Nitrogen fixation

Nitrogen fixation is a process in which nitrogen (N2) in the atmosphere is converted into ammonium (NH4+).[1] Atmospheric nitrogen or molecular nitrogen (N2) is relatively inert: it does not easily react with other chemicals to form new compounds. The fixation process frees up the nitrogen atoms from their triply bonded diatomic form, N≡N, to be used in other ways.

Nitrogen fixation, natural and synthetic, is essential for all forms of life because nitrogen is required to biosynthesize basic building blocks of plants, animals and other life forms, e.g., nucleotides for DNA and RNA and amino acids for proteins. Therefore, nitrogen fixation is essential for agriculture and the manufacture of fertilizer. It is also an important process in the manufacture of explosives (e.g. gunpowder, dynamite, TNT, etc.). Nitrogen fixation occurs naturally in the air by means of lightning.[2][3]

Nitrogen fixation also refers to other biological conversions of nitrogen, such as its conversion to nitrogen dioxide. All biological nitrogen fixation is done by way of nitrogenase metalo-enzymes which contain iron, molybdenum, or vanadium. Microorganisms that can fix nitrogen are prokaryotes (both bacteria and archaea, distributed throughout their respective kingdoms) called diazotrophs. Some higher plants, and some animals (termites), have formed associations (symbiosis) with diazotrophs.

Biological nitrogen fixation

Schematic representation of the nitrogen cycle. Abiotic nitrogen fixation has been omitted.

Biological nitrogen fixation was discovered by the German agronomist Hermann Hellriegel and Dutch microbiologist Martinus Beijerinck. Biological nitrogen fixation (BNF) occurs when atmospheric nitrogen is converted to ammonia by an enzyme called nitrogenase.[1] The overall reaction for BNF is:

N2 + 8 H+ + 8 e → 2 NH3 + H2

The process is coupled to the hydrolysis of 16 equivalents of ATP and is accompanied by the co-formation of one molecule of H2.[4] The conversion of N2 into ammonia occurs at a cluster called FeMoco, an abbreviation for the iron-molybdenum cofactor. The mechanism proceeds via a series of protonation and reduction steps wherein the FeMoco active site hydrogenates the N2 substrate.[5]

In free-living diazotrophs, the nitrogenase-generated ammonium is assimilated into glutamate through the glutamine synthetase/glutamate synthase pathway.

The microbial genes required for nitrogen fixation are widely distributed in diverse environments.[6][7]

Enzymes responsible for nitrogenase action are very susceptible to destruction by oxygen. For this reason, many bacteria cease production of the enzyme in the presence of oxygen. Many nitrogen-fixing organisms exist only in anaerobic conditions, respiring to draw down oxygen levels, or binding the oxygen with a protein such as leghemoglobin.[1]

Microorganisms that fix nitrogen

Main article: Diazotroph

Diazotrophs are a diverse group of prokaryotes that includes cyanobacteria (e.g. the highly significant Trichodesmium and Cyanothece), green sulfur bacteria, Azotobacteraceae, rhizobia and Frankia.

Cyanobacteria inhabit nearly all illuminated environments on Earth and play key roles in the carbon and nitrogen cycle of the biosphere. In general, cyanobacteria are able to utilize a variety of inorganic and organic sources of combined nitrogen, like nitrate, nitrite, ammonium, urea, or some amino acids. Several cyanobacterial strains are also capable of diazotrophic growth, an ability that may have been present in their last common ancestor in the Archean eon.[8] Nitrogen fixation by cyanobacteria in coral reefs can fix twice the amount of nitrogen than on land—around 1.8 kg of nitrogen is fixed per hectare per day. The colonial marine cyanobacterium Trichodesmium is thought to fix nitrogen on such a scale that it accounts for almost half of the nitrogen-fixation in marine systems on a global scale.[9]

Root nodule symbioses

Legume family

Plants that contribute to nitrogen fixation include the legume family  Fabaceae  with taxa such as kudzu, clovers, soybeans, alfalfa, lupines, peanuts, and rooibos. They contain symbiotic bacteria called rhizobia within nodules in their root systems, producing nitrogen compounds that help the plant to grow and compete with other plants. When the plant dies, the fixed nitrogen is released, making it available to other plants and this helps to fertilize the soil.[1][10] The great majority of legumes have this association, but a few genera (e.g., Styphnolobium) do not. In many traditional and organic farming practices, fields are rotated through various types of crops, which usually includes one consisting mainly or entirely of clover or buckwheat (non-legume family Polygonaceae), which are often referred to as "green manure".

Inga alley farming relies on the leguminous genus Inga, a small tropical, tough-leaved, nitrogen-fixing tree.[11]

Non-leguminous

A sectioned alder tree root nodule.

Although by far the majority of plants able to form nitrogen-fixing root nodules are in the legume family Fabaceae, there are a few exceptions:

The ability to fix nitrogen is far from universally present in these families. For instance, of 122 genera in the Rosaceae, only 4 genera are capable of fixing nitrogen. All these families belong to the orders Cucurbitales, Fagales, and Rosales, which together with the Fabales form a clade of eurosids. In this clade, Fabales were the first lineage to branch off; thus, the ability to fix nitrogen may be plesiomorphic and subsequently lost in most descendants of the original nitrogen-fixing plant; however, it may be that the basic genetic and physiological requirements were present in an incipient state in the last common ancestors of all these plants, but only evolved to full function in some of them:

Family: Genera

Betulaceae: Alnus (alders)

Cannabaceae: Trema

Casuarinaceae:

Allocasuarina
Casuarina
Ceuthostoma
Gymnostoma

......


Coriariaceae: Coriaria

Datiscaceae: Datisca

Elaeagnaceae:

Elaeagnus (silverberries)
Hippophae (sea-buckthorns)
Shepherdia (buffaloberries)

......


Myricaceae:

Comptonia (sweetfern)
Morella
Myrica (bayberries)

......


Rhamnaceae:

Ceanothus
Colletia
Discaria
Kentrothamnus
Retanilla
Talguenea
Trevoa

......


Rosaceae:

Cercocarpus (mountain mahoganies)
Chamaebatia (mountain miseries)
Dryas
Purshia/Cowania (bitterbrushes/cliffroses)

There are also several nitrogen-fixing symbiotic associations that involve cyanobacteria (such as Nostoc):

Endosymbiosis in diatoms

Rhopalodia gibba, a diatom algae, is propably the only eukaryote with cyanobacterial N2-fixing endosymbiont organelles. The spheroid bodies reside in the cytoplasm of the diatoms and are inseparable from their hosts.[14][15]

Industrial nitrogen fixation

The possibility that atmospheric nitrogen reacts with certain chemicals was first observed by Desfosses in 1828. He observed that mixtures of alkali metal oxides and carbon react at high temperatures with nitrogen. With the use of barium carbonate as starting material the first commercially used process became available in the 1860s developed by Margueritte and Sourdeval. The resulting barium cyanide could be reacted with steam yielding ammonia. In 1898 Adolph Frank and Nikodem Caro decoupled the process and first produced calcium carbide and in a subsequent step reacted it with nitrogen to calcium cyanamide. The Ostwald process for the production of nitric acid was discovered in 1902. Frank-Caro process and Ostwald process dominated the industrial fixation of nitrogen until the discovery of the Haber process in 1909.[16][17] Prior to 1900, Nikola Tesla also experimented with the industrial fixation of nitrogen "by using currents of extremely high frequency or rate of vibration".[18][19]

Haber process

Main article: Haber process
Equipment for a study of nitrogen fixation by alpha rays, Fixed Nitrogen Research Laboratory, 1926

Artificial fertilizer production is now the largest source of human-produced fixed nitrogen in the Earth's ecosystem. Ammonia is a required precursor to fertilizers, explosives, and other products. The most common method is the Haber process. The Haber process requires high pressures (around 200 atm) and high temperatures (at least 400 °C), routine conditions for industrial catalysis. This highly efficient process uses natural gas as a hydrogen source and air as a nitrogen source.[20]

Much research has been conducted on the discovery of catalysts for nitrogen fixation, often with the goal of reducing the energy required for this conversion. However, such research has thus far failed to even approach the efficiency and ease of the Haber process. Many compounds react with atmospheric nitrogen to give dinitrogen complexes. The first dinitrogen complex to be reported was Ru(NH3)5(N2)2+.[21]

Ambient nitrogen reduction

Catalytic chemical nitrogen fixation at ambient conditions is an ongoing scientific endeavor. Guided by the example of nitrogenase, this area of homogeneous catalysis is ongoing, with particular emphasis on hydrogenation to give ammonia.[22]

It has long been known that metallic Lithium burns in an atmosphere of nitrogen, converting to lithium nitride. Hydrolysis of the resulting nitride gives ammonia. In a related process, trimethylsilyl chloride, lithium, and nitrogen react in the presence of a catalyst to give tris(trimethylsilyl)amine. Tris(trimethylsilyl)amine can then be used for reaction with α,δ,ω-triketones to give tricyclic pyrroles.[23] Processes involving Li metal are however of no practical interest since they are noncatalytic (since it is difficult to re-reduce Li+).

Beginning in the 1960s several homogeneous systems were identified that convert nitrogen to ammonia, sometimes even catalytically but often operating via ill-defined mechanisms. The original discovery is described in an early review, "Vol'pin and co-workers, using a non-protic Lewis acid, aluminium tribromide, were able to demonstrate the truly catalytic effect of titanium by treating dinitrogen with a mixture of titanium tetrachloride, metallic aluminium, and aluminium tribromide at 50 °C, either in the absence or in the presence of a solvent, e.g. benzene. As much as 200 mol of ammonia per mol of TiCl4 was obtained after hydrolysis...."[24]

The quest for well defined intermediates led to the characterization of many transition metal dinitrogen complexes. Few of these well defined complexes function catalytically, their behavior illuminated likely stages in nitrogen fixation. Most fruitful of all of these early studies focused on M(N2)2(dppe)2 (M = Mo, W). For example double protonation of such low valent complexes gave intermediates with the linkage M=N-NH2. In 1995, a molybdenum(III) amido complex was discovered that cleaved N2 to give the corresponding molybdenum(VI) nitride.[25] This and related terminal nitrido complexes have been used to make nitriles.[26]

Synthetic nitrogen reduction Yandulov 2003

In 2003 a related molybdenum amido complex was found to catalyze the reduction of N2. In addition to a source of protons, the catalyst requires a strong reducing agent.[27][28][29][30] However, this catalytic reduction fixates only a few nitrogen molecules. In these systems, like the biological one, hydrogen is provided to the substrate heterolytically, by means of protons and reducing equivalents rather than with H2 itself.

In 2011 Arashiba et al. reported yet another system with a catalyst again based on molybdenum but with a diphosphorus pincer ligand.[31] Photolytic nitrogen splitting is also considered.[32][33][34][35][36]

See also

References

  1. 1.0 1.1 1.2 1.3 Postgate, J. (1998). Nitrogen Fixation, 3rd Edition. Cambridge University Press, Cambridge UK.
  2. Slosson, Edwin (1919). Creative Chemistry. New York: The Century Co. pp. 19–37.
  3. http://journals.ametsoc.org/doi/abs/10.1175/1520-0469%281980%29037%3C0179%3AANFBL%3E2.0.CO%3B2
  4. Chi Chung, Lee; Markus W., Ribbe; Yilin, Hu (2014). "Chapter 7. Cleaving the N,N Triple Bond: The Transformation of Dinitrogen to Ammonia by Nitrogenases". In Peter M.H. Kroneck and Martha E. Sosa Torres. The Metal-Driven Biogeochemistry of Gaseous Compounds in the Environment. Metal Ions in Life Sciences 14. Springer. pp. 147–174. doi:10.1007/978-94-017-9269-1_6.
  5. Hoffman, B. M.; Lukoyanov, D.; Dean, D. R.; Seefeldt, L. C. (2013). "Nitrogenase: A Draft Mechanism". Acc. Chem. Res. 46: 587. doi:10.1021/ar300267m.
  6. Gaby, J.C.; Buckley, D.H. (2011). "A global census of nitrogenase diversity". Environmental Microbiology 13 (7): 1790–1799. doi:10.1111/j.1462-2920.2011.02488.x. PMID 21535343.
  7. Hoppe, B.; Kahl, T.; Karasch, P.; Wubet, T.; Bauhus, J.; Buscot, F.; Krüger, D. (2014). "Network analysis reveals ecological links between N-fixing bacteria and wood-decaying fungi". PLoS ONE 9 (2): e88141. Bibcode:2014PLoSO...988141H. doi:10.1371/journal.pone.0088141. PMC 3914916. PMID 24505405.
  8. Latysheva, N.; Junker, V. L.; Palmer, W. J.; Codd, G. A.; Barker, D. (2012). "The evolution of nitrogen fixation in cyanobacteria". Bioinformatics 28 (5): 603–606. doi:10.1093/bioinformatics/bts008.
  9. Bergman, B.; Sandh, G.; Lin, S.; Larsson, H.; and Carpenter, E. J. (2012). "Trichodesmium – a widespread marine cyanobacterium with unusual nitrogen fixation properties". FEMS Microbiology Reviews 37 (3): 1–17. doi:10.1111/j.1574-6976.2012.00352.x.
  10. Smil, V (2000). Cycles of Life. Scientific American Library.
  11. Elkan, Daniel. "Slash-and-burn farming has become a major threat to the world's rainforest". The Guardian, 21 April 2004.
  12. Op den Camp, Rik; Streng, A. et al. (2010). "LysM-Type Mycorrhizal Receptor Recruited for Rhizobium Symbiosis in Nonlegume Parasponia". Science 331 (6019): 909–912. Bibcode:2011Sci...331..909O. doi:10.1126/science.1198181.
  13. Dawson, J. O. (2008). "Ecology of actinorhizal plants". Nitrogen-fixing Actinorhizal Symbioses 6. Springer. pp. 199–234. doi:10.1007/978-1-4020-3547-0_8.
  14. Prechtl, J. et al. (2004). Intracellular spheroid bodies of Rhopalodia gibba have nitrogen-fixing apparatus of cyanobacterial origin. Molecular biology and evolution, 21(8), 1477-1481, .
  15. Nakayama, T., & Inagaki, Y. (2014). Unique genome evolution in an intracellular N2-fixing symbiont of a rhopalodiacean diatom. Acta Societatis Botanicorum Poloniae, 83 (4), 409-413, .
  16. Heinrich, H.; Nevbner, Rolf (1934). "Die Umwandlungsgleichung Ba(Cn)2 → BaCN2 + C Im Temperaturgebiet von 500 Bis 1000 °C". Zeitschrift für Elektrochemie und angewandte physikalische Chemie 40 (10): 693–698. doi:10.1002/bbpc.19340401005 (inactive 2015-02-01).
  17. Curtis, Harry Alfred (1932). Fixed nitrogen.
  18. http://www.tfcbooks.com/tesla/1900-06-00.htm
  19. Tesla, Nikola (1900). "The Problem of Increasing Human Energy". The Century Magazine. 60 (n.s. v. 38) (1900 May–Oct): 175.
  20. http://cfpub.epa.gov/watertrain/pdf/issue1.pdf US Environmental Protection Agency: Human Alteration of the Global Nitrogen Cycle: Causes and Consequences by Peter M. Vitousek, Chair, John Aber, Robert W. Howarth, Gene E. Likens, Pamela A. Matson, David W. Schindler, William H. Schlesinger, and G. David Tilman
  21. A. D. Allen, C. V. Senoff; Senoff (1965). "Nitrogenopentammineruthenium(II) complexes". Journal of the Chemical Society, Chemical Communications (24): 621. doi:10.1039/C19650000621.
  22. "Reduction of dinitrogen" Richard R. Schrock PNAS 14 November 2006 vol. 103 no. 46 17087 doi:10.1073/pnas.0603633103
  23. Brook, Michael A. (2000). Silicon in Organic, Organometallic, and Polymer Chemistry. New York: John Wiley & Sons, Inc. pp. 193–194.
  24. Chatt, J.; Leigh, G. J., "Nitrogen Fixation", Chem. Soc. Rev. 1972, vol. 1, 121.
  25. "Dinitrogen Cleavage by a Three-Coordinate Molybdenum(III) Complex" Catalina E. Laplaza and Christopher C. Cummins Science 12 May 1995: 861–863.10.1126/science.268.5212.861
  26. "A Cycle for Organic Nitrile Synthesis via Dinitrogen Cleavage" John J. Curley, Emma L. Sceats, and Christopher C. Cummins J. Am. Chem. Soc., 2006, 128 (43), pp. 14036–14037 doi:10.1021/ja066090a
  27. Synthesis and Reactions of Molybdenum Triamidoamine Complexes Containing Hexaisopropylterphenyl Substituents Dmitry V. Yandulov, Richard R. Schrock, Arnold L. Rheingold, Christopher Ceccarelli, and William M. Davis Inorg. Chem.; 2003; 42(3) pp 796813; (Article) doi:10.1021/ic020505l
  28. "Catalytic Reduction of Dinitrogen to Ammonia at a Single Molybdenum Center" Dmitry V. Yandulov and Richard R. Schrock Science 4 July 2003: Vol. 301. no. 5629, pp. 7678 doi:10.1126/science.1085326
  29. The catalyst is derived from molybdenum(V) chloride and tris(2-aminoethyl)amine N-substituted with three very bulky hexa-isopropylterphenyl (HIPT) groups. Nitrogen adds end-on to the molybdenum atom, and the bulky HIPT substituents prevent the formation of the stable and nonreactive Mo-N=N-Mo dimer. In this isolated pocket the Mo-N2 . The proton donor is a pyridinium salt of weakly coordinating counter anion. The reducing agent is decamethylchromocene. All ammonia formed is collected as the HCl salt by trapping the distillate with a HCl solution
  30. Note also that, although the dinitrogen complex is shown in brackets, this species can be isolated and characterized. Here the brackets do not indicate that the intermediate is not observed.
  31. Arashiba, Kazuya; Yoshihiro; Yoshiaki Nishibayashi, Miyake (2011). "A molybdenum complex bearing PNP-type pincer ligands leads to the catalytic reduction of dinitrogen into ammonia". Nature Chemistry 3: 120–125. doi:10.1038/nchem.906.
  32. Rebreyend, C. and de Bruin, B. (2014), Photolytic N2 Splitting: A Road to Sustainable NH3 Production?. Angew. Chem. Int. Ed.doi:10.1002/anie.201409727
  33. Solari, E., Da Silva, C., Iacono, B., Hesschenbrouck, J., Rizzoli, C., Scopelliti, R. and Floriani, C. (2001) "Photochemical Activation of the N≡N Bond in a Dimolybdenum–Dinitrogen Complex: Formation of a Molybdenum Nitride" Angew. Chem. Int. Ed 40: 3907–3909 doi:10.1002/1521-3773(20011015)40:20<3907::AID-ANIE3907>3.0.CO;2-#
  34. Huss, Adam S.; Curley, John J.; Cummins, Christopher C.; Blank, David A. (2013). "Relaxation and Dissociation Following Photoexcitation of the (?-N2)[Mo(N[t-Bu]Ar)3]2 Dinitrogen Cleavage Intermediate". The Journal of Physical Chemistry B 117 (5): 1429–1436. doi:10.1021/jp310122x.
  35. Kunkely, H.; Vogler, A. (2010). "Photolysis of Aqueous [(NH3)5Os(μ-N2)Os(NH3)5]5+: Cleavage of Dinitrogen by an Intramolecular Photoredox Reaction". Angew. Chem. Int. Ed. 49: 1591–1593. doi:10.1002/anie.200905026.
  36. Miyazaki, T.; Tanaka, H.; Tanabe, Y.; Yuki, M.; Nakajima, K.; Yoshizawa, K.; Nishibayashi, Y. (2014). "Cleavage and Formation of Molecular Dinitrogen in a Single System Assisted by Molybdenum Complexes Bearing Ferrocenyldiphosphine". Angew. Chem. Int. Ed 53: 11488–11492. doi:10.1002/anie.201405673.

External links