Human impact on the nitrogen cycle

Last updated
Figure 1: The nitrogen cycle in a soil-plant system. One potential pathway: N is fixed by microbes into organic compounds, which are mineralized (i.e., ammonification) and then oxidized to inorganic forms (i.e., nitrification) that are assimilated by plants (NO3 ). NO3 may also be denitrified by bacteria, producing N2, NOx, and N2O. Nitrogen Cycle.svg
Figure 1: The nitrogen cycle in a soil-plant system. One potential pathway: N is fixed by microbes into organic compounds, which are mineralized (i.e., ammonification) and then oxidized to inorganic forms (i.e., nitrification) that are assimilated by plants (NO3 ). NO3 may also be denitrified by bacteria, producing N2, NOx, and N2O.
Estimated nitrogen surplus (the difference between inorganic and organic fertilizer application, atmospheric deposition, fixation and uptake by crops) for the year 2005 across Europe. Estimated nitrogen surplus across Europe 2005.png
Estimated nitrogen surplus (the difference between inorganic and organic fertilizer application, atmospheric deposition, fixation and uptake by crops) for the year 2005 across Europe.

Human impact on the nitrogen cycle is diverse. Agricultural and industrial nitrogen (N) inputs to the environment currently exceed inputs from natural N fixation. [1] As a consequence of anthropogenic inputs, the global nitrogen cycle (Fig. 1) has been significantly altered over the past century. Global atmospheric nitrous oxide (N2O) mole fractions have increased from a pre-industrial value of ~270 nmol/mol to ~319 nmol/mol in 2005. [2] Human activities account for over one-third of N2O emissions, most of which are due to the agricultural sector. [2] This article is intended to give a brief review of the history of anthropogenic N inputs, and reported impacts of nitrogen inputs on selected terrestrial and aquatic ecosystems.

Contents

History of anthropogenic nitrogen inputs

Mean acidifying emissions (air pollution) of different foods per 100g of protein [3]
Food TypesAcidifying Emissions (g SO2eq per 100g protein)
Beef
343.6
Cheese
165.5
Pork
142.7
Lamb and Mutton
139.0
Farmed Crustaceans
133.1
Poultry
102.4
Farmed Fish
65.9
Eggs
53.7
Groundnuts
22.6
Peas
8.5
Tofu
6.7

Approximately 78% of Earth's atmosphere is N gas (N2), which is an inert compound and biologically unavailable to most organisms. In order to be utilized in most biological processes, N2 must be converted to reactive nitrogen (Nr), which includes inorganic reduced forms (NH3 and NH4+), inorganic oxidized forms (NO, NO2, HNO3, N2O, and NO3), and organic compounds (urea, amines, and proteins). [1] N2 has a strong triple bond, and so a significant amount of energy (226 kcal mol−1) is required to convert N2 to Nr. [1] Prior to industrial processes, the only sources of such energy were solar radiation and electrical discharges. [1] Utilizing a large amount of metabolic energy and the enzyme nitrogenase, some bacteria and cyanobacteria convert atmospheric N2 to NH3, a process known as biological nitrogen fixation (BNF). [4] The anthropogenic analogue to BNF is the Haber-Bosch process, in which H2 is reacted with atmospheric N2 at high temperatures and pressures to produce NH3. [5] Lastly, N2 is converted to NO by energy from lightning, which is negligible in current temperate ecosystems, or by fossil fuel combustion. [1]

Until 1850, natural BNF, cultivation-induced BNF (e.g., planting of leguminous crops), and incorporated organic matter were the only sources of N for agricultural production. [5] Near the turn of the century, Nr from guano and sodium nitrate deposits was harvested and exported from the arid Pacific islands and South American deserts. [5] By the late 1920s, early industrial processes, albeit inefficient, were commonly used to produce NH3. [1] Due to the efforts of Fritz Haber and Carl Bosch, the Haber-Bosch process became the largest source of nitrogenous fertilizer after the 1950s, and replaced BNF as the dominant source of NH3 production. [5] From 1890 to 1990, anthropogenically created Nr increased almost ninefold. [1] During this time, the human population more than tripled, partly due to increased food production.

Since the industrial revolution, an additional source of anthropogenic N input has been fossil fuel combustion, which is used to release energy (e.g., to power automobiles). As fossil fuels are burned, high temperatures and pressures provide energy to produce NO from N2 oxidation. [1] Additionally, when fossil fuel is extracted and burned, fossil N may become reactive (i.e., NOx emissions). [1] During the 1970s scientists began to recognize that N inputs were accumulating in the environment and affecting ecosystems. [1]

Impacts of anthropogenic inputs on the nitrogen cycle

Between 1600 and 1990, global reactive nitrogen (Nr) creation had increased nearly 50%. [6] During this period, atmospheric emissions of Nr species reportedly increased 250% and deposition to marine and terrestrial ecosystems increased over 200%. [6] Additionally, there was a reported fourfold increase in riverine dissolved inorganic N fluxes to coasts. [6] Nitrogen is a critical limiting nutrient in many systems, including forests, wetlands, and coastal and marine ecosystems; therefore, this change in emissions and distribution of Nr has resulted in substantial consequences for aquatic and terrestrial ecosystems. [7] [8]

Atmosphere

Mean greenhouse gas emissions for different food types [9]
Food TypesGreenhouse Gas Emissions (g CO2-Ceq per g protein)
Ruminant Meat
62
Recirculating Aquaculture
30
Trawling Fishery
26
Non-recirculating Aquaculture
12
Pork
10
Poultry
10
Dairy
9.1
Non-trawling Fishery
8.6
Eggs
6.8
Starchy Roots
1.7
Wheat
1.2
Maize
1.2
Legumes
0.25

Atmospheric N inputs mainly include oxides of N (NOx), ammonia (NH3), and nitrous oxide (N2O) from aquatic and terrestrial ecosystems, [4] and NOx from fossil fuel and biomass combustion. [1]

In agroecosystems, fertilizer application has increased microbial nitrification (aerobic process in which microorganisms oxidize ammonium [NH4+] to nitrate [NO3]) and denitrification (anaerobic process in which microorganisms reduce NO3 to atmospheric nitrogen gas [N2]). Both processes naturally leak nitric oxide (NO) and nitrous oxide (N2O) to the atmosphere. [4] Of particular concern is N2O, which has an average atmospheric lifetime of 114–120 years, [10] and is 300 times more effective than CO2 as a greenhouse gas. [4] NOx produced by industrial processes, automobiles and agricultural fertilization and NH3 emitted from soils (i.e., as an additional byproduct of nitrification) [4] and livestock operations are transported to downwind ecosystems, influencing N cycling and nutrient losses. Six major effects of NOx and NH3 emissions have been cited: [1] 1) decreased atmospheric visibility due to ammonium aerosols (fine particulate matter [PM]); 2) elevated ozone concentrations; 3) ozone and PM affects human health (e.g. respiratory diseases, cancer); 4) increases in radiative forcing and global climate change; 5) decreased agricultural productivity due to ozone deposition; and 6) ecosystem acidification [11] and eutrophication.

Biosphere

Terrestrial and aquatic ecosystems receive Nr inputs from the atmosphere through wet and dry deposition. [1] Atmospheric Nr species can be deposited to ecosystems in precipitation (e.g., NO3, NH4+, organic N compounds), as gases (e.g., NH3 and gaseous nitric acid [HNO3]), or as aerosols (e.g., ammonium nitrate [NH4NO3]). [1] Aquatic ecosystems receive additional nitrogen from surface runoff and riverine inputs. [8]

Increased N deposition can acidify soils, streams, and lakes and alter forest and grassland productivity. In grassland ecosystems, N inputs have produced initial increases in productivity followed by declines as critical thresholds are exceeded. [1] Nitrogen effects on biodiversity, carbon cycling, and changes in species composition have also been demonstrated. In highly developed areas of near shore coastal ocean and estuarine systems, rivers deliver direct (e.g., surface runoff) and indirect (e.g., groundwater contamination) N inputs from agroecosystems. [8] Increased N inputs can result in freshwater acidification and eutrophication of marine waters.

Terrestrial ecosystems

Impacts on productivity and nutrient cycling

Much of terrestrial growth in temperate systems is limited by N; therefore, N inputs (i.e., through deposition and fertilization) can increase N availability, which temporarily increases N uptake, plant and microbial growth, and N accumulation in plant biomass and soil organic matter. [12] Incorporation of greater amounts of N in organic matter decreases C:N ratios, increasing mineral N release (NH4+) during organic matter decomposition by heterotrophic microbes (i.e., ammonification). [13] As ammonification increases, so does nitrification of the mineralized N. Because microbial nitrification and denitrification are "leaky", N deposition is expected to increase trace gas emissions. [14] Additionally, with increasing NH4+ accumulation in the soil, nitrification processes release hydrogen ions, which acidify the soil. NO3, the product of nitrification, is highly mobile and can be leached from the soil, along with positively charged alkaline minerals such as calcium and magnesium. [4] In acid soils, mobilized aluminium ions can reach toxic concentrations, negatively affecting both terrestrial and adjacent aquatic ecosystems.

Anthropogenic sources of N generally reach upland forests through deposition. [15] A potential concern of increased N deposition due to human activities is altered nutrient cycling in forest ecosystems. Numerous studies have demonstrated both positive and negative impacts of atmospheric N deposition on forest productivity and carbon storage. Added N is often rapidly immobilized by microbes, [16] and the effect of the remaining available N depends on the plant community's capacity for N uptake. [17] In systems with high uptake, N is assimilated into the plant biomass, leading to enhanced net primary productivity (NPP) and possibly increased carbon sequestration through greater photosynthetic capacity. However, ecosystem responses to N additions are contingent upon many site-specific factors including climate, land-use history, and amount of N additions. For example, in the Northeastern United States, hardwood stands receiving chronic N inputs have demonstrated greater capacity to retain N and increase annual net primary productivity (ANPP) than conifer stands. [18] Once N input exceeds system demand, N may be lost via leaching and gas fluxes. When available N exceeds the ecosystem's (i.e., vegetation, soil, and microbes, etc.) uptake capacity, N saturationoccurs and excess N is lost to surface waters, groundwater, and the atmosphere. [12] [17] [18] N saturation can result in nutrient imbalances (e.g., loss of calcium due to nitrate leaching) and possible forest decline. [13]

A 15-year study of chronic N additions at the Harvard Forest Long Term Ecological Research (LTER) program has elucidated many impacts of increased nitrogen deposition on nutrient cycling in temperate forests. It found that chronic N additions resulted in greater leaching losses, increased pine mortality, and cessation of biomass accumulation. [18] Another study reported that chronic N additions resulted in accumulation of non-photosynthetic N and subsequently reduced photosynthetic capacity, supposedly leading to severe carbon stress and mortality. [17] These findings negate previous hypotheses that increased N inputs would increase NPP and carbon sequestration.

Impacts on plant species diversity

Many plant communities have evolved under low nutrient conditions; therefore, increased N inputs can alter biotic and abiotic interactions, leading to changes in community composition. Several nutrient addition studies have shown that increased N inputs lead to dominance of fast-growing plant species, with associated declines in species richness. [19] [20] [21] Fast growing species have a greater affinity for nitrogen uptake, and will crowd out slower growing plant species by blocking access to sunlight with their higher above ground biomass. [22] Other studies have found that secondary responses of the system to N enrichment, including soil acidification and changes in mycorrhizal communities have allowed stress-tolerant species to out-compete sensitive species. [11] [23] Trees that have arbuscular mycorrhizal associations are more likely to benefit from an increase in soil nitrogen, as these fungi are unable to break down soil organic nitrogen. [24] Two other studies found evidence that increased N availability has resulted in declines in species-diverse heathlands. Heathlands are characterized by N-poor soils, which exclude N-demanding grasses; however, with increasing N deposition and soil acidification, invading grasslands replace lowland heath. [25] [26]

In a more recent experimental study of N fertilization and disturbance (i.e., tillage) in old field succession, it was found that species richness decreased with increasing N, regardless of disturbance level. Competition experiments showed that competitive dominants excluded competitively inferior species between disturbance events. With increased N inputs, competition shifted from belowground to aboveground (i.e., to competition for light), and patch colonization rates significantly decreased. These internal changes can dramatically affect the community by shifting the balance of competition-colonization tradeoffs between species. [21] In patch-based systems, regional coexistence can occur through tradeoffs in competitive and colonizing abilities given sufficiently high disturbance rates. [27] That is, with inverse ranking of competitive and colonizing abilities, plants can coexist in space and time as disturbance removes superior competitors from patches, allowing for establishment of superior colonizers. However, as demonstrated by Wilson and Tilman, increased nutrient inputs can negate tradeoffs, resulting in competitive exclusion of these superior colonizers/poor competitors. [21]

Aquatic ecosystems

Aquatic ecosystems also exhibit varied responses to nitrogen enrichment. NO3 loading from N saturated, terrestrial ecosystems can lead to acidification of downstream freshwater systems and eutrophication of downstream marine systems. Freshwater acidification can cause aluminium toxicity and mortality of pH-sensitive fish species. Because marine systems are generally nitrogen-limited, excessive N inputs can result in water quality degradation due to toxic algal blooms, oxygen deficiency, habitat loss, decreases in biodiversity, and fishery losses. [8]

Acidification of freshwaters

Atmospheric N deposition in terrestrial landscapes can be transformed through soil microbial processes to biologically available nitrogen, which can result in surface-water acidification, and loss of biodiversity. NO3 and NH4+ inputs from terrestrial systems and the atmosphere can acidify freshwater systems when there is little buffering capacity due to soil acidification. [8] N pollution in Europe, the Northeastern United States, and Asia is a current concern for freshwater acidification. [28] Lake acidification studies in the Experimental Lake Area (ELA) in northwestern Ontario clearly demonstrated the negative effects of increased acidity on a native fish species: lake trout (Salvelinus namaycush) recruitment and growth dramatically decreased due to extirpation of its key prey species during acidification. [29] Reactive nitrogen from agriculture, animal-raising, fertilizer, septic systems, and other sources have raised nitrate concentrations in waterways of most industrialized nations. Nitrate concentrations in 1,000 Norwegian lakes had doubled in less than a decade. Rivers in the northeastern United States and the majority of Europe have increased ten to fifteen fold over the last century. Reactive nitrogen can contaminate drinking water through runoff into streams, lakes, rivers, and groundwater. In the United States alone, as much as 20% of groundwater sources exceed the World Health Organization's limit of nitrate concentration in potable water. These high concentrations can cause "blue baby disease" where nitrate ions weaken the blood's capacity to carry oxygen. Studies have also linked high concentrations of nitrates to reproductive issues and proclivity for some cancers, such as bladder and ovarian cancer. [30]

Eutrophication of marine systems

Urbanization, deforestation, and agricultural activities largely contribute sediment and nutrient inputs to coastal waters via rivers. [8] Increased nutrient inputs to marine systems have shown both short-term increases in productivity and fishery yields, and long-term detrimental effects of eutrophication. Tripling of NO3 loads in the Mississippi River in the last half of the 20th century have been correlated with increased fishery yields in waters surrounding the Mississippi delta; [31] however, these nutrient inputs have produced seasonal hypoxia (oxygen concentrations less than 2–3 mg L−1, "dead zones") in the Gulf of Mexico. [1] [8] In estuarine and coastal systems, high nutrient inputs increase primary production (e.g., phytoplankton, sea grasses, macroalgae), which increase turbidity with resulting decreases in light penetration throughout the water column. Consequently, submerged vegetation growth declines, which reduces habitat complexity and oxygen production. The increased primary (i.e., phytoplankton, macroalgae, etc.) production leads to a flux of carbon to bottom waters when decaying organic matter (i.e., senescent primary production) sinks and is consumed by aerobic bacteria lower in the water column. As a result, oxygen consumption in bottom waters is greater than diffusion of oxygen from surface waters. Additionally, certain algal blooms termed harmful algal blooms (HABs) produce toxins that can act as neuromuscular or organ damaging compounds. These algal blooms can be harmful to other marine life as well as to humans. [32] [33]

Integration

The above system responses to reactive nitrogen (Nr) inputs are almost all exclusively studied separately; however, research increasingly indicates that nitrogen loading problems are linked by multiple pathways transporting nutrients across system boundaries. [1] This sequential transfer between ecosystems is termed the nitrogen cascade. [6] (see illustration from United Nations Environment Programme). During the cascade, some systems accumulate Nr, which results in a time lag in the cascade and enhanced effects of Nr on the environment in which it accumulates. Ultimately, anthropogenic inputs of Nr are either accumulated or denitrified; however, little progress has been made in determining the relative importance of Nr accumulation and denitrification, which has been mainly due to a lack of integration among scientific disciplines. [1] [34]

Most Nr applied to global agroecosystems cascades through the atmosphere and aquatic and terrestrial ecosystems until it is converted to N2, primarily through denitrification. [1] Although terrestrial denitrification produces gaseous intermediates (nitric oxide [NO] and nitrous oxide [N2O]), the last step—microbial production of N2— is critical because atmospheric N2 is a sink for Nr. [34] Many studies have clearly demonstrated that managed buffer strips and wetlands can remove significant amounts of nitrate (NO3) from agricultural systems through denitrification. [35] Such management may help attenuate the undesirable cascading effects and eliminate environmental Nr accumulation. [1]

Human activities dominate the global and most regional N cycles. [36] N inputs have shown negative consequences for both nutrient cycling and native species diversity in terrestrial and aquatic systems. In fact, due to long-term impacts on food webs, Nr inputs are widely considered the most critical pollution problem in marine systems. [8] In both terrestrial and aquatic ecosystems, responses to N enrichment vary; however, a general re-occurring theme is the importance of thresholds (e.g., nitrogen saturation) in system nutrient retention capacity. In order to control the N cascade, there must be integration of scientific disciplines and further work on Nr storage and denitrification rates. [34]

See also

Related Research Articles

<span class="mw-page-title-main">Fertilizer</span> Substance added to soils to supply plant nutrients for a better growth

A fertilizer or fertiliser is any material of natural or synthetic origin that is applied to soil or to plant tissues to supply plant nutrients. Fertilizers may be distinct from liming materials or other non-nutrient soil amendments. Many sources of fertilizer exist, both natural and industrially produced. For most modern agricultural practices, fertilization focuses on three main macro nutrients: nitrogen (N), phosphorus (P), and potassium (K) with occasional addition of supplements like rock flour for micronutrients. Farmers apply these fertilizers in a variety of ways: through dry or pelletized or liquid application processes, using large agricultural equipment or hand-tool methods.

<span class="mw-page-title-main">Nitrogen cycle</span> Biogeochemical cycle by which nitrogen is converted into various chemical forms

The nitrogen cycle is the biogeochemical cycle by which nitrogen is converted into multiple chemical forms as it circulates among atmospheric, terrestrial, and marine ecosystems. The conversion of nitrogen can be carried out through both biological and physical processes. Important processes in the nitrogen cycle include fixation, ammonification, nitrification, and denitrification. The majority of Earth's atmosphere (78%) is atmospheric nitrogen, making it the largest source of nitrogen. However, atmospheric nitrogen has limited availability for biological use, leading to a scarcity of usable nitrogen in many types of ecosystems.

<span class="mw-page-title-main">Fen</span> Type of wetland fed by mineral-rich ground or surface water

A fen is a type of peat-accumulating wetland fed by mineral-rich ground or surface water. It is one of the main types of wetlands along with marshes, swamps, and bogs. Bogs and fens, both peat-forming ecosystems, are also known as mires. The unique water chemistry of fens is a result of the ground or surface water input. Typically, this input results in higher mineral concentrations and a more basic pH than found in bogs. As peat accumulates in a fen, groundwater input can be reduced or cut off, making the fen ombrotrophic rather than minerotrophic. In this way, fens can become more acidic and transition to bogs over time.

<span class="mw-page-title-main">Denitrification</span> Microbially facilitated process

Denitrification is a microbially facilitated process where nitrate (NO3) is reduced and ultimately produces molecular nitrogen (N2) through a series of intermediate gaseous nitrogen oxide products. Facultative anaerobic bacteria perform denitrification as a type of respiration that reduces oxidized forms of nitrogen in response to the oxidation of an electron donor such as organic matter. The preferred nitrogen electron acceptors in order of most to least thermodynamically favorable include nitrate (NO3), nitrite (NO2), nitric oxide (NO), nitrous oxide (N2O) finally resulting in the production of dinitrogen (N2) completing the nitrogen cycle. Denitrifying microbes require a very low oxygen concentration of less than 10%, as well as organic C for energy. Since denitrification can remove NO3, reducing its leaching to groundwater, it can be strategically used to treat sewage or animal residues of high nitrogen content. Denitrification can leak N2O, which is an ozone-depleting substance and a greenhouse gas that can have a considerable influence on global warming.

In atmospheric chemistry, NOx is shorthand for nitric oxide and nitrogen dioxide, the nitrogen oxides that are most relevant for air pollution. These gases contribute to the formation of smog and acid rain, as well as affecting tropospheric ozone.

Denitrifying bacteria are a diverse group of bacteria that encompass many different phyla. This group of bacteria, together with denitrifying fungi and archaea, is capable of performing denitrification as part of the nitrogen cycle. Denitrification is performed by a variety of denitrifying bacteria that are widely distributed in soils and sediments and that use oxidized nitrogen compounds such as nitrate and nitrite in the absence of oxygen as a terminal electron acceptor. They metabolize nitrogenous compounds using various enzymes, including nitrate reductase (NAR), nitrite reductase (NIR), nitric oxide reductase (NOR) and nitrous oxide reductase (NOS), turning nitrogen oxides back to nitrogen gas or nitrous oxide.

<span class="mw-page-title-main">Paleolimnology</span> Scientific study of ancient lakes and streams

Paleolimnology is a scientific sub-discipline closely related to both limnology and paleoecology. Paleolimnological studies focus on reconstructing the past environments of inland waters using the geologic record, especially with regard to events such as climatic change, eutrophication, acidification, and internal ontogenic processes.

In the study of air pollution, a critical load is defined as "a quantitative estimate of an exposure to one or more pollutants below which significant harmful effects on specified sensitive elements of the environment do not occur according to present knowledge".

<span class="mw-page-title-main">Leaching (agriculture)</span> Loss of water-soluble plant nutrients from soil due to rain and irrigation

In agriculture, leaching is the loss of water-soluble plant nutrients from the soil, due to rain and irrigation. Soil structure, crop planting, type and application rates of fertilizers, and other factors are taken into account to avoid excessive nutrient loss. Leaching may also refer to the practice of applying a small amount of excess irrigation where the water has a high salt content to avoid salts from building up in the soil. Where this is practiced, drainage must also usually be employed, to carry away the excess water.

<span class="mw-page-title-main">Agricultural pollution</span> Type of pollution caused by agriculture

Agricultural pollution refers to biotic and abiotic byproducts of farming practices that result in contamination or degradation of the environment and surrounding ecosystems, and/or cause injury to humans and their economic interests. The pollution may come from a variety of sources, ranging from point source water pollution to more diffuse, landscape-level causes, also known as non-point source pollution and air pollution. Once in the environment these pollutants can have both direct effects in surrounding ecosystems, i.e. killing local wildlife or contaminating drinking water, and downstream effects such as dead zones caused by agricultural runoff is concentrated in large water bodies.

Aerobic denitrification or co-respiration the simultaneous use of both oxygen (O2) and nitrate (NO3) as oxidizing agents, performed by various genera of microorganisms. This process differs from anaerobic denitrification not only in its insensitivity to the presence of oxygen, but also in that it has a higher potential to create the harmful byproduct nitrous oxide.

Daycent is a daily time series biogeochemical model used in agroecosystems to simulate fluxes of carbon and nitrogen between the atmosphere, vegetation, and soil. It is a daily version of the CENTURY biogeochemical model.

<span class="mw-page-title-main">Freshwater acidification</span>

Freshwater acidification occurs when acidic inputs enter a body of fresh water through the weathering of rocks, invasion of acidifying gas, or by the reduction of acid anions, like sulfate and nitrate within a lake. Freshwater acidification is primarily caused by sulfur oxides (SOx) and nitrogen oxides (NOx) entering the water from atmospheric depositions and soil leaching. Carbonic acid and dissolved carbon dioxide can also enter freshwaters, in a similar manner associated with runoff, through carbon dioxide-rich soils. Runoff that contains these compounds may incorporate acidifying hydrogen ions and inorganic aluminum, which can be toxic to marine organisms. Acid rain is also a contributor to freshwater acidification. It is created when SOx and NOx react with water, oxygen, and other oxidants within the clouds.

<span class="mw-page-title-main">Cattle urine patches</span> Grass damage by cattle urine

Urine patches in cattle pastures generate large concentrations of the greenhouse gas nitrous oxide through nitrification and denitrification processes in urine-contaminated soils. Over the past few decades, the cattle population has increased more rapidly than the human population. Between the years 2000 and 2050, the cattle population is expected to increase from 1.5 billion to 2.6 billion. When large populations of cattle are packed into pastures, excessive amounts of urine soak into soils. This increases the rate at which nitrification and denitrification occur and produce nitrous oxide. Currently, nitrous oxide is one of the single most important ozone-depleting emissions and is expected to remain the largest throughout the 21st century.

Dissimilatory nitrate reduction to ammonium (DNRA), also known as nitrate/nitrite ammonification, is the result of anaerobic respiration by chemoorganoheterotrophic microbes using nitrate (NO3) as an electron acceptor for respiration. In anaerobic conditions microbes which undertake DNRA oxidise organic matter and use nitrate (rather than oxygen) as an electron acceptor, reducing it to nitrite, then ammonium (NO3→NO2→NH4+).

Some types of lichen are able to fix nitrogen from the atmosphere. This process relies on the presence of cyanobacteria as a partner species within the lichen. The ability to fix nitrogen enables lichen to live in nutrient-poor environments. Lichen can also extract nitrogen from the rocks on which they grow.

An oxygen minimum zone (OMZ) is characterized as an oxygen-deficient layer in the world's oceans. Typically found between 200m to 1500m deep below regions of high productivity, such as the western coasts of continents. OMZs can be seasonal following the spring-summer upwelling season. Upwelling of nutrient-rich water leads to high productivity and labile organic matter, that is respired by heterotrophs as it sinks down the water column. High respiration rates deplete the oxygen in the water column to concentrations of 2 mg/L or less forming the OMZ. OMZs are expanding, with increasing ocean deoxygenation. Under these oxygen-starved conditions, energy is diverted from higher trophic levels to microbial communities that have evolved to use other biogeochemical species instead of oxygen, these species include Nitrate, Nitrite, Sulphate etc. Several Bacteria and Archea have adapted to live in these environments by using these alternate chemical species and thrive. The most abundant phyla in OMZs are Pseudomonadota, Bacteroidota, Actinomycetota, and Planctomycetota.

<span class="mw-page-title-main">Viral shunt</span>

The viral shunt is a mechanism that prevents marine microbial particulate organic matter (POM) from migrating up trophic levels by recycling them into dissolved organic matter (DOM), which can be readily taken up by microorganisms. The DOM recycled by the viral shunt pathway is comparable to the amount generated by the other main sources of marine DOM.

<span class="mw-page-title-main">Ammonia pollution</span> Chemical contamination

Ammonia pollution is pollution by the chemical ammonia (NH3) – a compound of nitrogen and hydrogen which is a byproduct of agriculture and industry. Common forms include air pollution by the ammonia gas emitted by rotting agricultural slurry and fertilizer factories while natural sources include the burning coal mines of Jharia, the caustic Lake Natron and the guano of seabird colonies. Gaseous ammonia reacts with other pollutants in the air to form fine particles of ammonium salts, which affect human breathing. Ammonia gas can also affect the chemistry of the soil on which it settles and will, for example, degrade the conditions required by the sphagnum moss and heathers of peatland.

Seventeen elements or nutrients are essential for plant growth and reproduction. They are carbon (C), hydrogen (H), oxygen (O), nitrogen (N), phosphorus (P), potassium (K), sulfur (S), calcium (Ca), magnesium (Mg), iron (Fe), boron (B), manganese (Mn), copper (Cu), zinc (Zn), molybdenum (Mo), nickel (Ni) and chlorine (Cl). Nutrients required for plants to complete their life cycle are considered essential nutrients. Nutrients that enhance the growth of plants but are not necessary to complete the plant's life cycle are considered non-essential, although some of them, such as silicon (Si), have been shown to improve nutrent availability, hence the use of stinging nettle and horsetail macerations in Biodynamic agriculture. With the exception of carbon, hydrogen and oxygen, which are supplied by carbon dioxide and water, and nitrogen, provided through nitrogen fixation, the nutrients derive originally from the mineral component of the soil. The Law of the Minimum expresses that when the available form of a nutrient is not in enough proportion in the soil solution, then other nutrients cannot be taken up at an optimum rate by a plant. A particular nutrient ratio of the soil solution is thus mandatory for optimizing plant growth, a value which might differ from nutrient ratios calculated from plant composition.

References

PD-icon.svg This article incorporates public domain material from Environmental Health Perspectives. National Institutes of Health.
  1. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 Galloway, J. N.; Aber, J. D.; Erisman, J. N. W.; Seitzinger, S. P.; Howarth, R. W.; Cowling, E. B.; Cosby, B. J. (2003). "The Nitrogen Cascade". BioScience. 53 (4): 341. doi: 10.1641/0006-3568(2003)053[0341:TNC]2.0.CO;2 . S2CID   3356400.
  2. 1 2 Alley et al. 2007. IPCC Climate Change 2007: The Physical Science Basis. Contribution of Working Group I in the Third Assessment Report of Intergovernmental Panel on Climate Change. Report Summary for Policy Makers (SPM) Archived 2011-07-16 at the Wayback Machine .
  3. Nemecek, T.; Poore, J. (2018-06-01). "Reducing food's environmental impacts through producers and consumers". Science. 360 (6392): 987–992. Bibcode:2018Sci...360..987P. doi: 10.1126/science.aaq0216 . ISSN   0036-8075. PMID   29853680. S2CID   206664954.
  4. 1 2 3 4 5 6 Schlesinger, W. H. 1997. Biogeochemistry: An analysis of global change, San Diego, CA.
  5. 1 2 3 4 Smil, V. 2001. Enriching the earth: Fritz Haber, Carl Bosch, and the transformation of world food production. MIT Press, Cambridge, MA.
  6. 1 2 3 4 Galloway, James N.; Cowling, Ellis B. (2002). "Reactive Nitrogen and the World: 200 Years of Change". Ambio: A Journal of the Human Environment. 31 (2): 64–71. Bibcode:2002Ambio..31...64G. doi:10.1579/0044-7447-31.2.64. PMID   12078011. S2CID   8104525.
  7. Vitousek, P.; Howarth, R. (1991). "Nitrogen limitation on land and in the sea: How can it occur?". Biogeochemistry. 13 (2). doi:10.1007/BF00002772. S2CID   93106377.
  8. 1 2 3 4 5 6 7 8 Rabalais, Nancy N. (2002). "Nitrogen in Aquatic Ecosystems". Ambio: A Journal of the Human Environment. 31 (2): 102–12. doi:10.1639/0044-7447(2002)031[0102:NIAE]2.0.CO;2. PMID   12077998.
  9. Michael Clark; Tilman, David (November 2014). "Global diets link environmental sustainability and human health". Nature. 515 (7528): 518–522. Bibcode:2014Natur.515..518T. doi:10.1038/nature13959. ISSN   1476-4687. PMID   25383533. S2CID   4453972.
  10. John T. Houghton, Y. Ding, D. J. Griggs, M. Noguer, P. J. van der Linden, X. Dai, K. Maskell, and C. A. Johnson. 2001. IPCC Climate Change 2001: The Scientific Basis. Contribution of Working Group I in the Third Assessment Report of Intergovernmental Panel on Climate Change. Cambridge University Press]
  11. 1 2 Houdijk, A. L. F. M.; Verbeek, P. J. M.; Dijk, H. F. G.; Roelofs, J. G. M. (1993). "Distribution and decline of endangered herbaceous heathland species in relation to the chemical composition of the soil". Plant and Soil. 148 (1): 137–143. Bibcode:1993PlSoi.148..137H. doi:10.1007/BF02185393. S2CID   22600629.
  12. 1 2 Aber, J. D., K. J. Nadelhoffer, P. Steudler, and J. M. Melillo. 1989. "Nitrogen saturation in northern forest ecosystems". Bioscience 39:378–386
  13. 1 2 Aber, J. D. (1992). "Nitrogen cycling and nitrogen saturation in temperate forest ecosystems". Trends in Ecology & Evolution. 7 (7): 220–224. doi:10.1016/0169-5347(92)90048-G. PMID   21236013.
  14. Matson, P; Lohse, KA; Hall, SJ (2002). "The globalization of nitrogen deposition: Consequences for terrestrial ecosystems". Ambio. 31 (2): 113–9. doi:10.1639/0044-7447(2002)031[0113:tgondc]2.0.co;2. JSTOR   4315223. PMID   12077999.
  15. Aber, John D.; Goodale, Christine L.; Ollinger, Scott V.; Smith, Marie-Louise; Magill, Alison H.; Martin, Mary E.; Hallett, Richard A.; Stoddard, John L. (2003). "Is Nitrogen Deposition Altering the Nitrogen Status of Northeastern Forests?". BioScience. 53 (4): 375. doi: 10.1641/0006-3568(2003)053[0375:INDATN]2.0.CO;2 . JSTOR   1314369. S2CID   54037020.
  16. Nadelhoffer, K. J.; Downs, M. R.; Fry, B. (1999). "Sinks For15N-Enriched Additions to an Oak Forest and a Red Pine Plantation". Ecological Applications. 9: 72–86. doi:10.1890/1051-0761(1999)009[0072:SFNEAT]2.0.CO;2.
  17. 1 2 3 Bauer, G. A.; Bazzaz, F. A.; Minocha, R.; Long, S.; Magill, A.; Aber, J.; Berntson, G. M. (2004). "Effects of chronic N additions on tissue chemistry, photosynthetic capacity, and carbon sequestration potential of a red pine (Pinus resinosa Ait.) stand in the NE United States". Forest Ecology and Management. 196: 173–186. doi:10.1016/j.foreco.2004.03.032.
  18. 1 2 3 Magill, A. H.; Aber, J. D.; Currie, W. S.; Nadelhoffer, K. J.; Martin, M. E.; McDowell, W. H.; Melillo, J. M.; Steudler, P. (2004). "Ecosystem response to 15 years of chronic nitrogen additions at the Harvard Forest LTER, Massachusetts, USA". Forest Ecology and Management. 196: 7–28. doi:10.1016/j.foreco.2004.03.033.
  19. Laura Foster Huenneke; Steven P. Hamburg; Roger Koide; Harold A. Mooney; Peter M. Vitousek (1990). "Effects of Soil Resources on Plant Invasion and Community Structure in Californian Serpentine Grassland". Ecology. 71 (2): 478–491. Bibcode:1990Ecol...71..478H. doi:10.2307/1940302. JSTOR   1940302.
  20. Tilman, D. (1997). "Community Invasibility, Recruitment Limitation, and Grassland Biodiversity". Ecology. 78: 81–83. doi:10.1890/0012-9658(1997)078[0081:CIRLAG]2.0.CO;2.
  21. 1 2 3 Wilson, S. D.; Tilman, D. (2002). "Quadratic Variation in Old-Field Species Richness Along Gradients of Disturbance and Nitrogen". Ecology. 83 (2): 492. doi:10.1890/0012-9658(2002)083[0492:QVIOFS]2.0.CO;2.
  22. Wamelink, G.W.W.; van Dobben, H.F.; Mol-Dijkstra, J.P.; Schouwenberg, E.P.A.G.; Kros, J.; de Vries, W.; Berendse, F. (September 2009). "Effect of nitrogen deposition reduction on biodiversity and carbon sequestration". Forest Ecology and Management. 258 (8): 1774–1779. doi:10.1016/j.foreco.2008.10.024.
  23. Egerton-Warburton, L. M.; Allen, E. B. (2000). "Shifts in Arbuscular Mycorrhizal Communities Along an Anthropogenic Nitrogen Deposition Gradient". Ecological Applications. 10 (2): 484. doi:10.1890/1051-0761(2000)010[0484:SIAMCA]2.0.CO;2.
  24. Quinn Thomas, R.; Canham, Charles D.; Weathers, Kathleen C.; Goodale, Christine L. (22 December 2009). "Increased tree carbon storage in response to nitrogen deposition in the US". Nature Geoscience. 3 (1): 13–17. doi:10.1038/ngeo721. ISSN   1752-0908.
  25. Aerts, Rien; Berendse, Frank (August 1988). "The effect of increased nutrient availability on vegetation dynamics in wet heathlands". Vegetatio. 76 (1–2): 63–69. doi:10.1007/BF00047389. ISSN   0042-3106. S2CID   34882407.
  26. Bobbink, R.; Heil, G. W.; Raessen, M. B. (1992). "Atmospheric deposition and canopy exchange processes in heathland ecosystems". Environmental Pollution. 75 (1): 29–37. doi:10.1016/0269-7491(92)90053-D. PMID   15092046.
  27. Hastings, A. (1980). "Disturbance, coexistence, history, and competition for space". Theoretical Population Biology. 18 (3): 363–373. doi:10.1016/0040-5809(80)90059-3.
  28. Driscoll, C. T., G. B. Lawrence, A. J. Bulger, T. J. Butler, C. S. Cronan, C. Eagar, K. F. Lambert, G. E. Likens, J. L. Stoddard, and K. C. Weathers. 2001. Acidic Deposition in the Northeastern United States: Sources and Inputs, Ecosystem Effects, and Management Strategies. pp. 180-198
  29. Mills, K. H.; Chalanchuk, S. M.; Allan, D. J. (2000). "Recovery of fish populations in Lake 223 from experimental acidification". Canadian Journal of Fisheries and Aquatic Sciences. 57: 192–204. doi:10.1139/f99-186.
  30. Fields, Scott (July 2004). "Global Nitrogen: Cycling out of Control". Environmental Health Perspectives. 112 (10): A556–A563. doi:10.1289/ehp.112-a556. PMC   1247398 . PMID   15238298.
  31. Grimes, Churchill B. (2001). "Fishery Production and the Mississippi River Discharge". Fisheries. 26 (8): 17–26. doi:10.1577/1548-8446(2001)026<0017:FPATMR>2.0.CO;2.
  32. Skulberg, Olav M.; Codd, Geoffrey A.; Carmichael, Wayne W. (1984-01-01). "Toxic Blue-Green Algal Blooms in Europe: A Growing Problem". Ambio. 13 (4): 244–247. JSTOR   4313034.
  33. Smith, V.H.; Tilman, G.D.; Nekola, J.C. (1999). "Eutrophication: impacts of excess nutrient inputs on freshwater, marine, and terrestrial ecosystems". Environmental Pollution. 100 (1–3): 179–196. doi:10.1016/s0269-7491(99)00091-3. PMID   15093117. S2CID   969039.
  34. 1 2 3 Davidson, E. A.; Seitzinger, S. (2006). "The Enigma of Progress in Denitrification Research". Ecological Applications. 16 (6): 2057–2063. doi:10.1890/1051-0761(2006)016[2057:TEOPID]2.0.CO;2. PMID   17205889.
  35. Jackson, R. D.; Allen-Diaz, B.; Oates, L. G.; Tate, K. W. (2006). "Spring-water Nitrate Increased with Removal of Livestock Grazing in a California Oak Savanna". Ecosystems. 9 (2): 254. Bibcode:2006Ecosy...9..254J. doi:10.1007/s10021-005-0166-7. S2CID   24450808.
  36. Galloway, J. N.; Dentener, F. J.; Capone, D. G.; Boyer, E. W.; Howarth, R. W.; Seitzinger, S. P.; Asner, G. P.; Cleveland, C. C.; Green, P. A.; Holland, E. A.; Karl, D. M.; Michaels, A. F.; Porter, J. H.; Townsend, A. R.; Vöosmarty, C. J. (2004). "Nitrogen Cycles: Past, Present, and Future" (PDF). Biogeochemistry. 70 (2): 153. Bibcode:2004Biogc..70..153G. doi:10.1007/s10533-004-0370-0. JSTOR   4151466. S2CID   98109580. Archived from the original (PDF) on 2005-11-11.

Further reading