TEX86

Source: Wikipedia, the free encyclopedia.
Molecular structures and HPLC detection of GDGTs. Retrieved from Tierney and Tingley (2015).[1]

TEX86 is an organic

mesophilic marine Nitrososphaerota (formerly "Thaumarchaeota", "Marine Group 1 Crenarchaeota").[2][3]

Basics

The membrane lipids of

glycerol dialkyl glycerol tetraethers (GDGTs) which contain 0-3 cyclopentane moieties (commonly annotated as GDGT-n where n = numbers of cyclopentane moieties). Nitrososphaerota also synthesise crenarchaeol (cren) which contains four cyclopentane moieties and a single cyclohexane moiety and a regio-isomer (cren'). The cyclohexane and cyclopentane rings, formed by internal cyclisation of one of the biphytane chains,[4] have a pronounced effect on the thermal transition points of the Nitrososphaerota cell membrane. Mesocosm studies demonstrate that the degree of cyclisation is generally governed by growth temperature.[5]

Calibrations

Based upon the relative distribution of isoprenoidal GDGTs, Schouten et al. (2002)[6] proposed the tetraether index of 86 carbon atoms (TEX86) as a proxy for sea surface temperature (SST). GDGT-0 is excluded from the calibration as it can have multiple sources [7] while cren is omitted as it exhibits no correlation with SST and is often an order of magnitude more abundant than its isomer and the other GDGTs. The most recent TEX86 calibration invokes two separate indices and calibrations:[8] TEX86H uses the same combination of GDGTs as in the original TEX86 relationship:

GDGT ratio-2 is correlated to SST using the calibration equation:

TEX86H = 68.4×log(GDGT ratio-2) + 38.6.

TEX86H has a calibration error of ±2.5 °C and is based upon 255 core-top sediments.

TEX86L employs a combination of GDGTs that is different from TEX86H, removing GDGT-3 from the numerator and excluding cren' entirely:

GDGT ratio-1 is correlated to SST using the calibration equation:

TEX86L = 67.5×log(GDGT ratio-1) + 46.9.

TEX86Lhas a calibration error of ±4 °C and is based upon 396 core-top sediment samples.

Other calibrations exist (including 1/TEX86,[9] TEX86'[10] and pTEX86 [11]) and should be considered when reconstructing temperature.

Caveats

There are several caveats to this proxy and this list is by no means exhaustive. For more information, consult Schouten et al. 2013.[12]

Terrestrial input

The branched vs isoprenoidal tetratether (BIT) index can used to measure the relative fluvial input of terrestrial organic matter (TOM) into the marine realm.[13] The BIT index is based upon the premise that crenarchaeol is derived from marine-dwelling Nitrososphaerota and branched GDGTs are derived from terrestrial soil bacteria. When BIT values exceed 0.4, a deviation of >2 °C is incorporated into TEX86-based SST estimates. However, isoprenoidal GDGTs can be synthesised on land (by terrestrial archaea) and can render BIT values unreliable; isoGDGT becomes more abundant with higher soil pH.[12]: §7.1.2  A strong co-variation between GDGT-4 and branched GDGTs in modern marine and freshwater environments also suggests a common or mixed source for isoprenoidal and branched GDGTs (Fietz et al., 2012).[full citation needed]

Anaerobic oxidation of methane (AOM)

The Methane Index (MI) was proposed to help distinguish the relative input of methanotrophic Euryarchaeota in settings characterised by diffuse methane flux and anaerobic oxidation of methane (AOM).[14] These sites are characterised by a distinct GDGT distribution, namely the predominance of GDGT-1. -2 and -3. High MI values (>0.5) reflect high rates of gas-hydrate-related AOM.

Degradation

Thermal maturity is only thought to affect GDGTs when temperature exceed 240 °C. This can be tested using a ratio of specific

Oxic
degradation, which is a selective process and degrades compounds at different rates, has been shown to affect TEX86 values and can bias SST values by up to 6 °C.

Application

The oldest TEX86 record is from the middle

alkenones[19]
).

Eocene

TEX86 has been extensively used to reconstruct

Mg/Ca). During the middle and late Eocene, high southern latitude sites cooled while the tropics remained stable and warm. Possible reasons for this cooling include long-term changes in carbon dioxide and/or changes in gateway reorganisation (e.g. Tasman Gateway, Drake Passage
).

References

  1. .
  2. .
  3. .
  4. .
  5. .
  6. .
  7. ^ Koga, Y., Nishihara, M., Morii, H., and Akagawa-Matsushita, M., 1993, Ether polar lipids of methanogenic bacteria: structures, comparative aspects, and biosyntheses: Microbiological Reviews, v. 57, no. 1, p. 164-182
  8. ^ Kim, J.-H., van der Meer, J., Schouten, S., Helmke, P., Willmott, V., Sangiorgi, F., Koç, N., Hopmans, E. C., and Damsté, J. S. S., 2010, New indices and calibrations derived from the distribution of crenarchaeal isoprenoid tetraether lipids: Implications for past sea surface temperature reconstructions: Geochimica et Cosmochimica Acta, v. 74, no. 16, p. 4639-4654.
  9. ^ Liu, Z., Pagani, M., Zinniker, D., DeConto, R., Huber, M., Brinkhuis, H., Shah, S. R., Leckie, R. M., and Pearson, A., 2009, Global Cooling During the Eocene-Oligocene Climate Transition: Science, v. 323, no. 5918, p. 1187-1190
  10. ^ Sluijs, A., Schouten, S., Pagani, M., Woltering, M., Brinkhuis, H., Damsté, J. S. S., Dickens, G. R., Huber, M., Reichart, G.-J., Stein, R., Matthiessen, J., Lourens, L. J., Pedentchouk, N., Backman, J., Moran, K., and the Expedition, S., 2006, Subtropical Arctic Ocean temperatures during the Palaeocene/Eocene thermal maximum: Nature, v. 441, no. 7093, p. 610-613.
  11. ^ Hollis, C. J., Taylor, K. W. R., Handley, L., Pancost, R. D., Huber, M., Creech, J. B., Hines, B. R., Crouch, E. M., Morgans, H. E. G., Crampton, J. S., Gibbs, S., Pearson, P. N., and Zachos, J. C., 2012, Early Paleogene temperature history of the Southwest Pacific Ocean: Reconciling proxies and models: Earth and Planetary Science Letters, v. 349–350, no. 0, p. 53-66.
  12. ^
    ISSN 0146-6380
    .
  13. .
  14. .
  15. ^ Jenkyns, H., Schouten-Huibers, L., Schouten S. and Sinninghe-Damste, J.S., 2012, Warm Middle Jurassic-early Cretaceous high-latitude sea surface temperature from the Southern Ocean. Climate of the Past, v. 8, p.215-226
  16. ^ Sluijs, A., Schouten, S., Donders, T. H., Schoon, P. L., Rohl, U., Reichart, G.-J., Sangiorgi, F., Kim, J.-H., Sinninghe Damste, J. S., and Brinkhuis, H., 2009, Warm and wet conditions in the Arctic region during Eocene Thermal Maximum 2: Nature Geosci, v. 2, no. 11, p. 777-780.
  17. ^ Zachos, J. C., Schouten, S., Bohaty, S., Quattlebaum, T., Sluijs, A., Brinkhuis, H., Gibbs, S. J., and Bralower, T. J., 2006, Extreme warming of mid-latitude coastal ocean during the Paleocene-Eocene Thermal Maximum: Inferences from TEX86 and isotope data: Geology, v. 34, no. 9, p. 737-740.
  18. ^ Pearson, P. N., van Dongen, B. E., Nicholas, C. J., Pancost, R. D., Schouten, S., Singano, J. M., and Wade, B. S., 2007, Stable warm tropical climate through the Eocene Epoch: Geology, v. 35, no. 3, p. 211-214.
  19. ^ Bijl, P. K., Schouten, S., Sluijs, A., Reichart, G.-J., Zachos, J. C., and Brinkhuis, H., 2009, Early Palaeogene temperature evolution of the southwest Pacific Ocean: Nature, v. 461, no. 7265, p. 776-779.

Further reading

This page is based on the copyrighted Wikipedia article: TEX86. Articles is available under the CC BY-SA 3.0 license; additional terms may apply.Privacy Policy