Mass-independent fractionation

From Wikipedia, the free encyclopedia

Mass-independent isotope fractionation or Non-mass-dependent fractionation (NMD),[1] refers to any chemical or physical process that acts to separate isotopes, where the amount of separation does not scale in proportion with the difference in the masses of the isotopes. Most isotopic fractionations (including typical kinetic fractionations and equilibrium fractionations) are caused by the effects of the mass of an isotope on atomic or molecular velocities, diffusivities or bond strengths. Mass-independent fractionation processes are less common, occurring mainly in photochemical and spin-forbidden reactions. Observation of mass-independently fractionated materials can therefore be used to trace these types of reactions in nature and in laboratory experiments.

Mass-independent fractionation in nature[edit]

The most notable examples of mass-independent fractionation in nature are found in the isotopes of oxygen and sulfur. The first example was discovered by Robert N. Clayton, Toshiko Mayeda, and Lawrence Grossman in 1973,[2] in the oxygen isotopic composition of refractory calcium–aluminium-rich inclusions in the Allende meteorite. The inclusions, thought to be among the oldest solid materials in the Solar System, show a pattern of low 18O/16O and 17O/16O relative to samples from the Earth and Moon. Both ratios vary by the same amount in the inclusions, although the mass difference between 18O and 16O is almost twice as large as the difference between 17O and 16O. Originally this was interpreted as evidence of incomplete mixing of 16O-rich material (created and distributed by a large star in a supernova) into the Solar nebula. However, recent measurement of the oxygen-isotope composition of the Solar wind, using samples collected by the Genesis spacecraft, shows that the most 16O-rich inclusions are close to the bulk composition of the solar system. This implies that Earth, the Moon, Mars, and asteroids all formed from 18O- and 17O-enriched material. Photodissociation of carbon monoxide in the Solar nebula has been proposed to explain this isotope fractionation.

Mass-independent fractionation also has been observed in ozone. Large, 1:1 enrichments of 18O/16O and 17O/16O in ozone were discovered in laboratory synthesis experiments by Mark Thiemens and John Heidenreich in 1983,[3] and later found in stratospheric air samples measured by Konrad Mauersberger.[4] These enrichments were eventually traced to the three-body ozone formation reaction.[5]

O + O2 → O3* + M → O3 + M*

Theoretical calculations[6] by Rudolph Marcus and others suggest that the enrichments are the result of a combination of mass-dependent and mass-independent kinetic isotope effects (KIE) involving the excited state O3* intermediate related to some unusual symmetry properties. The mass-dependent isotope effect occurs in asymmetric species, and arises from the difference in zero-point energy of the two formation channels available (e.g., 18O16O + 16O vs 18O + 16O16O for formation of 18O16O16O.) These mass-dependent zero-point energy effects cancel one another out and do not affect the enrichment in heavy isotopes observed in ozone.[7] The mass-independent enrichment in ozone is still not fully understood, but may be due to isotopically symmetric O3* having a shorter lifetime than asymmetric O3*, thus not allowing a statistical distribution of energy throughout all the degrees of freedom, resulting in a mass-independent distribution of isotopes.

Mass-independent carbon dioxide fractionation[edit]

The mass-independent distribution of isotopes in stratospheric ozone can be transferred to carbon dioxide (CO2).[8] This anomalous isotopic composition in CO2 can be used to quantify gross primary production, the uptake of CO2 by vegetation through photosynthesis. This effect of terrestrial vegetation on the isotopic signature of atmospheric CO2 was simulated with a global model[9] and confirmed experimentally.[10]

Mass-independent sulfur fractionation[edit]

Mass-independent fractionation of sulfur can be observed in ancient sediments,[11] where it preserves a signal of the prevailing environmental conditions. The creation and transfer of the mass-independent signature into minerals would be unlikely in an atmosphere containing abundant oxygen, constraining the Great Oxygenation Event to some time after 2,450 million years ago. Prior to this time, the MIS record implies that sulfate-reducing bacteria did not play a significant role in the global sulfur cycle, and that the MIS signal is due primarily to changes in volcanic activity.[12]

See also[edit]

References[edit]

  1. ^ Timothy W. Lyons; Christopher T. Reinhard; Noah J. Planavsky (February 19, 2014). "The rise of oxygen in Earth's early ocean and atmosphere". Nature. 506 (7488): 307–315. Bibcode:2014Natur.506..307L. doi:10.1038/nature13068. PMID 24553238. S2CID 4443958. The disappearance of distinctive non-mass-dependent (NMD) sulphur isotope fractionations in sedimentary rocks deposited after about 2.4–2.3 Gyr ago16 (Fig. 2). Almost all fractionations among isotopes of a given element scale to differences in their masses; NMD fractionations deviate from this typical behaviour. The remarkable NMD signals are tied to photochemical reactions at short wavelengths involving gaseous sulphur compounds released from volcanoes into the atmosphere.
  2. ^ Clayton, R. N.; Grossman, L.; Mayeda, T. K. (1973). "A Component of Primitive Nuclear Composition in Carbonaceous Meteorites". Science. 182 (4111): 485–488. Bibcode:1973Sci...182..485C. doi:10.1126/science.182.4111.485. PMID 17832468. S2CID 22386977.
  3. ^ Thiemens, M. H.; Heidenreich, J. E. (1983). "The Mass-Independent Fractionation of Oxygen: A Novel Isotope Effect and Its Possible Cosmochemical Implications". Science. 219 (4588): 1073–1075. Bibcode:1983Sci...219.1073T. doi:10.1126/science.219.4588.1073. PMID 17811750. S2CID 26466899.
  4. ^ Mauersberger, K (1987). "Ozone isotope measurements in the stratosphere". Geophysical Research Letters. 14 (1): 80–83. Bibcode:1987GeoRL..14...80M. doi:10.1029/gl014i001p00080.
  5. ^ Morton, J.; Barnes, J.; Schueler, B.; Mauersberger, K. (1990). "Laboratory Studies of Heavy Ozone". Journal of Geophysical Research. 95 (D1): 901. Bibcode:1990JGR....95..901M. doi:10.1029/JD095iD01p00901.
  6. ^ Gao, Y.; Marcus, R. (2001). "Strange and unconventional isotope effects in ozone formation". Science. 293 (5528): 259–263. Bibcode:2001Sci...293..259G. doi:10.1126/science.1058528. PMID 11387441. S2CID 867229.
  7. ^ Janssen, Carl (2001). "Kinetic origin of the ozone isotope effect: a critical analysis of enrichments and rate coefficients". Physical Chemistry Chemical Physics. 3 (21): 4718. Bibcode:2001PCCP....3.4718J. doi:10.1039/b107171h.
  8. ^ Yung, Y. L.; DeMore, W. B.; Pinto, J. P. (1991). "Isotopic exchange between carbon dioxide and ozone via O(1D) in the stratosphere". Geophysical Research Letters. 18 (1): 13–16. Bibcode:1991GeoRL..18...13Y. doi:10.1029/90GL02478. PMID 11538378.
  9. ^ Koren, G.; Schneider, L.; Velde, I. R.; Schaik, E.; Gromov, S. S.; Adnew, G. A.; Mrozek Martino, D. J.; Hofmann, M. E. G.; Liang, M.‐C.; Mahata, S.; Bergamaschi, P.; Laan‐Luijkx, I. T.; Krol, M. C.; Röckmann, T.; Peters, W. (16 August 2019). "Global 3‐D Simulations of the Triple Oxygen Isotope Signature Δ17O in Atmospheric CO2". Journal of Geophysical Research: Atmospheres. 124 (15): 8808–8836. Bibcode:2019JGRD..124.8808K. doi:10.1029/2019JD030387. PMC 6774299. PMID 31598450.
  10. ^ Adnew, G. A.; Pons, T. L.; Koren, G.; Peters, W.; Röckmann, T. (31 July 2020). "Leaf-scale quantification of the effect of photosynthetic gas exchange on Δ17O of atmospheric CO2". Biogeosciences. 17 (14): 3903–3922. Bibcode:2020BGeo...17.3903A. doi:10.5194/bg-17-3903-2020.
  11. ^ Farquhar, J.; Bao, H.; Thiemens, M. (2000). "Atmospheric Influence of Earth's Earliest Sulfur Cycle". Science. 289 (5480): 756–758. Bibcode:2000Sci...289..756F. doi:10.1126/science.289.5480.756. PMID 10926533. S2CID 12287304.
  12. ^ Halevy, I.; Johnston, D.; Schrag, D. (2010). "Explaining the structure of the Archean mass-independent sulfur isotope record". Science. 329 (5988): 204–207. Bibcode:2010Sci...329..204H. doi:10.1126/science.1190298. PMID 20508089. S2CID 45825809.