You are on page 1of 9

Phosphorus cycle

The phosphorus cycle is the


biogeochemical cycle that describes the
movement of phosphorus through the
lithosphere, hydrosphere, and biosphere.
Unlike many other biogeochemical cycles,
the atmosphere does not play a significant
role in the movement of phosphorus,
because phosphorus and phosphorus-based
compounds are usually solids at the typical
ranges of temperature and pressure found
on Earth. The production of phosphine gas
occurs in only specialized, local conditions.
Phosphorus cycle
Therefore, the phosphorus cycle should be
viewed from whole Earth system and then
specifically focused on the cycle in terrestrial and aquatic systems.

Living organisms require phosphorus, a vital component of DNA, RNA, ATP, etc, for their proper
functioning. Plants assimilate phosphorus as phosphate and incorporate it into organic compounds and in
animals, phosphorus is a key component of bones, teeth, etc. On the land, phosphorus gradually becomes
less available to plants over thousands of years, since it is slowly lost in runoff. Low concentration of
phosphorus in soils reduces plant growth and slows soil microbial growth, as shown in studies of soil
microbial biomass. Soil microorganisms act as both sinks and sources of available phosphorus in the
biogeochemical cycle.[1] Short-term transformation of phosphorus is chemical, biological, or
microbiological. In the long-term global cycle, however, the major transfer is driven by tectonic movement
over geologic time.[2]

Humans have caused major changes to the global phosphorus cycle through shipping of phosphorus
minerals, and use of phosphorus fertilizer, and also the shipping of food from farms to cities, where it is lost
as effluent.

Phosphorus in the environment

Ecological function

Phosphorus is an essential nutrient for plants and animals. Phosphorus is a limiting nutrient for aquatic
organisms. Phosphorus forms parts of important life-sustaining molecules that are very common in the
biosphere. Phosphorus does enter the atmosphere in very small amounts when the dust is dissolved in
rainwater and seaspray but remains mostly on land and in rock and soil minerals. Eighty percent of the
mined phosphorus is used to make fertilizers. Phosphates from fertilizers, sewage and detergents can cause
pollution in lakes and streams. Over-enrichment of phosphate in both fresh and inshore marine waters can
lead to massive algae blooms. In fresh water, the death and decay of these blooms leads to eutrophication.
An example of this is the Canadian Experimental Lakes Area.
These freshwater algal blooms should not
be confused with those in saltwater
environments. Recent research suggests that
the predominant pollutant responsible for
algal blooms in saltwater estuaries and
coastal marine habitats is nitrogen.[3]

Phosphorus occurs most abundantly in


nature as part of the orthophosphate ion
(PO4 )3−, consisting of a P atom and 4
oxygen atoms. On land most phosphorus is
found in rocks and minerals. Phosphorus-
rich deposits have generally formed in the
ocean or from guano, and over time,
geologic processes bring ocean sediments to
land. Weathering of rocks and minerals Phosphorus cycle on land
release phosphorus in a soluble form where
it is taken up by plants, and it is transformed
into organic compounds. The plants may
then be consumed by herbivores and the
phosphorus is either incorporated into their
tissues or excreted. After death, the animal
or plant decays, and phosphorus is returned
to the soil where a large part of the
phosphorus is transformed into insoluble
compounds. Runoff may carry a small part
The aquatic phosphorus cycle
of the phosphorus back to the ocean.
Generally with time (thousands of years)
soils become deficient in phosphorus
leading to ecosystem retrogression.[4]

Major pools in aquatic systems

There are four major pools of phosphorus in freshwater ecosystems: dissolved inorganic phosphorus (DIP),
dissolved organic phosphorus (DOP), particulate inorganic phosphorus (PIP) and particulate organic
phosphorus (POP). Dissolved material is defined as substances that pass through a 0.45 μm filter.[5] DIP
consists mainly of orthophosphate (PO4 3-) and polyphosphate, while DOP consists of DNA and
phosphoproteins. Particulate matter are the substances that get caught on a 0.45 μm filter and do not pass
through. POP consists of both living and dead organisms, while PIP mainly consists of hydroxyapatite,
Ca5 (PO4 )3 OH .[5] Inorganic phosphorus comes in the form of readily soluble orthophosphate. Particulate
organic phosphorus occurs in suspension in living and dead protoplasm and is insoluble. Dissolved organic
phosphorus is derived from the particulate organic phosphorus by excretion and decomposition and is
soluble.

Biological function

The primary biological importance of phosphates is as a component of nucleotides, which serve as energy
storage within cells (ATP) or when linked together, form the nucleic acids DNA and RNA. The double
helix of our DNA is only possible because of the phosphate ester bridge that binds the helix. Besides
making biomolecules, phosphorus is also found in bone and the enamel of mammalian teeth, whose
strength is derived from calcium phosphate in the form of hydroxyapatite. It is also found in the exoskeleton
of insects, and phospholipids (found in all biological membranes).[6] It also functions as a buffering agent in
maintaining acid base homeostasis in the human body.[7]

Phosphorus cycling

Phosphates move quickly through plants and animals; however, the processes that move them through the
soil or ocean are very slow, making the phosphorus cycle overall one of the slowest biogeochemical
cycles.[2][8]

The global phosphorus cycle includes four major processes:

(i) tectonic uplift and exposure of phosphorus-bearing rocks such as apatite to surface
weathering;[9]
(ii) physical erosion, and chemical and biological weathering of phosphorus-bearing rocks
to provide dissolved and particulate phosphorus to soils,[10] lakes and rivers;
(iii) riverine and subsurface transportation of phosphorus to various lakes and run-off to the
ocean;
(iv) sedimentation of particulate phosphorus (e.g., phosphorus associated with organic
matter and oxide/carbonate minerals) and eventually burial in marine sediments (this
process can also occur in lakes and rivers).[11]

In terrestrial systems, bioavailable P (‘reactive P’) mainly comes from weathering of phosphorus-containing
rocks. The most abundant primary phosphorus-mineral in the crust is apatite, which can be dissolved by
natural acids generated by soil microbes and fungi, or by other chemical weathering reactions and physical
erosion.[12] The dissolved phosphorus is bioavailable to terrestrial organisms and plants and is returned to
the soil after their decay. Phosphorus retention by soil minerals (e.g., adsorption onto iron and aluminum
oxyhydroxides in acidic soils and precipitation onto calcite in neutral-to-calcareous soils) is usually viewed
as the most important process in controlling terrestrial P-bioavailability in the mineral soil.[13] This process
can lead to the low level of dissolved phosphorus concentrations in soil solution. Various physiological
strategies are used by plants and microorganisms for obtaining phosphorus from this low level of
phosphorus concentration.[14]

Soil phosphorus is usually transported to rivers and lakes and can then either be buried in lake sediments or
transported to the ocean via river runoff. Atmospheric phosphorus deposition is another important marine
phosphorus source to the ocean.[15] In surface seawater, dissolved inorganic phosphorus, mainly
orthophosphate (PO4 3-), is assimilated by phytoplankton and transformed into organic phosphorus
compounds.[11][15] Phytoplankton cell lysis releases cellular dissolved inorganic and organic phosphorus to
the surrounding environment. Some of the organic phosphorus compounds can be hydrolyzed by enzymes
synthesized by bacteria and phytoplankton and subsequently assimilated.[15] The vast majority of
phosphorus is remineralized within the water column, and approximately 1% of associated phosphorus
carried to the deep sea by the falling particles is removed from the ocean reservoir by burial in
sediments.[15] A series of diagenetic processes act to enrich sediment pore water phosphorus
concentrations, resulting in an appreciable benthic return flux of phosphorus to overlying bottom waters.
These processes include

(i) microbial respiration of organic matter in sediments,


(ii) microbial reduction and dissolution of iron and manganese (oxyhydr)oxides with
subsequent release of associated phosphorus, which connects the phosphorus cycle to
the iron cycle,[16] and
(iii) abiotic reduction of iron (oxyhydr)oxides by hydrogen sulfide and liberation of iron-
associated phosphorus.[11]

Additionally,

(iv) phosphate associated with calcium carbonate and


(v) transformation of iron oxide-bound phosphorus to vivianite play critical roles in
phosphorus burial in marine sediments.[17][18]

These processes are similar to phosphorus cycling in lakes and rivers.

Although orthophosphate (PO4 3-), the dominant inorganic P species in nature, is oxidation state (P5+),
certain microorganisms can use phosphonate and phosphite (P3+ oxidation state) as a P source by oxidizing
it to orthophosphate.[19] Recently, rapid production and release of reduced phosphorus compounds has
provided new clues about the role of reduced P as a missing link in oceanic phosphorus.[20]

Phosphatic minerals

The availability of phosphorus in an ecosystem is restricted by the rate of release of this element during
weathering. The release of phosphorus from apatite dissolution is a key control on ecosystem productivity.
The primary mineral with significant phosphorus content, apatite [Ca5 (PO4 )3 OH] undergoes
carbonation.[2][21]

Little of this released phosphorus is taken up by biota (organic form), whereas a larger proportion reacts
with other soil minerals. This leads to precipitation into unavailable forms in the later stage of weathering
and soil development. Available phosphorus is found in a biogeochemical cycle in the upper soil profile,
while phosphorus found at lower depths is primarily involved in geochemical reactions with secondary
minerals. Plant growth depends on the rapid root uptake of phosphorus released from dead organic matter
in the biochemical cycle. Phosphorus is limited in supply for plant growth. Phosphates move quickly
through plants and animals; however, the processes that move them through the soil or ocean are very slow,
making the phosphorus cycle overall one of the slowest biogeochemical cycles.[2][8]

Low-molecular-weight (LMW) organic acids are found in soils. They originate from the activities of
various microorganisms in soils or may be exuded from the roots of living plants. Several of those organic
acids are capable of forming stable organo-metal complexes with various metal ions found in soil solutions.
As a result, these processes may lead to the release of inorganic phosphorus associated with aluminum,
iron, and calcium in soil minerals. The production and release of oxalic acid by mycorrhizal fungi explain
their importance in maintaining and supplying phosphorus to plants.[2][22]

The availability of organic phosphorus to support microbial, plant and animal growth depends on the rate of
their degradation to generate free phosphate. There are various enzymes such as phosphatases, nucleases
and phytase involved for the degradation. Some of the abiotic pathways in the environment studied are
hydrolytic reactions and photolytic reactions. Enzymatic hydrolysis of organic phosphorus is an essential
step in the biogeochemical phosphorus cycle, including the phosphorus nutrition of plants and
microorganisms and the transfer of organic phosphorus from soil to bodies of water.[1] Many organisms
rely on the soil derived phosphorus for their phosphorus nutrition.

Eutrophication
Eutrophication is an enrichment of water by
nutrient that lead to structural changes to the
aquatic ecosystem such as algae bloom,
deoxygenation, reduction of fish species. The
primary source that contributes to the
eutrophication is considered as nitrogen and
phosphorus. When these two elements exceed the
capacity of the water body, eutrophication occurs.
Phosphorus that enters lakes will accumulate in
the sediments and the biosphere, it also can be
recycled from the sediments and the water
system.[23] Drainage water from agricultural land
also carries phosphorus and nitrogen.[24] Since a Nitrogen and phosphorus cycles in a wetland
large amount of phosphorus is in the soil contents,
so the overuse of fertilizers and over-enrichment
with nutrients will lead to increasing the amount of phosphorus concentration in agricultural runoff. When
eroded soil enters the lake, both phosphorus and the nitrogen in the soil contribute to eutrophication, and
erosion caused by deforestation which also results from uncontrolled planning and urbanization.[25]

Wetland

Wetlands are frequently applied to solve the issue of eutrophication. Nitrate is transformed in wetlands to
free nitrogen and discharged to the air. Phosphorus is adsorbed by wetland soils which are taken up by the
plants. Therefore, wetlands could help to reduce the concentration of nitrogen and phosphorus to remit and
solve the eutrophication. However, wetland soils can only hold a limited amount of phosphorus. To remove
phosphorus continually, it is necessary to add more new soils within the wetland from remnant plant stems,
leaves, root debris, and undecomposable parts of dead algae, bacteria, fungi, and invertebrates.[24]

Human influences
Nutrients are important to the growth and survival
of living organisms, and hence, are essential for
development and maintenance of healthy
ecosystems. Humans have greatly influenced the
phosphorus cycle by mining phosphorus,
converting it to fertilizer, and by shipping fertilizer
and products around the globe. Transporting
phosphorus in food from farms to cities has made
a major change in the global Phosphorus cycle.
However, excessive amounts of nutrients,
particularly phosphorus and nitrogen, are
detrimental to aquatic ecosystems. Waters are
enriched in phosphorus from farms' run-off, and
from effluent that is inadequately treated before it Phosphorus fertilizer application
is discharged to waters. The input of P in
agricultural runoff can accelerate the
eutrophication of P-sensitive surface waters. [26] Natural eutrophication is a process by which lakes
gradually age and become more productive and may take thousands of years to progress. Cultural or
anthropogenic eutrophication, however, is water pollution caused by excessive plant nutrients; this results in
excessive growth in the algal population; when this algae dies its putrefaction depletes the water of oxygen.
Such eutrophication may also give rise to toxic
algal bloom. Both these effects cause animal and
plant death rates to increase as the plants take in
poisonous water while the animals drink the
poisoned water. Surface and subsurface runoff
and erosion from high-phosphorus soils may be
major contributing factors to this fresh water
eutrophication. The processes controlling soil
Phosphorus release to surface runoff and to
subsurface flow are a complex interaction
between the type of phosphorus input, soil type
and management, and transport processes
depending on hydrological conditions.[27][28]

Repeated application of liquid hog manure in Phosphorus in manure production


excess to crop needs can have detrimental effects
on soil phosphorus status. Also, application of
biosolids may increase available phosphorus in soil.[29] In poorly drained soils or in areas where snowmelt
can cause periodic waterlogging, reducing conditions can be attained in 7–10 days. This causes a sharp
increase in phosphorus concentration in solution and phosphorus can be leached. In addition, reduction of
the soil causes a shift in phosphorus from resilient to more labile forms. This could eventually increase the
potential for phosphorus loss. This is of particular concern for the environmentally sound management of
such areas, where disposal of agricultural wastes has already become a problem. It is suggested that the
water regime of soils that are to be used for organic wastes disposal is taken into account in the preparation
of waste management regulations.[30]

Human interference in the phosphorus cycle occurs by overuse or careless use of phosphorus fertilizers.
This results in increased amounts of phosphorus as pollutants in bodies of water resulting in eutrophication.
Eutrophication devastates water ecosystems by inducing anoxic conditions.[25]

See also
Peak phosphorus
Planetary boundaries
Oceanic carbon cycle

References
1. Turner, B.L.; Frossard, E.; Baldwin, D.S. (2005). Organic phosphorus in the environment.
CABI Publishing. ISBN 978-0-85199-822-0.
2. Schlesinger, W.H. (1991). Biogeochemistry: An analysis of global change.
3. "Eutrophication" (https://web.archive.org/web/20140416123353/https://www.soils.org/discov
er-soils/soils-in-the-city/green-infrastructure/important-terms/eutrophication). www.soils.org.
Soil Science Society of America. Archived from the original (https://www.soils.org/discover-s
oils/soils-in-the-city/green-infrastructure/important-terms/eutrophication) on 2014-04-16.
Retrieved 2014-04-14.
4. Peltzer, D.A.; Wardle, D.A.; Allison, V.J.; Baisden, W.T.; Bardgett, R.D.; Chadwick, O.A.; et al.
(November 2010). "Understanding ecosystem retrogression". Ecological Monographs. 80
(4): 509–529. doi:10.1890/09-1552.1 (https://doi.org/10.1890%2F09-1552.1).
5. Wetzel, R.G. (2001). Limnology: Lake and river ecosystems. San Diego, CA: Academic
Press.
6. "Phosphorus Cycle" (http://www.enviroliteracy.org/article.php/480.html). enviroliteracy.org.
The Environmental Literacy Council. Archived (https://web.archive.org/web/2006110812161
2/http://www.enviroliteracy.org/article.php/480.html) from the original on 2006-11-08.
7. Voet, D.; Voet, J.G. (2003). Biochemistry. pp. 607–608.
8. Oelkers, E.H.; Valsami-Jones, E.; Roncal-Herrero, T. (February 2008). "Phosphate mineral
reactivity: From global cycles to sustainable development". Mineralogical Magazine. 72 (1):
337–40. Bibcode:2008MinM...72..337O (https://ui.adsabs.harvard.edu/abs/2008MinM...72..3
37O). doi:10.1180/minmag.2008.072.1.337 (https://doi.org/10.1180%2Fminmag.2008.072.1.
337). S2CID 97795738 (https://api.semanticscholar.org/CorpusID:97795738).
9. Buendía, C.; Kleidon, A.; Porporato, A. (2010-06-25). "The role of tectonic uplift, climate, and
vegetation in the long-term terrestrial phosphorous cycle" (https://bg.copernicus.org/articles/
7/2025/2010/). Biogeosciences. 7 (6): 2025–2038. Bibcode:2010BGeo....7.2025B (https://ui.
adsabs.harvard.edu/abs/2010BGeo....7.2025B). doi:10.5194/bg-7-2025-2010 (https://doi.org/
10.5194%2Fbg-7-2025-2010). ISSN 1726-4170 (https://www.worldcat.org/issn/1726-4170).
10. Adediran, Gbotemi A.; Tuyishime, J.R. Marius; Vantelon, Delphine; Klysubun, Wantana;
Gustafsson, Jon Petter (October 2020). "Phosphorus in 2D: Spatially resolved P speciation
in two Swedish forest soils as influenced by apatite weathering and podzolization" (https://d
oi.org/10.1016%2Fj.geoderma.2020.114550). Geoderma. 376: 114550.
Bibcode:2020Geode.376k4550A (https://ui.adsabs.harvard.edu/abs/2020Geode.376k4550
A). doi:10.1016/j.geoderma.2020.114550 (https://doi.org/10.1016%2Fj.geoderma.2020.1145
50). ISSN 0016-7061 (https://www.worldcat.org/issn/0016-7061).
11. Ruttenberg, K.C. (2014). "The global phosphorus cycle". Treatise on Geochemistry. Elsevier.
pp. 499–558. doi:10.1016/b978-0-08-095975-7.00813-5 (https://doi.org/10.1016%2Fb978-0-
08-095975-7.00813-5). ISBN 978-0-08-098300-4.
12. Slomp, C.P. (2011). "Phosphorus cycling in the estuarine and coastal zones". Treatise on
Estuarine and Coastal Science. Vol. 5. Elsevier. pp. 201–229. doi:10.1016/b978-0-12-
374711-2.00506-4 (https://doi.org/10.1016%2Fb978-0-12-374711-2.00506-4). ISBN 978-0-
08-087885-0.
13. Arai, Y.; Sparks, D.L. (2007). "Phosphate reaction dynamics in soils and soil components: A
multiscale approach". Advances in Agronomy. Elsevier. 94: 135–179. doi:10.1016/s0065-
2113(06)94003-6 (https://doi.org/10.1016%2Fs0065-2113%2806%2994003-6). ISBN 978-0-
12-374107-3.
14. Shen, J.; Yuan, L.; Zhang, J.; Li, H.; Bai, Z.; Chen, X.; Zhang, W.; Zhang, F (July 2011).
"Phosphorus dynamics: From soil to plant" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3
135930). Plant Physiology. 156 (3): 997–1005. doi:10.1104/pp.111.175232 (https://doi.org/1
0.1104%2Fpp.111.175232). PMC 3135930 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC
3135930). PMID 21571668 (https://pubmed.ncbi.nlm.nih.gov/21571668).
15. Paytan, A.; McLaughlin, K. (February 2007). "The oceanic phosphorus cycle". Chemical
Reviews. 107 (2): 563–576. doi:10.1021/cr0503613 (https://doi.org/10.1021%2Fcr0503613).
PMID 17256993 (https://pubmed.ncbi.nlm.nih.gov/17256993). S2CID 1872341 (https://api.se
manticscholar.org/CorpusID:1872341).
16. 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).
17. Kraal, P.; Dijkstra, N.; Behrends, T.; Slomp, C.P. (May 2017). "Phosphorus burial in
sediments of the sulfidic deep Black Sea: Key roles for adsorption by calcium carbonate and
apatite authigenesis". Geochimica et Cosmochimica Acta. 204: 140–158.
Bibcode:2017GeCoA.204..140K (https://ui.adsabs.harvard.edu/abs/2017GeCoA.204..140K).
doi:10.1016/j.gca.2017.01.042 (https://doi.org/10.1016%2Fj.gca.2017.01.042).
18. Defforey, D.; Paytan, A. (2018). "Phosphorus cycling in marine sediments: Advances and
challenges". Chemical Geology. 477: 1–11. Bibcode:2018ChGeo.477....1D (https://ui.adsab
s.harvard.edu/abs/2018ChGeo.477....1D). doi:10.1016/j.chemgeo.2017.12.002 (https://doi.or
g/10.1016%2Fj.chemgeo.2017.12.002).
19. Figueroa, I.A.; Coates, J.D. (2017). "Microbial phosphite oxidation and its potential role in
the global phosphorus and carbon cycles". Advances in Applied Microbiology. 98: 93–117.
doi:10.1016/bs.aambs.2016.09.004 (https://doi.org/10.1016%2Fbs.aambs.2016.09.004).
ISBN 978-0-12-812052-1. PMID 28189156 (https://pubmed.ncbi.nlm.nih.gov/28189156).
20. Van Mooy, B. A. S.; Krupke, A.; Dyhrman, S. T.; Fredricks, H. F.; Frischkorn, K. R.; Ossolinski,
J. E.; Repeta, D. J.; Rouco, M.; Seewald, J. D.; Sylva, S. P. (15 May 2015). "Major role of
planktonic phosphate reduction in the marine phosphorus redox cycle" (https://doi.org/10.11
26%2Fscience.aaa8181). Science. 348 (6236): 783–785. Bibcode:2015Sci...348..783V (http
s://ui.adsabs.harvard.edu/abs/2015Sci...348..783V). doi:10.1126/science.aaa8181 (https://d
oi.org/10.1126%2Fscience.aaa8181). PMID 25977548 (https://pubmed.ncbi.nlm.nih.gov/259
77548).
21. Filippelli GM (2002). "The Global Phosphorus Cycle". Reviews in Mineralogy and
Geochemistry. 48 (1): 391–425. Bibcode:2002RvMG...48..391F (https://ui.adsabs.harvard.ed
u/abs/2002RvMG...48..391F). doi:10.2138/rmg.2002.48.10 (https://doi.org/10.2138%2Frmg.2
002.48.10).
22. Harrold SA, Tabatabai MA (June 2006). "Release of inorganic phosphorus from soils by low‐
molecular‐weight organic acids". Communications in Soil Science and Plant Analysis. 37
(9–10): 1233–45. doi:10.1080/00103620600623558 (https://doi.org/10.1080%2F001036206
00623558). S2CID 84368363 (https://api.semanticscholar.org/CorpusID:84368363).
23. Carpenter SR (July 2005). "Eutrophication of aquatic ecosystems: bistability and soil
phosphorus" (https://www.pnas.org/content/pnas/102/29/10002.full.pdf) (PDF). Proceedings
of the National Academy of Sciences of the United States of America. 102 (29): 10002–5.
Bibcode:2005PNAS..10210002C (https://ui.adsabs.harvard.edu/abs/2005PNAS..10210002
C). doi:10.1073/pnas.0503959102 (https://doi.org/10.1073%2Fpnas.0503959102).
PMC 1177388 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC1177388). PMID 15972805
(https://pubmed.ncbi.nlm.nih.gov/15972805).
24. "Where Nutrients Come From and How They Cause Entrophication" (http://www.unep.or.jp/i
etc/publications/short_series/lakereservoirs-3/3.asp). Lakes and Reservoirs. United Nations
Environment Programme. 3.
25. Conley DJ, Paerl HW, Howarth RW, Boesch DF, Seitzinger SP, Havens KE, Lancelot C,
Likens GE (February 2009). "Ecology. Controlling eutrophication: nitrogen and phosphorus".
Science. 323 (5917): 1014–5. doi:10.1126/science.1167755 (https://doi.org/10.1126%2Fscie
nce.1167755). PMID 19229022 (https://pubmed.ncbi.nlm.nih.gov/19229022).
S2CID 28502866 (https://api.semanticscholar.org/CorpusID:28502866).
26. Daniel TC, Sharpley AN, Lemunyon JL (1998). "Agricultural phosphorus and eutrophication:
A symposium overview" (https://pubag.nal.usda.gov/pubag/downloadPDF.xhtml?id=20276&
content=PDF). Journal of Environmental Quality. 27 (2): 251–7.
doi:10.2134/jeq1998.00472425002700020002x (https://doi.org/10.2134%2Fjeq1998.00472
425002700020002x).
27. Branom JR, Sarkar D (March 2004). "Phosphorus bioavailability in sediments of a sludge-
disposal lake". Environmental Geosciences. 11 (1): 42–52. doi:10.1306/eg.10200303021 (ht
tps://doi.org/10.1306%2Feg.10200303021).
28. Schelde K, de Jonge LW, Kjaergaard C, Laegdsmand M, Rubæk GH (January 2006).
"Effects of manure application and plowing on transport of colloids and phosphorus to tile
drains". Vadose Zone Journal. 5 (1): 445–58. doi:10.2136/vzj2005.0051 (https://doi.org/10.2
136%2Fvzj2005.0051). S2CID 140621985 (https://api.semanticscholar.org/CorpusID:14062
1985).
29. Hosseinpur A, Pashamokhtari H (June 2013). "The effects of incubation on phosphorus
desorption properties, phosphorus availability, and salinity of biosolids-amended soils".
Environmental Earth Sciences. 69 (3): 899–908. doi:10.1007/s12665-012-1975-6 (https://do
i.org/10.1007%2Fs12665-012-1975-6). S2CID 140537340 (https://api.semanticscholar.org/C
orpusID:140537340).
30. Ajmone-Marsan F, Côté D, Simard RR (April 2006). "Phosphorus transformations under
reduction in long-term manured soils". Plant and Soil. 282 (1–2): 239–50.
doi:10.1007/s11104-005-5929-6 (https://doi.org/10.1007%2Fs11104-005-5929-6).
S2CID 23704883 (https://api.semanticscholar.org/CorpusID:23704883).

External links
Holding, B.V. (2006). "Matter Cycles" (http://www.lenntech.com/phosphorus-cycle.htm).
Lenntech. Water treatment & air purification.
"Phosphorus Cycle" (http://www.enviroliteracy.org/article.php/480.html). Environmental
Literacy Council. 10 July 2023.
"section 5.6 Phosphorus" (http://www.epa.gov/volunteer/stream/vms56.html). Monitoring and
assessing water quality. U.S. Environmental Protection Agency.
Miller, Kenneth R.; Levine, Joseph (2001). Biology (https://web.archive.org/web/2008081221
1636/http://www.phschool.com/atschool/biology/Dragonfly/Student_Area/PHB_S_BK_inde
x.html). Prentice Hall. Archived from the original (http://www.phschool.com/atschool/biology/
Dragonfly/Student_Area/PHB_S_BK_index.html) on 2008-08-12.
Corbin, Katie. "The Phosphorus Cycle" (https://web.archive.org/web/20080914101357/http://
filebox.vt.edu/users/chagedor/biol_4684/Cycles/Pcycle.html). Biogeochemical Cycles – Soil
Microbiology. Virginia Polytechnic Institute and State University. Archived from the original (h
ttp://filebox.vt.edu/users/chagedor/biol_4684/Cycles/Pcycle.html) on 2008-09-14.

Retrieved from "https://en.wikipedia.org/w/index.php?title=Phosphorus_cycle&oldid=1184308678"

You might also like