Isotopic signature

From Wikipedia, the free encyclopedia

An isotopic signature (also isotopic fingerprint) 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.

Stable isotopes[edit]

The atomic mass of different isotopes affect their chemical kinetic behavior, leading to natural isotope separation processes.

Carbon isotopes[edit]

Algal group δ13C range[1]
HCO3-using red algae −22.5‰ to −9.6‰
CO2-using red algae −34.5‰ to −29.9‰
Brown algae −20.8‰ to −10.5‰
Green algae −20.3‰ to −8.8‰

For example, different sources and sinks of methane have different affinity for the 12C and 13C isotopes, which allows distinguishing between different sources by the 13C/12C ratio in methane in the air. In geochemistry, paleoclimatology and paleoceanography this ratio is called δ13C. The ratio is calculated with respect to Pee Dee Belemnite (PDB) standard:

Similarly, carbon in inorganic carbonates shows little isotopic fractionation, while carbon in materials originated by photosynthesis is depleted of the heavier isotopes. In addition, there are two types of plants with different biochemical pathways; the C3 carbon fixation, where the isotope separation effect is more pronounced, C4 carbon fixation, where the heavier 13C is less depleted, and Crassulacean Acid Metabolism (CAM) plants, where the effect is similar but less pronounced than with C4 plants. Isotopic fractionation in plants is caused by physical (slower diffusion of 13C in plant tissues due to increased atomic weight) and biochemical (preference of 12C by two enzymes: RuBisCO and phosphoenolpyruvate carboxylase) factors.[2] The different isotope ratios for the two kinds of plants propagate through the food chain, thus it is possible to determine if the principal diet of a human or an animal consists primarily of C3 plants (rice, wheat, soybeans, potatoes) or C4 plants (corn, or corn-fed beef) by isotope analysis of their flesh and bone collagen (however, to obtain more accurate determinations, carbon isotopic fractionation must be also taken into account, since several studies have reported significant 13C discrimination during biodegradation of simple and complex substrates).[3][4] Within C3 plants processes regulating changes in δ13C are well understood, particularly at the leaf level,[5] but also during wood formation.[6][7] Many recent studies combine leaf level isotopic fractionation with annual patterns of wood formation (i.e. tree ring δ13C) to quantify the impacts of climatic variations and atmospheric composition[8] on physiological processes of individual trees and forest stands.[9] The next phase of understanding, in terrestrial ecosystems at least, seems to be the combination of multiple isotopic proxies to decipher interactions between plants, soils and the atmosphere, and predict how changes in land use will affect climate change.[10] Similarly, marine fish contain more 13C than freshwater fish, with values approximating the C4 and C3 plants respectively.

The ratio of carbon-13 and carbon-12 isotopes in these types of plants is as follows:[11]

  • C4 plants: -16 to -10 ‰
  • CAM plants: -20 to -10 ‰
  • C3 plants: -33 to -24 ‰

Limestones formed by precipitation in seas from the atmospheric carbon dioxide contain normal proportion of 13C. Conversely, calcite found in salt domes originates from carbon dioxide formed by oxidation of petroleum, which due to its plant origin is 13C-depleted. The layer of limestone deposited at the Permian extinction 252 Mya can be identified by the 1% drop in 13C/12C.

The 14C isotope is important in distinguishing biosynthetized materials from man-made ones. Biogenic chemicals are derived from biospheric carbon, which contains 14C. Carbon in artificially made chemicals is usually derived from fossil fuels like coal or petroleum, where the 14C originally present has decayed below detectable limits. The amount of 14C currently present in a sample therefore indicates the proportion of carbon of biogenic origin.

Nitrogen isotopes[edit]

Nitrogen-15, or 15N, is often used in agricultural and medical research, for example in the Meselson–Stahl experiment to establish the nature of DNA replication.[12] An extension of this research resulted in development of DNA-based stable-isotope probing, which allows examination of links between metabolic function and taxonomic identity of microorganisms in the environment, without the need for culture isolation.[13][14] Proteins can be isotopically labelled by cultivating them in a medium containing 15N as the only source of nitrogen, e.g., in quantitative proteomics such as SILAC.

Nitrogen-15 is extensively used to trace mineral nitrogen compounds (particularly fertilizers) in the environment. When combined with the use of other isotopic labels, 15N is also a very important tracer for describing the fate of nitrogenous organic pollutants.[15][16] Nitrogen-15 tracing is an important method used in biogeochemistry.

The ratio of stable nitrogen isotopes, 15N/14N or δ15N, tends to increase with trophic level, such that herbivores have higher nitrogen isotope values than plants, and carnivores have higher nitrogen isotope values than herbivores. Depending on the tissue being examined, there tends to be an increase of 3-4 parts per thousand with each increase in trophic level.[17] The tissues and hair of vegans therefore contain significantly lower δ15N than the bodies of people who eat mostly meat. Similarly, a terrestrial diet produces a different signature than a marine-based diet. Isotopic analysis of hair is an important source of information for archaeologists, providing clues about the ancient diets and differing cultural attitudes to food sources.[18]

A number of other environmental and physiological factors can influence the nitrogen isotopic composition at the base of the food web (i.e. in plants) or at the level of individual animals. For example, in arid regions, the nitrogen cycle tends to be more 'open' and prone to the loss of 14N, increasing δ15N in soils and plants.[19] This leads to relatively high δ15N values in plants and animals in hot and arid ecosystems relative to cooler and moister ecosystems.[20] Furthermore, elevated δ15N have been linked to the preferential excretion of 14N and reutilization of already enriched 15N tissues in the body under prolonged water stress conditions or insufficient protein intake.[21][22]

δ15N also provides a diagnostic tool in planetary science as the ratio exhibited in atmospheres and surface materials "is closely tied to the conditions under which materials form".[23]

Oxygen isotopes[edit]

Oxygen comes in three variants, but the 17O is so rare that it is very difficult to detect (~0.04% abundant).[24] The ratio of 18O/16O in water depends on the amount of evaporation the water experienced (as 18O is heavier and therefore less likely to vaporize). As the vapor tension depends on the concentration of dissolved salts, the 18O/16O ratio shows correlation on the salinity and temperature of water. As oxygen gets built into the shells of calcium carbonate secreting organisms, such sediments provide a chronological record of temperature and salinity of the water in the area.

Oxygen isotope ratio in atmosphere varies predictably with time of year and geographic location; e.g. there is a 2% difference between 18O-rich precipitation in Montana and 18O-depleted precipitation in Florida Keys. This variability can be used for approximate determination of geographic location of origin of a material; e.g. it is possible to determine where a shipment of uranium oxide was produced. The rate of exchange of surface isotopes with the environment has to be taken in account.[25]

The oxygen isotopic signatures of solid samples (organic and inorganic) are usually measured with pyrolysis and mass spectrometry.[26] Researchers need to avoid improper or prolonged storage of the samples for accurate measurements.[26]

Sulfur Isotopes[edit]

Sulfur has four stable isotopes, 32S, 33S, 34S, and 36S, of which 32S is the most abundant by a large margin due to the fact it is created by the very common 12C in supernovas. Sulfur isotope ratios are almost always expressed as ratios relative to 32S due to this major relative abundance (95.0%). Sulfur isotope fractionations are usually measured in terms of δ34S due to its higher abundance (4.25%) compared to the other stable isotopes of sulfur, though δ33S is also sometimes measured. Differences in sulfur isotope ratios are thought to exist primarily due to kinetic fractionation during reactions and transformations.

Sulfur isotopes are generally measured against standards; prior to 1993, the Canyon Diablo troilite standard (abbreviated to CDT), which has a 32S:34S equal to 22.220, was used as both a reference material and the zero point for the isotopic scale. Since 1993, the Vienna-CDT standard has been used as a zero point, and there are several materials used as reference materials for sulfur isotope measurements. Sulfur fractionations by natural processes measured against these standards have been shown to exist between -72‰ and +147‰,[27][28] as calculated by the following equation:

Natural sulfur isotope values
Natural Source δ34S range
Petroleum[29] -32‰ to -8‰
River water[30] -8‰ to 10‰
Lunar rocks[30] -2‰ to 2.5‰
Meteorites[30] 0‰ to 2‰
Ocean water[30] 17‰ to 20‰
Most relevant isotopes of sulfur
Isotope Abundance Half-life
32S 94.99% Stable
33S 0.75% Stable
34S 4.25% Stable
35S <0.1% 87.4 days
36S 0.01% Stable

As a very redox-active element, sulfur can be useful for recording major chemistry-altering events throughout Earth's history, such as marine evaporites which reflect the change in the atmosphere's redox state brought about by the Oxygen Crisis.[31][32]

Radiogenic isotopes[edit]

Lead isotopes[edit]

Lead consists of four stable isotopes: 204Pb, 206Pb, 207Pb, and 208Pb. Local variations in uranium/thorium/lead content cause a wide location-specific variation of isotopic ratios for lead from different localities. Lead emitted to the atmosphere by industrial processes has an isotopic composition different from lead in minerals. Combustion of gasoline with tetraethyllead additive led to formation of ubiquitous micrometer-sized lead-rich particulates in car exhaust smoke; especially in urban areas the man-made lead particles are much more common than natural ones. The differences in isotopic content in particles found in objects can be used for approximate geolocation of the object's origin.[25]

Radioactive isotopes[edit]

Hot particles, radioactive particles of nuclear fallout and radioactive waste, also exhibit distinct isotopic signatures. Their radionuclide composition (and thus their age and origin) can be determined by mass spectrometry or by gamma spectrometry. For example, particles generated by a nuclear blast contain detectable amounts of 60Co and 152Eu. The Chernobyl accident did not release these particles but did release 125Sb and 144Ce. Particles from underwater bursts will consist mostly of irradiated sea salts. Ratios of 152Eu/155Eu, 154Eu/155Eu, and 238Pu/239Pu are also different for fusion and fission nuclear weapons, which allows identification of hot particles of unknown origin.

Uranium has a relatively constant isotope ratio in all natural samples with ~0.72% 235
U
some 55 ppm 234
U
(in secular equilibrium with its parent nuclide 238
U
) and the balance made up by 238
U
. Isotopic compositions that diverge significantly from those values are evidence for the uranium having been subject to depletion or enrichment in some fashion or of (part of it) having participated in a nuclear fission reaction. While the latter is almost as universally due to human influence as the former two, the natural nuclear fission reactor at Oklo, Gabon was detected through a significant diversion of 235
U
concentration in samples from Oklo compared to those of all other known deposits on earth. Given that 235
U
is a material of proliferation concern then as now every IAEA-approved supplier of Uranium fuel keeps track of the isotopic composition of uranium to ensure none is diverted for nefarious purposes. It would thus become apparent quickly if another Uranium deposit besides Oklo proves to have once been a natural nuclear fission reactor.

Applications[edit]

Archaeological studies[edit]

In archaeological studies, stable isotope ratios have been used to track diet within the time span formation of analyzed tissues (10–15 years for bone collagen and intra-annual periods for tooth enamel bioapatite) from individuals; "recipes" of foodstuffs (ceramic vessel residues); locations of cultivation and types of plants grown (chemical extractions from sediments); and migration of individuals (dental material).[citation needed]

Forensics[edit]

With the advent of stable isotope ratio mass spectrometry, isotopic signatures of materials find increasing use in forensics, distinguishing the origin of otherwise similar materials and tracking the materials to their common source. For example, the isotope signatures of plants can be to a degree influenced by the growth conditions, including moisture and nutrient availability. In case of synthetic materials, the signature is influenced by the conditions during the chemical reaction. The isotopic signature profiling is useful in cases where other kinds of profiling, e.g. characterization of impurities, are not optimal. Electronics coupled with scintillator detectors are routinely used to evaluate isotope signatures and identify unknown sources.

A study was published demonstrating the possibility of determination of the origin of a common brown PSA packaging tape by using the carbon, oxygen, and hydrogen isotopic signature of the backing polymer, additives, and adhesive.[33]

Measurement of carbon isotopic ratios can be used for detection of adulteration of honey. Addition of sugars originated from corn or sugar cane (C4 plants) skews the isotopic ratio of sugars present in honey, but does not influence the isotopic ratio of proteins; in an unadulterated honey the carbon isotopic ratios of sugars and proteins should match.[34] As low as 7% level of addition can be detected.[35]

Nuclear explosions form 10Be by a reaction of fast neutrons with 13C in the carbon dioxide in air. This is one of the historical indicators of past activity at nuclear test sites.[36]

Solar System origins[edit]

Isotopic fingerprints are used to study the origin of materials in the Solar System.[37] For example, the Moon's oxygen isotopic ratios seem to be essentially identical to Earth's.[38] Oxygen isotopic ratios, which may be measured very precisely, yield a unique and distinct signature for each Solar System body.[39] Different oxygen isotopic signatures can indicate the origin of material ejected into space.[40] The Moon's titanium isotope ratio (50Ti/47Ti) appears close to the Earth's (within 4 ppm).[41][42] In 2013, a study was released that indicated water in lunar magma was 'indistinguishable' from carbonaceous chondrites and nearly the same as Earth's, based on the composition of water isotopes.[37][43]

Records of Early Life on Earth[edit]

Isotope biogeochemistry has been used to investigate the timeline surrounding life and its earliest iterations on Earth. Isotopic fingerprints typical of life, preserved in sediments, have been used to suggest, but do not necessarily prove, that life was already in existence on Earth by 3.85 billion years ago.[44]

Sulfur isotope evidence has also been used to corroborate the timing of the Great Oxidation Event, during which the Earth's atmosphere experienced a measurable rise in oxygen (to about 9% of modern values[45]) for the first time about 2.3-2.4 billion years ago. Mass-independent sulfur isotope fractionations are found to be widespread in the geologic record before about 2.45 billion years ago, and these isotopic signatures have since ceded to mass-dependent fractionation, providing strong evidence that the atmosphere shifted from anoxic to oxygenated at that threshold.[46]

Modern sulfate-reducing bacteria are known to favorably reduce lighter 32S instead of 34S, and the presence of these microorganisms can measurably alter the sulfur isotope composition of the ocean.[31] Because the δ34S values of sulfide minerals is primarily influenced by the presence of sulfate-reducing bacteria,[47] the absence of sulfur isotope fractionations in sulfide minerals suggests the absence of these bacterial processes or the absence of freely available sulfate. Some have used this knowledge of microbial sulfur fractionation to suggest that minerals (namely pyrite) with large sulfur isotope fractionations relative to the inferred seawater composition may be evidence of life.[48][49] This claim is not clear-cut, however, and is sometimes contested using geologic evidence from the ~3.49 Ga sulfide minerals found in the Dresser Formation of Western Australia, which are found to have δ34S values as negative as -22‰.[50] Because it has not been proven that the sulfide and barite minerals formed in the absence of major hydrothermal input, it is not conclusive evidence of life or of the microbial sulfate reduction pathway in the Archean.[51]

See also[edit]

References[edit]

  1. ^ Maberly, S. C.; Raven, J. A.; Johnston, A. M. (1992). "Discrimination between 12C and 13C by marine plants". Oecologia. 91 (4): 481. doi:10.1007/BF00650320. JSTOR 4220100.
  2. ^ Nobel, Park S. (7 February 2005). Physicochemical and Environmental Plant Physiology. Academic Press. p. 411. ISBN 978-0-12-520026-4.
  3. ^ Fernandez, Irene; Cadisch, Georg (2003). "Discrimination against13C during degradation of simple and complex substrates by two white rot fungi". Rapid Communications in Mass Spectrometry. 17 (23): 2614–2620. Bibcode:2003RCMS...17.2614F. doi:10.1002/rcm.1234. ISSN 0951-4198. PMID 14648898.
  4. ^ Fernandez, I.; Mahieu, N.; Cadisch, G. (2003). "Carbon isotopic fractionation during decomposition of plant materials of different quality". Global Biogeochemical Cycles. 17 (3): n/a. Bibcode:2003GBioC..17.1075F. doi:10.1029/2001GB001834. ISSN 0886-6236.
  5. ^ Farquhar, G D; Ehleringer, J R; Hubick, K T (1989). "Carbon Isotope Discrimination and Photosynthesis". Annual Review of Plant Physiology and Plant Molecular Biology. 40 (1): 503–537. doi:10.1146/annurev.pp.40.060189.002443. ISSN 1040-2519. S2CID 12988287.
  6. ^ McCarroll, Danny; Loader, Neil J. (2004). "Stable isotopes in tree rings". Quaternary Science Reviews. 23 (7–8): 771–801. Bibcode:2004QSRv...23..771M. CiteSeerX 10.1.1.336.2011. doi:10.1016/j.quascirev.2003.06.017. ISSN 0277-3791.
  7. ^ Ewe, Sharon M.L; da Silveira Lobo Sternberg, Leonel; Busch, David E (1999). "Water-use patterns of woody species in pineland and hammock communities of South Florida". Forest Ecology and Management. 118 (1–3): 139–148. doi:10.1016/S0378-1127(98)00493-9. ISSN 0378-1127.
  8. ^ Cabaneiro, Ana; Fernandez, Irene (2015). "Disclosing biome sensitivity to atmospheric changes: Stable C isotope ecophysiological dependences during photosynthetic uptake in Maritime pine and Scots pine ecosystems from southwestern Europe". Environmental Technology & Innovation. 4: 52–61. doi:10.1016/j.eti.2015.04.007. ISSN 2352-1864.
  9. ^ Silva, Lucas C. R.; Anand, Madhur; Shipley, Bill (2013). "Probing for the influence of atmospheric CO2and climate change on forest ecosystems across biomes". Global Ecology and Biogeography. 22 (1): 83–92. doi:10.1111/j.1466-8238.2012.00783.x. ISSN 1466-822X.
  10. ^ Gómez-Guerrero, Armando; Silva, Lucas C. R.; Barrera-Reyes, Miguel; Kishchuk, Barbara; Velázquez-Martínez, Alejandro; Martínez-Trinidad, Tomás; Plascencia-Escalante, Francisca Ofelia; Horwath, William R. (2013). "Growth decline and divergent tree ring isotopic composition (δ13C and δ18O) contradict predictions of CO2 stimulation in high altitudinal forests". Global Change Biology. 19 (6): 1748–1758. Bibcode:2013GCBio..19.1748G. doi:10.1111/gcb.12170. ISSN 1354-1013. PMID 23504983. S2CID 39714321.
  11. ^ O'Leary, M. H. (1988). "Carbon Isotopes in Photosynthesis". BioScience. 38 (5): 328–336. doi:10.2307/1310735. JSTOR 1310735. S2CID 29110460.
  12. ^ Meselson, M.; Stahl, F. W. (1958). "The replication of DNA in E. coli". Proceedings of the National Academy of Sciences of the United States of America. 44 (7): 671–682. Bibcode:1958PNAS...44..671M. doi:10.1073/pnas.44.7.671. PMC 528642. PMID 16590258.
  13. ^ Radajewski, S.; McDonald, I. R.; Murrell, J. C. (2003). "Stable-isotope probing of nucleic acids: a window to the function of uncultured microorganisms". Current Opinion in Biotechnology. 14 (3): 296–302. doi:10.1016/s0958-1669(03)00064-8. PMID 12849783.
  14. ^ Cupples, A. M.; Shaffer, E. A.; Chee-Sanford, J. C.; Sims, G. K. (2007). "DNA buoyant density shifts during 15N DNA stable isotope probing". Microbiological Research. 162 (4): 328–334. doi:10.1016/j.micres.2006.01.016. PMID 16563712.
  15. ^ Marsh, K. L.; Sims, G. K.; Mulvaney, R. L. (2005). "Availability of urea to autotrophic ammonia-oxidizing bacteria as related to the fate of 14C- and 15N-labeled urea added to soil". Biology and Fertility of Soils. 42 (2): 137–145. doi:10.1007/s00374-005-0004-2. S2CID 6245255.
  16. ^ Bichat, F.; Sims, G. K.; Mulvaney, R. L. (1999). "Microbial utilization of heterocyclic nitrogen from atrazine". Soil Science Society of America Journal. 63 (1): 100–110. Bibcode:1999SSASJ..63..100B. doi:10.2136/sssaj1999.03615995006300010016x.
  17. ^ Adams, Thomas S.; Sterner, Robert W. (2000). "The effect of dietary nitrogen content on trophic level 15N enrichment". Limnol. Oceanogr. 45 (3): 601–607. Bibcode:2000LimOc..45..601A. doi:10.4319/lo.2000.45.3.0601.
  18. ^ Richards, M. P.; Trinkaus, E. (2009). "Isotopic evidence for the diets of European Neanderthals and early modern humans". Proceedings of the National Academy of Sciences. 106 (38): 16034–16039. doi:10.1073/pnas.0903821106. PMC 2752538. PMID 19706482.
  19. ^ Handley, L.L; Austin, A. T.; Stewart, G.R.; Robinson, D.; Scrimgeour, C.M.; Raven, J.A.; Heaton, T.H.E.; Schmidt, S. (1999). "The 15N natural abundance (δ15N) of ecosystem samples reflects measures of water availability". Aust. J. Plant Physiol. 26 (2): 185–199. doi:10.1071/pp98146. ISSN 0310-7841.Closed access icon
  20. ^ Szpak, Paul; White, Christine D.; Longstaffe, Fred J.; Millaire, Jean-Francois; Vásquez Sánchez, Victor F. (2013). "Carbon and Nitrogen Isotopic Survey of Northern Peruvian Plants: Baselines for Paleodietary and Paleoecological Studies". PLOS ONE. 8 (1): e53763. Bibcode:2013PLoSO...853763S. doi:10.1371/journal.pone.0053763. PMC 3547067. PMID 23341996.
  21. ^ Ambrose, Stanley H.; DeNiro, Michael J. (1986). "The isotopic ecology of East African mammals". Oecologia. 69 (3): 395–406. Bibcode:1986Oecol..69..395A. doi:10.1007/bf00377062. PMID 28311342. S2CID 22660367.
  22. ^ Hobson, Keith A.; Alisauskas, Ray T.; Clark, Robert G. (1993). "Stable-Nitrogen Isotope Enrichment in Avian Tissues Due to Fasting and Nutritional Stress: Implications for Isotopic Analyses of Diet". The Condor. 95 (2): 388. doi:10.2307/1369361. JSTOR 1369361.
  23. ^ Dyches, Preston; Clavin, Whitney (June 23, 2014). "Titan's Building Blocks Might Pre-date Saturn" (Press release). Jet Propulsion Laboratory. Retrieved June 28, 2014.
  24. ^ de Laeter, John Robert; Böhlke, John Karl; De Bièvre, Paul; Hidaka, Hiroshi; Peiser, H. Steffen; Rosman, Kevin J. R.; Taylor, Philip D. P. (2003). "Atomic weights of the elements. Review 2000 (IUPAC Technical Report)". Pure and Applied Chemistry. 75 (6): 683–800. doi:10.1351/pac200375060683.
  25. ^ a b Moody, Kenton J.; Hutcheon, Ian D.; Grant, Patrick M. (28 February 2005). Nuclear Forensic Analysis. CRC Press. p. 399. ISBN 978-0-203-50780-3.
  26. ^ a b Tsang, Man-Yin; Yao, Weiqi; Tse, Kevin (2020). Kim, Il-Nam (ed.). "Oxidized silver cups can skew oxygen isotope results of small samples". Experimental Results. 1: e12. doi:10.1017/exp.2020.15. ISSN 2516-712X.
  27. ^ Lever, Mark A.; Rouxel, Olivier; Alt, Jeffrey C.; Shimizu, Nobumichi; Ono, Shuhei; Coggon, Rosalind M.; Shanks, Wayne C.; Lapham, Laura; Elvert, Marcus; Prieto-Mollar, Xavier; Hinrichs, Kai-Uwe (2013-03-01). "Evidence for Microbial Carbon and Sulfur Cycling in Deeply Buried Ridge Flank Basalt". Science. 339 (6125): 1305–1308. Bibcode:2013Sci...339.1305L. doi:10.1126/science.1229240. ISSN 0036-8075. PMID 23493710. S2CID 10728606.
  28. ^ Drake, Henrik; Roberts, Nick M. W.; Reinhardt, Manuel; Whitehouse, Martin; Ivarsson, Magnus; Karlsson, Andreas; Kooijman, Ellen; Kielman-Schmitt, Melanie (2021-06-03). "Biosignatures of ancient microbial life are present across the igneous crust of the Fennoscandian shield". Communications Earth & Environment. 2 (1): 1–13. doi:10.1038/s43247-021-00170-2. ISSN 2662-4435. S2CID 235307116.
  29. ^ Hannan, Keith (1998), "Sulfur isotopes in geochemistry", Geochemistry, Encyclopedia of Earth Science, Dordrecht: Springer Netherlands, pp. 610–615, doi:10.1007/1-4020-4496-8_309, ISBN 978-1-4020-4496-0
  30. ^ a b c d Stable Isotope Geochemistry (PDF). Springer Textbooks in Earth Sciences, Geography and Environment. 2021. doi:10.1007/978-3-030-77692-3. ISBN 978-3-030-77691-6. S2CID 238480248.
  31. ^ a b Seal, Robert R. II (2006-01-01). "Sulfur Isotope Geochemistry of Sulfide Minerals". Reviews in Mineralogy and Geochemistry. 61 (1): 633–677. Bibcode:2006RvMG...61..633S. doi:10.2138/rmg.2006.61.12. ISSN 1529-6466.
  32. ^ Farquhar, James; Bao, Huiming; Thiemens, Mark (2000-08-04). "Atmospheric Influence of Earth's Earliest Sulfur Cycle". Science. 289 (5480): 756–758. Bibcode:2000Sci...289..756F. doi:10.1126/science.289.5480.756. ISSN 0036-8075. PMID 10926533.
  33. ^ Carter, James F.; Grundy, Polly L.; Hill, Jenny C.; Ronan, Neil C.; Titterton, Emma L.; Sleeman, Richard (2004). "Forensic isotope ratio mass spectrometry of packaging tapes". Analyst. 129 (12): 1206–1210. Bibcode:2004Ana...129.1206C. doi:10.1039/b409341k. PMID 15565219.
  34. ^ González Martín, I.; Marqués Macías, E.; Sánchez Sánchez, J.; González Rivera, B. (1998). "Detection of honey adulteration with beet sugar using stable isotope methodology". Food Chemistry. 61 (3): 281–286. doi:10.1016/S0308-8146(97)00101-5.
  35. ^ "Tracking Nature: Geographical fingerprints in food ingredients add transparency to organic chain" (PDF). Canadian Honey Council. November 2004. pp. 10–11. Archived from the original (PDF) on 2014-01-01. Retrieved 30 April 2021.
  36. ^ Whitehead, Ne; Endo, S; Tanaka, K; Takatsuji, T; Hoshi, M; Fukutani, S; Ditchburn, Rg; Zondervan, A (2008). "A preliminary study on the use of (10)Be in forensic radioecology of nuclear explosion sites". Journal of Environmental Radioactivity. 99 (2): 260–70. doi:10.1016/j.jenvrad.2007.07.016. PMID 17904707.
  37. ^ a b Spudis, Paul D. (May 14, 2013). "Earth-Moon: A Watery "Double-Planet"". Archived from the original on 2013-08-07. Retrieved April 30, 2021.
  38. ^ Wiechert, U.; et al. (October 2001). "Oxygen Isotopes and the Moon-Forming Giant Impact". Science. 294 (12): 345–348. Bibcode:2001Sci...294..345W. doi:10.1126/science.1063037. PMID 11598294. S2CID 29835446.
  39. ^ Scott, Edward R. D. (December 3, 2001). "Oxygen Isotopes Give Clues to the Formation of Planets, Moons, and Asteroids". Planetary Science Research Discoveries Report: 55. Bibcode:2001psrd.reptE..55S. Retrieved 2014-01-01.
  40. ^ Nield, Ted (September 2009). "Moonwalk". Geological Society of London. p. 8. Retrieved 2014-01-01.
  41. ^ Zhang, Junjun; Nicolas Dauphas; Andrew M. Davis; Ingo Leya; Alexei Fedkin (25 March 2012). "The proto-Earth as a significant source of lunar material". Nature Geoscience. 5 (4): 251–255. Bibcode:2012NatGe...5..251Z. doi:10.1038/ngeo1429. S2CID 38921983.
  42. ^ Koppes, Steve (March 28, 2012). "Titanium paternity test fingers Earth as moon's sole parent". Zhang, Junjun. The University of Chicago. Retrieved 2014-01-01.
  43. ^ Saal, A. E.; Hauri, E. H.; Van Orman, J. A.; Rutherford, M. J. (2013). "Hydrogen Isotopes in Lunar Volcanic Glasses and Melt Inclusions Reveal a Carbonaceous Chondrite Heritage". Science. 340 (6138): 1317–1320. Bibcode:2013Sci...340.1317S. doi:10.1126/science.1235142. PMID 23661641. S2CID 9092975.
  44. ^ Mojzsis, S. J.; Arrhenius, G.; McKeegan, K. D.; Harrison, T. M.; Nutman, A. P.; Friend, C. R. L. (November 1996). "Evidence for life on Earth before 3,800 million years ago". Nature. 384 (6604): 55–59. Bibcode:1996Natur.384...55M. doi:10.1038/384055a0. hdl:2060/19980037618. ISSN 1476-4687. PMID 8900275. S2CID 4342620.
  45. ^ Holland, Heinrich D (2006-06-29). "The oxygenation of the atmosphere and oceans". Philosophical Transactions of the Royal Society B: Biological Sciences. 361 (1470): 903–915. doi:10.1098/rstb.2006.1838. PMC 1578726. PMID 16754606.
  46. ^ Papineau, Dominic; Mojzsis, Stephen J.; Schmitt, Axel K. (2007-03-15). "Multiple sulfur isotopes from Paleoproterozoic Huronian interglacial sediments and the rise of atmospheric oxygen". Earth and Planetary Science Letters. 255 (1): 188–212. Bibcode:2007E&PSL.255..188P. doi:10.1016/j.epsl.2006.12.015. ISSN 0012-821X.
  47. ^ Canfield, D. E. (2001-01-01). "Biogeochemistry of Sulfur Isotopes". Reviews in Mineralogy and Geochemistry. 43 (1): 607–636. Bibcode:2001RvMG...43..607C. doi:10.2138/gsrmg.43.1.607. ISSN 1529-6466.
  48. ^ Archer, Corey; Vance, Derek (2006-03-01). "Coupled Fe and S isotope evidence for Archean microbial Fe(III) and sulfate reduction". Geology. 34 (3): 153–156. Bibcode:2006Geo....34..153A. doi:10.1130/G22067.1. ISSN 0091-7613.
  49. ^ Wacey, David; McLoughlin, Nicola; Whitehouse, Martin J.; Kilburn, Matt R. (2010-12-01). "Two coexisting sulfur metabolisms in a ca. 3400 Ma sandstone". Geology. 38 (12): 1115–1118. Bibcode:2010Geo....38.1115W. doi:10.1130/G31329.1. ISSN 0091-7613.
  50. ^ Philippot, Pascal; Zuilen, Mark; Lepot, Kevin; Thomazo, Christophe; Farquhar, James; Van Kranendonk, Martin (2007-09-14). "Early Archaean Microorganisms Preferred Elemental Sulfur, Not Sulfate". Science. 317 (5844): 1534–1537. Bibcode:2007Sci...317.1534P. doi:10.1126/science.1145861. PMID 17872441. S2CID 41254565.
  51. ^ Early Life on Earth (PDF). Topics in Geobiology. Vol. 31. 2009. doi:10.1007/978-1-4020-9389-0. ISBN 978-1-4020-9388-3.

Further reading[edit]