You are on page 1of 17

Nitrogen cycle

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,[16]
making it the largest source of
Global cycling of reactive nitrogen [1] including industrial fertilizer
nitrogen. However, atmospheric
production,[2] nitrogen fixed by natural ecosystems,[3] nitrogen fixed by
nitrogen has limited availability for
oceans,[4] nitrogen fixed by agricultural crops,[5] NOx emitted by
biological use, leading to a scarcity
biomass burning,[6] NOx emitted from soil,[7] nitrogen fixed by
of usable nitrogen in many types of
lightning,[8] NH3 emitted by terrestrial ecosystems,[9] deposition of
ecosystems.
nitrogen to terrestrial surfaces and oceans,[10][11] NH3 emitted from
The nitrogen cycle is of particular oceans,[12][13][11] ocean NO2 emissions from the atmosphere,[14]
interest to ecologists because denitrification in oceans,[4][15][11] and reactive nitrogen burial in
nitrogen availability can affect the oceans.[5]
rate of key ecosystem processes,
including primary production and
decomposition. Human activities such as fossil fuel combustion, use of artificial nitrogen fertilizers, and
release of nitrogen in wastewater have dramatically altered the global nitrogen cycle.[17][18][19] Human
modification of the global nitrogen cycle can negatively affect the natural environment system and also
human health.[20][21]

Processes
Nitrogen is present in the environment in a wide variety of chemical forms including organic nitrogen,
+ − −
ammonium (NH4 ), nitrite (NO2 ), nitrate (NO3 ), nitrous oxide (N2 O), nitric oxide (NO) or inorganic
nitrogen gas (N2 ). Organic nitrogen may be in the form of a living organism, humus or in the intermediate
products of organic matter decomposition. The processes in the nitrogen cycle is to transform nitrogen from
one form to another. Many of those processes are carried out by microbes, either in their effort to harvest
energy or to accumulate nitrogen in a form needed for their growth. For example, the nitrogenous wastes in
animal urine are broken down by nitrifying bacteria in the soil to be used by plants. The diagram alongside
shows how these processes fit together to form the nitrogen cycle.

Nitrogen fixation

The conversion of nitrogen gas (N2 ) into nitrates and nitrites through atmospheric, industrial and biological
processes is called nitrogen fixation. Atmospheric nitrogen must be processed, or "fixed", into a usable
form to be taken up by plants. Between 5 and 10 billion kg per year are fixed by lightning strikes, but most
fixation is done by free-living or symbiotic bacteria known as diazotrophs. These bacteria have the
nitrogenase enzyme that combines gaseous nitrogen with hydrogen to produce ammonia, which is
converted by the bacteria into other organic compounds. Most biological nitrogen fixation occurs by the
activity of molybdenum (Mo)-nitrogenase, found in a wide variety of bacteria and some Archaea. Mo-
nitrogenase is a complex two-component enzyme that has multiple metal-containing prosthetic groups.[22]
An example of free-living bacteria is Azotobacter. Symbiotic nitrogen-fixing bacteria such as Rhizobium
usually live in the root nodules of legumes (such as peas, alfalfa, and locust trees). Here they form a
mutualistic relationship with the plant, producing ammonia in exchange for carbohydrates. Because of this
relationship, legumes will often increase the nitrogen content of nitrogen-poor soils. A few non-legumes can
also form such symbioses. Today, about 30% of the total fixed nitrogen is produced industrially using the
Haber-Bosch process,[23] which uses high temperatures and pressures to convert nitrogen gas and a
hydrogen source (natural gas or petroleum) into ammonia.[24]

Assimilation

Plants can absorb nitrate or ammonium from the soil by their root hairs. If nitrate is absorbed, it is first
reduced to nitrite ions and then ammonium ions for incorporation into amino acids, nucleic acids, and
chlorophyll. In plants that have a symbiotic relationship with rhizobia, some nitrogen is assimilated in the
form of ammonium ions directly from the nodules. It is now known that there is a more complex cycling of
amino acids between Rhizobia bacteroids and plants. The plant provides amino acids to the bacteroids so
ammonia assimilation is not required and the bacteroids pass amino acids (with the newly fixed nitrogen)
back to the plant, thus forming an interdependent relationship.[25] While many animals, fungi, and other
heterotrophic organisms obtain nitrogen by ingestion of amino acids, nucleotides, and other small organic
molecules, other heterotrophs (including many bacteria) are able to utilize inorganic compounds, such as
ammonium as sole N sources. Utilization of various N sources is carefully regulated in all organisms.

Ammonification

When a plant or animal dies or an animal expels waste, the initial form of nitrogen is organic. Bacteria or
+
fungi convert the organic nitrogen within the remains back into ammonium (NH4 ), a process called
ammonification or mineralization. Enzymes involved are:

GS: Gln Synthetase (Cytosolic & Plastic)


GOGAT: Glu 2-oxoglutarate aminotransferase (Ferredoxin & NADH-dependent)
GDH: Glu Dehydrogenase:
Minor Role in ammonium assimilation.
Important in amino acid catabolism.

Nitrification

The conversion of ammonium to nitrate is


performed primarily by soil-living bacteria and
other nitrifying bacteria. In the primary stage of
+
nitrification, the oxidation of ammonium (NH4 ) is
performed by bacteria such as the Nitrosomonas
species, which converts ammonia to nitrites

(NO2 ). Other bacterial species such as
Nitrobacter, are responsible for the oxidation of
− −
the nitrites (NO2 ) into nitrates (NO3 ). It is
important for the ammonia (NH3 ) to be converted
to nitrates or nitrites because ammonia gas is toxic
to plants.

Due to their very high solubility and because soils


are highly unable to retain anions, nitrates can
enter groundwater. Elevated nitrate in Microbial nitrogen cycle [26][27]
groundwater is a concern for drinking water use ANAMMOX is anaerobic ammonium oxidation, DNRA is
because nitrate can interfere with blood-oxygen dissimilatory nitrate reduction to ammonium, and
levels in infants and cause methemoglobinemia or COMMAMOX is complete ammonium oxidation.
blue-baby syndrome. [28] Where groundwater
recharges stream flow, nitrate-enriched
groundwater can contribute to eutrophication, a process that leads to high algal population and growth,
especially blue-green algal populations. While not directly toxic to fish life, like ammonia, nitrate can have
indirect effects on fish if it contributes to this eutrophication. Nitrogen has contributed to severe
eutrophication problems in some water bodies. Since 2006, the application of nitrogen fertilizer has been
increasingly controlled in Britain and the United States. This is occurring along the same lines as control of
phosphorus fertilizer, restriction of which is normally considered essential to the recovery of eutrophied
waterbodies.

Denitrification

Denitrification is the reduction of nitrates back into nitrogen gas (N2 ), completing the nitrogen cycle. This
process is performed by bacterial species such as Pseudomonas and Paracoccus, under anaerobic
conditions. They use the nitrate as an electron acceptor in the place of oxygen during respiration. These
facultatively (meaning optionally) anaerobic bacteria can also live in aerobic conditions. Denitrification
happens in anaerobic conditions e.g. waterlogged soils. The denitrifying bacteria use nitrates in the soil to
carry out respiration and consequently produce nitrogen gas, which is inert and unavailable to plants.
Denitrification occurs in free-living microorganisms as well as obligate symbionts of anaerobic ciliates.[29]
Classical representation of nitrogen Flow of nitrogen through the ecosystem. Bacteria
cycle are a key element in the cycle, providing different
forms of nitrogen compounds able to be
assimilated by higher organisms

Dissimilatory nitrate reduction to


ammonium

Dissimilatory nitrate reduction to ammonium (DNRA), or


nitrate/nitrite ammonification, is an anaerobic respiration
process. Microbes which undertake DNRA oxidise organic
matter and use nitrate as an electron acceptor, reducing it to
nitrite, then ammonium (NO−3 → NO−2 → NH+4).[30] Both
denitrifying and nitrate ammonification bacteria will be
competing for nitrate in the environment, although DNRA
acts to conserve bioavailable nitrogen as soluble ammonium
rather than producing dinitrogen gas.[31]

Anaerobic ammonia oxidation

In this biological process, nitrite and ammonia are converted


directly into molecular nitrogen (N2 ) gas. This process
makes up a major proportion of nitrogen conversion in the
oceans. The balanced formula for this "anammox" chemical
+ −
reaction is: NH4 + NO2 → N2 + 2 H2 O (ΔG° =
−357 kJ⋅mol−1 ).[32] Simple representation of the nitrogen
cycle. Blue represent nitrogen storage,
green is for processes moving nitrogen
Other processes from one place to another, and red is
for the bacteria involved
Though nitrogen fixation is the primary source of plant-
available nitrogen in most ecosystems, in areas with
nitrogen-rich bedrock, the breakdown of this rock also serves as a nitrogen source.[33][34][35] Nitrate

reduction is also part of the iron cycle, under anoxic conditions Fe(II) can donate an electron to NO3 and is
− − +
oxidized to Fe(III) while NO3 is reduced to NO2 , N2 O, N2 , and NH4 depending on the conditions and
microbial species involved.[36] The fecal plumes of cetaceans also act as a junction in the marine nitrogen
cycle, concentrating nitrogen in the epipelagic zones of ocean environments before its dispersion through
various marine layers, ultimately enhancing oceanic primary productivity.[37]

Marine nitrogen cycle


The nitrogen cycle is an important
process in the ocean as well. While the
overall cycle is similar, there are
different players[40] and modes of
transfer for nitrogen in the ocean.
Nitrogen enters the water through the
precipitation, runoff, or as N2 from the
atmosphere. Nitrogen cannot be
utilized by phytoplankton as N2 so it
must undergo nitrogen fixation which
is performed predominately by
Marine nitrogen cycle
cyanobacteria.[41] Without supplies of
fixed nitrogen entering the marine
cycle, the fixed nitrogen would be used up in about 2000 years.[42] Phytoplankton need nitrogen in
biologically available forms for the initial synthesis of organic matter. Ammonia and urea are released into
the water by excretion from plankton. Nitrogen sources are removed from the euphotic zone by the
downward movement of the organic matter. This can occur from sinking of phytoplankton, vertical mixing,
or sinking of waste of vertical migrators. The sinking results in ammonia being introduced at lower depths
below the euphotic zone. Bacteria are able to convert ammonia to nitrite and nitrate but they are inhibited
by light so this must occur below the euphotic zone.[43] Ammonification or Mineralization is performed by
bacteria to convert organic nitrogen to ammonia. Nitrification can then occur to convert the ammonium to
nitrite and nitrate.[44] Nitrate can be returned to the euphotic zone by vertical mixing and upwelling where
it can be taken up by phytoplankton to continue the cycle. N2 can be returned to the atmosphere through
denitrification.

Ammonium is thought to be the preferred source of fixed nitrogen for phytoplankton because its
assimilation does not involve a redox reaction and therefore requires little energy. Nitrate requires a redox
reaction for assimilation but is more abundant so most phytoplankton have adapted to have the enzymes
necessary to undertake this reduction (nitrate reductase). There are a few notable and well-known
exceptions that include most Prochlorococcus and some Synechococcus that can only take up nitrogen as
ammonium.[42]

The nutrients in the ocean are not uniformly distributed. Areas of upwelling provide supplies of nitrogen
from below the euphotic zone. Coastal zones provide nitrogen from runoff and upwelling occurs readily
along the coast. However, the rate at which nitrogen can be taken up by phytoplankton is decreased in
oligotrophic waters year-round and temperate water in the summer resulting in lower primary
production.[45] The distribution of the different forms of nitrogen varies throughout the oceans as well.

Nitrate is depleted in near-surface water except in upwelling regions. Coastal upwelling regions usually
have high nitrate and chlorophyll levels as a result of the increased production. However, there are regions
of high surface nitrate but low chlorophyll that are referred to as HNLC (high nitrogen, low chlorophyll)
regions. The best explanation for HNLC regions relates to iron scarcity in the ocean, which may play an
important part in ocean dynamics and
nutrient cycles. The input of iron
varies by region and is delivered to the
ocean by dust (from dust storms) and
leached out of rocks. Iron is under
consideration as the true limiting
element to ecosystem productivity in
the ocean.

Ammonium and nitrite show a


maximum concentration at 50–80 m
(lower end of the euphotic zone) with
decreasing concentration below that
depth. This distribution can be
accounted for by the fact that nitrite
and ammonium are intermediate
species. They are both rapidly
produced and consumed through the
water column.[42] The amount of
ammonium in the ocean is about 3
orders of magnitude less than
nitrate.[42] Between ammonium,
nitrite, and nitrate, nitrite has the fastest
turnover rate. It can be produced
during nitrate assimilation, nitrification,
and denitrification; however, it is
immediately consumed again.

New vs. regenerated


nitrogen

Nitrogen entering the euphotic zone is


The main studied processes of the N cycle in different marine
referred to as new nitrogen because it
environments. Every coloured arrow represents a N transformation:
is newly arrived from outside the
productive layer.[41] The new nitrogen N2 fixation (red), nitrification (light blue), nitrate reduction (violet),
DNRA (magenta), denitrification (aquamarine), N-damo (green), and
can come from below the euphotic
anammox (orange). Black curved arrows represent physical
zone or from outside sources. Outside
sources are upwelling from deep water processes such as advection and diffusion.[38]
and nitrogen fixation. If the organic
matter is eaten, respired, delivered to
the water as ammonia, and re-incorporated into organic matter by phytoplankton it is considered
recycled/regenerated production.
New production is an important
component of the marine environment.
One reason is that only continual input
of new nitrogen can determine the total
capacity of the ocean to produce a
sustainable fish harvest.[45] Harvesting
fish from regenerated nitrogen areas
will lead to a decrease in nitrogen and
therefore a decrease in primary
production. This will have a negative
effect on the system. However, if fish
are harvested from areas of new
nitrogen the nitrogen will be
replenished.

Future acidification

As illustrated by the diagram on the Marine nitrogen cycle under future ocean acidification [39]
right, additional carbon dioxide is
absorbed by the ocean and reacts with
+
water, carbonic acid is formed and broken down into both bicarbonate (H2 CO3 ) and hydrogen (H ) ions
(gray arrow), which reduces bioavailable carbonate and decreases ocean pH (black arrow). This is likely to
+
enhance nitrogen fixation by diazotrophs (gray arrow), which utilize H ions to convert nitrogen into
+
bioavailable forms such as ammonia (NH3 ) and ammonium ions (NH4 ). However, as pH decreases, and
more ammonia is converted to ammonium ions (gray arrow), there is less oxidation of ammonia to nitrite

(NO2 ), resulting in an overall decrease in nitrification and denitrification (black arrows). This in turn would
lead to a further build up of fixed nitrogen in the ocean, with the potential consequence of eutrophication.
Gray arrows represent an increase while black arrows represent a decrease in the associated process.[39]

Human influences on the nitrogen cycle


As a result of extensive cultivation of legumes (particularly soy,
alfalfa, and clover), growing use of the Haber–Bosch process in the
creation of chemical fertilizers, and pollution emitted by vehicles
and industrial plants, human beings have more than doubled the
annual transfer of nitrogen into biologically available forms.[28] In
addition, humans have significantly contributed to the transfer of
nitrogen trace gases from Earth to the atmosphere and from the land
to aquatic systems. Human alterations to the global nitrogen cycle
are most intense in developed countries and in Asia, where vehicle
emissions and industrial agriculture are highest.[46] Nitrogen fertilizer application

Generation of Nr, reactive nitrogen, has increased over 10 fold in


the past century due to global industrialisation.[2][47] This form of nitrogen follows a cascade through the
biosphere via a variety of mechanisms, and is accumulating as the rate of its generation is greater than the
rate of denitrification.[48]
Nitrous oxide (N2 O) has risen in the atmosphere as a result of
agricultural fertilization, biomass burning, cattle and feedlots, and
industrial sources.[49] N2 O has deleterious effects in the
stratosphere, where it breaks down and acts as a catalyst in the
destruction of atmospheric ozone. Nitrous oxide is also a
greenhouse gas and is currently the third largest contributor to
global warming, after carbon dioxide and methane. While not as
abundant in the atmosphere as carbon dioxide, it is, for an
equivalent mass, nearly 300 times more potent in its ability to warm
the planet.[50] Nitrogen in manure production

Ammonia (NH3 ) in the atmosphere has tripled as the result of


human activities. It is a reactant in the atmosphere, where it acts as an aerosol, decreasing air quality and
clinging to water droplets, eventually resulting in nitric acid (HNO3 ) that produces acid rain. Atmospheric
ammonia and nitric acid also damage respiratory systems.

The very high temperature of lightning naturally produces small amounts of NOx , NH3 , and HNO3 , but
high-temperature combustion has contributed to a 6- or 7-fold increase in the flux of NOx to the
atmosphere. Its production is a function of combustion temperature - the higher the temperature, the more
NOx is produced. Fossil fuel combustion is a primary contributor, but so are biofuels and even the burning
of hydrogen. However, the rate that hydrogen is directly injected into the combustion chambers of internal
combustion engines can be controlled to prevent the higher combustion temperatures that produce NOx .

Ammonia and nitrous oxides actively alter atmospheric chemistry. They are precursors of tropospheric
(lower atmosphere) ozone production, which contributes to smog and acid rain, damages plants and
increases nitrogen inputs to ecosystems. Ecosystem processes can increase with nitrogen fertilization, but
anthropogenic input can also result in nitrogen saturation, which weakens productivity and can damage the
health of plants, animals, fish, and humans.[28]

Decreases in biodiversity can also result if higher nitrogen availability increases nitrogen-demanding
grasses, causing a degradation of nitrogen-poor, species-diverse heathlands.[51]

Consequence of human modification of the nitrogen cycle

Impacts on natural systems

Increasing levels of nitrogen deposition are shown to have a


number of negative effects on both terrestrial and aquatic
ecosystems.[52][53] Nitrogen gases and aerosols can be directly
toxic to certain plant species, affecting the aboveground physiology
and growth of plants near large point sources of nitrogen pollution.
Estimated nitrogen surplus (the
Changes to plant species may also occur, as accumulation of
difference between inorganic and
nitrogen compounds increase its availability in a given ecosystem,
organic fertilizer application,
eventually changing the species composition, plant diversity, and atmospheric deposition, fixation and
nitrogen cycling. Ammonia and ammonium - two reduced forms of uptake by crops) for the year 2005
nitrogen - can be detrimental over time due to an increased toxicity across Europe.
toward sensitive species of plants,[54] particularly those that are
accustomed to using nitrate as their source of nitrogen, causing
poor development of their roots and shoots. Increased nitrogen deposition also leads to soil acidification,
which increases base cation leaching in the soil and amounts of aluminum and other potentially toxic
metals, along with decreasing the amount of nitrification occurring and increasing plant-derived litter. Due
to the ongoing changes caused by high nitrogen deposition, an environment's susceptibility to ecological
stress and disturbance - such as pests and pathogens - may increase, thus making it less resilient to situations
that otherwise would have little impact to its long-term vitality.

Additional risks posed by increased availability of inorganic nitrogen in aquatic ecosystems include water
acidification; eutrophication of fresh and saltwater systems; and toxicity issues for animals, including
humans.[55] Eutrophication often leads to lower dissolved oxygen levels in the water column, including
hypoxic and anoxic conditions, which can cause death of aquatic fauna. Relatively sessile benthos, or
bottom-dwelling creatures, are particularly vulnerable because of their lack of mobility, though large fish
kills are not uncommon. Oceanic dead zones near the mouth of the Mississippi in the Gulf of Mexico are a
well-known example of algal bloom-induced hypoxia.[56][57] The New York Adirondack Lakes, Catskills,
Hudson Highlands, Rensselaer Plateau and parts of Long Island display the impact of nitric acid rain
deposition, resulting in the killing of fish and many other aquatic species.[58]

Ammonia (NH3 ) is highly toxic to fish and the level of ammonia discharged from wastewater treatment
facilities must be closely monitored. To prevent fish deaths, nitrification via aeration prior to discharge is
often desirable. Land application can be an attractive alternative to the aeration.

Impacts on human health: nitrate accumulation in drinking water

Leakage of Nr (reactive nitrogen) from human activities can cause nitrate accumulation in the natural water
environment, which can create harmful impacts on human health. Excessive use of N-fertilizer in
agriculture has been one of the major sources of nitrate pollution in groundwater and surface water.[59][60]
Due to its high solubility and low retention by soil, nitrate can easily escape from the subsoil layer to the
groundwater, causing nitrate pollution. Some other non-point sources for nitrate pollution in groundwater
are originated from livestock feeding, animal and human contamination and municipal and industrial waste.
Since groundwater often serves as the primary domestic water supply, nitrate pollution can be extended
from groundwater to surface and drinking water in the process of potable water production, especially for
small community water supplies, where poorly regulated and unsanitary waters are used.[61]
− −
The WHO standard for drinking water is 50 mg NO3 L−1 for short-term exposure, and for 3 mg NO3 L−1
chronic effects.[62] Once it enters the human body, nitrate can react with organic compounds through
nitrosation reactions in the stomach to form nitrosamines and nitrosamides, which are involved in some
types of cancers (e.g., oral cancer and gastric cancer).[63]

Impacts on human health: air quality

Human activities have also dramatically altered the global nitrogen cycle via production of nitrogenous
gases, associated with the global atmospheric nitrogen pollution. There are multiple sources of atmospheric
reactive nitrogen (Nr) fluxes. Agricultural sources of reactive nitrogen can produce atmospheric emission of
ammonia (NH3 ), nitrogen oxides (NOx ) and nitrous oxide (N2 O). Combustion processes in energy
production, transportation and industry can also result in the formation of new reactive nitrogen via the
emission of NOx , an unintentional waste product. When those reactive nitrogens are released to the lower
atmosphere, they can induce the formation of smog, particulate matter and aerosols, all of which are major
contributors to adverse health effects on human health from air pollution.[64] In the atmosphere, NO2 can
be oxidized to nitric acid (HNO3 ), and it can further react with NH3 to form ammonium nitrate, which
facilitates the formation of particular nitrate. Moreover, NH3 can react with other acid gases (sulfuric and
hydrochloric acids) to form ammonium-containing particles, which are the precursors for the secondary
organic aerosol particles in photochemical smog.[65]

See also
Planetary boundaries – Limits not to be exceeded if humanity wants to survive in a safe
ecosystem
Phosphorus cycle – Biogeochemical movement

References
1. Fowler, David; Coyle, Mhairi; Skiba, Ute; Sutton, Mark A.; Cape, J. Neil; Reis, Stefan;
Sheppard, Lucy J.; Jenkins, Alan; Grizzetti, Bruna; Galloway, JN; Vitousek, P; Leach, A;
Bouwman, AF; Butterbach-Bahl, K; Dentener, F; Stevenson, D; Amann, M; Voss, M (5 July
2013). "The global nitrogen cycle in the twenty-first century" (https://www.ncbi.nlm.nih.gov/p
mc/articles/PMC3682748). Philosophical Transactions of the Royal Society of London.
Series B, Biological Sciences. 368 (1621): 20130164. doi:10.1098/rstb.2013.0164 (https://do
i.org/10.1098%2Frstb.2013.0164). PMC 3682748 (https://www.ncbi.nlm.nih.gov/pmc/articles/
PMC3682748). PMID 23713126 (https://pubmed.ncbi.nlm.nih.gov/23713126).
2. Galloway, J. N.; Townsend, A. R.; Erisman, J. W.; Bekunda, M.; Cai, Z.; Freney, J. R.;
Martinelli, L. A.; Seitzinger, S. P.; Sutton, M. A. (2008). "Transformation of the Nitrogen Cycle:
Recent Trends, Questions, and Potential Solutions" (http://156.35.33.98/ranadon/Ricardo_A
nadon/docencia/DoctoradoEconomia/241eGallowayNitrogenTrend08.pdf) (PDF). Science.
320 (5878): 889–892. Bibcode:2008Sci...320..889G (https://ui.adsabs.harvard.edu/abs/2008
Sci...320..889G). doi:10.1126/science.1136674 (https://doi.org/10.1126%2Fscience.113667
4). ISSN 0036-8075 (https://www.worldcat.org/issn/0036-8075). PMID 18487183 (https://pub
med.ncbi.nlm.nih.gov/18487183). S2CID 16547816 (https://api.semanticscholar.org/CorpusI
D:16547816). Archived (https://web.archive.org/web/20111108115937/http://156.35.33.98/ra
nadon/Ricardo_Anadon/docencia/DoctoradoEconomia/241eGallowayNitrogenTrend08.pdf)
(PDF) from the original on 2011-11-08. Retrieved 2019-09-23.
3. Vitousek, P. M.; Menge, D. N. L.; Reed, S. C.; Cleveland, C. C. (2013). "Biological nitrogen
fixation: rates, patterns and ecological controls in terrestrial ecosystems" (https://www.ncbi.nl
m.nih.gov/pmc/articles/PMC3682739). Philosophical Transactions of the Royal Society B:
Biological Sciences. 368 (1621): 20130119. doi:10.1098/rstb.2013.0119 (https://doi.org/10.1
098%2Frstb.2013.0119). PMC 3682739 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC368
2739). PMID 23713117 (https://pubmed.ncbi.nlm.nih.gov/23713117).
4. Voss, M.; Bange, H. W.; Dippner, J. W.; Middelburg, J. J.; Montoya, J. P.; Ward, B. (2013).
"The marine nitrogen cycle: recent discoveries, uncertainties and the potential relevance of
climate change" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3682741). Philosophical
Transactions of the Royal Society B: Biological Sciences. 368 (1621): 20130121.
doi:10.1098/rstb.2013.0121 (https://doi.org/10.1098%2Frstb.2013.0121). PMC 3682741 (http
s://www.ncbi.nlm.nih.gov/pmc/articles/PMC3682741). PMID 23713119 (https://pubmed.ncbi.
nlm.nih.gov/23713119).
5. Fowler, David; Coyle, Mhairi; Skiba, Ute; Sutton, Mark A.; Cape, J. Neil; Reis, Stefan;
Sheppard, Lucy J.; Jenkins, Alan; Grizzetti, Bruna; Galloway, JN; Vitousek, P; Leach, A;
Bouwman, AF; Butterbach-Bahl, K; Dentener, F; Stevenson, D; Amann, M; Voss, M (5 July
2013). "The global nitrogen cycle in the twenty-first century" (https://www.ncbi.nlm.nih.gov/p
mc/articles/PMC3682748). Philosophical Transactions of the Royal Society of London.
Series B, Biological Sciences. 368 (1621): 20130164. doi:10.1098/rstb.2013.0164 (https://do
i.org/10.1098%2Frstb.2013.0164). PMC 3682748 (https://www.ncbi.nlm.nih.gov/pmc/articles/
PMC3682748). PMID 23713126 (https://pubmed.ncbi.nlm.nih.gov/23713126).
6. Vuuren, Detlef P van; Bouwman, Lex F; Smith, Steven J; Dentener, Frank (2011). "Global
projections for anthropogenic reactive nitrogen emissions to the atmosphere: an assessment
of scenarios in the scientific literature". Current Opinion in Environmental Sustainability. 3
(5): 359–369. Bibcode:2011COES....3..359V (https://ui.adsabs.harvard.edu/abs/2011COE
S....3..359V). doi:10.1016/j.cosust.2011.08.014 (https://doi.org/10.1016%2Fj.cosust.2011.08.
014). hdl:1874/314192 (https://hdl.handle.net/1874%2F314192). ISSN 1877-3435 (https://w
ww.worldcat.org/issn/1877-3435). S2CID 154935568 (https://api.semanticscholar.org/Corpu
sID:154935568).
7. Pilegaard, K. (2013). "Processes regulating nitric oxide emissions from soils" (https://www.n
cbi.nlm.nih.gov/pmc/articles/PMC3682746). Philosophical Transactions of the Royal Society
B: Biological Sciences. 368 (1621): 20130126. doi:10.1098/rstb.2013.0126 (https://doi.org/1
0.1098%2Frstb.2013.0126). PMC 3682746 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC
3682746). PMID 23713124 (https://pubmed.ncbi.nlm.nih.gov/23713124).
8. Levy, H.; Moxim, W. J.; Kasibhatla, P. S. (1996). "A global three-dimensional time-dependent
lightning source of tropospheric NOx". Journal of Geophysical Research: Atmospheres. 101
(D17): 22911–22922. Bibcode:1996JGR...10122911L (https://ui.adsabs.harvard.edu/abs/19
96JGR...10122911L). doi:10.1029/96jd02341 (https://doi.org/10.1029%2F96jd02341).
ISSN 0148-0227 (https://www.worldcat.org/issn/0148-0227).
9. Sutton, M. A.; Reis, S.; Riddick, S. N.; Dragosits, U.; Nemitz, E.; Theobald, M. R.; Tang, Y. S.;
Braban, C. F.; Vieno, M. (2013). "Towards a climate-dependent paradigm of ammonia
emission and deposition" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3682750).
Philosophical Transactions of the Royal Society B: Biological Sciences. 368 (1621):
20130166. doi:10.1098/rstb.2013.0166 (https://doi.org/10.1098%2Frstb.2013.0166).
PMC 3682750 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3682750). PMID 23713128
(https://pubmed.ncbi.nlm.nih.gov/23713128).
10. Dentener, F.; Drevet, J.; Lamarque, J. F.; Bey, I.; Eickhout, B.; Fiore, A. M.; Hauglustaine, D.;
Horowitz, L. W.; Krol, M. (2006). "Nitrogen and sulfur deposition on regional and global
scales: a multimodel evaluation" (https://hal.archives-ouvertes.fr/hal-00342326/file/Dentener
_et_al-2006-Global_Biogeochemical_Cycles.pdf) (PDF). Global Biogeochemical Cycles. 20
(4): n/a. Bibcode:2006GBioC..20.4003D (https://ui.adsabs.harvard.edu/abs/2006GBioC..20.
4003D). doi:10.1029/2005GB002672 (https://doi.org/10.1029%2F2005GB002672).
S2CID 839759 (https://api.semanticscholar.org/CorpusID:839759).
11. Duce, R. A.; LaRoche, J.; Altieri, K.; Arrigo, K. R.; Baker, A. R.; Capone, D. G.; Cornell, S.;
Dentener, F.; Galloway, J. (2008). "Impacts of Atmospheric Anthropogenic Nitrogen on the
Open Ocean" (https://semanticscholar.org/paper/c662d779132f5d8d12ed259621cee3fff148
804f). Science. 320 (5878): 893–7. Bibcode:2008Sci...320..893D (https://ui.adsabs.harvard.e
du/abs/2008Sci...320..893D). doi:10.1126/science.1150369 (https://doi.org/10.1126%2Fscie
nce.1150369). hdl:21.11116/0000-0001-CD7A-0 (https://hdl.handle.net/21.11116%2F0000-
0001-CD7A-0). PMID 18487184 (https://pubmed.ncbi.nlm.nih.gov/18487184).
S2CID 11204131 (https://api.semanticscholar.org/CorpusID:11204131).
12. Bouwman, L.; Goldewijk, K. K.; Van Der Hoek, K. W.; Beusen, A. H. W.; Van Vuuren, D. P.;
Willems, J.; Rufino, M. C.; Stehfest, E. (2011-05-16). "Exploring global changes in nitrogen
and phosphorus cycles in agriculture induced by livestock production over the 1900-2050
period" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3876211). Proceedings of the
National Academy of Sciences. 110 (52): 20882–7. doi:10.1073/pnas.1012878108 (https://d
oi.org/10.1073%2Fpnas.1012878108). PMC 3876211 (https://www.ncbi.nlm.nih.gov/pmc/arti
cles/PMC3876211). PMID 21576477 (https://pubmed.ncbi.nlm.nih.gov/21576477).
13. Solomon, Susan (2007). Climate change 2007 : the physical science basis. Published for
the Intergovernmental Panel on Climate Change [by] Cambridge University Press.
ISBN 9780521880091. OCLC 228429704 (https://www.worldcat.org/oclc/228429704).
14. Sutton, Mark A., ed. (2011-04-14). The European nitrogen assessment : sources, effects, and
policy perspectives. Cambridge University Press. ISBN 9781107006126. OCLC 690090202
(https://www.worldcat.org/oclc/690090202).
15. Deutsch, Curtis; Sarmiento, Jorge L.; Sigman, Daniel M.; Gruber, Nicolas; Dunne, John P.
(2007). "Spatial coupling of nitrogen inputs and losses in the ocean" (https://www.semantics
cholar.org/paper/52dc58be6fca3997f3bfd2f2f1f79c7b956c57ab). Nature. 445 (7124): 163–
167. Bibcode:2007Natur.445..163D (https://ui.adsabs.harvard.edu/abs/2007Natur.445..163
D). doi:10.1038/nature05392 (https://doi.org/10.1038%2Fnature05392). ISSN 0028-0836 (htt
ps://www.worldcat.org/issn/0028-0836). PMID 17215838 (https://pubmed.ncbi.nlm.nih.gov/1
7215838). S2CID 10804715 (https://api.semanticscholar.org/CorpusID:10804715).
16. Steven B. Carroll; Steven D. Salt (2004). Ecology for gardeners (https://books.google.com/b
ooks?id=aM4W9e5nmsoC&pg=PA93). Timber Press. p. 93. ISBN 978-0-88192-611-8.
17. Kuypers, MMM; Marchant, HK; Kartal, B (2011). "The Microbial Nitrogen-Cycling Network"
(https://www.semanticscholar.org/paper/ec5210fe0ef4ec95865bbd684127bf77abd782f1).
Nature Reviews Microbiology. 1 (1): 1–14. doi:10.1038/nrmicro.2018.9 (https://doi.org/10.10
38%2Fnrmicro.2018.9). hdl:21.11116/0000-0003-B828-1 (https://hdl.handle.net/21.11116%2
F0000-0003-B828-1). PMID 29398704 (https://pubmed.ncbi.nlm.nih.gov/29398704).
S2CID 3948918 (https://api.semanticscholar.org/CorpusID:3948918).
18. Galloway, J. N.; et al. (2004). "Nitrogen cycles: past, present, and future generations" (https://
www.semanticscholar.org/paper/4634b9c9305e9405c84d2b2f0f2012b825278378).
Biogeochemistry. 70 (2): 153–226. doi:10.1007/s10533-004-0370-0 (https://doi.org/10.100
7%2Fs10533-004-0370-0). S2CID 98109580 (https://api.semanticscholar.org/CorpusID:981
09580).
19. Reis, Stefan; Bekunda, Mateete; Howard, Clare M; Karanja, Nancy; Winiwarter, Wilfried;
Yan, Xiaoyuan; Bleeker, Albert; Sutton, Mark A (2016-12-01). "Synthesis and review:
Tackling the nitrogen management challenge: from global to local scales" (https://doi.org/10.
1088%2F1748-9326%2F11%2F12%2F120205). Environmental Research Letters. 11 (12):
120205. Bibcode:2016ERL....11l0205R (https://ui.adsabs.harvard.edu/abs/2016ERL....11l02
05R). doi:10.1088/1748-9326/11/12/120205 (https://doi.org/10.1088%2F1748-9326%2F11%
2F12%2F120205). ISSN 1748-9326 (https://www.worldcat.org/issn/1748-9326).
20. Gu, Baojing; Ge, Ying; Ren, Yuan; Xu, Bin; Luo, Weidong; Jiang, Hong; Gu, Binhe; Chang,
Jie (2012-08-17). "Atmospheric Reactive Nitrogen in China: Sources, Recent Trends, and
Damage Costs". Environmental Science & Technology. 46 (17): 9420–9427.
Bibcode:2012EnST...46.9420G (https://ui.adsabs.harvard.edu/abs/2012EnST...46.9420G).
doi:10.1021/es301446g (https://doi.org/10.1021%2Fes301446g). ISSN 0013-936X (https://w
ww.worldcat.org/issn/0013-936X). PMID 22852755 (https://pubmed.ncbi.nlm.nih.gov/228527
55).
21. Kim, Haryun; Lee, Kitack; Lim, Dhong-Il; Nam, Seung-Il; Kim, Tae-Wook; Yang, Jin-Yu T.; Ko,
Young Ho; Shin, Kyung-Hoon; Lee, Eunil (2017-05-11). "Widespread Anthropogenic
Nitrogen in Northwestern Pacific Ocean Sediment". Environmental Science & Technology.
51 (11): 6044–52. Bibcode:2017EnST...51.6044K (https://ui.adsabs.harvard.edu/abs/2017En
ST...51.6044K). doi:10.1021/acs.est.6b05316 (https://doi.org/10.1021%2Facs.est.6b05316).
ISSN 0013-936X (https://www.worldcat.org/issn/0013-936X). PMID 28462990 (https://pubme
d.ncbi.nlm.nih.gov/28462990).
22. Moir, JWB, ed. (2011). Nitrogen Cycling in Bacteria: Molecular Analysis. Caister Academic
Press. ISBN 978-1-904455-86-8.
23. Smith, B.; Richards, R.L.; Newton, W.E. (2013) [2004]. Catalysts for nitrogen fixation:
nitrogenases, relevant chemical models and commercial processes (https://books.google.co
m/books?id=gJnSBwAAQBAJ&pg=PR5). Kluwer. ISBN 9781402036118.
24. Smil, V (1997). Cycles of Life. Scientific American Library. ISBN 9780716750796.
25. Willey, Joanne M. (2011). Prescott's Microbiology (8th ed.). McGraw Hill. p. 705. ISBN 978-
0-07-337526-7.
26. Sparacino-Watkins, Courtney; Stolz, John F.; Basu, Partha (2013-12-16). "Nitrate and
periplasmic nitrate reductases" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4080430).
Chem. Soc. Rev. 43 (2): 676–706. doi:10.1039/c3cs60249d (https://doi.org/10.1039%2Fc3cs
60249d). PMC 4080430 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4080430).
PMID 24141308 (https://pubmed.ncbi.nlm.nih.gov/24141308).
27. Simon, Jörg; Klotz, Martin G. (2013). "Diversity and evolution of bioenergetic systems
involved in microbial nitrogen compound transformations" (https://doi.org/10.1016%2Fj.bbab
io.2012.07.005). Biochimica et Biophysica Acta (BBA) - Bioenergetics. 1827 (2): 114–135.
doi:10.1016/j.bbabio.2012.07.005 (https://doi.org/10.1016%2Fj.bbabio.2012.07.005).
PMID 22842521 (https://pubmed.ncbi.nlm.nih.gov/22842521).
28. Vitousek, PM; Aber, J; Howarth, RW; Likens, GE; Matson, PA; Schindler, DW; Schlesinger,
WH; Tilman, GD (1997). "Human alteration of the global nitrogen cycle: Sources and
consequences" (https://ecommons.cornell.edu/bitstream/1813/60830/1/Vitousek_et_al-1997
-Ecological_Applications.pdf) (PDF). Ecological Applications. 1 (3): 1–17. doi:10.1890/1051-
0761(1997)007[0737:HAOTGN]2.0.CO;2 (https://doi.org/10.1890%2F1051-0761%281997%
29007%5B0737%3AHAOTGN%5D2.0.CO%3B2). hdl:1813/60830 (https://hdl.handle.net/18
13%2F60830). ISSN 1051-0761 (https://www.worldcat.org/issn/1051-0761).
29. Graf, Jon S.; Schorn, Sina; Kitzinger, Katharina; Ahmerkamp, Soeren; Woehle, Christian;
Huettel, Bruno; Schubert, Carsten J.; Kuypers, Marcel M. M.; Milucka, Jana (3 March 2021).
"Anaerobic endosymbiont generates energy for ciliate host by denitrification" (https://www.nc
bi.nlm.nih.gov/pmc/articles/PMC7969357). Nature. 591 (7850): 445–450.
Bibcode:2021Natur.591..445G (https://ui.adsabs.harvard.edu/abs/2021Natur.591..445G).
doi:10.1038/s41586-021-03297-6 (https://doi.org/10.1038%2Fs41586-021-03297-6).
ISSN 0028-0836 (https://www.worldcat.org/issn/0028-0836). PMC 7969357 (https://www.ncb
i.nlm.nih.gov/pmc/articles/PMC7969357). PMID 33658719 (https://pubmed.ncbi.nlm.nih.gov/
33658719).
30. Lam, Phyllis; Kuypers, Marcel M.M. (2011). "Microbial Nitrogen Processes in Oxygen
Minimum Zones". Annual Review of Marine Science. 3: 317–345.
Bibcode:2011ARMS....3..317L (https://ui.adsabs.harvard.edu/abs/2011ARMS....3..317L).
doi:10.1146/annurev-marine-120709-142814 (https://doi.org/10.1146%2Fannurev-marine-1
20709-142814). hdl:21.11116/0000-0001-CA25-2 (https://hdl.handle.net/21.11116%2F0000
-0001-CA25-2). PMID 21329208 (https://pubmed.ncbi.nlm.nih.gov/21329208).
31. Marchant, H.K.; Lavik, G.; Holtappels, M.; Kuypers, M.M.M. (2014). "The Fate of Nitrate in
Intertidal Permeable Sediments" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4134218).
PLOS ONE. 9 (8): e104517. Bibcode:2014PLoSO...9j4517M (https://ui.adsabs.harvard.edu/
abs/2014PLoSO...9j4517M). doi:10.1371/journal.pone.0104517 (https://doi.org/10.1371%2F
journal.pone.0104517). PMC 4134218 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4134
218). PMID 25127459 (https://pubmed.ncbi.nlm.nih.gov/25127459).
32. "Anammox" (https://microbewiki.kenyon.edu/index.php/Anammox). Anammox - MicrobeWiki.
MicrobeWiki. Archived (https://web.archive.org/web/20150927080721/http://microbewiki.ken
yon.edu/index.php/Anammox) from the original on 2015-09-27. Retrieved 5 July 2015.
33. "Nitrogen Study Could 'Rock' A Plant's World" (https://www.npr.org/2011/09/06/140206913/d
iscovery-forces-scientists-to-rethink-nitrogen). NPR.org. 2011-09-06. Archived (https://web.ar
chive.org/web/20111205114315/http://www.npr.org/2011/09/06/140206913/discovery-forces
-scientists-to-rethink-nitrogen) from the original on 2011-12-05. Retrieved 2011-10-22.
34. Schuur, E. A. G. (2011). "Ecology: Nitrogen from the deep" (https://doi.org/10.1038%2F4770
39a). Nature. 477 (7362): 39–40. Bibcode:2011Natur.477...39S (https://ui.adsabs.harvard.ed
u/abs/2011Natur.477...39S). doi:10.1038/477039a (https://doi.org/10.1038%2F477039a).
PMID 21886152 (https://pubmed.ncbi.nlm.nih.gov/21886152). S2CID 2946571 (https://api.se
manticscholar.org/CorpusID:2946571).
35. Morford, S. L.; Houlton, B. Z.; Dahlgren, R. A. (2011). "Increased forest ecosystem carbon
and nitrogen storage from nitrogen rich bedrock" (https://www.semanticscholar.org/paper/a9
3c66e5b6cd5bbc4b6281ca50f331bfc860f584). Nature. 477 (7362): 78–81.
Bibcode:2011Natur.477...78M (https://ui.adsabs.harvard.edu/abs/2011Natur.477...78M).
doi:10.1038/nature10415 (https://doi.org/10.1038%2Fnature10415). PMID 21886160 (https://
pubmed.ncbi.nlm.nih.gov/21886160). S2CID 4352571 (https://api.semanticscholar.org/Corp
usID:4352571).
36. Burgin, Amy J.; Yang, Wendy H.; Hamilton, Stephen K.; Silver, Whendee L. (2011). "Beyond
carbon and nitrogen: how the microbial energy economy couples elemental cycles in
diverse ecosystems". Frontiers in Ecology and the Environment. 9 (1): 44–52.
doi:10.1890/090227 (https://doi.org/10.1890%2F090227). hdl:1808/21008 (https://hdl.handl
e.net/1808%2F21008). ISSN 1540-9309 (https://www.worldcat.org/issn/1540-9309).
37. Roman, J.; McCarthy, J.J. (2010). "The Whale Pump: Marine Mammals Enhance Primary
Productivity in a Coastal Basin" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC2952594).
PLOS ONE. 5 (10): e13255. Bibcode:2010PLoSO...513255R (https://ui.adsabs.harvard.edu/
abs/2010PLoSO...513255R). doi:10.1371/journal.pone.0013255 (https://doi.org/10.1371%2
Fjournal.pone.0013255). PMC 2952594 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC295
2594). PMID 20949007 (https://pubmed.ncbi.nlm.nih.gov/20949007).
38. Pajares Moreno, S.; Ramos, R. (2019). "Processes and Microorganisms Involved in the
Marine Nitrogen Cycle: Knowledge and Gaps" (https://doi.org/10.3389%2Ffmars.2019.0073
9). Frontiers in Marine Science. 6: 739. doi:10.3389/fmars.2019.00739 (https://doi.org/10.338
9%2Ffmars.2019.00739). Material was copied from this source, which is available
under a Creative Commons Attribution 4.0 International License (https://creativecommons.or
g/licenses/by/4.0/).
39. O'Brien, Paul A.; Morrow, Kathleen M.; Willis, Bette L.; Bourne, David G. (2016).
"Implications of Ocean Acidification for Marine Microorganisms from the Free-Living to the
Host-Associated" (https://doi.org/10.3389%2Ffmars.2016.00047). Frontiers in Marine
Science. 3. doi:10.3389/fmars.2016.00047 (https://doi.org/10.3389%2Ffmars.2016.00047).
Material was copied from this source, which is available under a Creative Commons
Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/).
40. Moulton, Orissa M; Altabet, Mark A; Beman, J Michael; Deegan, Linda A; Lloret, Javier;
Lyons, Meaghan K; Nelson, James A; Pfister, Catherine A (May 2016). "Microbial
associations with macrobiota in coastal ecosystems: patterns and implications for nitrogen
cycling". Frontiers in Ecology and the Environment. 14 (4): 200–8. doi:10.1002/fee.1262 (htt
ps://doi.org/10.1002%2Ffee.1262). hdl:1912/8083 (https://hdl.handle.net/1912%2F8083).
ISSN 1540-9295 (https://www.worldcat.org/issn/1540-9295).
41. Miller, Charles (2008). Biological Oceanography. Blackwell. pp. 60–62. ISBN 978-0-632-
05536-4.
42. Gruber, Nicolas (2008). Nitrogen in the Marine Environment. Elsevier. pp. 1–35. ISBN 978-0-
12-372522-6.
43. Miller, Charles (2008). Biological oceanography. Blackwell. pp. 60–62. ISBN 978-0-632-
05536-4.
44. Boyes, Elliot, Susan, Michael. "Learning Unit: Nitrogen Cycle Marine Environment" (https://w
eb.archive.org/web/20120415123617/http://www.chemgapedia.de/vsengine/vlu/vsc/en/ch/1
6/uc/vlus/nitrogenmarine.vlu.html). Archived from the original (http://www.chemgapedia.de/vs
engine/vlu/vsc/en/ch/16/uc/vlus/nitrogenmarine.vlu.html) on 15 April 2012. Retrieved
22 October 2011.
45. Lalli, Parsons, Carol, Timothy (1997). Biological oceanography: An introduction. Butterworth-
Heinemann. ISBN 978-0-7506-3384-0.
46. Holland, Elisabeth A.; Dentener, Frank J.; Braswell, Bobby H.; Sulzman, James M. (1999).
"Contemporary and pre-industrial global reactive nitrogen budgets". Biogeochemistry. 46
(1–3): 7. doi:10.1007/BF01007572 (https://doi.org/10.1007%2FBF01007572).
S2CID 189917368 (https://api.semanticscholar.org/CorpusID:189917368).
47. Gu, Baojing; Ge, Ying; Ren, Yuan; Xu, Bin; Luo, Weidong; Jiang, Hong; Gu, Binhe; Chang,
Jie (2012-09-04). "Atmospheric Reactive Nitrogen in China: Sources, Recent Trends, and
Damage Costs". Environmental Science & Technology. 46 (17): 9420–7.
Bibcode:2012EnST...46.9420G (https://ui.adsabs.harvard.edu/abs/2012EnST...46.9420G).
doi:10.1021/es301446g (https://doi.org/10.1021%2Fes301446g). ISSN 0013-936X (https://w
ww.worldcat.org/issn/0013-936X). PMID 22852755 (https://pubmed.ncbi.nlm.nih.gov/228527
55).
48. Cosby, B. Jack; Cowling, Ellis B.; Howarth, Robert W.; Seitzinger, Sybil P.; Erisman, Jan
Willem; Aber, John D.; Galloway, James N. (2003-04-01). "The Nitrogen Cascade" (https://do
i.org/10.1641%2F0006-3568%282003%29053%5B0341%3ATNC%5D2.0.CO%3B2).
BioScience. 53 (4): 341–356. doi:10.1641/0006-3568(2003)053[0341:TNC]2.0.CO;2 (https://
doi.org/10.1641%2F0006-3568%282003%29053%5B0341%3ATNC%5D2.0.CO%3B2).
ISSN 0006-3568 (https://www.worldcat.org/issn/0006-3568). S2CID 3356400 (https://api.se
manticscholar.org/CorpusID:3356400).
49. Chapin, S.F.; Matson, P.A.; Mooney, H.A. (2002). Principles of Terrestrial Ecosystem Ecology
(https://books.google.com/books?id=OOH1H779-7EC&pg=PA345). Springer. p. 345.
ISBN 0-387-95443-0.
50. Proceedings of the Scientific Committee on Problems of the Environment (SCOPE)
International Biofuels Project Rapid Assessment, 22–25 September 2008, Gummersbach,
Germany, R. W. Howarth and S. Bringezu, editors. 2009 Executive Summary, p. 3 (http://cip.c
ornell.edu/biofuels/) Archived (https://web.archive.org/web/20090606093439/http://cip.cornel
l.edu/biofuels/) 2009-06-06 at the Wayback Machine
51. Aerts, Rien & Berendse, Frank (1988). "The Effect of Increased Nutrient Availability on
Vegetation Dynamics in Wet Heathlands". Vegetatio. 76 (1/2): 63–69.
doi:10.1007/BF00047389 (https://doi.org/10.1007%2FBF00047389). JSTOR 20038308 (http
s://www.jstor.org/stable/20038308). S2CID 34882407 (https://api.semanticscholar.org/Corpu
sID:34882407).
52. Bobbink, R.; Hicks, K.; Galloway, J.; Spranger, T.; Alkemade, R.; Ashmore, M.; Bustamante,
M.; Cinderby, S.; Davidson, E. (2010-01-01). "Global assessment of nitrogen deposition
effects on terrestrial plant diversity: a synthesis" (http://eprints.whiterose.ac.uk/10814/1/0Bob
binketal2010.pdf) (PDF). Ecological Applications. 20 (1): 30–59. doi:10.1890/08-1140.1 (http
s://doi.org/10.1890%2F08-1140.1). ISSN 1939-5582 (https://www.worldcat.org/issn/1939-55
82). PMID 20349829 (https://pubmed.ncbi.nlm.nih.gov/20349829). S2CID 4792945 (https://a
pi.semanticscholar.org/CorpusID:4792945). Archived (https://web.archive.org/web/20190930
190859/http://eprints.whiterose.ac.uk/10814/1/0Bobbinketal2010.pdf) (PDF) from the original
on 2019-09-30. Retrieved 2019-09-30.
53. Liu, Xuejun; Duan, Lei; Mo, Jiangming; Du, Enzai; Shen, Jianlin; Lu, Xiankai; Zhang, Ying;
Zhou, Xiaobing; He, Chune (2011). "Nitrogen deposition and its ecological impact in China:
An overview". Environmental Pollution. 159 (10): 2251–2264.
doi:10.1016/j.envpol.2010.08.002 (https://doi.org/10.1016%2Fj.envpol.2010.08.002).
PMID 20828899 (https://pubmed.ncbi.nlm.nih.gov/20828899).
54. Britto, Dev T.; Kronzucker, Herbert J. (2002). "NH4+ toxicity in higher plants: a critical
review". Journal of Plant Physiology. 159 (6): 567–584. doi:10.1078/0176-1617-0774 (http
s://doi.org/10.1078%2F0176-1617-0774).
55. Camargoa, Julio A.; Alonso, Álvaro (2006). "Ecological and toxicological effects of inorganic
nitrogen pollution in aquatic ecosystems: A global assessment". Environment International.
32 (6): 831–849. doi:10.1016/j.envint.2006.05.002 (https://doi.org/10.1016%2Fj.envint.2006.
05.002). hdl:10261/294824 (https://hdl.handle.net/10261%2F294824). PMID 16781774 (http
s://pubmed.ncbi.nlm.nih.gov/16781774).
56. Rabalais, Nancy N.; Turner, R. Eugene; Wiseman, William J. Jr. (2002). "Gulf of Mexico
Hypoxia, aka "The Dead Zone" ". Annu. Rev. Ecol. Syst. 33: 235–63.
doi:10.1146/annurev.ecolsys.33.010802.150513 (https://doi.org/10.1146%2Fannurev.ecolsy
s.33.010802.150513). JSTOR 3069262 (https://www.jstor.org/stable/3069262).
57. Dybas, Cheryl Lyn. (2005). "Dead Zones Spreading in World Oceans" (https://doi.org/10.164
1%2F0006-3568%282005%29055%5B0552%3ADZSIWO%5D2.0.CO%3B2). BioScience.
55 (7): 552–557. doi:10.1641/0006-3568(2005)055[0552:DZSIWO]2.0.CO;2 (https://doi.org/1
0.1641%2F0006-3568%282005%29055%5B0552%3ADZSIWO%5D2.0.CO%3B2).
58. New York State Environmental Conservation, Environmental Impacts of Acid Deposition:
Lakes [1] (http://www.dec.ny.gov/chemical/8631.html) Archived (https://web.archive.org/web/
20101124171015/http://www.dec.ny.gov/chemical/8631.html) 2010-11-24 at the Wayback
Machine
59. Power, J.F.; Schepers, J.S. (1989). "Nitrate contamination of groundwater in North America".
Agriculture, Ecosystems & Environment. 26 (3–4): 165–187. doi:10.1016/0167-
8809(89)90012-1 (https://doi.org/10.1016%2F0167-8809%2889%2990012-1). ISSN 0167-
8809 (https://www.worldcat.org/issn/0167-8809).
60. Strebel, O.; Duynisveld, W.H.M.; Böttcher, J. (1989). "Nitrate pollution of groundwater in
western Europe". Agriculture, Ecosystems & Environment. 26 (3–4): 189–214.
doi:10.1016/0167-8809(89)90013-3 (https://doi.org/10.1016%2F0167-8809%2889%299001
3-3). ISSN 0167-8809 (https://www.worldcat.org/issn/0167-8809).
61. Fewtrell, Lorna (2004). "Drinking-Water Nitrate, Methemoglobinemia, and Global Burden of
Disease: A Discussion" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC1247562).
Environmental Health Perspectives. 112 (14): 1371–4. doi:10.1289/ehp.7216 (https://doi.org/
10.1289%2Fehp.7216). PMC 1247562 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC1247
562). PMID 15471727 (https://pubmed.ncbi.nlm.nih.gov/15471727).
62. Global Health Observatory : (GHO). World Health Organization. OCLC 50144984 (https://ww
w.worldcat.org/oclc/50144984).
63. Canter, Larry W. (2019-01-22), "Illustrations of Nitrate Pollution of Groundwater", Nitrates in
Groundwater, Routledge, pp. 39–71, doi:10.1201/9780203745793-3 (https://doi.org/10.120
1%2F9780203745793-3), ISBN 9780203745793, S2CID 133944481 (https://api.semanticsc
holar.org/CorpusID:133944481)
64. Kampa, Marilena; Castanas, Elias (2008). "Human health effects of air pollution".
Environmental Pollution. 151 (2): 362–367. doi:10.1016/j.envpol.2007.06.012 (https://doi.or
g/10.1016%2Fj.envpol.2007.06.012). PMID 17646040 (https://pubmed.ncbi.nlm.nih.gov/176
46040). S2CID 38513536 (https://api.semanticscholar.org/CorpusID:38513536).
65. Erisman, J. W.; Galloway, J. N.; Seitzinger, S.; Bleeker, A.; Dise, N. B.; Petrescu, A. M. R.;
Leach, A. M.; de Vries, W. (2013-05-27). "Consequences of human modification of the global
nitrogen cycle" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3682738). Philosophical
Transactions of the Royal Society B: Biological Sciences. 368 (1621): 20130116.
doi:10.1098/rstb.2013.0116 (https://doi.org/10.1098%2Frstb.2013.0116). PMC 3682738 (http
s://www.ncbi.nlm.nih.gov/pmc/articles/PMC3682738). PMID 23713116 (https://pubmed.ncbi.
nlm.nih.gov/23713116).

Retrieved from "https://en.wikipedia.org/w/index.php?title=Nitrogen_cycle&oldid=1170051813"

You might also like