stringtranslate.com

Detrital zircon geochronology

Fig. 1 – Zircon grains in real life (Coin for scale)

Detrital zircon geochronology is the science of analyzing the age of zircons deposited within a specific sedimentary unit by examining their inherent radioisotopes, most commonly the uranium–lead ratio. Zircon is a common accessory or trace mineral constituent of most granite and felsic igneous rocks. Due to its hardness, durability and chemical inertness, zircon persists in sedimentary deposits and is a common constituent of most sands. Zircons contain trace amounts of uranium and thorium and can be dated using several modern analytical techniques.

Detrital zircon geochronology has become increasingly popular in geological studies from the 2000s mainly due to the advancement in radiometric dating techniques.[1][2] Detrital zircon age data can be used to constrain the maximum depositional age, determine provenance,[3] and reconstruct the tectonic setting on a regional scale.[4]

Detrital zircon

Origin

Detrital zircons are part of the sediment derived from weathering and erosion of pre-existing rocks. Since zircons are heavy and highly resistant at Earth's surface,[5] many zircons are transported, deposited and preserved as detrital zircon grains in sedimentary rocks.[3]

Fig. 2 – Simple diagram illustrating the formation of igneous zircon, the processes of them becoming detrital zircons and the differences between igneous and detrital zircons

Properties

Detrital zircons usually retain similar properties as their parent igneous rocks, such as age, rough size and mineral chemistry.[6][7] However, the composition of detrital zircons is not entirely controlled by the crystallization of the zircon mineral. In fact, many of them are modified by later processes in the sedimentary cycle. Depending on the degree of physical sorting, mechanical abrasion and dissolution, a detrital zircon grain may lose some of its inherent features and gain some over-printed properties like rounded shape and smaller size.[5] On a larger scale, two or more tribes of detrital zircons from different origins may deposit within the same sedimentary basin. This give rise to a natural complexity of associating detrital zircon populations and their sources.[3]

Zircon is a strong tool for uranium-lead age determination because of its inherent properties:[8]

  1. Zircon contains high amount of uranium for machine recognition, commonly 100–1000 ppm.[8]
  2. Zircon has a low amount of lead during crystallization, in parts per trillion.[8] Thus, lead found in zircon can be assumed as daughter nuclei from parent uranium.
  3. Zircon crystals grow between 600 and 1100 °C, while lead is retained within the crystal structure below 800 °C (see Closure temperature). So once zircon has cooled below 800 °C it retains all the lead from the radioactive decay. Therefore, U-Pb age can be treated as the age of crystallization,[8] if the mineral/sample itself has not undergone high temperature metamorphism after formation.
  4. Zircon commonly crystallizes in felsic igneous rocks, with greater than 60% silica (SiO2) content.[4] These rocks are generally less dense and more buoyant. They sit high in the Earth's (continental crust), and have good preservation potential.
  5. Zircon is physically and chemically resistant, so it is more likely to be preserved in the sedimentary cycle.[8]
  6. Zircon contains other elements which gives supplementary information, such as hafnium (Hf), uranium/thorium (U/Th) ratio.[8]

Sample collection

There are no set rules for sample selection in detrital zircon geochronology studies. The objective and scale of the research project govern the type and number of samples taken. In some cases, the sedimentary rock type and depositional setting can significantly affect the result.[3] Examples include:

Detrital zircon extraction

After rock samples are collected, they are cleaned, chipped, crushed and milled through standardized procedures. Then, detrital zircons are separated from the fine rock powder by three different ways, namely gravity separation using water, magnetic separation, and gravity separation using heavy liquid.[11] In the process, grains are also sieved according to their size. The commonly used grain size for detrital zircon provenance analysis is 63–125 μm, which is equivalent to fine sand grain size.[12]

Type of detrital zircon analysis

There are two main types of detrital zircon analysis: qualitative analysis and quantitative analysis. The biggest advantage of qualitative analysis is being able to uncover all possible origin of the sedimentary unit, whereas quantitative analysis should allow meaningful comparison of proportions in the sample.[3]

Qualitative analysis

Qualitative approach examines all the available detrital zircons individually regardless of their abundance among all grains.[13][14] This approach is usually conducted with high precision thermal ionization mass spectrometry (TIMS) and sometimes secondary ion mass spectrometry (SIMS).[3] Optical examination and classification of detrital zircon grains are commonly included in qualitative studies through back-scatter electrons (BSE) or cathodoluminescence (CL) imagery,[3] despite the relationship between the age and optical classification of detrital zircon grains is not always reliable.[15]

Quantitative analysis

Quantitative approach requires large number of grain analyses within a sample rock in order to represent the overall detrital zircon population[3] statistically (i.e. the total number of analyses should achieve an appropriate level of confidence).[16] Because of the large sample size, secondary ion mass spectrometry (SIMS) and laser ablation-inductively coupled plasma mass spectrometry (LA-ICPMS) are used instead of thermal ionization mass spectrometry (TIMS). In this case, BSE and CL imagery are applied to select the best spot on a zircon grain for acquiring reliable age.[17]

Methods

Different methods in detrital zircon analysis yield different results. Generally, researchers would include the methods/ analytical instruments they used within their studies. There are generally three categories, which are the instrument(s) used for zircon analysis, their calibration standards and instrument(s) used for zircon imagery. Details are listed in Table 1.

Detrital zircon data

Depending on the detrital zircon study, there should be different variables included for analysis. There are two main types of data, analyzed zircon data (quantifiable data and imagery/descriptive data), and sample (where they extract the zircon grains) data. Details are listed in Table 2.

Filtering detrital zircon data

All data acquired first-hand should be cleansed before using to avoid error, normally by computer.

By U-Pb age discordance

Before applying detrital zircon ages, they should be evaluated and screened accordingly. In most cases, data are compared with U-Pb Concordia graphically. For a large dataset, however, data with high U-Pb age discordance (>10 – 30%) are filtered out numerically. The acceptable discordance level is often adjusted with the age of the detrital zircon since older population should experience higher chances of alteration and project higher discordance.[19] (See Uranium–lead dating)

By choosing the best age

Because of the intrinsic uncertainties within the three yield U-Pb ages (207Pb/235U, 206Pb/238U and 207Pb/206Pb), the age at ~1.4 Ga has the poorest resolution. An overall consensus for age with higher accuracy is to adopt:

By data clustering

Given the possibility of concordant yet incorrect detrital zircon U-Pb ages associated with lead loss or inclusion of older components, some scientists apply data selection through clustering and comparing the ages. Three or more data overlapping within ±2σ uncertainty would be classified as a valid age population of a particular source origin.[19]

By age uncertainty (±σ)

There are no set limit for age uncertainty and the cut-off value varies with different precision requirement. Although excluding data with huge age uncertainty would enhance the overall zircon grain age accuracy, over elimination may lower overall research reliability (decrease in size of the database). The best practice would be to filter accordingly, i.e. setting the cut-off error to eliminate reasonable portion of the dataset (say <5% of the total ages available[6])

By applied analytical methods

Depending on the required analytical accuracy, researchers may filter data via their analytical instruments. Generally, researchers use only the data from sensitive high-resolution ion microprobe (SHRIMP), inductively coupled plasma mass spectrometry (LA-ICPMS) and thermal ionization mass spectrometry (TIMS) because of their high precision (1–2%, 1–2% and 0.1% respectively[17]) in spot analysis. An older analytical technique, lead-lead evaporation,[37] is no longer used since it cannot determine the U-Pb concordance of the age data.[38]

By spot nature

Apart from analytical methods, researchers would isolate core or rim ages for analysis. Normally, core ages would be used as crystallization age as they are first generated and least disturbed part in zircon grains. On the other hand, rim ages can be used to track peak metamorphism as they are first in contact with certain temperature and pressure condition.[39] Researchers may utilize these different spot natures to reconstruct the geological history of a basin.

Application of detrital zircon ages

Maximum depositional age

Some of the most important information we can get from detrital zircon ages is the maximum depositional age of the referring sedimentary unit. The sedimentary unit cannot be older than the youngest age of the analyzed detrital zircons because the zircon should have existed before the rock formation. This provides useful age information to rock strata where fossils are unavailable, such as the terrestrial successions during Precambrian or pre-Devonian times. [40][3] Practically, maximum depositional age is averaged from a cluster of youngest age data or the peak in age probability because the youngest U-Pb age within a sample is almost always younger with uncertainty.[17]

Tectonic studies

Using detrital zircon age abundance

In a global scale, detrital zircon age abundance can be used as a tool to infer significant tectonic events in the past.[4] In Earth's history, the abundance of magmatic age peaks during periods of supercontinent assembly.[6] This is because supercontinent provides a major crustal envelop selectively preserve the felsic magmatic rocks, resulting from partial melts.[41] Thus, many detrital zircons originate from these igneous provence, resulting similar age peak records.[6] For instance, the peak at about 0.6–0.7 Ga and 2.7 Ga (Figure 6) may correlate the break-up of Rodinia and supercontinent Kenorland respectively.[26]

Fig. 6 – Global detrital zircon age distribution in a frequency versus geological age diagram. Modified from Voice et al. (2011)

Using difference between detrital zircons crystallisation ages and their corresponding maximum depositional age

Apart from the detrital zircon age abundance, difference between detrital zircons crystallisation ages (CA) and their corresponding maximum depositional age (DA) can be plotted in cumulative distribution function to correlate particular tectonic regime in the past. The effect of different tectonic settings on the difference between CA and DA is illustrated in Figure 7 and summarized in Table. 3.[4]

Fig. 7 – Schematic diagram showing the source rock nature and their proximity to the sedimentary basins in multiple tectonic settings. Modified from Cawood et al. (2012)

References

  1. ^ Davis, Donald W.; Williams, Ian S.; Krogh, Thomas E. (2003). Hanchar, J.M.; Hoskin, P.W.O. (eds.). "Historical development of U-Pb geochronology" (PDF). Zircon: Reviews in Mineralogy and Geochemistry. 53: 145–181. doi:10.2113/0530145.
  2. ^ Kosler, J.; Sylvester, P.J. (2003). Hanchar, J.M.; Hoskin, P.W.O. (eds.). "Present trends and the future of zircon in U-Pb geochronology: laser ablation ICPMS". Zircon: Reviews in Mineralogy and Geochemistry. 53 (1): 243–275. Bibcode:2003RvMG...53..243K. doi:10.2113/0530243.
  3. ^ a b c d e f g h i Fedo, C. M.; Sircombe, K. N.; Rainbird, R. H. (2003). "Detrital zircon analysis of the sedimentary record". Reviews in Mineralogy and Geochemistry. 53 (1): 277–303. Bibcode:2003RvMG...53..277F. doi:10.2113/0530277.
  4. ^ a b c d e f Cawood, P. A.; Hawkesworth, C. J.; Dhuime, B. (22 August 2012). "Detrital zircon record and tectonic setting". Geology. 40 (10): 875–878. Bibcode:2012Geo....40..875C. doi:10.1130/G32945.1. hdl:10023/3575.
  5. ^ a b Morton, Andrew C; Hallsworth, Claire R (March 1999). "Processes controlling the composition of heavy mineral assemblages in sandstones". Sedimentary Geology. 124 (1–4): 3–29. Bibcode:1999SedG..124....3M. doi:10.1016/S0037-0738(98)00118-3.
  6. ^ a b c d Condie, Kent C.; Belousova, Elena; Griffin, W.L.; Sircombe, Keith N. (June 2009). "Granitoid events in space and time: Constraints from igneous and detrital zircon age spectra". Gondwana Research. 15 (3–4): 228–242. Bibcode:2009GondR..15..228C. doi:10.1016/j.gr.2008.06.001.
  7. ^ Hawkesworth, C. J.; Dhuime, B.; Pietranik, A. B.; Cawood, P. A.; Kemp, A. I. S.; Storey, C. D. (1 March 2010). "The generation and evolution of the continental crust". Journal of the Geological Society. 167 (2): 229–248. Bibcode:2010JGSoc.167..229H. doi:10.1144/0016-76492009-072. S2CID 131052922.
  8. ^ a b c d e f g Gehrels, G. (12 August 2010). UThPb analytical methods for Zircon. Arizona LaserChron Center. Retrieved 10 November 2016, from https://drive.google.com/file/d/0B9ezu34P5h8eMzkyMGFlNjgtMDU0Zi00MTQyLTliZDMtODU2NGE0MDQ2NGU2/view?hl=en.
  9. ^ Smith, Moira; Gehrels, George (July 1994). "Detrital zircon geochronology and the provenance of the Harmony and Valmy Formations, Roberts Mountains allochthon, Nevada". Geological Society of America Bulletin. 106 (7): 968–979. Bibcode:1994GSAB..106..968S. doi:10.1130/0016-7606(1994)106<0968:DZGATP>2.3.CO;2.
  10. ^ DeGraaff-Surpless, K., McWilliams, M. O., Wooden, J. L., & Ireland, T. R. (2000). Limitations of detrital zircon data for provenance analysis: an example from the Methow Basin, Washington and British Columbia. In Geol Soc Am Abstr Progr (Vol. 32, No. 9).
  11. ^ Chisholm, E. I., Sircombe, K. N. and DiBugnara, D. L. 2014. Handbook of Geochronology Mineral Separation Laboratory Techniques. Record 2014/46. Geoscience Australia, Canberra. doi:10.11636/Record.2014.046
  12. ^ Morton, A. C.; Claoué-Long, J. C.; Berge, C. (1996). "SHRIMP constraints on sediment provenance and transport history in the Mesozoic Statfjord Formation, North Sea". Journal of the Geological Society. 153 (6): 915–929. Bibcode:1996JGSoc.153..915M. doi:10.1144/gsjgs.153.6.0915. S2CID 130260438.
  13. ^ Gehrels, G. E.; Dickinson, W. R.; Ross, G. M.; Stewart, J. H.; Howell, D. G. (1995). "Detrital zircon reference for Cambrian to Triassic miogeoclinal strata of western North America". Geology. 23 (9): 831–834. Bibcode:1995Geo....23..831G. doi:10.1130/0091-7613(1995)023<0831:dzrfct>2.3.co;2.
  14. ^ a b Gehrels, G. E. (2000). "Introduction to detrital zircon studies of Paleozoic and Triassic strata in western Nevada and northern California". Special Paper of the Geological Society of America. 347: 1–17.
  15. ^ Roback, R. C.; Walker, N. W. (1995). "Provenance, detrital zircon U-Pb geochronometry, and tectonic significance of Permian to Lower Triassic sandstone in southeastern Quesnellia, British Columbia and Washington". Geological Society of America Bulletin. 107 (6): 665–675. Bibcode:1995GSAB..107..665R. doi:10.1130/0016-7606(1995)107<0665:pdzupg>2.3.co;2.
  16. ^ Dodson, M. H.; Compston, W.; Williams, I. S.; Wilson, J. F. (1988). "A search for ancient detrital zircons in Zimbabwean sediments". Journal of the Geological Society. 145 (6): 977–983. Bibcode:1988JGSoc.145..977D. doi:10.1144/gsjgs.145.6.0977. S2CID 140654427.
  17. ^ a b c d Gehrels, G (2014). "Detrital zircon U-Pb geochronology applied to tectonics". Annual Review of Earth and Planetary Sciences. 42 (1): 127–149. Bibcode:2014AREPS..42..127G. doi:10.1146/annurev-earth-050212-124012.
  18. ^ a b c d e f g h i Corfu, F.; Hanchar, J. M.; Hoskin, P. W.; Kinny, P. (2003). "Atlas of zircon textures". Reviews in Mineralogy and Geochemistry. 53 (1): 469–500. Bibcode:2003RvMG...53..469C. doi:10.2113/0530469.
  19. ^ a b c Gehrels, G. (2011). Detrital zircon U‐Pb geochronology: Current methods and new opportunities. Tectonics of sedimentary basins: recent advances, 45–62.
  20. ^ Chakoumakos, BC; Murakami, T; Lumpkin, GR; Ewing, RC (1987). "Alpha-decay-induced fracturing in zircon: the transition from the crystalline to the metamict state". Science. 236 (4808): 1556–1559. Bibcode:1987Sci...236.1556C. doi:10.1126/science.236.4808.1556. PMID 17835739. S2CID 44648291.
  21. ^ Murakami, T; Chakoumakos, BC; Ewing, RC; Lumpkin, GR; Weber, WJ (1991). "Alpha-decay event damage in zircon". Am Mineral. 76: 1510–1532.
  22. ^ Krogh TE, Davis GL (1975) Alteration in zircons and differential dissolution of altered and metamict zircon. Carnegie Inst Washington Yrbk74:619–623
  23. ^ Crookes, W (1879). "Contributions to Molecular Physics in High Vacua. Magnetic Deflection of Molecular Trajectory. Laws of Magnetic Rotation in High and Low Vacua. Phosphorogenic Properties of Molecular Discharge". Philosophical Transactions of the Royal Society of London. 170: 641–662. Bibcode:1879RSPT..170..641C. doi:10.1098/rstl.1879.0076.
  24. ^ Ohnenstetter, D.; Cesbron, F.; Remond, G.; Caruba, R.; Claude, J. M. (1991). "Émissions de cathodoluminescence de deux populations de zircons naturels: tentative d'interprétation". Comptes Rendus de l'Académie des Sciences. 313 (6): 641–647.
  25. ^ Hanchar, JM; Miller, CF (1993). "Zircon zonation patterns as revealed by cathodoluminescence and backscattered electron images: Implications for interpretation of complex crustal histories". Chem Geol. 110 (1–3): 1–13. Bibcode:1993ChGeo.110....1H. doi:10.1016/0009-2541(93)90244-D.
  26. ^ a b Voice, P. J.; Kowalewski, M.; Eriksson, K. A. (2011). "Quantifying the timing and rate of crustal evolution: global compilation of radiometrically dated detrital zircon grains". The Journal of Geology. 119 (2): 109–126. Bibcode:2011JG....119..109V. doi:10.1086/658295. S2CID 128408445.
  27. ^ Welcome to Geochron | EarthChem. (n.d.). Retrieved 15 November 2016, from http://www.geochron.org/
  28. ^ Hiess, J.; Condon, D. J.; McLean, N.; Noble, S. R. (2012). "238U/235U systematics in terrestrial uranium-bearing minerals" (PDF). Science. 335 (6076): 1610–1614. Bibcode:2012Sci...335.1610H. doi:10.1126/science.1215507. PMID 22461608. S2CID 206538233.
  29. ^ Decay systems and geochronology II: U and Th. (4 December 2013). Retrieved 15 November 2016, from http://www.geo.cornell.edu/geology/classes/Geo656/656notes13/IsotopeGeochemistry Chapter3.pdf
  30. ^ Jaffey, A. H.; Flynn, K. F.; Glendenin, L. E.; Bentley, W. T.; Essling, A. M. (1971). "Precision measurement of half-lives and specific activities of U 235 and U 238". Physical Review C. 4 (5): 1889. doi:10.1103/physrevc.4.1889.
  31. ^ Steiger, R. H., & Jager, E. (1978). Subcommission on Geochronology: Convention on the use of decay constants in geochronology and cosmochronology.
  32. ^ Pupin, J. P. (1980). "Zircon and granite petrology". Contributions to Mineralogy and Petrology. 73 (3): 207–220. Bibcode:1980CoMP...73..207P. doi:10.1007/bf00381441. S2CID 96470918.
  33. ^ Wang, X.; Zhou, D. (2001). "A new equilibrium form of zircon crystal". Science in China Series B: Chemistry. 44 (5): 516–523. doi:10.1007/bf02880682.
  34. ^ Mattinson, J. M., Graubard, C. M., Parkinson, D. L., & McClelland, W. C. (1996). U‐Pb Reverse Discordance in Zircons: The Role of Fine‐Scale Oscillatory Zoning and Sub‐Micron Transport of Pb. Earth Processes: Reading the Isotopic Code, 355–370.
  35. ^ a b Dickinson, W. R.; Gehrels, G. E. (2009). "Use of U–Pb ages of detrital zircons to infer maximum depositional ages of strata: a test against a Colorado Plateau Mesozoic database". Earth and Planetary Science Letters. 288 (1): 115–125. Bibcode:2009E&PSL.288..115D. doi:10.1016/j.epsl.2009.09.013.
  36. ^ Gehrels, G.E.; Valencia, V.; Ruiz, J. (2008). "Enhanced precision, accuracy, efficiency, and spatial resolution of U-Pb ages by laser ablation–multicollector–inductively coupled plasma–mass spectrometry". Geochemistry, Geophysics, Geosystems. 9 (3): n/a. Bibcode:2008GGG.....9.3017G. doi:10.1029/2007GC001805.
  37. ^ Kober, B (1986). "Whole-grain evaporation for 207Pb/206Pb-age-investigations on single zircons using a double-filament thermal ion source". Contributions to Mineralogy and Petrology. 93 (4): 482–490. Bibcode:1986CoMP...93..482K. doi:10.1007/bf00371718. S2CID 129728272.
  38. ^ Hirata, T.; Nesbitt, R. W. (1995). "U-Pb isotope geochronology of zircon: Evaluation of the laser probe-inductively coupled plasma mass spectrometry technique". Geochimica et Cosmochimica Acta. 59 (12): 2491–2500. Bibcode:1995GeCoA..59.2491H. doi:10.1016/0016-7037(95)00144-1.
  39. ^ Nicoli, G., Moyen, J. F., & Stevens, G. (2016). Diversity of burial rates in convergent settings decreased as Earth aged. Scientific reports, 6.
  40. ^ Rugen, Elias J.; Pastore, Guido; Vermeesch, Pieter; Spencer, Anthony M.; Webster, David; Smith, Adam G. G.; Carter, Andrew; Shields, Graham A. (2 September 2024). "Glacially influenced provenance and Sturtian affinity revealed by detrital zircon U–Pb ages from sandstones in the Port Askaig Formation, Dalradian Supergroup". Journal of the Geological Society. 181 (5). doi:10.1144/jgs2024-029. ISSN 0016-7649.
  41. ^ a b Hawkesworth, C. J.; Dhuime, B.; Pietranik, A. B.; Cawood, P. A.; Kemp, A. I. S.; Storey, C. D. (2010). "The generation and evolution of the continental crust". Journal of the Geological Society. 167 (2): 229–248. Bibcode:2010JGSoc.167..229H. doi:10.1144/0016-76492009-072. S2CID 131052922.
  42. ^ Storey, B. C. (1995). "The role of mantle plumes in continental breakup: case histories from Gondwanaland". Nature. 377 (6547): 301–308. Bibcode:1995Natur.377..301S. doi:10.1038/377301a0. S2CID 4242617.

External links