and A. Jernelov. 1978. Ecological aspects of mercury selenium interactions in the marine environment. Environ. Health Perspect. 25: 43-45. Belser, L.W., and E.L. Mays. 1982. Use of nitrifier, activity measurements to estimate the efficiency of viable nitrifier counts in soils and sediments. Appl. Environ. Microbiol. 43: 945-948. Bengtsson, B.E. 1979. Biological variables, especially skeletal deformities in fish, for monitoring marine pollution. Philos. Trans. R. Soc. Lond. B. Biol. Sci. 286: 457-464. Bengtsson, B.E., A. Bengtsson and M. Himberg. 1985. Fish deformities and pollution in some Swedish waters. Ambio. 14: 32-35. Bennett, J.H., and A.C. Hill. 1974. Acute inhibition of apparent photosynthesis by phytotoxic air pollutants, pp. 115-127 in Dugger, M., ed., Air Pollutant Effects on Plant Growth. ACS Symposium Series 3. Amer. Chem. Soc. Bergersen, F.J. 1980. Methods for Evaluating Biological Nitrogen Fixation. John Wiley and Sons, New York. Berner, T., Z. Dubinsky, F. Schanz, U.J. Grobbelaar, H. Rai, U. Uehlinger, and P.G. Falkowski. 1986. The measurement of primary productivity in a high-rate oxidation pond (HROP). J. Plankton Res. 8(4): 659-672. Beyers, R.J. 1963. The metabolism of twelve aquatic laboratory microecosystems. Ecol. Monogr. 33: 281-306. Billiard, R. 1978. Effects of heat pollution and organo-chlorinated pesticides on fish reproduction, pp. 265-267 in Final Reports on Research Sponsored Under the First Environmental Research Program, Commission of the European Communities, Publ. by CEC Brussels, IBN 99-825-0/85-C, Brussels, Belgium. (Cited in: Effects of Pollutants at the Ecosystem Level, P.J. Sheehan, D.R. Miller, G.C. Butler and P. Bourdeau, eds., John Wiley and Sons, New York, pp. 23-50, 1984.) Birge, W.J., J.A. Black, and R.A. Kuehne. 1980. Effects of organic compounds on amphibian reproduction. University of Kentucky, Water Resources Research Institute, Research Report 121, Lexington, Kentucky. 39 pp. (Cited in: Effects of Pollutants at the Ecosystem Level, P.J. Sheehan, D.R. Miller, G.C. Butler and P. Bourdeau, eds. John Wiley and Sons, New York, pp. 23-50, 1984.) Birge, W.J., J.A. Black, and D.M. Stern. 1979. Toxicity of organic chemicals to embryo-larval stages of fish. U.S. EPA-560-11-79-007; PB80-101637. 152 ------- Bishop, W.E., and L.C. Valentine. 1982. Use of the central mudminnow (Umbra limi) in the development and evaluation of a sister chromatid exchange test for detecting mutagens in vivo, pp. 99-108 in Pearson, J.C., R.B. Foster, and W.E. Bishop, eds., Aquatic Toxicology and Hazard Assessment: Fifth Symposium. STP 766. Am. Soc. Testing Materials, Philadelphia. Blus, L.J., C.D. Gish, A.A. Belisle and R.M. Prouty. 1972. Logarithmic relationship of DDE residues to eggshell thinning. Nature. 235: 376-377. Bollag, J.M., and W. Barabasz. 1979. Effect of heavy metals on the denitrification process in soil. J. Environ. Qual. 8(2): 196-201. Borgmann, U., R. Cove, and C. Loveridge. 1980. Effect of metals on the biomass production kinetics of freshwater copepods. Can. J. Fish. Aquat. Sci. 37: 567-575. Bormann, F.H., and G.E. Likens. 1967. Nutrient cycling. Science. 155: 424-428. Bormann, F.H., and G.E. Likens. 1970. The nutrient cycles of an ecosystem. Scientific American 223(4): 92-101. Bowmer, T., R.G.V. Boelens, B.F. Keegan, and J.O. O'Neill. 1986. The use of marine benthic "key" species in ecotoxicological testing: Amphiura filiformis. Aquat. Toxicol. (NY) 8: 93-109. Boyce, M.S. 1977. Population growth and stochastic fluctuations in the life table. Theor. Popul. Biol. 12: 366-373. Brinson, M.M., A.E. Lugo, and S. Brown. 1981. Primary productivity decomposition and consumer activity in freshwater wetlands. Annu. Rev. Ecol. Syst. 12: 123-161. Bull, C., and J. Mclnemey. 1974. Behavior of juvenile coho salmon Oncorhvnchus kisutch exposed to Sumithion (fenitrothion) an organophosphate insecticide. J. Fish Res. Board Can. 31: 1867-1872. Busby, D.G., P.A. Pearce, and N.R. Garrity. 1981. Brain cholinesterase response in songbirds exposed to experimental fenitrothion spraying in New Brunswick, Canada. Bull. Environ. Contam. Toxicol. 26: 401-406. Cairns, J., ed. 1986. Community Toxicity Testing. STP 920, Am. Soc. Testing Materials, Philadelphia. Cairns, J., and K.L. Dickson. 1977. Recovery of streams and spills of hazardous materials. pp. 24-44 in Cairns, J., K.L. Dickson, and E.E. Henricks, eds., Recovery and Restoration of Damaged Ecosystems. Univ. of Virginia Press. Call, D.J., L.T. Brooke, M.L. Knuth, S.H. Poirier, and M.D. Hoglund. 1985. Fish subchronic toxicity prediction model for industrial organic chemicals that produce narcosis. Environ. Toxicol. Chem. 4: 335-341. Campbell, P.G.C., and P.M. Stokes. 1985. Acidification and toxicity of metals to aquatic biota. Can. J. Fish Aquat. Sci. 42: 2034-2049. 153 ------- Carlson, R.W. 1979. Reduction in the photosynthetic rate of Acer auercus and fraxinus species caused by sulphur dioxide and ozone. Environ. Pollut. 18: 159- 170. Christensen, E.R., and N. Nyholm. 1984. Ecotoxicological assays with algae: Weibull dose-response curves. Environ. Sci. Technol. 18: 713-718. Clapham, W.B. 1973. Natural Ecosystems. Case Western Reserve Univ., Cleveland, OH. Colbourn, P., M.M. Iqbal, and I.W. Harper. 1984. Estimation of the total gaseous nitrogen losses from clay soils under laboratory and field conditions. J. Soil. Sci. 35: 11-22. Cole, D.W., S.P. Gessel, and E.E. Held. 1961. Tension lysimeter studies of ion and moisture movement in glacial till and coral atoll soils. Soil Sci. Soc. Am. Proc. 25:321-325. Connell, J.H. 1978. Diversity in tropical rain forests and coral reefs. Science 199: 1302-1310. Constantinidou, H.A., and T.T. Kozlowski. 1979. Effects of sulfur dioxide and ozone on Ulmus americana seedlings. I. Visible injury and growth. Can. J. Bot. 57: 170-175. Cook, R.B. 1984. Man and the biogeochemical cycles: Interacting with the elements. Environ. 26(7): 11-40. Cooke, A.S. 1972. The effects of DDT, dieldrin and 2,4,-D on amphibian spawn and tadpoles. Environ. Pollut. 3: 51-68. Cooke, A.S. 1973. Shell thinning in avian eggs by environmental pollutants. Environ. Pollut. 4: 85-152. Cooke, A.S. 1981. Tadpoles as indicators of harmful levels of pollution in the field. Environ. Pollut. Ser. A. Ecol. Biol. 25: 123-133. Cooper, D.C., and B.J. Copeland. 1973. Responses of continuous-series estuarine microecosystems to point-source input variations. Ecol. Monographs 43: 213-236. Couch, J.A., and J.C. Harshbarger. 1985. Effects of carcinogenic agents on aquatic animals: An environmental and experimental overview. Environ. Carcin. Revs. 3: 63-105. Cummins, K.W. 1974. Structure and function of stream ecosystems. Bioscience. 24: 631-641. Czuba, M., and D.P. Ormrod. 1984. Cadmium concentration in cress shoots in relation to cadmium-enhanced ozone phytotoxicity. Environ. Pollut. 25: 67-76. 154 ------- Davey, F.B., H. Kleerekoper, and P. Gensler. 1972. Effect of exposure to sublethal DDT on the locomotor behavior of the goldfish (Carassius auratus). J. Fish Res Board Can. 29: 1333-1336. Dawson, J.L., and T.H. Nash. 1980. Effects of air pollution from copper smelters on a desert grassland community. Environ. Exp. Bot. 20: 61-72. Donaldson, E.M. 1981. The pituitary-interrenal axis and as indicator of stress in fish. pp. 11-17 in Pickering, A.D., ed., Stress and Fish. Academic Press, New York. Donaldson, E.M., and H.M. Dye. 1975. Corticosteroid concentration in sockeye salmon (Oncorhvnchus nerka) exposed to low concentration of copper. J. Fish Res. Board Can. 32: 533-539. Drummond, R.A., W.A. Spoor, and G.F. Olson. 1973. Some short-term indicators of sublethal effects of copper on brook trout, Salvelinus fontinalis. J. Fish Res. Board Can. 30: 698-701. Edwards, R.H., and R.M. Overstreet. 1976. Mesenchymal tumors of some estuarine fishes of the northern Gulf of Mexico. I. Subcutaneous tumors, probably fibrosarcomas, in the striped mullet, Mueil cephalus. Bull. Mar. Sci. 26(1): 33-40. Eisenberg, J.F. 1980. The density and biomass of tropical mammals, pp. 35-55 in Soule, M.E. and B.A. Wilcox, eds., Conservation Biology, an Evolutionary- Ecological Perspective. Sinauer Assoc. Publ., Sunderland, MA. Elton, C. 1933. The Ecology of Animals. Methuen, London. Estes, J.A., and J.F. Palmisano. 1974. Sea otters: Their role in structuring nearshore communities. Science 185: 1058-1060. Farr, J.A. 1978. The effect of methyl parathion on predator choice of two estuarine prey species. Trans. Am. Fish Soc. 107: 87-91. Feder, W.A. 1973. Cumulative effects of chronic exposure of plants to low levels of air pollutants, in Naegele, J.A., ed., Air Pollution Damage to Vegetation. Adv. Chem. Ser. 122: 21-30. Fenchel, T., and T.H. Blackburn. 1979. The nitrogen cycle. pp. 101-126 in Fenchel, T., and T.H. Blackburn, eds., Bacteria and Mineral Cycling. Academic Press, New York. Fitzwater, S.E., G.A. Knauer, and J.H. Martin. 1982. Metal contamination and its effects of primary production measurements. Limnol. Oceanogr. 27(3): 544-551. Fleming, W.J., M.A.R. McLane, and E. Cromartie. 1982. Endrin decreases screech owl productivity. J. Wildl. Manage. 46: 462-468. Folmar, L.C. 1976. Overt avoidance reaction of rainbow trout fry to nine herbicides. Bull. Environ. Contam. Toxicol. 15: 509-514. 155 ------- Forbes, A.M., and J.J. Magnuson. 1980. Decomposition and microbial colonization and leaves in a stream modified by coal ash effluent. Hydrobiology 76: 263-267. Fowler, C.W. 1981. Density dependence as related to life history strategy. Ecology 62(3): 602-610. Fox, G.A., A.P. Gilman, D.B. Peakall, and F.W. Anderka. 1978. Behaviorial abnormalities of nesting Lake Ontario herring gulls. J. Wildl. Manage. 42: 477-483. Franson, J.C., J.W. Spann, G.H. Heinz, C. Bunck, and T. Lamont. 1983. Effects of dietary ABATE on reproductive success, duckling survival, behavioral, and clinical pathology in game-farm mallards. Arch. Environ. Contam. Toxicol. 12: 529-534. Freedman, B., and T.C. Hutchinson. 1980a. Long-term effects of smelter pollution at Sudbury, Ontario on forest community composition. Can, J. Bot. 58: 2123-2140. Freedman, B., and T.C. Hutchinson. 1980b. Smelter pollution near Sudbury, Ontario, Canada, and effects on forest litter decomposition. pp. 395-434 in Hutchinson, T.C., and M. Havas, eds., Effects of Acid Precipitation on Terrestrial Ecosystems. Plenum Publ., New York. Fulton, R.S., and H.W. Paerl. 1987. Toxic and inhibitory effects of the blue- green alga Microcystis aeruginosa to herbivorous zooplankton. J. Plankton Res. 9: 837-855. Gaarder, T., and H.H. Gran. 1927. Production of plankton in the Oslo Fjord. Rapp. Proc. Verb. Cons. Perm. Int. Explor. Mer. 42: 1-48. Garten, C.T., Jr., G.W. Suter II, and B.G. Blaylock. 1985. Development and evaluation of multispecies test protocols for assessing chemical toxicity, ORNL/TM-9225. Martin Marietta, Oak Ridge National Laboratory, Oak Ridge, TN. Gentile, J.H., S.M. Gentile, N.G. Hairston, Jr., and B.K. Sullivan. 1982. The use of life-tables for evaluating the chronic toxicity of pollutants to Mvsidopsis bahia. Hydrobiologia 93: 179-187. Giddings, J., and G.K. Eddlemon. 1978. Photosynthesis respiration ratios in aquatic microcosms under arsenic stress. Water, Air and Soil Pollut. 9: 207-212. Goh, K.M., and R.J. Haynes. 1986. Nitrogen and agronomic practice, pp. 379-468 in Haynes, R.J., ed., Mineral Nitrogen in the Plant System. Acad. Press, Orlando, FL. Goodyear, P.C. 1972. A simple technique for detecting effects of toxicants or other stresses on a predator-prey interaction. Trans. Am. Fish Soc. 101: 367-370. Gorden, R.W. 1972. Field and Laboratory Microbial Ecology. Wm. C. Brown Co., Publishers, Dubuque, IA. Gorden, R.W, R.J. Beyers, E.P. Odum, and R.G. Eagon. 1969. Studies of a simple laboratory microecosystem: Bacterial activities in a heterotrophic succession. Ecology 50: 86-100. 156 ------- Gray, J.S. 1979. Pollution-induced changes in populations. Philos. Trans. R. Soc. Lond. 286: 545-561. Gray, J.S., and T.H. Pearson. 1982. Objective selection of sensitive species indicative of pollution-induced change in benthic communities. I. Comparative methodology. Mar. Ecol. Prog. Ser. 9: 111-119. Green, R.H. 1979. Sampling design and Statistical Methods for Environmental Biologists. John Wiley and Sons, New York. Grime, J.P. 1977. Evidence for the existence of three primary strategies in plants and relevance to ecological and evolutionary theory. Am. Natl. Ill: 1169- 1194. Haegele, M.A., and R.H. Hudson. 1977. Reduction of courtship behavior induced by DDE in male ringed turtle doves. Wilson Bull. 89: 593-601. Hall, G.H. 1984. Measurement of nitrification rates in lake sediments: Comparison of the nitrification inhibitors nitrapyrin and allylthiourea. Microb. Ecol. 10: 25-36. Hamaker, J.W. 1975. The interpretation of soil leaching experiments in Haque, pp. 115-133 in Haque, R., and V.H. Freed, eds., Environmental Science Research, Vol. 6. Environmental Dynamics of Pesticides. Plenum Press, New York. Harrison, F.L., and I.M. Jones. 1982. An in vivo sister-chromatid exchange assay in the larvae of the mussel Mvtilus edulis: Response to 3 mutagens. Mutat. Res. 105: 235-242. Hatakeyama, S., and M. Yasuno. 1981. Effects of cadmium on the periodicity of parturition and blood size Moina macrocopa (Cladocera). Environ. Pollut. Ser. A. Ecol. Biol. 26: 111-120. Havas, M. 1985. Aluminum bioaccumulation and toxicity to Daohnia magna in soft water at low pH. Can. J. Fish. Aquat. Sci. 42: 1741-1748. Havas, M., and T.C. Hutchinson. 1982. Aquatic invertebrates from the Smoking Hills, N.W.T.: Effect of pH and metals on mortality. Can. J. Fish Aquat. Sci. 39: 890-903. Haynes, R.J., and R.R. Sherlock. 1986. Gaseous losses of nitrogen, pp. 242-302 in Haynes, R.J., ed., Mineral Nitrogen in the Plant-Soil System. Academic Press, Orlando, FL Heath, R.L. 1975. Ozone, pp. 23-55 in Mudd, J.B. and T.T. Kozlowski, eds., Response of Plants to Air Pollution. Academic Press, New York. Heinz, G.H. 1976. Behavior of mallard ducklings from parents fed 3 ppm DDE. Bull. Environ. Contam. Toxicol. 16: 640-645. Heinz, G.H. 1979. Methylmercury: Reproductive and behaviorial effects on three generations of mallard ducks. J. Wildl. Manage. 43(2): 394-401. 157 ------- Heinz, G.H., S.D. Haseltine, W.L. Reichel, and G.L. Hensler. 1983. Relationships of environmental contaminants to reproductive success in red-breasted mergansers Mergus serrator from Lake Michigan. Environ. Pollut. Ser. A. Ecol. and Biol. 32: 211-232. Hellawell, J.M. 1977. Change in natural and managed ecosystems: Detection, measurement and assessment. Proc. R. Soc. B. Biol. Sci. 197: 31-56. Hendry, G.R., K.. Baalsrud, T.S. Traaen, M. Laake, and G. Raddum. 1976. Acid precipitation: Some hydrobiological changes. Ambio 5: 224-227. Herricks, E.E., and J. Cairns, Jr. 1982. Biological monitoring. Part Ill-Receiving system methodology based on community structure. Water Res. 16: 141-153. Hetrick, F.M., M.D. Knittel, and J.L. Fryer. 1979. Increased susceptibility of rainbrow trout to infectious hematopoietic necrosis virus after exposure to copper. Appl. Environ. Microbiol. 37: 198-201. Hixon, M.A., and W.N. Brostoff. 1983. Damselfish as keystone species in reverse: intermediate disturbance and diverity of reef algae. Science 220: 511-513. Hoch, G., O.V.H. Owens, and B. Kok. 1963. Photosynthesis and respiration. Arch. Biochem. Biophys. 101: 171-180. Hoffman, D.J., O.H. Pattee, S.N. Wiemeyer, and B. Mulhern. 1981. Effects of lead shot ingestion on delta-aminolevulinic acid dehydratase activity, hemoglobin concentration, and serum chemistry in bald eagles. J. Wildl. Dis. 17: 423-431. Hoffman, D.J., L. Sileoa, and H.C. Murray. 1984. Subchronic organophosphorus ester-induced delayed neurotoxicity in mallards. Toxicol. Appl. Pharmacol. 75: 128-136. Holcombe, G.W., D.A. Benoit, E.N. Leonard, and J.M. McKim. 1976. Long-term effects of lead exposure on three generations of brook trout (Salvelinus fontinalis). J. Fish Res. Board Can. 33: 1731-1741. Holling, C.S. 1973. Resilence and stability of ecological systems. Annu. Rev. Ecol. Syst. 4: 1-23. Home, A.J. and C.R. Goldman. 1974. Suppression of nitrogen fixation by blue- green algae in a eutrophic lake with trace additions of copper. Science 83: 409- 411. Hose, J.E., J. Hannah, H. Puffer, and M.L. Landolt. 1984. Histologic and skeletal abnormalities in benzo(a)pyrene-treated rainbow trout alevins. Arch. Environ. Contam. Toxicol. 13: 675-684. Hughes, P.R., J.E. Potter, and L.H. Weinstein. 1982. Effects of air pollution on plant/insect interactions: Increased susceptibility of greenhouse grown soybeans to the Mexican bean beetle following plant exposure to SO. Environ. Entmol. 11: 173-176. 158 ------- Hurlbert, S.H. 1975. Secondary effects of pesticides on aquatic ecosystems. Residue Rev. 57: 81-148. Hurlbert, S.H. 1984. Pseudoreplication and the design of ecological field experiments. Ecol. Monogr. 54: 187-211. Hutton, M. 1980. Metal contamination of feral pigeons Columbia livia from the London area: Part 2. Biological effects of lead exposure. Environ. Pollut. Ser. A. Ecol. Biol. 22: 281-293. Hynes, R.K., and R. Knowles. 1983. Inhibition of chemoautotrophic nitrification by sodium chlorate and sodium chlorite: a reexamination. Appl. Environ. Microbiol. 45: 1178-1182. Jackim, R. 1973. Influence of lead and other metals on fish delta-aminolevulinate dehydrase activity. J. Fish Res. Board Can. 30: 560-562. Jensen, A.L., and J.S. Marshall. 1983. Toxicant-induced fecundity compensation: A model of population responses. Environ. Manage. 7(2): 171-175. Johnson, R.A. 1981. Application of the guild concept to environmental impact analysis of terrestrial vegetation. J. Environ. Management 13: 205-222. Jones, M.B. 1975. Synergistic effects of salinity, temperature and heavy metals on mortality and osmoregulation in marine and estuarine isopods (Crustacea). Mar. Biol. 30: 13-20. Kania, H.J., and J. O'Hara. 1974. Behaviorial alterations in a simple predator- prey system due to sublethal exposure to mercury. Trans. Am. Fish. Soc. 103: 134-136. Keeney, D.R. 1980. Prediction of soil nitrogen availability in forest ecosystems: A literature review. Forest Sci. 26: 159-171. Keeney, D.R. 1982. Nitrogen-availability indices, pp. 71 1-733 in Page, C.A.L., ed., Methods of Soil Analysis. Part 2. Chemical and Microbiology Properties. Am. Soc. Agron., Madison, WI. Kelso, J.R.M. 1977. Density, distribution and movement of Nipigon Bay Fishes in relation to a pulp and paper mill effluent. J. Fish Res. Board Can. 34: 879-885. King, K.A., D.H. White, and C.A. Mitchell. 1984. Nest defense behavior and reproductive success of laughing gulls sublethally dosed with parathion. Bull. Environ. Contam. Toxicol. 33: 499-504. King, P.H., and P.L. McCarty. 1968. A chromatographic model for predicting pesticide migration in soil. Soil Sci. 106: 248. Koenig, C.C., R.J. Livingston, and C.R. Cripe. 1976. Blue crab mortality: Interaction of temperature and DDT residues. Arch. Environ. Contam. Toxicol. 4: 119-128. 159 ------- Koike, I., and A. Hattori. 1978. Simultaneous determinations of nitrification and IS nitrate reduction in coastal sediments by a JJN dilution technique. Appl. Environ. Microbiol. 35: 853-857. Kormondy, E.J. 1969. Concepts of Ecology. Prentice-Hall Biological Science Series. Prentice-Hall, Inc., Englewood Cliffs, NJ. Krebs, C.J. 1985. Ecology: The Experimental Analysis of Distribution and Abundance. 3rd Edition. Harper and Row, New York. Kynard, B. 1974. Avoidance behavior of insecticide susceptible and resistant populations of mosquitofish to four insecticides. Trans. Am. Fish Soc. 103: 557- 561. Lack, D. 1954. The Natural Regulation of Animal Numbers. Clarendon Press, Oxford. Landres, P.B. 1983. Use of the guild concept in environmental impact assessment. Environ. Manage. 7(5): 393-398. Leivestad, H., G. Hendry, I.P. Muniz, and E. Snekvik. 1976. Effects of acid precipitation on freshwater organisms, pp. 87-111 in Braekke F.H., ed., Impact of Acid Precipitation on Forest and Freshwater Ecosystems in Norway. Research Rep. 6/76. The SNSF Project. Levin, S.A., K.D. Kimball, W.H. McDowell, and S.F. Kimball. 1984. New perspectives in ecotoxicology. Environ. Manage. 8(5): 375-442. Liang, C.N., and M.A. Tabatabai. 1978. Effects of trace elements on nitrification in soils. J. Environ. Qual. 7(2): 291-293. Lighthart, B. 1980. Effects of certain cadmium species on pure and litter populations of microorganisms. Antonie van Leeuwenhoek. 46: 161-167. Likens, G.E., F.H. Bormann, N.M. Johnson, D.W. Fisher, and R.S. Pierce. 1970. Effects of forest cutting and herbicide treatment on nutrient budgets in the Hubbard Brook watershed ecosystem. Ecol. Monogr. 40(1): 23-47. Limburg, K.E., C.C. Harwell, and S.A. Levin. 1984. Principles for estuarine impact assessment: Lessons learned from the Hudson River and other estuarine experiences. Ecosystem Res. Center. Cornell Univ., Ithaca, NY. ECR-024. 75 pp. Livingston, R.J. 1977. Review of current literature concerning the acute and chronic effects of pesticides on aquatic organisms. CRC Crit. Rev. Environ. Control. 7: 325-351. Lomnicki, A., E. Bandola, and K. Kankowska. 1968. Modification of the Wiegert- Evans method for estimation of net primary production. Ecology. 49(1): 147-149. Lorz, H.W., and B.R. McPherson. 1976. Effects of copper or zinc in fresh water on the adaption to sea water and ATPase activity, and the effects of copper on 160 ------- migratory disposition of Coho salmon (Qncorhvnchus kisutch). J. Fish Res. Board Can. 33: 2023-2030. Lotti, M., and M.K. Johnson. 1978. Neurotoxicity of organophosphorus pesticides: Predictions can be based on in vitro studies with hen and human enzyme. Arch. Toxicol. 41: 215-221. Ludke, J.L., E.F. Hill, and M.P. Dieter. 1975. Cholinesterase (ChE) response and related mortality among birds fed ChE inhibitors. Arch. Environ. Contam. Toxicol. 3: 1-21. Macek, K., W. Birge, F.L. Mayer, A.L. Buikema, Jr., and A.W. Maki. 1978. Discussion session synposis, pp. 27-32 in Cairns, J., K.L. Dickson, and A.W. Maki, eds., Estimating the Hazard of Chemical Substances to Aquatic Life. Am. Soc. Testing Materials, Philadelphia, PA. Maki, A.W. 1979. Respiratory activity of fish as a predictor of chronic fish toxicity values for surfactants, pp. 77-95 in Marking, L.L., and R.A. Kimmerle, eds., Aquatic Toxicology. ASTM STP 667. Amer. Soc. for Testing Materials, Philadelphia, PA. Maki, A.W., and H.E. Johnson. 1976. Evaluation of a toxicant on the metabolism of model stream communities. J. Fish. Res. Board Can. 33: 2740-2746. Martin, A. 1983. Assessment of the effects of chemicals on the reproductive functions of reptiles and amphibians, pp. 405-413 in Vouk, V.B., and P.J. Sheehan, eds., Methods for Assessing the Effects of Chemicals on Reproductive Functions. John Wiley and Sons, New York. Mauck, W., P. Mehrle, and F.L. Mayer. 1978. Effects of polychlorinated biphenyl Aroclor 1254 on growth, survival and bone development in brook trout (Salvelinus fontinalisl. J. Fish Res. Board Can. 35: 1084-1088. May, R.M. 1986. The search for pattern in the balance of nature: Advances and retreats. Ecology 67: 1115-1126. McCain, B.B., K.V. Pierce, S.R. Willings, and B.S. Miller. 1977. Hepatomas in marine fish from an urban estuary. Bull. Environ. Contam. Toxicol. 18: 1-2. McClenahen, J.R. 1978. Community changes in a deciduous forest exposed to air pollution. Can. J. For. Res. 8: 432-438. McGreer, E.R. 1979. Sublethal effects of heavy metal contaminated sediments on the bivalve Macoma balthica (L.). Mar. Pollut. Mar. 10: 259-262. McKim, J.M. 1985. Early life stage toxicity tests, pp. 58-95 in Rand, G.M., and S.R. Petrocelli, eds., Fundamentals of Aquatic Toxicology. Hemisphere Publ. Co., Washington, DC. McKim, J.M., and D.A. Benoit. 1971. Effects of long-term exposures to copper on survival, growth, and reproduction of brook trout (Salvelinus fontinalis). J. Fish Res. Board Can. 28: 655-662. 161 ------- McLaughlin, S., and O.U. Braker. 1985. Methods for evaluating and predicting forest growth responses to air pollution. Experientia 41: 310-319. McLaughlin, S.B. 1985. Effects of air pollution on forests. A critical review. J. Air Pollut. Control Assoc. 35: 512-534. McLenahen, J.R. 1978. Community changes in a deciduous forest exposed to air pollution. Can. J. For. Res. 8: 432-438. McNaughton, S.J. 1977. Diversity and stability of ecological communities. A comment on the role of empiricism in ecology. Amer. Nat. Ill: 515-525. McNaughton, S.J., and L.L. Wolf. 1979. General Ecology. Holt, Rinehart and Winston, New York. pp. 1-702. Mearns, A.J., and M.J. Sherwood. 1974. Environmental aspects of fin erosion and tumors in southern California dover sole. Trans. Am. Fish Soc. 103: 799-810. Meyers, T.R., and J.D. Hendricks. 1982. A summary of tissue lesions in aquatic animals induced by controlled exposures to environmental contaminants, chemotherapeutic agents, and potential carcinogens. Mar. Fish Rev. 44: 1-17. Miles, L.J., and G.R. Parker. 1980. Effects of cadmium and a one-time drought stress on survival, growth, and yield of native plant species. J. Environ. Qual. 9: 278-282. Miller, J.E., J.J. Hassett, and D.E. Koeppe. 1977. Interaction of lead and cadmium on metal uptake and growth of plant species. J. Environ. Qual. 6: 18-20. Monk, D.C. 1983. The uses and abuses of ecotoxicology. Mar. Pollut. Bull. 14(8): 284-288. Moore, J.C., and H.W. Hunt. 1988. Resource compartmentation and the stability of real ecosystems. Nature 333: 261-263. Moore, M.N., and A.R.D. Stebbing. 1976. The quanitative cytochemical effects of three metals on a lysosomal hydrolase of a hydroid. J. Mar. Biol. Ass. U.K. 56: 995-1005. Moriarty, F. 1983. Ecotoxicology. Academic Press, New York. Murray, B.G. 1979. Population Dynamics: Alternative Models. Academic Press, New York. Newbould, P.J. 1967. Methods for Estimating the Primary Production of Forests. IBP Handbook No. 2. Blackwell Scientific Publications. Oxford. Nur, N. 1984. The consequences of brood size for breeding blue tits. II. Nestling weight, offspring survival and optimal brood size. J. Anim. Ecol. 53: 497-517. 162 ------- O'Neill, R.V., B.S. Ausmus, D.R. Jackson, R.I. Van Hook, P. Van Voris, C. Washburne, and A.P. Watson. 1977. Monitoring terrestrial ecosystems by analysis of nutrient export. Water Air Soil Pollut. 8: 271-277. Odum, E.P. 1971. Fundamentals of Ecology. W.B. Saunders Co., Philadelphia, PA. pp. 1-574. Pagano, G., M. Cipollaro, G. Corsale, A. Esposito, E. Ragucci, G.G. Giordano, and N.M. Trieff. 1986. The sea urchin: Bioassay for the assessment of damage from environment contaminants, pp. 66-91 in Cairns, J., ed., Community Toxicity Testing, A Symposium Sponsored by ASTM Committee D-19 on Water. Colorado Springs, CO. 6-7 May 1985. STP 920. Am. Soc. Testing Materials, Philadelphia, PA. Paine, R.T. 1966. Food web complexity and species diversity. Am. Nat. 100(910): 65-75. Paine, R.T. 1974. Intertidal community structure: experimental studies on the relationship between a dominant competitor and its principal predator. Oecologia. 15: 93-120. Patil, S.G., D.K. Harshey, and D.F. Singh. 1985. The effects of DDT on primary production in a pond ecosystem. Comp. Physiol. Ecol. 10: 139-140. Peakall, D.B. 1983. Methods for assessment of the effects of pollutants on avian reproduction, pp. 345-363 in Vouk, V.B., and P.J. Sheehan, eds., Methods of Assessing the Effects of Chemicals on Reproductive Functions. SCOPE 20. John Wiley and Sons, New York. Pearson, T.H., J.S. Gray, and P.J. Johannessen. 1983. Objective selection of sensitive species indicative of pollution-induced change in benthic communities. 2. Data analyses. Mar. Ecol. Prog. Ser. 12: 237-255. Peet, R.K. 1974. The measurement of species diversity. Ann. Rev. Ecol. Syst. 5: 285-307. Pesch, G.G., and C.E. Pesch. 1980. Neanthes arenaceodentata (Polychaeta: Annelida), a proposed model for marine genetic toxicology. Can. J. Fish Aquat. Sci. 37: 1225-1228. Petersen, R.C. 1986. Population and guild analysis for interpretation of heavy metal pollution streams, pp. 180-198 in Cairns, J. ed., Community Toxicity Testing. STP 920. Am. Soc. Testing Materials, Philadelphia, PA. Peterson, B.J. 1980. Aquatic primary productivity and the ^C-C02 method: A history of the productivity problem. Annu. Rev. Ecol. Syst. 11: 359-385. Pielou, E.C. 1966. The measurement of diversity in different types of biological collections. J. Theor. Biol. 13: 131-144. Pimm, S.L. 1984. The complexity and stability of ecosystems. Nature 307: 321 - 326. 163 ------- Poysa, H. 1983. Resource utilization pattern and guild structure in a waterfowl community. Oikos. 40: 295-307. Rattner, B.A., V.P. Eroschenko, G.A. Fox, D.M. Fry, and J. Gorsline. 1984. Avian endocrine response to environmental pollutants. J. Exp. Zool. 232: 683-689. Read, P.A., T. Renshaw, and K.J. Anderson. 1978. Pollution effects on intertidal macrobenthic communities. J. Appl. Ecol. 15: 15-31. Reinert, R.A., and H.W. Spurr. 1972. Differential effects of fungicides on ozone injury and brown spot disease of tobacco. J. Environ. Qual. 1: 450-452. Renfro, J.L., B. Schmidt-Nielsen, D. Miller, D. Benos, and J. Allen. 1974. Methyl mercury and inorganic mercury: Uptake, distribution and effect on osmoregulatory mechanisms in fishes, pp. 101-122 in Vernberg, F.J., and W.B. Vernberg, eds., Pollution and Physiology of Marine Organisms. Academic Press, New York. Reynolds, C.S., and A.E. Walsby. 1975. Water-blooms. Biol. Rev. 50: 437-481. Ricklefs, R.E. 1973. Ecology. Chiron Press, New York. Ricklefs, R.E. 1987. Community diversity: Relative roles of local and regional processes. Science 235: 167-171. Rodsaether, C.M., J.A. Olafsen, J. Raa, K. Myhre, and J.B. Steen. 1977. Copper as initiating factor of vibriosis (Vibrio aneuillarum) in eel (Anguilla anguilla). J. Fish Biol. 10: 17-21. Rolston, D.E., A.N. Sharpley, O.W. Toy, and F.E. Broadbent. 1982. Field measurements of denitrification. III. Rates during irrigation cycles. Soil Sci. Soc. Am. J. 42: 863-869. Root, R.B. 1967. The niche exploitation pattern of the blue-gray gnatcatcher. Ecol. Monogr. 37(4): 318-350. Rose, F.L., and J.C. Harshbarger. 1977. Neoplastic and possible related skin lesions in neotenic tiger salamanders from a sewage lagoon. Science 196: 315-317. Rosenberg, R., and J.D. Costlow. 1976. Synergistic effects of cadmium and salinity with constant and cycling temperatures on the larval development of two estuarine crab species. Mar. Biol. 38: 291-303. Royce, W.F. 1972. Introduction to the Fishery Sciences. Academic Press, New York. Sahrawat, K.L. 1983. Nitrogen availability indexes for submerged rice soils. Adv. Agron. 36: 415-451. Sakamoto, M., M.M. Tilzer, R. Gachter, H. Rai, Y. Collos, P. Tschumi, P. Berner, D. Zbaren, J. Zbaren, M. Dokulil, P. Bossard, U. Uehlinger, and E.A. Nusch. 1984. Joint field experiments for comparisons of measuring methods of photosynthetic production. J. Plankton Res. 6(2): 365-383. 164 ------- Saunders, R.L., and J.B. Sprague. 1967. Effects of copper-zinc mining pollution on a spawning migration of Altantic salmon (Salmo salar). Water Res. 1: 419-432. Schindler, D.W. 1987. Detecting ecosystem responses to anthropogenic stress. Can. J. Fish Aquat. Sci. 44(Suppl. 1): 6-25. Schindler, D.W., K.H. Mills, D.F. Malley, D.L. Findlay, J.A. Shearer, I.J. Davis, M.A. Turner, G.A. Linsey, and D.R. Cruikshank. 1985. Long-term ecosystem stress: the effects of years of experimental acidification on a small lake. Science 228: 1395-1401. Schmidt, R.H. 1986. Community-level effects of coyote population reduction, pp. 49-65 in Cairns, J., Jr., ed., Community Toxicity Testing. STP 920. Am. Soc. Testing Materials, Philadelphia, PA. Schofield, C.L. 1976. Acid precipitation effects on fish. Ambio 5: 228-230. Schreck, C.B. 1981. Stress and compensation in teleostean fishes: response to social and physical factors, pp. 295-321 in Pickering, A.D., ed., Stress and Fish. Academic Press, New York. Severinghaus, W.D. 1981. Guild theory development as a mechanism for assessing environment impact. Environ. Manage. 5(3): 187-190. Sheehan, P.J. 1984a. Effects on individuals and populations, pp. 23-50 in Sheehan, P.J., D.R. Miller, G.C. Butler, and P. Bourdeau, eds., Effects of Pollutants at the Ecosystem Level. John Wiley and Sons, New York. Sheehan, P.J. 1984b. Effects on community and ecosystem structure and dynamics, pp. 51-99 in Sheehan, P.J., D.R. Miller, G.C. Butler and P. Bourdeau, eds., Effects of Pollutants at the Ecosystem Level. John Wiley and Sons, New York. Sheehan, P.J. 1984c. Functional changes in the ecosystem, pp. 101-145 in Sheehan, P.J., D.R. Miller, G.C. Butler, and P. Bourdeau, eds., Effects of Pollutants at the Ecosystem Level. John Wiley and Sons, New York. Silvester, W.B. 1983. Analysis of nitrogen fixation, pp. 173-212 in Gordon, J.C., and C.T. Wheeler, eds., Biological Nitrogen Fixation in Forest Ecosytems: Foundations and Applications. Nijhoff/Junk Publ., The Hague. Sindermann, C.T. 1980. The use of pathological effects of pollutants in marine environment monitoring programs in Mclntyre, A.D., and J.B. Pearce, eds., Biological Effects of Marine Pollution and the Problems of Monitoring. Rapp. P.- V. Reun. Cons. Int. Explor. Mer. 179: 135-151. Sladecek, V. 1963. The future of the saprobity system. Hydrobiologia 25: 518- 537. Smith, V.H. 1983. Low nitrogen to phosphorus ratios favor dominance by blue- green algae in lake phytoplankton. Science 221: 669-671. 165 ------- Solbe, J.F. 1977. Water quality, fish and invertebrates in a zinc polluted stream, pp. 97-105 in Alabaster, J.S., ed., Biological monitoring of Inland Fisheries. Applied Science Publ., London. Sousa, W.P. 1984. The role of disturbance in natural communities. Annu. Rev. Ecol. Syst. 15: 353-391. Southwood, T.R.E. 1978. Ecological Methods. Chapman and Hall. Sprague, J.B., P.F. Elson, and R.L. Saunders. 1965. Sublethal copper-zinc pollution in a salmon river - a field and laboratory study, pp. 65-82 in Jaag, O., ed., Advances in Water Pollution Research. Pergamon Press, New York. (Cited in: Sheehan, P.J., D.R. Miller, G.C. Butler, and P. Bordeau, eds., 1984. Effects of Pollutants at the Ecosystem Level. John Wiley and Sons, New York. Stanford, G. 1982. Assessment of soil nitrogen availability, pp. 651-688 in Stevenson, F.J., ed., Nitrogen in Agricultural Soils. Am. Soc. Agron., Madison, WI. Stauffer, J.R., and C.H. Hocutt. 1980. Inertia and recovery: An approach to stream classification and stress evaluation. Water Res. Bull. 16(1): 72-78. Stegeman, J.J. 1980. Mixed-function oxygenase studies in monitoring for effects of organic pollution, in Mclntyre, A.D., and J.B. Pearce, eds., Biological Effects of Marine Pollution and Problems of Monitoring. Rapp. P-V Reun. Cons. Int. Explor. Mer. 179: 33-38. Stewart-Oaten, A., W.W. Murdoch, and K.R. Parker. 1986. Environmental impact assessment: "Pseudoreplication" in time? Ecology 67: 929-940. Stokes, P.M. 1986. Ecological effects of acidification on primary producers in aquatic systems. Water Air Soil Pollut. 30(1-2): 421-438. Stoner, A.W., and R.J. Livingston. 1978. Respiration, growth and food conversion efficiency and pinfish (Laeodon rhomboides) exposed to sublethal concentrations of bleached kraft mill effluent. Environ. Pollut. 17: 207-217. Sullivan, J.F., and G.J. Atchison. 1978. Predator-prey behaviour of fathead minnows, Pimephales promelas and largemouth bass, Micropterus salmoides in a model ecosystem. J. Fish Biol. 13: 249-253. Sylvester, J.R. 1972. Effect of thermal stress on predator avoidance in sockeye salmon. J. Fish Res. Board Can. 29: 601-603. Szaro, R.C. 1986. Guild management: An evaluation of avian guilds as a predictive tool. Environ. Manage. 10(5): 681-688. Tagatz, M.E. 1976. Effect of mirex on predator-prey interaction in an experimental estuarine ecosystem. Trans. Am. Fish Soc. 105: 546-549. Tanner, J.T. 1978. Guide to the Study of Animals Populations. Univ. Tenn. Press, Knoxville, TN. 166 ------- Tingey, D.T., R.A. Reinert, J.A. Dunning, and W.W. Heck. 1973. Foliar injury responses of eleven plant species to ozone/sulfur dioxide mixtures. Atmos. Environ. 7: 201-208. TRI. 1988a. Ecological endpoint selection criteria, Office of Research and Development, U.S. EPA, September, 1988, Washington, D.C. TRI. 1988b. Ecological model selection criteria. Office of Research and Development, U.S. EPA, September 1988, Washington, D.C. TRI. 1988c. Ecological risk assessment issues Volume I - Ecological measurements used in risk assessment: State of the science. Office of Research and Development. U.S. EPA, September 1988, Washington, D.C. TRI. 1988d. Ecological risk assessment issues Volume II - Ecosystem stability/recovery report. Office of Research and Development, U.S. EPA, September 1988, Washington, DC. Usher, M.B. 1972. Developments in the Leslie matrix model, pp. 29-60 in Jeffers, J.N.R, ed., Mathematical Models in Ecology. Blackwell Scientific Publ. Oxford. Van Voris, P., R.V. O'Neill, W.R. Emanuel, and H.H. Shugart. 1980. Functional complexity and ecosystem stability. Ecology 61(6): 1352-1360. Verma, S.R., I.P. Tonk A.K. Gupta, and R.C. Dalela. 1981. In vivo enzymatic alterations in certain tissues of Saccobranchus fossils following exposure to four toxic substances. Environ. Pollut. Ser. A. Ecol. Biol. 26: 121-127. Verner, J. 1984. The guild concept applied to management of bird populations. Environ. Manage. 8(1): 1-13. Visser, S., J. Fujikawa, C.L. Griffiths, and D. Parkinson. 1984. Effect of topsoil storage on microbial activity, primary production and decomposition potential. Plant and Soil 82: 41-50. Walley, K.A., M.S.I. Khan, and A.D. Bradshaw. 1974. The potential for evolution and heavy metal tolerance in plants. I. Copper and zinc tolerance in Aerostis tenuis. Heredity 32: 309-319. Warfar, W.V.M., P. LeCorre, and J.L. Birrien. 1983. Nutrients and primary production in permanently well-mixed temperate coastal waters. Est. Cstl. Shelf Sci. 17: 431-446. Warren, C.E. 1971. Biology and Water Pollution Control. W.B. Saunders Co., Philadelphia, PA. Weis, J., and P. Weis. 1987. Pollutants as developmental toxicants in aquatic organisms. Environ. Health Perspect. 71: 77-85. Wellings, S.R., C.E. Alpers, B.B. McCain, and B.S. Miller. 1976. Fin erosion disease of the starry flounder (Platichthvs stellatusl in the estuary of the Duwamish River, Seattle, Washington. J. Fish Res. Board Can. 33: 2577-2586. 167 ------- Wentsel, R., A. Mcintosh, W.P. McCafferty, G. Atchinson, and V. Anderson. 1977. Avoidance response of midge larvae (Chironomus tentans't to sediment containing heavy metals. Hydrobiology. 55: 171-175. Westman, W.E. 1985. Ecology, Impact Assessment, and Environmental Planning. John Wiley and Sons, New York. pp. 1-532. Wiegert, R.G., and F.C. Evans. 1964. Primary production and the disappearance of dead vegetation on an old field in southeastern Michigan. Ecology 45(1): 49- 63. Winner, R.W., T. Kelling, R. Yeager, and M.P. Farrell. 1977. Effect of food type on the acute and chronic toxicity of copper to Daphnia magna. Freshwater Biol. 7: 343-349. Woodwell, G.M. 1970. The energy cycle of the biosphere. Sci. Amer. 223: 64- 74. Woodwell, G.M., and D.B. Botkin. 1970. Metabolism of terrestrial ecosystems by gas exchange techniques: The Brookhaven approach, pp. 73-85 in Reichle, D.E. ed., Analysis of Temperate Forest Ecosystems. Springer, NY. Wu, R.S.S. and C.D. Levings. 1980. Mortality, growth and fecundity of transplanted mussell and barnacle populations near a pulp mill outfall. Mar. Pollut. Bull. 11:11-15. 168 ------- |