Hostname: page-component-848d4c4894-hfldf Total loading time: 0 Render date: 2024-05-19T14:43:52.246Z Has data issue: false hasContentIssue false

A downcore increase in time averaging is the null expectation from the transit of death assemblages through a mixed layer

Published online by Cambridge University Press:  19 January 2023

Adam Tomašových*
Affiliation:
Earth Science Institute, Slovak Academy of Sciences, 84005 Bratislava, Slovakia. E-mail: geoltoma@savba.sk
Susan M. Kidwell
Affiliation:
Department of Geophysical Sciences, University of Chicago, 5734 S. Ellis Avenue, Chicago, Illinois 60637, U.S.A. E-mail: skidwell@uchicago.edu
Ran Dai
Affiliation:
Department of Biostatistics, University of Nebraska Medical Center, Omaha, Nebraska 68198-4375, U.S.A. E-mail: ran.dai@unmc.edu
*
*Corresponding author.

Abstract

Understanding how time averaging changes during burial is essential for using Holocene and Anthropocene cores to analyze ecosystem change, given the many ways in which time averaging affects biodiversity measures. Here, we use transition-rate matrices to explore how the extent of time averaging changes downcore as shells transit through a taphonomically complex mixed layer into permanently buried historical layers: this is a null model, without any temporal changes in rates of sedimentation or bioturbation, to contrast with downcore patterns that might be produced by human activity. Assuming stochastic burial and exhumation movements of shells between increments within the mixed layer and stochastic disintegration within increments, we find that almost all combinations of net sedimentation, mixing, and disintegration produce a downcore increase in time averaging (interquartile range [IQR] of shell ages), this trend is typically associated with a decrease in kurtosis and skewness and by a shift from right-skewed to symmetrical age distributions. A downcore increase in time averaging is thus the null expectation wherever bioturbation generates an internally structured mixed layer (i.e., a surface, well-mixed layer is underlain by an incompletely mixed layer): under these conditions, shells are mixed throughout the entire mixed layer at a slower rate than they are buried below it by sedimentation. This downcore trend created by mixing is further amplified by the downcore decline in disintegration rate. We find that transition-rate matrices accurately reproduce the downcore changes in IQR, skewness, and kurtosis observed in bivalve assemblages from the southern California shelf. The right-skewed shell age-frequency distributions typical of surface death assemblages—the focus of most actualistic research—might be fossilized under exceptional conditions of episodic anoxia or sudden burial. However, such right-skewed assemblages will typically not survive transit through the surface mixed layer into subsurface historical layers: they are geologically transient. The deep-time fossil record will be dominated instead by the more time-averaged assemblages with weakly skewed age distributions that form in the lower parts of the mixed layer.

Type
Article
Copyright
Copyright © The Author(s), 2023. Published by Cambridge University Press on behalf of The Paleontological Society

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

Literature Cited

Agiadi, K., Azzarone, M., Hua, Q., Kaufman, D. S., Thivaiou, D., and Albano, P. G.. 2022. The taphonomic clock in fish otoliths. Paleobiology 48:154170.10.1017/pab.2021.30CrossRefGoogle Scholar
Albano, P. G., Hua, Q., Kaufman, D. S., Tomašových, A., Zuschin, M., and Agiadi, K.. 2020. Radiocarbon dating supports bivalve-fish age coupling along a bathymetric gradient in high-resolution paleoenvironmental studies. Geology 48:589593.10.1130/G47210.1CrossRefGoogle Scholar
Albano, P. G., Steger, J., Bošnjak, M., Dunne, B., Guifarro, Z., Turapova, E., Hua, Q., Kaufman, D. S., Rilov, G., and Zuschin, M.. 2021. Native biodiversity collapse in the eastern Mediterranean. Proceedings of the Royal Society of London B 288:20202469.Google ScholarPubMed
Aller, R. C. 1982. Carbonate dissolution in nearshore terrigenous muds: the role of physical and biological reworking. Journal of Geology 90:7995.10.1086/628652CrossRefGoogle Scholar
Aller, R. C. 1994. Bioturbation and remineralization of sedimentary organic matter: effects of redox oscillation. Chemical Geology 114:331345.10.1016/0009-2541(94)90062-0CrossRefGoogle Scholar
Anderson, L. C., and McBride, R. A.. 1996. Taphonomic and paleoenvironmental evidence of Holocene shell-bed genesis and history on the northeastern Gulf of Mexico shelf. Palaios 11:532–549.10.2307/3515189CrossRefGoogle Scholar
Aquino, T., Roche, K. R., Aubeneau, A., Packman, A. I., and Bolster, D.. 2017. A process-based model for bioturbation-induced mixing. Scientific Reports 7:14287.10.1038/s41598-017-14705-1CrossRefGoogle ScholarPubMed
Aronson, R. B., Macintyre, I. G., and Precht, W. F.. 2005. Event preservation in lagoonal reef systems. Geology 33:717720.CrossRefGoogle Scholar
Baldwin, C. T., and McCave, I. N.. 1999. Bioturbation in an active deep-sea area; implications for models of trace fossil tiering. Palaios, 14:375388.10.2307/3515463CrossRefGoogle Scholar
Barker, S., Broecker, W., Clark, E. and Hajdas, I.. 2007. Radiocarbon age offsets of foraminifera resulting from differential dissolution and fragmentation within the sedimentary bioturbated layer. Paleoceanography 22:PA2205.10.1029/2006PA001354CrossRefGoogle Scholar
Bentley, S. J., and Nittrouer, C. A.. 1999. Physical and biological influences on the formation of sedimentary fabric in an oxygen-restricted depositional environment; Eckernforde Bay, southwestern Baltic Sea. Palaios 14:585600.10.2307/3515315CrossRefGoogle Scholar
Bentley, S. J., Sheremet, A., and Jaeger, J. M.. 2006. Event sedimentation, bioturbation, and preserved sedimentary fabric: field and model comparisons in three contrasting marine settings. Continental Shelf Research 26:21082124.CrossRefGoogle Scholar
Berger, W. H., and Heath, G. R.. 1968. Vertical mixing in pelagic sediments. Journal of Marine Research 26:134143.Google Scholar
Bergmann, K. D., Grotzinger, J. P., and Fischer, W. W.. 2013. Biological influences on seafloor carbonate precipitation. Palaios 28:99115.CrossRefGoogle Scholar
Bertics, V. J., and Ziebis, W.. 2010. Bioturbation and the role of microniches for sulfate reduction in coastal marine sediments. Environmental Microbiology 12:30223034.10.1111/j.1462-2920.2010.02279.xCrossRefGoogle ScholarPubMed
Best, M. M., Ku, T. C., Kidwell, S. M., and Walter, L. M.. 2007. Carbonate preservation in shallow marine environments: unexpected role of tropical siliciclastics. Journal of Geology 115:437456.10.1086/518051CrossRefGoogle Scholar
Bottjer, D. J., and Ausich, W. I.. 1986. Phanerozoic development of tiering in soft substrata suspension-feeding communities. Paleobiology 12:400420.10.1017/S0094837300003134CrossRefGoogle Scholar
Boudreau, B. P. 1998. Mean mixed depth of sediments: the wherefore and the why. Limnology and Oceanography 43:524526.10.4319/lo.1998.43.3.0524CrossRefGoogle Scholar
Bradshaw, C., and Scoffin, T. P.. 2001. Differential preservation of gravel-sized bioclasts in alpheid-versus callianassid-bioturbated muddy reefal sediments. Palaios 16:185191.10.1669/0883-1351(2001)016<0185:DPOGSB>2.0.CO;22.0.CO;2>CrossRefGoogle Scholar
Braga, J. C., Puga-Bernabéu, Á., Heindel, K., Patterson, M. A., Birgel, D., Peckmann, J., Sánchez-Almazo, I. M., Webster, J. M., Yokoyama, Y., and Riding, R.. 2019. Microbialites in last glacial maximum and deglacial reefs of the Great Barrier Reef (IODP Expedition 325, NE Australia). Palaeogeography, Palaeoclimatology, Palaeoecology 514:117.CrossRefGoogle Scholar
Brett, C. E., IV, J. J. Zambito, Hunda, B. R., and Schindler, E.. 2012. Mid-Paleozoic trilobite Lagerstätten: models of diagenetically enhanced obrution deposits. Palaios 27:326345.CrossRefGoogle Scholar
Broecker, W., and Clark, E.. 2011. Radiocarbon age differences among coexisting planktic foraminifera shells: the Barker effect. Paleoceanography 26:PA2222.10.1029/2011PA002116CrossRefGoogle Scholar
Brush, G. S. 2001. Natural and anthropogenic changes in Chesapeake Bay during the last 1000 years. Human and Ecological Risk Assessment 7:12831296.CrossRefGoogle Scholar
Burdige, D. J., Hu, X., and Zimmerman, R. C.. 2010. The widespread occurrence of coupled carbonate dissolution/reprecipitation in surface sediments on the Bahamas Bank. American Journal of Science 310:492521.CrossRefGoogle Scholar
Bush, A. M., Powell, M. G., Arnold, W. S., Bert, T. M., and Daley, G. M.. 2002. Time-averaging, evolution, and morphologic variation. Paleobiology 28:9-25.10.1666/0094-8373(2002)028<0009:TAEAMV>2.0.CO;22.0.CO;2>CrossRefGoogle Scholar
Chen, T., Li, S., Zhao, J., and Feng, Y.. 2021. Uranium-thorium dating of coral mortality and community shift in a highly disturbed inshore reef (Weizhou Island, northern South China Sea). Science of the Total Environment 752:141866.10.1016/j.scitotenv.2020.141866CrossRefGoogle Scholar
Cherns, L., Wheeley, J. R., and Wright, V. P.. 2008. Taphonomic windows and molluscan preservation. Palaeogeography, Palaeoclimatology, Palaeoecology 270:220229.10.1016/j.palaeo.2008.07.012CrossRefGoogle Scholar
Clark, T. R., Roff, G., Zhao, J. X., Feng, Y. X., Done, T. J., McCook, L. J., and Pandolfi, J. M.. 2017. U-Th dating reveals regional-scale decline of branching Acropora corals on the Great Barrier Reef over the past century. Proceedings of the National Academy of Sciences USA 114:1035010355.CrossRefGoogle ScholarPubMed
Cramer, K. L., Jackson, J. B., Angioletti, C. V., Leonard-Pingel, J., and Guilderson, T. P.. 2012. Anthropogenic mortality on coral reefs in Caribbean Panama predates coral disease and bleaching. Ecology Letters 15:561567.10.1111/j.1461-0248.2012.01768.xCrossRefGoogle ScholarPubMed
Davies, D. J., Powell, E. N., and Stanton, R. J. Jr. 1989. Relative rates of shell dissolution and net sediment accumulation—a commentary: can shell beds form by the gradual accumulation of biogenic debris on the sea floor? Lethaia 22:207212.10.1111/j.1502-3931.1989.tb01683.xCrossRefGoogle Scholar
Dillon, E. M., McCauley, D. J., Morales-Saldaña, J. M., Leonard, N. D., Zhao, J. X., and O'Dea, A.. 2021. Fossil dermal denticles reveal the preexploitation baseline of a Caribbean coral reef shark community. Proceedings of the National Academy of Sciences USA 118:2017735118.CrossRefGoogle ScholarPubMed
Dominguez, J. G., Kosnik, M. A., Allen, A. P., Hua, Q., Jacob, D. E., Kaufman, D. S., and Whitacre, K.. 2016. Time-averaging and stratigraphic resolution in death assemblages and Holocene deposits: Sydney Harbour's molluscan record. Palaios 31:564575.CrossRefGoogle Scholar
Eganhouse, R. P., and Pontolillo, J.. 2000. Depositional history of organic contaminants on the Palos Verdes Shelf, California. Marine Chemistry 70:317338.10.1016/S0304-4203(00)00033-5CrossRefGoogle Scholar
Flessa, K. W., and Kowalewski, M.. 1994. Shell survival and time-averaging in nearshore and shelf environments: estimates from the radiocarbon literature. Lethaia 27:153165.10.1111/j.1502-3931.1994.tb01570.xCrossRefGoogle Scholar
Flessa, K. W., Cutler, A. H., and Meldahl, K. H.. 1993. Time and taphonomy: quantitative estimates of time-averaging and stratigraphic disorder in a shallow marine habitat. Paleobiology 19:266286.10.1017/S0094837300015918CrossRefGoogle Scholar
Fürsich, F. T., 1978. The influence of faunal condensation and mixing on the preservation of fossil benthic communities. Lethaia 11:243250.10.1111/j.1502-3931.1978.tb01231.xCrossRefGoogle Scholar
Ge, Y., Pederson, C. L., Lokier, S. W., Traas, J. P., Nehrke, G., Neuser, R. D., Goetschl, K. E., and Immenhauser, A.. 2020. Late Holocene to Recent aragonite-cemented transgressive lag deposits in the Abu Dhabi lagoon and intertidal sabkha. Sedimentology 67:24262454.10.1111/sed.12707CrossRefGoogle Scholar
Glover, C. P., and Kidwell, S. M.. 1993. Influence of organic matrix on the post-mortem destruction of molluscan shells. Journal of Geology 101:729747.10.1086/648271CrossRefGoogle Scholar
Griffis, R. B., and Suchanek, T. H.. 1991. A model of burrow architecture and trophic modes in thalassinidean shrimp (Decapoda: Thalassinidea). Marine Ecology Progress Series 79:171183.CrossRefGoogle Scholar
Guinasso, N. L. Jr., and Schink, D. R.. 1975. Quantitative estimates of biological mixing rates in abyssal sediments. Journal of Geophysical Research 80:30323043.10.1029/JC080i021p03032CrossRefGoogle Scholar
Guttorp, P. 1995. Stochastic modeling of scientific data. Chapman and Hall, London.10.1007/978-1-4899-4449-8CrossRefGoogle Scholar
Hannisdal, B. 2006. Phenotypic evolution in the fossil record: numerical experiments. Journal of Geology 114:133153.CrossRefGoogle Scholar
Hohmann, N. 2021. Incorporating information on varying sedimentation rates into paleontological analyses. Palaios 36:5367.CrossRefGoogle Scholar
Holland, S. M. 2000. The quality of the fossil record: a sequence stratigraphic perspective. Paleobiology 26(S4):148168.10.1666/0094-8373(2000)26[148:TQOTFR]2.0.CO;2CrossRefGoogle Scholar
Holland, S. M., and Patzkowsky, M. E.. 2015. The stratigraphy of mass extinction. Palaeontology 58:903924.10.1111/pala.12188CrossRefGoogle Scholar
Hover, V. C., Walter, L. M., and Peacor, D. R.. 2001. Early marine diagenesis of biogenic aragonite and Mg-calcite: new constraints from high-resolution STEM and AEM analyses of modern platform carbonates. Chemical Geology 175:221248.10.1016/S0009-2541(00)00326-0CrossRefGoogle Scholar
Hull, P. M., Franks, P. J., and Norris, R. D.. 2011. Mechanisms and models of iridium anomaly shape across the Cretaceous–Paleogene boundary. Earth and Planetary Science Letters 301:98106.10.1016/j.epsl.2010.10.031CrossRefGoogle Scholar
Hülse, D., Vervoort, P., van de Velde, S. J., Kanzaki, Y., Boudreau, B., Arndt, S., Bottjer, D. J., Hoogakker, B., Kuderer, M., Middelburg, J. J., and Volkenborn, N.. 2022. Assessing the impact of bioturbation on sedimentary isotopic records through numerical models. Earth-Science Reviews 234:104213.CrossRefGoogle Scholar
Hupp, B. N., Kelly, D. C., Zachos, J. C., and Bralower, T. J.. 2019. Effects of size-dependent sediment mixing on deep-sea records of the Paleocene–Eocene Thermal Maximum. Geology 47:749752.10.1130/G46042.1CrossRefGoogle Scholar
Hyman, A. C., Frazer, T. K., Jacoby, C. A., Frost, J. R., and Kowalewski, M.. 2019. Long-term persistence of structured habitats: seagrass meadows as enduring hotspots of biodiversity and faunal stability. Proceedings of the Royal Society of London B 286:20191861.Google Scholar
Jarochowska, E., Nohl, T., Grohganz, M., Hohmann, N., Vandenbroucke, T. R. A., and Munnecke, A.. 2020. Reconstructing depositional rates and their effect on paleoenvironmental proxies: the case of the Lau Carbon Isotope Excursion in Gotland, Sweden. Paleoceanography and Paleoclimatology 35:2020PA003979.CrossRefGoogle Scholar
Jilbert, T., and Slomp, C. P.. 2013. Rapid high-amplitude variability in Baltic Sea hypoxia during the Holocene. Geology 41:11831186.CrossRefGoogle Scholar
Jørgensen, B. B. 1977. Bacterial sulfate reduction within reduced microniches of oxidized marine sediments. Marine Biology 41:717.10.1007/BF00390576CrossRefGoogle Scholar
Kabel, K., Moros, M., Porsche, C., Neumann, T., Adolphi, F., Andersen, T. J., Siegel, H., Gerth, M., Leipe, T., Jansen, E., and Damste, J. S. S.. 2012. Impact of climate change on the Baltic Sea ecosystem over the past 1,000 years. Nature Climate Change 2:871874.CrossRefGoogle Scholar
Katz, T., Katsman, R., Kalman, A., and Goodman-Tchernov, B.. 2022. Evolution of sediment grain-size profiles on a sheltered, continental shelf in response to punctuated flood deposition. Continental Shelf Research 251:104868.10.1016/j.csr.2022.104868CrossRefGoogle Scholar
Keil, R. 2017. Anthropogenic forcing of carbonate and organic carbon preservation in marine sediments. Annual Review of Marine Science 9:151172.CrossRefGoogle ScholarPubMed
Kemnitz, N., Berelson, W., Hammond, D., Morine, L., Figueroa, M., Lyons, T. W., Scharf, S., Rollins, N., Petsios, E., Lemieux, S., and Treude, T.. 2020. Evidence of changes in sedimentation rate and sediment fabric in a low oxygen setting: Santa Monica Basin, CA. Biogeosciences 17:23812396.10.5194/bg-17-2381-2020CrossRefGoogle Scholar
Kemp, D. B., Fraser, W. T., and Izumi, K.. 2018. Stratigraphic completeness and resolution in an ancient mudrock succession. Sedimentology 65:18751890.CrossRefGoogle Scholar
Kidwell, S. M. 1986. Models for fossil concentrations: paleobiologic implications. Paleobiology 12:624.CrossRefGoogle Scholar
Kidwell, S. M. 1989. Stratigraphic condensation of marine transgressive records: origin of major shell deposits in the Miocene of Maryland. Journal of Geology 97:124.CrossRefGoogle Scholar
Kidwell, S. M. 1991. The stratigraphy of shell concentrations. Pp. 211290 in Allison, P. A. and Briggs, D. E. G., eds. Taphonomy: releasing the data locked in the fossil record. Plenum Press, New York.CrossRefGoogle Scholar
Kidwell, S. M., 1993. Influence of subsidence on the anatomy of marine siliciclastic sequences and on the distribution of shell and bone beds. Journal of the Geological Society 150:165167.CrossRefGoogle Scholar
Kidwell, S. M. 1997. Time-averaging in the marine fossil record: overview of strategies and uncertainties. Geobios 30:977995.10.1016/S0016-6995(97)80219-7CrossRefGoogle Scholar
Kidwell, S. M., and Bosence, D. W.. 1991. Taphonomy and time-averaging of marine shelly faunas. Pp. 115209 in Allison, P. A., and Briggs, D. E. G. eds. 1991. Taphonomy: releasing the data locked in the fossil record. Plenum Press, New York.CrossRefGoogle Scholar
Kidwell, S. M., and Brenchley, P. J.. 1994. Patterns in bioclastic accumulation through the Phanerozoic: changes in input or in destruction? Geology 22:11391143.2.3.CO;2>CrossRefGoogle Scholar
Kidwell, S. M., and Tomašových, A.. 2013. Implications of time-averaged death assemblages for ecology and conservation biology. Annual Review of Ecology, Evolution, and Systematics 44:539563.10.1146/annurev-ecolsys-110512-135838CrossRefGoogle Scholar
Kidwell, S. M., Meadows, C. A., and Edelman-Furstenberg, Y.. 2018. SEM evidence of the early diagenesis associated with aragonitic shells that survive 100s-1000s years of time-averaged residence in bioturbated seabeds. Geological Society of America Abstracts with Programs 50, doi: 10.1130/abs/2018AM-320131.CrossRefGoogle Scholar
Kirtland Turner, S., and Ridgwell, A.. 2013. Recovering the true size of an Eocene hyperthermal from the marine sedimentary record. Paleoceanography 28:700712.CrossRefGoogle Scholar
Kirtland Turner, S., Hull, P. M., Kump, L. R., and Ridgwell, A.. 2017. A probabilistic assessment of the rapidity of PETM onset. Nature Communications 8:110.10.1038/s41467-017-00292-2CrossRefGoogle ScholarPubMed
Kokesh, B. S., and Stemann, T. A.. 2023. Dead men still tell tales: bivalve death assemblages record dynamics and consequences of recent biological invasions in Kingston Harbour, Jamaica. In Nawrot R, R.., Dominici, S., Tomašových, A., and Zuschin, M., eds. Conservation palaeobiology of marine ecosystems. Geological Society of London Special Publication 529, doi: 10.1144/SP529-2022-28.Google Scholar
Kosnik, M. A., Hua, Q., Jacobsen, G. E., Kaufman, D. S., and Wüst, R. A.. 2007. Sediment mixing and stratigraphic disorder revealed by the age-structure of Tellina shells in Great Barrier Reef sediment. Geology 35:811814.CrossRefGoogle Scholar
Kosnik, M. A., Hua, Q., Kaufman, D. S., and Wüst, R. A.. 2009. Taphonomic bias and time-averaging in tropical molluscan death assemblages: differential shell half-lives in Great Barrier Reef sediment. Paleobiology 35:565586.10.1666/0094-8373-35.4.565CrossRefGoogle Scholar
Kosnik, M. A., Kaufman, D. S., and Hua, Q.. 2013. Radiocarbon-calibrated multiple amino acid geochronology of Holocene molluscs from Bramble and Rib Reefs (Great Barrier Reef, Australia). Quaternary Geochronology 16:7386.10.1016/j.quageo.2012.04.024CrossRefGoogle Scholar
Kosnik, M. A., Hua, Q., Kaufman, D. S., and Zawadzki, A.. 2015. Sediment accumulation, stratigraphic order, and the extent of time-averaging in lagoonal sediments: a comparison of 210Pb and 14C/amino acid racemization chronologies. Coral Reefs 34:215229.CrossRefGoogle Scholar
Kowalewski, M. 1996. Time-averaging, overcompleteness, and the geological record. Journal of Geology 104:317326.CrossRefGoogle Scholar
Kowalewski, M., Goodfriend, G. A., and Flessa, K. W.. 1998. High-resolution estimates of temporal mixing within shell beds: the evils and virtues of time-averaging. Paleobiology 24:287304.Google Scholar
Kowalewski, M., Casebolt, S., Hua, Q., Whitacre, K. E., Kaufman, D. S., and Kosnik, M. A.. 2018. One fossil record, multiple time resolutions: disparate time-averaging of echinoids and mollusks on a Holocene carbonate platform. Geology 46:5154.CrossRefGoogle Scholar
Kristensen, E. 2000. Organic matter diagenesis at the oxic/anoxic interface in coastal marine sediments, with emphasis on the role of burrowing animals. Pp. 124 in Liebezeit, G., Dittmann, S., and Kröncke, I., eds. Life at interfaces and under extreme conditions. Springer, Dordrecht, Netherlands.Google Scholar
Ku, T. C. W., Walter, L. M., Coleman, M. L., Blake, R. E., and Martini, A. M.. 1999. Coupling between sulfur recycling and syndepositional carbonate dissolution: evidence from oxygen and sulfur isotope composition of pore water sulfate, South Florida Platform, USA. Geochimica et Cosmochimica Acta 63:25292546.CrossRefGoogle Scholar
Kunz, T., Dolman, A. M., and Laepple, T.. 2020. A spectral approach to estimating the timescale-dependent uncertainty of paleoclimate records—Part 1: Theoretical concept. Climate of the Past 16:14691492.10.5194/cp-16-1469-2020CrossRefGoogle Scholar
Kusnerik, K. M., Means, G. H., Portell, R. W., Brenner, M., Hua, Q., Kannai, A., Means, R., Monroe, M. A., and Kowalewski, M.. 2020. Live, dead, and fossil mollusks in Florida freshwater springs and spring-fed rivers: taphonomic pathways and the formation of multisourced, time-averaged death assemblages. Paleobiology 46:356378.CrossRefGoogle Scholar
Lee, H. J., Sherwood, C. R., Drake, D. E., Edwards, B. D., Wong, F., and Hamer, M.. 2002. Spatial and temporal distribution of contaminated, effluent-affected sediment on the Palos Verdes margin, southern California. Continental Shelf Research 22:859880.10.1016/S0278-4343(01)00108-XCrossRefGoogle Scholar
Lehto, N., Glud, R. N., á Norði, G., Zhang, H., and Davison, W.. 2014. Anoxic microniches in marine sediments induced by aggregate settlement: biogeochemical dynamics and implications. Biogeochemistry 119:307327.Google Scholar
Leonard-Pingel, J. S., Kidwell, S. M., Tomašových, A., Alexander, C. R., and Cadien, D. B.. 2019. Gauging benthic recovery from 20th century pollution on the southern California continental shelf using bivalves from sediment cores. Marine Ecology Progress Series 615:101119.CrossRefGoogle Scholar
Leshno, Y., Edelman-Furstenberg, Y., Mienis, H., and Benjamini, C.. 2015. Molluscan live and dead assemblages in an anthropogenically stressed shallow-shelf: Levantine margin of Israel. Palaeogeography, Palaeoclimatology, Palaeoecology 433:4959.10.1016/j.palaeo.2015.05.008CrossRefGoogle Scholar
Liu, H., Meyers, S. R., and Marcott, S. A.. 2021. Unmixing deep-sea paleoclimate records: a study on bioturbation effects through convolution and deconvolution. Earth and Planetary Science Letters 564:116883.10.1016/j.epsl.2021.116883CrossRefGoogle Scholar
Lougheed, B. C., and Metcalfe, B.. 2022. Testing the effect of bioturbation and species abundance upon discrete-depth individual foraminifera analysis. Biogeosciences 19:11951209.10.5194/bg-19-1195-2022CrossRefGoogle Scholar
Martin, R. E., Hippensteel, S. P., Nikitina, D., and Pizzuto, J. E.. 2002. Artificial time-averaging of marsh foraminiferal assemblages: linking the temporal scales of ecology and paleoecology. Paleobiology 28:263277.2.0.CO;2>CrossRefGoogle Scholar
Martinelli, J. C., Madin, J. S., and Kosnik, M. A.. 2016. Dead shell assemblages faithfully record living molluscan assemblages at One Tree Reef. Palaeogeography, Palaeoclimatology, Palaeoecology 457:158169.CrossRefGoogle Scholar
McCave, I. N. 1988. Biological pumping upwards of the coarse fraction of deep-sea sediments. Journal of Sedimentary Research 58:148158.10.1306/212F8D3C-2B24-11D7-8648000102C1865DCrossRefGoogle Scholar
McGann, M. 2009. Review of impacts of contaminated sediment on microfaunal communities in the Southern California Bight. Geological Society of America Special Papers 454:413455.Google Scholar
Mekik, F. 2014. Radiocarbon dating of planktonic foraminifer shells: a cautionary tale. Paleoceanography and Paleoclimatology 29:1329.10.1002/2013PA002532CrossRefGoogle Scholar
Meldahl, K. H. 1987. Sedimentologic and taphonomic implications of biogenic stratification. Palaios 2:350358.CrossRefGoogle Scholar
Meysman, F. J., Boudreau, B. P., and Middelburg, J. J.. 2003. Relations between local, nonlocal, discrete and continuous models of bioturbation. Journal of Marine Research 61:391410.CrossRefGoogle Scholar
Meysman, F. J., Malyuga, V. S., Boudreau, B. P., and Middelburg, J. J.. 2008. A generalized stochastic approach to particle dispersal in soils and sediments. Geochimica et Cosmochimica Acta, 72:34603478.10.1016/j.gca.2008.04.023CrossRefGoogle Scholar
F. J. R., Meysman, Boudreau, B. P., and Middelburg, J. J.. 2010. When and why does bioturbation lead to diffusive mixing. Journal of Marine Research 68:881920.Google Scholar
Miller, J. H., Behrensmeyer, A. K., Du, A., Lyons, S. K., Patterson, D., Tóth, A., Villaseñor, A., Kanga, E., and Reed, D.. 2014. Ecological fidelity of functional traits based on species presence-absence in a modern mammalian bone assemblage (Amboseli, Kenya). Paleobiology 40:560583.10.1666/13062CrossRefGoogle Scholar
Moros, M., Andersen, T. J., Schulz-Bull, D., Häusler, K., Bunke, D., Snowball, I., Kotilainen, A., Zillén, L., Jensen, J. B., Kabel, K., and Hand, I.. 2017. Towards an event stratigraphy for Baltic Sea sediments deposited since AD 1900: approaches and challenges. Boreas 46:129142.10.1111/bor.12193CrossRefGoogle Scholar
Nawrot, R., Scarponi, D., Azzarone, M., Dexter, T. A., Kusnerik, K. M., Wittmer, J. M., Amorosi, A., and Kowalewski, M.. 2018. Stratigraphic signatures of mass extinctions: ecological and sedimentary determinants. Proceedings of the Royal Society of London B 285:20181191.Google ScholarPubMed
Nawrot, R., Berensmeier, M., Gallmetzer, I., Haselmair, A., Tomašových, A., and Zuschin, M.. 2022. Multiple phyla, one time resolution? Similar time averaging in benthic foraminifera, mollusk, echinoid, crustacean, and otolith fossil assemblages. Geology 50:902906.10.1130/G49970.1CrossRefGoogle Scholar
Niedoroda, A. W., Swift, D. J. P., Reed, C. W., and Stull, J. K.. 1996. Contaminant dispersal on the Palos Verdes continental margin: III. Processes controlling transport, accumulation and re-emergence of DDT-contaminated sediment particles. Science of the Total Environment 179:109133.CrossRefGoogle Scholar
Olszewski, T. D. 2004. Modeling the influence of taphonomic destruction, reworking, and burial on time-averaging in fossil accumulations. Palaios 19:3950.2.0.CO;2>CrossRefGoogle Scholar
Olszewski, T. D., and Kaufman, D. S.. 2015. Tracing burial history and sediment recycling in a shallow estuarine setting (Copano Bay, Texas) using postmortem ages of the bivalve Mulinia lateralis. Palaios 30:224237.10.2110/palo.2014.063CrossRefGoogle Scholar
Olszewski, T. D., and Kidwell, S. M.. 2007. The preservational fidelity of evenness in molluscan death assemblages. Paleobiology 33:123.10.1666/05059.1CrossRefGoogle Scholar
Ozalas, K., Savrda, C. E., and Fullerton, R. R. Jr. 1994. Bioturbated oxygenation-event beds in siliceous facies: Monterey Formation (Miocene), California. Palaeogeography, Palaeoclimatology, Palaeoecology 112:6383.CrossRefGoogle Scholar
Parker, W. G., Yanes, Y., Mesa Hernández, E., Hernández Marrero, J. C., Pais, J., and Surge, D.. 2020. Scale of time-averaging in archaeological shell middens from the Canary Islands. Holocene 30:258271.CrossRefGoogle Scholar
Parsons-Hubbard, K. M., Callender, W. R., Powell, E. N., Brett, C. E., Walker, S. E., Raymond, A. L., and Staff, G. M.. 1999. Rates of burial and disturbance of experimentally-deployed molluscs; implications for preservation potential. Palaios 14:337351.CrossRefGoogle Scholar
Parsons-Hubbard, K., Hubbard, D., Tems, C., and Burkett, A.. 2014. The relationship between modern mollusk assemblages and their expression in subsurface sediment in a carbonate lagoon, St. Croix, US Virgin Islands. Pp. 143167 in Hembree, D. I., Platt, B. F., and Smith, J. J., eds. Experimental approaches to understanding fossil organisms. Springer, Dordrecht, Netherlands.CrossRefGoogle Scholar
Peng, T. H., and Broecker, W. S.. 1984. The impacts of bioturbation on the age difference between benthic and planktonic foraminifera in deep sea sediments. Nuclear Instruments and Methods in Physics Research Section B 233:346352.10.1016/0168-583X(84)90540-8CrossRefGoogle Scholar
Poirier, C., Chaumillon, E., and Arnaud, F.. 2011. Siltation of river-influenced coastal environments: respective impact of late Holocene land use and high-frequency climate changes. Marine Geology 290:5162.10.1016/j.margeo.2011.10.008CrossRefGoogle Scholar
Powell, E. N., Kraeuter, J. N., and Ashton-Alcox, K. A.. 2006. How long does oyster shell last on an oyster reef? Estuarine, Coastal and Shelf Science 69:531542.10.1016/j.ecss.2006.05.014CrossRefGoogle Scholar
Raiswell, R. 1988. Chemical model for the origin of minor limestone-shale cycles by anaerobic methane oxidation. Geology 16:641644.10.1130/0091-7613(1988)016<0641:CMFTOO>2.3.CO;22.3.CO;2>CrossRefGoogle Scholar
Raiswell, R., and Fisher, Q. J.. 2004. Rates of carbonate cementation associated with sulphate reduction in DSDP/ODP sediments: implications for the formation of concretions. Chemical Geology 211:7185.10.1016/j.chemgeo.2004.06.020CrossRefGoogle Scholar
Reid, R. P., and Macintyre, I. G.. 1998. Carbonate recrystallization in shallow marine environments: a widespread diagenetic process forming micritized grains. Journal of Sedimentary Research 68:928946.10.2110/jsr.68.928CrossRefGoogle Scholar
Rhoads, D. C. 1967. Biogenic reworking of intertidal and subtidal sediments in Barnstable Harbor and Buzzards Bay, Massachusetts. Journal of Geology 75:461476.CrossRefGoogle Scholar
Rhoads, D. C., and Stanley, D. J.. 1965. Biogenic graded bedding. Journal of Sedimentary Research 35:956963.Google Scholar
Ridgwell, A. 2007. Interpreting transient carbonate compensation depth changes by marine sediment core modeling. Paleoceanography 22:PA001372.CrossRefGoogle Scholar
Ritter, M. D. N., Erthal, F., Kosnik, M. A., Coimbra, J. C., and Kaufman, D. S.. 2017. Spatial variation in the temporal resolution of subtropical shallow-water molluscan death assemblages. Palaios 32:572583.10.2110/palo.2017.003CrossRefGoogle Scholar
Ryan, E. K., Soreghan, M. J., McGlue, M. M., Todd, J. A., Michel, E., Kaufman, D. S., and Kimirei, I.. 2020. Paleoenvironmental implications of time-averaging and taphonomic variation of shell beds in Lake Tanganyika. Palaios 35:4966.CrossRefGoogle Scholar
Sadler, P. M. 1993. Models of time-averaging as a maturation process: how soon do sedimentary sections escape reworking? Short Courses in Paleontology 6:188209.CrossRefGoogle Scholar
Santschi, P. H., Guo, L., Asbill, S., Allison, M., Kepple, A. B., and Wen, L. S.. 2001. Accumulation rates and sources of sediments and organic carbon on the Palos Verdes shelf based on radioisotopic tracers (137Cs, 239,240Pu, 210Pb, 234Th, 238U and 14C). Marine Chemistry 73:125152.CrossRefGoogle Scholar
Savranskaia, T., Egli, R., and Valet, J. P.. 2022. Multiscale Brazil nut effects in bioturbated sediment. Scientific Reports 12:19.CrossRefGoogle ScholarPubMed
Savrda, C. E., and Bottjer, D. J.. 1989. Anatomy and implications of bioturbated beds in “black shale” sequences: examples from the Jurassic Posidonienschiefer (southern Germany). Palaios 4:330342.10.2307/3514557CrossRefGoogle Scholar
Savrda, C. E., and Ozalas, K.. 1993. Preservation of mixed-layer ichnofabrics in oxygenation-event beds. Palaios 8:609613.CrossRefGoogle Scholar
Scarponi, D., Kaufman, D., Amorosi, A., and Kowalewski, M.. 2013. Sequence stratigraphy and the resolution of the fossil record. Geology 41:239242.CrossRefGoogle Scholar
Scarponi, D., Nawrot, R., Azzarone, M., Pellegrini, C., Gamberi, F., Trincardi, F., and Kowalewski, M.. 2022. Resilient biotic response to long-term climate change in the Adriatic Sea. Global Change Biology 28:40414053.10.1111/gcb.16168CrossRefGoogle ScholarPubMed
Schiffelbein, P. 1986. The interpretation of stable isotopes in deep-sea sediments: an error analysis case study. Marine Geology 70:313320.CrossRefGoogle Scholar
Schink, D. R., and Guinasso, N. L. Jr. 1977. Effects of bioturbation on sediment—seawater interaction. Marine Geology 23:133154.CrossRefGoogle Scholar
Schwarzacher, W. 1972. The semi-Markov process as a general sedimentation model. Pp. 247268 in Merriam, D. F., ed. Mathematical models of sedimentary processes. Springer, Boston.CrossRefGoogle Scholar
Seilacher, A. 1982. General remarks about event deposit. Pp. 161173 in Einsele, G. and Seilacher, A., eds. Cyclic and event stratification. Springer, Berlin.10.1007/978-3-642-75829-4_11CrossRefGoogle Scholar
Seilacher, A. 1985. The Jeram model: event condensation in a modern intertidal environment. Pp. 335341 in Bayer, U. and Seilacher, A., ed. Sedimentary and evolutionary cycles. Springer, Berlin.CrossRefGoogle Scholar
Shull, D. H. 2001. Transition-matrix model of bioturbation and radionuclide diagenesis. Limnology and Oceanography 46:905916.CrossRefGoogle Scholar
Shull, D. H., and Yasuda, M.. 2001. Size-selective downward particle transport by cirratulid polychaetes. Journal of Marine Research 59:453473.10.1357/002224001762842271CrossRefGoogle Scholar
Smith, J. A., Auerbach, D. A., Flessa, K. W., Flecker, A. S., and Dietl, G. P.. 2016. Fossil clam shells reveal unintended carbon cycling consequences of Colorado River management. Royal Society Open Science 3:160170.CrossRefGoogle ScholarPubMed
Soetaert, K. 2012. diagram: functions for visualising simple graphs (networks), plotting flow diagrams, R package version 1.6.5. https://CRAN.R-project.org/package=diagram, accessed August 2022.Google Scholar
Soetaert, K., Herman, P. M., Middelburg, J. J., Heip, C., de Stigter, H. S., van Weering, T. C., Epping, E., and Helder, W.. 1996. Modeling 210Pb-derived mixing activity in ocean margin sediments: diffusive versus nonlocal mixing. Journal of Marine Research 54:12071227.CrossRefGoogle Scholar
Soetaert, K., Hofmann, A. F., Middelburg, J. J., Meysman, F. J. R., and Greenwood, J.. 2007. The effect of biogeochemical processes on pH. Marine Chemistry 105:3051.CrossRefGoogle Scholar
Solan, M., Ward, E. R., White, E. L., Hibberd, E. E., Cassidy, C., Schuster, J. M., Hale, R., and Godbold, J. A.. 2019. Worldwide measurements of bioturbation intensity, ventilation rate, and the mixing depth of marine sediments. Scientific Data 6:16.CrossRefGoogle ScholarPubMed
Sorrel, P., Tessier, B., Demory, F., Baltzer, A., Bouaouina, F., Proust, J. N., Menier, D., and Traini, C.. 2010. Sedimentary archives of the French Atlantic coast (inner Bay of Vilaine, south Brittany): depositional history and late Holocene climatic and environmental signals. Continental Shelf Research 30:12501266.CrossRefGoogle Scholar
Stegner, M. A., Ratajczak, Z., Carpenter, S. R., and Williams, J. W.. 2019. Inferring critical transitions in paleoecological time series with irregular sampling and variable time-averaging. Quaternary Science Reviews 207:4963.CrossRefGoogle Scholar
Steiner, Z., Lazar, B., Levi, S., Tsroya, S., Pelled, O., Bookman, R., and Erez, J.. 2016. The effect of bioturbation in pelagic sediments: lessons from radioactive tracers and planktonic foraminifera in the Gulf of Aqaba, Red Sea. Geochimica et Cosmochimica Acta 194:139152.CrossRefGoogle Scholar
Straub, K. M., and Foreman, B. Z.. 2018. Geomorphic stasis and spatiotemporal scales of stratigraphic completeness. Geology 46:311314.CrossRefGoogle Scholar
Stull, J. K., Haydock, C. I., Smith, R. W., and Montagne, D. E.. 1986. Long-term changes in the benthic community on the coastal shelf of Palos Verdes, Southern California. Marine Biology 91:539–511.CrossRefGoogle Scholar
Stull, J. K., Swift, D. J., and Niedoroda, A. W.. 1996. Contaminant dispersal on the Palos Verdes continental margin: I. Sediments and biota near a major California wastewater discharge. Science of the Total Environment 179:7390.CrossRefGoogle Scholar
Suchanek, T. H., Colin, P. L., McMurtry, G. M., and Suchanek, C. S.. 1986. Bioturbation and redistribution of sediment radionuclides in Enewetak Atoll lagoon by callianassid shrimp: biological aspects. Bulletin of Marine Science 38:144154.Google Scholar
Swift, D. J. P., Stull, J. K., Niedoroda, A. W., Reed, C. W., and Wong, G. T. F.. 1996. Contaminant dispersal on the Palos Verdes continental margin: II. Estimates of biodiffusion coefficient, Db, from composition of the benthic infaunal community. Science of the Total Environment 179:91107.CrossRefGoogle Scholar
Syvitski, J. P., and Kettner, A.. 2011. Sediment flux and the Anthropocene. Philosophical Transactions of the Royal Society of London A 369:957975.Google ScholarPubMed
Tedesco, L. P., and Aller, R. C.. 1997. 210Pb chronology of sequences affected by burrow excavation and infilling; examples from shallow marine carbonate sediment sequences, Holocene South Florida and Caicos Platform, British West Indies. Journal of Sedimentary Research 67:3646.Google Scholar
Tedesco, L. P., and Wanless, H. R.. 1991. Generation of sedimentary fabrics and facies by repetitive excavation and storm infilling of burrow networks, Holocene of South Florida and Caicos Platform, BWI. Palaios 6:326343.CrossRefGoogle Scholar
Terry, R. C. 2010. The dead do not lie: using skeletal remains for rapid assessment of historical small-mammal community baselines. Proceedings of the Royal Society of London B 277:11931201.Google Scholar
Terry, R. C., and Novak, M.. 2015. Where does the time go? Mixing and the depth-dependent distribution of fossil ages. Geology 43:487490.CrossRefGoogle Scholar
Tomašových, A., and Kidwell, S. M.. 2009. Fidelity of variation in species composition and diversity partitioning by death assemblages: time-averaging transfers diversity from beta to alpha levels. Paleobiology 35:94118.CrossRefGoogle Scholar
Tomašových, A., and Kidwell, S. M.. 2010. The effects of temporal resolution on species turnover and on testing metacommunity models. American Naturalist 175:587606.10.1086/651661CrossRefGoogle ScholarPubMed
Tomašových, A., and Kidwell, S. M.. 2017. Nineteenth-century collapse of a benthic marine ecosystem on the open continental shelf. Proceedings of the Royal Society of London B 284:20170328.Google ScholarPubMed
Tomašových, A., Fürsich, F. T., and Olszewski, T. D.. 2006. Modeling shelliness and alteration in shell beds: variation in hardpart input and burial rates leads to opposing predictions. Paleobiology 32:278298.CrossRefGoogle Scholar
Tomašových, A., Kidwell, S. M., Barber, R. F., and Kaufman, D. S.. 2014. Long-term accumulation of carbonate shells reflects a 100-fold drop in loss rate. Geology 42:819822.CrossRefGoogle Scholar
Tomašových, A., Kidwell, S. M., and Barber, R. F.. 2016. Inferring skeletal production from time-averaged assemblages: skeletal loss pulls the timing of production pulses towards the modern period. Paleobiology 42:5476.CrossRefGoogle Scholar
Tomašových, A., Gallmetzer, I., Haselmair, A., Kaufman, D. S., Vidović, J., and Zuschin, M.. 2017. Stratigraphic unmixing reveals repeated hypoxia events over the past 500 yr in the northern Adriatic Sea. Geology 45:363366.CrossRefGoogle Scholar
Tomašových, A., Gallmetzer, I., Haselmair, A., Kaufman, D. S., Kralj, M., Cassin, D., Zonta, R., and Zuschin, M.. 2018. Tracing the effects of eutrophication on molluscan communities in sediment cores: outbreaks of an opportunistic species coincide with reduced bioturbation and high frequency of hypoxia in the Adriatic Sea. Paleobiology 44:575602.CrossRefGoogle Scholar
Tomašových, A., Gallmetzer, I., Haselmair, A., Kaufman, D. S., Mavrič, B., and Zuschin, M.. 2019a. A decline in molluscan carbonate production driven by the loss of vegetated habitats encoded in the Holocene sedimentary record of the Gulf of Trieste. Sedimentology 66:781807.CrossRefGoogle ScholarPubMed
Tomašových, A., Kidwell, S. M., Alexander, C. R., and Kaufman, D. S.. 2019b. Millennial-scale age offsets within fossil assemblages: result of bioturbation below the taphonomic active zone and out-of-phase production. Paleoceanography and Paleoclimatology 34:954977.CrossRefGoogle Scholar
Tomašových, A., Albano, P. G., Fuksi, T., Gallmetzer, I., Haselmair, A., Kowalewski, M., Nawrot, R., Nerlović, V., Scarponi, D., and Zuschin, M.. 2020. Ecological regime shift preserved in the Anthropocene stratigraphic record. Proceedings of the Royal Society of London B 287:20200695.Google ScholarPubMed
Tomašových, A., Gallmetzer, I., Haselmair, A., and Zuschin, M.. 2022. Inferring time averaging and hiatus durations in the stratigraphic record of high-frequency depositional sequences. Sedimentology 69:10831118.CrossRefGoogle Scholar
Tomašových, A., García-Ramos, D. A., Nawrot, R., Nebelsick, J. H., and Zuschin, M.. 2023. Millennial-scale changes in abundance of brachiopods in bathyal environments detected by postmortem age distributions in death assemblage (Bari Canyon, Adriatic Sea). In Nawrot R, R.., Dominici, S., Tomašových, A., and Zuschin, M., eds. Conservation palaeobiology of marine ecosystems. Geological Society of London Special Publication 529, 10.1144/SP529-2022-117.Google Scholar
Toth, L. T., Kuffner, I. B., Stathakopoulos, A., and Shinn, E. A.. 2018. A 3,000-year lag between the geological and ecological shutdown of Florida's coral reefs. Global Change Biology 24:54715483.CrossRefGoogle ScholarPubMed
Trauth, M. H. 1998. TURBO: a dynamic-probabilistic simulation to study the effects of bioturbation on paleoceanographic time series. Computers and Geosciences 24:433441.CrossRefGoogle Scholar
Tudhope, A. W., and Scoffin, T. P.. 1984. The effects of Callianassa bioturbation on the preservation of carbonate grains in Davies Reef Lagoon, Great Barrier Reef, Australia. Journal of Sedimentary Research 54:10911096.Google Scholar
Yanes, Y., Kowalewski, M., Ortiz, J. E., Castillo, C., de Torres, T., and de la Nuez, J.. 2007. Scale and structure of time-averaging (age mixing) in terrestrial gastropod assemblages from Quaternary eolian deposits of the eastern Canary Islands. Palaeogeography, Palaeoclimatology, Palaeoecology 251:283299.CrossRefGoogle Scholar
van de Velde, S., and Meysman, F. J.. 2016. The influence of bioturbation on iron and sulphur cycling in marine sediments: a model analysis. Aquatic Geochemistry 22:469504.CrossRefGoogle Scholar
Walter, L. M., and Burton, E. A.. 1990. Dissolution of recent platform carbonate sediments in marine pore fluids. American Journal of Science 290:601643.CrossRefGoogle Scholar
Wetzel, A. 2013. Formation of methane-related authigenic carbonates within the bioturbated zone—an example from the upwelling area off Vietnam. Palaeogeography, Palaeoclimatology, Palaeoecology 386:2333.CrossRefGoogle Scholar
Wheatcroft, R. A. 2006. Time-series measurements of macrobenthos abundance and sediment bioturbation intensity on a flood-dominated shelf. Progress in Oceanography 71:88122.CrossRefGoogle Scholar
Wheatcroft, R. A., and Jumars, P. A.. 1987. Statistical re-analysis for size dependency in deep-sea mixing. Marine Geology 77:157163.CrossRefGoogle Scholar
Wright, V. P., and Cherns, L.. 2016. Leaving no stone unturned: the feedback between increased biotic diversity and early diagenesis during the Ordovician. Journal of the Geological Society 173:241244.10.1144/jgs2015-043CrossRefGoogle Scholar
Zillén, L., Conley, D. J., Andrén, T., Andrén, E., and Björck, S.. 2008. Past occurrences of hypoxia in the Baltic Sea and the role of climate variability, environmental change and human impact. Earth-Science Reviews 91:7792.CrossRefGoogle Scholar
Zimmt, J. B., Holland, S. M., Finnegan, S., and Marshall, C. R.. 2021. Recognizing pulses of extinction from clusters of last occurrences. Palaeontology 64:120CrossRefGoogle Scholar
Zimmt, J. B., Kidwell, S. M., Lockwood, R., and Thirlwall, M.. 2022. Strontium isotope stratigraphy reveal 100 ky-scale condensation, beveling, and internal shingling of transgressive shell beds in the Maryland Miocene. Palaios 37:553573.CrossRefGoogle Scholar