Phytane

From Wikipedia, the free encyclopedia
Skeletal formula of phytane
Names
IUPAC name
2,6,10,14-Tetramethylhexadecane[1]
Identifiers
3D model (JSmol)
1744639
ChEBI
ChemSpider
ECHA InfoCard 100.010.303 Edit this at Wikidata
EC Number
  • 211-332-2
MeSH phytane
UNII
  • InChI=1S/C20H42/c1-7-18(4)12-9-14-20(6)16-10-15-19(5)13-8-11-17(2)3/h17-20H,7-16H2,1-6H3 ☒N
    Key: GGYKPYDKXLHNTI-UHFFFAOYSA-N ☒N
  • CCC(C)CCCC(C)CCCC(C)CCCC(C)C
Properties
C20H42
Molar mass 282.556 g·mol−1
Appearance Colourless liquid
Odor Odourless
Density 791 mg mL−1 (at 20 °C)
Boiling point 301.41 °C (574.54 °F; 574.56 K) at 100 mPa
Related compounds
Related alkanes
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
☒N  (what is checkY☒N ?)
Infobox references

Phytane is the isoprenoid alkane formed when phytol, a constituent of chlorophyll, loses its hydroxyl group.[2] When phytol loses one carbon atom, it yields pristane.[2] Other sources of phytane and pristane have also been proposed than phytol.[3][4]

Pristane and phytane are common constituents in petroleum and have been used as proxies for depositional redox conditions, as well as for correlating oil and its source rock (i.e. elucidating where oil formed). In environmental studies, pristane and phytane are target compounds for investigating oil spills.

Chemistry[]

Phytane is a non-polar organic compound that is a clear and colorless liquid at room temperature. It is a head-to-tail linked regular isoprenoid with chemical formula C20H42.[2]

Phytane has many structural isomers. Among them, crocetane is a tail-to-tail linked isoprenoid and often co-elutes with phytane during gas chromatography (GC) due to its structural similarity.

Phytane also has many stereoisomers because of its three stereo carbons, C-6, C-10 and C-14. Whereas pristane has two stereo carbons, C-6 and C-10. Direct measurement of these isomers has not been reported using gas chromatography.[2]

Chemical structure of an archaeol, with two phytanyl groups.
Chemical structure of α-tocopherol.
Chemical structure of trimethyl 2-methyl-2-(4,8,12-trimethyltridecyl)chroman, a type of MTTCs.

The substituent of phytane is phytanyl. Phytanyl groups are frequently found in archaeal membrane lipids of methanogenic and halophilic archaea[4] (e.g., in archaeol). Phytene is the singly unsaturated version of phytane. Phytene is also found as the functional group phytyl in many organic molecules of biological importance such as chlorophyll, tocopherol (vitamin E), and phylloquinone (vitamin K1). Phytene's corresponding alcohol is phytol. Geranylgeranene is the fully unsaturated form of phytane, and its corresponding substituent is geranylgeranyl.

Sources[]

The major source of phytane and pristane is thought to be chlorophyll.[5] Chlorophyll is one of the most important photosynthetic pigments in plants, algae, and cyanobacteria, and is the most abundant tetrapyrrole in the biosphere.[6] Hydrolysis of chlorophyll a, b, d, and f during diagenesis in marine sediments, or during invertebrate feeding[7] releases phytol, which is then converted to phytane or pristane.

Structure of chlorophyll a, with a side chain containing a phytyl group.

Another possible source of phytane and pristane is archaeal ether lipids. Laboratory studies show that thermal maturation of methanogenic archaea generates pristane and phytane from diphytanyl glyceryl ethers (archaeols).[8][9][10]

In addition, pristane can be derived from tocopherols[11] and methyltrimethyltridecylchromans (MTTCs).[12]

Preservation[]

In suitable environments, biomolecules like chlorophyll can be transformed and preserved in recognizable forms as biomarkers. Conversion during diagenesis often causes the chemical loss of functional groups like double bonds and hydroxyl groups.

Pristane and phytane are formed by diagenesis of phytol under oxic and anoxic conditions, respectively.

Studies suggested that pristane and phytane are formed via diagenesis of phytol under different redox conditions.[13] Pristane can be formed in oxidizing conditions by phytol oxidation to phytenic acid, which may then undergo decarboxylation to pristene, before finally being reduced to pristane. In contrast, phytane is likely from reduction and dehydration of phytol (via dihydrophytol or phytene) under relatively anoxic conditions.[13] However, various biotic and abiotic processes may control the diagenesis of chlorophyll and phytol, and the exact reactions are more complicated and not strictly-correlated to redox conditions.[3][4]

In thermally immature sediments, pristane and phytane has a configuration dominated by 6R,10S stereochemistry (equivalent to 6S, 10R), which is inherited from C-7 and C-11 in phytol. During thermal maturation, isomerization at C-6 and C-10 leads to a mixture of 6R, 10S, 6S, 10S, and 6R, 10R.[2]

Geochemical parameters[]

Pristane/Phytane ratio[]

Pristane/phytane (Pr/Ph) is the ratio of abundances of pristane and phytane. It is a proxy for redox conditions in the depositional environments. The Pr/Ph index is based on the assumption that pristane is formed from phytol by an oxidative pathway, while phytane is generated through various reductive pathways.[13][14] In non-biodegraded crude oil, Pr/Ph less than 0.8 indicates saline to hypersaline conditions associated with evaporite and carbonate deposition, whereas organic-lean terrigenous, fluvial,and deltaic sediments under oxic to suboxic conditions usually generate crude oil with Pr/Ph above 3.[15] Pr/Ph is commonly applied because pristane and phytane are measured easily using gas chromatography.

However, the index should be used with caution, as pristane and phytane may not result from degradation of the same precursor (see *Source*). Also, pristane, but not phytane, can be produced in reducing environments by clay-catalysed degradation of phytol and subsequent reduction.[16] Additionally, during catagenesis, Pr/Ph tends to increase.[17] This variation may be due to preferential release of sulfur-bound phytols from source rocks during early maturation.[18]

Pristane/nC17 and phytane/nC18 ratios[]

Pristane/n-heptadecane (Pr/nC17) and phytane/n-octadecane (Ph/C18) are sometimes used to correlate oil and its source rock (i.e. to elucidate where oil formed). Oils from rocks deposited under open-ocean conditions showed Pr/nC17< 0.5, while those from inland peat swamp had ratios greater than 1.[19]

The ratios should be used with caution for several reasons. Both Pr/nC17and Ph/nC18 decrease with thermal maturity of petroleum because isoprenoids are less thermally stable than linear alkanes. In contrast, biodegradation increases these ratios because aerobic bacteria generally attack linear alkanes before the isoprenoids. Therefore, biodegraded oil is similar to low-maturity non-degraded oil in the sense of exhibiting low abundance of n-alkanes relative to pristane and phytane.[15]

Biodegradation scale[]

Pristane and phytane are more resistant to biodegradation than n-alkanes, but less so than steranes and hopanes. The substantial depletion and complete elimination of pristane and phytane correspond to a Biomarker Biodegradation Scale of 3 and 4, respectively.[20]

Compound specific isotope analyses[]

Carbon isotopes[]

The carbon isotopic composition of pristane and phytane generally reflects the kinetic isotope fractionation that occurs during photosynthesis. For example, δ13C(PDB) of phytane in marine sediments and oils has been used to reconstruct ancient atmospheric CO2levels, which affects the carbon isotopic fractionation associated with photosynthesis, over the past 500 million years.[21] In this study,[21] partial pressure of CO2 reached more than 1000 ppm at maxima compared to 410 ppm today.

Carbon isotope compositions of pristane and phytane in crude oil can also help to constrain their source. Pristane and phytane from a common precursor should have δ13C values differing by no more than 0.3‰.[22]

Hydrogen isotopes[]

Hydrogen isotope composition of phytol in marine phytoplankton and algae starts out as highly depleted, with δD (VSMOW) ranging from -360 to -280‰.[23] Thermal maturation preferentially releases light isotopes, causing and pristane and phytane to become progressively heavier with maturation.

Case study: limitation of Pr/Ph as a redox indicator[]

Inferences from Pr/Ph on the redox potential of source sediments should always be supported by other geochemical and geological data, such as sulfur content or the C35 homohopane index (i.e. the abundance of C35 homohopane relative to that of C31-C35 homohopanes). For example, the Baghewala-1 oil from India has low Pr/Ph (0.9), high sulfur (1.2 wt.%) and high C35 homohopane index, which are consistent with anoxia during deposition of the source rock.[24]

However, drawing conclusion on the oxic state of depositional environments only from Pr/Ph ratio can be misleading because salinity often controls the Pr/Ph in hypersaline environments. In another example, the decrease in Pr/Ph during deposition of the PermianKupferschiefer sequence in Germany is in coincidence with an increase in trimethylated 2-methyl-2-(4,8,12-trimethyltridecyl)chromans, an aromatic compound believed to be markers of salinity.[25] Therefore, this decrease in Pr/Ph should indicate an increase in salinity, instead of an increase in anoxia.

See also[]

References[]

  1. ^ "phytane - Compound Summary". PubChem Compound. USA: National Center for Biotechnology Information. 27 March 2005. Identification and Related Records. Retrieved 14 March 2012.
  2. ^ a b c d e Moldowan, J. M.; Walters, C. C.; Peters, K. E. (December 2004). "Organic chemistry". The Biomarker Guide. The Biomarker Guide. pp. 18–44. doi:10.1017/CBO9780511524868.004. ISBN 9780511524868.
  3. ^ a b Rontani, Jean-François; Volkman, John K. (2003-01-01). "Phytol degradation products as biogeochemical tracers in aquatic environments". Organic Geochemistry. 34 (1): 1–35. doi:10.1016/S0146-6380(02)00185-7. ISSN 0146-6380.
  4. ^ a b c Rontani, Jean-François; Bonin, Patricia (November 2011). "Production of pristane and phytane in the marine environment: role of prokaryotes". Research in Microbiology. 162 (9): 923–933. doi:10.1016/j.resmic.2011.01.012. PMID 21288485.
  5. ^ Dean, R. A.; Whitehead, E. V. (1961-01-01). "The occurrence of phytane in petroleum". Tetrahedron Letters. 2 (21): 768–770. doi:10.1016/S0040-4039(01)99264-0. ISSN 0040-4039.
  6. ^ Baker, E.W.; Louda, J.W. (1986). "Porphyrins in the geological record". In Johns, R.B. (ed.). Biological Markers in the Sedimentary Record. Elsevier. pp. 125–224.
  7. ^ Blumer, Max; Avigan, Joel (1968-05-01). "On the origin of pristane in marine organisms". Journal of Lipid Research. 9 (3): 350–352. ISSN 0022-2275. PMID 5646185.
  8. ^ Rowland, S. J. (1990-01-01). "Production of acyclic isoprenoid hydrocarbons by laboratory maturation of methanogenic bacteria". Organic Geochemistry. 15 (1): 9–16. doi:10.1016/0146-6380(90)90181-X. ISSN 0146-6380.
  9. ^ Navale, Vivek (1994-06-01). "Comparative study of low and high temperature hydrous pyrolysis products of monoglycerol diether lipid from archaebacteria". Journal of Analytical and Applied Pyrolysis. 29 (1): 33–43. doi:10.1016/0165-2370(93)00786-M. ISSN 0165-2370.
  10. ^ Pease, T. K.; Van Vleet, E. S.; Barre, J. S.; Dickins, H. D. (1998-01-01). "Simulated degradation of glyceryl ethers by hydrous and flash pyrolysis". Organic Geochemistry. 29 (4): 979–988. doi:10.1016/S0146-6380(98)00047-3. ISSN 0146-6380.
  11. ^ Brassell, S. C.; P. A. Schenck; de Leeuw, J. W.; Goossens, H. (November 1984). "Tocopherols as likely precursors of pristane in ancient sediments and crude oils". Nature. 312 (5993): 440–442. Bibcode:1984Natur.312..440G. doi:10.1038/312440a0. ISSN 1476-4687. S2CID 4329068.
  12. ^ Li, Maowen; Larter, Steve R.; Taylor, Paul; Jones, D. Martin; Bowler, Bernard; Bjorøy, Malvin (1995-02-01). "Biomarkers or not biomarkers? A new hypothesis for the origin of pristane involving derivation from methyltrimethyltridecylchromans (MTTCs) formed during diagenesis from chlorophyll and alkylphenols". Organic Geochemistry. 23 (2): 159–167. doi:10.1016/0146-6380(94)00112-E. ISSN 0146-6380.
  13. ^ a b c Eglinton, G.; S. C. Brassell; Simoneit, B. R. T.; Didyk, B. M. (March 1978). "Organic geochemical indicators of palaeoenvironmental conditions of sedimentation". Nature. 272 (5650): 216–222. Bibcode:1978Natur.272..216D. doi:10.1038/272216a0. ISSN 1476-4687. S2CID 128737515.
  14. ^ D. M. McKIRDY; Powell, T. G. (May 1973). "Relationship between Ratio of Pristane to Phytane, Crude Oil Composition and Geological Environment in Australia". Nature Physical Science. 243 (124): 37–39. Bibcode:1973NPhS..243...37P. doi:10.1038/physci243037a0. ISSN 2058-1106.
  15. ^ a b Peters, K. E.; Walters, C. C.; Moldowan, J. M. (2004), "Source- and age-related biomarker parameters", The Biomarker Guide, Cambridge University Press, pp. 483–607, doi:10.1017/cbo9781107326040.004, ISBN 9781107326040
  16. ^ Schenck, P. A.; Lange, F. de; Boon, J. J.; Rijpstra, C.; Irene, W.; Leeuw, J. W. de (1977). "relationship between lipids from Fontinalis antipyretica, its detritus and the underlying sediment: the fate of waxesters and sterolesters". Interactions Between Sediments and Fresh Water; Proceedings of an International Symposium.
  17. ^ VOLKMAN, J. K. (1986). "Acyclic isoprenoids as biological markers". Biological Markers in the Sedimentary Record.: 1817–1828.
  18. ^ De Graaf, Wim; Damsté, Jaap S. Sinninghe; de Leeuw, Jan W (1992-12-01). "Laboratory simulation of natural sulphurization: I. Formation of monomeric and oligomeric isoprenoid polysulphides by low-temperature reactions of inorganic polysulphides with phytol and phytadienes". Geochimica et Cosmochimica Acta. 56 (12): 4321–4328. Bibcode:1992GeCoA..56.4321D. doi:10.1016/0016-7037(92)90275-N. ISSN 0016-7037.
  19. ^ Lijmbach, W. M. (1975-01-01). "SP (1) On the Origin of Petroleum". World Petroleum Congress. Cite journal requires |journal= (help)
  20. ^ Peters, K. E.; Walters, C. C.; Moldowan, J. M. (2004), "Biodegradation parameters", The Biomarker Guide, Cambridge University Press, pp. 645–708, doi:10.1017/cbo9781107326040.007, ISBN 9781107326040
  21. ^ a b Damsté, Jaap S. Sinninghe; Schouten, Stefan; Blais, Brian; Weijers, Johan W. H.; Witkowski, Caitlyn R. (2018-11-01). "Molecular fossils from phytoplankton reveal secular Pco2 trend over the Phanerozoic". Science Advances. 4 (11): eaat4556. Bibcode:2018SciA....4.4556W. doi:10.1126/sciadv.aat4556. ISSN 2375-2548. PMC 6261654. PMID 30498776.
  22. ^ Hayes, J. M.; Freeman, Katherine H.; Popp, Brian N.; Hoham, Christopher H. (1990-01-01). "Compound-specific isotopic analyses: A novel tool for reconstruction of ancient biogeochemical processes". Organic Geochemistry. Proceedings of the 14th International Meeting on Organic Geochemistry. 16 (4): 1115–1128. doi:10.1016/0146-6380(90)90147-R. ISSN 0146-6380. PMID 11540919.
  23. ^ Sessions, Alex L.; Burgoyne, Thomas W.; Schimmelmann, Arndt; Hayes, John M. (1999-09-01). "Fractionation of hydrogen isotopes in lipid biosynthesis". Organic Geochemistry. 30 (9): 1193–1200. doi:10.1016/S0146-6380(99)00094-7. ISSN 0146-6380.
  24. ^ K. E. Peters (2), M. E. Clark (3) (1995). "Recognition of an Infracambrian Source Rock Based on Biomarkers in the Baghewala-1 Oil, India". AAPG Bulletin. 79 (10). doi:10.1306/7834da12-1721-11d7-8645000102c1865d. ISSN 0149-1423.
  25. ^ Schwark, L; Vliex, M; Schaeffer, P (1998-12-01). "Geochemical characterization of Malm Zeta laminated carbonates from the Franconian Alb, SW-Germany (II)". Organic Geochemistry. 29 (8): 1921–1952. doi:10.1016/S0146-6380(98)00192-2. ISSN 0146-6380.
Retrieved from ""