Hostname: page-component-848d4c4894-8bljj Total loading time: 0 Render date: 2024-06-30T22:38:37.408Z Has data issue: false hasContentIssue false

Ion microprobe–measured stable isotope evidence for ammonite habitat and life mode during early ontogeny

Published online by Cambridge University Press:  03 September 2018

Benjamin J. Linzmeier
Affiliation:
Department of Earth and Planetary Sciences, Northwestern University, Evanston, Illinois 60208, USA; and Department of Geoscience, University of Wisconsin–Madison, Madison, Wisconsin 53703, USA. E-mail: benl@earth.northwestern.edu
Neil H. Landman
Affiliation:
Division of Paleontology, American Museum of Natural History, New York, New York 10024, USA
Shanan E. Peters
Affiliation:
Department of Geoscience, University of Wisconsin–Madison, Madison, Wisconsin 53703, USA
Reinhard Kozdon
Affiliation:
Lamont-Doherty Earth Observatory, Columbia University, Palisades, New York 10964, USA; and Department of Geoscience, and WiscSIMS Laboratory, University of Wisconsin–Madison, Madison, Wisconsin 53703, USA
Kouki Kitajima
Affiliation:
Department of Geoscience, and WiscSIMS Laboratory, University of Wisconsin–Madison, Madison, Wisconsin 53703, USA
John W. Valley
Affiliation:
Department of Geoscience, and WiscSIMS Laboratory, University of Wisconsin–Madison, Madison, Wisconsin 53703, USA

Abstract

Ammonites have disparate adult morphologies indicative of diverse ecological niches, but ammonite hatchlings are small (~1 mm diameter), which raises questions about the similarity of egg incubation and hatchling life mode in ammonites. Modern Nautilus is sometimes used as a model organism for understanding ammonites, but despite their outward similarities, the groups are only distantly related. Trends in ammonite diversity and extinction vulnerability in the fossil record contrast starkly with those of nautilids, and embryonic shells from Late Cretaceous ammonites are two orders of magnitude smaller than nautilid embryonic shells. To investigate possible environmental changes experienced by ammonite hatchlings, we used secondary ion mass spectrometry to analyze the oxygen and carbon isotope composition of the embryonic shells and early postembryonic whorls of five juveniles of Hoploscaphites comprimus obtained from a single concretion in the Fox Hills Formation of South Dakota. Co-occurring bivalves and diagenetic calcite were also analyzed to provide a benthic baseline for comparison. The oxygen isotope ratios of embryonic shells are more like those of benthic bivalves, suggesting that ammonite eggs were laid on the bottom. Ammonite shell immediately after hatching has more negative δ18O, suggesting movement to more shallow water that is potentially warmer and/or fresher. After approximately one whorl of postembryonic growth, the values of δ18O become more positive in three of the five individuals, suggesting that these animals transitioned to a more demersal mode of life. Two other individuals transition to even lower δ18O values that could suggest movement to nearshore brackish water. These data suggest that ammonites, like many modern coleoids, may have spawned at different times of the year. Because scaphites were one of the short-term Cretaceous–Paleogene extinction survivors, it is possible that this characteristic allowed them to develop a broader geographic range and, consequently, a greater resistance to extinction.

Type
Articles
Copyright
© 2018 The Paleontological Society. All rights reserved 

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

Aguiar, D. C. de, Rossi-Wongtschowski, C. L. D. B., and Perez, J. A. A.. 2012. Validation of daily growth increments of statoliths of Brazilian squid Doryteuthis plei and D. sanpaulensis (Cephalopoda. Loliginidae). Bioikos 26:1321.Google Scholar
Allison, P. A., and Pye, K.. 1994. Early diagenetic mineralization and fossil preservation in modern carbonate concretions. Palaios 9:561575.Google Scholar
Amiotte-Suchet, P., Aubert, D., Probst, J. L., Gauthier-Lafaye, F., Probst, A., Andreux, F., and Viville, D.. 1999. δ13C pattern of dissolved inorganic carbon in a small granitic catchment: the Strengbach case study (Vosges Mountains, France). Chemical Geology 159:129145.Google Scholar
Arkhipkin, A. I., and Laptikhovsky, V. V.. 2012. Impact of ocean acidification on plankton larvae as a cause of mass extinctions in ammonites and belemnites. Neues Jahrbuch für Geologie und Paläontologie–Abhandlungen 266:3950.Google Scholar
Aubert, M., Williams, I. S., Boljkovac, K., Moffat, I., Moncel, M.-H., Dufour, E., and Grün, R.. 2012. In situ oxygen isotope micro-analysis of faunal material and human teeth using a SHRIMP II: a new tool for palaeo-ecology and archaeology. Journal of Archaeological Science 39:31843194.Google Scholar
Bhattacharya, J. P., and MacEachern, J. A.. 2009. Hyperpycnal rivers and prodeltaic shelves in the Cretaceous seaway of North America. Journal of Sedimentary Research 79:184209.Google Scholar
Boletzky, S. v. 1983. Embyronic phase. Pp. 531. in P. R. Boyle, ed. Cephalopod life cycles Vol. 2. Comparative reviews. Academic Press, London.Google Scholar
Boletzky, S. von. 1978. Nos connaissances actuelles sur le développement des Octopodes. Vie Milieu 28:85120.Google Scholar
Bucher, H., Landman, N. H., Klofak, S. M., and Guex, J.. 1996. Mode and rate of growth in ammonoids. Pp. 407461. in N. H. Landman, K. Tanabe, and R. A. Davis, eds. Ammonoid paleobiology. Topics in Geobiology Vol. 13. Plenum, New York.Google Scholar
Cherns, L., and Wright, V. P.. 2000. Missing molluscs as evidence of large-scale, early skeletal aragonite dissolution in a Silurian sea. Geology 28:791794.Google Scholar
Clements, T., Colleary, C., De Baets, K., and Vinther, J.. 2017. Buoyancy mechanisms limit preservation of coleoid cephalopod soft tissues in Mesozoic Lagerstätten. Palaeontology 60:114.Google Scholar
Cobban, W. A., Merewether, E. A., Fouch, T. D., and Obradovich, J. D.. 1994. Some Cretaceous shorelines in the Western Interior of the United States. Pp. 393425. in M. V. Caputo, J. A. Peterson, and K. J. Franczyk, eds. Mesozoic systems of the Rocky Mountain region, USA. SEPM Rocky Mountain Section, Denver.Google Scholar
Cochran, J. K., Landman, N. H., Turekian, K. K., Michard, A., and Schrag, D. P.. 2003. Paleoceanography of the Late Cretaceous (Maastrichtian) Western Interior Seaway of North America: evidence from Sr and O isotopes. Palaeogeography, Palaeoclimatology, Palaeoecology 191:4564.Google Scholar
Cusack, M., England, J., Dalbeck, P., Tudhope, A. W., Fallick, A. E., and Allison, N.. 2008. Electron backscatter diffraction (EBSD) as a tool for detection of coral diagenesis. Coral Reefs 27:905911.Google Scholar
De Baets, K., Klug, C., and Korn, D.. 2011. Devonian pearls and ammonoid-endoparasite co-evolution. Acta Palaeontologica Polonica 56:159180.Google Scholar
De Baets, K., Klug, C., Korn, D., and Landman, N. H.. 2012. Early evolutionary trends in ammonoid embryonic development. Evolution 66:17881806.Google Scholar
De Baets, K., Landman, N. H., and Tanabe, K.. 2015. Ammonoid embryonic development. Pp. 113205. in C. Klug, D. Korn, K. De Baets, I. Kruta, and R. H. Mapes, eds. Ammonoid paleobiology: From anatomy to ecology. Springer, Dordrecht, Netherlands.Google Scholar
Dennis, K. J., Cochran, J. K., Landman, N. H., and Schrag, D. P.. 2013. The climate of the Late Cretaceous: new insights from the application of the carbonate clumped isotope thermometer to Western Interior Seaway macrofossil. Earth and Planetary Science Letters 362:5165.Google Scholar
Dutton, A., Huber, B. T., Lohmann, K. C., and Zinsmeister, W. J.. 2007. High-resolution stable isotope profiles of a dimitobelid belemnite: implications for paleodepth habitat and Late Maastrichtian climate seasonality. Palaios 22:642650.Google Scholar
Etches, S., Clarke, J., and Callomon, J.. 2009. Ammonite eggs and ammonitellae from the Kimmeridge Clay Formation (Upper Jurassic) of Dorset, England. Lethaia 42:204217.Google Scholar
Fatherree, J. W., Harries, P. J., and Quinn, T. M.. 1998. Oxygen and carbon isotopic “dissection” of Baculites compressus (Mollusca: Cephalopoda) from the Pierre Shale (upper Campanian) of South Dakota; implications for paleoenvironmental reconstructions. Palaios 13:376385.Google Scholar
Fisher, C. G., and Arthur, M. A.. 2002. Water mass characteristics in the Cenomanian US Western Interior seaway as indicated by stable isotopes of calcareous organisms. Palaeogeography, Palaeoclimatology, Palaeoecology 188:189213.Google Scholar
Forsythe, J. W., and Van Heukelem, W. J.. 1987. Growth. Pp. 135156. in P. R. Boyle, ed. Cephalopod life cycles Vol. 2. Comparative reviews. Academic Press, London.Google Scholar
Gill, J. R., and Cobban, W. A.. 1966. The Red Bird section of the Upper Cretaceous Pierre Shale in Wyoming, with a section on a new echinoid from the Cretaceous Pierre Shale of eastern Wyoming. U.S. Geological Survey Professional Paper 393-A, 173.Google Scholar
Gillikin, D. P., Lorrain, A., Meng, L., and Dehairs, F.. 2007. A large metabolic carbon contribution to the δ13C record in marine aragonitic bivalve shells. Geochimica et Cosmochimica Acta 71:29362946.Google Scholar
Grossman, E. L., and Ku, T.-L.. 1986. Oxygen and carbon isotope fractionation in biogenic aragonite: temperature effects. Chemical Geology: Isotope Geoscience 59:5974.Google Scholar
Hain, M. P., Sigman, D. M., and Haug, G. H.. 2014. The biological pump in the past. Pp. 485517. in H. D. Holland, and K. K. Turekian, eds. Treatise on geochemistry, 2nd ed. Elsevier, Oxford.Google Scholar
Hein, J. R., Normark, W. R., McIntyre, B. R., Lorenson, T. D., and Powell, C. L.. 2006. Methanogenic calcite, 13C-depleted bivalve shells, and gas hydrate from a mud volcano offshore southern California. Geology 34:109112.Google Scholar
Ivany, L. C. 2012. Reconstructing paleoseasonality from accretionary skeletal carbonates - Challenges and opportunities. In L. C. Ivany and B. T. Huber, eds. Earth’s deep-time climate—the state of the art in 2012, Paleontological Society Short Course, November 3, 2012. Paleontological Society Papers 18:133–165.Google Scholar
Ivany, L. C., Wilkinson, B. H., and Jones, D. S.. 2003. Using stable isotopic data to resolve rate and duration of growth throughout ontogeny: an example from the surf clam. Spisula solidissima. Palaios 18:126137.Google Scholar
Jablonski, D., and Hunt, G.. 2006. Larval ecology, geographic range, and species survivorship in Cretaceous mollusks: organismic versus species-level explanations. American Naturalist 168:556564.Google Scholar
Jacobs, D. K., and Chamberlain, J. A.. 1996. Buoyancy and hydrodynamics in ammonoids. Pp. 169224. in N. H. Landman, K. Tanabe, and R. A. Davis, eds. Ammonoid paleobiology. Topics in Geobiology Vol. 13. Plenum, New York.Google Scholar
Jacobs, D. K., and Landman, N. H.. 1993. Nautilus—a poor model for the function and behavior of ammonoids? Lethaia 26:101111.Google Scholar
Jacobs, D. K., Landman, N. H., and Chamberlain, J. A.. 1994. Ammonite shell shape covaries with facies and hydrodynamics: iterative evolution as a response to changes in basinal environment. Geology 22:905908.Google Scholar
Jones, D. S. 1983. Sclerochronology: reading the record of the molluscan shell. American Scientist 71:384391.Google Scholar
Killick, R., and Eckley, I.. 2014. changepoint: an R package for changepoint analysis. Journal of Statistical Software 58:119.Google Scholar
Killingley, J. S., and Berger, W. H.. 1979. Stable isotopes in a mollusk shell: detection of upwelling events. Science 205:186188.Google Scholar
Kim, S.-T., Mucci, A., and Taylor, B. E.. 2007a. Phosphoric acid fractionation factors for calcite and aragonite between 25 and 75 °C: revisited. Chemical Geology 246:135146.Google Scholar
Kim, S.-T., O’Neil, J. R., Hillaire-Marcel, C., and Mucci, A.. 2007b. Oxygen isotope fractionation between synthetic aragonite and water: influence of temperature and Mg2+ concentration. Geochimica et Cosmochimica Acta 71:47044715.Google Scholar
Kita, N. T., Ushikubo, T., Fu, B., and Valley, J. W.. 2009. High precision SIMS oxygen isotope analysis and the effect of sample topography. Chemical Geology 264:4357.Google Scholar
Klug, C., Riegraf, W., and Lehmann, J.. 2012. Soft-part preservation in heteromorph ammonites from the Cenomanian–Turonian Boundary Event (OAE 2) in north-west Germany. Palaeontology 55:13071331.Google Scholar
Knoll, A. H., Bambach, R. K., Payne, J. L., Pruss, S., and Fischer, W. W.. 2007. Paleophysiology and end-Permian mass extinction. Earth and Planetary Science Letters 256:295313.Google Scholar
Kozdon, R., Ushikubo, T., Kita, N. T., Spicuzza, M., and Valley, J. W.. 2009. Intratest oxygen isotope variability in the planktonic foraminifer N. pachyderma: real vs. apparent vital effects by ion microprobe. Chemical Geology 258:327337.Google Scholar
Kozdon, R., Kelly, D. C., Kitajima, K., Strickland, A., Fournelle, J. H., and Valley, J. W.. 2013. In situ δ18O and Mg/Ca analyses of diagenetic and planktic foraminiferal calcite preserved in a deep-sea record of the Paleocene–Eocene thermal maximum. Paleoceanography 28:517528.Google Scholar
Kröger, B., Vinther, J., and Fuchs, D.. 2011. Cephalopod origin and evolution: a congruent picture emerging from fossils, development and molecules. BioEssays 33:602613.Google Scholar
Kroopnick, P., Weiss, R. F., and Craig, H.. 1972. Total CO2,13C, and dissolved oxygen 18O at Geosecs II in the North Atlantic. Earth and Planetary Science Letters 16:103110.Google Scholar
Kroopnick, P. M. 1985. The distribution of 13C of ΣCO2 in the world oceans. Deep-Sea Research, part A (Oceanographic Research Papers) 32:5784.Google Scholar
Kruta, I., Landman, N. H., and Cochran, J. K.. 2014. A new approach for the determination of ammonite and nautilid habitats. PLoS ONE 9:e87479.Google Scholar
Kulicki, C. 1996. Ammonoid shell microstructure. Pp. 65101. in N. H. Landman, K. Tanabe, and R. A. Davis, eds. Ammonoid paleobiology. Topics in Geobiology Vol. 13. Plenum, New York.Google Scholar
Kump, L. R., and Slingerland, R. L.. 1999. Circulation and stratification of the early Turonian Western Interior Seaway: sensitivity to a variety of forcings. Pp. 181–190 in E. Barrera and C. C. Johnson, eds. Evolution of the Cretaceous ocean–climate system. Geological Society of America Special Paper 332.Google Scholar
Landman, N. H. 1987. Ontogeny of Upper Cretaceous (Turonian–Santonian) scaphitid ammonites from the Western Interior of North America: systematics, developmental patterns, and life history. Bulletin of the American Museum of Natural History 185:117241.Google Scholar
Landman, N. H., and Cobban, W. A.. 2007. Redescription of the Late Cretaceous (Late Santonian) ammonite Desmoscaphites bassleri Reeside, 1927, from the Western Interior of North America. Rocky Mountain. Geology 42:6794.Google Scholar
Landman, N. H., and Cochran, J. K.. 2010. Growth and longevity of Nautilus . Pp. 401420. in W. B. Saunders, and N. H. Landman, eds. Nautilus: the biology and paleobiology of a living fossil. Topics in Geobiology Vol. 6. Springer, Dordrecht, Netherlands.Google Scholar
Landman, N. H., and Klofak, S. M.. 2012. Anatomy of a concretion: life, death and burial in the Western Interior Seaway. Palaios 27:672693.Google Scholar
Landman, N. H., and Waage, K. M.. 1993. Scaphitid ammonites of the Upper Cretaceous (Maastrichtian) Fox Hills Formation in South Dakota and Wyoming. Bulletin of the American Museum of Natural History 215:1257.Google Scholar
Landman, N. H., Rye, D. M., and Shelton, K. L.. 1983. Early ontogeny of Eutrephoceras compared to recent Nautilus and Mesozoic ammonites: evidence from shell morphology and light stable isotopes. Paleobiology 9:269279.Google Scholar
Landman, N. H., Cochran, J. K., Rye, D. M., Tanabe, K., and Arnold, J. M.. 1994. Early life history of Nautilus: evidence from isotopic analyses of aquarium-reared specimens. Paleobiology 20:4051.Google Scholar
Landman, N. H., Tanabe, K., and Shigeta, Y.. 1996. Ammonoid embryonic development. Pp. 343405. in N. H. Landman, K. Tanabe, and R. A. Davis, eds. Ammonoid paleobiology. Topics in Geobiology Vol. 13. Plenum, New York.Google Scholar
Landman, N. H., Klofak, S. M., and Sarg, K. B.. 2008. Variation in adult size of scaphitid ammonites from the Upper Cretaceous Pierre Shale and Fox Hills Formation. Pp. 149194. in. High-resolution approaches in stratigraphic paleontology. Springer, Dordrecht, Netherlands.Google Scholar
Landman, N. H., Cobban, W. A., and Larson, N. L.. 2012a. Mode of life and habitat of scaphitid ammonites. Geobios 45:8798.Google Scholar
Landman, N. H., Cochran, J. K., Larson, N. L., Brezina, J., Garb, M. P., and Harries, P. J.. 2012b. Methane seeps as ammonite habitats in the U.S. Western Interior Seaway revealed by isotopic analyses of well-preserved shell material. Geology 40:507510.Google Scholar
Landman, N. H., Remin, Z., Garb, M. P., and Chamberlain, J. A. Jr. 2013. Cephalopods from the Badlands National Park area, South Dakota: reassessment of the position of the Cretaceous/Paleogene boundary. Cretaceous Research 42:127.Google Scholar
Landman, N. H., Goolaerts, S., Jagt, J. W. M., Jagt-Yazykova, E. A., Machalski, M., and Yacobucci, M. M.. 2014. Ammonite extinction and nautilid survival at the end of the Cretaceous. Geology 42:707710.Google Scholar
Landman, N. H., Grier, J. C., Grier, J. W., Cochran, J. K., and Klofak, S. M.. 2015. 3-D orientation and distribution of ammonites in a concretion from the Upper Cretaceous Pierre Shale of Montana. Swiss. Journal of Palaeontology 134:257279.Google Scholar
Lécuyer, C., and Bucher, H.. 2006. Stable isotope compositions of a late Jurassic ammonite shell: a record of seasonal surface water temperatures in the southern hemisphere? eEarth Discussions 1:119.Google Scholar
Lemanis, R., Zachow, S., Fusseis, F., and Hoffmann, R.. 2015. A new approach using high-resolution computed tomography to test the buoyant properties of chambered cephalopod shells. Paleobiology 41:313329.Google Scholar
Linzmeier, B. J., Kozdon, R., Peters, S. E., and Valley, J. W.. 2016. Oxygen isotope variability within Nautilus shell growth bands. PLoS ONE 11:e0153890.Google Scholar
Linzmeier, B. J., Kitajima, K., Denny, A. C., and Cammack, J. N.. 2018. Making maps on a micrometer scale. Eos 99.Google Scholar
Lukeneder, A. 2015. Ammonoid habitats and life history. Pp. 689791. in C. Klug, D. Korn, K. D. Baets, I. Kruta, and R. H. Mapes, eds. Ammonoid paleobiology: From anatomy to ecology. Springer, Dordrecht, Netherlands.Google Scholar
Lukeneder, A., Harzhauser, M., Müllegger, S., and Piller, W. E.. 2010. Ontogeny and habitat change in Mesozoic cephalopods revealed by stable isotopes (δ18O, δ13C). Earth and Planetary Science Letters 296:103114.Google Scholar
Macintyre, I. G., and Reid, R. P.. 1995. Crystal alteration in a living calcareous alga (Halimeda): Implications for studies in skeletal diagenesis. Journal of Sedimentary Research 65:143153.Google Scholar
Mapes, R. H., and Nützel, A.. 2009. Late Palaeozoic mollusc reproduction: cephalopod egg-laying behavior and gastropod larval palaeobiology. Lethaia 42:341356.Google Scholar
McConnaughey, T. A., and Gillikin, D. P.. 2008. Carbon isotopes in mollusk shell carbonates. Geo-Marine Letters 28:287299.Google Scholar
McCorkle, D. C., Emerson, S. R., and Quay, P. D.. 1985. Stable carbon isotopes in marine porewaters. Earth and Planetary Science Letters 74:1326.Google Scholar
McCorkle, D. C., Corliss, B. H., and Farnham, C. A.. 1997. Vertical distributions and stable isotopic compositions of live (stained) benthic foraminifera from the North Carolina and California continental margins. Deep-Sea Research, part I (Oceanographic Research Papers) 44:9831024.Google Scholar
Melzner, F., Gutowska, M. A., Langenbuch, M., Dupont, S., Lucassen, M., Thorndyke, M. C., Bleich, M., and Pörtner, H. O.. 2009. Physiological basis for high CO2 tolerance in marine ectothermic animals: pre-adaptation through lifestyle and ontogeny? Biogeosciences 6:23132331.Google Scholar
Miller, A. I., and Foote, M.. 2009. Epicontinental seas versus open-ocean settings: the kinetics of mass extinction and origination. Science 326:11061109.Google Scholar
Miller, R. G. 1981. Simultaneous statistical inference. Springer Series in Statistics. Springer-Verlag, New York.Google Scholar
Mironenko, A. A., and Rogov, M. A.. 2015. First direct evidence of ammonoid ovoviviparity. Lethaia 49:245260.Google Scholar
Moriya, K. 2015. Isotope signature of ammonoid shells. Pp. 793836. in C. Klug, D. Korn, K. D. Baets, I. Kruta, and R. H. Mapes, eds. Ammonoid paleobiology: From anatomy to ecology. Springer, Dordrecht, Netherlands.Google Scholar
Moriya, K., Nishi, H., Kawahata, H., Tanabe, K., and Takayanagi, Y.. 2003. Demersal habitat of Late Cretaceous ammonoids: evidence from oxygen isotopes for the Campanian (Late Cretaceous) northwestern Pacific thermal structure. Geology 31:167170.Google Scholar
Myers, C. E., MacKenzie, R. A., and Lieberman, B. S.. 2012. Greenhouse biogeography: the relationship of geographic range to invasion and extinction in the Cretaceous Western Interior Seaway. Paleobiology 39:135148.Google Scholar
Orland, I. J. 2012. Seasonality from speleothems: high-resolution ion microprobe studies at Soreq Cave, Israel. Ph.D. dissertation. University of Wisconsin–Madison. 304 pp.Google Scholar
Orland, I. J., Bar-Matthews, M., Kita, N. T., Ayalon, A., Matthews, A., and Valley, J. W.. 2009. Climate deterioration in the Eastern Mediterranean as revealed by ion microprobe analysis of a speleothem that grew from 2.2 to 0.9 ka in Soreq Cave, Israel. Quaternary Research 71:2735.Google Scholar
Orland, I. J., Edwards, R. L., Cheng, H., Kozdon, R., Cross, M., and Valley, J. W.. 2015a. Direct measurements of deglacial monsoon strength in a Chinese stalagmite. Geology 43:555558.Google Scholar
Orland, I. J., Kozdon, R., Linzmeier, B. J., Wycech, J., Śliwiński, M. G., Kitajima, K., Kita, N. T., and Valley, J. W.. 2015b. Enhancing the accuracy of carbonate δ18O and δ13C measurements by SIMS. AGU fall meeting, San Francisco. Abstract PP52B-03.Google Scholar
Payne, J. L., and Finnegan, S.. 2007. The effect of geographic range on extinction risk during background and mass extinction. Proceedings of the National Academy of Sciences USA 104:10506–10511.Google Scholar
Peters, S. E. 2007. The problem with the Paleozoic. Paleobiology 33:165181.Google Scholar
Peters, S. E., Husson, J. M., and Czaplewski, J.. 2018. Macrostrat: a platform for geological data integration and deep-time Earth crust research. Geochemistry, Geophysics, Geosystems 19:13931409.Google Scholar
Petersen, S. V., Tabor, C. R., Lohmann, K. C., Poulsen, C. J., Meyer, K. W., Carpenter, S. J., Erickson, J. M., Matsunaga, K. K. S., Smith, S. Y., and Sheldon, N. D.. 2016. Temperature and salinity of the Late Cretaceous Western Interior Seaway. Geology 44:903906.Google Scholar
R Core Team 2017. R: a language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria. https://www.R-project.org.Google Scholar
Ritterbush, K. A., Hoffmann, R., Lukeneder, A., and De Baets, K.. 2014. Pelagic palaeoecology: the importance of recent constraints on ammonoid palaeobiology and life history. Journal of Zoology 292:229241.Google Scholar
Rocha, F., and Guerra, Á.. 1996. Signs of an extended and intermittent terminal spawning in the squids Loligo vulgaris Lamarck and Loligo forbesi Steenstrup (Cephalopoda: Loliginidae). Journal of Experimental Marine Biology and Ecology 207:177189.Google Scholar
Rocha, F., Guerra, Á., and González, Á. F.. 2001. A review of reproductive strategies in cephalopods. Biological Reviews 76:291304.Google Scholar
Rollion-Bard, C., Mangin, D., and Champenois, M.. 2007. Development and application of oxygen and carbon isotopic measurements of biogenic carbonates by ion microprobe. Geostandards and Geoanalytical Research 31:3950.Google Scholar
Romanek, C. S., Grossman, E. L., and Morse, J. W.. 1992. Carbon isotopic fractionation in synthetic aragonite and calcite: effects of temperature and precipitation rate. Geochimica et Cosmochimica Acta 56:419430.Google Scholar
Rude, P. D., and Aller, R. C.. 1991. Fluorine mobility during early diagenesis of carbonate sediment: an indicator of mineral transformations. Geochimica et Cosmochimica Acta 55:24912509.Google Scholar
Sessa, J. A., Patzkowsky, M. E., and Bralower, T. J.. 2009. The impact of lithification on the diversity, size distribution, and recovery dynamics of marine invertebrate assemblages. Geology 37:115118.Google Scholar
Sessa, J. A., Ivany, L. C., Schlossnagle, T. H., Samson, S. D., and Schellenberg, S. A.. 2012. The fidelity of oxygen and strontium isotope values from shallow shelf settings: Implications for temperature and age reconstructions. Palaeogeography, Palaeoclimatology, Palaeoecology 342–343:2739.Google Scholar
Sessa, J. A., Larina, E., Knoll, K., Garb, M., Cochran, J. K., Huber, B. T., MacLeod, K. G., and Landman, N. H.. 2015. Ammonite habitat revealed via isotopic composition and comparisons with co-occurring benthic and planktonic organisms. Proceedings of the National Academy of Sciences USA 112:15562–15567.Google Scholar
Shackleton, N. J., and Kennett, J. P.. 1975. Paleotemperature history of the Cenozoic and the initiation of Antarctic glaciation: oxygen and carbon isotope analyses in DSDP Sites 277, 279, and 281. Initial Reports of the Deep Sea Drilling Project 29:743–755. Government Printing Office, Washington, D.C.Google Scholar
Shigeta, Y. 1993. Post-hatching early life history of Cretaceous ammonoidea. Lethaia 26:133145.Google Scholar
Śliwiński, M. G., Kitajima, K., Kozdon, R., Spicuzza, M. J., Fournelle, J. H., Denny, A., and Valley, J. W.. 2016a. Secondary ion mass spectrometry bias on isotope ratios in dolomite–ankerite, part 1: δ18O matrix effects. Geostandards and Geoanalytical Research 40:157172.Google Scholar
Śliwiński, M. G., Kitajima, K., Kozdon, R., Spicuzza, M. J., Fournelle, J. H., Denny, A., and Valley, J. W.. 2016b. Secondary ion mass spectrometry bias on isotope ratios in dolomite–ankerite, part 2: δ13C matrix effects. Geostandards and Geoanalytical Research 40:173184.Google Scholar
Speden, I. G. 1970. The type Fox Hills Formation, Cretaceous (Maestrichtian), South Dakota, Part 2. Systematics of the Bivalvia. Bulletin of the Peabody Museum of Natural History 33:1222.Google Scholar
Stephen, D. A., Bylund, K. G., Garcia, P., McShinsky, R. D., and Carter, H. J.. 2012. Taphonomy of dense concentrations of juvenile ammonoids in the Upper Cretaceous Mancos Shale, east-central Utah, USA. Geobios 45:121128.Google Scholar
Stevens, K., Mutterlose, J., and Wiedenroth, K.. 2015. Stable isotope data (δ18O, δ13C) of the ammonite genus Simbirskites—implications for habitat reconstructions of extinct cephalopods. Palaeogeography, Palaeoclimatology, Palaeoecology 417:164175.Google Scholar
Stinnesbeck, W., Frey, E., and Zell, P.. 2016. Evidence for semi-sessile early juvenile life history in Cretaceous ammonites. Palaeogeography, Palaeoclimatology, Palaeoecology 457:186194.Google Scholar
Summers, W. C. 1983. Loligo pealei (Pp. 115142. in P. R. Boyle, ed. Cephalopod life cycles Vol. 1. Species accounts. Academic Press, London.Google Scholar
Tanabe, K., Landman, N. H., and Kruta, I.. 2012. Microstructure and mineralogy of the outer calcareous layer in the lower jaws of Cretaceous Tetragonitoidea and Desmoceratoidea (Ammonoidea). Lethaia 45:191199.Google Scholar
Tobin, T. S., Schauer, A. J., and Lewarch, E.. 2011. Alteration of micromilled carbonate δ18O during Kiel device analysis. Rapid Communications in Mass Spectrometry 25:21492152.Google Scholar
Tomašových, A., Schlögl, J., Biroň, A., Hudáčková, N., and Mikuš, T.. 2017. Taphonomic clock and bathymetric dependence of cephalopod preservation in bathyal, sediment-starved environments. Palaios 32:135152.Google Scholar
Tsujita, C. J., and Westermann, G. E. G.. 1998. Ammonoid habitats and habits in the Western Interior Seaway: a case study from the Upper Cretaceous Bearpaw Formation of southern Alberta, Canada. Palaeogeography, Palaeoclimatology, Palaeoecology 144:135160.Google Scholar
Urey, H. C., Lowenstam, H. A., Epstein, S., and McKinney, C. R.. 1951. Measurement of paleotemperatures and temperatures of England, Denmark, and the southeastern United States. GSA Bulletin 62:399416.Google Scholar
Valley, J. W., and Kita, N. T.. 2009. In situ oxygen isotope geochemistry by ion microprobe. In M. Fayek, ed. Secondary ion mass spectrometry in the earth sciences. Mineralogical Association of Canada Short Course Series 41:19–63.Google Scholar
Vetter, L., Kozdon, R., Valley, J. W., Mora, C. I., and Spero, H. J.. 2014. SIMS measurements of intrashell δ13C in the cultured planktic foraminifer Orbulina universa . Geochimica et Cosmochimica Acta 139:527539.Google Scholar
Waage, K. M. 1965. The Late Cretaceous coleoid cephalopod Actinosepia canadensis Whiteaves. In Postilla, Vol. 94. Peabody Museum of Natural History, Yale University, New Haven, Conn., pp. 1–33.Google Scholar
Waage, K. M. 1968. The type Fox Hills Formation, Cretaceous (Maestrichtian), South Dakota, part 1. Stratigraphy and paleoenvironments. Bulletin of the Peabody Museum of Natural History 27:33.Google Scholar
Walter, L. M., Bischof, S. A., Patterson, W. P., Lyons, T. W., O’Nions, R. K., Gruszczynski, M., Sellwood, B. W., and Coleman, M. L.. 1993. Dissolution and recrystallization in modern shelf carbonates: evidence from pore water and solid phase chemistry [and discussion]. Philosophical Transactions of the Royal Society of London A 344:2736.Google Scholar
Wani, R. 2007. How to recognize in situ fossil cephalopods: evidence from experiments with modern Nautilus . Lethaia 40:305311.Google Scholar
Wani, R. 2011. Sympatric speciation drove the macroevolution of fossil cephalopods. Geology 39:10791082.Google Scholar
Ward, P. D., and Bandel, K.. 1987. Life history strategies in fossil cephalopods. Pp. 329350. in P. R. Boyle, ed. Cephalopod life cycles Vol. 2. Comparative reviews. Academic Press, London.Google Scholar
Westermann, G. E. G. 1996. Ammonid life and habitat. Pp. 607707. in N. H. Landman, K. Tanabe, and R. A. Davis, eds. Ammonoid paleobiology. Topics in Geobiology Vol. 13. Plenum, New York.Google Scholar
Whitney, N. M., Wanamaker, A. D., Kreutz, K. J., and Introne, D. S.. 2017. Spatial and temporal variability in the δ18Ow and salinity compositions of Gulf of Maine coastal surface waters. Continental Shelf Research 137:163171.Google Scholar
Wierzbowski, H. 2007. Effects of pre-treatments and organic matter on oxygen and carbon isotope analyses of skeletal and inorganic calcium carbonate. International Journal of Mass Spectrometry 268:1629.Google Scholar
Worms, J. 1983. Loligo vulgaris. Pp. 143157. in P. R. Boyle, ed. Cephalopod life cycles Vol. 1. Species accounts. Academic Press, London.Google Scholar
Wycech, J., Kelly, D. C., Kozdon, R., Orland, I. J., Spero, H. J., and Valley, J. W.. 2018. Comparison of δ18O analyses on individual planktic foraminifer (Orbulina universa) shells by SIMS and gas-source mass spectrometry. Chemical Geology 483:119130.Google Scholar
Zakharov, Y., Shigeta, Y., Smyshlyaeva, O., Popov, A., and Ignatiev, A.. 2006. Relationship between δ13C and δ18O values of the Recent Nautilus and brachiopod shells in the wild and the problem of reconstruction of fossil cephalopod habitat. Geosciences Journal 10:331345.Google Scholar
Zakharov, Y. D., Shigeta, Y., Popov, A. M., Velivetskaya, T. A., and Afanasyeva, T. B.. 2011. Cretaceous climatic oscillations in the Bering area (Alaska and Koryak Upland): Isotopic and palaeontological evidence. Sedimentary Geology 235:122131.Google Scholar
Zumholz, K., Hansteen, T. H., Klügel, A., and Piatkowski, U.. 2006. Food effects on statolith composition of the common cuttlefish (Sepia officinalis). Marine Biology 150:237244.Google Scholar