Hostname: page-component-76fb5796d-vvkck Total loading time: 0 Render date: 2024-04-26T05:09:38.542Z Has data issue: false hasContentIssue false

Effects of mass extinction and recovery dynamics on long-term evolutionary trends: a morphological study of Strophomenida (Brachiopoda) across the Late Ordovician mass extinction

Published online by Cambridge University Press:  31 August 2018

Judith A. Sclafani
Affiliation:
Department of Geosciences, Pennsylvania State University, University Park, Pennsylvania16802, U.S.A. E-mail: jas1169@psu.edu, crc29@psu.edu, azk102@psu.edu, mep12@psu.edu
Curtis R. Congreve
Affiliation:
Department of Geosciences, Pennsylvania State University, University Park, Pennsylvania16802, U.S.A. E-mail: jas1169@psu.edu, crc29@psu.edu, azk102@psu.edu, mep12@psu.edu
Andrew Z. Krug
Affiliation:
Department of Geosciences, Pennsylvania State University, University Park, Pennsylvania16802, U.S.A. E-mail: jas1169@psu.edu, crc29@psu.edu, azk102@psu.edu, mep12@psu.edu
Mark E. Patzkowsky
Affiliation:
Department of Geosciences, Pennsylvania State University, University Park, Pennsylvania16802, U.S.A. E-mail: jas1169@psu.edu, crc29@psu.edu, azk102@psu.edu, mep12@psu.edu

Abstract

Mass extinctions affect the history of life by decimating existing diversity and ecological structure and creating new evolutionary and ecological pathways. Both the loss of diversity during these events and the rebound in diversity following extinction had a profound effect on Phanerozoic evolutionary trends. Phylogenetic trees can be used to robustly assess the evolutionary implications of extinction and origination.

We examine both extinction and origination during the Late Ordovician mass extinction. This mass extinction was the second largest in terms of taxonomic loss but did not appear to radically alter Paleozoic marine assemblages. We focus on the brachiopod order Strophomenida, whose evolutionary relationships have been recently revised, to explore the disconnect between the processes that drive taxonomic loss and those that restructure ecological communities.

A possible explanation for this disconnect is if extinction and origination were random with respect to morphology. We define morphospace using principal coordinates analysis (PCO) of character data from 61 Ordovician–Devonian taxa and their 45 ancestral nodes, defined by a most parsimonious reconstruction in Mesquite. A bootstrap of the centroid of PCO values indicates that genera were randomly removed from morphospace by the Late Ordovician mass extinction, and new Silurian genera were clustered within a smaller previously unoccupied region of morphospace. Diversification remained morphologically constrained throughout the Silurian and into the Devonian. This suggests that the recovery from the Late Ordovician mass extinction resulted in a long-term shift in strophomenide evolution. More broadly, recovery intervals may hold clues to understanding the evolutionary impact of mass extinctions.

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

Anderson, M. J., and Willis, T. J.. 2003. Canonical analysis of principal coordinates: a useful method of constrained ordination for ecology. Ecology 84:511525.Google Scholar
Ausich, W. I. 1987. Revision of Rowley’s Ordovician (?) and Silurian crinoids from Missouri. Journal of Paleontology 61:563578.Google Scholar
Ausich, W. I., and Deline, B.. 2012. Macroevolutionary transition in crinoids following the Late Ordovician extinction event (Ordovician to Early Silurian). Palaeogeography, Palaeoclimatology, Palaececology 361:3848.Google Scholar
Bapst, D. W. 2014. Preparing paleontological datasets for phylogenetic comparative methods. Pp. 515544 in L. Z. Garamszegi, ed. Modern phylogenetic comparative methods and their application in evolutionary biology. Springer, Berlin.Google Scholar
Bapst, D. W., and Hopkins, M. J.. 2017. Comparing cal3 and other a posteriori time-scaling approaches in a case study with the pterocephalid trilobites. Paleobiology 43:4967.Google Scholar
Barnes, C. R, and Zhang, S.. 1999. Pattern of conodont extinction and recovery across the Ordovician-Silurian boundary interval. In P. Kraft and O. Fatka, eds. Quo vadis Ordovician? Acta Universitatis Carolinae Geologica 43:211212.Google Scholar
Berry, W. B. N., and Boucot, A. J.. 1973. Glacio-eustatic control of Late Ordovician–Early Silurian platform sedimentation and faunal changes. Geological Society of America Bulletin 84:275284.Google Scholar
Benton, M. J. 1987. Progress and competition in macroevolution. Biological Review 62:305338.Google Scholar
Benton, M. J. 1996. On the nonprevalence of competitive replacement in the evolution of tetrapods. Pp. 185210 in D. Jablonski, D. H. Erwin, and J. Lipps, eds. Evolutionary paleobiology. Chicago: University of Chicago Press.Google Scholar
Brenchley, P. J., Carden, G. A., Hints, L., Kaljo, D., Marshall, J. D., Martma, T., Meidla, T., and Nõlvak, J.. 2003. High resolution stable isotope stratigraphy of Upper Ordovician sequences: constraints on the timing of bioevents and environmental changes associated with mass extinction and glaciation. Geological Society of American Bulletin 115:89104.Google Scholar
Brusatte, S. L., Benton, M. J., Ruta, M., and Lloyd, G. T.. 2008. Superiority, competition, and opportunism in the evolutionary radiation of dinosaurs. Science 321:14851488.Google Scholar
Brusatte, S. L., Montanari, S., Yi, H.-Y., and Norell, M. A.. 2011. Phylogenetic corrections for morphological disparity analysis: new methodology and case studies. Paleobiology 31:122.Google Scholar
Cailliez, F. 1983. The analytical solution of the additive constant problem. Psychometrika 48:305308.Google Scholar
Carlson, S. J. 1991. A phylogenetic perspective on articulate brachiopod diversity and the Permo-Triassic extinctions. In E. C. Dudley, ed. The unity of evolutionary biology. Proceedings of the Fourth International Congress of Systematic and Evolutionary Biology 1:119–142. Dioscorides Press, Portland, Ore.Google Scholar
Cavin, L., and Forey, P. L.. 2007. Using ghost lineages to identify diversification events in the fossil record. Biology Letters 2007:201204.Google Scholar
Chatterton, B. D. E., and Speyer, S. E.. 1989. Larval ecology, life history strategies, and patterns of extinction and survivorship among Ordovician trilobites. Paleobiology 15:118132.Google Scholar
Christie, M., Holland, S. M., and Bush, A. M.. 2013. Contrasting the ecological and taxonomic consequences of extinction. Paleobiology 39:538559.Google Scholar
Ciampaglio, C. N. 2004. Measuring changed in articulate brachiopod morphology before and after the Permian mass extinction event: do developmental constraints limit morphological innovation? Evolution and Development 6:260274.Google Scholar
Ciampaglio, C. N., Kemp, M., and McShea, D. W.. 2001. Detecting changes in morphospace occupation patterns in the fossil record: characterization and analysis of measures of disparity. Paleobiology 27:695715.Google Scholar
Cocks, L. R. M., and Rong, J.-Y.. 1989. Classification and review of the brachiopod superfamily Plectambonitacea. Bulletin of the British Museum (Natural History) Geology 45:77163.Google Scholar
Cocks, L. R. M., and Rong, J.-Y.. 2000. Strophomenida. Pp. 216348 in Brachiopoda. Part H of R. L. Kaesler, ed. Treatise on invertebrate paleontology. Geological Society of America, Boulder, Colo., and University of Kansas Press, Lawrence.Google Scholar
Cocks, L. R. M., and Rong, J.-Y.. 2008. Earliest Silurian faunal survival and recovery after the end Ordovician glaciation: evidence from the brachiopods. Earth and Environmental Science Transactions of the Royal Society of Edinburgh 98:291301.Google Scholar
Congreve, C. R. 2013. Cladal turnover: the end-Ordovician as a large-scale analogues of species turnover. Palaeontology 56:12851296.Google Scholar
Congreve, C. R., and Lieberman, B. S.. 2010. Phylogenetic and biogeographic analysis of deiphonine trilobites. Journal of Paleontology 84:128136.Google Scholar
Congreve, C. R., and Lieberman, B. S. 2011. Phylogenetic and biogeographic analysis of sphaerexochine trilobites. PLoS ONE 6:e21304.Google Scholar
Congreve, C. R., Falk, A. R., and Lamsdell, J. C.. 2018. Biological hierarchies and the nature of extinction. Biological Reviews. https://doi.org10.1111/brv.12368.Google Scholar
Congreve, C. R., Krug, A. Z., and Patzkowsky, M. E.. 2015. Phylogenetic revision of the Strophomenida, a diverse and ecologically important Palaeozoic brachiopod order. Palaeontology 58:743758.Google Scholar
Copeland, M. J. 1981. Latest Ordovician and Silurian ostracode faunas from Anticosti Island Québec. Pp. 185–195 in Subcommission on Silurian stratigraphy, Ordovician-Silurian Boundary Working Group, Field Meeting, Anticosti-Gaspé, Quebec.Google Scholar
Crônier, C. 2013. Morphological disparity and developmental patterning: contribution of phacopid trilobites. Palaeontology 56:12631271.Google Scholar
Droser, M. L., Bottjer, D. J., Sheehan, P. M., and McGhee, G. R. J.. 2000. Decoupling of taxonomic and ecologic severity of Phanerozoic marine mass extinctions. Geology 28:675678.Google Scholar
Eckert, J. D. 1988. Late Ordovician extinction of North American and British crinoids. Lethaia 21:147167.Google Scholar
Efron, B. 1979. Bootstrap methods: another look at the jackknife. Annals of Statistics 7:126.Google Scholar
Erwin, D. H. 1998. The end and the beginning: recoveries from mass extinctions. Trends in Ecology and Evolution Reviews 13:344349.Google Scholar
Erwin, D. H. 2000. Macroevolution is more than repeated rounds of microevolution. Evolution and Development 2:7884.Google Scholar
Erwin, D. H. 2001. Lessons from the past: biotic recoveries from mass extinctions. Proceedings of the National Academy of Sciences USA 105:1152011527.Google Scholar
Erwin, D. H. 2007. Disparity: morphological pattern and developmental context. Palaeontology 50:5773.Google Scholar
Erwin, D. H. 2008. Extinction as the loss of evolutionary history. Proceedings of the National Academy of Sciences USA 98:53995403.Google Scholar
Faith, D. P. 1992. Systematics and conservation: on predicting the feature diversity of subsets of taxa. Cladistics 8:361373.Google Scholar
Finnegan, S., Bergmann, K., Eiler, J. M., Jones, D. S., Fike, D. A., Eisenman, I., Hughes, N. C., Tripati, A. K., and Fischer, W. W.. 2011. The magnitude and duration of Late Ordovician–Early Silurian glaciation. Science 331:903906.Google Scholar
Finnegan, S., Heim, N. A., Peters, S. E., and Fischer, W. W.. 2012. Climate change and the selective signature of the Late Ordovician mass extinction. Proceedings of the National Academy of Sciences USA 109:68296834.Google Scholar
Finnegan, S., Rasmussen, C. M. Ø., and Harper, D. A. T.. 2016. Biogeographic and bathymetric determinants of brachiopod extinction and survival during the Late Ordovician mass extinction. Proceedings of the Royal Society of London B 283:20160007. https://doi.org10.1098/rspb.2016.0007.Google Scholar
Foote, M. 1991. Morphological and taxonomic diversity in a clade’s history: the blastoid record and stochastic simulation. Contributions from the Museum of Paleontology, University of Michigan 28:101140.Google Scholar
Foote, M. 1993. Contributions of individual taxa to overall morphological disparity. Paleobiology 19:403419.Google Scholar
Foote, M. 1994. Morphological disparity in Ordovician–Devonian crinoids and the early saturation of morphological space. Paleobiology 20:320344.Google Scholar
Foote, M. 1997. The evolution of morphological diversity. Annual Review of Ecology and Systematics 28:129152.Google Scholar
Foote, M. 1999. Morphological diversity in the evolutionary radiation of Paleozoic and post-Paleozoic crinoids. Paleobiology 25:1116.Google Scholar
Foote, M. 2000. Origination and extinction components of taxonomic diversity: general problems. Pp. 74102 in D. H. Erwin and S. L. Wing, eds. Deep time: paleobiology’s perspective. University of Chicago Press, Chicago, Ill.Google Scholar
Ghienne, J.-F., Desrochers, A., Vandenbroucke, T. R. A., Achab, A., Asselin, E., Dabard, M.-P., Farley, C., Loi, A., Paris, F., Wickson, S., and Veizer, J.. 2014. A Cenozoic-style scenario for the end-Ordovician glaciation. Nature Communications 4485. https://doi.org10.1038/ncomms5485.Google Scholar
Ghobadi Pour, M., Kebriaee-Zadeh, M. R., and Popov, L. E.. 2011. Early Ordovician (Tremadocian) brachiopods from the Eastern Alborz Mountains, Iran. Estonian Journal of Earth Sciences 60(2):6582.Google Scholar
Gould, S. J., and Lewontin, R. C.. 1979. The spandrels of San Marco and the Panglossian paradigm: a critique of the adaptionist program. Proceedings of the Royal Society of London B 205:581598.Google Scholar
Gower, J. C. 1966. Some distance properties of latent root and vector methods used in multivariate analysis. Biometrics 53:325338.Google Scholar
Gower, J. C. 2005. Principal coordinates analysis. Pp. 42334237 in Encyclopedia of biostatistics, 2nd ed. Wiley, New York.Google Scholar
Harper, D. A. T., and Rong, J.. 1995. Patterns of change in the brachiopod faunas through the Ordovician–Silurian interface. Modern Geology 20:83100.Google Scholar
Hethering, A. J., Sherratt, E., Ruta, M., Wilkinson, M., Deline, B., and Donoghue, P. C. J.. 2015. Do cladistic and morphometric data capture common patterns of morphological disparity? Palaeontology 58:393399.Google Scholar
Hopkins, M. J. 2014. The environmental structure of trilobite morphological disparity. Paleobiology 40:352373.Google Scholar
Hopkins, M. J. 2017. How well does a part represent the whole? A comparison of cranidial shape evolution with exoskeletal character evolution in the trilobite family Pterocephaliidae. Palaeontology 60:309318.Google Scholar
Holland, S. M., and Patzkowsky, M. E.. 2015. The stratigraphy of mass extinction. Palaeontology 58:903924.Google Scholar
Hughes, M., Gerber, S., and Wills, M. A.. 2013. Clades reach highest morphological disparity early in their evolution. Proceedings of the National Academy of Sciences USA 110:1387513879.Google Scholar
Jablonski, D. 1986. Evolutionary consequences of mass extinctions. Pp. 313329 in D. M. Raup and D. Jablonski, eds. Patterns and processes in the history of life. Springer, Berlin.Google Scholar
Jablonski, D. 2000. Micro- and macroevolution: scale and hierarchy in evolutionary biology and paleobiology. Paleobiology 26:1552.Google Scholar
Jablonski, D. 2002. Survival without recovery after mass extinctions. Proceedings of the National Academy of Sciences USA 99:81398144.Google Scholar
Jablonski, D. 2005. Mass extinction and macroevolution. Paleobiology 31:192210.Google Scholar
Jablonski, D., and Raup, D. M.. 1995. Selectivity of end-Cretaceous marine bivalve extinctions. Science 268:389391.Google Scholar
Korn, D., Hopkins, M. J., and Walton, S. A.. 2013. Extinction space: a method for the quantification and classification of changes in morphospace across extinction boundaries. Evolution 67:27952810.Google Scholar
Kowalewski, M., and Novack-Gottshall, P.. 2010. Resampling methods in paleontology. Pp. 19–54 in J. Alroy and G. Hunt, eds. Quantitative methods in paleobiology. Paleontological Society Papers 16. Yale University Printing and Publishing Services, New Haven, Conn.Google Scholar
Krug, A. Z., and Patzkowsky, M. E.. 2004. Rapid recovery from the Late Ordovician mass extinction. Proceedings of the National Academy of Sciences USA 101:1760517610.Google Scholar
Krug, A. Z., and Patzkowsky, M. E.. 2007. Geographic variation in turnover and recovery from the Late Ordovician mass extinction. Paleobiology 33:435454.Google Scholar
Krug, A. Z., and Patzkowsky, M. E.. 2015. Phylogenetic clustering of origination and extinction across the Late Ordovician mass extinction. PLoS ONE. https://doi.org10.1371/journal.pone.0144354.Google Scholar
Lamsdell, J. C. 2016. Horseshoe crab phylogeny and independent colonizations of fresh water: ecological invasion as a driver for morphological innovation. Paleontology 59:181194.Google Scholar
Lamsdell, J. C., and Selden, P. A.. 2017. From success to persistence: identifying an evolutionary regime shift in the diverse Paleozoic aquatic arthropod group Eurypterida, driven by the Devonian biotic crisis. Evolution 71:95110.Google Scholar
Lamsdell, J. C., Congreve, C. R., Hopkins, M. J., Krug, A. Z., and Patzkowsky, M. E.. 2017. Phylogenetic paleoecology: tree-thinking and ecology in deep time. Trends in Ecology and Evolution 32:452463.Google Scholar
Lane, A., Janis, C. M., and Sepkoski, J. J.. 2005. Estimating paleodiversities: a test of the taxic and phylogenetic methods. Paleobiology 31:2134.Google Scholar
Lloyd, G. T. 2016. Estimating morphological diversity and tempo with discrete character-taxon matrices: implementation, challenges, progress, and future directions Biological Journal of the Linnean Society 118:131151.Google Scholar
Lockwood, R. 2004. The K/T event and infaunality: morphological and ecological patterns of extinction and recovery in veneered bivalves. Paleobiology 30:507521.Google Scholar
Maddison, W. P., and Maddison, D. R.. 2015. Mesquite: a modular system for evolutionary analysis. Version 3.02. http://mesquiteproject.org, accessed 2011.Google Scholar
McCune, B., and Grace, J. B.. 2002. Analysis of ecological communities. MjM Software Design, Gleneden Beach, Ore.Google Scholar
McGhee, G. R. Jr., Sheehan, P. M., Bottjer, D. J., and Droser, M. L.. 2004. Ecological ranking of Phanerozoic biodiversity crises: ecological and taxonomic severities are decoupled. Palaeogeography, Palaeoclimatology, Palaeoecology 211:289297.Google Scholar
McGhee, G. R. Jr., Sheehan, P. M., Bottjer, D. J., and Droser, M. L.. 2012. Ecological ranking of Phanerozoic biodiversity crises: the Serpukhovian (early Carboniferous) crisis had a greater ecological impact than the end-Ordovician. Geology 40:147150.Google Scholar
McGhee, G. R. Jr., Clapham, M. E., Sheehan, P. M., Bottjer, D. J., and Droser, M. L.. 2013. A new ecological-severity ranking of major Phanerozoic biodiversity crises. Palaeogeography, Palaeoclimatology, Palaeoecology 370: 260270.Google Scholar
Nee, S., and May, R. M.. 1997. Extinction and the loss of evolutionary history. Science 278:692694.Google Scholar
Norell, M. A. 1993. Tree-based approaches to understanding history: comments on ranks, rules, and the quality of the fossil record. American Journal of Science 293-A:407417.Google Scholar
Norell, M. A., and Novacek, M. J.. 1992. Congruence between superposition and phylogenetic patterns: comparing cladistic patterns with fossil records. Cladistics 8:319337.Google Scholar
Peters, S. E., and Foote, M.. 2001. Biodiversity in the Phanerozoic: a reinterpretation. Paleobiology 27:583601.Google Scholar
O’Leary, M. A., and Kaufman, S. G.. 2012. Morphobank 3.0: web application for morphological phylogenetics and taxonomy. http://www.morphobank.org, accessed 2015.Google Scholar
Popov, L. E., and Cocks, L. R. M.. 2017. The world’s second oldest strophomenoid-dominated benthic assemblage in the first Dapingian (Middle Ordovician) brachiopod fauna identified from Iran. Journal of Asian Earth Sciences 140:112.Google Scholar
Raup, D. M. 1987. Biological extinction in earth history. Science 231:15281533.Google Scholar
Raup, D. M. 1991. Extinction: bad genes or bad luck? Norton, New York.Google Scholar
Raup, D. M. 1994. The role of extinction in evolution. Proceedings of the National Academy of Sciences USA 91:67586763.Google Scholar
Raup, D. M., and Sepkoski, J. J.. 1982. Mass extinctions in the marine fossil record. Science 215:15011503.Google Scholar
R Core Team. 2014. R: A language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria. http://www.R-project.org.Google Scholar
Rong, J.-Y., and Cocks, L. R. M.. 1994. True Strophomena and a revision of the classification and evolution of Strophomenoid and Strophodontoid brachiopods. Palaeontology 37:651694.Google Scholar
Rong, J.-Y., and Cocks, L. R. M.. 2014. Global diversity and endemism in Early Silurian (Aeronian) brachiopods. Lethaia 47:77106.Google Scholar
Rong, J.-Y., Boucot, A. J., Harper, D. A. T., Zhan, R.-B., and Neuman, R. B.. 2006. Global analyses of brachiopod faunas through the Ordovician and Silurian transition: reducing the role of the Lazarus effect. Canadian Journal of Earth Sciences 43:2339.Google Scholar
Roy, K., Hunt, G., and Jablonski, D.. 2009. Phylogenetic conservatism of extinctions in marine bivalves. Science 325:733737.Google Scholar
Sepkoski, J. J. 1996. Patterns of Phanerozoic extinction: a perspective from global databases. Pp. 3551 in O.H. Walliser, ed. Global events and event stratigraphy. Springer, Berlin.Google Scholar
Sheehan, P. M. 1973. The relation of Late Ordovician glaciation to the Ordovician–Silurian changeover in North American brachiopod faunas. Althaea 6:147154.Google Scholar
Sheehan, P. M. 2001. The Late Ordovician mass extinction. Annual Review of Earth and Planetary Science 29:331364.Google Scholar
Simpson, C., and Harnik, P. G.. 2009. Assessing the role of abundance in marine bivalve extinction over the post-Paleozoic. Paleobiology 35:631647.Google Scholar
Simpson, G. G. 1944. Tempo and mode in evolution. Columbia University Press, New York.Google Scholar
Simpson, G. G. 1953. The major features of evolution. Columbia University Press, New York.Google Scholar
Thorne, P. M., M. Ruta, M., and Benton, M. J.. 2011. Resetting the evolution of marine reptiles at the Triassic-Jurassic boundary. Proceedings of the National Academy of Sciences USA 108:83398344.Google Scholar
Westrop, S. R. 1989. Macroevolutionary implications of mass extinction—evidence from an Upper Cambrian stage boundary. Paleobiology 15:4652.Google Scholar
Wills, M. A. 2001. Morphological disparity: a primer. Pp. 55144 in J. M. Adrain, G. D. Edgecombe, and B. S. Lieberman, eds. Fossils, phylogeny, and form: an analytical approach. Springer-Science+Business, New York.Google Scholar
Wright, D. R., and Stigall, A. L.. 2013. Geologic drivers of Late Ordovician faunal change in Laurentia: investigating links between tectonics, speciation, and biotic invasions. PLoS ONE 8:e68353.Google Scholar
Wright, D. R., and Toom, U.. 2017. New crinoids from the Baltic region (Estonia): fossil tip-dating phylogenetics constrains the origin and Ordovician–Silurian diversification of the Flexibilia (Echinodermata). Palaeontology. https://doi.org10.1111/pala.12324.Google Scholar
Zhan, R., Huang, B., Liang, Y., and Jin, J.. 2013. Pulses of the Early Ordovician brachiopod radiation in South China. In A. Lundskog and K. Mehlqvist, eds., Proceedings of the Third IGCP 591 Annual Meeting, Lund, Sweden, 9–19 June.Google Scholar
Zhou, L., Algeo, T. J., Shen, J., Hu, Z., Gong, H., Xie, S., Huang, J., and Gao, S.. 2015. Changes in marine productivity and redox conditions during the Late Ordovician Hirnantian glaciation. Palaeogeography, Palaeoclimatology, Palaeoecology 20:223234.Google Scholar
Supplementary material: PDF

Sclafani et al. supplementary material

Sclafani et al. supplementary material 1

Download Sclafani et al. supplementary material(PDF)
PDF 885.6 KB
Supplementary material: File

Sclafani et al. supplementary material

Sclafani et al. supplementary material 2

Download Sclafani et al. supplementary material(File)
File 206.3 KB
Supplementary material: File

Sclafani et al. supplementary material

Sclafani et al. supplementary material 3

Download Sclafani et al. supplementary material(File)
File 17.6 KB
Supplementary material: File

Sclafani et al. supplementary material

Sclafani et al. supplementary material 4

Download Sclafani et al. supplementary material(File)
File 12.6 KB