Hostname: page-component-8448b6f56d-cfpbc Total loading time: 0 Render date: 2024-04-25T00:37:12.005Z Has data issue: false hasContentIssue false

Environmental influence on growth history in marine benthic foraminifera

Published online by Cambridge University Press:  18 September 2018

Caitlin R. Keating-Bitonti
Affiliation:
Department of Geological Sciences, Stanford University, Stanford, California 94305, U.S.A. E-mail: Keating-BitontiC@si.edu
Jonathan L. Payne
Affiliation:
Department of Geological Sciences, Stanford University, Stanford, California 94305, U.S.A. E-mail: Keating-BitontiC@si.edu

Abstract

Energy availability influences natural selection on the ontogenetic histories of organisms. However, it remains unclear whether physiological controls on size remain constant throughout ontogeny or instead shift as organisms grow larger. Benthic foraminifera provide an opportunity to quantify and interpret the physicochemical controls on both initial (proloculus) and adult volumes across broad environmental gradients using first principles of cell physiology. Here, we measured proloculus and adult test dimensions of 129 modern rotaliid species from published images of holotype specimens, using holotype size to represent the maximum size of all species’ occurrences across the North American continental margin. We merged size data with mean annual temperature, dissolved oxygen concentration, particulate organic carbon flux, and seawater calcite saturation for 718 unique localities to quantify the relationship between physicochemical variables and among-species adult/proloculus size ratios. We find that correlation of community mean adult/proloculus size ratios with environmental parameters reflects covariation of adult test volume with environmental conditions. Among-species proloculus sizes do not covary identifiably with environmental conditions, consistent with the expectation that environmental constraints on organism size impose stronger selective pressures on adult forms due to lower surface area-to-volume ratios at larger sizes. Among-species adult/proloculus size ratios of foraminifera occurring in resource-limited environments are constrained by the limiting resource in addition to temperature. Identified limiting resources are food in oligotrophic waters and oxygen in oxygen minimum zones. Because among-species variations in adult/proloculus size ratios from the North American continental margin are primarily driven by the local environment’s influence on adult sizes, the evolution of foraminiferal sizes over the Phanerozoic may have been strongly influenced by changing oceanographic conditions. Furthermore, lack of correspondence between among-species proloculus sizes and environmental conditions suggests that offspring sizes in foraminifera are rarely limited by physiological constraints and are more susceptible to selection related to other aspects of fitness.

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.)

Footnotes

*

Present address: Department of Paleobiology, National Museum of Natural History Smithsonian Institution, Washington, D.C. 20013, U.S.A.

References

Literature Cited

Alegret, L., Ortiz, S., Arenillas, I., and Molina, E.. 2010. What happens when the ocean is overheated? The foraminiferal response across the Paleocene–Eocene Thermal Maximum at the Alamedilla section (Spain). Geological Society of America Bulletin 122:16161624.Google Scholar
Antonov, J. I., Seidov, D., Boyer, T. P., Locarnini, R. A., Mishonov, A. V., Garcia, H. E., Baranova, O. K., Zweng, M. M., and Johnson, D. R.. 2010. World ocean atlas 2009, Vol. 2. Salinity. P. 184 in S. Levitus, ed. NOAA atlas NESDIS 69. Government Printing Office, Washington, D.C.Google Scholar
Beavington-Penney, S., and Racey, A.. 2004. Ecology of extant nummulitids and other larger benthic foraminifera: applications in palaeoenvironmental analysis. Earth-Science Reviews 67:219265.Google Scholar
Berger, W. H. 1969. Planktonic foraminifera: basic morphology and ecological implications. Journal of Paleontology 43:13691383.Google Scholar
Bernhard, J. M. 1986. Characteristic assemblages and morphologies of benthic foraminifera from anoxic, organic-rich deposits—Jurassic through Holocene. Journal of Foraminiferal Research 16:207215.Google Scholar
Bernhard, J. M., and Alve, E.. 1996. Survival, ATP pool, and ultrastructural characterization of benthic foraminifera from Drammensfjord (Norway): response to anoxia. Marine Micropaleontology 28:517.Google Scholar
Bernhard, J. M., and Bowser, S. S.. 2008. Peroxisome proliferation in foraminifera inhabiting the chemocline: an adaption to reactive oxygen species exposure? Journal of Eukaryotic Microbiology 2008:135144.Google Scholar
Bernhard, J. M., Casciotti, K. L., McIlvin, M. R., Beaudoin, D. J., Visscher, P. T., and Edgcomb, V. P.. 2012a. Potential importance of physiologically diverse benthic foraminifera in sedimentary nitrate storage and respiration. Journal of Geophysical Research 117:G03002.Google Scholar
Bernhard, J. M., Edgcomb, V. P., Casciotti, K. L., McIlvin, M. R., and Beaudoin, D. J.. 2012b. Denitrification likely catalyzed by endobionts in an allogromiid foraminifer. ISME Journal 6:951960.Google Scholar
Bijma, J., Faber, W. W. Jr., and Hemleben, C.. 1990. Temperature and salinity limits for growth and survival of some planktonic foraminifers in laboratory cultures. Journal of Foraminiferal Research 20:95116.Google Scholar
Bozinovic, F., Calosi, P., and Spicer, J. I.. 2011. Physiological correlates of geographic range in animals. Annual Review of Ecology, Evolution, and Systematics 42:155179.Google Scholar
Brown, J. H. 1995. Macroecology. University of Chicago Press, Chicago.Google Scholar
Brown, J. H., Gillooly, J. F., Allen, A. P., Savage, V. M., and West, G. B.. 2004. Toward a metabolic theory of ecology. Ecology 85:17711789.Google Scholar
Buick, D. P., and Ivany, L. C.. 2004. 100 years in the dark: extreme longevity of Eocene bivalves from Antarctica. Geology 32:921924.Google Scholar
Button, D. J., Llyod, G. T., Ezcurra, M. D., and Butler, R. J.. 2017. Mass extinctions drove increased global faunal cosmopolitanism on the supercontinent Pangaea. Nature Communications 8:17.Google Scholar
CARINA Group. 2009. Carbon in the Arctic Mediterranean Seas Region—the CARINA project: Results and DATA. Carbon Dioxide Information Analysis Center, Oak Ridge National Laboratory, U.S. Department of Energy, Oak Ridge, Tenn.Google Scholar
Caval-Holme, F., Payne, J. L., and Skotheim, J. M.. 2013. Constraints on the adult-offspring size relationship in protists. Evolution 67:35373544.Google Scholar
Chen, J., and Strous, M.. 2013. Denitrification and aerobic respiration, hybrid electron transport chains and co-evolution. Biochimica et Biophysica Acta 1827:136144.Google Scholar
Clarke, A. 1979. Living in cold water—k-strategies in antarctic benthos. Marine Biology 55:111119.Google Scholar
Culver, S. J., and Buzas, M. A.. 1980. Distribution of recent benthic foraminifera off the North American Atlantic coast. Smithsonian Institution Press, Washington, D.C.Google Scholar
Culver, S. J., and Buzas, M. A.. 1981. Distribution of recent benthic foraminifera in the Gulf of Mexico. Smithsonian Institution Press, Washington, D.C.Google Scholar
Culver, S. J., and Buzas, M. A.. 1982. Distribution of recent benthic foraminifera in the Caribbean region. Smithsonian Institution Press, Washington, D.C.Google Scholar
Culver, S. J., and Buzas, M. A.. 1985. Distribution of recent benthic foraminifera off the North American Pacific coast from Oregon to Alaska. Smithsonian Institution Press, Washington, D.C.Google Scholar
Culver, S. J., and Buzas, M. A.. 1986. Distribution of recent benthic foraminifera off the North American Pacific coast from California to Baja. Smithsonian Institution Press, Washington, D.C.Google Scholar
Culver, S. J., and Buzas, M. A.. 1987. Distribution of recent benthic foraminifera off the Pacific coast of Mexico and Central America. Smithsonian Institution Press, Washington, D.C.Google Scholar
Cushman, J. A. 1905. Developmental stages in the Lagenidae. American Naturalist 39:537553.Google Scholar
DeLong, J. P., Okie, J. G., Moses, M. E., Sibly, R. M., and Brown, J. H.. 2010. Shifts in metabolic scaling, production, and efficiency across major evolutionary transitions of life. Proceedings of the National Academy of Sciences USA 107:1294112945.Google Scholar
Douglas, R., and Staines-Urias, F.. 2007. Dimorphism, shell Mg/Ca ratios and stable isotope content in species of Bolivina (benthic foraminifera) in the Gulf of California, Mexico. Journal of Foraminiferal Research 37:189203.Google Scholar
Douglas, R. G. 1987. Paleoecology of the continental margin basins: a modern case history from the borderland of Southern California. Pp. 81117 in D. S. Gorsline, ed. Depositional systems in active margin basins. Pacific Section of the Society of Economic Paleontologists and Mineralogists, Los Angeles, Calif.Google Scholar
Eder, W., Briguglio, A., and Hohenegger, J.. 2016. Growth of Heterostegina depressa under natural and laboratory conditions. Marine Micropaleontology 122:2743.Google Scholar
Eldredge, N., Thompson, J. N., Brakefield, P. M., Gavrilets, S., Jablonski, D., Jackson, J. B. C., Lenski, R. E., Lieberman, B. S., McPeek, M. A., and Miller, W. III. 2005. The dynamics of evolutionary stasis. Paleobiology 31:133145.Google Scholar
Ellis, B. F., and Messina, A. R.. 1940–2006. Catalogue of foraminifera. American Museum of Natural History, New York.Google Scholar
Environmental Systems Research Institute. 2011. ArcGIS Desktop. ESRI, Redlands, Calif.Google Scholar
Findlay, H. S., Wood, H. F., Kendall, M. A., Spicer, J. I., Twitchett, R. J., and Widdicombe, S.. 2009. Calcification, a physiological process to be considered in the context of the whole organism. Biogeosciences Disscussions 6:22672284.Google Scholar
Garcia, H. E., Locarnini, R. A., Boyer, T. P., Antonov, J. I., Baranova, O. K., Zweng, M. M., and Johnson, D. R.. 2010. World ocean atlas 2009, Vol. 3. Dissolved oxygen, apparent oxygen utilization, and oxygen saturation. P. 344 in S. Levitus, ed. NOAA atlas NESDIS 70. U.S. Government Printing Office, Washington, D.C.Google Scholar
Gattuso, J.-P., Epitalon, J. M., and Lavigne, H.. 2015 seacarb: seawater carbonate chemistry, R package, Version 3.0.11.Google Scholar
Geslin, E., Barras, C., Langlet, D., Nardelli, M. P., Kim, J.-H., Bonnin, J., Metzger, E., and Jorissen, F.. 2014. Survival, reproduction and calcification of three benthic foraminiferal species in response to experimentally induced hypoxia. Pp. 163193 in H. Kitazato and J. M. Bernhard, eds. Approaches to study living foraminifera: collection, maintenance and experimentation. Springer, Japan.Google Scholar
Gillooly, J. F., Brown, J. H., West, G. B., Savage, V. M., and Charnov, E. L.. 2001. Effects of size and temperature on metabolic rate. Science 293:22482251.Google Scholar
Gillooly, J. F., Charnov, E. L., West, G. B., Savage, V. M., and Brown, J. H.. 2002. Effects of size and temperature on developmental time. Nature 417:7073.Google Scholar
Goldstein, S. T. 1999. Foraminifera: a biological overview. Pp. 3756 in B. K. Sen Gupta, ed. Modern foraminifera. Kluwer Academic, London.Google Scholar
Gooday, A. J., Levin, L. A., da Silva, A. A., Bett, B. J., Cowie, G. L., Dissard, D., Gage, J. D., Hughes, D. J., Jeffreys, R., Lamont, P. A., Larkin, K. E., Murty, S. J., Schumacher, S., Whitcraft, C., and Woulds, C.. 2009. Faunal responses to oxygen gradients on the Pakistan margin: a comparison of foraminiferans, macrofauna and megafauna. Deep-Sea Research, Part II (Topical Studies in Oceanography) 56:488502.Google Scholar
Groves, J. R., Rettori, R., Payne, J. L., Boyce, M. D., and Altiner, D.. 2007. End-Permian mass extinction of lagenide foraminifers in the Southern Alps (Northern Italy). Journal of Paleontology 81:415434.Google Scholar
Hallock, P. 1985. Why are larger foraminifera large? Paleobiology 11:195208.Google Scholar
Hansen, H. J. 1999. Shell construction in modern calcareous foraminifera. Pp. 5770 in B. K. Sen Gupta, ed. Modern foraminifera. Kluwer Academic, London.Google Scholar
Helly, J. J., and Levin, L. A.. 2004. Global distribution of naturally occurring marine hypoxia on continental margins. Deep-Sea Research Part I (Oceanographic Research Papers) 51:11591168.Google Scholar
Hohenegger, J., and Briguglio, A.. 2014. Methods for estimating growth pattern and lifetime of foraminifera based on chamber volumes. Pp. 2954 in H. Kitazato and J. M. Bernhard, eds. Approaches to study living foraminifera: collection, maintenance and experimentation. Springer, Japan.Google Scholar
Hottinger, L. 1982. Larger foraminifera, giant-cells with a historical background. Naturwissenschaften 69:361371.Google Scholar
Hunt, G., and Roy, K.. 2006. Climate change, body size evolution, and Cope’s Rule in deep-sea ostracodes. Proceedings of the National Academy of Sciences USA 103:13471352.Google Scholar
Hunt, G., Wicaksono, S. A., Brown, J. E., and MacLeod, K. G.. 2010. Climate-driven body-size trends in the ostracod fauna of the deep Indian Ocean. Palaeontology 53:12551268.Google Scholar
IOC, IHO, and BODC. 2003. Centenary Edition of the GEBCO Digital Atlas. British Oceanographic Data Centre, Liverpool, U.K.Google Scholar
Kaiho, K. 1994. Benthic foraminiferal dissolved-oxygen index and dissolved-oxygen levels in the modern ocean. Geology 22:719722.Google Scholar
Kaiho, K. 1998. Global climatic forcing of deep-sea benthic foraminiferal test size during the past 120 m.y. Geology 26:491494.Google Scholar
Kaiho, K. 1999. Evolution in the test size of deep-sea benthic foraminifera during the past 120 m.y. Marine Micropaleontology 37:5365.Google Scholar
Kaiho, K., Takeda, K., Petrizzo, M. R., and Zachos, J. C.. 2006. Anomalous shifts in tropical Pacific planktonic and benthic foraminiferal size during the Paleocene–Eocene thermal maximum. Palaeogeography, Palaeoclimatology, Palaeoecology 237:456464.Google Scholar
Keating-Bitonti, C. R., and Payne, J. L.. 2016. Physicochemical controls on biogeographic variation of benthic foraminiferal test size and shape. Paleobiology 42:595611.Google Scholar
Keating-Bitonti, C. R., and Payne, J. L.. 2017. Ecophenotypic responses of benthic foraminifera to oxygen availability along an oxygen gradient in the California Borderland. Marine Ecology 38:e12430.Google Scholar
Key, R. M., Kozyr, A., Sabine, C. L., Lee, K., Wanninkhof, R., Bullister, J., Feely, R. A., Millero, F., Mordy, C., and Peng, T.-H.. 2004. A global ocean carbon climatology: results from GLODAP. Global Biogeochemical Cycles 18:GB4031.Google Scholar
Koho, K. A., Piña-Ochoa, E., Geslin, E., and Risgaard-Petersen, N.. 2011. Vertical migration, nitrate uptake and denitrification: survival mechanisms of foraminifers (Globobulimina turgida) under low oxygen conditions. FEMS Microbiology Ecology 75:273283.Google Scholar
Korsun, S., Hald, M., Panteleeva, N., and Tarasov, G.. 1998. Biomass of foraminifera in the St. Anna Trough, Russian Arctic continental margin. Sarsia 83:419431.Google Scholar
Kosnik, M. A., Jablonski, D., Lockwood, R., and Novack-Gottshall, P. M.. 2006. Quantifying molluscan body size in evolutionary and ecological analyses: maximizing the return on data-collection efforts. Palaios 21:588597.Google Scholar
Legendre, P. 2014. lmodel2: model II regression, R package, Version 1.7-2.Google Scholar
Levin, L. A. 2003. Oxygen minimum zone benthos: adaptation and community response to hypoxia. Oceanography and Marine Biology 41:145.Google Scholar
Locarnini, R. A., Mishonov, A. V., Antonov, J. I., Boyer, T. P., Garcia, H. E., Baranova, O. K., Zweng, M. M., and Johnson, D. R.. 2010. World ocean atlas 2009, Vol. 1: Temperature. P. 184 in S. Levitus, ed. NOAA atlas NESDIS 68. U.S. Government Printing Office, Washington, D.C.Google Scholar
Loeblich, A. R. Jr., and Tappan, H.. 1988. Foraminiferal genera and their classification. Van Nostrand Reinhold, New York.Google Scholar
Lombard, F., Labeyrie, L., Michel, E., Spero, H. J., and Lea, D. W.. 2009. Modelling the temperature dependent growth rates of planktic foraminifera. Marine Micropaleontology 70:17.Google Scholar
Lutz, M. J., Caldeira, K., Dunbar, R. B., and Behrenfeld, M. J.. 2007. Seasonal rhythms of net primary production and particulate organic carbon flux to depth describe the efficiency of biological pump in the global ocean. Journal of Geophysical Research (Oceans) 112(C10011).Google Scholar
Lutze, G. 1964. Statistical investigation on the variability of Bolivina argentea . Contributions from the Cushman Laboratory for Foraminiferal Research 15:105116.Google Scholar
Morel, A., Claustre, H., and Gentili, B.. 2010. The most oligotrophic subtropical zones of the global ocean: similarities and differences in terms of chlorophyll and yellow substance. Biogeosciences 7:31393151.Google Scholar
Moss, D. K., Ivany, L. C., Judd, E. J., Cummings, P. W., Bearden, C. E., Kim, W.-J., Artruc, E. G., and Driscoll, J. R.. 2016. Lifespan, growth rate and body size across latitude in marine Bivalvia, with implications for Phanerozoic evolution. Proceedings of the Royal Society of London B 283:20161364.Google Scholar
Myers, E. H. 1943. The biology, ecology, and morphologenesis of a pelagic foraminifer. Stanford University Press, Stanford, Calif.Google Scholar
Neuheimer, A. B., Hartvig, M., Heuschele, J., Hylander, S., Kiørboe, T., Olsson, K. H., Sainmont, J., and Andersen, K. H.. 2015. Adult and offspring size in the ocean over 17 orders of magnitude follows two life hisotry strategies. Ecology 96:33033311.Google Scholar
Nigam, R. 1986. Dimorphic forms of recent foraminifera: an additional tool in paleoclimatic studies. Palaeogeography, Palaeoclimatology, Palaeoecology 53:239244.Google Scholar
Nigam, R., and Khare, N.. 1995. Significance of correspondence between river discharge and proloculus size of benthic foraminifera in paleomonsoonal studies. Geo-Marine Letters 15:4550.Google Scholar
Nigam, R., and Rao, A. S.. 1987. Proloculus size variation in recent benthic foraminifera: implication for paleoclimatic studies. Estuarine, Coastal and Shelf Science 24:649655.Google Scholar
Novack-Gottshall, P. M. 2008. Using simple body-size metrics to estimate fossil body volume: empirical validation using diverse Paleozoic invertebrates. Palaios 23:163173.Google Scholar
Park, M. Y., and Hastie, T.. 2013. glmpath: L1 regularization path for generalized linear models and cox proportional hazards model, R package. Version 0:97.Google Scholar
Paulmier, A., Ruiz-Pino, D., and Garcon, V.. 2011. CO2 maximum in the oxygen minumum zone (OMZ). Biogeosciences 8:239252.Google Scholar
Payne, J. L., Summers, M., Rego, B. L., Altiner, D., Wei, J., Yu, M., and Lehrmann, D. L.. 2011. Early and Middle Triassic trends in diversity, evenness, and size of foraminifers on a carbonate platform in south China: implications for tempo and mode of biotic recovery from the end Permian mass extinction. Paleobiology 37:409425.Google Scholar
Payne, J. L., Groves, J. R., Jost, A. B., Nguyen, T., Moffitt, S. E., Hill, T. M., and Skotheim, J. M.. 2012. Late Paleozoic fusulinoidean gigantism driven by atmospheric hyperoxia. Evolution 66:29292939.Google Scholar
Payne, J. L., Jost, A. B., Wang, S. C., and Skotheim, J. M.. 2013. A shift in the long-term mode of foraminiferan size evolution caused by the end-Permian mass extinction. Evolution 67:816827.Google Scholar
Pianka, E. R. 1970. On r-selection and K-selection. American Naturalist 104:592597.Google Scholar
Picken, G. B. 1979. Growth, production and biomass of the antarctic gastropod Laevilacunaria antarctica Martens 1885. Journal of Experimental Marine Biology and Ecology 40:7179.Google Scholar
Piña-Ochoa, E., Høgslund, S., Geslin, E., Cedhagen, T., Revsbech, N. P., Nielsen, L. P., Schweizer, M., Jorissen, F., Rysgaard, S., and Risgaard-Petersen, N.. 2010. Widespread occurrence of nitrate storage and denitrification among foraminifera and Gromiida. Proceedings of the National Academy of Sciences USA 107:11481153.Google Scholar
Racey, A. 2001. A review of Eocene nummulite accumulations: structure, formation and reservior potential. Journal of Petroleum Geology 24:79100.Google Scholar
R Core Team. 2015. R: a language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria.Google Scholar
Rego, B. L., Wang, S. C., Altiner, D., and Payne, J. L.. 2012. Within- and among-genus components of size evolution during mass extinction, recovery, and background intervals: a case study of Late Permian through Late Triassic foraminifera. Paleobiology 38:627643.Google Scholar
Rex, M. A., Etter, R. J., Morris, J. S., Crouse, J., McClain, C. R., Johnson, N. A., Stuart, C. T., Deming, J. W., Thies, R., and Avery, R.. 2006. Global bathymetric patterns of standing stock and body size in the deep-sea benthos. Marine Ecology Progress Series 317:18.Google Scholar
Risgaard-Petersen, N., Langezaal, A. M., Ingvardsen, S., Schmid, M. C., Jetten, M. S. M., Op den Camp, H. J. M., Derksen, J. W. M., Piña-Ochoa, E., Eriksson, S. P., Nielsen, L. P., Revsbech, N. P., Cedhagen, T., and van der Zwaan, G. J.. 2006. Evidence for complete denitrification in a benthic foraminifer. Nature 443:9396.Google Scholar
Röttger, R., Irwan, A., and Schmaljohann, R.. 1980. Growth of symbiont-bearing foraminifera Amphistegina lessonii D’Orbigny and Heterostegina depressa D’Orbigney (Protozoa). Walter de Gruyter, Berlin.Google Scholar
Saraswat, R., Nigam, R., and Barreto, L.. 2005. Palaeoceanographic implications of abundance and mean proloculus diameter of benthic foraminiferal species Epistominella exigua in sub-surface sediments from distal Bay of Bengal fan. Journal of Earth System Sciences 114:453458.Google Scholar
Schenck, H. G. 1944. Proloculus in foraminifera. Journal of Paleontology 18:275282.Google Scholar
Schmidt, D. N., Renaud, S., Bollmann, J., Schiedbel, R., and Thierstein, H. R.. 2004. Size distribution of Holocene planktic foraminifer assemblages: biogeography, ecology and adaptation. Marine Micropaleontology 50:319338.Google Scholar
Schöne, B. R., Tanabe, K., Dettman, D. L., and Sato, S.. 2003. Environmental controls on shell growth rates and d18O of the shallow-marine bivalve mollusk Phacosoma japonicum in Japan. Marine Biology 142:473485.Google Scholar
Seiglie, G. A. 1976. Significance of proloculus size in the foraminifer Fursenkoina punctata (d’Orbigny). Micropaleontology 22:485490.Google Scholar
Sen Gupta, B. K., and Machain-Castillo, M. L.. 1993. Benthic foraminifera in oxygen-poor habitats. Marine Micropaleontology 20:183201.Google Scholar
Song, H., Tong, J., and Chen, Z. Q.. 2011. Evolutionary dynamics of the Permian–Triassic foraminifer size: evidence for Lilliput effect in the end-Permian mass extinction and its aftermath. Palaeogeography Palaeoclimatology Palaeoecology 308:98110.Google Scholar
Spalding, C., Fischer, W. W., and Finnegan, S.. 2015. The sensitivity of calcification energetics to ocean acidification. Geological Society of America Abstracts with Programs 47:639.Google Scholar
Speakman, J. R. 2005. Body size, energy metabolism and lifespan. Journal of Experimental Biology 208:17171730.Google Scholar
Spero, H. J., and Lea, D. W.. 1996. Experimental determination of stable isotope variability in Globigerina bulloides: implications for paleoceanographic reconstructions. Marine Micropaleontology 28:231246.Google Scholar
Strohm, T. O., Griffin, B., Zumft, W. G., and Schink, B.. 2007. Growth yields in bacterial denitrification and nitrate ammonification. Applied and Environmental Microbiology 73:14201424.Google Scholar
Thomas, E. 1980. The development of Uvigerina in the Cretan Mio-Pliocene. Utrecht Micropaleontological Bulletin 23:168.Google Scholar
Travis, J. L., and Bowser, S. S.. 1991. The motility of foraminifera. Pp. 91155 in J. J. Lee and O. R. Anderson, eds. Biology of the foraminifera. Academic, London.Google Scholar
Vance, R. R. 1973a. More on reproductive strategies in marine benthic invertebrates. American Naturalist 107:353361.Google Scholar
Vance, R. R. 1973b. Reproductive strategies in marine benthic invertebrates. American Naturalist 107:339352.Google Scholar
Van Gorsel, J. T. 1978. Late Cretaceous orbitoidal foraminifera. Academic, New York.Google Scholar
Van Voorhies, W. A. 2001. Metabolism and lifespan. Experimental Gerontology 36:5564.Google Scholar
Winguth, A. M. E., Thomas, E., and Winguth, C.. 2012. Global decline in ocean ventilation, oxygenation, and productivity during the Paleocene–Eocene Thermal Maximum: Implications for the benthic extinction. Geology 40:263266.Google Scholar
Wood, H. F., Spicer, J. I., and Widdicombe, S.. 2008. Ocean acidification may increase calcification rates, but at a cost. Proceedings of the Royal Society of London B 275:17671773.Google Scholar