Skip to main content

Glasses as sources of condensed phosphates on the early earth

Abstract

Procedures for the analysis of phosphorus in geological material normally aims for the determination of the total amount of P expressed as orthophosphate PO 4 3 - or the differentiation between inorganic and organic P. This is probably due to analytical difficulties but also to the prevalent opinion that the chemistry of phosphorus in geological environments is almost entirely restricted to the mineral apatite. Because of the low solubility of apatite it is, therefore, commonly argued that little P was around for prebiotic chemistry and that pre-biological processes would essentially have had to do without this indispensable element unless it was provided by alternative sources or mechanisms (such as reduction and activation by lightning or delivery to Earth by celestial bodies). It is a paradox that the potential existence of reactive phosphorus compounds, such as the mineral schreibersite - iron phosphide, in geological material on Earth is seldom considered although we are aware of the existence of such compounds in meteorite material. The content of Al2O3 in rocks appears to be important for the speciation of phosphorus and for how strongly it binds to silicates. In general, low alumina seems to promote the existence of isolated charge-balanced phosphorus complexes.

Introduction

It is conventional wisdom that phosphorus has always been a scarce element on Earth, that the little P that was around was locked up in insoluble apatite, and that no P in the form of condensed phosphates, like pyrophosphate, occurs in the geosphere [1–3]. This is not necessarily true. We know little of the speciation of P in the lithosphere, especially in melts and non-crystalline materials like glasses and gels [4]. The content of Al2O3 appears to be important for phosphorus chemistry, since increases in the alumina/alkali ratio leads to a decreased effect of phosphorus on silicate polymerization [5]. The suggestion that condensed phosphates are dissolved in glasses is supported by experiments carried out by Cody, Mysen, and colleagues [5–7]. Their studies showed that pyrophosphate and trimetaphosphate ((P3O9)3-) are stable in alkaline glasses at low concentrations of Al2O3. At about 2 mol% Al2O3 the phosphate speciation is dominated by pyrophosphate. At 5 mol% Al2O3 trimetaphosphate appears to be present. Phosphorus shows a stronger preference for the basaltic – mafic – member of coexisting immicible granitic and basaltic melts [8]. Yamagata and coworkers [9] have, accordingly, reported the occurrence of pyrophosphate (0.45 μM) and triphosphate (0.37 μM) in fumaroles presumably formed through partial hydrolysis of P4O10 emanating from basaltic melts of Mount Usu on the island of Hokkaido, Japan. Ultramafic lavas with an even lower content of SiO2 (about 40–45 wt%) than basalt (45-52 wt%), a CaO:Al2O3 ratio >1, and more than 18 wt% MgO, has been given the term komatiites [10, 11]. Abundant deposition of ultramafic volcanic glasses due to komatiitic volcanism in the Hadean and early Archean is probable [12, 13], and oceanic crust is likely to have been komatiitic and fed from komatiitic liquids [14]. Ultramafic glasses may have played an important role in prebiotic chemistry of P on the early Earth.

Condensed phosphates in prebiotic chemistry

Keefe and Miller discussed about 20 years ago whether condensed phosphates like pentavalent pyrophosphate ((P2O7)4-) were ever likely prebiotic reagents on Earth [2]. At the time they argued that a widespread occurrence of protonated orthophosphates, which are known to serve as precursors of pyrophosphate, is doubtful in nature (cf. [15]), and that heating of such protonated phosphates in a closed system does not give polyphosphates because water cannot escape. However, pyrophosphate could have been formed during early subduction of oceanic lithosphere by dehydration of known protonated orthophosphates like whitlockite (Ca18Mg2H2(PO4)14) and newberyite (MgHPO4.3H2O) [16–18]. Whitlockite forms in a sterile (non-biotic) seawater medium in the pH range 7–9 [19]. Newberyite crystallizes at lower pH [20]. Prior to the development of life on Earth, newberyite may have preceded organogenic apatite as the dominant inorganic phosphate mineral being formed in the primordial ocean floor [16, 21]. The key to the pyrophosphate formation in these geological environments is low water to rock ratio or low rate of hydrolysis, i.e. low local activity of water, which has the effect that the water can be removed [4]. Pyrophosphate may have preceded ATP (adenosine triphosphate) as energy ‘currency’ in connection with the origin and early evolution of life on Earth [22, 23]. It is possible that pyrophosphate was originally incorporated into prebiotic systems because of its energy-transfer capacity and that the information coding function of P as a constituent of polynucleotides is a secondary property [24]. Another condensed phosphate, the ring-formed trimetaphosphate ((P3O9)3-), has, on the other hand, been shown to be the most effective condensing agent among polyphosphates, particularly because it can serve as a phosphorylating agent reagent in strongly alkaline aqueous solutions (pH 12) [18, 25–28]. Such high pH values can be found in, for instance, the fairly recently discovered modern sediment-starved subduction zones with serpentinite mud volcanoes, like the Mariana Trench [29, 30]. Serpentine mud volcanoes similar to those of the Mariana forearc next to the trench occur at Isua, Greenland, and have been dated to early Archean (3.81-3.70 Ga) [31]. Prieur has shown that it is possible to phosphorylate the carbohydrate ribose with trimetaphosphate to ribose-5-phosphate – which is a central part of nucleotides like ATP and RNA (ribonucleic acid) - in the presence of borate [32, 33]. According to Prieur, pyrophosphate does not appear to have as good phosphorylating effect.

Delivery of phosphorus by meteorites

Since Keefe and Miller published their paper in 1995, Pasek and Block [34] have analyzed lightning-derived glass compounds, termed fulgerites, and shown that phosphate can be reduced to phosphite and phosphide by cloud-to-ground lightning. The reduction of phosphate to phosphite by lightning discharge had been shown a few years earlier in laboratory experiments carried out by Glindemann et al. [35]. Pasek and co-workers also suggested that phosphite derived from the iron-nickel phosphide mineral schreibersite ((Fe,Ni)3P) in meteorites was a plausible reagent in the prebiotic synthesis of phosphorylated biomolecules on the early Earth [36–38]. Schreibersite is particularly abundant in some iron meteorites [39]. The phosphite that Pasek and co-workers postulate to have been prebiotically active would be an oxidation product of phosphide that was delivered to Earth during the Hadean-Archean heavy meteorite bombardment >3.9 Ga [37]. This scenario is a potential solution of the ‘phosphate problem’ as discussed by Schwartz [40, 41], i.e. solubilization of phosphate compounds is necessary before activation can occur. Schreibersite oxidizes slowly in contact with fluid water as the surrounding mineral matrix gets weathered, and forms several phosphorus species of mixed oxidation states like orthophosphate, pyrophosphate, hypophosphate, phosphite, etc. [36–38, 42, 43].

Condensed phosphorus compounds in minerals and glasses on earth

Environments with condensed phosphorus compounds are not absent on our own planet, although such compounds are normally not searched for. Standard analytical procedures of P in geological material include the hydrolysis of all potential forms of P to orthophosphate PO 4 3 - . The first occurrence on Earth of a natural pyrophosphate mineral, canaphite, was not reported in the scientific literature until 1985 [44, 45]. The second mineral, wooldridgeite, was discovered in 1999 [46]. The canaphite was found on crystals of the zeolite mineral stilbite in cavities of Triassic basalts in New Jersey, USA, whereas wooldridgeite was found in an unconformity between Precambrian igneous rocks and overlying Cambrian sedimentary rocks in Warwickshire, England. The reason why such minerals were not found earlier maybe reflects discussions in the chemistry community. It was actually not generally accepted in chemistry that polyphosphates larger than pyrophosphate existed until mid-20th century [47]. The idea of triphosphate as a crystalline entity was, for instance, not accepted until around 1940 [47].

In modern oceanic sediments in areas of low-temperature hydrothermal activity phosphate is strongly enriched as authigenic phases in the basal sedimentary layer on top of the basaltic basement, with the source of phosphorus being primarily the basalts underneath [48]. The phosphorus reported in the study of Bodeï and co-workers was determined in bulk and occurred in association with zeolites and Fe-Mn oxide hydroxides. Silicate minerals can account for as much as 25–30% of the total content of P in igneous and metamorphic rocks [49], i.e. the rocks may be undersatured with respect to separate phosphate minerals. Studies have shown that partitioning of phosphorus between silicate solids preferentially favors glasses, alkaline glasses in particular, relative to olivine, amphibole, and clinopyroxene [50]. Experiments by Milman-Barris and co-workers show that glass adjacent to olivine phenocrysts is enriched relative to the far-field glass by up to 20% in P2O5, so the glass appears to extract phosphorus from the olivine [51]. Glass of phosphate is widely distributed in the lithospheric mantle and is formed during rapid quenching of mantle material [52].

Phosphorus of different speciation may be released from the basalts and precipitated as secondary protonated phases during water convection and weathering in the surface parts of marine basement. Magnesium pyrophosphate can be formed easily under mild hydrothermal conditions (165-180°C), like those existing in subduction zones, from protonated Mg salts and orthophosphate, for example newberyite [53–55]. The reason is probably that the size of the Mg2+ ion makes it possible to simultaneously coordinate and stabilize negatively charged oxygen of two adjacent phosphorus atoms, i.e. Mg(II) has the function of being a ‘bidentate clamp’ [56]. Condensed phosphates have stronger binding energies to hydroxide minerals (e.g. brucite, smectite) than orthophosphate [17]. It has also been shown that the presence of free Fe(II) ions as well as Fe(II) minerals retard the hydrolysis of pyrophosphate [57]. Therefore, it is likely that any condensed phosphates existing in oceanic basement will stay attached to authigenic minerals until the plate is processed in a subduction zone [23]. Nitschke and Russell have proposed that pyrophosphate is dissolved in basaltic glasses and is released upon alteration of the glass into palagonite [58], i.e. a metabasite which contains a mixture of palagonitized glass, authigenic minerals like smectite, corrensite, zeolites, carbonates, and Fe-Ti oxides and phosphates, as well as primary minerals like plagioclase feldspars, clinopyroxene and olivine [59]. The rate of dissolution of phosphate glass is reduced with increased proportions of multiply charged metal ions like Mg(II) [47]. It has, on the other hand, been known for quite some time that the presence of divalent cations like Mg(II) accelerates the formation of ring-formed metaphosphates from polyphosphates [60]. This is probably due to greater ease of chelation within the polyphosphate chains than at the chain ends.

Mitsis and Economou-Eliopoulos concluded in their study of hydroxylapatite associated with massive magnetite in the Othrys ophiolite complex, Greece, that the apatite was formed from re-mobilization of phosphorus and deposition during circulation of hydrothermal fluids [61]. Even though ridge-axis and ridge-flank hydrothermal processes in the ocean floor are estimated to remove about 50% of the global input of P into oceanic crust [62], hydrothermal fluids are obviously able to re-distribute P within hydrothermal systems. However, it is not yet known whether the ‘primary source’ of P after removal from seawater exists entirely in the form of orthophosphate, is bound to the minerals of the mafic rocks, or occur as more soluble condensed P species.

Conclusions

Phosphorus is an extremely important element for the biological coding of information as well as for the transfer of energy and information in all living organisms on Earth. It is possible that the energy-transfer capacity of phosphorus compounds is a primary property for life processes and that the information coding function is secondary. It is often claimed that phosphorus is a rare and inaccessible element on our planet and that the earliest biosphere, therefore, would have been almost devoid of this indispensable element unless it was provided by alternative sources or mechanisms, such as activation by lightning or delivery as more accessible reduced phosphorus species in iron meteorites. Marine scientists often emphasize that we know more of the surface of the other terrestrial planets than we know of the ocean floor on Earth. This may be true also for the grand scale element cycling of essential nutrients in oceanic basement. The building blocks of the first biochemicals like phosphorus may, after all, have been around and available for prebiotic chemistry during Earth’s entire life span.

References

  1. De Duve C: Vital Dust; the Origin and Evolution of Life on Earth. 1995, New York: Basic Books

    Google Scholar 

  2. Keefe AD, Miller SL: Are polyphosphates or phosphate esters prebiotic reagents?. J Mol Evol. 1995, 41: 693-702.

    Article  Google Scholar 

  3. Zubay G: Origins of Life on the Earth and in the Cosmos. 2000, San Diego, CA: Harcourt, 2

    Google Scholar 

  4. Da Silva JAL, Holm NG: Borosilicates and silicophosphates as plausible contributors to the origin of life. J Colloid Interface Sci. 2014, early view at http://www.sciencedirect.com/science/article/pii/S0021979714001179,

    Google Scholar 

  5. Mysen BO, Cody GD: Silicate-phosphate interactions in silicate glasses and melts: II. Quantitative, high-temperature structure of P-bearing alkali aluminosilicate melts. Geochim Cosmochim Acta. 2001, 65: 2413-2431. 10.1016/S0016-7037(01)00598-1.

    Article  Google Scholar 

  6. Cody GD, Mysen B, Sághi-Szabó G, Tosell JA: Silicate-phosphate interactions in silicate glasses and melts: I. A multinuclear (27Al,29Si,31P) MAS NMR and ab initio chemical shielding (31P) study of phosphorus speciation in silicate glasses. Geochim Cosmochim Acta. 2001, 65: 2395-2411. 10.1016/S0016-7037(01)00597-X.

    Article  Google Scholar 

  7. Mysen BO: An experimental study of phosphorus and aluminosilicate speciation in and partitioning between aqueous fluids and silicate melts determined in-situ at high temperature and pressure. Am Mineral. 2011, 96: 1636-1649. 10.2138/am.2011.3728.

    Article  Google Scholar 

  8. Toplis MJ, Libourel G, Carroll MR: The role of phosphorus in crystallization processes of basalt: an experimental study. Geochim Cosmochim Acta. 1994, 58: 797-810. 10.1016/0016-7037(94)90506-1.

    Article  Google Scholar 

  9. Yamagata Y, Watanabe H, Saithoh M, Namba T: Volcanic production of polyphosphates and its relevance to prebiotic evolution. Nature. 1991, 352: 516-519. 10.1038/352516a0.

    Article  Google Scholar 

  10. Arndt NT, Nisbet EG: What is a Komatiite?. Komatiites. Edited by: Arndt NT, Nisbet EG. 1982, London: Allen & Unwin, 19-27.

    Google Scholar 

  11. Grove TL, Parman SW: Thermal evolution of the Earth as recorded by komatiites. Earth Planet Sci Lett. 2004, 219: 173-187. 10.1016/S0012-821X(04)00002-0.

    Article  Google Scholar 

  12. Arndt NT, Nisbet EG: Processes of the young Earth and the habitats of early life. Annu Rev Earth Planet Sci. 2012, 40: 521-549. 10.1146/annurev-earth-042711-105316.

    Article  Google Scholar 

  13. Taylor SR, McLennan SM: Planetary Crusts: Their Composition, Origin and Evolution. 2009, Cambridge: Cambridge University Press

    Google Scholar 

  14. Nisbet EG: The Young Earth: an Introduction to Archean Geology. 1987, Boston MA: Allen & Unwin

    Book  Google Scholar 

  15. Nriagu JO, Moore PB: Phosphate minerals. 1984, Berlin: Springer

    Book  Google Scholar 

  16. Sales BC, Chakoumakos BC, Boatner LA, Ramey JO: Structural properties of amorphous phases produced by heating crystalline MgHPO4.3H2O. J Non-Cryst Solids. 1993, 159: 121-139. 10.1016/0022-3093(93)91289-F.

    Article  Google Scholar 

  17. Arrhenius GO, Sales B, Mojzsis S, Lee T: Entropy and charge in molecular evolution – the case of phosphate. J Theor Biol. 1997, 187: 503-522. 10.1006/jtbi.1996.0385.

    Article  Google Scholar 

  18. Kolb V, Zhang S, Xu Y, Arrhenius G: Mineral induced phosphorylation of glycolate ion – a metaphor in chemical evolution. Orig Life Evol Biosph. 1997, 27: 485-503. 10.1023/A:1006582526535.

    Article  Google Scholar 

  19. Gedulin B, Arrhenius G: Sources and Geochemical Evolution of RNA Precursor Molecules: The Role of Phosphate. Early Life on Earth. Edited by: Bengtson S. 1994, New York: Columbia University Press, 90-106.

    Google Scholar 

  20. Boistelle R, Abbona F, Lundager Madsen HE: On the transformation of struvite into newberyite in aqueous systems. Phys Chem Miner. 1983, 9: 216-222. 10.1007/BF00311958.

    Article  Google Scholar 

  21. Assaaoudi H, Fang Z, Butler IS, Ryan DH, Kozinski JA: Characterization of a new magnesium hydrogen orthophosphate salt, Mg3.5H2(PO4)3, synthesized in supercritical water. Solid State Sci. 2007, 9: 385-393. 10.1016/j.solidstatesciences.2007.03.015.

    Article  Google Scholar 

  22. Baltscheffsky H, Baltscheffsky M: Molecular Origin and Evolution of Early Energy Conversion. Early Life on Earth. Edited by: Bengtson S. 1994, New York: Columbia University Press, 81-90.

    Google Scholar 

  23. Holm NG, Baltscheffsky H: Links between hydrothermal environments, pyrophosphate, Na+, and early evolution. Origins Life Evol Biosphere. 2011, 41: 483-493. 10.1007/s11084-011-9235-4.

    Article  Google Scholar 

  24. Holm NG, Neubeck A, Ivarsson M, Konn C: Abiotic Organic Synthesis Beneath the Ocean Floor. Astrobiology: Emergence, Search and Detection of Life. Edited by: Basiuk VA. 2010, Stevenson Ranch, CA: American Scientific Publishers, 187-198.

    Google Scholar 

  25. Etaix E, Orgel LE: Phosphorylation of nucleosides in aqueous solution using trimetaphosphate: formation of nucleoside triphosphates. J Carbohydr Nucleosides Nucleotides. 1978, 5: 91-110.

    Google Scholar 

  26. Yamagata Y, Inoue H, Inomata K: Specific effect of magnesium ion on 2’,3’-cyclic AMP synthesis from adenosine and trimeta phosphate in aqueous solution. Orig Life Evol Biosph. 1995, 25: 47-52. 10.1007/BF01581572.

    Article  Google Scholar 

  27. Kolb V, Orgel LE: Phosphorylation of glyceric acid in aqueous solution using trimetaphosphate. Orig Life Evol Biosph. 1996, 26: 7-13. 10.1007/BF01808156.

    Article  Google Scholar 

  28. Ozawa K, Nemoto A, Imai EI, Honda H, Hatori K, Matsuno K: Phosphorylation of nucleotide molecules in hydrothermal environments. Orig Life Evol Biosph. 2004, 34: 465-471.

    Article  Google Scholar 

  29. Fryer P, Wheat CG, Mottl MJ: Mariana blueschist mud volcanism: implications for conditions within the subduction zone. Geology. 1999, 27: 103-106. 10.1130/0091-7613(1999)027<0103:MBMVIF>2.3.CO;2.

    Article  Google Scholar 

  30. Mottl MJ: Pore waters from serpentinite seamounts in the Mariana and Izu-Bonin forearcs, Leg 125. Proc Ocean Drill Prog Sci Results. 1992, 125: 373-385.

    Google Scholar 

  31. Pons ML, Quitté G, Fujii T, Rosing MT, Reynard B, Moynier F, Douchet C, Albarède F: Early Archean serpentine mud volcanoes at Isua, Greenland, as a niche for early life. Proc Natl Acad Sci U S A. 2011, 108: 17639-17643. 10.1073/pnas.1108061108.

    Article  Google Scholar 

  32. Prieur B: Etude de l’activité prébiotique potentielle de l’acide borique. Comptes Rendus de l’Académie des Sciences, Chemie/Chemistry. 2001, 4: 667-670.

    Google Scholar 

  33. Prieur BE: Phosphorylation of ribose in the presence of borate salts. Origins Life Evol Biosphere. 2009, 39: 264-265.

    Google Scholar 

  34. Pasek M, Block K: Lightning-induced reduction of phosphorus oxidation state. Nat Geosci. 2009, 2: 553-556. 10.1038/ngeo580.

    Article  Google Scholar 

  35. Glindemann D, De Graaf RM, Schwartz AW: Chemical reduction of phosphate on the primitive Earth. Orig Life Evol Biosph. 1999, 29: 555-561. 10.1023/A:1006622900660.

    Article  Google Scholar 

  36. Pasek MA, Dworkin JP, Lauretta DS: A radical pathway for organic phosphorylation during schreibersite corrosion with implications for the origin of life. Geochim Cosmochim Acta. 2007, 71: 1721-1736. 10.1016/j.gca.2006.12.018.

    Article  Google Scholar 

  37. Pasek MA, Harnmeijer JP, Buick R, Gull M, Atlas Z: Evidence for reactive reduced phosphorus species in the early Archean ocean. Proc Natl Acad Sci U S A. 2013, 110: 10089-10094. 10.1073/pnas.1303904110.

    Article  Google Scholar 

  38. Pasek MA: Rethinking early Earth geochemistry. Proc Natl Acad Sci U S A. 2008, 105: 853-858. 10.1073/pnas.0708205105.

    Article  Google Scholar 

  39. Hazen RM: Paleomineralogy of the Hadean Eon: a preliminary species list. Am J Sci. 2013, 313: 807-843. 10.2475/09.2013.01.

    Article  Google Scholar 

  40. Schwartz AW, van der Veen M, Bisseling T, Chittenden GJF: Prebiotic nucleotide synthesis - demonstration of a geologically plausible pathway. Orig Life. 1975, 6: 163-168. 10.1007/BF01372401.

    Article  Google Scholar 

  41. Schwartz AW: Phosphorus in prebiotic chemistry. Philos Trans R Soc B. 2006, 361: 1743-1749. 10.1098/rstb.2006.1901.

    Article  Google Scholar 

  42. Pasek MA, Lauretta DS: Aqueous corrosion of phosphide minerals from iron meteorites: a highly reactive source of prebiotic phosphorus on the surface of the early Earth. Astrobiology. 2005, 5: 515-535. 10.1089/ast.2005.5.515.

    Article  Google Scholar 

  43. Pasek MA, Kee TP, Bryant DE, Pavlov AA, Lunine JI: Production of potentially condensed phosphates by phosphorus redox chemistry. Angew Chem Int Ed. 2008, 47: 7918-7920. 10.1002/anie.200802145.

    Article  Google Scholar 

  44. Peacor DR, Dunn PJ, Simmons WB, Wicks FJ: Canaphite, a new sodium calcium phosphate hydrate from the Paterson area, New Jersey. Mineralogical Rec. 1985, 16: 467-468.

    Google Scholar 

  45. Rouse RC, Peacor DR, Freed RL: Pyrophosphate groups in the structure of canaphite, CaNa2P2O7.4H2O: the first occurrence of a condensed phosphate as a mineral. Am Mineral. 1988, 73: 168-171.

    Google Scholar 

  46. Hawthorne FC, Cooper MA, Green DI, Starkey RE, Roberts AC, Grice JD: Wooldrigite, Na2CaCu22+(P2O7)(H2O): a new mineral from Judkins quarry, Warwickshire, England. Mineral Mag. 1999, 63: 13-16. 10.1180/002646199548268.

    Article  Google Scholar 

  47. Van Wazer JR: Phosphorus and its Compounds: Volume I Chemistry. 1965, New York: Interscience

    Google Scholar 

  48. Bodëi S, Buatier M, Steinmann M, Adatte T, Wheat CG: Characterization of metalliferous sediment from a low-temperature hydrothermal environment on the Eastern Flank of the East-Pacific Rise. Mar Geol. 2008, 250: 128-141. 10.1016/j.margeo.2008.01.003.

    Article  Google Scholar 

  49. Koritnig S: Geochemistry of phosphorus: Part I. The replacement of Si4+ by P5+ in rock forming silicate minerals. Geochim Cosmochim Acta. 1965, 29: 361-371. 10.1016/0016-7037(65)90033-5.

    Article  Google Scholar 

  50. Brunet F, Chazot G: Partitioning of phosphorus between olivine, clinopyroxene and silicate glass in a lherzolite xenolith from Yemen. Chem Geol. 2001, 176: 51-72. 10.1016/S0009-2541(00)00351-X.

    Article  Google Scholar 

  51. Milman-Barris MS, Beckett JR, Baker MB, Hofmann A, Morgan Z, Crowley MR, Vielzeuf D, Stolper E: Zoning of phosphorus in igneous olivine. Contrib Mineral Petrol. 2008, 155: 739-765. 10.1007/s00410-007-0268-7.

    Article  Google Scholar 

  52. Wenlan Z, Shao Jian X, Xisheng WR, Lihui C: Mantle metasomatism by P- and F-rich melt/fluids: evidence from phosphate glass in spinel lherzolite xenolith in Keluo, Heilonjiang Province. Chin Sci Bull. 2007, 52: 1827-1835. 10.1007/s11434-007-0223-z.

    Article  Google Scholar 

  53. Seel F, Klos KP, Schuh J: Hydrothermale Kondensation von Magnesium-hydrogenphosphaten zu Magnesiumdiphosphaten. Naturwissenschaften. 1985, 72: 658-10.1007/BF00497440.

    Article  Google Scholar 

  54. Seel F, Klos KP, Rechtenwald D, Schuh J: Non-enzymatic formation of condensed phosphates under prebiotic conditions. Z Naturforsch B. 1986, 41B: 815-824.

    Google Scholar 

  55. Kongshaug KO, Fjellvåg H, Lillerud KP: Synthesis and crystal structure of hydrated magnesium diphosphate Mg2P2O7.3.5H2O and its high temperature variant Mg2P2O7.H2O. Solid State Sci. 2000, 2: 205-214. 10.1016/S1293-2558(00)00133-3.

    Article  Google Scholar 

  56. Petrov AS, Bowman JC, Harvey SC, Williams LD: Bidentate RNA-magnesium clamps: on the origin of the special role of magnesium in RNA folding. RNA. 2011, 17: 291-297. 10.1261/rna.2390311.

    Article  Google Scholar 

  57. De Zwart II, Meade SJ, Pratt AJ: Biomimetic phosphoryl transfer catalysed by iron(II)-mineral precipitates. Geochim Cosmochim Acta. 2004, 68: 4093-4098. 10.1016/j.gca.2004.01.028.

    Article  Google Scholar 

  58. Nitschke W, Russell MJ: Hydrothermal focusing of chemical and chemiosmotic energy, supported by delivery of catalytic Fe, Ni, Mo/W, Co, S and Se, forced life to emerge. J Mol Evol. 2009, 69: 481-496. 10.1007/s00239-009-9289-3.

    Article  Google Scholar 

  59. Schiffman P, Southard RJ: Cation exchange capacity of layer silicates and palagonitized glass in mafic volcanic rocks: a comparative study of bulk extraction and in situ techniques. Clay Clay Miner. 1996, 44: 624-634. 10.1346/CCMN.1996.0440505.

    Article  Google Scholar 

  60. Thilo E: The structural chemistry of condensed inorganic phosphates. Angew Chem Int Ed. 1965, 4: 1061-1071. 10.1002/anie.196510611.

    Article  Google Scholar 

  61. Mitsis I, Economou-Eliopoulos M: On the origin of hydroxyapatite associated with pure massive magnetite in the Othrys ophiolite complex, Greece. Ophioliti. 2003, 28: 25-32.

    Google Scholar 

  62. Wheat CG, Feely RA, Mottl MJ: Phosphate removal by oceanic hydrothermal processes: an update of the phosphorus budget of the oceans. Geochim Cosmochim Acta. 1996, 60: 3593-3608. 10.1016/0016-7037(96)00189-5.

    Article  Google Scholar 

Download references

Acknowledgments

This work was supported by the Stockholm University Faculty of Science and the Stockholm University Astrobiology Centre.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Nils G Holm.

Additional information

Competing interests

The author declares no competing interest.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0), which permits use, duplication, adaptation, distribution, and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Holm, N.G. Glasses as sources of condensed phosphates on the early earth. Geochem Trans 15, 8 (2014). https://doi.org/10.1186/1467-4866-15-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1467-4866-15-8

Keywords