Mass-independent fractionation

Last updated

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.

Contents

Mass-independent fractionation in nature

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

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

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

Related Research Articles

<span class="mw-page-title-main">Isotope analysis</span> Analytical technique used to study isotopes

Isotope analysis is the identification of isotopic signature, abundance of certain stable isotopes of chemical elements within organic and inorganic compounds. Isotopic analysis can be used to understand the flow of energy through a food web, to reconstruct past environmental and climatic conditions, to investigate human and animal diets, for food authentification, and a variety of other physical, geological, palaeontological and chemical processes. Stable isotope ratios are measured using mass spectrometry, which separates the different isotopes of an element on the basis of their mass-to-charge ratio.

<span class="mw-page-title-main">Magnesite</span> Type of mineral

Magnesite is a mineral with the chemical formula MgCO
3
. Iron, manganese, cobalt, and nickel may occur as admixtures, but only in small amounts.

Isotope geochemistry is an aspect of geology based upon the study of natural variations in the relative abundances of isotopes of various elements. Variations in isotopic abundance are measured by isotope ratio mass spectrometry, and can reveal information about the ages and origins of rock, air or water bodies, or processes of mixing between them.

<span class="mw-page-title-main">Sulfur cycle</span> Biogeochemical cycle of sulfur

The sulfur cycle is a biogeochemical cycle in which the sulfur moves between rocks, waterways and living systems. It is important in geology as it affects many minerals and in life because sulfur is an essential element (CHNOPS), being a constituent of many proteins and cofactors, and sulfur compounds can be used as oxidants or reductants in microbial respiration. The global sulfur cycle involves the transformations of sulfur species through different oxidation states, which play an important role in both geological and biological processes. Steps of the sulfur cycle are:

Kinetic fractionation is an isotopic fractionation process that separates stable isotopes from each other by their mass during unidirectional processes. Biological processes are generally unidirectional and are very good examples of "kinetic" isotope reactions. All organisms preferentially use lighter isotopic species, because "energy costs" are lower, resulting in a significant fractionation between the substrate (heavier) and the biologically mediated product (lighter). As an example, photosynthesis preferentially takes up the light isotope of carbon 12C during assimilation of an atmospheric CO2 molecule. This kinetic isotope fractionation explains why plant material (and thus fossil fuels, which are derived from plants) is typically depleted in 13C by 25 per mil (2.5 per cent) relative to most inorganic carbon on Earth.

An isotopic signature is a ratio of non-radiogenic 'stable isotopes', stable radiogenic isotopes, or unstable radioactive isotopes of particular elements in an investigated material. The ratios of isotopes in a sample material are measured by isotope-ratio mass spectrometry against an isotopic reference material. This process is called isotope analysis.

<span class="mw-page-title-main">Oxygen isotope ratio cycle</span> Cyclical variations in the ratio of the abundance of oxygen

Oxygen isotope ratio cycles are cyclical variations in the ratio of the abundance of oxygen with an atomic mass of 18 to the abundance of oxygen with an atomic mass of 16 present in some substances, such as polar ice or calcite in ocean core samples, measured with the isotope fractionation. The ratio is linked to ancient ocean temperature which in turn reflects ancient climate. Cycles in the ratio mirror climate changes in the geological history of Earth.

The environmental isotopes are a subset of isotopes, both stable and radioactive, which are the object of isotope geochemistry. They are primarily used as tracers to see how things move around within the ocean-atmosphere system, within terrestrial biomes, within the Earth's surface, and between these broad domains.

<span class="mw-page-title-main">Atmosphere of Mars</span> Layer of gases surrounding planet Mars

The atmosphere of Mars is the layer of gases surrounding Mars. It is primarily composed of carbon dioxide (95%), molecular nitrogen (2.8%), and argon (2%). It also contains trace levels of water vapor, oxygen, carbon monoxide, hydrogen, and noble gases. The atmosphere of Mars is much thinner than Earth's. The average surface pressure is only about 610 pascals (0.088 psi) which is less than 1% of the Earth's value. The currently thin Martian atmosphere prohibits the existence of liquid water on the surface of Mars, but many studies suggest that the Martian atmosphere was much thicker in the past. The higher density during spring and fall is reduced by 25% during the winter when carbon dioxide partly freezes at the pole caps. The highest atmospheric density on Mars is equal to the density found 35 km (22 mi) above the Earth's surface and is ≈0.020 kg/m3. The atmosphere of Mars has been losing mass to space since the planet's core slowed down, and the leakage of gases still continues today. The atmosphere of Mars is colder than Earth's. Owing to the larger distance from the Sun, Mars receives less solar energy and has a lower effective temperature, which is about 210 K. The average surface emission temperature of Mars is just 215 K, which is comparable to inland Antarctica. Although Mars' atmosphere consists primarily of carbon dioxide, the greenhouse effect in the Martian atmosphere is much weaker than Earth's: 5 °C (9.0 °F) on Mars, versus 33 °C (59 °F) on Earth. This is because the total atmosphere is so thin that the partial pressure of carbon dioxide is very weak, leading to less warming. The daily range of temperature in the lower atmosphere is huge due to the low thermal inertia; it can range from −75 °C (−103 °F) to near 0 °C (32 °F) near the surface in some regions. The temperature of the upper part of the Martian atmosphere is also significantly lower than Earth's because of the absence of stratospheric ozone and the radiative cooling effect of carbon dioxide at higher altitudes.

In geochemistry, paleoclimatology and paleoceanography δ18O or delta-O-18 is a measure of the ratio of stable isotopes oxygen-18 (18O) and oxygen-16 (16O). It is commonly used as a measure of the temperature of precipitation, as a measure of groundwater/mineral interactions, and as an indicator of processes that show isotopic fractionation, like methanogenesis. In paleosciences, 18O:16O data from corals, foraminifera and ice cores are used as a proxy for temperature.

<span class="mw-page-title-main">Isotope fractionation</span> Processes for the separation of isotopes

Isotope fractionation describes fractionation processes that affect the relative abundance of isotopes, phenomena which are taken advantage of in isotope geochemistry and other fields. Normally, the focus is on stable isotopes of the same element. Isotopic fractionation can be measured by isotope analysis, using isotope-ratio mass spectrometry or cavity ring-down spectroscopy to measure ratios of isotopes, an important tool to understand geochemical and biological systems. For example, biochemical processes cause changes in ratios of stable carbon isotopes incorporated into biomass.

Oxygen-17 (17O) is a low-abundance, natural, stable isotope of oxygen.

The Dole effect, named after Malcolm Dole, describes an inequality in the ratio of the heavy isotope 18O to the lighter 16O, measured in the atmosphere and seawater. This ratio is usually denoted δ18O.

Equilibrium isotope fractionation is the partial separation of isotopes between two or more substances in chemical equilibrium. Equilibrium fractionation is strongest at low temperatures, and forms the basis of the most widely used isotopic paleothermometers : D/H and 18O/16O records from ice cores, and 18O/16O records from calcium carbonate. It is thus important for the construction of geologic temperature records. Isotopic fractionations attributed to equilibrium processes have been observed in many elements, from hydrogen (D/H) to uranium (238U/235U). In general, the light elements are most susceptible to fractionation, and their isotopes tend to be separated to a greater degree than heavier elements.

<span class="mw-page-title-main">Global meteoric water line</span>

The Global Meteoric Water Line (GMWL) describes the global annual average relationship between hydrogen and oxygen isotope (Oxygen-18 and Deuterium) ratios in natural meteoric waters. The GMWL was first developed in 1961 by Harmon Craig, and has subsequently been widely used to track water masses in environmental geochemistry and hydrogeology.

Clumped isotopes are heavy isotopes that are bonded to other heavy isotopes. The relative abundance of clumped isotopes (and multiply-substituted isotopologues) in molecules such as methane, nitrous oxide, and carbonate is an area of active investigation. The carbonate clumped-isotope thermometer, or "13C–18O order/disorder carbonate thermometer", is a new approach for paleoclimate reconstruction, based on the temperature dependence of the clumping of 13C and 18O into bonds within the carbonate mineral lattice. This approach has the advantage that the 18O ratio in water is not necessary (different from the δ18O approach), but for precise paleotemperature estimation, it also needs very large and uncontaminated samples, long analytical runs, and extensive replication. Commonly used sample sources for paleoclimatological work include corals, otoliths, gastropods, tufa, bivalves, and foraminifera. Results are usually expressed as Δ47 (said as "cap 47"), which is the deviation of the ratio of isotopologues of CO2 with a molecular weight of 47 to those with a weight of 44 from the ratio expected if they were randomly distributed.

Carbonate-associated sulfates (CAS) are sulfate species found in association with carbonate minerals, either as inclusions, adsorbed phases, or in distorted sites within the carbonate mineral lattice. It is derived primarily from dissolved sulfate in the solution from which the carbonate precipitates. In the ocean, the source of this sulfate is a combination of riverine and atmospheric inputs, as well as the products of marine hydrothermal reactions and biomass remineralisation. CAS is a common component of most carbonate rocks, having concentrations in the parts per thousand within biogenic carbonates and parts per million within abiogenic carbonates. Through its abundance and sulfur isotope composition, it provides a valuable record of the global sulfur cycle across time and space.

<span class="mw-page-title-main">Microbial oxidation of sulfur</span>

Microbial oxidation of sulfur is the oxidation of sulfur by microorganisms to build their structural components. The oxidation of inorganic compounds is the strategy primarily used by chemolithotrophic microorganisms to obtain energy to survive, grow and reproduce. Some inorganic forms of reduced sulfur, mainly sulfide (H2S/HS) and elemental sulfur (S0), can be oxidized by chemolithotrophic sulfur-oxidizing prokaryotes, usually coupled to the reduction of oxygen (O2) or nitrate (NO3). Anaerobic sulfur oxidizers include photolithoautotrophs that obtain their energy from sunlight, hydrogen from sulfide, and carbon from carbon dioxide (CO2).

Mark Howard Thiemens is a distinguished professor and the John Doves Isaacs Endowed Chair in Natural Philosophy of Physical Sciences in the department of chemistry and biochemistry at the University of California San Diego. He is best known for the discovery of a new physical chemical phenomena termed the mass independent isotope effect.

Sulfur isotope biogeochemistry is the study of the distribution of sulfur isotopes in biological and geological materials. In addition to its common isotope, 32S, sulfur has three rare stable isotopes: 34S, 36S, and 33S. The distribution of these isotopes in the environment is controlled by many biochemical and physical processes, including biological metabolisms, mineral formation processes, and atmospheric chemistry. Measuring the abundance of sulfur stable isotopes in natural materials, like bacterial cultures, minerals, or seawater, can reveal information about these processes both in the modern environment and over Earth history.

References

  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.