Phosphorus Bike

The phosphorus bicycle includes mineralization and immobilization reactions mediated by the microbial biomass pool, and the solubilization of inorganic phosphorus-containing minerals by organic acids, inorganic acids, and chelating agents produced by soil organisms.

From: Principles and Applications of Soil Microbiology (Tertiary Edition) , 2021

Marine Biogeochemistry

K.C. Ruttenberg , in Encyclopedia of Ocean Sciences (Third Edition), 2019

Introduction

The global phosphorus cycle has four major components: (i) tectonic uplift and exposure of phosphorus-begetting rocks to the forces of weathering; (two) physical erosion and chemical weathering of rocks producing soils and providing dissolved and particulate phosphorus to rivers; (three) riverine transport of phosphorus to lakes and the ocean; and (four) sedimentation of phosphorus associated with organic and mineral matter and burial in sediments ( Fig. 1). The cycle begins anew with uplift of sediments into the weathering authorities.

Fig. 1

Fig. ane. Cartoon illustrating the major reservoirs and fluxes of phosphorus described in the text and summarized in Tables ane and 2. The oceanic photic zone, idealized in the cartoon, is typically thinner in littoral environments attributable to turbidity from continental terrigenous input, and deepens as the water column clarifies with altitude away from the continental margins. The distribution of phosphorus among different chemical/mineral forms in marine sediments is given in the pie diagrams, where the abbreviations used are: Porg, organic phosphorus; PIron, iron-bound phosphorus; Pdetr, detrital apatite; Pauth, authigenic/biogenic apatite. The Porg, PFe, and Pauth reservoirs stand for potentially reactive phosphorus pools (come across text and Tables ii and 3 for give-and-take), whereas the Pdetr puddle reflects mainly detrital apatite weathered off the continents and passively deposited in marine sediments (note that Pdetr is non an important sedimentary phosphorus component in abyssal sediments, far from continents). Continental margin phosphorus speciation data were compiled from Louchouarn, P., Lucotte, Chiliad., Duchemin, E., and de Vernal, A. (1997). Early diagenetic processes in recent sediments of the Gulf of St-Lawrence: Phosphorus, carbon and fe burying rates. Marine Geology 139(1/four), 181–200; Ruttenberg, K.C. and Berner, R.A. (1993). Authigenic apatite formation and burying in sediments from non-upwelling continental margin environments. Geochimica et Cosmochimica Acta 57, 991–1007. Abyssal sediment phosphorus speciation information were compiled from Filippelli, G.M. and Delaney, M.L. (1996). Phosphorus geochemistry of equatorial Pacific sediments. Geochimica et Cosmochimica Acta lx, 1479–1495; Ruttenberg, Grand.C. (1990). Diagenesis and burying of phosphorus in marine sediments: implications for the marine phosphorus budget. PhD thesis, Yale Academy. The global phosphorus cycle cartoon is from Ruttenberg (2002). The global phosphorus bicycle. In: The Encyclopedia of Global Modify. Oxford University Press. (2014), with permission. Note that newly identified authigenic phosphorus phases (vivianite, aluminophosphates) are not represented in this cartoon (meet text). Vertical h2o column distributions of phosphate typically observed in the iii ocean basins are shown in the panel to the right of the global phosphorus bicycle drawing, and are from Sverdrup, H.V., Johnson, M.Westward. and Fleming, R.H. (1942). The oceans, their physics, chemistry and general biological science. New York: Prentice Hall, copyright 1942 Prentice Hall; used with permission.

Phosphorus is an essential food for all life forms. It is a key component of primal biochemicals, including genetic material (DNA, RNA), free energy transferal molecules (e.m., adenosine triphosphate: ATP), and compounds that provide structural support to organisms in the form of membranes (phospholipids) and bone (the biomineral hydroxyapatite). Photosynthetic organisms at the base of the food web in both terrestrial and aquatic ecosystems require dissolved phosphorus, forth with carbon and other essential nutrients, to build their tissues using energy from the sun. Biological productivity is contingent upon the availability of phosphorus to these organisms.

Phosphorus locked upwardly in bedrock, soils, and sediments is not directly available to organisms. Conversion of unavailable forms to bioavailable forms (principally dissolved orthophosphate: PO4 3   ), which tin can be directly assimilated, occurs through geochemical and biochemical reactions at diverse stages in the global phosphorus cycle. Production of biomass fueled by phosphorus bioavailability results in the deposition of organic matter in soil and sediments, where information technology acts as a source of fuel and nutrients to microbial communities. Microbial action in soils and sediments, in turn, strongly influences the concentration and chemical course of phosphorus incorporated into the geological tape.

This article begins with a brief overview of the diverse components of the global phosphorus cycle. Estimates of the mass of important phosphorus reservoirs, transport rates (fluxes) betwixt reservoirs, and residence times are given in Tables 1 and ii. Big uncertainties are associated with these estimates, reflecting the challenges associated with constraining global reservoir size and flux magnitudes. These uncertainties underscore the fact that many aspects of the global phosphorus cycle remain poorly understood. The 2d half of the article describes recent advances in our understanding of the oceanic phosphorus cycle. These include: (i) new insights into the role of phosphorus as a bio-limiting nutrient in the ocean, with focus on dissolved organic phosphorus (DOP); (ii) new observations of phosphorus redox chemistry in the body of water; (iii) application of cutting edge molecular techniques to probe the genetic underpinnings of microbial phosphorus cycling; (iv) novel insights into the limerick of particulate phosphorus, including authigenic phosphate mineral paragenesis (i.e., mode of formation); (five) the use of phosphate oxygen isotopes to reveal pathways of microbial transformations of phosphorus; (vi) reevaluation of the oceanic residence time of phosphorus; and (vii) rethinking the global phosphorus-wheel on geological timescales, with implications for atmospheric oxygen and phosphorus limitation of primary productivity in the body of water.

Table 1. Major reservoirs active in the global phosphorus bicycle and associated residence times

Reservoir # Reservoir clarification Reservoir size (mole P   ×   1012) References Residence time (years)
R1 Sediments (crustal rocks and soil >   60   cm deep and marine sediments 0.27   ×   x8–1.3   ×   10eight b, a   =   c   =   d 42–201   ×   10half dozen
R2 Land (≈   total soil <   60   cm deep: organic   +   inorganic) 3100–6450 b, a   =   c   =   d 425–2311
R3 Land biota 83.nine–96.8 b, a   =   c   =   d xiii–48
R4 Surface ocean, 0–300   m (full dissolved P) 87.iv a   =   c 2.46–4.39
R5 Deep sea, 300–3300   m (total dissolved P) 2810 a   =   c&d 1502
R6 Oceanic biota i.61–4.45 b&d, a   =   c&d 0.044–0.217 (16–78 days)
R7 Minable P 323–645 a   =   c, b&d 718–1654
R8 Atmospheric P 0.0009 b   =   c   =   d 0.009 (80   h)

(1) Ranges are reported for those reservoirs for which a consensus on a single all-time estimate of reservoir size does non exist. Maximum and minimum estimates found in a survey of the literature are reported. References cited earlier the comma refer to the first (everyman) gauge, those after the comma refer to the second (college) judge. References that give identical values are designated past an equality sign, references giving similar values are indicated past an ampersand. As indicated by the wide ranges reported for some reservoirs, all calculations of reservoir size have associated with them a big degree of uncertainty. Methods of calculation, underlying assumptions, and sources of error are given in the references cited.

(2) Residence times are calculated by dividing the concentration of phosphorus contained in a given reservoir by the sum of fluxes out of the reservoir. Where ranges are reported for reservoir size and flux, maximum and minimum residence time values are given; these ranges reverberate the uncertainties inherent in reservoir size and flux estimates. Fluxes used to summate residence times for each reservoir are every bit follows: R1 (F12), R2 (F23  +   F28  +   F24(d)+F24(p)), R3 (F32), R4 (F45  +   F46), R5 (F54), R6 (F64  +   F65), R7 (F72), R8 (F82  +   F84). Flux estimates are given in Table ii. The residence time of R5 is decreased to 1492 years by inclusion of the scavenged flux of deep-sea phosphate at hydrothermal MOR systems, by and large onto ferric oxide and oxyhydroxide phases, later on Wheat et al. (1996).

(3) Estimates for the partitioning of the oceanic reservoir between dissolved inorganic phosphorus and particulate phosphorus are given in references b and d as follows: (2581–2600)   ×   1012  mol dissolved inorganic phosphorus (b, d) and (20–21)   ×   ten12  mol particulate phosphorus (d, b).

(4) The residence times estimated for the minable phosphorus reservoir reflect estimates of current mining rates; if mining activity increases or diminishes, the residence time volition change accordingly.

References:

(a) Lerman, A., Mackenzie, F.T. and Garrels, R.Thousand. (1975). Geological Gild of America Memoir 142, 205.

(b) Richey, J.E. (1983). In: Bolin, B. and Cook, R.B. (eds) The Major Biogeochemical Cycles and Their Interactions, Telescopic 21, Chichester: Wiley, pp. 51–56.

(c) Jahnke, R.A. (1992). In: Butcher, S.S. et al. (eds) Global Geochemical Cycles, San Diego: Bookish Press, pp. 301–315. (Values are identical to Lehrman et al., except the inclusion of the atmospheric reservoir estimate taken from Graham, W.F. and Duce, R.A. (1979). Geochimica et Cosmochimica Acta 43, 1195.)

(d) Mackenzie, F.T., Ver, Fifty.M., Sabine, C., Lane, M. and Lerman, A. (1993). In: Wollast, R., Mackenzie, F.T. and Chou, L. (eds) Interactions of C, N, P and Southward Biogeochemical Cycles and Global Alter. NATO ASI Series 1, vol. 4, Berlin: Springer-Verlag, pp. one–61.

Table 2. Fluxes between the major phosphorus reservoirs

Flux# Description of flux Flux (moles P   ×   1012year  1) References and comments
Reservoir fluxes
F12 Rocks/sediments     soils (erosion/weathering, soil accumulation) 0.645 a   =   c&d
F21 Soils     rocks/sediments (deep burial, lithification) 0.301–0.603 d, a   =   c
F23 Soils     land biota ii.03–6.45 a   =   c, b&d
F32 Country biota     soils 2.03–6.45 a   =   c, b&d
F24(d) Soil     surface sea (river full dissolved P flux) 0.032–0.058 east, a   =   c; ca. >   50% of TDP is DOP (eastward)
F24(p) Soil     surface sea (river particulate P flux) 0.59–0.65 d, eastward; ca. 40% of RSPM-P is org. P (e); it is estimated that between 25% and 45% of RSPM-P is reactive once it enters the ocean (f)
F46 Surface body of water     oceanic biota 19.35–35 b, d; a   =   c   =   33.5, b reports upper limit of 32.iii; d reports lower limit of 28.2
F64 Oceanic biota     surface ocean xix.35–35 b, d; a&c   =   32.2, b reports upper limit of 32.iii, d reports lower limit of 28.2
F65 Oceanic biota     deep sea (particulate rain) one.13–1.35 d, a   =   c
F45 Surface body of water     deep sea (downwelling) 0.581 a   =   c
F54 Deep bounding main     surface ocean (upwelling) 1.87 a   =   c
F42 Surface ocean     state (fisheries) 0.01 d
F72 Minable P     land (soil) 0.39–0.45 a   =   c   =   d, b
F28 Land (soil)→   temper 0.14 b   =   c   =   d
F82 Temper     state (soil) 0.one b   =   c   =   d
F48 Surface sea     atmosphere 0.01 b   =   c   =   d
F84 Atmosphere     surface bounding main 0.02–0.05 c, b; d gives 0.04; ca. 30% of atmospheric aerosol-P is soluble (chiliad)
Subreservoir fluxes: marine sediments
sFms Marine sediment accumulation (full) 0.265–0.280 i, j; for higher judge (j), utilize of sediment P-concentration below the diagenesis zone implicitly accounts for P-loss via benthic remineralization flux and yields preanthropogenic cyberspace burial flux. For estimates of reactive-P burial see notation (j).
sFcs Continental margin body of water sediments     burying 0.150–0.223 j, i; values reported reflect total-P; reactive-P burial constitutes from xl%–75% of total-P (h). These values reflect preagricultural fluxes; modern value is estimated every bit 0.33 (d).
sFequally Abyssal (deep sea) sediments     burial 0.042–0.130 i, j; a   =   c gives a value of 0.055. It is estimated that 90%–100% of this flux is reactive-P (h). These values reverberate preagricultural fluxes; modern estimates range from 0.32 (d) to 0.419 (b).
sFMOR mid-ocean ridge systems     burial 0.036 l; value represents most recent estimate of dissolved inorganic phosphate removal to sedimented and unsedimented mid-ocean ridge systems, both due to scavenging from basement fluids during hydrothermally-driven circulation into the chaff (0.028), and to scavenging onto MOR plume particles that subsequently settle to the seafloor (0.008). 100% of this pool is considered reactive-P because its origin is dissolved P that is scavenged from seawater onto mineral surfaces.
sFcbf Coastal sediments     coastal waters (remineralization, benthic flux) 0.51–0.84 d, m; these values reflect preagricultural fluxes, modern value estimated is as ane.21 with uncertainties ±   40% (k)
sFabf Abyssal sediments     deep bounding main (remineralization, benthic flux) 0.41 one thousand; this value reflects preagricultural fluxes, modernistic value is estimated every bit 0.52, uncertainty ±   thirty% (k)

Notation 1: Reservoir Fluxes (F) represent the P-flux betwixt reservoirs #R1–R8 divers in Table 1. The Subreservoir fluxes (sF) refer to the flux of P into the marine sediment portion of Reservoir #1 via sediment burial, and the flux of diagenetically-mobilized P out of marine sediments via benthic return flux. These subfluxes have been calculated equally described in references h–chiliad. Note that the large magnitude of these subfluxes relative to those into and out of Reservoir #1 as a whole, and the short oceanic-P residence time they imply (Tables 1 and iii), highlight the dynamic nature of the marine phosphorus cycle.

Note 2: Ranges are reported where concensus on a single all-time estimate does not exist. References cited earlier the comma refer to the first (lowest) judge, those after the comma refer to the second (higher) estimate. References that requite identical values are designated by an equality sign, references giving similar values are indicated by an ampersand. Maximum and minimum estimates plant in a survey of the literature are reported. In some cases this range subsumes ranges reported in the main references. As indicated by the wide ranges reported, all flux calculations have associated with them a large degree of uncertainty. Methods of adding, underlying assumptions, and sources of error are given in the references cited.

a = Lerman, A., Mackenzie, F.T., and Garrels, R.G. (1975). Modeling of geochemical cycles: Phosphorus as an example. Geological Society of America Memoir 142, 205–217.

b = Richey, J.E. (1983). The phosphorus cycle. In: The Major Biogeochemical Cycles and Their Interactions. Bolin, B. and Cook, R.B. (eds.) Telescopic 21, John Wiley and Sons: Chichester, pp. 51–56.

c = Jahnke, R. A. (1992). The phosphorus bicycle. In: Global Geochemical Cycles. Butcher, South.S. et al. (eds.) Academic: San Diego, pp. 301–315. (Values are identical to those found in Lerman, et al. (1975). Except for atmospheric P fluxes taken from Graham, W.F. and Duce, R.A. (1979). Geochimica et Cosmochimica Acta 43, 1195–1208.)

d = Mackenzie, F.T., Ver, L.Thou., Sabine, C., Lane, G., and Lerman, A. (1993). C, N, P, South Global biogeochemical cycles and modeling of global change. In: Interactions of C, N, P and S Biogeochemical Cycles and Global Modify. Wollast, R., Mackenzie, F.T., and Chou, L. (eds.) NATO ASI Series1, vol. four, pp. 1–61. Springer-Verlag: Berlin.

east = Meybeck, Grand. (1982). Carbon, nitrogen, and phosphorus transport by earth rivers. American Journal of Scientific discipline 282(4), 401–450.

f = The range of riverine suspended particulate matter that may be solubilized one time it enters the marine realm (e.g., then-called "reactive-P") is derived from three sources. Colman, A.S. and Holland, H.D. (2000). The global diagenetic flux of phosphorus from marine sediments to the oceans: Redox sensitivity and the control of atmospheric oxygen levels. In: Marine Authigenesis: From Global to Microbial. Glenn, C.R., Prévôt-Lucas, Fifty. and Lucas, J. (eds.) SEPM Spec. Pub. #66. pp. 53–75 judge that 45% may be reactive, based on RSPM-P compositional information from a number of rivers and estimated burial efficiency of this material in marine sediments. Berner, R.A. and Rao, J.-Fifty. (1994). Phosphorus in sediments of the Amazon River and estuary: Implications for the global flux of phosphorus to the ocean. Geochimica et Cosmochimica Acta 58, 2333–2339 and Ruttenberg, K.C. and Canfield, D.East. (1994). Chemical distribution of phosphorus in suspended particulate matter from 12 North American rivers: Bear witness for bioavailability of particulate P, EOS. Trans. Amer. Geophys. Union. 75: 110. These sources estimate that 35% and 31% of RSPM-P is released upon inbound the ocean, based on comparing of RSPM-P and adjacent deltaic surface sediment P in the Amazon and Mississippi systems, respectively. This range is besides consequent with a more recent judge from the Mississippi River (Sutula, 1000., Bianchi, T. S. and McKee, B. A. (2004). Outcome of seasonal sediment storage in the lower Mississippi River on the flux of reactive particulate phosphorus to the Gulf of United mexican states. Limnology and Oceanography 49 (6), 2223–2235. doi: x.4319/lo.2004.49.half dozen.2223). Lower estimates have been published (8%: Ramirez, A.J. and Rose, A.W. (1992). Analytical geochemistry of organic phosphorus and its correlation with organic carbon in marine and fluvial sediments and soils. American Periodical of Science 292, 421–454; 18%: Froelich, P.N. (1988). Kinetic command of dissolved phosphate in natural rivers and estuaries. Limnology & Oceanography 33, 649–558; 18%: Compton, J., Mallinson, D., Glenn, C.R., Filippelli, Grand, Follmi, K., Shields, G. and Zanin, Y. (2000). Variations in the global phosphorus cycle, in: Marine Authigenesis: From Global to Microbial. Glenn, C.R., Prévôt-Lucas, Fifty. and Lucas, J. (eds.) SEPM Spec. Pub. #66. pp. 21–33). College estimates take also been published (69%: Howarth, R.Westward., Jensen, H.S., Marine, R. and Postma, H. (1995). Transport to and processing of P in about-shore and oceanic waters. In: Phosphorus in the Global Environment. Tiessen, H. (ed.) SCOPE 54. J. Wiley and Sons: Chichester, pp. 323–345). Howarth et al. (1995) also estimate the total flux of riverine particulate P to the oceans at 0.23   ×   x12 moles P   year  1, an gauge likely as well low because information technology uses the suspended sediment flux from Milliman, J.D. and Meade, R.H. (1983). Earth-wide commitment of river sediment to the oceans. The Journal of Geology 91, 1–21, which does not include the high sediment flux rivers from tropical mountainous terranes (Milliman, J. D. and Syvitski, J.P.Chiliad. (1992). Geomorphic/tectonic control of sediment belch to the ocean: The importance of minor mountainous rivers. The Journal of Geology 100, 525–544).

thousand = Duce, R.A., Liss, P.Southward., Merrill, J.T., Atlans, E.L., Buat-Menard, P., Hicks, B.B., Miller, J.M., Prospero, J.M., Arimoto, R. Church, T.M., Ellis, Due west., Galloway, J.N., Hansen, L., Jickells, T.D., Knap, A.H., Reinhardt, K.H., Schneider, B., Soudine, A., Tokos, J.J., Tsunogai, S., Wollast, R., and Zhou, M. (1991). The atmospheric input of trace species to the world ocean. Global Biogeochemical Cycles v, 193–259.

h = Ruttenberg, K.C. (1993). Reassessment of the oceanic residence fourth dimension of phosphorus. Chemical Geology 107, 405–409.

i = Howarth, R.W., Jensen, H.Southward., Marino, R. and Postma, H. (1995). Ship to and processing of P in nigh-shore and oceanic waters. In: Phosphorus in the Global Environment. Tiessen, H. (ed.) SCOPE 54. John Wiley and Sons: Chichester, pp. 323–345.

j = P-burial flux estimates equally reported in Ruttenberg (1993). Reassessment of the oceanic residence time of phosphorus. Chemical Geology 107, 405–409, modified using preagricultural sediment fluxes updated by Colman, A.S. and Kingdom of the netherlands, H.D. (2000). The global diagenetic flux of phosphorus from marine sediments to the oceans: Redox sensitivity and the control of atmospheric oxygen levels. In: Marine Authigenesis: From Global to Microbial. Glenn, C.R., Prévôt-Lucas, L. and Lucas, J. (eds.) SEPM Spec. Pub. #66. pp. 53–75. Using these full P burial fluxes and the ranges of probable reactive P given in the table, the best estimate for reactive P burial flux in the oceans lies betwixt 0.177 and 0.242   ×   1012 moles P twelvemonth  1. Other estimates of whole-ocean reactive P burial fluxes range from, at the low end: 0.032–0.081   ×   1012 moles P twelvemonth  i (Compton, J., Mallinson, D., Glenn, C.R., Filippelli, G, Follmi, K., Shields, G. and Zanin, Y. (2000). Variations in the global phosphorus bike, In: Marine Authigenesis: From Global to Microbial. Glenn, C.R., Prévôt-Lucas, L. and Lucas, J. (eds.) SEPM Spec. Pub. #66. pp. 21–33), and 0.09   ×   ten12 moles P twelvemonth  1 (Wheat, C.One thousand., Feely, R.A. and Mottl, M.J. (1996). Phosphate removal past oceanic hydrothermal processes: An update of the phosphorus upkeep in the oceans. Geochimica et Cosmochimica Acta threescore, 3593–3608; but see updated estimates in Wheat et al., 2003.); to values more comparable to those derived from the table above (0.21   ×   1012 moles P year  1: Filippelli, K.M. and Delaney, M.L. (1996). Phosphorus geochemistry of equatorial Pacific sediments. Geochimica et Cosmochimica Acta lx, 1479–1495; 0.12–0.22   ×   ten12 moles P year  1: Wheat et al. (2003). Oceanic phosphorus imbalance: Magnitude of the mid-sea ridge flank hydrothermal sink. Geophysical Research Letters 30(17). https://doi.org/10.1029/2003GL0017318.). Recently, Wallmann (2010) estimated essentially higher oceanic P burying fluxes (0.46 and 1.211012 year 1) than those summarized in Tabular array 2. The lower stop of his range is based upon full-P burial fluxes from a contempo compilation of sediment P concentrations, and may overestimate reactive-P burying flux. The college end is derived from model calculations of P regenerated from water cavalcade particulates and benthic sediments. If Wallman's (2010) P-burial estimates are borne out by further scrutiny, the relatively contempo revision of the P residence fourth dimension downwards to values in the 10–17 ky range (e.g., Ruttenberg, K.C. (2004). The Global Phosphorus Wheel. In: Treatise on Geochemistry. H.D. Holland and Thou.Yard. Turekian, eds. Vol. viii (Biogeochemistry: Due west. H. Schlesinger, volume editor), Chapter 13, Elsevier Science, pp. 585–643), will drop to even lower values (ii–6 ky: Wallmann, 1000. (2010). Phosphorus imbalance in the global body of water? Global Biogeochemical Cycles 24, https://doi.org/10.1029/2009GB003643.)

k = Colman, A.Due south. and Holland, H.D. (2000). The global diagenetic flux of phosphorus from marine sediments to the oceans: Redox sensitivity and the control of atmospheric oxygen levels. In: Marine Authigenesis: From Global to Microbial. Glenn, C.R., Prévôt-Lucas, L. and Lucas, J. eds. SEPM Spec. Pub. #66. pp. 53–75.

l = Wheat, C.Thousand., McManus, J., Mottl, M.J., and Giambalvo, E. (2003). Oceanic phosphorus imbalance: Magnitude of the mid-bounding main ridge flank hydrothermal sink. Geophysical Research Letters 30(17). https://doi.org/10.1029/2003GL0017318.

Read full chapter

URL:

https://www.sciencedirect.com/science/article/pii/B9780124095489108073

Nutrient recovery from municipal waste stream: status and prospects

Vaibhav Srivastava , ... Rajeev Pratap Singh , in Urban Ecology, 2020

2.ane.2 Phosphorus selective

The anthropogenic phosphorus bike is represented by greater loss through diffusion, which results in transfer of P from lithosphere to hydrosphere causing eutrophication. Around 27 million tonne per year of P is applied to the agricultural fields, and only x%–11% is consumed past man beings and rest is dissipated ( van Enk and van der Vee, 2011). Therefore, it is imperative to recover P in an efficient and more sustainable way. 1 of the promising approaches to P accumulation is by employing phosphate accumulating organisms (PAOs), which uptake and stock high amount of P than typical ones. Thereafter, enhanced biological phosphorus removal (EBPR) select PAOs using alternating anaerobic/aerobic conditions during wastewater treatment. Subsequently, EBPR sludge could have five times greater P in comparison with typical activated sludge treatments (Rittmann and McCarty, 2001), and P can exist recovered from EMBR sludge either nether anaerobic atmospheric condition or through physicochemical processing. Furthermore, different microbes such as cyanobacteria, nonsulfur bacteria and algae take potential to bioaccumulate polyphosphates or proteins (Mehta et al., 2015). Phosphate is the nigh recoverable class of P when recovered from aqueous stream as solid precipitates. Nonetheless, phosphorus extraction as struvite (MgNH4PO4·6HiiO) and hydroxylapatite is currently favored as compared with mineral fertilizers (Johnston and Richards, 2004) (Fig. 15.2). Although, struvite contains considerable amount of Due north and Mg, it is used for phosphate recovery and an alternative to rock phosphate equally the recovering charge per unit of phosphate is very loftier through struvite crystallization (Rahman et al., 2014). Wastewater containing loftier amount of P and North is a adept source of struvite. Besides hydroxylapatite used equally fertilizer, it can be used as substitution for chemical production of rock phosphate (Cornel and Schaum, 2009). Another innovative approach for phosphorus recovery from wastewater is utilizing algal pond or wetlands. Many researchers state that the preconsideration for P removal from pond, or constructed wetland is effective biomass growth (Korner et al., 2003). Through this process, high uptake of P content up to 2.nine% can be accomplished (Chaiprapat et al., 2005). Another promising applied science for phosphate recovery is hybrid anion exchange (HAIX) resin, although research on this approach is in infancy. The HAIX resins accept higher selectivity for phosphate equally compared with other competing anions such equally sulfate (Blaney et al., 2007; Pan et al., 2009; Sengupta and Pandit, 2011). Thus, these underflow waste material stream had huge prospective to recover high phosphorus as fertilizer supplements (Sutton et al., 2011).

Read total chapter

URL:

https://www.sciencedirect.com/scientific discipline/article/pii/B978012820730700015X

Geologic History and Energy

C. Ludwig , W. Steffen , in Encyclopedia of the Anthropocene, 2018

Phosphorus

Human activities accept contradistinct the phosphorus cycle to a large degree, also to produce fertilizer to support an expanding homo population, with near of the increasing mining and application of phosphorus occurring subsequently the middle of the 20th century. Preindustrial weathering processes produced 15–twenty  Tg   year  1 of mobilized phosphorus, but the human mining of phosphorus is now estimated to be 23.5   Tg   yr  1, more than than doubling the amount of phosphorus flowing through the environment compared to the baseline of the Holocene. This additional phosphorus has significant impacts on the functioning of the Earth System, primarily eutrophication of freshwater lakes but also potentially contributing to the germination of anoxic zones in the ocean (Carpenter and Bennett, 2011).

Read full affiliate

URL:

https://www.sciencedirect.com/science/article/pii/B9780128096659099407

Estuarine and Coastal Ecosystem Modelling

Grand. Arhonditsis , ... Due west. Zhang , in Treatise on Estuarine and Coastal Science, 2011

9.ten.3.i.vi Phosphorus cycle

2 state variables of the phosphorus cycle are considered in the model: phosphate (PO 4) and OP ( Effigy 5 ). The phosphate equation considers the phytoplankton uptake, the proportion of phytoplankton and zooplankton mortality/higher predation that is directly supplied into the system in inorganic class, the bacteria-mediated mineralization of OP, and the net diffusive fluxes betwixt epilimnion and hypolimnion. Nosotros also accounted for the phosphorus precipitation to sediment due to the fe loadings from the two steel mills, based on an empirical equation originally implemented to correct for the observed Hamilton Harbour phosphorus concentrations (Hamilton Harbour Technical Team – Water Quality, 2007). The OP equation besides considers the corporeality of OP that is redistributed through phytoplankton and zooplankton basal metabolism. A fraction of OP settles to the sediment and another fraction is mineralized to phosphate through a get-go-order reaction. We as well consider external phosphorus loads to the arrangement and losses via the exchanges with Lake Ontario.

Read full chapter

URL:

https://www.sciencedirect.com/scientific discipline/article/pii/B9780123747112009104

Nonpersistent Inorganic Pollution

Ricardo Beiras , in Marine Pollution, 2018

3.one Natural and Anthropogenic Sources of Nitrogen and Phosphorus in the Water

Primary production is express by N and P

Eq. (2.1) describes primary production in a quite simplistic style, taking into business relationship the cycling of carbon and oxygen remainder simply. In fact, carbon is very arable in the water, either equally dissolved CO2 or as bicarbonate and carbonate ions, and this element does not limit algal growth. Other essential macronutrients such as Grand and S are as well rather arable in the seawater compared to the biological requirements. The nutrients essential for all forms of life more than ofttimes limiting primary production are N and P. The C:Northward:P stoichiometric ratios of the living organisms are fairly abiding, and termed Redfield ratios. For both phytoplankton and zooplankton organic matter 106:xvi:i is applicative. This means that if the N:P atomic ratio of inorganic nutrients dissolved in the water exceed sixteen the algal growth may exist limited by phosphorus, and otherwise by nitrogen. In the ocean, limitation of phytoplankton biomass by iron has also been demonstrated, and the growth of sure taxa such every bit diatoms only not the overall production, may be express past silicate. It is generally causeless that Northward is normally less harmful than P in inland waters where main production is often express by the latter, simply information technology may cause problems when discharged to the seas.

Mineralization of organic matter past heterotrophic decomposers produces ammonium that in aerobic environments is oxidized to nitrites and nitrates by the chemotrophic nitrifying leaner. Nitrate is the thermodynamically stable form of nitrogen in the well-oxygenated seawater. Except for a few species of cyanobacteria capable of N2 fixation, nitrates are the most abundant source of nitrogen for primary producers in the sea. Other inorganic salts of Due north also available for phytoplankton are nitrites and ammonium, the latter the preferred source since cellular nitrogen is by and large in the same oxidation state (III) in the amine (-NHii) groups of proteins. i Some forms of organic nitrogen such every bit urea and amino acids are also utilizable by phytoplankton to satisfy their nitrogen demands.

In anaerobic environments the denitrifying bacteria can mediate the reduction of nitrate to nitrogen gas, which in sediments displace the oxygen and contributes to further anoxia.

Compared to nitrogen, the phosphorus cycle is much simpler, with a single inorganic course, the salts of the phosphoric acid, normally orthophosphate ions, HPO 4 2 .

The vertical distribution of N and P shows depleted values in the photic zone and enrichment in deep waters

Mineral dissolved nutrients are depleted from the photic zone by master producers and regenerated in deep waters by decomposers. As a result the vertical distribution of both nitrates and phosphates show a surface minimum, sharply increases with depth during the first 100–500   k, and information technology is approximately steady hence on. This typical food-similar behavior reflecting a dominant role for biological cycling is common to the distributions of several trace metals in the ocean (see Section 9.3). Table three.1 shows typical nitrate and phosphate concentrations in natural rivers and deep seawater.

Tabular array iii.ane. Levels of Dissolved North and P Mutual in Unlike Kinds of Water Bodies

Nutrient Inland Waters a Atlantic Deep Seawater b Threshold for Poor/Bad Quality Status in Surface Seawater (EEA) Raw Urban Wastewater c Treated Urban Wastewater d
N 0.1–one   mg   Northward/L 20   μM NOthree >ix/>16   μM NOiii 25–twoscore   mg/L N 10–15   mg/L Due north
P 10–50   μg   P/50; hypereutrophic: >96   μg/L ane.5   μM PO4 >0.7/>1.1   μM POfour 7–ten   mg/L P 1–2   mg/50 P
a
Henry & Heinke (1996) op. cit.
b
Chester & Jickells (2012) op. cit.
c
EEA (2007) op. cit.; hypereutrophic: Carlson (2007) op. cit.
d
Directive 98 /15 /EC.

In marine surface waters there is a very high variability depending on geographical factors and oceanographic conditions. In coastal areas nutrient levels are enhanced by the supply from on land and from upwelling, while in central oceanic gyres concentrations of nutrients may be undetectable fifty-fifty for modernistic analytical methods. Therefore, there is not a constant background level of nutrients in surface marine waters, the range of natural scenarios overlaps with the water quality criteria established by different international agencies (Table 3.1), and the interpretation of measured concentrations must exist conducted in lite of complete environmental information to endeavour to make the difference between anthropogenic impact and natural variability.

The main anthropogenic sources of N and P are fertilizers and detergents

The chief source of anthropogenic nitrogen comes from the utilize of artificial fertilizers in agriculture, typically contributing 50–80 percent of the full load. 2 Nitrogen surplus from agricultural land, computed equally fertilizer input minus harvest output, ranges in Europe from ten to 200   Kg/Ha/year depending on the country. 3 Nitrate from fertilizers is very mobile in soil, and is easily leached to groundwater or surface water. Due to the implementation of national and international regulations such as the European union Nitrate Directive (1991), the levels of nitrate in inland waters take been reduced in some regions over the past 15   years, 4 but this did non reverberate in reduced N contents in coastal environments (run into Fig. three.2). Indeed, agriculture is a lengthened source, difficult to control and to quantify, and N levels in h2o courses are highly dependent on rainfall.

Figure 3.2. Temporal trends in the nitrate and phosphate levels in the N Sea and Baltic Sea. Notice the positive development of phosphates compared to nitrates and the resulting overall trend to increase the Due north/P ratios in both areas.

Source: European Surround Bureau.

In dissimilarity, the nearly of import contributors to pollution by phosphorous are generally the point sources, such as wastewater treatment plants and industrial outlets. Detergents are the major source of P in municipal wastewater, although in many countries the phosphorus content of the detergents has been lowered by substitution with other substances (encounter Section 3.4). Industries producing fertilizers may emit quantities of P equal to the full emissions from minor countries, though these emissions decreased significantly every bit a outcome of improved engineering and wastewater handling. 5

Dissimilar N pools may be identified by their 15Northward ratio

The importance of different nitrogen sources and the identification of anthropogenic inputs in inert compartments (east.k., DIN, POM) or marine biota (e.thou., phytoplankton, macroalgae) can be assessed from an isotopic standpoint. Nitrogen has two stable isotopes, the most common xivN, and the heavier isotope with 1 additional neutron, the fifteenDue north. The ratio of the N stable isotopes is termed δxvN and this ratio changes equally nitrogen circulates through certain metabolic routes since the lite isotope may be mobilized faster; a procedure termed isotopic fractionation. Thus, nitrogen from sewage, groundwater, or fish subcontract discharges is often more enriched in 15North than nitrogen from seawater. This is due to isotopic fractionation during nitrification and volatilization in the case of NH four + , or denitrification in the case of NO 3 . half-dozen In dissimilarity, nitrogen pools from almost agronomical facilities are characterized past depleted δ15North values, as they are synthesized from atmospheric Due northtwo. 7 Besides nutrients from anthropogenic origin, different natural processes also affect inorganic nitrogen concentrations and their isotopic ratio. For case, algae from mangrove habitats that were exposed to nitrogen derived from Due northtwo fixation were depleted in 15North while those in habitats with frequent coastal upwelling were relatively enriched. 8

Read full chapter

URL:

https://www.sciencedirect.com/scientific discipline/commodity/pii/B9780128137369000039

Phosphorus Bike☆

Y. Liu , J. Chen , in Encyclopedia of Environmental (2d Edition), 2014

Organic Cycles

Imposed on the inorganic cycle are two organic cycles that motility phosphorus through living organisms every bit part of the food chain. These are a land-based phosphorus cycle that transfers it from soil to plants, to animals, and back to soil once more and a water-based organic wheel that circulates it among the creatures living in rivers, lakes, and seas. The land-based bike takes a yr on average and the water-based bicycle organic cycle simply weeks. Information technology is the corporeality of phosphorus in these two cycles that governs the biomass of living forms that state and sea can sustain.

The amount of phosphorus in the earth׳due south soils is roughly (90–200)   ×   ten3  MMT   P according to various estimates (Emsley, 1980; Filippelli, 2002). While the total phosphorus content of soils is large, but a small fraction is available to biota in most soils. This constitutes an available phosphorus pool containing 1805–3000   MMT   P, most likely 2000–2600   MMT   P (Emsley, 1980; Richey, 1983). A larger amount, in the range (27–840)   ×   106  MMT   P, tin can be found in the oceans. The surface seawater contains (80–3097)   ×   103  MMT   P and the rest is accumulated in deep seawater (Paytan and McLaughlin, 2007).

It is estimated that the oceanic residence time of dissolved P is between 20 and 100 kiloyears, although P is extensively cycled inside the ocean on much shorter timescale. The ocean h2o loses phosphorus continually in a steady drizzle of detritus to the lesser, where it builds up in the sediments as insoluble calcium phosphate. Despite the geologic remobilization, there is a internet annual loss of millions of tons of phosphate a yr from the marine biosphere. Thus, the sea sediments are by far the largest stock in the biogeochemical cycles of phosphorus. Estimates of total P burial in open up bounding main marine sediments range from 2.8   ×   103 to x.five   ×   103  MMT   P per twelvemonth, and the full marine sediments is estimated to be 840   ×   ten3  MMT   P(Paytan and McLaughlin, 2007). The majority of this burial flux is reactive P, with well-nigh of the nonreactive P having been deposited in the continental shelves.

Read full chapter

URL:

https://www.sciencedirect.com/science/article/pii/B9780124095489090436

Dynamics of Dissolved Organic Phosphorus

David M. Karl , Karin M. Björkman , in Biogeochemistry of Marine Dissolved Organic Matter (Second Edition), 2015

Abstract

Phosphorus is an essential macronutrient for all marine microorganisms. Dissolved organic phosphorus (DOP) is an integral, dynamic role of the marine organic affair pool and of the phosphorus cycle, and is of fundamental significance to the understanding of microbial oceanography and marine biogeochemistry. Although much remains to be understood, several new discoveries take been made over the past decade with respect to our knowledge of the chemical composition and microbial processing of DOP in the oceans.

In this affiliate, we cover the history of the early on years of marine phosphorus measurements to novel analytical techniques and the many challenges remaining into the twenty-get-go century. We describe the temporal and spatial variability of the global DOP pool inventories, both in terms of regional gradients as well equally into the deep bounding main interior. We further examine the tools available to characterize the chemically complex DOP pool and the gains made to place specific DOP compounds, or compound classes. As is often the case, the picture show is incomplete, but significant progress has been in the past decade, with refined belittling methodology. Additionally, nosotros explore the role of DOP in the marine P-bicycle, as DOP is produced, utilized by the extant microbial community and eventually remineralized.

Finally, we are just first to understand the, previously underappreciated, reduction-oxidation pathways of phosphorus, and their relationships to cellular bioenergetics, biogeochemistry and ecology. The application of recently developed genomics based applied science gives new opportunities to farther our understanding of the dynamics and regulation of the marine microbial P-cycle.

Read full chapter

URL:

https://world wide web.sciencedirect.com/science/article/pii/B9780124059405000054

Book 1

Eugenia Valsami-Jones , in Encyclopedia of Geology (2d Edition), 2021

Phosphates in the Surround

Phosphate minerals, whether in phosphate ores, rocks or soils, act as the primary source of phosphorus. Their abundance, availability and reactivity are therefore primal to the entire phosphorus cycle on Earth, and through regulation of biological activity also impact on the Globe'south carbon cycle. Phosphorus, every bit an essential nutrient, is extensively recycled inside soils where its availability oftentimes dominates biological processes, notably found growth. Soils are finer a temporary reservoir of phosphorus for plants. Soil phosphate undergoes several biochemical transformations, including conversion from mineral phosphate to organic phosphate. When present naturally, phosphorus occurs in soils at relatively depression concentrations, and thus, in social club to maintain high crop yields in modern day intensive agriculture, addition of phosphorus along with other nutrients in the form of fertilizers is essential. Phosphorus in fertilizers is present in a soluble chemic form ("superphosphate") produced by acid treatment of chief phosphate ore. In this form, phosphorus has a very short life in soil, and either becomes incorporated into the crops, leaches out of the soil, or converts into the less soluble apatite. Manure, another substance used usually every bit fertilizer, contains both soluble inorganic and organic phosphate; the latter is hands convertible to soluble inorganic phosphate and releases phosphate in a like manner to fertilizers.

Phosphorus, carbon and nitrogen are the key regulators of biological activity on Earth. Although phosphorus requirement is the lowest of the three, it has no gaseous presence in the atmosphere, and thus can be the limiting nutrient, in terrestrial environments. The global phosphorus bike has inevitably been modified extensively by modern human practices, notably agriculture, but also urbanization and industrialization. Primal to the global changes of P flux in the environment is the rate of its extraction from phosphate deposits and incorporation in fertilizers, detergents and beast feeds. There are presently many unknowns when phosphorus global fluxes are calculated, not to the lowest degree of which is reliable rates of solubility of chief phosphate minerals. Despite such uncertainties, current predictions cause concerns well-nigh the long-term sustainability of phosphate ore exploitation.

Read full affiliate

URL:

https://www.sciencedirect.com/scientific discipline/article/pii/B9780081029084000321

Phosphorus and selected metals and metalloids

Shiping Deng , in Principles and Applications of Soil Microbiology (3rd Edition), 2021

Phosphorus cycling

Soil phosphorus pools and cycling are intimately connected to the mining of stone phosphate (Fig. xix.3 ) and transporting of terrestrial phosphorus to the sea sediments. A model of the terrestrial phosphorus bike, illustrated in Fig. 19.4, shows the predominant pools and major transformation processes. In contrast to nitrogen, the phosphorus bike is sedimentary with no major gaseous forms of phosphorus produced. The cycling of phosphorus is affected by both biological and chemical reactions. In this model, the solution phosphorus pool serves as the primal indicate in the overall cycle.

Figure 19.iii. Relationships of major terrestrial phosphorus pools.

Effigy 19.four. The phosphorus cycle: predominant pools and major transformation processes.

Plants take up primarily phosphate from the solution P portion of the soil P pool.

In the biological phosphorus subcycle, soil solution phosphorus can be assimilated by plants and soil microorganisms, leading to the germination of orthophosphate and biota phosphorus, such every bit phospholipids, nucleotides (Dna, RNA), ATP, phytate, and biomass phosphorus. These processes result in immobilization of phosphorus, which is the reverse process of mineralization. As constitute residues and animal remains and wastes are returned to soil, the organic phosphorus may exist directly incorporated into persistent soil organic affair (incorporation), mineralized to orthophosphate, or immobilized (assimilated) into the microbial biomass. Biomass phosphorus is subject area to incorporation into soil organic affair besides as mineralization and immobilization reactions. The turnover or cycling of biomass contributes significantly to the labile organic phosphorus pool. Crop removal and erosion are two processes that lead to the loss of organic phosphorus.

In the chemical phosphorus subcycle, orthophosphate is released into the soil solution by chemical and biochemical weathering processes from rock phosphate, minerals, inorganic phosphorus fertilizers, and Al/Fe/Ca/Mn phosphates occluded by secondary minerals. Dissolution and desorption of phosphate minerals releases orthophosphate into the soil solution. Solubilization occurs through organic acids produced by plants and microbes. Orthophosphate can exist adsorbed to aluminum and iron oxides (labile inorganic phosphorus) or precipitated every bit aluminum, iron, or calcium phosphates (secondary minerals). Through time, phosphorus finds its way into streams, lakes, rivers, and somewhen the oceans where it ends up in sediments.

There is growing prove that microbes also produce reduced P species, phosphite (HPOiii 2−) and phosphine gas (PH3), in soil under anaerobic conditions. Some microorganisms have been demonstrated to produce phosphine gas nether highly anaerobic conditions. Likewise, studies accept shown phosphine to be produced under wetland environments such as rice paddy fields. In some aquatic systems, phosphite can account for ten%–thirty% of dissolved P (Figueroa and Coates, 2017). Although high levels of naturally occurring reduced P species are unlikely to exist in near aerobic, agricultural soils, phosphite is sometimes added to these soils every bit a fungicide, fertilizer, and/or biostimulant. Phosphite can positively influence plants by disease suppression or providing phosphorus nutrition through oxidation to phosphate by microorganisms. Added phosphite appears to oxidize slowly in soil with a reported half-life of weeks to months. Yet, the ability to oxidize phosphite appears to be widespread among microorganisms, possibly because of reduced P being the dominant grade on early Earth. Microbial oxidation of phosphite to phosphate by assimilatory phosphite oxidation has been known since the 1950s. More recently, a bacterium (Desulfotignum phosphitoxidans strain FiPS-3) was shown to possess dissimilatory phosphite oxidation capabilities (Schink and Friedrick, 2000). This bacterium was isolated from brackish sediments and is capable of growth by coupling phosphite oxidation to the reduction of either sulfate or carbon dioxide. And so far, this is the only bacterium known to conduct out dissimilatory phosphite oxidation.

Read full chapter

URL:

https://www.sciencedirect.com/scientific discipline/commodity/pii/B9780128202029000198

Volume 5

Alexis Godet , Karl B. Föllmi , in Encyclopedia of Geology (Second Edition), 2021

Abstract

The element phosphorus plays an of import office in the biosphere and the lithosphere: it is involved in the transcription of genes and is one of the most of import micronutrients for life forms. Modernistic sites of phosphogenesis include upwelling coastal zones where large amounts of organic matter and sedimentary phosphate deposits are formed, reflecting links between the phosphorus and carbon cycles. Periods of time during the Earth'southward history witnessed enhanced transfer of phosphorus from continental landmasses to the oceans, or from deep oceanic settings into shallower environments where the stimulated primary productivity favored the deposition on phosphorites—sedimentary rocks with more than than 18% of phosphates. Phosphorites formed in shallow environments and often display evidence of bottom current activeness, such equally reworked grains and nodules, and intraformational erosional contacts. In deeper settings, phosphate-rich deposits are often finely laminated and associated with organic material devoid of bioturbation, a characteristic of oxygen-depleted environments. The geochemistry of phosphorites is influenced by depositional environments and redox conditions that favor the precipitation of phosphate minerals through microbial arbitration. Assessing the nature, geochemistry and evolution of sedimentary phosphate rocks offers an opportunity to trace oceanographic and ecology changes back in geologic time.

Read full chapter

URL:

https://www.sciencedirect.com/science/article/pii/B978008102908400045X