National Academies Press: OpenBook

Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion (1981)

Chapter: 6 Biological Accumulation and Effect of Atmospheric Contaminants

« Previous: 5 Atmospheric Transport, Transformation, and Deposition Processes
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 87
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 88
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 89
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 90
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 91
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 92
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 93
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 94
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 95
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 96
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 97
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 98
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 99
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 100
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 101
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 102
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 103
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 104
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 105
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 106
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 107
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 108
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 109
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 110
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 111
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 112
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 113
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 114
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 115
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 116
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 117
Suggested Citation:"6 Biological Accumulation and Effect of Atmospheric Contaminants." National Research Council. 1981. Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion. Washington, DC: The National Academies Press. doi: 10.17226/135.
×
Page 118

Below is the uncorrected machine-read text of this chapter, intended to provide our own search engines and external engines with highly rich, chapter-representative searchable text of each book. Because it is UNCORRECTED material, please consider the following text as a useful but insufficient proxy for the authoritative book pages.

6. BIOLOGICAL ACCUMULATION AND EFFECT OF ATMOSPHERIC CONTAMINANTS As we saw in the previous chapter, the particular structure of aquatic or terrestrial ecosystems can impose constraints upon interactions between the atmosphere and the biosphere. Terrestrial ecosystems are readily accessible to gaseous and suspended substances in the atmosphere. They present large, rough surface areas for exposure to gases, and where forest canopies impede wind transport, the wind velocity is reduced sufficiently to induce substantial dry deposition of fine particles. Evergreen canopies can be active in this manner year-round, whereas deciduous canopies are less effective in winter. In contrast, aquatic ecosystems are somewhat protected from gases by the air-water interface. The rate of gas exchange between the atmosphere and a body of water depends upon the temperature and turbulence of the air and the water, and the diffusivity and concentration gradient of the gas across the air-water interface (Danckwerts 1970~. The rate of exchange, however, is slow compared to aerial diffusion throughout terrestrial vegetation canopies. ACCUMULATION IN TERRESTRIAL ECOSYSTEMS In terrestrial ecosystems, atmospheric contaminants and pollutants are captured by plant surfaces in a variety of ways. Hairy and glandular plant surfaces can physically trap suspended matter in the form of wet or dry aerosols. Trace metals are an example of substances that tend to be tranported through the atmosphere and deposited in this way. Wet precipitation washes suspended matter from the air and also facilitates the eventual transfer of atmospheric contaminants captured by the vegetative canopy to the soil surface. Some metals, such as lead and cadmium, can even be incorporated as particles into plant cuticles (Hemphill and Rule 1975), and if the aerosol particles are less than 0.1 micron in diameter they can enter directly inside the leaf tissues through the stomata! pores. Ultimately all such material, whether attached to or incorporated into 7

88 vegetation, is transferred into the soil surface by rainwater wash or by leaf fall and the decomposition of vegetation. Aerosols may enter leaves through epidermal pores, trichomes, or wounds, as well as stomata! pores. In some cases direct penetration of the waxy cuticle is possible, or secretions by the plant may dissolve the aerosols, facilitating their transport and absorption by plant tissue (Hosker and Lindberg 1981~. Factors that affect the size of stomata! apertures can determine how readily gases and aerosols reach the internal tissues of a leaf. The layer of moisture surrounding the palisade and mesophyll cells inside leaves also is crucial to the rate of reaction of the substances in the aerosols or gas with the living protoplasm. Among the substances that enter an ecosystem as aerosols are sulfur, nitrogen, and several of the more volatile heavy metals, such as mercury, selenium, and arsenic. Gaseous pollutants generally react with plant surfaces more readily than aerosols. Some substances, such as hydrochloric acid, are so reactive that they will be absorbed by any plant surface they contact (W. Heck, USDA/SEA, North Carolina State University, Raleigh, NC, personal communication, 1980) while other gases, such as monoxides of carbon and nitrogen and volatile organic compounds, react very slowly with plants. Gaseous substances are also more readily incorporated into plant tissues and translocated by vascular systems than aerosol pollutants. Most studies of pollutant uptake by plants have been of single pollutants on individual leaves or whole plants under laboratory conditions. Estimates of fluxes of pollutants into plant tissues under field conditions have been made using the eddy correlations technique (Eastman and Stedman 1977, Garland et al. 19731. At present there is no satisfactory way of extrapolating such data to whole ecosystems or to mixtures of pollutants. Figure 6.1 shows schematically some of the ways gaseous pollutants may react with plant surfaces, and Table 6.1 summarizes the types of reactions a number of gases have with plant tissues. Lindberg and McLaughlin (1981) have reviewed problems associated with collecting and interpreting data on the interaction of air pollutants with vegetation. ACCUMULATION IN AQUATIC ECOSYSTEMS The pathway followed by atmospheric contaminants entering surface waters depends on several factors, including thermal stratification in the water, the physical form of the contaminant at entry, its solubility, the interactions of organisms with the contaminant, and the affinity of the pollutant for sestonic (suspended) particles or sediment surfaces. Aerosols and large particulate components of the atmosphere tend to become trapped in the surface film of natural waters. As a result, surface films often become enriched in atmospheric contaminants such as trace metals and organic micropollutants by as much as orders of magnitude more than the concentration observed in the bulk water beneath (Duce et al. 1972, MacIntyre 1974, LiSS 1975, Andren et al.

89 rc1 rs1 IT' ' - / ~~~ ~ EPIDERI\/ IS I NTE RCE LLU LA R ~ rL 1 · , · -. rL l. . SPACE ~ t.t I l', FL1 XL1 / ' Fa1 ~ GUAR D CE LL 1 kd I 1\ 1 1} 1\ Am. VASCU LAY ~'~.-'T1 l. .; —I PALLISADE FLII, ~ PARENCHYMA 1l C E L L I'W-1 ~3' ~~ 1l:7$r.2 ~ ~ v. ~ ~~ AAA_ ~ XL2 STOMATE rc2 Fs2 SPONGY PARENCHYMA ~MESOPHY LL CELL . ~ ra2 . Fa2 FIGURE 6.1 Electrical analog of pollutant exchange between leaf and surrounding air. The circuitry is superposed on a cross-section of an amphistomatous leaf (stomates on both surfaces). Xair, XL1 and xL2, and xint denote gaseous pollutant concentrations in well-mixed surrounding air, at the upper and lower leaf surfaces, and average gas phase concentrations within the leaf mesophyll, respectively. The (variable) resistances ra1 and ra2, rL1 and rL2, and rL1 and rL2 are the upper and lower boundary layer resis- tances, stomata! plus intercellular resistances, and cuticular plus internal resistances, respectively. Resistances rCl and rC2 denote resistance to chemical reaction at the upper and lower leaf surfaces. The fluxes Fa, Fa1 and Fa2, FSl and FS2, FL1 and FL2, and F'L1 and FL2 denote the total flux to both surfaces (Fa = Fa1 + Fa2), surface uptake at the upper and lower surfaces, and the fluxes through the upper and lower stomata and cuticles. The electrical symbols depicting grounds ( _ ) represent the termi- nation of fluxes due to reaction with leaf surface materials. The (variable) capacitor sym- bols (- ~) represent surface adsorption or retention (based on Bennett et al. (1973)). SOURCE: Hosker and Lindberg (1981). a ~ XA I R

go TABLE 6.1 Types of Interactions of Gases with Plant Surfacesa Type of Reaction Atmospheric Contaminant Primarily react with or are absorbed onto outer surfaces Primarily react within the leaf tissue React both on outer surfaces and in internal tissues React slowly but have potentially important interactions with plants H2O2, HNO3, H2 SO4 SO2, NH3, NO2, PAN, metabolically reactive cO2a, O2a ethylene,b formaldehyde HE, HCl, O3, C12 N2O NO, CO, organic molecules such as hydrocarbons, pheromones, and terpenes aNormal large-scale exchange in photosynthesis and respiration. bEthylene is unusual in that it is a plant hormone and can have a strong effect in plant cells even at very low concentrations. NOTE: Aerosols tend to react more with outer surfaces because they do not pass through stomata as easily as gases. Gases interact more readily with the internal tissues of plants because they can rapidly diffuse into the leaf through stomata! openings and the moist unprotected surfaces of internal tissues present more reaction sites for the gases. SOURCE: After Hosker and Lindbery (1981).

91 1976, Elzerman and Armstrong 1979~. The fates of such contaminants in surface films are poorly known. Thermal stratification, either in the form of a thermocline or as ice cover, is a barrier to gas exchange and the dispersal of pollutants into aquatic ecosystems. Ice cover totally inhibits exchange of gases between air and water (Schindler 1971) and may be of major significance in temperate lakes, which can be ice covered for as much as 6 months, and in arctic lakes, on which ice cover can last more than 11 months of the year. In winter, oxygen consumption in ice-covered, eutrophic aquatic ecosystems may cause the entire water body to become anoxic. An experiment by Welch and coworkers (1980), simulating a natural gas pipeline break, found that methane discharged into an arctic lake in winter was trapped in the lake under ice cover until spring and that oxygen in the lake was slowly consumed by methane-oxidizing bacteria. On the other hand, atmospheric pollutants that accumulate in the ice cover of a lake over several months are discharged into the lake as a large pulse during ice and snow melt. Figure 6.2 shows one such pulse caused by the acid precipitation that accumulated in winter snow. A larger-scale example of thermal stratification is the thermocline that exists between the surface ocean and the deep ocean. Except at the poles where mixing can occur the thermocline acts as an effective barrier against transfer of substances between surface and deep ocean water. For example, the time needed to achieve equilibrium in CO2 transfer across the ocean thermocline is several hundred years, while the time required for equilibrium between the atmosphere and the surface water of the ocean is only about 8 years. If the ocean surface waters had sufficient capacity for CO2, the equilibration time is short enough that the CO2 released by anthropogenic activities would not accumulate in the atmosphere. The surface ocean, however, cannot absorb all the CO2 that is released. Exchange with the deep ocean is therefore the process that limits the rate of exchange and the achievement of equilibrium for CO2 between the atmosphere and the oceans, and the exchange is so slow that excess CO2 does indeed accumulate in the atmosphere. Likewise, once a substance enters the deep ocean, residence time is very long, not only for carbon dioxide but also for other contaminants emitted into the atmosphere as combustion products (Broecker 1974~. Carbon dioxide, sulfate, nitrate, and ammonium are all required to some degree by algae and other aquatic plants. The carbon, sulfur, and nitrogen are either actively metabolized and passed up the food chain or egested as fecal pellets. Although much of the fecal material is Demineralized in the water column, some of the carbon, nitrogen, and sulfur eventually collect in sediments. Nitrogen, which is usually in greatest demand by plants, tends to be transferred most rapidly. Reactive pollutants adsorbed or absorbed by plankton or other suspended articles travel similar pathways. Less reactive pollutants, which remain in solution, tend to remain trapped above the thermocline, because diffusive transfer across the thermocline is much slower than the settling of particles.

92 pH r 7 6 5 1 976/77 M FIGURE 6.2 Seasonal changes of pH of lake water taken from the outlet of Little Moose Lake, Adirondacks, New York, and from a 3-m-deep pipe in the lake. The spring acid pulse is greater at the surface than at the 3-m-deep level. SOURCE: Schofield (1980). Reprinted with permission from Ann Arbor Science Publishers, Inc., Ann Arbor, Michigan. A. I· ~ fit A

93 For example, Hesslein et al. (1980) studied the pathways of trace metals by adding radioactive tracers to a whole lake. Iron-59 and cobalt-60, which are quickly adsorbed by particulate matter, settled rapidly from the epilimnion (surface water) to sediments. Iron redissolved in anoxic surface sediments underlying the hypolimnion (deep water). In contrast, cesium-134 remained in solution and was removed from the epilimnion at a much slower rate, chiefly by direct reaction with littoral sediments. Removal times were roughly twice as long as for cobalt and iron. Other metallic tracers had intermediate removal times. Substances that are relatively inert can accumulate in the deep ocean. Examples are the radioactive pollutants argon-39 and krypton- 85, or the halocarbon pollutants CF3C1, CF4, and CC14. All of the above enter surface waters as gases. Cesium-137, which is a product of weapons testing and the nuclear fuel cycle, enters the ocean as an aerosol. It accumulates in the lipid fractions of marine organisms, and from the lipid phase it is ultimately released by metabolism or by rapid mineralization upon the death of the organism. Hence, cesium is removed from salt water as slowly as from fresh water. The volume of the deep ocean is 50 times that of the layer above the thermocline. Pollutants entering the deep ocean are thus diluted 50-fold from their concentration in the surface layer. Some will enter the sediments associated with the depositing inorganic or biological debris. Those with longer residence times--i.e., the less reactive pollutants--will remain in the ocean deeps for millennia. ACCUMULATION IN Sol LS AND AQUATIC SEDIMENTS In both terrestrial and aquatic communities, there is a tendency for most pollutants to accumulate in soils or lake sediments, where they become concentrated by absorption, direct cation exchange, or by chemical precipitation. There are both striking similarities and substantial differences in the response of soils and aquatic sediments to pollutants. Both are depositories for biological and mineral matter and can concentrate any pollutants bound to deposited materials. In terrestrial ecosystems, however, plant roots tend to resorb and recycle materials to some degree, and there is some long-term storage in wood. With the exception of sphagnum bogs and wetlands, there is no such recycling in aquatic ecosystems, and there are no long-lived plants to form long-term living sinks for pollutants although dead organic debris may accumulate in aquatic sediments that are long-term sinks for many substances in the form of organic molecules, insoluble oxides, hydroxides of redox-sensitive metals, or as sulfides under anoxic conditions. In anoxic environments, there may be substantial recycling of many elements that are not mobile in oxic environments from sediments into the hydrosphere and atmosphere. For example, sulfur may be released as hydrogen sulfide or clime thyl sulfide; several metals are released in a methylated form; and nitrogen is released as ammonia, nitrous oxide, and molecular nitrogen. These processes, widespread in aquatic

94 ecosystems, also occur in waterlogged soils (Tusneem and Patrick 1971, Reddy et al. 1978~. Flux rates for the above substances are poorly known and are highly variable even within one ecosystem. In aquatic ecosystems, some pollutants are rapidly locked into sediments, where they cease to affect planktonic or nektonic organisms. For example, the addition of phosphorus to lakes is objectionable largely because it causes increased growth of phytoplankton. Once deposited in oxic sediments, it is usually no longer a threat. Sediment-water partition coefficients for phosphorus range from 104 to 106. Likewise, most trace metals are rapidly transferred from water to sediments (Hesslein et al. 1980), as are nitrogen and sulfur, although they are less efficiently concentrated in sediments because microbial reduction can rapidly release them (Schindler et al. 1977, 1980a). Under certain circumstances, the high sediment-to-water partition coefficient may become an undesirable feature for the aquatic ecosystem as a whole. Once incorporated into sediments, a pollutant cannot be flushed rapidly from a water body. Some materials may be transferred directly from sediments to pelagic biota; for example, mercury, selenium, and perhaps lead are methylated in sediments and transferred to the water column. Owing to their high affinity for lipids, methylated metals can be taken up and accumulate In fish and other pelagic forms, even though concentrations in water remain low. DOT and many other pesticides, petroleum hydrocarbons, polynuclear aromatic compounds, and polybrominated or polychlorinated biphenyls (PBBs and PCBs) also are known to be highly concentrated in both sediments and the lipids of living organisms. Bottom-living aquatic organisms that ingest sediments many receive very high concentrations of pollutants. The incidence of deformities among the larvae of bottom dwelling chironomids is one of the most sensitive indicators of contamination in the aquatic environment (Warwick 1980a, 1980b). Likewise, surface soils, where high accumulation of pollutants occurs, are occupied by innumerable small invertebrates, including mites, springtails, beetles, earthworms, Crustacea, and midge larvae, and by a variety of fungi, bacteria, and other organisms. The fungi include both important plant pathogens and mycorrhizal fungi crucial for recycling nutrients from soil to plants. All these organisms perform decomposition via many complex pathways, recycling crucial nutrients in forms usable by plants. The seeds, spores, and eggs of many organisms also begin development in these layers. Saprophytic organisms that burrow throughout soil and sediments may cause deposited pollutants to be diluted, or spread throughout the upper layers. As a result, substances deposited in one year may be dispersed through a layer of soil or sediment representing decades or even centuries of accumulation (Davis 1968, Berger and Heath 1968, Rhoads 19731. Such processes will delay both the development of high concentrations of pollutants in surface sediments and the burial of substances after inputs have ceased (Schindler et al. 1977~. This "bioturbation" may be responsible for the fact that concentrations of DDT and PCBs in the fishes and sediments of the Great Lakes have not

95 decreased greatly in the past decade, despite stringent control of inputs (International Joint Commission 1980~. Acid conditions and high trace-metal concentrations in excess of 500 ppm are known to inhibit litter decomposition, causing undecomposed organic matter to accumulate in both terrestrial and aquatic ecosystems (Watson et al. 1976, Freedman and Hutchin son 1980, Strojan 1980, Hendrey et al. 1976, Traaen 1980~. A number of key enzyme systems appear to be affected (Tyler 1974, 1976a,b). This accumulation assuredly slows nutrient cycling. Soil and aquatic fungi are usually more acid tolerant than bacteria, and thus they become the dominant decomposers in acidic ecosystems. But even fungi appear to be susceptible to extremely high concentrations of metals derived from the atmosphere (Jordan and Lechevalier 1975, Williams et al. 1977~. Fungi with increased tolerance of acids and heavy metals have, however, been obtained from several sites near smelters (Hartman 1976, Freedman 1978, Carter 1978~. Aquatic Sediments as Historical Records Aquatic sediments provide a "history" of changes in the aquatic ecosystem, its terrestrial watershed, and its airshed. The increase in heavy metals in lake sediments in this century is an excellent example (Robbing and Edgington 1977, Norton et al. 1978, Galloway and Likens 1979~. A study of sediment cores from the mouths of tributary streams around the St. Lawrence River and Great Lakes confirms the relationship between such increases and terrestrial watersheds contaminated by human activity (Fitchko and Hutchinson 1975~. Mackereth (1966) and Warwick (1980) were able to show historical evidence for increased weathering of terrestrial watersheds owing to human disturbance. Davis et al. (1980) have been able to decipher an acidification history of lakes in Norway and Maine, using a combination of recent dating methods and changes in the abundance of acidophilic diatoms. No study has yet used paleolimnological methods to construct a history of terrestrial responses to acidification. In most cases, such methods offer at present the only possibility of documenting and dating the course of acidification in either terrestrial or aquatic ecosystems, because of the paucity of regional background data collected prior to the development of the acid rain problem. TRANSFORMATIONS IN Sol LS AND WATERS A number of volatile components of the atmosphere can react with waters or soil constituents and be incorporated into the compounds in the biosphere. How volatile components are removed, the responsible organisms or abiotic mechanisms, and frequently the factors that can affect removal are as yet unclear or wholly unknown. Some removal mechanisms are abiotic, but more often they are entirely biological.

96 Many organic compounds present in the atmosphere undoubtedly are transformed and decomposed in soils and waters. The evidence for such transformations, however, is derived usually from studies in which the chemical has not been supplied in the volatile form and in which the chemicals were present in water or soil samples at concentrations far higher than are present in nature. Tests with concentrations of contaminants far in excess of ambient levels must be interpreted very cautiously, because chemicals that are biodegraded at high concentrations may degrade slowly or even not at all when present at low concentrations (Boethling and Alexander 1979a,b). If the transformation yields inorganic products (CO2 and H2O), the reaction detoxifies the molecules. Among the organic compounds destroyed--at least at higher concentrations--are certain polycylic aromatic hydrocarbons, phenols, alkanes, esters, and simple aromatic molecules. In some instances, however, the transformations may yield not inorganic products but other organic products, which may themselves be toxic. Certain organic compounds are resistant to microbial transformation and thus persist, for example certain chlorinated aromatics, aliphatics, and several nitrogen heterocycles. The biodegradability of many of the organic atmospheric pollutants has not been directly evaluated, and none has been tested at realistic concentrations. Hence it is not possible at the present time to assess their rate of transformation when introduced into waters or soils or even whether such a transformation does take place. Some of the problems in assessing such kinds of transformations and the types of reactions that occur are considered by Alexander (1981~. Information on the role of soils in removing or generating atmospheric substances is growing constantly, and current assessments will need to be modified as information is obtained from new studies. The following examples have been relatively well researched and are offered as an indication of the current level of understanding for several atmospheric constituents. Methane. A number of bacteria present in soils and lake waters are able to oxidize methane. Although there has been considerable study of which organisms can do so in laboratory media and of the biochemistry of the process, surprisingly little attention has been paid to the rates of biological methane oxidation under natural conditions. Studies of the process in paddy fields indicate that the rate of oxidation is limited by the rate of methane diffusion to the soil (de Bont et al. 1978~. In lakes, the rate of oxidation is usually limited by diffusion of oxygen and methane, as well as by the concentration of nitrogen (Rudd et al. 1976~. During summer thermal stratification, activity is restricted to the metalimnion (the part of the water column where the thermal gradient occurs), although after the fall overturn destroys thermal stratification methane oxidation may occur throughout the water column. Often this activity persists throughout the following winter (Rudd and Hamilton 1975, Welch et al. 1980~.

97 Carbon monoxide. An increase in ambient levels of carbon monoxide would be expected on the basis of present rates of global emission, but no such major increases are evident; therefore a significant sink for this gas must exist. Many investigators have pointed to soils as a means for removing carbon monoxide from the gas phase. Because carbon monoxide is oxidized in nonsterile but not in sterile soil (K. Smith et al. 1973), the removal mechanism is apparently microbial. Estimates have been made of global uptake by soil of carbon monoxide based on its rapid loss from the head space of a container when soils are incubated in the laboratory with this volatile compound. For example, based upon tests in which several soils were exposed to 100 ppm carbon monoxide, it was calculated that soils of the world remove 14.3 x 1015 g of carbon monoxide per year (Ingersoll et al. 1974~. Such figures are undoubtedly a gross overestimate, because the rate of removal is markedly dependent upon concentration and appears to follow Michaelis-Menten kinetics. The rate of removal at concentrations equivalent to those in the atmosphere is far lower than when abnormally high concentrations are used in the tests (G. W. Bartholomew and M. Alexander, Cornell University, personal communication). Other estimates suggest that the earth's land surface removes 8 x 109 g of carbon monoxide per hour (Heichel 1973) and that the rate of removal ranges from 4.5 to 14 x 1014 g per year (Nozhevnikova and Yurganov 1978~. Despite the disparity in estimates for removal, it is generally believed that soils are a major sink and that microorganisms in soil are chiefly responsible for the removal (NRC 1977d, Liebl and Seller 1976~. Many individual microbial species are able to metabolize carbon monoxide in laboratory culture (Nozhevnikova and Yurganov 1978, Bartholomew and Alexander 1979~. Investigations with soils suggest that, in the few soil samples investigated, the microorganisms consuming carbon monoxide are neither heterotrophs using carbon monoxide as a carbon source nor autotrophs using it as a source of energy; rather the responsible organisms co-metabolize the gas (Bartholomew and Alexander 1979, and personal communication). Ethylene. Soils also remove ethylene from the gas phase. Little of this removal occurs in sterilized or air-dried soil, but the reaction is rapid in moist and nonsterile soil. Such data indicate that the process is microbial. The organisms involved are apparently aerobic species, probably bacteria. The few rates that have been calculated in laboratory studies indicate that from 0.14 to about 14 nmol can be removed per gram of soil per day (K. Smith et al. 1973, Abeles et al. 1971~. In aquatic ecosystems ethylene is oxidized by the same bacteria that oxidize methane. Under similar conditions ethylene usually is used preferentially (Flett et al. 1975~. Acetylene. Acetylene can apparently be readily destroyed, again by microbial processes, since it does not occur in sterilized or air-dried soil. The rate of removal of acetylene at a moisture level of 50 percent of the water-holding capacity of the soil is 0.24 to 3.12 nmol per gram of soil per day, depending on the composition of

98 the soil (K. Smith et al. 19731. Acetylene can be reduced to ethylene by nitrogen-fixing bacteria in both aquatic and terrestrial ecosystems--a fact that is utilized to estimate amounts of nitrogen fixation (Stewart et al. 1967~. Nitrogen. Biological nitrogen fixation leads to N2 incorporation into soil, and one global estimate for this process is that the annual amount of molecular nitrogen fixed into soil is 99 x 1012 grams of nitrogen per year (Delwiche and Likens 19771. Both blue-green algae and methane-oxidizing bacteria are known to fix N2 in aquatic ecosystems (Davis et al. 1964~. While there have been some measurements of nitrogen fixation in aquatic ecosystems (Johannes et al. 1972, Mague and Holm-Hansen 1975) data are presently too scarce to establish their contribution to global fixation of nitrogen. Nitrous oxide may also be removed from the gas phase. This removal is also apparently related to microbial activity and is part of the sequence involving the reduction of inorganic nitrogen compounds to molecular nitrogen. The process is promoted by anaerobiosis and by the presence of organic materials that stimulate microbial proliferation. But soils are a more significant source than a sink for nitrous oxide (Blackmer and Bremner 1976~. Nitric oxide is sorbed from the air. This sorption may take place even in dry soil, and the process leads to an increase in acidity and a rise in the nitrate concentration as the sorbed gas is oxidized in soil to nitrate (Prather et al. 1973a). Nitrogen dioxide is also readily removed from the gas phase, and it is converted in soil to nitrite and nitrate. Nitrogen dioxide is sorbed rapidly onto air-dried soils, and the removal is apparently nonmicrobial, because it takes place in both nonsterile and sterile soil (Prather et al 1973b). On the other hand, the nitrite that is formed is converted to nitrate--and is thus detoxified--by microorganisms (Ghiorse and Alexander 1976~. Soils remove a smaller portion of ammonia by chemical reaction because much of the ammonia is removed by rainfall and dry deposition (Rodgers 1978~. Sulfur. Sulfur dioxide is readily removed from the atmosphere in contact with soil. The removal appears to be non-biological because rates are the same whether the soil is sterile or nonsterile. Much of the sulfur dioxide that is thus removed by soil is converted to inorganic sulfur products, and much of it is probably ultimately convertedito sulfate. Nevertheless, part of the sulfur derived from the gas phase is apparently converted to an organic sulfur compound or compounds (Ghiorse and Alexander 1976~. The removal is affected by the moisture content of the air and the sulfur dioxide concentration of the gas phase (Yee et al. 1975~. Soils also remove hydrogen sulfide and methyl mercaptan from the air. The addition of water to a dry soil decreases the rate of methyl mercaptan removal and has little effect on hydrogen sulfide removal. Both air-dried and wet soils are able to remove clime thyl disulfide, carbonyl sulfide, and carbon disulfide. In the process, small amounts of carbonyl sulfide are converted in moist soils to carbon disulfide.

99 Microorganisms are apparently involved in the removal of dimethyl sulfide, dimethyl disulfide, carbonyl sulfide, and carbon disulfide by moist soils, but they may not be significant in the removal of hydrogen sulfide, which reacts chemically (K. Smith et al. 1973, Bremner and Banwart 1976~. The ultimate destruction of the organic portions of any sulfur compounds that are generated probably requires microbial intervention, as is true of most organic compounds. TRANSFER OF SUBSTANCES FROM TERRESTRIAL TO AQUATIC ECOSYSTEMS Because aquatic ecosystems are repositories not only for direct atmospheric deposition but also for many of the substances leached from or initially deposited upon terrestrial drainages, aquatic ecosystems may be contaminated to a greater degree than terrestrial ecosystems. The degree to which materials entering from the atmosphere are retained by terrestrial soils and vegetation is therefore an important consideration when aquatic contamination is studied, particularly because the terrestrial portion of a watershed is usually much larger than the lake or stream which drains it. Many substances have a high affinity for clay minerals in soils and become tightly bound in terrestrial soils. Large amounts of such substances will not be transferred to lakes as long as soil erosion is well controlled. This includes most trace metals and the hydrogen ion, as well as phosphorus. Other substances, such as nitrogen, are retained due to biological demands. At Hubbard Brook, New Hampshire, over a 10-year period, 88 percent of the ammonium that entered the watershed was retained, but only 13 percent of the nitrate was retained (Likens et al. 1977~. Still other substances leave watersheds in amounts equal to or higher than that in the precipitation that enters the watershed. At Hubbard Brook, stream flow removed more sulfate than entered as bulk precipitation. The discrepancy may be due to dry deposition, which was not included in estimates of substances entering the system. In Norway, inflow and outflow of sulfate from lakes are roughly balanced (Abrahamsen 1980~. The difficulty in measuring entry of a substance to a watershed via dry deposition makes more precise estimates of watershed balances impossible. Many cations may leave terrestrial soils after exchanging for H+. These include both soil nutrients, such as calcium, and substances which may be toxic to aquatic life, such as aluminum. One study of substances deposited from the atmosphere to terrestrial soils (Hutchinson et al. 1975) showed that nickel, copper, and cobalt from the Sudbury smelter had not only accumulated in soils to concentrations toxic for seedlings but had also poisoned nearby lakes through runoff from the watersheds. River sediments in drainage from the area were contaminated up to 100 km away. Land management practices that allow land erosion or leaching are thus likely to enhance the movement of atmospheric substances into aquatic ecosystems. It now is recognized that fish mortality attributed to acid precipitation is in large part caused by the aluminum leached from

100 terrestrial soils during spring melt (Cronan and Schofield 1979! Acid precipitation is also identified as a contributor to the increased runoff of nitrate, which is causing concentrations of nitrate in Adirondack surface waters to become uncomfortably close to maximum acceptable levels for drinking water. While much of the nitrate undoubtedly results from precipitation inputs in excess of biological demands, it is possible that the suppression of denitrification by acidification of soils may favor nitrate accumulation (Alexander 1980~. Seip (1980) calculated that the actual input of hydrogen ions to Norwegian lakes from direct precipitation as well as its runoff from the lake catchment basin, could explain only a negligible part of the acidification that takes place. The rate of acidification is better correlated with sulfate deposition. Sulfate output from terrestrial ecosystems is roughly equal to input (Abrahamsen 1980), and thus if Seip's observations are correct, increased sulfur deposition in terrestrial areas could contribute to the acidification of receiving waters. A complete understanding of the natural processes and mechanisms involved in the acidification of freshwaters is still needed. On the other hand, few lakes are as acid as the precipitation falling in their watersheds, which leads some authors to believe that buffering by terrestrial ecosystems may protect poorly buffered lakes from acidifying as rapidly as they might under the influence of unaltered acid precipitation upon the entire drainage basin (Gorham and McFee 1980~. The relatively high pH levels of calcareous soils neutralize the acid, while in soils below pH 5, aluminum species may be the major source of buffering (Johannessen 19801. The magnitude of transfers of acidifying substances from terrestrial ecosystems to aquatic ones is usually unknown, and this hinders the prediction of rates of acidification and recovery of aquatic ecosystems. Henriksen (1979) and Almer et al. (1978) have developed simple models, based on expected ratios of alkalinity to concentrations of calcium or calcium plus magnesium for predicting the degree of acidification of lakes. Henriksen (1980) has extended this logic to predict the degree of acidification which will take place under precipitation regimes of different acidity. The model assumes that as precipitation becomes more acid, the calcium concentration of lakes does not increase. Several studies support this assumption. Maimer {1974) and Schofield (cited in Henriksen 1979) found no increase in calcium concentrations of waters in acidified areas. Schindler et al. (1980a) found no increase in calcium release from lake sediments as pH decreased. On the other hand, Gordon and Gorham (1963), Almer et al. (1974) and Dillon et al. (1980) have observed that calcium losses are higher from terrestrial ecosystems in areas where the pH of precipitation is from 4.0 to 4.5 than in areas with precipitation of normal acidity. It is probable that these observations reflect differences among soil types. Abrahamsen's (1980) experimental acidification of forest plots in Norway supports the latter argument. If the relationship between calcium leaching and acid inputs can be quantified, Henriksen's model may have predictive

101 value (Figure 6.3), although different soils may respond differently to acid inputs. The surface film in aquatic ecosystems is an area of concern. AS noted earlier, organic pollutants and heavy metals may concentrate there at orders of magnitude higher than in the atmosphere above or the bulk water beneath. The ecology of the biota in such layers (the neuston) is little known. However, it is known that a wide variety of organisms congregate near the surface including microscopic autotrophs (algae) and heterotrophs (e.g., amoebae, bacteria), as well as larger organisms such as pelagic fish eggs and crustacean larvae (Makarov 1976~. Oil spills as well as atmospheric deposition may add toxic organic molecules to the surface film and thus affect the survival of the inhabitants of this unique community. For example, recent studies by Hardy and Creselius (1981) have shown urban aerosols to be six times more toxic than rural particles, with toxicity to marine phytoplankton being due primarily to soluble Pb, V, Cd, Cu. Zn, and Ni, in that order. While present deposition rates of atmospheric particulate matter do not appear sufficient to inhibit marine primary productivity, they may have serious effects on the sea-surface microlayer--the neuston. Wetlands are transitional ecosystems that incorporate many of the features of both terrestrial and aquatic ecosystems, such as canopy capture and air-water exchanges. In addition, gaseous substances may be regenerated by both oxic and anoxic means. There are vast tracts of wetlands in northern latitudes and coastal marine areas, but data for such ecosystems are too scarce. This shortcoming must be rectified to provide an accurate evaluation of their role in global biogeochemical cycles, and to permit natural global exchanges between the atmosphere and the biosphere to be quantified. POLLUTANTS AND INDIVIDUAL ORGANISMS Several possible types of interactions may occur between pollutants and individual organisms. The pollutant may be actively taken up by the organism--bioconcentrated--or, avoided or excreted--bioexcluded; neither occurs if the molecule is biologically inert. In general, substances required for the growth of the organism will be bioconcentrated. For example, aquatic plants require nitrogen in much higher concentrations than exist in surrounding waters, and they may concentrate the element up to 30,000 times with respect to the water in which they grow (Vallentyne 1974~. Phosphorus and carbon, essential nutrients to all living forms, are also bioconcentrated by plants. Some other elements, nutrients not essential to most organisms, are bioconcentrated by organisms with special requirements. For example, only diatoms and a few other Chrysophyta will concentrate silicon, which they require for their frustules. Molluscs and vertebrates require much more calcium per unit weight than organisms that have no heavily calcified structures in their bodies.

102 250 300 c 200 ' c: 100 o 200' * HCO3- Lakes A. 150 O 100 50 04 me' Qua' · 9~ ~' Acid Lakes ~ ~ . ~ ~ I I rat 0 1 00 200 ~ I SO4* j . . in lakewater, peq/l - 7.0 5.0 4.7 4.5 4.4 4.3 4.2 pH of precipitation - 4.1 4.0 FIGURE 6.3 Nomogram used to predict changes in the acidity of lakes as the acidity of precipitation changes. If changes in the acidity of precipita- tion cause no increase in the weathering of the basic cations, Ca++ and Mg++, the relationship between the concentration of basic cations and the concentration of sulfate in lake water would be expected to move horizontally as a lake is acidified (arrow A). If increased acidity is accom- panied by increased leaching of basic cations, the lake water chemistry changes as indicated by the arrow B. causing the degree of acidification to be underestimated. SOURCE: After Henriksen (1980).

103 Some rare elements are bioconcentrated because an organism's physiology cannot distinguish them from more abundant and useful elements that are chemically similar. The fact that molluscs and vertebrates take up strontium, lead, and radium along with calcium is well-known. Incorporation of toxic or radioactive substances with calcium is particularly dangerous, because substances incorporated in bone, shells, and other calcified tissues are not metabolized quickly--i.e., they have a long biological residence time. In general, elements with similar chemical properties may be expected to be dealt with in a similar fashion by organisms. A special form of bioconcentration, called biomagnification, occurs when organisms bioconcentrate substances via the food chain. The accumulation and concentration of DOT in predatory birds is a well-known example of the biomagnification of this toxic substance. More commonly, toxic elements are bioexcluded, that is, organisms contain them at concentrations much lower than those in their surroundings or their food. For example, vertebrates appear to exclude lead selectively from absorption by their bodies; the concentration of lead in their tissue is lower than in food organisms (Settle and Patterson 1980~. Gachter and Geiger (1979) using radioactive tracers found that 5 toxic metals, including mercury and zinC, decreased in concentration as they were passed up an aquatic food chain. Their results contrast with other published reports that show mercury to be bioconcentrated (National Research Council Canada 1979~. For all metals, phytoplankton and periphyton contained the highest concentrations, zooplankton and insects somewhat less, and fishes least of all. Microorganisms can make methylated derivatives of mercury, selenium, and other trace metals (Wood 1974, Chan et al. 1976, Braman and Forbach 1973~. The decomposers can then bioexclude these trace metals owing to the solubility and volatility of their derivatives, which are rapidly carried away by the water column. On the other hand, methylated metals are known to be bioconcentrated in pelagic vertebrates and other organisms where they accumulate in lipid deposits {Wood 1974, Benson and Summons 1981~. In between bioconcentration and bioexclusion are biointermediate elements, which are concentrated only slightly more in organisms than in water, and biologically inert substances, which are found at similar concentrations in both the organism and its surroundings. Oxygen and chloride are nearly biologically inert; potassium and sulfur tend to be biointermediate. Bioconcentrated substances may be stimulatory or very toxic. Nitrogen and phosphorus concentrated by plants from their surroundings stimulate plant growth, if some other factor is not limiting. On the other hand, high concentrations of arsenic are known to inhibit the uptake of phosphorus by phytoplankton slowing the growth of some organisms (Brunskill et al. 1980~. In some cases, a substance may stimulate one group of organisms while being toxic to another. When two or more pollutants enter an ecosystem together, one may enhance or lessen the effects of the other. For example, brook trout are more susceptible to aluminum poisoning at high concentrations of

104 hydrogen ions, owing to the effects of the latter on the chemical form of aluminum in solution (Cronan and Schofield 1979~. On the other hand, the toxic effects of mercury on a wide variety of organisms are suppressed by high concentrations of selenium, through the latter is itself a toxic substance (Rudd et al. 1980~. The same nonmetabolic absorption-adsorption reactions that occur in clays and inert particles take place on the outer surface of some aquatic organisms. The silicious outer frustules of diatoms are known to concentrate radium by adsorption (Havlik 1971, Edgington et al. 1970, Emerson and Hesslein 1973~. Physical factors, such as temperature and light, directly and indirectly affect biological responses to pollutants. Phytoplankton depend on light for photosynthesis and growth, which means that the uptake of any element by plankton is ultimately dependent on light. All of the physiological parameters in biota are temperature dependent to some degree. The solubility of gases and the kinetics of chemical reactions are also dependent on temperature. In some cases, near-lethal concentrations of substances occur in food chains, even under natural conditions. In older carnivorous fish, bioconcentration may lead to 1 to 2 milligrams of mercury per kilogram of flesh, even in remote areas of the Arctic. Under such circumstances, even small anthropogenic additions to the ecosystem can lead to concentrations exceeding standards for human consumption. Armstrong (1979) mentions that fish with concentrations of mercury greater than 10 mg/kg fresh body weight are seldom encountered even in the most heavily polluted environments and hypothesizes that such concentrations may be lethal to the fish themselves. Toxicity of Deposited Substances to Organisms A number of the substances deposited in ecosystems as a consequence of fossil fuel combustion are known to be extremely toxic to plants or animals. These include: SO2, NO2, SO3, H2SO4, NO2, HNO3, HE, H2S, and various trace organics and metals. In most situations, the amount deposited is lower than concentrations reported to have toxic effects, but there are some exceptions, and because few toxicity studies cover entire life cycles for sensitive test organisms or combinations of several pollutants, predictions should be very conservative. Terrestrial Ecosystems In general, primary gaseous pollutants affect only terrestrial organisms and ecosystems. The distance over which effects are noted depends, of course, on weather conditions, the height of the stack

105 emitting the gaseous pollutant into the atmosphere, and the sensitivity of the receptor organisms. Of most concern is sulfur dioxide, which affects sensitive species at extremely low concentrations or after very short exposures. AS discussed later, SO2 may also interact synergistically with other gaseous pollutants--especially ozone--causing increased damage. Linzon (1971) found that Pinus strobes, eastern white pine, was damaged after several years' exposure to an average SO2 concentration of 8 x 10-3 ppm. This concentration is approached in many rural and wilderness areas of the United States and southern Canada, for example in northern Minnesota and northwestern Ontario (Glass and Loucks 1980) and the eastern United States (Costonis 1972~. In other studies, fumigations of only one to several hours with 2.5 to 5 pphm SO2 caused visible damage to new white pine needles (Costonis 1970, 1972, 1973; Houston 1974~. Germination and seed production are also reduced at low SO2 concentrations, particularly when other gaseous pollutants are present (Houston and Dochinger 1977~. Lichens and bryophytes have been described as especially sensitive to SO2. Many species of these plants are absent in urban areas and around point sources where SO2 concentrations for extended periods exceed 0.02 ppm SO2 (Hawksworth and Rose 1976~. The effects of acid precipitation on terrestrial ecosystems are covered in several recent publications (Hutchinson and Havas 1980, Drablos and Tollan 1980, Overrein et al. 1980~. Although sulfur dioxide is the pollutant of primary concern with vegetation, nitrogen oxides have been shown to cause some damage when they are combined with sulfur dioxide, a typical situation near many sources of emissions. Nitrogen oxides can also interact with hydrocarbons to form secondary pollutants such as ozone and peroxyacetylnitrate (PAN), which are more toxic than the nitrogen oxides at low concentrations {Dvorak et al. 1978, Cleveland and Graedel 1979~. The Academy report on ozone and other photochemical oxidants (NRC 1977b) provides some detailed data on damage by those pollutants to a number of plant species, and another report (NRC 1978a) summarizes similar information for sulfr oxides. Responses of sensitive vegetation to nitrogen oxides are given in Figure 6.4. The developmental stage of a plant influences its sensitivity; many plants appear to be most susceptible to damage from gases during flowering. There are no definitive studies of the effects of gaseous emissions on animals at concentrations observed in the environment, although effects have been predicted. Animal populations in fumigated ecosystems may be affected by the disappearance of sensitive plants that are their usual food or shelter (Dvorak et al. 1978~. Along roadsides, for example, contamination from atmospheric pollutants has caused loss of some organisms and accumulation of lead in others (Ward and Brooks 1978~. Acid precipitation resulting from oxides of sulfur and nitrogen can affect terrestrial ecosystems (Drablos and Tollan 1980), but the extent of these effects is difficult to assess. The subject is discussed more thoroughly in Chapter 8.

106 ~W / ZON 6w o o o o o o - o o o cn ~r 0 ~ — o 1. ~ ~ ~ ~ 1 · ~ 1. ~ . ~ ~ . ~ ~ 1. o o o o o o o ~_ ~n / / / o o ... . . o - l I I I I l_ I / ~1 o I / / ~ C / / o / ~ I / / . . ...... . . ( 6/6 Wdd ) Z ON JO NOIlVUlN3ONOO i _ _ o o ~n CC OI o I2J ~r ~ s~ ~ 3O ~ O ~ ~ ~ ° ~ ,4 X '~ X 111 ~ ~ Ct s: ~ .o Ct GO s Z O C~ _ ~o - o o C~ C) s ~ Ct X o ._ s ~ o I,., U:' ·— —o Sa' s ~ a, O C~ _ · a., ~4 Ct C~ _ ~ h_ C) ~ ~,

107 Aquatic Ecosystems The effects of acid precipitation on aquatic ecosystems are well documented, and several symposia have recently been held on the problem (Shriner et al. 1980, Drablos and Tollan 1980~. The effects are largely on organisms in poorly buffered lakes, where alkalinity is less than 100 microequivalents per liter. In such lakes, decreases in alkalinity and pH have been detected at several localities where the precipitation pH is 4.7 or less, both in Scandinavia (Henriksen 1980) and in North America. Chapter 8 gives a detailed account of the acid precipitation problem, and it will not be repeated here. Trace metals and other trace substances are known to be toxic at concentrations found in soils and fresh waters within a few kilometers of smelters (Hutchinson and Whitby 1977~. Toxic effects farther afield are more difficult to predict, because of inadequate background data. Apparently most of the trace organic pollutants transported via the atmosphere that have proved to be troublesome in ecosystems remote from sources are lipid-soluble. Examples include mercury in the methylated form, chlorinated hydrocarbons, and PCBs. In the upper Great Lakes, PCB concentrations in Coho and Chinook salmon, catfish, and eel are unacceptable for human consumption in some localities. Reproductive failure and deformities have been reported for herring gulls as a result (International Joint Commission 1977~. The widespread dispersal of PCBs appears to have resulted largely from atmospheric inputs (International Joint Commission 1980, Murphy and Rzesz~tko 1977~. Concentrations of mercury in fish considered unacceptable for human consumption are often no more than 2 to 3 times natural concentrations, and the increase in release of mercury to the atmosphere caused by man is well over that (Lantzy and Mackenzie 1979~. As mentioned in Chapter 1, cadmium and zinc are approaching toxic concentrations in Lake Michigan (Muhlbaier and Tisue 19801. Other lakes have not been studied, so it is not known how widespread this problem may be. The toxicity of a substance to an organism may be displayed in different ways at different concentrations. For example, copper is known to affect behavior of freshwater fishes at concentrations as low as 4 micrograms per liter, but at this concentration there is no measurable chemical or physiological sign of stress. At 10 micrograms per liter, growth, reproduction, and mortality are affected. Blood chemistry and respiration are not affected until still higher concentrations. The effects of lead are known to be similar, with neurological damage occurring at lower concentrations than could be detected chemically or physiologically. Significant effects were noted at 8 micrograms per liter (Spry et al. 1981~. Interactions among Pollutants or between Pollutants and Other Stresses The effects on the biosphere of many pollutants are influenced by other pollutants and elements normally present in the atmosphere,

108 water, and soil. Sometimes the influences are indirect; for example, the solubility of many heavy metals is enhanced at low pH--i.e., high hydrogen ion concentration--and toxicity increases correspondingly. When direct interactions enhance the effects of one or both pollutants, the relationship is termed synergism; the result is more than additive. When interactions mitigate harmful effects, the relationship is termed antagonism. - Terrestrial Ecosystems Menser and Heggestad (1966) were the first to demonstrate the synergistic effect of two gaseous pollutants upon plants. They reported injury to several tobacco cultivars when these cultivars were exposed simultaneously to concentrations of SO2 and O3 that individually would not injure the plants. Dochinger et al. (1970) and Houston (1974) reported a similar effect of SO2 and O3 in a tree species, Pinus strobus. Table 6.2 gives a summary of the synergistic effects of SO2 with other gaseous pollutants on native North American trees. SO2 and NO2, and NO2 and NO are known to have additive effects on the pea, Pisum sativum (Figure 6.5~. Reinert et al. (1975) review pollutant interaction effects on terrestrial vegetation, including examples of synergisms and antagonisms. The authors point out that such pollutant interactions depend on the genetics and physiology of the target organism, the environmental conditions, and the concentration of pollutants--which may explain some of the contradictory findings reported in the literature. Holdgate (1979a) has made a more recent review; he believes different investigations have yielded different conclusions because of a lack of standardized methods and experimental conditions. Genetic, physiological, and environmental factors must all be standardized before exact comparisons are possible. Interesting synergistic interactions have been described between ozone and both cadmium and nickel (Czuba and Ormood 1975), in their toxic effects on a number of horticultural crops. In contrast, the presence of copper in Sudbury area soils was reported to reduce the toxicity of fumigations by SO2 for some grass species (Toivonen and Hofstra 1979~. The mechanisms of these metal-gas interactions are not known. Affects through stomata! activity appear likely. A number of authors have observed an increased susceptibility of forests to insect outbreaks after exposure to SO2 and other atmospheric pollutants. This subject is reviewed by Renwick and Potter (1981) who attribute the infestations to an increased output of insect attracting terpenes. Aquatic Ecosystems High contents of hydrogen ion in freshwater bodies enhance the solubility of a number of elements normally present in soils and sediments. These include aluminum, manganese, zinc, and iron. Also

109 C) C) C) U. ·C) a Ct P" Cal Ct _, o 5 - o ._ 3 o ._ Ct ._ o C) I: ._ Cal o V, Cat o .~ Ct of A) .~ He o Cal C) ¢ o Ct bO o o o C.) .° X Ct ~ ._ ~ Cal o £ . - d c C) ~ ~ ~ _ ~ ~ so ~ ~ U: ~ ~ ~ _ 50 o o Cal Cal ~ ~ ~ ^ ·= S ·= .=, ~ ~ ° ~ ~ ·O _ ~ 3 ~ ~ 3 o ~ Ct Ct _ ~ ~ ~ . ~ . . 3 ~ , , ~ ~ ~ _ ~ ~ ~ ~ ~ ~ ~ ~ ~ ~ ~o ~ o 0 ~ ~ ~ ~ ~ ~ ~ ,, E X ~ ~ o E -~ ~ ~ ~ =, == D ~ ~ X C ~ ~ .~0 ~ ~ O Z ~ 0m 0 0m P4 ', O ~ ~ O O O ~/ O O ~ ~ . . . . . . . O O O O O O O C~ ,: 3 ~ 3 - , — oo o oo C~ _ U, C) o U3 cn Ct C~ C~ .. V O U,

110 2D _ E 15 ~ - (a) ° '° 5 (b) / ' ~m~ ~ 2S 25 'at ''K2S \ l E~:::...:~-.. )\ > > ..? ~ ~ 0 26~ ~0~'' 50 50 2 ~ FIGURE 6.5 Interactive effects of pollutants. (a) Effects of SO2 and NO2 pollution on the rate of net photosynthesis in pea, Pisum s~ztivum (from Bull and Mansfield 1974). (b) Effects of NO and NO2 pollution on the rate of net photosynthesis in tomato (from Capron and Mansfield 1977). Note that in both cases combinations of pollutants have a greater effect than either gas does alone. SOURCE: Holdgate (1979b).

111 increased is the toxicity of metals, including copper, lead, zinc, and nickel {Spry et al. 1981~. The presence of elevated calcium at low pH, however, reduces the otherwise high toxicity of aluminum--i.e., it acts as an ameliorative for both terrestrial higher plants and for aquatic freshwater plankton (Hutchinson and Collins 19781. This appears to be due to competitive inhibition of aluminum uptake by calcium. Higher calcium concentrations also appear to cause Crustacea to be less sensitive to hydrogen ions (Figure 6.6~. Among toxic metals a rather wide range of interactions occur. Nickel and copper in lake waters contaminated by the Sudbury smelters act synergistically in their effects on various green algae (Hutchinson 1973, Hutchinson and Stokes 1975~. Stokes (1975) has shown that this effect relates to the initial separation of copper and nickel in the cell, with copper moving across the plasmalemma while nickel is bound to the cell wall. Eventually, sufficient damage occurs to the cell membrane that it becomes leaky and nickel enters also, this being the synergistic phase. Nickel-copper synergism has also been described for higher plants (Whitby and Hutchinson 1974~. In studies of the floating aquatic plants Lemna valdiviana and Salvinia natans it was found that, in addition, cadmium and zinc also - acted synergistically. Frond numbers increased with the tissues' concentrations of the two elements, and these were mutually enhanced by the presence of the second element (Hutchinson and Czyrska 1973~. Both light intensity and the specific range of concentration influenced the outcome. Water quality criteria based on laboratory tests of single pollutants appear to be inadequate for situations where several pollutants enter a water body. Wong et al. (1978) found that a mixture of heavy metals added to Great Lakes water or culture medium at the maximum permissible concentration levels for each was very toxic to freshwater algae. Diatoms were even more sensitive than either green or blue-green algae. Similar findings were reported by the ME:LIMEX study in Switzerland, where a mixture of metals was added to lake water at maximum permissible concentrations for each. Toxicity to phytoplankton (Gachter and Mares 1979) and heterotrophic microorganisms (Bossard and Gachter 1979) was reported for the mixture of metals, although singly the metals caused no observable damage. On the other hand, the metals caused populations of benthic organisms to increase, presumably because inhibition of the decomposer microbiota allowed more oxygen to accumulate at the mud-water interface (Lang and Lang-Dobler 1979~. Another example of synergistic effects of metals in animals is the strong teratogenicity of the combined injection of cadmium and lead compared to the effects of administration of either of the metals separately. In contrast, metal interactions can protect against the normal toxic effects of a single metal. Such interactions have been described for mammals and fish (Cherian and Goyer 1978, Nordberg 1978, Parizek 1978, Levander 1977, Groth et al. 1976, Ganther et al. 1973~; for invertebrates (Talbot and Magee 1978~; and for unicellular green algae (Hutchinson 1973~. The biochemical mechanism of the protective

112 I ° 1.00 an Lid at ~ 0.50 ° 0.30 0.23 B. .................. · · · e ~ · · . . AS E L LU S AD UAT I C U S: - : · ~ - - - ~ P R E S ~ N T . . · . · · . · · · . · · · · . · · · · · · ~ · ~ · · . . . ~ ~ . . · · · · · · · · ~ · · · · . . . · · · · · · . . · · · . . . . . . . · . · . . . . C1 · ~ I, n U. o 4.2 4.8 5.2 6.0 7.0 HYDROGEN- ION CONCENTRATION (pH) FIGURE 6.6 A model for tolerance limits of Asellus aquaticus toward low water hardness and low pH. The species is absent in lakes belonging to areas B. C, and D. Arrows indicate that limits may be moved through synergism with other factors. One degree dissolved hardness (1° dH) = 10 mg CaO/l. SOURCE: Okland (1980).

113 effect is often the induction of metallothionein, which acts nonspecifically to chelate metals (Kagi and Nordberg 1979~. For example, exposure of an organism to any one of cadmium, mercury, silver, copper, or zinc causes increased tolerance to any of the other metals upon subsequent exposure (Winge et al. 1975~. When the protective effect involves selenium (Ganther 1978), the ameliorative action is thought to involve the diversion of metals from the biochemical sites of toxic action (Bark et al. 1974, Komsta-Szumska et al. 19761. Previous exposure also affects the organisms' response to toxic concentrations of metals. Some metals, such as silver, arsenic, cadmium, mercury, lead, and tin are less toxic when sub-lethal doses are given prior to a dose that normally produces mortality. Others, such as barium, chromium, and iron, have toxicity enhanced by pretreatment (Yoshikawa 1970~. In cases where tolerance increases, the underlying mechanism may involve metallothionein (Leber and Miya 1976, Probst et al. 1977), selenoproteins (Sandholm 1974, Prohaska and Ganther 1977), glutathione (Congiu et al. 1979), or other mechanisms (Tendon et al. 1980~. Increased tolerance to metals may also be due to physiological mechanisms, such as increased excretion of metals induced by pre-exposures {Levander and Argrett 1969, McConnell and Carpenter 1971~. A mixture of detergents and mercury or cadmium was found by Calamari and Marchetti (1973) to have effects on rainbow trout different from the effects of detergents or metals alone. Anionic detergents plus metal had more than additive effects, while nonionic detergents plus metal had less than additive effects. On the other hand, phosphorus and nitrogen were found to increase the tolerance of algae for heavy metals, as demonstrated in the tolerance of Stigeoclonium tenue Kutz to heavy metals in South Wales (McLean 1974) and the increased tolerance of microbiota and algae to arsenic in waters enriched with nutrients (Brunskill et al. 1980~. As is well known from experience with human illnesses, stress often renders organisms susceptible to diseases of various types. For example, Hetrick et al. (1979) found that exposure of rainbow trout to copper increased their susceptibility to infectious hematopoetic necrosis and other viral diseases. The general relationship between pollution stress and the incidence of infectious diseases in fish was reviewed by Snieszko (1974~. In the marine environment, mercury contamination is accompanied by enhanced concentrations of selenium in all investigated species of mammals, birds, and fish--possibly due to a normal homeostatic regulation. It seems likely that selenium exerts a protective action against mercury toxicity in the marine environment, decreasing the detrimental effects of mercury on reproduction behavior, growth, etc., and thus protecting the population and the ecosystem. In freshwater ecosystems, high selenium retarded the uptake of mercury by freshwater organisms (Rudd et al. 1980~. On the other hand, Beijer and Jernelov (1978) have shown that high selenium increases the retention of mercury by aquatic organisms leading to bioaccumulation of mercury in fish and a higher body burden in the individual. This might

114 counteract the positive effect of selenium in decreasing mercury uptake and toxicity. In the environment, the presence of several pollutants and other kinds of stress is the norm. We must develop a fuller understanding of the mechanisms of interactions. Often, synergisms and antagonisms may occur simultaneously, so that no apparent effect may be detectable. For example, Eaton (1973) exposed Pimephales promelas to a mixture of copper, cadmium, and zinc for one year. First analyses showed the mixture to be no more toxic than zinc alone had been in earlier studies. When various effects were evaluated, however, it appeared that zinc toxicity was unchanged, while copper toxicity had increased and cadmium toxicity had decreased. Resistance to Pollutants There is considerable evidence that both individuals and populations can develop resistance to a wide variety of pollutants. At the individual level, increased resistance appears most often to be acquired through enzyme induction and the appearance of toxicant-binding proteins. The degree of resistance that can be acquired appears to vary both from one species to another and one toxicant to another, ranging from no increased tolerance to increases of several thousand times. Resistance to several toxic metals by a wide variety of aquatic organisms appears to be the result of synthesis of metallothionein (G. Brown 1976), a protein that reportedly binds metals to sulfhydryl groups and prevents them from diffusing across cellular membranes into sensitive tissues or interacting with critical enzymes (Cherian and Goyer 1978, Brown and Parsons 19781. This protein develops in the organism in response to exposure to sublethal concentrations of several metals--Hg, Cd, Cu. Zn, Ag, and Sn--(Winge et al. 1975) after which resistance to a number of metals is increased. The amount of resistance that can be developed is high in most fishes but low in zooplankton (Brown and Parsons 1978~. The protein has been shown to occur in most phyla of plants and animals (Brown 19801. Similar multipollutant resistance appears to occur with pesticides and chlorinated hydrocarbons. Exposure to sublethal levels of one chlorinated hydrocarbon often increases tolerance to others. In the case of DOT, three mechanisms have been found to be involved: (1) a membrane barrier to DOT passage, (2) a structural difference in myelin, and (3) an efficient blood-brain barrier. Wells and Yarbrough (1972) and Moffett and Yarbrough (1972) discuss the details of these mechanisms. A high intrinsic genetic variability in a population increases its chances of having some resistant survivors to a broad range of pollutant stresses. Species that occupy a wide variety of habitats under natural conditions tend to have high genetic variability. For example, the ubiquitous algal genera Chlorella and Scenedesmus will survive concentrations of heavy metals lethal to most other algal species (Stokes et al. 1973~.

115 Structures or behaviors of some organisms protect them from atmospheric pollutants (Jacobson 1980~. Waxy cuticles, surface hairs with hydrophobic properties, flat or narrow shapes, and vertical orientation tend to protect leaf surfaces from penetration of aqueous solutions of pollutants. Reproductive organs may be protected by flowers with inverted openings or flowers that close during cloudy periods. Other flowers open only when triggered by the weight of insects. Duration of pollination and number of pollinating flowers are also considerations. Salts on leaf surfaces or internal buffering reactions may also offer protection (Table 6.3~. Different genotypes in single plant or animal species are known to vary in their susceptibility to pollutants, as do different species within the same genus (Taylor 1978~. Animal populations also evolve characteristics consistent with a polluted environment. For example, in industrial areas of Britain, elimination of pollutant-sensitive lichens from the stems of trees caused the disappearance of light-colored peppered moths, which had relied on the lichens for camouflage from insectivorous birds. A previously rare soot-colored "melanic" variant of the same species replaced the light form as the dominant genotype. The melanic form was also better adapted to eating pollutant-laden vegetation (Kettlewell 1955, 1956~. Some resistance, including that of the melanic peppered moths and tolerance of ryegrass to coal-smoke pollutants, is known to have developed within the past century or two. The gene pools of organisms are well prepared to adapt to changing environmental conditions, as long as the changes occur slowly enough for adaptation to occur. This is usually a function of the length of the generation time for a given species. Microbes can complete a life cycle in hours; larger mammals and tree species may take decades to centuries. In areas of airborne fallout of contaminating metals and on old mine waste dumps, numerous examples of genotypes highly tolerant of normally toxic metal levels have been found. Such selection has been reported for a wide variety of higher plants, ferns, green or blue-green algae, bacteria, and fungi (e.g. Bradshaw 1952; Jowett 1958; Stokes et al. 1973; Tatsuyamo et al. 1975; Allen and Sheppard 1971; Cox and Hutchinson 1979, 1980) and for a number of metals and combinations of metals. Past Efforts to Predict Widespread Toxic Effects Both scientific induction and unplanned-for catastrophes have led to an understanding of biological effects of pollutants. Perhaps most illustrative is the example of DOT. Before its extensive broadcast about the environment as an insecticide following World War II, its potential effects upon nontarget organisms were recognized by several scientists, including Cottam and Higgins (1946~. They pointed out that agricultural applications would affect wildlife and game and were especially concerned about its entry to streams, lakes, and coastal bays, where the crabs and fish would be sensitive to this toxic halogenated hydrocarbon.

116 TABLE 6.3 Plant Processes and Characteristics That May Increase Tolerance to Acid Precipitation - Exclusion Leaf orientation and morphology Chemical composition of cuticle Flower orientation Protection of sexual organs Pollination mechanism Neutralization Salts on leaf surfaces Buffering capacity of leaves Metabolic Feedback Reactions Enzymatic reactions that consume hydrogen ions or yield alkaline products SOURCE: Jacobson ( 19 80).

117 Subsequent events have confirmed that higher trophic levels are jeopardized by exposure to DDT. One of the more publicized effects upon the marine ecosystem involved the reproductive failures of the brown pelican population on Anacapa Island, off the California coast, from 1969 to 1972. The accumulation of DDT and its degradation products, primarily DOE, in marine organisms that were the food for the pelicans initiated the problem. The source of the pesticide was allegedly wastes from a manufacturing plant in Los Angeles, which were introduced to the oceans in sewer outfalls. The pelicans produced thin, friable egg shells, which broke easily (Risebrough 1972~. DDT usage has been severely restricted since the early 1970s by the United States and many nations of the Northern Hemisphere. The brown pelican population on Anacapa Island has shown breeding recovery since 1972 when the DDT dumping was stopped. Despite ODE residues in the sediments, the success rate for young per nest had risen from 0.04 in 1972 to 0.88 in 1975 (Schreiber 1980~. This recovery is true in other locations also. With nuclear weapons development in the early 1950s, scientists from several countries became concerned about the entry of artificially produced radionuclides from power plants to the atmosphere and to the oceans. The ingestion of food containing radionuclides from the sea and the exposure of individuals to environmental radioactivity were minimized through limitations of the amounts of wastes emitted from nuclear plants to their surroundings. The greatest amounts of artificial radionuclides introduced to the environment come from the nuclear processing plant at Windscale, England. British environmental scientists have developed protocols to regulate the discharges on the basis of the "critical pathways technique," through which nuclides posing dangers to human health are identified and the critical population for exposure is protected. As a consequence of the resultant regulatory measures, there appears to be no danger to the population liable to exposure at the present time. SUMMARY Accumulation of pollutants in ecosystems is affected by a number of physical factors at the surface of the ecosystem, such as the amount of surface area presented by a forest canopy or turbulence near the air-water interface. The form of the pollutant--gas, aerosol, or large particulates--is also important. Ice and snow cover and thermocline formation may protect freshwater ecosystems from dissolved pollutants at certain times of the year, but can cause higher than average inputs at others. The thermocline regulates the passage of pollutants to and from the deep ocean. Soils and aquatic sediments tend to accumulate many pollutants, thereby lowering toxicity to pelagic or above-ground forms but raising toxicity for benthic or soil organisms. The binding in soils and sediments for some elements depends on oxidation-reduction potentials, which are lowered when anoxic conditions develop. Many organic constituents are transformed by microbes in soils or sediments into

118 other organic compounds, which might be either toxic or harmless, or into inorganic products, CO2 and H2O. Substances that are not degraded in terrestrial ecosystems will eventually be transferred to aquatic ecosystems via water flow or soil erosion; hence, higher concentrations of toxicants are likely to occur in aquatic ecosystems. Aquatic sediments often provide a reservoir of datable toxic deposits and fossil organisms, from which a history of pollution might be reconstructed. At the species or individual level, pollutants may be bioconcentrated, bioexcluded, or biomagnified by passage up the food chain. Gaseous pollutants have their greatest effect on terrestrial ecosystems, because the air-water interface offers to aquatic organisms some degree of protection from episodic events. When many pollutants are introduced to an ecosystem, effects are in many cases synergistic--i.e., effects are greater than additive. In some other cases less than additive effects have been noted. Natural selection has caused some species to develop resistance to pollutants. Such genetic adaptation requires many life cycles, and no examples are known for species with life cycles over one year long. Individual resistance may, however, develop in individuals exposed to sublethal concentrations of pollutants.

Next: 7 Studying the Effects of Atmospheric Deposition on Ecosystems »
Atmosphere-Biosphere Interactions: Toward a Better Understanding of the Ecological Consequences of Fossil Fuel Combustion Get This Book
×
Buy Paperback | $80.00
MyNAP members save 10% online.
Login or Register to save!
Download Free PDF
  1. ×

    Welcome to OpenBook!

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

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

    No Thanks Take a Tour »
  2. ×

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

    « Back Next »
  3. ×

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

    « Back Next »
  4. ×

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

    « Back Next »
  5. ×

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

    « Back Next »
  6. ×

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

    « Back Next »
  7. ×

    View our suggested citation for this chapter.

    « Back Next »
  8. ×

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

    « Back Next »
Stay Connected!