Introduction edit

Hydrogen isotope biogeochemistry is the scientific study of biological, geological, and chemical processes in the environment using the distribution and relative abundance of hydrogen isotopes. There are two stable isotopes of hydrogen, 1H and 2H, which vary in relative abundance on the order of hundreds of permil. The ratio between these two species can be considered the hydrogen isotopic fingerprint of a substance. Understanding isotopic fingerprints and the sources of fractionation that lead to variation between them can be applied to address a diverse array of questions ranging from ecology and hydrology to geochemistry and paleoclimate reconstructions. 

Paleoreconstructions edit

The isotopic compositions of biomolecules preserved in the sedimentary record can be used as a proxy for paleoenvironment reconstructions. Since water is the primary hydrogen source for photoautotrophs, the hydrogen isotope composition of their biomass can be related to the composition of their growth water and thereby used to gain insight into some properties of ancient environments[1]. Studying hydrogen isotopes can be very valuable, as hydrogen is more directly related to climate than other relevant stable isotope systems. However, hydrogen atoms bonded to oxygen, nitrogen, or sulfur atoms are exchangeable with environmental hydrogen, which makes this system less straightforward[2] [ref to earlier H exchange section]. To study the hydrogen isotope composition of biomolecules, it is preferable to use compounds where the hydrogen is largely bound to carbon, and therefore not exchangeable on experimental timescales. By this criteria, lipids are a much better subject for hydrogen isotope studies than sugars or amino acids.

The net fractionation between source water and lipids is denoted as εl/w, and can be represented as

   

where w refers to the water, and l refers to the lipids.  

While the δD of source water is the biggest influence on the δD of lipids[3], discrepancies between fractionation factor values obtained from the slope and from the intercept of the regression suggest that the relationship is more complex than a two-pool fractionation[4]. In other words, there are multiple fractionation steps that must be taken into account in understanding the isotopic composition of lipids.  

Plant leaf waxes edit

 
Photograph of a leaf, with sunlight reflecting off of the waxy surface.

Terrestrial plants make leaf waxes to coat the surfaces of their leaves as an adaptation to minimize water loss. These waxes are comprised largely of straight-chain n-alkyl lipids. They are insoluble, non-volatile, chemically inert, and resistant to degradation, making them easily preserved in the sedimentary record, and therefore good targets as biomarkers[5]

Soil water and leaf water edit

The starting δD value of source water is highly variable, and is influenced by temperature, amount, and geography (the continental effect)[6]. The main water source for terrestrial plants is soil water, which largely resembles the hydrogen isotope composition of rain water, but varies between environments and with enrichment by precipitation, depletion by evaporation, and exchange with atmospheric water vapor. [ref to Sang's hydrology section]

There can be a significant offset between the δD value of source water and the δD value of leaf water at the site of lipid biosynthesis. No fractionation is associated with water uptake by roots, a process usually driven by capillary tension, with the one exception of xerophytes that burn ATP to pump water in extremely arid environments (with a roughly 10‰ depletion)[7]. However, leaf water can be substantially enriched relative to soil water due to transpiration, an evaporative process which is influenced by temperature, humidity, and the composition of surrounding water vapor. The leaf water hydrogen isotope composition can be described with a modified Craig-Gordon model[8], where ΔDe is the steady state enrichment of leaf water, εeq is the temperature-dependent equilibrium fractionation between liquid water and vapor, εk is the kinetic isotope effect from diffusion between leaf internal air space and the atmosphere, ΔDv is the leaf/air disequilibrium, ea is atmospheric vapor pressure, and ei is internal leaf vapor pressure.

 

The Péclet effect, which describes the opposing forces of advection and diffusion can be added to the model as

 

where E is the transpiration rate, L is the length scale of transport, C is the concentration of water, and D is the diffusion coefficient.

 

[ref to Sujung's ecohydrology section]

While the role of rain water δD as the fundamental control on the final δD of lipids is well documented[9], the importance of fractionation effects from rain water to soil water and leaf water on εl/w is appreciated but remains poorly understood[10][1].

Biochemical effects edit

Organic biomolecules are generally depleted relative to the δD of leaf water[1]. However, differences between organisms, biosynthetic pathways, and biological roles of different molecules can lead to huge variability in fractionation; the diversity of lipid biomarkers spans a 600‰ range of δD values [11].  

Lipid biosynthesis is biochemically complex, involving multiple enzyme-dependent steps that can lead to isotope fractionations. There are three major pathways of lipid biosynthesis, known as the mevalonate pathway, the acetogenic pathway, and the 1-deoxyD-xylulose-5-phosphate/2-methylerythroyl-4-phosphate pathway[12]. The acetogenic pathway is responsible for the production of n-alkyl lipids like leaf waxes, and is associated with a smaller δD depletion relative to source water than the other two lipid biosynthesis pathways[13][14]. While leaf water is the main source of hydrogen in leaf biomolecules, relatively depleted hydrogen from acetate or NADPH is often added during biosynthesis, and contributes to the hydrogen composition of the final molecule. Secondary hydrogen exchange reactions, meaning hydrogenation and dehydrogenation reactions outside of the primary biosynthetic pathway, also contribute substantially to the variability of lipid hydrogen isotope composition[15].                  

It is important to note that biological differences in fractionation stem not only from biochemical differences between different molecules, but also from physiological differences between different organisms. For example, the δD values of multiple leaf wax molecules are enriched in shrubs (median ~ -90‰) relative to trees (median ~ -135‰), which themselves are enriched relative to both C3 (median ~ -160‰) and C4 grasses (median ~ -140‰)[1]. Between indiviudal species, substantial variation of δD values have been documented[16][17][18][19]. Other physiological factors that contribute to variable leaf wax δD values include the seasonal timing of leaf development[20], response to external stress or environmental variability[21], and the presence or absence of stomata[9].                  

It can be difficult to distinguish between physiological factors and environmental factors, when many physiological adaptations are directly related to environment.                  

Environmental effects edit

Several environmental factors have been shown to contribute to leaf wax δD variability, in addition to environmental effects on the δD of source water. Humidity is known to impact lipid δD values at moderate humidity levels, but not at particularly high (>80%) or low (<40%) humidity levels, and a broad trend of enriched δD values, meaning smaller εl/w, is seen in arid regions[16][3][1]. Temperature and sunlight intensity, both correlated to geographic latitude, have strong effects on the rates of metabolism and transpiration, and by extension on εl/w[22]. Additionally, the average chain length of leaf wax molecules varies with geographic latitude, and εl/w has been shown to increase with increasing chain length[9].

When using biomarkers as a proxy for reconstructing ancient environments, it is important to be aware of the biases inherent in the sedimentary record. Leaf matter incorporated into sediment is largely deposited during the autumn, so seasonal variations in leaf waxes must be considered accordingly[9]. Furthermore, sediments average leaf waxes over lots of different plants in both space and time, making it difficult to calibrate the biological constraints on εl/w[1]. Finally, preservation of biomolecules in the geologic record does not faithfully represent whole ecosystems, and there is always the threat of hydrogen exchange, particularly if the sediments are subjected to high temperatures.

Summary of leaf wax δD as a paleoenvironment proxy edit

 
The hydrogen isotope relationships between rain water and leaf wax lipids. Adapted from Sachse et al., 2012[23].

The hydrogen isotope composition of leaf waxes can be summarized as the δD of rain water, with three main fractionation steps- evaporation from soil water, transpiration from leaf water, and lipid biosynthesis, which can be combined and measured as the net fractionation, or εl/w[1]. With the application of ever-improving measurement techniques for single molecules, and correlation with other independent proxies in the geological record that can help constrain some variables, investigating the hydrogen isotope composition of leaf waxes can be extremely productive. Leaf wax δD data has been successfully applied to improving our understanding of climate driven changes in terrestrial hydrology, by demonstrating that ocean circulation and surface temperature have a significant effect on continental precipitation[24][25]. Leaf wax δD values have also been used as records of paleoaltimetry to reconstruct the elevation gradients in ancient mountain ranges based on the effect of altitude on rain water δD[26][27].

Alkenones edit

 
Scanning electron micrograph of a single Emiliana huxleyi cell.

Another group of molecules frequently used in paleoreconstructions are alkenones, long chain largely unsaturated lipids produced exclusively by coccolithophores. Coccolithophores are marine haptophyte algae, and include the globally iconic species Emiliania huxleyi, one of the main CaCO3 producers in the ocean. The δD values of alkenones are highly correlated to the δD values of sea water, and therefore can be used to reconstruct paleoenvironmental properties that constrain the isotopic composition of sea water. The most notable reconstruction that alkenone δD values are applied to is the salinity of ancient oceans.

Alkenones as a paleosalinity proxy edit

Both the δD values of sea water and the fractionations associated with hyptophyte biochemistry (εbio) are fairly well understood, so alkenones can be readily used to observe the secondary effect of salinity on δD[28]. There is a well established positive linear correlation between salinity and εl/w, on the order of a ~3‰ change in fractionation per salinity unit[29]. Hypothesized mechanisms for this effect include enrichment of D in intracellular water due to reduced exchange with extracellular water at higher salinity[30], removal of H from intracellular water due to increased production of solutes to maintain osmotic pressure at higher salinity[31], and lower haptophyte growth rates at higher salinity[1].

Alkenone δD values have been used successfully to reconstruct past salinity changes in the Mediterranean Sea[32], Black Sea[33][34], Panama Basin[35], and Mozambique Channel[28]. As an extension of salinity, this data was also used to draw further conclusions about ancient environments, such as ancient freshwater flooding events[32][33], and the evolution of plankton in response to environmental changes[34].

References edit

  1. ^ a b c d e f g h Sachse, Dirk; Billault, Isabelle; Bowen, Gabriel J.; Chikaraishi, Yoshito; Dawson, Todd E.; Feakins, Sarah J.; Freeman, Katherine H.; Magill, Clayton R.; McInerney, Francesca A. (2012-01-01). "Molecular Paleohydrology: Interpreting the Hydrogen-Isotopic Composition of Lipid Biomarkers from Photosynthesizing Organisms". Annual Review of Earth and Planetary Sciences. 40 (1): 221–249. doi:10.1146/annurev-earth-042711-105535.
  2. ^ Schimmelmann, Arndt; Sessions, Alex L.; Mastalerz, Maria (2006-01-01). "Hydrogen Isotopic (d/H) Composition of Organic Matter During Diagenesis and Thermal Maturation". Annual Review of Earth and Planetary Sciences. 34 (1): 501–533. doi:10.1146/annurev.earth.34.031405.125011.
  3. ^ a b Hou, Juzhi; D’Andrea, William J.; Huang, Yongsong (2008-07-15). "Can sedimentary leaf waxes record D/H ratios of continental precipitation? Field, model, and experimental assessments". Geochimica et Cosmochimica Acta. 72 (14): 3503–3517. doi:10.1016/j.gca.2008.04.030.
  4. ^ Sessions, Alex L.; Hayes, John M. (2005-02-01). "Calculation of hydrogen isotopic fractionations in biogeochemical systems". Geochimica et Cosmochimica Acta. 69 (3): 593–597. doi:10.1016/j.gca.2004.08.005.
  5. ^ Eglinton, Timothy I.; Eglinton, Geoffrey (2008-10-30). "Molecular proxies for paleoclimatology". Earth and Planetary Science Letters. 275 (1–2): 1–16. doi:10.1016/j.epsl.2008.07.012.
  6. ^ Gat, J. R. (1996-01-01). "Oxygen and Hydrogen Isotopes in the Hydrologic Cycle". Annual Review of Earth and Planetary Sciences. 24 (1): 225–262. doi:10.1146/annurev.earth.24.1.225.
  7. ^ Ellsworth, Patrick Z.; Williams, David G. (2007-01-17). "Hydrogen isotope fractionation during water uptake by woody xerophytes". Plant and Soil. 291 (1–2): 93–107. doi:10.1007/s11104-006-9177-1. ISSN 0032-079X.
  8. ^ Farquhar, Graham D.; Cernusak, Lucas A.; Barnes, Belinda (2007-01-01). "Heavy water fractionation during transpiration". Plant Physiology. 143 (1): 11–18. doi:10.1104/pp.106.093278. ISSN 0032-0889. PMC 1761995. PMID 17210909.
  9. ^ a b c d Sachse, Dirk; Radke, Jens; Gleixner, Gerd (2006-04-01). "δD values of individual n-alkanes from terrestrial plants along a climatic gradient – Implications for the sedimentary biomarker record". Organic Geochemistry. 37 (4): 469–483. doi:10.1016/j.orggeochem.2005.12.003.
  10. ^ Sauer, Peter E; Eglinton, Timothy I; Hayes, John M; Schimmelmann, Arndt; Sessions, Alex L (2001-01-15). "Compound-specific D/H ratios of lipid biomarkers from sediments as a proxy for environmental and climatic conditions1". Geochimica et Cosmochimica Acta. 65 (2): 213–222. doi:10.1016/S0016-7037(00)00520-2.
  11. ^ Zhang, Xinning; Gillespie, Aimee L.; Sessions, Alex L. (2009-08-04). "Large D/H variations in bacterial lipids reflect central metabolic pathways". Proceedings of the National Academy of Sciences. 106 (31): 12580–12586. doi:10.1073/pnas.0903030106. ISSN 0027-8424. PMC 2722351. PMID 19617564.
  12. ^ Chikaraishi, Yoshito; Naraoka, Hiroshi; Poulson, Simon R. (2004-05-01). "Hydrogen and carbon isotopic fractionations of lipid biosynthesis among terrestrial (C3, C4 and CAM) and aquatic plants". Phytochemistry. 65 (10): 1369–1381. doi:10.1016/j.phytochem.2004.03.036.
  13. ^ Chikaraishi, Yoshito; Naraoka, Hiroshi (2003-06-01). "Compound-specific δD–δ13C analyses of n-alkanes extracted from terrestrial and aquatic plants". Phytochemistry. 63 (3): 361–371. doi:10.1016/S0031-9422(02)00749-5.
  14. ^ Sessions, Alex L.; Burgoyne, Thomas W.; Schimmelmann, Arndt; Hayes, John M. (1999-09-01). "Fractionation of hydrogen isotopes in lipid biosynthesis". Organic Geochemistry. 30 (9): 1193–1200. doi:10.1016/S0146-6380(99)00094-7.
  15. ^ Behrouzian, B.; Buist, P.H. (2003-02-01). "Mechanism of fatty acid desaturation: a bioorganic perspective". Prostaglandins, Leukotrienes and Essential Fatty Acids. 68 (2): 107–112. doi:10.1016/s0952-3278(02)00260-0. ISSN 0952-3278.
  16. ^ a b Feakins, Sarah J.; Sessions, Alex L. (2010-12-01). "Crassulacean acid metabolism influences D/H ratio of leaf wax in succulent plants". Organic Geochemistry. 41 (12): 1269–1276. doi:10.1016/j.orggeochem.2010.09.007.
  17. ^ Chikaraishi, Yoshito; Naraoka, Hiroshi (2007-02-01). "δ13C and δD relationships among three n-alkyl compound classes (n-alkanoic acid, n-alkane and n-alkanol) of terrestrial higher plants". Organic Geochemistry. 38 (2): 198–215. doi:10.1016/j.orggeochem.2006.10.003.
  18. ^ Hou, Juzhi; D’Andrea, William J.; MacDonald, Dana; Huang, Yongsong (2007-06-01). "Hydrogen isotopic variability in leaf waxes among terrestrial and aquatic plants around Blood Pond, Massachusetts (USA)". Organic Geochemistry. 38 (6): 977–984. doi:10.1016/j.orggeochem.2006.12.009.
  19. ^ Pedentchouk, Nikolai; Sumner, William; Tipple, Brett; Pagani, Mark (2008-08-01). "δ13C and δD compositions of n-alkanes from modern angiosperms and conifers: An experimental set up in central Washington State, USA". Organic Geochemistry. Advances in Organic Geochemistry 2007Proceedings of the 23rd International Meeting on Organic Geochemistry. 39 (8): 1066–1071. doi:10.1016/j.orggeochem.2008.02.005.
  20. ^ Kahmen, Ansgar; Dawson, Todd E.; Vieth, Andrea; Sachse, Dirk (2011-10-01). "Leaf wax n-alkane δD values are determined early in the ontogeny of Populus trichocarpa leaves when grown under controlled environmental conditions". Plant, Cell & Environment. 34 (10): 1639–1651. doi:10.1111/j.1365-3040.2011.02360.x. ISSN 1365-3040.
  21. ^ Sachse, Dirk; Kahmen, Ansgar; Gleixner, Gerd (2009-06-01). "Significant seasonal variation in the hydrogen isotopic composition of leaf-wax lipids for two deciduous tree ecosystems (Fagus sylvativa and Acerpseudoplatanus)". Organic Geochemistry. 40 (6): 732–742. doi:10.1016/j.orggeochem.2009.02.008.
  22. ^ Yang, Hong; Pagani, Mark; Briggs, Derek E. G.; Equiza, M. A.; Jagels, Richard; Leng, Qin; LePage, Ben A. (2009-04-08). "Carbon and hydrogen isotope fractionation under continuous light: implications for paleoenvironmental interpretations of the High Arctic during Paleogene warming". Oecologia. 160 (3): 461–470. doi:10.1007/s00442-009-1321-1. ISSN 0029-8549.
  23. ^ Sachse, Dirk; Billault, Isabelle; Bowen, Gabriel J.; Chikaraishi, Yoshito; Dawson, Todd E.; Feakins, Sarah J.; Freeman, Katherine H.; Magill, Clayton R.; McInerney, Francesca A. (2012-01-01). "Molecular Paleohydrology: Interpreting the Hydrogen-Isotopic Composition of Lipid Biomarkers from Photosynthesizing Organisms". Annual Review of Earth and Planetary Sciences. 40 (1): 221–249. doi:10.1146/annurev-earth-042711-105535.
  24. ^ Schefuß, Enno; Schouten, Stefan; Schneider, Ralph R. "Climatic controls on central African hydrology during the past 20,000 years". Nature. 437 (7061): 1003–1006. doi:10.1038/nature03945.
  25. ^ Tierney, Jessica E.; Russell, James M.; Huang, Yongsong; Damsté, Jaap S. Sinninghe; Hopmans, Ellen C.; Cohen, Andrew S. (2008-10-10). "Northern Hemisphere Controls on Tropical Southeast African Climate During the Past 60,000 Years". Science. 322 (5899): 252–255. doi:10.1126/science.1160485. ISSN 0036-8075. PMID 18787132.
  26. ^ Polissar, Pratigya J.; Freeman, Katherine H.; Rowley, David B.; McInerney, Francesca A.; Currie, Brian S. (2009-09-30). "Paleoaltimetry of the Tibetan Plateau from D/H ratios of lipid biomarkers". Earth and Planetary Science Letters. 287 (1–2): 64–76. doi:10.1016/j.epsl.2009.07.037.
  27. ^ Hren, M. T.; Pagani, M.; Erwin, D. M.; Brandon, M. "Biomarker reconstruction of the early Eocene paleotopography and paleoclimate of the northern Sierra Nevada". Geology. 38 (1): 7–10. doi:10.1130/g30215.1.
  28. ^ a b Kasper, Sebastian; van der Meer, Marcel T. J.; Castañeda, Isla S.; Tjallingii, Rik; Brummer, Geert-Jan A.; Sinninghe Damsté, Jaap S.; Schouten, Stefan (2015-01-01). "Testing the alkenone D/H ratio as a paleo indicator of sea surface salinity in a coastal ocean margin (Mozambique Channel)". Organic Geochemistry. 78: 62–68. doi:10.1016/j.orggeochem.2014.10.011.
  29. ^ Schouten, S.; Ossebaar, J.; Schreiber, K.; Kienhuis, M. V. M.; Langer, G.; Benthien, A.; Bijma, J. (2006-03-15). "The effect of temperature, salinity and growth rate on the stable hydrogen isotopic composition of long chain alkenones produced by Emiliania huxleyi and Gephyrocapsa oceanica". Biogeosciences. 3 (1): 113–119. doi:10.5194/bg-3-113-2006. ISSN 1726-4189.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  30. ^ Sachse, Dirk; Sachs, Julian P. (2008-02-01). "Inverse relationship between D/H fractionation in cyanobacterial lipids and salinity in Christmas Island saline ponds". Geochimica et Cosmochimica Acta. 72 (3): 793–806. doi:10.1016/j.gca.2007.11.022.
  31. ^ Sachs, Julian P.; Schwab, Valérie F. (2011-01-15). "Hydrogen isotopes in dinosterol from the Chesapeake Bay estuary". Geochimica et Cosmochimica Acta. 75 (2): 444–459. doi:10.1016/j.gca.2010.10.013.
  32. ^ a b van der Meer, Marcel T. J.; Baas, Marianne; Rijpstra, W. Irene C.; Marino, Gianluca; Rohling, Eelco J.; Sinninghe Damsté, Jaap S.; Schouten, Stefan (2007-10-30). "Hydrogen isotopic compositions of long-chain alkenones record freshwater flooding of the Eastern Mediterranean at the onset of sapropel deposition". Earth and Planetary Science Letters. 262 (3–4): 594–600. doi:10.1016/j.epsl.2007.08.014.
  33. ^ a b van der Meer, Marcel T. J.; Sangiorgi, Francesca; Baas, Marianne; Brinkhuis, Henk; Sinninghe Damsté, Jaap S.; Schouten, Stefan (2008-03-30). "Molecular isotopic and dinoflagellate evidence for Late Holocene freshening of the Black Sea". Earth and Planetary Science Letters. 267 (3–4): 426–434. doi:10.1016/j.epsl.2007.12.001.
  34. ^ a b Coolen, Marco J. L.; Orsi, William D.; Balkema, Cherel; Quince, Christopher; Harris, Keith; Sylva, Sean P.; Filipova-Marinova, Mariana; Giosan, Liviu (2013-05-21). "Evolution of the plankton paleome in the Black Sea from the Deglacial to Anthropocene". Proceedings of the National Academy of Sciences of the United States of America. 110 (21): 8609–8614. doi:10.1073/pnas.1219283110. ISSN 1091-6490. PMC 3666672. PMID 23650351.
  35. ^ Pahnke, Katharina; Sachs, Julian P.; Keigwin, Lloyd; Timmermann, Axel; Xie, Shang-Ping (2007-12-01). "Eastern tropical Pacific hydrologic changes during the past 27,000 years from D/H ratios in alkenones". Paleoceanography. 22 (4): PA4214. doi:10.1029/2007PA001468. ISSN 1944-9186.