Hostname: page-component-848d4c4894-wg55d Total loading time: 0 Render date: 2024-04-30T10:38:22.897Z Has data issue: false hasContentIssue false

Multi-Substrate Radiocarbon Data Constrain Detrital and Reservoir Effects in Holocene Sediments of the Great Salt Lake, Utah

Published online by Cambridge University Press:  14 June 2019

Gabriel J Bowen*
Affiliation:
Department of Geology and Geophysics and Global Change and Sustainability Center, University of Utah, Salt Lake City, UT 84112, USA Department of Earth, Atmospheric, and Planetary Sciences; Purdue University, West Lafayette, IN 47907, USA
Kristine E Nielson
Affiliation:
Department of Earth, Atmospheric, and Planetary Sciences; Purdue University, West Lafayette, IN 47907, USA Department of Earth Sciences, University of Adelaide, Adelaide, 5005, Australia
Timothy I Eglinton
Affiliation:
Geologisches Institut, ETH Zürich, 8092 Zürich, Switzerland Department of Marine Chemistry and Geochemistry, Woods Hole Oceanographic Institution, Woods Hole, MA 02543, USA
*
*Corresponding author. Email: gabe.bowen@utah.edu.

Abstract

The radiocarbon (14C) content of simultaneously deposited substrates in lacustrine archives may differ due to reservoir and detrital effects, complicating the development of age models and interpretation of proxy records. Multi-substrate 14C studies quantifying these effects remain rare, however, particularly for large, terminal lake systems, which are excellent recorders of regional hydroclimate change. We report 14C ages of carbonates, brine shrimp cysts, algal mat biomass, total organic carbon (TOC), terrestrial macrofossils, and n-alkane biomarkers from Holocene sediments of the Great Salt Lake (GSL), Utah. 14C ages for co-deposited aquatic organic substrates are generally consistent, with small offsets that may reflect variable terrestrial organic matter inputs to the system. Carbonates and long-chain n-alkanes derived from vascular plants, however, are ∼1000–4000 14C years older than other substrates, reflecting deposition of pre-aged detrital materials. All lacustrine substrates are 14C-depleted compared to terrestrial macrofossils, suggesting that the reservoir age of the GSL was > 1200 years throughout most of the Holocene, far greater than the modern reservoir age of the lake (∼300 years). These results suggest good potential for multi-substrate paleoenvironmental reconstruction from Holocene GSL sediments but point to limitations including reservoir-induced uncertainty in 14C chronologies and attenuation and time-shifting of some proxy signals due to detrital effects.

Type
Research Article
Copyright
© 2019 by the Arizona Board of Regents on behalf of the University of Arizona 

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

REFERENCES

Abbott, MB, Stafford, TW. 1996. Radiocarbon geochemistry of modern and ancient Arctic lake systems, Baffin Island, Canada. Quaternary Research 45(3):300311.CrossRefGoogle Scholar
Ascough, P, Cook, G, Dugmore, A. 2005. Methodological approaches to determining the marine radiocarbon reservoir effect. Progress in Physical Geography: Earth and Environment 29(4):532547.CrossRefGoogle Scholar
Ascough, P, Cook, G, Hastie, H, Dunbar, E, Church, M, Einarsson, Á, McGovern, T, Dugmore, A. 2011. An Icelandic freshwater radiocarbon reservoir effect: implications for lacustrine 14C chronologies. The Holocene 21(7):10731080.CrossRefGoogle Scholar
Balch, DP, Cohen, AS, Schnurrenberger, DW, Haskell, BJ, Garces, BLV, Beck, JW, Cheng, H, Edwards, RL. 2005. Ecosystem and paleohydrological response to Quaternary climate change in the Bonneville Basin, Utah. Palaeogeography, Palaeoclimatology, Palaeoecology 221(1–2):99122.CrossRefGoogle Scholar
Blaauw, M, Christen, JA. 2011. Flexible paleoclimate age-depth models using an autoregressive gamma process. Bayesian Analysis 6(3):457474.Google Scholar
Blair, NE, Leithold, EL, Aller, RC. 2004. From bedrock to burial: the evolution of particulate organic carbon across coupled watershed-continental margin systems. Marine Chemistry 92:141156.CrossRefGoogle Scholar
Blair, NE, Leithold, EL, Ford, ST, Peeler, KA, Holmes, JC, Perkey, DW. 2003. The persistence of memory: The fate of ancient sedimentary organic carbon in a modern sedimentary system. Geochimica et Cosmochimica Acta 67(1):6373.CrossRefGoogle Scholar
Blumer, M, Guillard, R, Chase, T. 1971. Hydrocarbons of marine phytoplankton. Marine Biology 8(3):183189.CrossRefGoogle Scholar
Broecker, WS, Kaufman, A. 1965. Radiocarbon chronology of Lake Lahontan and Lake Bonneville II, Great Basin. Geological Society of America Bulletin 76(5):537566.CrossRefGoogle Scholar
Broecker, WS, Walton, A. 1959. The geochemistry of C14 in fresh-water systems. Geochimica et Cosmochimica Acta 16(1–3):1538.CrossRefGoogle Scholar
Bush, RT, McInerney, FA. 2013. Leaf wax n-alkane distributions in and across modern plants: implications for paleoecology and chemotaxonomy. Geochimica et Cosmochimica Acta 117:161179.CrossRefGoogle Scholar
Caraco, N, Bauer, JE, Cole, JJ, Petsch, S, Raymond, P. 2010. Millennial-aged organic carbon subsidies to a modern river food web. Ecology 91(8):23852393.CrossRefGoogle ScholarPubMed
Collister, JW, Schamel, S. 2002. Lipid composition of recent sediments from the Great Salt Lake. In: Gwynn, JW, editor. Great Salt lake—an overview of change. Salt Lake City: Utah Department of Natural Resources. p. 127142.Google Scholar
Cook, GT, Bonsall, C, Hedges, RE, McSweeney, K, Boronean, V, Pettitt, PB. 2001. A freshwater diet-derived 14C reservoir effect at the Stone Age sites in the Iron Gates gorge. Radiocarbon 43(2A):453460.CrossRefGoogle Scholar
Cotton, JM, Cerling, TE, Hoppe, KA, Mosier, TM, Still, CJ. 2016. Climate, CO2, and the history of North American grasses since the Last Glacial Maximum. Science Advances 2(3):e1501346.CrossRefGoogle ScholarPubMed
Currey, DR, Atwood, G, Mabey, DR, Roy, JS, Brown, KD. 1984. Major levels of Great Salt Lake and Lake Bonneville. Utah Geological and Mineral Survey Map 73.Google Scholar
Dean, WE, Rosenbaum, J, Kaufman, D. 2009. Endogenic carbonate sedimentation in Bear Lake, Utah and Idaho, over the last two glacial-interglacial cycles. In: Rosenbaum, J, Kaufman, D, editors. Paleoenvironments of Bear Lake and its catchment: Geological Society of America Special Paper Special Paper 450. Boulder, CO: Geological Society of America. p. 169196.Google Scholar
Deevey, ES, Gross, MS, Hutchinson, GE, Kraybill, HL. 1954. The natural C14 contents of materials from hard-water lakes. Proceedings of the National Academy of Sciences 40(5):285288.CrossRefGoogle Scholar
DeLucia, EH, Schlesinger, WH. 1991. Resource-use efficiency and drought tolerance in adjacent Great Basin and Sierran plants. Ecology 72(1):5158.CrossRefGoogle Scholar
Diefendorf, AF, Freeman, KH, Wing, SL, Graham, HV. 2011. Production of n-alkyl lipids in living plants and implications for the geologic past. Geochimica et Cosmochimica Acta 75(23):74727485.CrossRefGoogle Scholar
Domagalski, JL, Orem, WH, Eugster, HP. 1989. Organic geochemistry and brine composition in Great Salt, Mono, and Walker Lakes. Geochimica et Cosmochimica Acta 53(11):28572872.CrossRefGoogle Scholar
Douglas, PMJ, Pagani, M, Eglinton, TI, Brenner, M, Hodell, DA, Curtis, JH, Ma, KF, Breckenridge, A. 2014. Pre-aged plant waxes in tropical lake sediments and their influence on the chronology of molecular paleoclimate proxy records. Geochimica et Cosmochimica Acta 141:346364.CrossRefGoogle Scholar
Drenzek, NJ, Montluçon, DB, Yunker, MB, Macdonald, RW, Eglinton, TI. 2007. Constraints on the origin of sedimentary organic carbon in the Beaufort Sea from coupled molecular 13C and 14C measurements. Marine Chemistry 103(1–2):146162.CrossRefGoogle Scholar
Edwards, KJ, Whittington, G. 2001. Lake sediments, erosion and landscape change during the Holocene in Britain and Ireland. Catena 42(2-4):143173.CrossRefGoogle Scholar
Eglinton, TI, Aluwihare, LI, Bauer, JE, Druffel, ER, McNichol, AP. 1996. Gas chromatographic isolation of individual compounds from complex matrices for radiocarbon dating. Analytical Chemistry 68(5):904912.CrossRefGoogle ScholarPubMed
Eglinton, TI, Benitez, Nelson BC, Pearson, A, McNichol, AP, Bauer, JE, Druffel, ERM. 1997. Variability in radiocarbon ages of individual organic compounds from marine sediments. Science 277(5327):796799.CrossRefGoogle Scholar
Eglinton, TI, Eglinton, G. 2008. Molecular proxies for paleoclimatology. Earth & Planetary Science Letters 275:116.CrossRefGoogle Scholar
Fernandes, R, Dreves, A, Nadeau, M-J, Grootes, PM. 2013. A freshwater lake saga: carbon routing within the aquatic food web of Lake Schwerin. Radiocarbon 55(3):11021113.CrossRefGoogle Scholar
Ficken, KJ, Li, B, Swain, DL, Eglinton, G. 2000. An n-alkane proxy for the sedimentary input of submerged/floating freshwater aquatic macrophytes. Organic Geochemistry 31(7-8):745749.CrossRefGoogle Scholar
Gelman, A, Rubin, DB. 1992. Inference from iterative simulation using multiple sequences. Statistical Science 7(4):457472.CrossRefGoogle Scholar
Gelpi, E, Schneider, H, Mann, J, Oro, J. 1970. Hydrocarbons of geochemical significance in microscopic algae. Phytochemistry 9(3):603612.CrossRefGoogle Scholar
Geyh, MA, Schotterer, U, Grosjean, M. 1997. Temporal changes of the 14C reservoir effect in lakes. Radiocarbon 40(2):921931.CrossRefGoogle Scholar
Gostling, NJ, Dong, X, Donoghue, PCJ. 2009. Ontogeny and taphonomy: an experimental taphonomy study of the development of the brine shrimp Artemia salina. Palaeontology 52(1):169186.CrossRefGoogle Scholar
Hajdas, I, Zolitschka, B, Ivy-Ochs, SD, Beer, J, Bonani, G, Leroy, SA, Negendank, JW, Ramrath, M, Suter, M. 1995. AMS radiocarbon dating of annually laminated sediments from Lake Holzmaar, Germany. Quaternary Science Reviews 14(2):137143.CrossRefGoogle Scholar
Hostetler, S, Giorgi, F, Bates, G, Bartlein, P. 1994. Lake-atmosphere feedbacks associated with paleolakes Bonneville and Lahontan. Science 263(5147):665668.CrossRefGoogle ScholarPubMed
Hua, Q, Barbetti, M, Rakowski, AZ. 2013. Atmospheric radiocarbon for the period 1950–2010. Radiocarbon 55(4):20592072.CrossRefGoogle Scholar
Ishikawa, NF, Uchida, M, Shibata, Y, Tayasu, I. 2014. Carbon storage reservoirs in watersheds support stream food webs via periphyton production. Ecology 95(5):12641271.CrossRefGoogle ScholarPubMed
Keaveney, EM, Reimer, PJ, Foy, RH. 2015. Young, old, and weathered carbon—part 2: using radiocarbon and stable isotopes to identify terrestrial carbon support of the food web in an alkaline, humic lake. Radiocarbon 57(3):425438.CrossRefGoogle Scholar
Kusch, S, Eglinton, TI, Mix, AC, Mollenhauer, G. 2010a. Timescales of lateral sediment transport in the Panama Basin as revealed by radiocarbon ages of alkenones, total organic carbon and foraminifera. Earth and Planetary Science Letters 290(3):340350.CrossRefGoogle Scholar
Kusch, S, Rethemeyer, J, Schefuß, E, Mollenhauer, G. 2010b. Controls on the age of vascular plant biomarkers in Black Sea sediments. Geochimica et Cosmochimica Acta 74(24):70317047.CrossRefGoogle Scholar
Kusumgar, S, Agrawal, DP, Bhandari, N, Deshpande, RD, Raina, A, Sharma, C, Yadava, MG. 1992. Lake sediments from the Kashmir Himalayas: inverted 14C chronology and its implications. Radiocarbon 34(3):561565.CrossRefGoogle Scholar
Leenheer, JA, Noyes, TI, Rostad, CE, Davisson, ML. 2004. Characterization and origin of polar dissolved organic matter from the Great Salt Lake. Biogeochemistry 69:125141.CrossRefGoogle Scholar
MacDonald, GM, Beukens, RP, Kieser, WE, Vitt, DH. 1987. Comparative radiocarbon dating of terrestrial plant macrofossils and aquatic moss from the “ice-free corridor” of western Canada. Geology 15(9):837840.2.0.CO;2>CrossRefGoogle Scholar
McKenzie, JA, Eberli, GP. 1985. Late Holocene lake-level fluctuations of the Great Salt Lake (Utah) as deduced from oxygen isotope and carbonate content of cored sediments. In: Center for Public Affairs and Administration UoU, editor. Problems of and Prospects for Predicting Great Salt Lake levels. p. 25–39.Google Scholar
McKenzie, JA, Eberli, GP. 1987. Indications for abrupt Holocene climatic change: Late Holocene oxygen isotope stratigraphy of the Great Salt Lake, Utah. In: Berger, WH, Labeyrie, LD, editors. Abrupt Climatic Change. D. Reidel Publishing Company. p. 127136.CrossRefGoogle Scholar
Mohebbi, F. 2010. The Brine Shrimp Artemia and hypersaline environments microalgal composition: a mutual interaction. International Journal of Aquatic Science 1(1):1927.Google Scholar
Mollenhauer, G, Inthorn, M, Vogt, T, Zabel, M, Sinninghe, Damste JS, Eglinton, T. 2007. Aging of marine organic matter during cross-shelf lateral transport in the Benguela upwelling system revealed by compound-specific radiocarbon dating.Geochemistry Geophysics Geosystems 8:Q09004.CrossRefGoogle Scholar
Mollenhauer, G, McManus, JF, Wagner, T, McCave, IN, Eglinton, TI. 2011. Radiocarbon and 230Th data reveal rapid redistribution and temporal changes in sediment focussing at a North Atlantic drift. Earth and Planetary Science Letters 301(1):373381.CrossRefGoogle Scholar
Moroney, JV, Somanchi, A. 1999. How do algae concentrate CO2 to increase the efficiency of photosynthetic carbon fixation? Plant Physiology 119(1):916.CrossRefGoogle ScholarPubMed
Murchison, SB. 1989. Fluctuation history of Great Salt Lake, Utah, during the last 13,000 years. [Salt Lake City, UT]: University of Utah.Google Scholar
Ohkouchi, N, Eglinton, TI, Keigwin, LD, Hayes, JM. 2002. Spatial and temporal offsets between proxy records in a sediment drift. Science 298(5596):12241227.CrossRefGoogle Scholar
Olsen, J, Heinemeier, J, Lübke, H, Lüth, F, Terberger, T. 2010. Dietary habits and freshwater reservoir effects in bones from a Neolithic NE German cemetery. Radiocarbon 52(2):635644.CrossRefGoogle Scholar
Oviatt, CG, Madsen, DB, Miller, DM, Thompson, RS, McGeehin, JP. 2015. Early Holocene Great Salt Lake, USA. Quaternary Research 84(1):5768.CrossRefGoogle Scholar
Pancost, RD, Boot, CS. 2004. The palaeoclimatic utility of terrestrial biomarkers in marine sediments. Marine Chemistry 92(1–4):239261.CrossRefGoogle Scholar
Pearson, A, McNichol, AP, Benitez-Nelson, BC, Hayes, JM, Eglinton, TI. 2001. Origins of lipid biomarkers in Santa Monica Basin surface sediment: a case study using compound-specific Δ14C analysis. Geochimica et Cosmochimica Acta 65(18):31233137.10.1016/S0016-7037(01)00657-3CrossRefGoogle Scholar
Prospero, JM, Ginoux, P, Torres, O, Nicholson, SE, Gill, TE. 2002. Environmental characterization of global sources of atmospheric soil dust identified with the Nimbus 7 Total Ozone Mapping Spectrometer. Reviews of Geophysics 40(1002):131.CrossRefGoogle Scholar
R Core Team. 2018. R: A language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria. https://www.R-project.org/.Google Scholar
Raymond, PA, Bauer, JE. 2001. Use of 14C and 13C natural abundances for evaluating riverine, estuarine, and coastal DOC and POC sources and cycling: a review and synthesis. Organic Geochemistry 32(4):469485.CrossRefGoogle Scholar
Reimer, PJ, Bard, E, Bayliss, A, Beck, JW, Blackwell, PG, Ramsey, CB, Buck, CE, Cheng, H, Edwards, RL, Friedrich, M. 2013. IntCal13 and Marine13 radiocarbon age calibration curves 0–50, 000 years cal BP. Radiocarbon 55(4):18691887.CrossRefGoogle Scholar
Schreuder, LT, Stuut, J-BW, Korte, LF, Sinninghe, Damsté JS, Schouten, S. 2018. Aeolian transport and deposition of plant wax n-alkanes across the tropical North Atlantic Ocean. Organic Geochemistry 115:113123.CrossRefGoogle Scholar
Simoneit, BR, Cardoso, J, Robinson, N. 1991. An assessment of terrestrial higher molecular weight lipid compounds in aerosol particulate matter over the South Atlantic from about 30–70 S. Chemosphere 23(4):447465.CrossRefGoogle Scholar
Soulet, G, Skinner, LC, Beaupré, SR, Galy, V. 2016. A Note on reporting of reservoir 14C disequilibria and age offsets. Radiocarbon 58(1):205211.CrossRefGoogle Scholar
Spencer, RJ, Baedecker, MJ, Eugster, HP, et al. 1984. Great Salt Lake, and precursors, Utah: the last 30, 000 years. Contributions to Mineralogy and Petrology 86:321334.CrossRefGoogle Scholar
Spencer, RJ, Eugster, HP, Jones, BF. 1985a. Geochemistry of Great Salt Lake, Utah: II: Pleistocene-Holocene evolution. Geochimica et Cosmochimica Acta 49:739747.CrossRefGoogle Scholar
Spencer, RJ, Eugster, HP, Jones, BF, Rettig, SL. 1985b. Geochemistry of Great Salt Lake, Utah: I: Hydrochemistry since 1850. Geochimica et Cosmochimica Acta 49:727737.CrossRefGoogle Scholar
Stephens, DW, Gillespie, DM. 1976. Phytoplankton production in the Great Salt Lake, Utah, and a laboratory study of algal response to enrichment. Limnology and Oceanography 21(1):7487.CrossRefGoogle Scholar
Thompson, RS, Toolin, LJ, Forester, RM, Spencer, RJ. 1990. Accelerator-mass spectrometer (AMS) radiocarbon dating of Pleistocene lake sediments in the Great Basin. Palaeogeography, Palaeoclimatology, Palaeoecology 78(3):301313.CrossRefGoogle Scholar
Villanueva, J, Hastings, DW. 2000. A century-scale record of the preservation of chlorophyll and its transformation products in anoxic sediments. Geochimica et Cosmochimica Acta 64(13):22812294.CrossRefGoogle Scholar
Wurtsbaugh, WA, Berry, TS. 1990. Cascading effects of decreased salinity on the plankton chemistry, and physics of the Great Salt Lake (Utah). Canadian Journal of Fisheries and Aquatic Sciences 47(1):100109.CrossRefGoogle Scholar
Wurtsbaugh, WA, Gliwicz, ZM. 2001. Limnological control of brine shrimp population dynamics and cyst production in the Great Salt Lake, Utah. Saline Lakes. Springer. p. 119132.Google Scholar