Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Lake Ecosystem Ecology: A Global Perspective
Lake Ecosystem Ecology: A Global Perspective
Lake Ecosystem Ecology: A Global Perspective
Ebook1,547 pages39 hours

Lake Ecosystem Ecology: A Global Perspective

Rating: 5 out of 5 stars

5/5

()

Read preview

About this ebook

A derivative of the Encyclopedia of Inland Waters, Lake Ecosystem Ecology examines the workings of the lake and reservoir ecosystems of our planet. Information and perspectives crucial to the understanding and management of current environmental problems are covered, such as eutrophication, acid rain and climate change. Because the articles are drawn from an encyclopedia, the articles are easily accessible to interested members of the public, such as conservationists and environmental decision makers.
  • Includes an up-to-date summary of global aquatic ecosystems and issues
  • Covers current environmental problems and management solutions
  • Features full-color figures and tables to support the text and aid in understanding
LanguageEnglish
Release dateMay 20, 2010
ISBN9780123820037
Lake Ecosystem Ecology: A Global Perspective

Related to Lake Ecosystem Ecology

Related ebooks

Earth Sciences For You

View More

Related articles

Related categories

Reviews for Lake Ecosystem Ecology

Rating: 5 out of 5 stars
5/5

1 rating0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Lake Ecosystem Ecology - Gene E. Likens

    Argentina

    Introduction to Lake Ecosystem Ecology: A Global Perspective

    Gene E. Likens Cary, Institute of Ecosystem Studies, Millbrook, NY December 2009

    The scientific discipline of limnology is focused on the study of inland waters, fresh and saline, large and small, and young and old. For example, the 5 Laurentian Great Lakes in North America contain some 20% of the Earth’s surface fresh water, and Lake Baikal in Siberia contains another 18% (Likens, 2009). The recently constructed Three Gorges Dam Reservoir in China, to be completely filled by the end of 2009, will contain about 40 km³, or about 8% of what Laurentian Lake Erie (484 km³) contains. Lake Baikal is the deepest and one of the oldest freshwater basins (> 20 Mya); Three Gorges Reservoir is one of the newest major basins.

    Whether large or small, young or old, shallow or deep, these water bodies function as ecosystems with strong and complicated interactions among all of the living and nonliving components of the ecosystem. Ecosystems are the basic units of nature (Tansley, 1935). In terms of area of water (< 1%) or volume of fresh water (< 0.3%) on our planet, inland lakes and reservoirs of the world are tiny, but in terms of interest and use for drinking, sanitation services, food, recreation, transportation and aesthetics, they provide hugely important resources and ecosystem services to humans.

    Limnology is a quintessential form of ecosystem science because comprehensive limnological understanding is based on information from biology, physics, chemistry, geology, hydrology, meteorology and so forth. The recent Encyclopedia of Inland Waters (Likens, 2009) incorporates all of these various disciplines and more in addressing and describing the inland-water ecosystems of the Earth.

    Limnology is a relatively old scientific discipline with a professional international society (the International Society of Limnology (SIL)) founded in 1922. From the beginning, SIL has focused on both theoretical and applied issues in aquatic ecosystems as do the topics of this volume.

    This volume focuses on 5 topics about the lentic, inland waters of our planet, that is, lake and reservoir ecosystems: 1. Introduction to Lake Ecosystem Ecology: A Global Perspective, 2. Lake Ecosystems: Structure, Function and Change, 3. Hydrodynamics and Mixing in Lakes, 4. Lakes and Reservoirs of the World, and 5. Lakes and Reservoirs: Pollution, Management and Services. The information and perspectives contained in this volume are highly relevant to the understanding and management of a variety of current environmental problems, such as eutrophication, acid rain and climate change.

    The articles in this volume are reproduced from the Encyclopedia of Inland Waters (Likens, 2009). I would like to acknowledge and thank the authors of the articles in this volume for their excellent and up-to-date coverage of these important topics in limnology.

    References

    Likens GE. Chapter 1 in Encyclopedia of Inland Waters. 2009a.

    Likens GE. Encyclopedia of Inland Waters. Elsevier; 2009b.

    Tansley AG. The use and abuse of vegetational concepts and terms. Ecology. 1935;16:284–307.

    Lake Ecosystems: Structure, Function and Change

    Lakes as Ecosystems

    W.M. Lewis    University of Colorado, Boulder, CO, USA

    Introduction of the Ecosystem Concept

    Scientific studies of lakes began as early as the seventeenth century, but at first were descriptive rather than analytical. Toward the end of the nineteenth century, measurements and observations on lakes became more directed. For example, the thermal layering of lakes was attributed to specific physical causes, and such phenomena as the movement of plankton in the water column were the subject of hypotheses that were tested with specific kinds of data.

    Comprehensive studies of lakes began with the work of Alphonse Forel (1841–1912) on Lac Léman (Lake Geneva), Switzerland, as well as other Swiss lakes. In a three-volume monograph (Le Léman: 1892, 1895, 1904), Forel presented data on a wide variety of subjects including sediments and bottom-dwelling organisms, fishes and fisheries, water movement, transparency and color, temperature, and others. Thus Forel, who introduced the term ‘limnology’ (originally the study of lakes, but later expanded to include other inland waters), demonstrated the holistic approach for understanding a lake as an environmental entity, but without application of an explicit ecosystem concept.

    The conceptual basis for studying lakes as ecosystems was first clearly given by Stephen Forbes (1844–1930; Figure 1) through a short essay, The Lake as a Microcosm (1887). Forbes was professor of biology at the University of Illinois in Champaign, IL, USA, and director of the Illinois Natural History Laboratory (subsequently the Illinois Natural History Survey), which was charged with describing and analyzing the flora and fauna of Illinois. Forbes realized that it was not possible to achieve a full understanding of a lake or, by implication, of any other environmental system such as a stream or forest, simply from knowledge of the resident species. Forbes proposed that the species in a particular environment, when interacting with each other and with the nonliving components of the environment, show collective (system) properties. The microcosm that Forbes described today would be called an ecosystem, although this term did not come into use until 48 years later through the work of the British botanist A. G. Tansley. By current usage, an ecosystem is any portion of the Earth’s surface that shows strong and constant interactions among its resident organisms and between these organisms and the abotic environment.

    Figure 1 Stephen A. Forbes. Reproduced from the Illinois Natural History Survey website, with permission from the Illinois Natural History Survey.

    Forbes not only described accurately the modern ecosystem concept, but also identified ways in which critical properties of ecosystems could be measured and analyzed. He named four common properties of lakes, each of which provides a cornerstone for the study of lakes and other ecosystems, as shown in Table 1. Although Forbes’s concepts have been renamed, they are easily visible in the modern study of lakes as ecosystems.

    Table 1

    Four key properties of ecosystems identified by S. A. Forbes, along with their modern nomenclatural counterparts and some examples

    Although the essay by Forbes now is considered a classic in limnology and in ecology generally, it caused no immediate change in the practices of limnologists or ecologists. Like many important discoveries in science, it was a seed that required considerable time to germinate.

    In studies of Cedar Bog Lake, Minnesota, for his Ph.D. at the University of Minnesota, Raymond Lindeman (1915–1942; Figure 2) gave limnologists the clearest early modern example of the study of lakes as ecosystems. Rather than focusing on a particular type of organism or group of organisms, which would have been quite typical for his era, Lindeman decided that he would attempt to analyze all of the feeding relationships (‘trophic’ relationships) among organisms in Cedar Bog Lake. Thus, his Ph.D. work extended from algae and aquatic vascular plants to herbivorous invertebrates, and then to carnivores, and conceptually even to bacteria, although there were no methods for quantifying bacterial abundance at that time. Lindeman’s descriptions of feeding relationships were voluminous but straightforward to write up and publish, but he sought some more general conclusions for which he needed a new concept.

    Figure 2 R. L. Lindeman and his famous study site, Cedar Bog Lake, MN. Reproduced from People of Cedar Creek website, with permission from Cedar Creek Natural History Area.

    Lindeman took a postdoctoral position with G. Evelyn Hutchinson at Yale University in 1942. Hutchinson had become a limnologist of note through his quantitatively oriented studies of plankton and biogeochemical processes in the small kettle lakes near New Haven. He would in subsequent decades become the world’s most influential limnologist, and part of his reputation grew out of his contributions to the field of biogeochemistry, an important tool of ecosystem science.

    Hutchinson made suggestions that no doubt were critical to Lindeman’s groundbreaking paper, The Trophic Dynamic Aspect of Ecology (1942, published in the journal Ecology), which now is recognized as a landmark in limnology and in ecology generally. Lindeman proposed a way of converting the tremendous mass of highly specific information for Cedar Bog Lake into a format that would allow comparisons with any other lake or even with other kinds of ecosystems. Building on the work of the German limnologist August Thienemann and the British ecologist Charles Elton, Lindeman organized the feeding relationships as a feeding hierarchy within which each kind of organism was assigned to a specific feeding level (trophic level). He then proposed that the feeding relationship represented by any given link in the food web be quantified as an energy flow. Thus, the total energy flow from level 1 (plants) to level 2 (herbivores) could be quantified as the summation of energy flows between all pairs of plants and herbivores; a similar estimate could be made for all other pairs of trophic levels. In this way, the flow of energy across the levels of the food web could be expressed in quantitative terms.

    The first important conclusion from Lindeman’s energy-based approach was that each transfer of energy between trophic levels is governed by the second law of thermodynamics, which requires that significant energy loss must occur each time energy is transferred. Thus, Lindeman demonstrated why food webs have relatively few trophic levels: progressive dissipation of energy as it passes through the food web from the bottom (plants) to the upper levels (upper-level carnivores) ultimately provides insufficient energy for expansion of the food web to further levels. Also, analysis of a food web in this way sets the stage for calculating efficiencies of energy transfer, comparison of efficiencies across different ecosystem types, and the identification and analysis of bottlenecks restricting the flow of energy within the food web.

    The contributions of G. Evelyn Hutchinson in the 1940s on the biogeochemistry of carbon in lakes also must be counted as landmarks in the development of ecosystem science. Even so, ecosystem science was scarcely represented in the research agenda of ecologists or in the academic curriculum as late as 1950. Penetration of the ecosystem thinking into research, teaching, and public awareness occurred first through the publication of a textbook, Fundamentals of Ecology (1953), written by Eugene P. Odum, and especially through the second edition of the same book (1959), written by E. P. Odum and Howard T. Odum. The Odums visualized ecology as best viewed from the top down, with ecosystems as a point of departure and studies of ecosystem components as infrastructure for the understanding of ecosystems.

    The ecosystem perspective has not displaced the more specialized branches of ecology that deal with particular kinds of organisms or specific kinds of physical phenomena, such as studies of water movement, optics, or heat exchange. Rather, ecosystem science has had a unifying effect on studies of ecosystem components (Figure 3). Study of a specific ecosystem component produces not only a better understanding of that component, but also a better understanding of the ecosystem, which is a final objective for the science of an ecosystem type, such as lakes.

    Figure 3 Ecosystem diagram of a lake.

    Metabolism in Lakes

    The dominant anabolic component of metabolism in lakes is photosynthesis based on carbon dioxide and water plus solar irradiance as an energy source. The dominant catabolic component is aerobic respiration based on the oxidation of organic matter with oxygen as an electron acceptor. Conversion of solar energy to stored chemical energy in the form of biomass seldom reaches 1% efficiency in lakes because of losses inherent in the wavelengths that can be intercepted by photosynthetic pigments, inefficiency in the interception process, and thermodynamic losses in the conversions leading to the production of biomass. Respiration also involves thermodynamic losses. Therefore, the solar energy source greatly exceeds photosynthetic output, and cellular capture of the energy released from organic matter by respiration greatly exceeds the energy stored in the organic matter.

    Aerobic photosynthesis is characteristic of the aquatic vascular plants, attached algae, and phytoplankton of lakes and is universal wherever light and oxygen are present. Aerobic respiration is characteristic of aerobic autotrophs as well as consumers and most bacteria; it occurs wherever oxygen is present (Table 2). Other categories of metabolism occur under either of two more restrictive conditions: (1) where light penetrates into an anaerobic zone, and (2) where there is an interface or mixing between oxidizing and reducing conditions, as is common near a sediment–water interface.

    Table 2

    Summary of metabolic processes in lake ecosystems

    A study of whole-ecosystem metabolism would require consideration of all of the metabolic categories listed in Table 2. The dominance of aerobic photosynthesis and aerobic respiration, however, often allows ecosystem studies to focus on these two metabolic components. In open water, photosynthesis and respiration are often measured with a vertical series of paired transparent and darkened incubation bottles filled with lake water; rate of decline of oxygen in the dark bottles indicates respiration rate, and rate increase of oxygen in the transparent bottle indicates net photosynthesis. In unproductive lakes, uptake of ¹⁴C-labeled CO2 can be used as an even more sensitive indicator of photosynthesis. Where macrophytes or attached algae are important, as is often the case in small lakes, separate measurements must be made of their photosynthesis, typically by the use of enclosures. The metabolic rates of microbes in deepwater sediments can be inferred from the rate of oxygen loss from the hypolimnion of a stratified lake, or can be measured with enclosures.

    Annual rate of photosynthesis and respiration per unit area (typically given as mg C m−2 year−1) is the most commonly used metabolic statistic for lakes. A complete metabolic accounting would also include processing of organic matter entering a lake from the surrounding terrestrial environment, primarily through stream flow.

    Total annual net production of autotrophs (production in excess of respiration) is a measure of the capacity for a lake to generate biomass at higher trophic levels (Figure 4).

    Figure 4 Example of an early metabolic (energy-flow) diagram for an aquatic ecosystem. Reproduced from Odum H T (1956) Limnology and Oceanography 1: 102–118, with permission from the American Society of Limnology and Oceanography.

    Also, the time-course of production, which shows seasonal and nonseasonal variation, is often a useful point of departure for the analysis of mechanisms that control biotic functions in a lake. Factors that may suppress production include deep mixing of the water column, exhaustion of key nutrients, grazing, and hydraulic removal of biomass.

    Photosynthesis and respiration often respond differently to a specific physical or chemical change.

    Food-Web Analysis

    Modern food-web analysis follows the example given by Lindeman. The sophistication of the analysis is much greater today than it was in 1942, however, and the uses of food-web analysis have diversified as well.

    At the base of the food web is organic matter generated within a lake or coming to a lake from its watershed and atmosphere above (Figure 5). All of these sources of organic matter can be quantified, as explained in the preceding section, but require the application of several methods and must take into account both spatial and temporal variability in the synthesis or transport of organic matter.

    Figure 5 A modern diagrammatic view of a lacustrine food web, including both microbial and nonmicrobial components. (from Weisse and Stockner, 1993 as modified by Kalff, 2002; with permission)

    Quantification of linkages in the food web begins with feeding relationships between autotrophs (primary producers) and herbivores (primary consumers). Analysis of gut contents is one method for establishing the linkage between a specific kind of consumer and one or more plant foods. A linkage drawn in this way has two disadvantages, however: (1) it is not quantitative because food items often cannot be identified fully, and (2) it is subject to errors of interpretation caused by the ingestion of foods that are not assimilated or only partially assimilated through the gut wall. The first problem can be overcome either by the use of feeding experiments or by the quantification of growth rate of the consumer, with some empirically-based assumptions about the growth efficiency of the consumer. The second problem can be resolved by experimental use of tissue labels (typically isotopes) or, more efficiently, by the use of stable isotopes as passive tracers (i.e., relying on the natural abundance of stable isotopes to infer food sources). Because primary producers of different categories may differ substantially in their concentration of the stable carbon isotope ¹³C, analysis of ¹³C content of protoplasm from the consumer may allow a quantitative estimation of the relative importance of several possible foods contributing to the synthesis of biomass by the consumer.

    Above the level of primary consumers are carnivores, which may be secondary, tertiary, or even quaternary consumers, depending on their food source. Measurements of growth rate, along with gut-content analysis and use of passive or active tracers, can be applied to carnivores just as they are to herbivores. Within the carnivore trophic levels, however, the assignment of consumers to a specific trophic level may be difficult because carnivores often consume foods belonging to more than one trophic level. The stable nitrogen isotope ¹⁵N is useful in assigning a species to a fractional position on the food web because ¹⁵N shows increasing tissue enrichment from one trophic level to the next.

    When completed, a food-web analysis shows the efficiency of energy transfer from one level to the next. Because of the thermodynamic limits on efficiency and the typical observed efficiency, which are a matter of record, a food-web analysis based on energy shows whether particular linkages are unusually weak or strong by comparison with expectations. Such observations in turn lead to hypotheses about mechanisms of control for energy transfer within the food web. For example, modern ecosystem theory includes the concept of ‘trophic cascades’ involving ‘top-down’ and ‘bottom-up’ effects on trophic dynamics, which is easily applicable to lakes. Change in one trophic level may be visible in other trophic level, in the manner of a cascade (Figure 6). Top-down effects pass from any trophic level to the next trophic level below. For example, an unusual abundance of algal biomass in a lake could be traced to unusually efficient removal of herbivores through predation at such a rate as to leave algae mostly uneaten (a top-down effect).

    Figure 6 Contrasts in nutrient response for lakes that have strong control by top carnivores (lower line) and lakes that have weak control by top carnivores (upper line). The top carnivore exercises top-down control by suppressing smaller fish that eat large grazing zooplankton. Survival of more large zooplankton holds phytoplankton in check, thus weakening response to plant nutrients (P, N). Reproduced from Carpenter S R (2003) Regime Shifts in Lake Ecosystems, with permission from the International Ecology Institute, Oldendorf/Luhe, Germany.

    Similarly, bottom-up control passes from any trophic level to the next level above. The inability of an oligotrophic (nutrient-poor) lake to grow substantial amounts of plant biomass, e.g., exerts a bottom-up effect on all higher trophic levels by restricting the amount of energy that is available at the base of the food chain.

    Trophic-dynamic analysis also leads to other fundamental questions involving the structure of biotic communities. For example, food webs might or might not be more productive or more efficient with a high diversity of herbivores than with strong dominance from a very successful, specific type of herbivore. Such questions bear not only on the analysis of natural ecosystems, but also on ecosystem management.

    Biogeochemistry

    Many of the functional attributes of lake ecosystems can be analyzed through biogeochemical studies. While metabolism and trophic dynamics are viewed in terms of energy flux, biogeochemistry typically is viewed in terms of mass flux. As with energy analysis, the foundation is basic physics: mass is neither created nor destroyed (under conditions that are compatible with the presence of life). The conservation of mass leads to the mass-balance equation, which can be given as follows for an ecosystem: I − O = ΔS, where I is input of mass of a particular type (carbon or phosphorus, for example), O is output of mass, and ΔS is change in storage of mass. Input and output pathways for lakes are either hydrologic (flow) or atmospheric (deposition or gas exchange). Over the short term, change in storage may involve changes in concentration of mass in the water column, but over the long term, change in storage mostly reflects accumulation of mass in sediments. Any element can be a target for biogeochemical studies, but the most frequently studied biogeochemical processes in aquatic ecosystems involve carbon, phosphorus, and nitrogen.

    The mass-balance equation applies not only to the entire ecosystem, but also to compartments within the ecosystem (Figure 7). For example, carbon in the water column could be partitioned into compartments, each of which would have its own mass-balance equation. Some of these compartments would have a high turnover rate, while others would not. The dynamics of compartments explain restrictions on specific biological processes such as photosynthesis, and account for differences in individual lakes or categories of lakes, such as those that differ greatly in nutrient supply.

    Figure 7 Compartment matrix for analysis of carbon cycling in the water column of a lake.

    Studies of the carbon cycle are ideal complements to studies of lake metabolism and food webs. CO2, which is the feedstock for photosynthesis, enters aquatic autotrophs either in the form of free CO2 (CO2 + H2CO3) or bicarbonate (HCO−3). CO2 is converted to organic matter by the reduction process of photosynthesis. Either within an autotroph or through consumption of autotroph or other biomass by consumers or decomposers, reduced carbon is reconverted to its oxidized form as CO2 through aerobic respiration.

    Organic carbon resides not only within organisms, but also in water or sediments in dissolved or nonliving particulate form. Dissolved organic carbon derives from the watershed, atmosphere, or organisms within the lake. Watershed contributions to dissolved organic carbon in lakes are composed to a large extent of humic and fulvic acids, which are the byproduct of the degradation of organic matter within soils. Humic and fulvic acids are refractory (resistant to breakdown), although they are slowly decomposed by microbes and can be broken down by ultraviolet light in a water column. Humic and fulvic acids are generally present in concentrations of 1 − 10 mgl−1 in lakes and, when present at concentrations above 5 mgl−1, generally impart a brown or orange color to the water.

    Organic compounds that are released to the water column of lakes by the resident organisms vary greatly in composition. All soluble organic compounds present in organisms can be found in the water column at measurable concentrations. Thus, a careful analysis of lake water would show a wide variety of amino acids, carbohydrates of varying complexity, and other metabolites of organisms. These compounds enter the water column through leakage (excretion), death, or the production of fecal matter. As might be expected, the turnover rate for metabolites is generally high because most of these compounds are labile (easily used), in contrast to humic and fulvic acids.

    Particulate organic matter present in a water column may be either living or nonliving. Frequently, these two compartments are physically joined, as all nonliving particulate organic matter in water is colonized by bacteria and, under some conditions, fungi. Most of the carbon attributable to living organisms is accounted for by phytoplankton and zooplankton; fish and bacteria make smaller contributions in the sense of mass but have important ecosystem effects. Carbon is continually stored in sediments, which accumulate at rates often about 1 mm per year, much of which is organic. As sediments become buried by additional sediments, their decomposition slows because of the lack of oxygen and chemically hostile conditions for active metabolism of bacteria. Thus, while some of the organic matter in sediments is remobilized into the water column, other organic matter in the sediments becomes long-term storage.

    A study of mass flux and storage, when viewed in terms of ecosystem compartments and subcompartments, tells a story about the mechanisms by which a lake functions. For example, lakes may differ greatly in the terrestrial contribution to carbon processing and carbon storage, and also may differ greatly in speed of carbon turnover in specific living and nonliving compartments.

    The cycling of phosphorus in lakes has been studied intensively, because phosphorus is one of the two elements (nitrogen is the other) most likely to be critically depleted by aquatic autotrophs. Thus, phosphorus at the ecosystem level is commonly viewed as an ecosystem regulator; lakes in which it is abundant (eutrophic lakes) have potential to produce high amounts of plant biomass (phytoplankton, attached algae, or aquatic vascular plants) and have water-quality characteristics that are often viewed as undesirable (low transparency, green color, potential for odor production, and severe loss of oxygen in deep water). In contrast, lakes having low supplies of phosphorus (oligotrophic lakes) may show constant suppression of plant growth through phosphorus scarcity. Such lakes have much lower amounts of carbon in the water, higher transparency, often appear blue, and retain oxygen in deep water consistent with the requirements of fish and other eukaryotes.

    Because the phosphorus requirement of plants is only approximately 1% of dry biomass, scarcity of phosphorus can be offset by relatively modest increases in the phosphorus additions to a lake. Addition of 1 kg of phosphorus per unit volume or area, e.g., could easily generate 100 kg of dry mass or 500 kg of wet mass of autotrophs. Thus, mobilization of phosphorus by humans has the potential to change lakes and has done so throughout the world wherever human populations liberate phosphorus through waste disposal, agriculture, and disturbance of soil. For this reason, the study of trophic state (nutrient status) has received more attention than any other ecosystem feature of lakes. The practical application of modeling or analysis of trophic status in lakes arises through the desire to prevent changes in trophic state or to reverse changes in trophic state that have already occurred, which requires ecosystem-level understanding of the lake and particularly of its nutrient budgets.

    Lakes may also be limited by low concentrations of the forms of inorganic nitrogen that are readily available to aquatic autotrophs (ammonium, nitrate). In this case, nitrogen limitation rather than phosphorus limitation may control plant growth. In fact, the two types of nutrient limitation may occur sequentially across seasons or across years in a single lake.

    Nitrogen cycling in lakes is much more complicated than is phosphorus cycling because nitrogen has a gaseous atmospheric component that phosphorus lacks, and because nitrogen exists in seven stable oxidation states within a lake, which sets the stage for the use of nitrogen as a substrate for oxidation reduction reactions supporting microbial metabolism (Table 2). Like phosphorus, nitrogen enters lakes through the watershed and, in small amounts, with precipitation. Unlike phosphorus, nitrogen also enters lakes as a gas by diffusion at the air-water interface in the form of N2. N2 is so inert chemically that it cannot be used as a nitrogen source by most organisms. Certain prokaryotes have the ability to use N2 by converting it to ammonium, which then can be used in organic synthesis, provided that a substantial energy supply is available. This process is called nitrogen fixation, in that it converts the gaseous nitrogen to a solid that is soluble in water (ammonium). The dominant nitrogen fixers in lakes are the cyanobacteria. Not all cyanobacteria fix nitrogen, but certain taxa that have a specialized nitrogen-fixing cell (heterocyst) grow commonly in lakes that show nitrogen depletion.

    Nitrogen fixers escape the limitation of growth associated with nitrogen depletion, thereby gaining an advantage over other autotrophs. Thus, lakes that are enriched with phosphorus but showing nitrogen depletion often have nuisance growths of nitrogen-fixing cyanobacteria, which produce all of the expected symptoms of nutrient enrichment, and sometimes also produce toxins. There is currently much interest in predicting and preventing the development of circumstances that lead to large and persistent blooms of nitrogen-fixing cyanobacteria.

    Nitrogen is a multiplier element, just as phosphorus is. It constitutes approximately 5% of dry mass in plants. Therefore, when it is added to lakes, provided that phosphorus is added at the same time, it supports a 20-fold multiplication of dry plant biomass.

    A detailed documentation of the carbon, nitrogen, and phosphorus cycles for a lake produces an understanding of factors that regulate the metabolic rates of lakes and the accumulation of biomass of various kinds in lakes. These ecosystem phenomena have numerous practical applications, ranging from interest in biomass production (e.g., fish) to interest in constraining water quality within boundaries that are either natural or that favor human purposes.

    Community Organization

    As foretold by Forbes, ecosystems have a strong degree of spatial organization involving the arrangement of organisms according to abotic constraints based on factors such as amount of irradiance, concentration of dissolved oxygen, or substrate characteristics. The requirement for irradiance is especially important because it dictates the distribution and growth potential for aquatic autotrophs. Phytoplankton grow strongly only within the euphotic zone, which corresponds approximately to water receiving at least 1% of the solar irradiance available at the water surface. Thus, vigorously growing phytoplanktons often are confined to the upper part of the water column (epilimnion and sometimes metaliminon). The same principle applies to periphyton and rooted aquatic plants, which can occupy substrates only over portions of the sediment that are exposed to at least 1% surface irradiance. The gradient of irradiance is controlled by the transparency of the water.

    Spatial organization of animal communities and microbial communities is dictated partly by physical conditions and partly by the distribution of autotrophs. Crustacean zooplankton, e.g., may migrate large distances vertically in the water, hiding in deep water during the daytime but rising to feed within surface waters at night. The zooplankton of the littoral zone differs from that of open water, in that the littoral zone offers food types (periphyton especially) that are not available in open water. Distribution of fishes may be dictated in part by the presence of structure associated with littoral zone. The effect of structure may even be related to life history in that immature or small fish may seek the structure of the littoral zone for shelter from predation.

    Because organisms conduct the business of an ecosystem, an understanding of their habitat requirements, reflected as spatial organization, leads to explanations of the total abundances of certain categories of organisms, the failure of certain organisms to fare well in one lake but not in another, and the consequences of habitat disturbances of natural or human origin.

    Synthetic Analysis of Ecosystems

    Synthetic analysis of ecosystems often involves the statistical study of empirical information, particularly if the objective is to test a particular hypothesis. A quantitative overview of multiple ecosystem functions or of the intricate detail for an ecosystem component typically involves modeling as a supplement to other types of quantitative analysis. When computers first became available, it was anticipated that ecosystem science would ultimately be driven almost entirely by mechanistic models that would predict ecosystem responses to natural or anthropogenic conditions. These expectations were not realized. Ecosystems, like other complex systems such as climate or economics, are moderately predictable from modeling that combines a modest number of well-quantified variables addressed to a specific question, but typically are unpredictable on the basis of large number of variables addressing more general questions. Even so, modeling of numerous variables in an ecosystem context can be useful in setting limits on expected outcomes or showing possible outcomes of multiple interactions that cannot be easily discerned from the study of individual variables. Therefore, modeling is useful in promoting the understanding of ecosystems.

    Conclusion

    Lakes first inspired the ecosystem concept, and have been a constant source of ideas about ecosystem structure and function. Ecosystem science as applied to lakes, is supported by and is consistent with, other kinds of ecological studies that are directed toward specific organisms, groups of organisms, or specific categories of abiotic phenomena in lakes. The strength of ecosystem science lies in its relevance to the understanding of all subordinate components in ecosystems, and to its often direct connection to human concerns in understanding and managing lakes.

    Glossary

    Biogeochemistry   Scientific study of the mass flux of any element, compound, or group of compounds within or across the boundaries of an ecosystem or any other spatial component of the environment. Mass flux within an ecosystem is often designated as nutrient cycling or element cycling.

    Cycling   A biogeochemical term referring to the movement of elements or compounds within an ecosystem or any other bounded environmental system.

    Ecosystem   Any portion of the earth’s surface that shows strong and constant interaction among its resident organisms and between those organisms and the abotic environment.

    Lake trophic state   Fertility of a lake, as measured either by its concentrations of key plant nutrients (especially phosphorus and nitrogen) or the annual production of plant biomass (aquatic vascular plants and algae). Trophic-state categories include oligotrophic (weakly nourished), mesotrophic (nutrition of intermediate status), and eutrophic (richly nourished).

    Mass flux   Movement of mass per unit time, often in ecosystem studies expressed as mass/area/time.

    Stable isotope   Any isotope of an element that does not decay spontaneously.

    Trophic dynamics   Fluxes of energy or mass caused by feeding relationships, including rates of grazing by herbivores on plant matter, rates of predation by carnivores on other animals, or rates of decomposition of organic matter by microbes.

    See also: Modeling of Lake Ecosystems; Trophic Dynamics in Aquatic Ecosystems.

    Further Reading

    Belgrano A, Scharler UM, Dunne J, Ulanowicz RE. Aquatic Food Webs: An Ecosystem Approach. New York: Oxford University Presss; 2006.

    Carpenter SR. Regime Shifts in Lake Ecosystems: Pattern and Variation. Oldendorf/Luhe, Germany: International Ecology Institute; 2003.

    Golley FB. A History of the Ecosystem Concept in Ecology: More than the Sum of the Parts. New Haven/London: Yale University Press; 1993.

    Kalff J. Limnology: Inland Water Ecosystems. Upper Saddle River, NJ: Prentice Hall; 2002.

    Ecological Zonation in Lakes

    W.M. Lewis    University of Colorado, Boulder, CO, USA

    Introduction

    Lakes show many kinds of spatial variation in both vertical and horizontal dimensions. Variation can be chemical, physical, or biotic, and is important to the understanding of ecosystem functions. Although some types of variation are unique to specific classes of lakes, others are common to most lakes, and correspond to an obvious spatial organization of the biota in lakes. The existence of certain common types of spatial organization in lakes has led to the naming of specific zones that have distinctive ecological characteristics.

    A complete list of all zones that have been named by limnologists would be lengthy and complex (Wetzel, 2001). Complex nomenclatures have been abandoned by modern limnologists, however. Modern limnology focuses on simple zonation systems that are easily applied by limnologists and others interested in lakes. Table 1 gives a summary of zonation systems that are currently in broad use.

    Table 1

    Summary of the four major zonation systems for lakes

    Horizontal Zonation: Littoral and Pelagic

    The nearshore area of a lake (littoral zone) differs from the offshore shore area (pelagic zone). The only group of autotrophs in the pelagic zone is the phytoplankton, which consists of very small algae that are suspended in the water column. The littoral zone also has phytoplankton (which move freely between littoral zone and pelagic zone), but also has two other categories of autotrophs (Figure 1): aquatic vascular plants (aquatic macrophytes), and films of attached algae (periphyton). Periphyton grow on the leaves of macrophytes and on other solid surfaces such as mud, sand, rocks, or wood.

    Figure 1 Depiction of the littoral and pelagic zones of a lake. The littoral zone extends outward from the shoreline to approximately the location at which the solar irradiance at the bottom of the lake corresponds to about 1% of the solar irradiance at the top of the water column. Within the littoral zone, growth of aquatic macrophytes and attached algae (periphyton) is possible. The pelagic zone begins at the outer margin of the littoral zone. Phytoplankton are exchanged freely between the littoral and pelagic zones as well.

    The outer margin of the littoral zone, beyond which is the pelagic zone, is the point at which significant growth of macrophytes and periphyton becomes impossible because of darkness. This boundary corresponds approximately to the location at which the amount of solar irradiance reaching the bottom of the lake is < 1% of surface irradiance. At bottom irradiances < 1%, there is little or no net photosynthesis, which prevents growth of the attached autotrophs (macrophytes and periphyton) that are typical of the littoral zone.

    For a lake of a given size and shape, the width of the littoral zone depends on transparency of the water as well as shoreline slope (Figure 2). In oligotrophic lakes, which have low nutrient concentrations and therefore develop very small amounts of the phytoplankton biomass that could shade the lower water column of lakes, the littoral zone extends to depths of 4–20 m or even more, depending on transparency of the water. In the eutrophic category, the depth of 1% irradiance ranges between 0.1 and about 2 m, and the mesotrophic category spans ~ 2–4 m. In the most extreme cases of eutrophication (lakes highly enriched with nutrients), where the depth of 1% light corresponds to only a few centimeters, the littoral zone as defined by light is virtually absent, and the littoral zone may be defined instead by the zone of influence for traveling waves and corresponding disturbance of the bottom (0.5–1.5 m). In general, small lakes have a higher percentage of surface area in the littoral zone than do large lakes (Figure 2), although some large, shallow lakes have large littoral zones (e.g., Lake Okeechobee, Florida).

    Figure 2 A plan view of the littoral and pelagic zones of lakes, illustrating variation in the width of the littoral zone associated with changes in slope along the margin of any given lake, and the tendency of smaller lakes to have a higher percent areal coverage of littoral zone for the total lake area.

    Although littoral zones are most easily defined on the basis of macrophytes and attached algae, a littoral zone can also be distinguished from a pelagic zone by its distinctive heterotrophic communities and by its food-web structure. Because littoral zones provide shelter, whereas pelagic zones do not, littoral zones often support dense populations of organisms that thrive when protected from predation. Larval and juvenile fish, for example, seek shelter within the littoral zone from predation by larger fish. Large invertebrates, such as dragonfly larvae or crayfish, typically are most abundant in littoral zones, where they are least likely to be consumed by fish. The littoral zone also has invertebrate communities that specialize in the consumption of attached algae by nipping or scraping the algal coatings on macrophytes or other solid surfaces. In the pelagic zone, there is no food source comparable to the periphyton of a littoral zone. In general, the communities of a littoral zone are more diverse than those of the pelagic zone, and the key species of the two zones differ.

    In the pelagic zone of a lake, the autotroph community is composed of phytoplankton (Figure 1), which are adapted for life in an environment that is free of solid surfaces. Consumers, such as zooplankton, living and reproducing in the pelagic zone must escape predators by avoiding the upper, illuminated part of the water column during the day, or must be agile or so small as to be impractical as a food for many predators.

    Vertical Zonation: Water Column, Sediments, and the Benthic Interface

    Lakes have a vertical zonation consisting of the water column, underlying lacustrine sediments (lake sediments), and the benthic zone, which occupies a few centimeters above and below the sediment-water interface (Figure 3). The water column extends across both the pelagic and littoral zones. The water column of the pelagic zone is driven by wind-generated currents into the littoral zone where water is displaced from the littoral zone into the pelagic zone. Thus, water-column constituents such as dissolved gases, dissolved solids, suspended solids, and suspended organisms are constantly exchanged between the pelagic zone and the littoral zone whenever there are currents in the top few meters of a lake. Chemical differences between the top few meters of the pelagic zone and the littoral zone may develop under the influence of biological processes, however, when currents are weak.

    Figure 3 Illustration of the division of a lake into three vertical components: water column, lacustrine sediments, and the boundary between the water column and sediments (benthic zone).

    Although the water column is shared by the pelagic zone and the littoral zone, lacustrine sediments always underlie the pelagic zone but may or may not cover all of the littoral zone. Sediments are produced by the settling of mineral and organic matter that is derived from the watershed of a lake, and from organic matter consisting of fecal pellets, organic debris (detritus), and skeletal fragments of organisms derived from the lake itself. A constant sedimentation of this fine mixed solid material occurs over the entire lake. Disturbance of sediments by moving water occurs primarily in shallow water, where most of the energy of wind-generated currents and traveling waves are expended against the bottom of the littoral zone. Thus, energy near the shore may cause fine sediments, such as those that are characteristic of lakes, to be swept to deeper water. For this reason, lacustrine sediments may not accumulate in all parts of a littoral zone. Alternatively, in lakes that are small or strongly sheltered from wind, and thus not subjected to extremes of wind-generated disturbance, lacustrine sediments may occupy most or all of the littoral zone.

    Lacustrine sediments are capable of supporting eukaryotic organisms (algae, protozoans, invertebrates, vertebrates) only when they are oxic. When the hypolimnion is oxic, the top few millimeters of sediment often (but not always) will be oxic. Below the top few millimeters, there typically is a decline in oxygen because microbial respiration supported by organic matter in the sediments leads to the depletion of oxygen, but some oxygen (e.g., 50%) may persist because invertebrates in the sediment pump oxygen through small tunnels into the sediment to as much as 10–20 cm within the sediments. Almost all of the deeper sediments (> 10–20 cm) in lakes, which may be many meters thick, are anoxic and can support only microbes that are capable of anaerobic metabolism. When the water of the hypolimnion is anoxic, the entire sediment profile is anoxic, and can support only anaerobic microbes. There are many such microbes, and anoxic sediments show strong evidence of their metabolism, including accumulation of reduced substances such as ferrous iron, sulfide, and methane. At progressively greater depths in sediments, however, the metabolism of microbial anaerobes slows because the easily used portions of organic matter are exhausted or because oxidizing agents such as sulfate or nitrate may be depleted. Thus, there is a decline in microbial metabolic rate from the upper sediments to the deepest sediments, which are almost inert biologically.

    The interface between the water column and the lacustrine sediments carries its own name (‘benthic zone’) because it is exceptionally important from the ecosystem perspective, despite its narrow dimensions. Organisms that live on the sediment surface or just below it (down to about 20 cm) carry the name ‘benthos’. In sediments below the pelagic zone, the benthos does not include autotrophs because there is no light reaching these sediments. The benthic zone is rich in invertebrates, provided that it is oxic at the surface, which is not always the case. Oxic benthic zones often support a number of important invertebrates, most of which are embedded within the sediment, as necessary to avoid predation. Examples include midge larvae and the larvae of other insects (Figure 3). Certain fish species (e.g., catfish) may be associated closely with the benthic zone, in that they are adapted to find and consume the embedded invertebrates by chemosensory means, without using vision. Oxic benthic zones also support protozoans and bacteria conducting oxic metabolism, including especially the oxic breakdown of organic matter. Anoxic benthic zones only support anaerobic bacteria and a few specialized protozoans.

    The benthic zone extends not only across the bottom of the pelagic zone but also across the bottom of the littoral zone (Figure 3). The benthic component of the littoral zone includes not only the interface between lacustrine sediments and the water column but also between the water column and any parts of the littoral zone that happen to be swept free of lacustrine sediments. Thus, the entire solid surface at the bottom of a lake lies within the benthic zone.

    Seasonal Zonation: Vertical Layering Based on Density

    Because the temperature of water affects its density, it is common for lakes to develop layers of different density corresponding to temperature differences across the layers. At temperate latitudes, all but the shallowest lakes develop a density stratification during spring that typically persists until late fall. Similar seasonal stratification is also common in subtropical and tropical lakes, but the duration of stratification is longer and the nonstratified period (mixing period) does not contain an interval of ice cover, as it often does at temperate latitudes.

    Density stratification causes an ecologically important vertical zonation of lakes (Figure 4). An upper layer, which contains the air–water interface, is the epilimnion of a stratified lake; it may also be referred to as the ‘mixed layer’. The epilimnion is the warmest and least dense of the three layers. Its thickness is strongly influenced by the size of the lake, in that larger lakes show a higher transfer of wind energy to water currents, which thickens the mixed layer during its period of formation. In sheltered water bodies, mixed layers may be as thin as two meters, but in larger (> 10 km²), windswept water bodies, they may be as thick as 15–20 m. In addition, the thickness of a mixed layer in a given lake increases over the fall cooling period, during which the bottom of the mixed layer erodes the layer below.

    Figure 4 Illustration of the three density (temperature) layers determined by seasonal stratification of a water column.

    The mixed layer often shows sufficient irradiance throughout its full thickness to support photosynthesis. Even in lakes that have very low transparency, the uppermost portion of the mixed layer is well illuminated.

    The mixed layer extends continuously across the pelagic zone and littoral zone. Only in highly transparent lakes does the littoral zone extend below the mixed layer. Within the pelagic portion of the mixed layer, zooplankton herbivores feed vigorously on phytoplankton, but may move downward out of the mixed layer during the day in order to avoid predation. Other small invertebrates often migrate out of the mixed layer during daylight hours as well. Most of the energy of wind transferred to water is dissipated within the mixed layer; this energy contributes to the high degree of uniformity in the mixed layer at small to intermediate distance scales (up to 10 km or more), except during extended calm weather.

    Below the mixed layer of a stratified lake is a thermal gradient, corresponding to a density gradient. The gradient makes a transition in temperature and density between the mixed layer and the coolest layer, which lies in contact with the bottom of the lake. The thermal transition is referred to as the thermocline, but the layer within which the thermocline lies is best referred to as the metalimnion (Figure 4). The metalimnion may or may not receive sufficient irradiance to support photosynthesis. In lakes that are quite transparent, phytoplankton often grows in the metalimnion, whereas lakes of low transparency seldom show growth of autotrophs in the metalimnion. In the pelagic zone, phytoplankton growing in the metalimnion of a transparent lake may be of different species composition than phytoplankton growing in the mixed layer. Some mass transfer occurs between the epilimnion and the metalimnion, the amount of which is dependent on the amount of turbulence at the interface of the two layers. The thickness of the metalimnion varies a great deal among lakes. It is seldom thinner than 2 m, and may be as thick as the mixed layer or even thicker.

    The deepest seasonal layer in lakes is the hypolimnion. The temperature and density of the hypolimnion typically reflect conditions that occur when seasonal stratification becomes established. It is common for stratified lakes at temperate latitudes to have hypolimnetic waters that are near 4 ° C, the temperature at which water is most dense, or slightly above 4 °C, reflecting the prevailing water temperature at the time of spring stratification. At lower latitudes, the minimum temperature increases, until it reaches approximately 24 °C within 10° latitude of the equator. Thus, a stratified lake might have a hypolimnion of 4 °C in Wisconsin and 24 °C in Venezuela.

    The hypolimnion, in contrast to the mixed layer, has a very low degree of turbulence, is too dark for photosynthesis, and is isolated from the surface. Also, except at low latitudes, it is much cooler than the mixed layer (Figure 4).

    Because of its physical isolation from photosynthesis and from atmospheric oxygen, the hypolimnion typically loses oxygen during the period of stratification. The loss of oxygen depends on the size of the hypolimnion, its temperature, the duration of stratification, and the amount of organic matter coming down to it from above, which is a byproduct of the trophic status of the lake. In lakes that are just deep enough to support stable stratification (e.g., 10–20 m), the hypolimnion may be absent, as the metalimnion reaches the bottom of the lake. In deeper lakes, the hypolimnion may equal the volume of the epilimnion, and in very deep lakes (e.g., > 100 m), the hypolimnion may be much larger than the epilimnion. Lakes with a very large hypolimnion often maintain hypolimnetic oxygen throughout the stratification season, especially at temperate latitudes where the hypolimnion is cool. Lakes with a very small hypolimnion typically lose most or all of their oxygen, even if they have low productivity, because the sediments of a lake contain enough organic matter to demand most or all the oxygen from a small hypolimnion. Oxygen concentration in lakes with a hypolimnion of intermediate size is quite sensitive to trophic state. Eutrophic lakes typically lose most or all of their hypolimnetic oxygen, thus producing an anoxic benthic zone, whereas lakes of lower productivity may retain an oxic hypolimnion overlying a benthic zone that has an oxidized surface.

    Loss of oxygen is of great importance to the metabolism of a lake because eukaryotes (most protozoa, invertebrates, fish, algae) cannot live in anoxic waters. Thus, loss of hypolimnetic oxygen excludes nonmicrobial components of the biota from the deep waters of a lake.

    Zones with Dynamic Dimensions: Euphotic and Aphotic

    Photosynthesis in the water column of lake is dependent on the availability of photosynthetically available radiation (PAR, wavelengths 350–700 nm). PAR, which corresponds closely to the spectrum of human vision, is removed exponentially as it travels through a water column. The availability of PAR is high during daylight hours at or near the surface of the water column. If nutrients are present, rates of photosynthesis are likely to be high near the surface. A progressive decline in PAR with depth is paraleled by a decline in rates of photosynthesis with depth. At a depth corresponding to ~ 1% of the surface irradiance, net photosynthesis reaches zero, which is a threshold beyond which accumulation of plant biomass is not possible. At depths below the 1% level, photosynthetic organisms (e.g., phytoplankton) lose mass and either die or become dormant unless they are returned to the surface by water currents, which commonly occurs in the mixed layer but not in the metalimnion or hypolimnion.

    The depth between the water surface and the depth of 1% irradiance is referred to as the euphotic zone (Figure 5). Below the depth of 1% irradiance is the aphotic zone. The euphotic zone may not occupy the entire epilimnion of a lake, or may extend to the full thickness of the epilimnion. In some cases, the euphotic zone may extend into the metalimnion, but its extension into the hypolimnion of a lake is unlikely.

    Figure 5 Illustration of the division of a lake into euphotic and aphotic zones. The euphotic zone, which is the lake volume within which positive net photosynthesis can occur, corresponds approximately to the depth at which ≥1% of incoming irradiance is found. Portions of the lake below this boundary have negative net photosynthesis or negligible total photosynthesis.

    The thickness of the euphotic zone is dependent on transparency of the water, which in turn is influenced by dissolved color (colored organic acids from soils), inorganic suspended matter (silt and clay), and living organisms, and especially those that contain chlorophyll, an efficient absorber of light. Because the concentrations of each of these constituents can vary on relatively short time scales (e.g., weekly), the thickness of the euphotic and aphotic zones is dynamic; it is subject to both seasonal and irregular change over time.

    The thickness of the euphotic zone may be small at times of high runoff, when suspended inorganic material and colored organic compounds enter lakes in the largest quantities. Other times when the euphotic zone may be thin it coincides with algal blooms, which can produce sufficient chlorophyll to reduce the transparency of the water column substantially. Conversely, the euphotic zone may thicken substantially when nutrients are exhausted because phytoplankton biomass is likely to decline, thus increasing transparency of the water. Also, strong grazing by zooplankton may thicken the euphotic zone by removing phytoplankton biomass.

    The euphotic zone extends across an entire lake, including both pelagic and littoral zones. In fact, the mean thickness of the euphotic zone determines the outer boundary of the littoral zone because of its effect on the attached vegetation that is characteristic of littoral zones.

    Overview

    The four sets of zones shown in Table 1 define distinctive habitats within lakes that are associated with specific categories of organisms and biogeochemical or metabolic processes. The zones reflect some of the most important physical and chemical factors that control biotically driven processes and biotic community structure. Zonation, although generally a qualitative rather a quantitative concept, reflects accumulation of experience and measurements across lakes of many kinds. Therefore, knowledge of zonal boundaries in a lake allows some general predictions to be made about the kinds of organisms and rates of biogeochemical processes that will occur in a given lake, and the spatial distribution of these organisms and processes.

    Glossary

    Aphotic   Without light, generally interpreted limno-logically as receiving less than 1% of solar irradiance reaching a lake surface.

    Epilimnion   The uppermost and warmest layer (also called the mixed layer) of a lake that experiences density stratification induced by seasonal warming at the lake surface.

    Euphotic   That portion of a lake receiving sufficient solar irradiance to support photosynthesis (typically more than 1% of full solar irradiance).

    Hypolimnion   The most dense, deepest, and coolest layer of a thermally stratified lake. The hypolimnion does not support photosynthesis, because it lacks solar irradiance, and in many cases shows partial or complete depletion of dissolved oxygen.

    Littoral   Near the shore of a waterbody, where irradiance reaching the bottom is above 1% of solar irradiance at the water surface.

    Metalimnion   A layer of transitional density and temperature that connects the epilimnion to the hypolimnion.

    Pelagic   Beyond the littoral zone of a lake.

    Benthic   The zone of a lake extending a few centimeters above and below the bottom of the lake.

    See also: Density Stratification and Stability; Geomorphology of Lake Basins.

    Further Reading

    Hutchinson GE. A Treatise on Limnology, Volume II: Introduction to Lake Biology and the Limnoplankton. New York: Wiley; 1967.

    Hutchinson GE. A Treatise on Limnology, Volume III: Limnological Botany. New York: Wiley; 1975.

    Hutchinson GE. A Treatise on Limnology, Volume IV: The Zoobenthos. New York: Wiley; 1993.

    Kalff J. Limnology: Inland Water Ecosystems. Upper Saddle River, NJ: Prentice Hall; 2002.

    Wetzel RG. Limnology. 3rd edn New York: Academic Press; 2001.

    Littoral Zone

    J.A. Peters; D.M. Lodge    University of Notre Dame, Notre Dame, IN, USA

    Introduction

    The littoral zone of a lake is the nearshore interface between the terrestrial ecosystem and the deeper pelagic zone of the lake. It is the area where at least one percent of the photosynthetically active light (400–700 nm) entering the water reaches the sediment, allowing primary producers (macrophytes and algae) to flourish. The littoral zone is structurally and functionally an important part of most lakes for several reasons. First, most lakes on Earth are small and therefore, the littoral zone comprises a large proportion of total lake area (Figure 1). Second, as an interface, the littoral zone influences the movement and processing of material flowing into the lake from terrestrial runoff, groundwater, or stream connections, thus affecting the physical and biological processes in this zone and the rest of the lake ecosystem. Third, the littoral zone is generally the most productive area of the lake, especially in terms of aquatic plants and invertebrates. Finally, human uses of aquatic systems (swimming, fishing, boating, power generation, irrigation, etc.) often focus on the littoral zone.

    Figure 1 Number of lakes of the world dominated by littoral or pelagic zones. Modified from Wetzel RG (2001) Limnology: Lake and River Ecosystems. New York: Elsevier, Academic Press.

    In the first section, the physical structure and nutrient dynamics within the littoral zone are described. In the second section, interactions among organisms in the littoral zone are discussed, in addition to interactions between the littoral and terrestrial ecosystems, and between the littoral and pelagic zones. In the final section, anthropogenic effects on the littoral zone are described.

    Factors Influencing the Physical Structure and Nutrient Dynamics of the Littoral Zone

    Many characteristics determine the percentage of the lake that is littoral zone and the type of lake-bottom substrates found there. Littoral zone area and substrate type, in turn, influence the processing of incoming nutrients, minerals, and organic matter and therefore the functioning of the entire lake.

    Physical Structure of the Littoral Zone

    Zonation

    In general, the littoral zone can be divided into upper, middle, and lower zones, extending from the shoreline area sprayed by waves to the bottom of the littoral zone, below which light does not penetrate (Figure 2). Emergent vegetation is rooted in

    Enjoying the preview?
    Page 1 of 1