<<

Phylogenetic : Phenotypic Diversification and Evolutionary Radiation in

Paleozoic

DISSERTATION

Presented in Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy in the Graduate School of The Ohio State University

By

David F. Wright

Graduate Program in Geological

The Ohio State University

2016

Dissertation Committee:

William I. Ausich, Advisor

Matthew R. Saltzman

Lawrence A. Krissek

John V. Freudenstein

Copyrighted by

David F. Wright

2016

Abstract

Phylogenetic paleobiology is an interdisciplinary research program at the nexus of

, , and evolutionary . Specifically, phylogenetic

paleobiology integrates the deep-time data, techniques, and geologic perspectives of paleontology with phylogeny-based statistical and computational approaches in macroevolutionary biology. Chapters contained within this dissertation revolve around two primary themes: (1) understanding patterns of change over geologic time, and (2) assembling a clearer picture of the phylogeny, , and geologic history of marine , especially the Crinoidea (Echinodermata).

Under the umbrella of phylogenetic paleobiology, the primary objective of this dissertation is to help bridge the disciplinary gap between specimen-based paleontology

and statistical approaches in phylogenetic comparative methods. The chapters herein use

advanced statistical phylogenetic methods, such as Bayesian “tip-dating” approaches, to

infer phylogenetic trees of , quantify rates of phenotypic evolution, and

document patterns of morphospace occupation among fossil members of the Crinoidea

(Echinodermata).

In addition to these broader studies, a major taxonomic revision of fossil and extant Crinoidea (Echinodermata) is proposed herein, as well as taxonomic description of

a new of fossil from the (Katian) of Ontario.

ii

This work is jointly (and lovingly) dedicated to (1) my parents, for being supportive of my decision to pursue academic interests; (2) to my wife Lena, for her encouragement, friendship, and kindness; and (3) to my Gluey, for showing me how beautiful the

world can be.

iii

Acknowledgments

This dissertation would not have been possible without the assistance and support

of many individuals. I would like to thank to following people for their gracious

assistance to museum collections: K. Hollis (United States Museum of Natural History),

T. Ewin (Natural History Museum), P. Mayer (Field Museum of Natural History), and K.

Riddigton (Lapworth Museum of Geology). J. Koniecki is thanked for collecting and generously donating specimens used in this dissertation. This dissertation was supported by numerous student research grants, including those from the Paleontological Society,

Sigma Xi, the Palaeontological Association, the American Federation of Mineralogical and Geological Societies, the Association of Applied Paleontological Sciences, Friends of

Orton Hall, and a Presidential Fellowship from The Ohio State University. Additional support for this research was provided by the National Foundation’s Assembling the Tree of (DEB 10364116) to W.I. Ausich.

I have been very fortunate in my academic career to have had wonderful and gracious academic mentors who have helped guide my path. Greg Farley is thanked for igniting my earliest interests in science (pretty late by most standards!), letting me borrow his personal copy of S.J. Gould’s The Flamingo’s Smile, and encouraging me to seek a university degree. Bruce Lieberman is thanked for warmly encouraging my interests in macroevolutionary biology and paleontology when I was an undergraduate at the

iv

University of . I thank Alycia Stigall for her delightful (and highly contagious) enthusiasm for specimen-based paleontology. I cannot overstate how much my academic trajectory benefited from her guidance and training when I was a M.Sc. student at Ohio

University. Alycia has continued to be a mentor and I thank her for her support and friendship. I would especially like to thank my Ph.D. advisor, Bill Ausich, for his invaluable (and seemingly encyclopaedic) knowledge of fossil crinoids and willingness to have a scientific discussion with me at a moment’s notice. Bill has been an excellent

Ph.D. advisor, mentor, and friend. Most of all, I thank Bill for instilling in me a passion for studying fossil crinoids and introducing me to the delectable wonder of Bell’s Two-

Hearted Ale.

Many ideas presented in this dissertation were greatly aided through informal conversations with friends and colleagues over the , including Dave Bapst, Tom

Baumiller, Lena Cole, Tom Kammer, David Polly, Dan Rabosky, Graham Slater, Jeff

Thompson, and Peter Wagner. I would also like to thank those who provided specific feedback on earlier drafts of the chapters presented herein and/or peer-reviewed the published portions of this dissertation, including W.I. Ausich, S.R. Cole, A.

Gavryushkina, T.W. Kammer, G.D. Sevastopulo, G.J. Slater, and two anonymous reviewers. Stig Bergström and Dale Gnidovic are thanked for contributing an atmosphere of fossil-based comraderie that makes me proud to have been a part of Ohio State paleontology.

I am thankful to have participated in several workshops that greatly improved my programming skills in R and led to my better understanding of analytical approaches in

v

the fields of paleontology and phylogenetic biology. I would like to thank the organizers,

instructors, and fellow participants of the 2013 Analytical Methods in Paleobiology

course held in Sydney, ; the National Evolutionary Synthesis Center sponsored

“Paleobiological and Phylogenetic Approaches to ” course held in

Durham, NC; the Computational Macroevolution workshop at the 2015 meeting of the

Society of Systematic Biologists in Ann Arbor, MI; and the “Phylogeny developeR

bootcamp” held at the University of Massachusetts Boston’s field station in Nantucket,

MA.

I thank Lena Cole for putting up with my perambulating (and at times

incomprehensible) vocalizations on various topics related to this dissertation. No one has listened to me ramble, rant, and think out loud about inferring phylogenies, statistical methods, , and macroevolution more than her. Lena also read and helped edit my manuscript drafts and generously provided assistance with specimen illustration and photography. I have greatly benefited from her company both in and out of the office and will always appreciate her accompanying me on my daily walks to get coffee. I especially thank Lena for being my best friend, office mate, colleague, and loving spouse.

I’ve never been able to relate to the well-circulated blog posts and comics on social media that promulgate the stereotype of unhappy, disgruntled graduate students.

My experience has been the exact opposite. These four years as a Ph.D. student at Ohio

State have been the best of my life. Many thanks to all of those whom I have not

mentioned yet were a part of this wonderful experience. Cheers. –Davey

vi

Vita

2010...... B.S. Geology, University of Kansas

2012...... M.S. Geological Sciences, Ohio University

2012 to 2016 ...... Graduate Teaching Associate, School of

Earth Sciences, The Ohio State University

2016 to present ...... Presidential Fellow, The Ohio State

University

Publications

Wright, D.F., W.I. Ausich, S.R. Cole, M.E. Peter, and E.C. Rhenberg, in press, Phylogenetic and classification of the Crinoidea (Echinodermata). Journal of Paleontology.

Wright, D.F. in press, Bayesian estimation of fossil phylogenies and the evolution of early to middle crinoids (Echinodermata). Journal of Paleontology.

Wright, D.F., 2015, , , and “Phylogenetic Paleo-ontogeny”: a reassessment of primary posterior plate homologies in fossil and living crinoids with insight from . Paleobiology, 41:570-591.

Ausich, W.I., T.W. Kammer, E.C. Rhenberg, and D.F. Wright, 2015, Early phylogeny of crinoids within the Pelmatozoan . Palaeontology, 58, 937-952.

Wright, D.F. and W.I. Ausich, 2015, From the stem to the crown: phylogeny and diversification of pan-cladid crinoids. Pp. 199-202 in Progress in Echinoderm

vii

Palaeobiology. S. Zamora and I. Rábono (eds.), Cuadernos del museo Geominero, 19, Instituto Geológico y Minero de España, 292 pp.

Ausich, W.I., T.W. Kammer, D.F. Wright, S.R. Cole, M.E. Peter, and E.C. Rhenberg, 2015, Toward a phylogenetic classification of the Crinoidea (Echinodermata). Pp. 29-32 in Progress in Echinoderm Palaeobiology. S. Zamora and I. Rábono (eds.), Cuadernos del museo Geominero, 19, Instituto Geológico y Minero de España, 292 pp.

Wright, D.F. and A.L. Stigall, 2014, Species-level phylogenetic revision of the Late Ordovician orthid genus Glyptorthis. Journal of Systematic Palaeontology, DOI 10.1080/14772019.2013.839584

Wright, D.F. and A.L. Stigall, 2013, Phylogenetic revision of the Late Ordovician orthid brachiopod genera Plaesiomys and Hebertella from . Journal of Paleontology, 87: 1107-1128.

Wright, D.F. and A.L. Stigall, 2013, Geological drivers of Late Ordovician faunal change in : investigating links between tectonics, , and biotic invasions. PLoS ONE, 8(7): e68353. doi:10.1371/journal.pone.0068353

Wright, D.F. and P.A. Selden, 2011, A Trigonotarbid from the of Kansas. Journal of Paleontology, 85: 871-876.

Fields of Study

Major Field: Geological Sciences

Paleontology,

viii

Table of Contents

Abstract ...... ii

Dedication ...... iii

Acknowledgements ...... iv

Vita ...... vii

List of Tables ...... xi

List of Figures ...... xii

Chapter 1: Overview ...... 1

Chapter 2: Fossil, homology, and “Phylogenetic Paleo-ontogeny”: a reassessment of primary posterior plate homologies among fossil and living crinoids with insights from developmental biology ...... 13

Chapter 3: Bayesian estimation of fossil phylogenies and the evolution of early to middle

Paleozoic crinoids (Echinodermata) ...... 77

Chapter 4: Phylogenetic taxonomy and classification of the Crinoidea (Echinodermata)

...... 150

ix

Chapter 5: Ecologic innovation and phenotypic constraint in the evolutionary radiation of

Paleozoic crinoids ...... 231

Chapter 6: Crinoids from the Bobcaygeon and Verulam Formations (Upper Ordovician) near Brechin, Ontario: new taxa and emended descriptions ...... 289

References ...... 343

Appendix A: Character list for phylogenetic analysis in Chapter 5 ...... 399

x

List of Tables

Table 3.1 Species included in phylogenetic analysis...... 139

Table 4.1 Species name, least inclusive clade, and first appearance interval for each depicted in Figure 4.1...... 222

Table 4.2 Revised rank-based classification of the Crinoidea...... 224

xi

List of Figures

Figure 1.1 The logical structure and interdependence of research domains within phylogenetic paleobiology...... 11

Figure 2.1 Summary of previously proposed phylogenetic relationships of pan-cladid crinoids...... 69

Figure 2.2 Generalized of cladid crinoids...... 71

Figure 2.3 An antinomy of alternative primary homology schemes for a single posterior plate in the genus Phanocrinus...... 73

Figure 2.4 Pentacrinoid development and morphogenesis of the extant comatulid crinoid meridionalis...... 75

Figure 3.1 Illustration of the fossilized birth- process...... 142

Figure 3.2 Parsimony-based estimate of rate variation among characters...... 144

Figure 3.3 Testing statistical associations between the amount of morphologic evolution inferred from an undated analysis and divergence times estimated under a relaxed morphologic- model...... 146

Figure 3.4 Maximum Clade Credibility tree of early to middle Paleozoic crinoids...... 148

Figure 4.1 Taxa representing major crinoid ...... 227

Figure 4.2 depicting phylogenetic relationships among species used to define major clades within the Crinoidea...... 229

Figure 5.1 Macroevolutionary models of trait evolution...... 269

Figure 5.2 Rates of phenotypic evolution in eucladid crinoids during the Paleozoic ...... 271

Figure 5.3 Disparity and diversity profiles for Paleozoic eucladid crinoids...... 273

xii

Figure 5.4 Time-varying diversification and sampling parameters estimated from phylogenetic analysis and stratigraphic ranges of genera...... 275

Figure 5.5 Rates of morphologic evolution among eucladid subclades ...... 277

Figure 5.6 Identification of adaptive zones based on principal coordinate analysis of the morphologic character matrix ...... 279

Figure 5.7 Phylomorphospace and adaptive zone occupation throughout the history of Paleozoic radiation...... 281

Figure 5.8 The relationship between taxonomic and morphologic diversification in Paleozoic eucladids...... 283

Figure 5.9 Additional plots depicting the radiation of eucladids in diversity-disparity space...... 285

Figure 5.10 Maximum Clade Credibility tree from Bayesian analysis of 81 Paleozoic eucladid crinoid species...... 287

Figure 6.1 Locality map of the Lake Simcoe region in Ontario ...... 315

Figure 6.2 of the Simcoe Group highlighting the position of the Bobcaygeon and Verulam Formations...... 317

Figure 6.3 Abundance distribution of specimens represented in the Brechin collections ...... 319

Figure 6.4 Konieckicrinus brechinensis n. gen. n. sp., specimen 5097.2 ...... 321

Figure 6.5 Konieckicrinus brechinensis n. gen. n. sp., specimen 5097.3 ...... 323

Figure 6.6 Konieckicrinus josephi n. gen. n. sp., specimen 5097.7JK ...... 325

Figure 6.7 Konieckicrinus josephi n. gen. n. sp., specimen 5097.1JK ...... 327

Figure 6.8 Konieckicrinus josephi n. gen. n. sp., specimen 5097.6 ...... 329

Figure 6.9 Carabocrinus radiatus Billings, 1857 ...... 331

Figure 6.10 Carabocrinus vancortlandti Billings, 1859 ...... 333

Figure 6.11 Grenprisia springeri Moore 1962 ...... 335

xiii

Figure 6.12 Illemocrinus amphiatus Eckert, 1987 ...... 337

Figure 6.13 Cupulocrinus humilis (Billings, 1857) ...... 339

Figure 6.14 Cupulocrinus jewetti (Billings, 1859) ...... 341

xiv

Chapter 1: Overview

Introduction

A fundamental aspect of biodiversity is its tremendous propensity to vary across

multiple scales and hierarchical levels. Components of biodiversity, such as morphologic

diversity and taxonomic richness, vary widely among clades and through geologic time.

Understanding the processes that regulate large- variation in Earth’s biodiversity

requires integrating quantitative models of phenotypic evolution with macroevolutionary

theories describing how large-scale Earth- processes and global change events

modify the structure and evolution of adaptive landscapes over geologic time.

For at least 3.5 billion years, episodes of diversification and have

shaped and pruned the vast evolutionary . Despite the spectacular array of

Earth’s present day biodiversity, the estimated ~8.7 million species inhabiting the planet today represent a small sample of all species that have ever existed (Foote and Miller,

2007; Mora et al., 2011).

Fortunately, the fossil record provides a wealth of data regarding the morphology, spatial distribution, and temporal duration of lineages over long timescales. Although also incompletely sampled, the fossil record details the geologic ’s biodiversity and offers the only direct, empirical record of evolution’s trajectory, long- term trends, intermediate forms, and former glory of extinct clades. Thus, the fossil

1

record provides a natural, model system for studying macroevolution—commonly

defined as either (1) evolutionary changes taking place over time scales inaccessible to

direct observation or experimentation on extant species, (2) evolutionary patterns among

species rather than within an ancestor-descendant lineage, or (3) any process(es) resulting

in large-scale evolutionary change (Simpson, 1944; Eldredge and Gould, 1972; Stanley,

1979; Erwin, 2000; Hunt and Rabosky, 2014).

Phylogenetic Paleobiology

Phylogenetic paleobiology is an emerging interdisciplinary research program at

the nexus of paleontology, systematics, and evolutionary biology. Specifically,

phylogenetic paleobiology integrates the deep-time data, techniques, and geologic perspectives of paleontology with phylogeny-based statistical and computational approaches in macroevolutionary biology. Phylogenetic paleobiology is primarily concerned with the development and application of mathematical methods and evolutionary models to (1) estimate evolutionary relationships among lineages of extinct taxa, and (2) to study the tempo and mode of trait evolution among lineages using phylogenetic trees of fossil species (Hunt and Slater, 2016). A schematic of this research program and how it unites the major features of this dissertation is depicted in Figure 1.1.

Crinoidea (Echinodermata)

Echinoderms are a diverse and ecologically significant of marine

with an exceptional fossil record and are represented by more than 7,000 species in

2

today’s oceans. The Echinodermata includes familiar animals such as , sea urchins, and sand dollars. The apparent diversity of alive today masks their more prodigious geologic history. For example, the subclass Crinoidea ( stars and

sea lilies) comprises an evolutionary lineage of -dwelling to deep water echinoderms represented by ~600 species today, yet more than 8,000 fossil species have been

described spanning some 480 million years of evolutionary history.

Fossil crinoids are an ideal taxon for studying trait evolution and diversification

because they are a long ranging, well sampled taxonomic group (Foote and Raup, 1996)

and preserve aspects of their feeding and functional morphology as fossils (Lane,

1971; Ausich, 1980; Brower, 2007, Baumiller, 2008). Further, crinoids are highly

amenable for morphology-based phylogenetic analyses because they are comprised of

many discrete, homologous morphological characters (Moore and Laudon, 1943; Ausich,

1988, 1998; Foote, 1999). The combination of a well-sampled geologic history and the

propensity to construct phylogenetic hypotheses makes fossil crinoids a well-suited taxon

for integrating phylogenetic comparative methods with paleontological approaches to

studying macroevolution (Slater and Harmon, 2013).

Summary of Contents

Using the framework of phylogenetic paleobiology and the fossil record of

crinoid echinoderms, this dissertation is comprised of six semi-independent research

projects concerning the phylogeny, classification, phenotypic evolution, and evolutionary

radiation of Paleozoic crinoids, with emphasis placed on the .

3

Each chapter represents a distinct contribution proportional to a stand alone

scientific paper (Figure 1.1). Indeed, the content in chapters 2-4 has either already been

published (Wright, 2015) or accepted for publication in peer-reviewed scientific journals

(Wright, in press; Wright et al., in press). Chapters 5-6 represent early versions of manuscripts to be submitted for publication following the defense of this dissertation.

The details of these projects are described more fully below.

Chapter 2: Fossils, homology, and “Phylogenetic Paleo-ontogeny”: a reassessment of primary posterior plate homologies among fossil and living crinoids with insights from developmental biology

Homology is a fundamental concept in biology because it provides an evolutionary explanation for why similar features are shared by different organisms—

they both inherited that feature from a common ancestor. However, the distributions of

“primary” homologies among species reflect a prior belief in what constitutes comparable

organismal elements and are the principal determinants of the outcome of phylogenetic

analysis. Thus, it is critical that primary homology statements are well justified when

conducting phylogeny-based research. This chapter combines information on fossil

morphology, phylogenetic systematics, and insights from evolutionary developmental

biology to reassess a set of historically contentious calyx plate homologies in non-

camerate crinoids. The content of this chapter was published in Paleobiology (Wright,

2015).

4

Chapter 3: Bayesian estimation of fossil phylogenies and the evolution of early to middle

Paleozoic crinoids (Echinodermata)

The application of computational methods to infer evolutionary relationships among fossil species is becoming increasingly common in paleobiological studies.

However, most methods of phylogenetic inference typically used by paleontologists do not accommodate the idiosyncrasies of fossil data and therefore do not take full advantage of the information provided by the fossil record. However, recent advances in fossil “tip-dating” offer paleontologists a statistical approach to combine morphologic and stratigraphic data with probabilistic models to infer phylogenies of fossil taxa. In this chapter, I describe the basics of Bayesian tip-dating , describe the fossilized birth‒death process model and its utility as a tree prior distribution, and present an empirical application of these techniques inferring the phylogeny of Ordovician through crinoids. The content of this chapter has been accepted for publication in the Journal of Paleontology (Wright, in press).

Chapter 4: Phylogenetic taxonomy and classification of the Crinoidea (Echinodermata)

Biological classification systems are most useful when they convey phylogenetic relationships and facilitate lucid communication among researchers. In the Linnaean system, species are arrayed into higher taxonomic ranks of increasing inclusivity based on various criteria for delimiting taxa and their ranks. Ideally, a taxon’s rank in the hierarchy provides a coarse statement of evolutionary relationships. In contrast, phylogenetic taxonomy eschews ranks and instead defines taxa based on phylogenetic

5

relationships alone and delineates their taxonomic content based on phylogenetic hypotheses. This chapter builds on a number of recent quantitative phylogenetic analyses and uses the principles of phylogenetic taxonomy to propose a of stem- and node- based definitions for all major taxa (fossil and extant) traditionally recognized within the

Crinoidea. In addition, a phylogeny-based revision to the traditional rank-based Linnaean

classification is also proposed. The content of this chapter has been accepted for

publication in the Journal of Paleontology (Wright et al., in press).

Chapter 5: Ecologic innovation and phenotypic constraint in the evolutionary radiation

of Paleozoic crinoids

The application of model-based approaches to estimating evolutionary trees and

the development of phylogeny-based statistical methods have greatly expanded

paleobiological research programs investigating the tempo and mode of morphologic

evolution. In this chapter, I test the relationships between phylogeny-based rates of

morphologic evolution, taxonomic diversification, and morphospace occupation in the

~200 million evolutionary radiation of Paleozoic crinoids. Results indicate

phenotypic diversification is more complex than models commonly assumed in

comparative biology, at least over geologic timescales—highlighting the need for

continued synthesis between fossil and phylogenetic approaches to macroevolution. This

chapter represents an early version of a manuscript currently in preparation for

submission to a peer-reviewed journal.

6

Chapter 6: Crinoids from the Bobcaygeon and Verulam Formations (Upper Ordovician) near Brechin, Ontario: new taxa and emended descriptions

The Upper Ordovician (lower Katian) Bobcaygeon and Verulam Formations near

Brechin, Ontario are comprised of a highly diverse, well-preserved crinoid fana. This provides an exceptional window into a taxonomically diverse Late Ordovician crinoid from the interval between the diversity peak and the end-

Ordovician extinction. A survey of newly collected material from the Bobcaygeon and

Verulam Formations reveals a large number of exceptionally preserved crinoid specimens with arms, stems, and attachment structures intact. Focusing on taxa placed within the

Cladida, this chapter contains taxonomic descriptions for a new genus, comprised of two new species, and additional descriptions for other species within the Brechin fauna. This chapter is a nascent version of a manuscript focusing on cladid species preserved in the

Brechin fauna and is part of a much broader work conducting a taxonomic re-evaluation of fossil crinoids from the Ordovician of Ontario.

7

References

Ausich. W.I., 1980, A model for niche differentiation in Lower Mississippian crinoid

communities: Journal of Paleontology, v. 54, p. 273-288.

Ausich. W.I., 1988, Evolutionary convergence and parallelism in crinoid calyx design:

Journal of Paleontology, v. 62. P. 906-916.

Ausich. W.I., 1998, Phylogeny of Arenig to Caradoc crinoids (Phylum Echinodermata)

and suprageneric classification of the Crinoidea: The University of Kansas

Paleontological Contributions, v. 9, p. 1-36.

Baumiller, T.K., 2008, Crinoid ecological morphology: Annual Review of Earth and

Planetary Sciences, v. 36, p. 221-249.

Brower, J.C., 2007,The application of filtration theory to food gathering in Ordovician

crinoids: Journal of Paleontology, v. 81, p. 1284-1300.

Eldredge, N., and Gould, S.J., 1972, Punctuated equilibria: an alternative to phyletic

gradualism, p. 82-115. In T. J. M. Schopf (ed.), Models in Paleobiology. Freeman,

Cooper, San Francisco.

Erwin, D.H., 2000, Macroevolution is more than repeated rounds of :

Evolution and Development, v. 2, p. 78-84.

Foote, M. 1999, Morphological diversity in the evolutionary radiation of Paleozoic and

post-Paleozoic crinoids: Paleobiology, v. 25, p. 1-115.

Foote, M., and Miller, A.I., 2007, Principles of Paleontology, 3rd edition: W.H. Freeman,

San Francisco, 354 p.

8

Foote, M., and Raup, D.M., 1996, Fossil preservation and the stratigraphic ranges of taxa:

Paleobiology, v. 22:121-140.

Hunt, G., and Rabosky, D.L., 2014, Phenotypic evolution in fossil species: pattern and

process: Annual Review of Earth and Planetary Sciences, v. 42, p. 421-441.

Hunt, G., and Slater, G., 2016, Integrating paleontological and phylogenetic approaches

to macroevolution: Annual Review of Ecology, Evolution, and Systematics, v. 47.

P. 189–213.

Lane, N.G., 1971, Crinoids and reefs: Proceedings of the North American Paleontological

Convention, v. 2, p.1430-1443.

Moore, R.C., and Laudon, L.R., 1943, Evolution and Classification of Paleozoic crinoids:

Geological Society of America Special Paper, v. 46, p. 1-154.

Mora, C., Tittensor. D.P., Adl, S., Simpson A.G.B., Worm, B., 2011, How Many Species

Are There on Earth and in the Ocean? PLoS Biol 9(8): e1001127. doi:

10.1371/journal.pbio.1001127

Simpson, G.G., 1944, Tempo and Mode in Evolution: Columbia University Press, 237 p.

Slater, G.J., and Harmon, L.J., 2013, Unifying fossils and phylogenies for comparative

analyses of diversification and trait evolution: Methods in Ecology and Evolution,

v. 4, p. 699-702.

Stanley, S.M., 1979, Macroevolution: pattern and process: Freeman, San Francisco, 332

p.

Wright, D.F., 2015, Fossils, homology, and “Phylogenetic Paleo-ontogeny”: a

reassessment of primary posterior plate homologies among fossil and living

9

crinoids with insights from developmental biology: Paleobiology, v. 41, p. 570‒

591.

Wright, D.F., in press, Bayesian estimation of fossil phylogenies and the evolution of

early to middle Paleozoic crinoids (Echinodermata): Journal of Paleontology.

Wright, D.F., Ausich, W.I., Cole, S.R., Peter, M.E., Rhenburg, E.C., in press,

Phylogenetic taxonomy and classification of the Crinoidea: Journal of

Paleontology.

10

Figure 1.1 The logical structure and interdependence of research domains within phylogenetic paleobiology. Major research areas emphasized in each chapter are depicted along with their cascading effects, where each arrow tip points toward an area that depends (in part) on information gathered within the domain located at the arrow tail.

Areas of emphasis within phylogenetic paleobiology are in bold, italicized text.

11

Fig. 1.1

12

Chapter 2: Fossils, homology, and “Phylogenetic Paleo-ontogeny”: a reassessment of

primary posterior plate homologies among fossil and living crinoids with insights from

developmental biology

Abstract.—Paleobiologists must propose a priori hypotheses of homology when

conducting a phylogenetic analysis of extinct taxa. The distributions of such “primary”

homologies among species are fundamental to phylogeny reconstruction because they

reflect a prior belief in what constitutes comparable organismal elements and are the

principal determinants of the outcome of phylogenetic analysis. Problems arise when

fossil morphology presents seemingly equivocal hypotheses of homology, herein referred

to as antinomies. In groups where homology recognition has been elusive, such as

echinoderms, these problems are commonly accompanied by the presence (and

persistence) of poor descriptive terminology in taxonomic literature that confounds an

understanding of characters and stymy phylogenetic research. This paper combines fossil

morphology, phylogenetic systematics, and insights from evolutionary developmental

biology to outline a research program in Phylogenetic Paleo-ontogeny. A “paleo”-

ontogenetic approach to character analysis provides a logical basis for homology recognition and discerning patterns of character evolution in a phylogenetic context. To illustrate the utility of the paleo-ontogenetic approach, I present a reassessment of

13

historically contentious plate homologies for “pan-cladid” crinoids (Cladida, Flexibilia,

Articulata). Developmental patterns in living crinoids were combined with the fossil record of pan-cladid morphologies to investigate primary posterior plate homologies.

Results suggest the sequence of morphologic transitions unfolding during the ontogeny of

extant crinoids are developmental relics of their Paleozoic precursors. Developmental

genetic modules controlling posterior plate development in pan-cladid crinoids have

likely experienced considerable constraint for over 250 million years and limited

morphologic diversity in the complexity of calyx characters. Future phylogenetic

analyses of pan-cladids are recommended to consider the presence of a single plate in the

posterior region homologous with the radianal, rather than the anal X, as is commonly

assumed.

Introduction

“…to avoid the taint of theory in morphology is impossible, however much it may be

wished. The whole science is riddled with theory. Not a specimen can be described

without the use of a terminology coloured by theory, implying the acceptance of some one

or other theory of homologies.” –William Bateson (1894: p. vii)

Paleobiologists study the rich, geologic history of fossilized morphologies to

examine patterns and processes in macroevolution. The ~520 million year history of

14

metazoan evolution revealed by fossils provides an empirical record of mass ,

morphologic innovations, and adaptive radiations otherwise unknowable from

observations on extant species alone (Sepkoski 1981; Alroy 2010). Such studies are of

interest to ecologists, developmental, and evolutionary biologists because they offer

insight into evolution on timescales inaccessible to direct observation or experimentation.

Integration between paleobiology and other biologic disciplines frequently leads to a

better understanding of evolutionary patterns and processes at multiple hierarchical

levels. For example, combining the stratigraphic record of fossil occurrences with recent

discoveries in evolutionary developmental biology (EDB, or “evo-devo”) has facilitated unprecedented insights into the causal mechanisms of morphologic diversity, evolution, and developmental links between micro- and macroevolution (Shubin and

Marshall 2000; Raff 2007; Wagner 2007; Carroll 2008; Erwin and Davidson 2009).

Many studies combining fossils with other biologic disciplines, such as developmental biology, utilize phylogenetic information at coarse taxonomic levels (Raff

2007). Unfortunately for paleobiologists, transformations between ancestor-descendent morphologies at low taxonomic levels are not always unambiguously arrayed in stratigraphic succession. The fossil record is notoriously incomplete and can obscure patterns of first and last appearances (Smith 2001). In addition to paleontologic incompleteness, and/or heterogeneous rates of phenotypic evolution frequently confound accurate interpretations of morphologic disparity and taxonomic evolution

(Erwin 2007). In an attempt to account for these difficulties, paleobiologists are increasingly using quantitative computational methods (e.g., maximum parsimony,

15

Bayesian inference) to construct phylogenetic hypotheses for fossil taxa rather than

relying on “expert” taxonomic opinion and/or the stratigraphic distribution of first and

last appearances. Reconstructed phylogenies of fossil taxa are becoming commonplace

for analyzing a broad swath of macroevolutionary topics ranging from extinction

dynamics (Purvis 2004; Harnik et al. 2014) to evolution (Wagner 1995;

Bapst 2014) and paleobiogeography (Lieberman 2001; Wright and Stigall 2014).

Importantly, phylogenetic information is necessary for fossil occurrences to be

meaningfully applied with other kinds of evolutionary data such as in EDB (Raff 2007).

The use of model phylogenies as a template for paleobiological studies is not

without concern (Wagner 2000a). Setting aside methodological issues involving

optimality criteria (Wright and Hillis 2014), the accuracy of recovered phylogenies are

fundamentally limited by the information content of their underlying morphologic data

and a researcher’s ability to codify fossil morphology into a set of a priori hypotheses of

homology. Therefore, choosing a robust set of primary homologies (sensu de Pinna 1991)

is critical to reconstructing model phylogenies. Given their importance for accurately

reconstructing phylogenetic trees, how should paleobiologists propose homologous

features in extinct lineages? What criteria form a logical basis for choosing among

hypotheses of homologies when alternatives are possible? How are these hypotheses tested to discover features that reflect “true” evolutionary homologies?

This paper presents a combined phylogenetic and “paleo”-ontogenetic approach

to these questions (cf. Mooi et al. 2005). A research program in Phylogenetic Paleo-

ontogeny combines fossils, phylogenetic systematics, and evolutionary developmental

16

biology to provide a logical basis for discovering homologies and discerning patterns of

character evolution among fossil species. As an example, I present a character analysis of posterior plate homologies among fossil and living crinoids (Echinodermata, Crinoidea) to illustrate how combining fossil morphology with developmental data can help resolve homology schemes and provide an ontogenetic basis for generating phylogenetic hypotheses of fossil taxa. The reassessment presented herein not only provides a step towards building a phylogeny that links extant crinoids with their Paleozoic ancestors but may also be useful guide for other researchers seeking a logical rationale when proposing homology statements for fossil taxa. As a prelude to my discussion on crinoid plate homologies, I first present an overview of theoretical and epistemological aspects of homology and discuss their relationship with phylogenetic systematics. Throughout, I integrate concepts from EDB when discussing homology and phylogenetic systematics to develop a general framework for a research program in Phylogenetic Paleo-ontogeny.

Homology: conceptual basis and operational definition

The concept of homology is fundamental to evolution and permeates all aspects of

biology (Laubichler 2000). Despite its significance, homology is a seemingly elusive concept because different disciplines use the word in different ways (Brigandt 2003). For example, an evolutionary (paleo)biologist might propose an historical definition of homology involving a comparison of morphologic structures among species to reveal a

17

sequence of adaptive transformations. In contrast, an evolutionary developmental biologist may instead prefer a mechanistic definition where homologous features are described in terms of their underlying developmental . Contrasts between perspectives are potentially problematic for relating fossilized morphologies to mechanistic definitions of homology because the overwhelming majority of species known from fossils are entirely extinct and therefore unavailable for study in an evo-devo

laboratory. More worrisome for paleobiologists is the knowledge that paralogous

may produce superficially “homologous” morphological structures, and truly

homologous loci may control non-homologous structures (e.g., Bolker and Raff 1996).

In light of evolution, differing perspectives of homology (i.e., historical and

mechanistic) among disciplines are united by a common theoretical basis: homology is

best viewed as a concept reflecting the continuity of information in the context of

phylogenetic history (van Valen 1982). Ever since Darwin (1859) the concept of

homology has been used intuitively among biologists to convey the “same” features in

different species arising from shared common ancestry. Recent insights from EDB

suggests that the “same-ness” underlying complex morphologic structures result from the

inheritance of regulatory networks (GRNs) with co-adapted transcription factors

rather than the cumulative expression of individual “homologous” genes (Wagner 2007;

Erwin and Davidson 2009). Indeed, most morphologic evolution likely results from

changes in cis-regulatory elements rather changes in gene number or function

(Carroll 2008). GRNs can be dissected into quasi-independent “developmental modules”

responsible for the occurrence, reoccurrence, and modification of a morphological

18

character across a phylogeny (Wagner 1996; Arnone and Davidson 1997). Thus, a unified

definition of homology combines developmental causality with phylogenetic continuity

(Hall 2003).

The phylogenetic distribution of homologous information is hierarchically nested

across multiple levels of biological organization—from genes to species (Hall 2003). The

notion of “deep” homology requires a decoupling between phylogenetic, phenotypic, and

genetic levels of homology. arises when identical sets of genes are

shared among phylogenetically disparate taxa despite great morphological differences

between them (Shubin and Marshall 2000). For example, the last decade of research in

EDB has discovered striking similarities among the gene families of analogous

morphological structures (Carroll 2008; Shubin and Marshall 2000). When considering

morphological traits, instances of convergence or parallelism are not homologous at the

phenotypic level even if such patterns are causally linked to “deep” homologies at the

genetic level (Hall 2003). Because this paper is about formulating homology statements

for realized fossil morphology, it is useful herein for the concept of “homology” to refer

to the subset of phylogenetic information expressed in the phenotypes of ancestor-

descendent lineages while recognizing that deeper genetic homologs (or paralogs) play an

important role in morphologic evolution.

The theory and practice of phylogenetic systematics unites ontological and

epistemological aspects of homology by providing the concept with a set of procedures that lead to its discovery in empirical data ( and Lieberman 2011). In fact, the advent of phylogenetic systematics has led many biologists to equate homology with

19

synapomorphy (Hennig 1966; Wiley 1975; Patterson 1982; Wheeler 2012). Any trait

present in an ancestor and all of its descendants is by definition a homologous trait via

continuity of descent (note that symplesiomorphies are synapomophies when considered

at a more inclusive level). Thus, “a homology is always a synapomorphy and a

synapomorphy is always a homology” (Wheeler 2012: p. 117).

Phylogenetic systematics and the discovery of homologous characters

Knowledge of “true” evolutionary homology requires knowledge of absolute truth, which is of course nonexistent in science. Phylogenetic empiricism requires observational, a priori hypotheses of homology (Wiley and Lieberman 2011). Hypotheses of homology are repeatedly tested and ultimately either corroborated or falsified on the basis of available evidence (Hennig 1966). The “discovery” of homologous characters among fossil lineages is therefore anchored by empirical observations and best approached when multiple lines of independent evidence are considered.

Ernst Haeckel claimed to have employed the “threefold parallelisms” of Louis

Agassiz to infer phylogeny: ontogenetic sequences, comparative , and the phyletic transformation of fossils in geologic succession (see Gould 1973). Although theoretical and computational aspects of modern phylogenetic methods bear little resemblance to Haeckel’s, the components of Agassiz’s parallelisms still comprise the basic character data used to propose homologous characters and build phylogenies. The

20

essence of Agassiz’s parallelisms can be combined with the ontological basis of

homology provided by EDB and the epistemological principles of phylogenetic

systematics to forge an interdisciplinary research program: Phylogenetic Paleo-ontogeny.

Characters and phylogenetic hypothesis testing—Homology statements in

phylogenetics originate as character data. Any observable, heritable organismal feature is

potentially a phylogenetically informative character (e.g., morphologic structures, a

sequence of nucleotides, developmental traits, or behavioral data). Because morphologic

characters are produced by developmental modules in the GRN, the phenotype of

organisms can similarly be atomized into components of semi-autonomous “morphologic modules”; where each module has a degree of independence despite some integration with other modules (Wagner 2006). The quasi-independent of characters underscores a common (if not ubiquitous) assumption in mathematical phylogenetic methods: a character must operationally be expressed as an independent random variable

X with N mutually exclusive transformational states (X0, X1, X2…XN) in quantitative

phylogenetic analysis (Sereno 1997).

De Pinna (1991) recognized two distinct levels of homology statements that arise

in phylogenetic systematics: primary and secondary homologies. A primary homology

refers to a proposition that two characters are homologous (de Pinna 1991). Primary

homologies are generated prior to phylogenetic analysis and therefore represent a priori

hypotheses. In practice, primary homology statements for characters X0, X1, X2…XN are

given by the distributions of transformational states among the columns of a character by

taxon matrix. Primary homology statements can be falsified as instances of or

21

corroborated by others characters via phylogenetic analysis. Patterson’s (1982)

suggestion to examine the degree of congruence among characters forms the most

decisive test of homology for morphologic characters (de Pinna 1991). Under Hennig’s

(1966) auxiliary principle, primary homologies represent inchoate hypotheses of common

descent. If a primary homology passes the Patterson’s congruence test, it matures into a

secondary homology (de Pinna 1991).

Quantitative phylogenetic methods form a natural test of Patterson’s (1982)

notion of character congruence because they utilize the covariance structure among

characters simultaneously when searching for optimal tree topologies. Once an analysis

results in a tree (or set of trees), characters can be mapped onto the tree(s) to delimit

monophyletic groups and discover instances of homoplasy. Congruence obtains when

multiple primary homology statements corroborate one another and diagnose the same

monophyletic group, viz. primary homologies become secondary homologies. However, if a primary homology statement requires independent derivations in different clades, then that primary homology has been falsified by the available evidence and cannot be considered a homology at a hierarchical level inclusive of those clades. Note that at one phylogenetic level may be considered synapomorphies when

considered at a less inclusive level. Observations of character congruence in real datasets

reflect a mixture of homology and homoplasy at different phylogenetic scales (Wagner

2000b).

Of course, erroneous assumptions about characters are likely to affect the

accuracy of phylogenetic inferences (Wagner 2000b). Although the proposition of

22 primary homology statements requires a priori assumptions about what constitutes the quality of same-ness between characters, it does not require a priori assumptions regarding inferences of actual patterns of character evolution. (Those require an independent set of assumptions involving computational and/or optimality criteria in phylogenetic analysis, not discussed here.) The accuracy of all evolutionary inferences in phylogenetic systematics is constrained by what is herein termed the phylogenetic uncertainty principle. The phylogenetic uncertainty principle simply states that the certainty of any output tree topology recovered using phylogenetic methods is fundamentally limited by the inherent uncertainty in the original choice of primary homologies. Primary homologies may be falsified during phylogenetic inference

(rendering phylogenetics an empirically rigorous and testable science), but the underlying character data must be assumed a priori to comprise comparable elements in the first place. The phylogenetic uncertainty principle is an epistemological constraint imposed by the nature of historical data in biology and cannot be circumvented by mathematical or other methodological techniques, although such procedures may provide useful heuristics. The identification of comparable elements in molecular phylogenetics has benefited greatly from the advent of alignment techniques for DNA sequences, but no analogous framework currently exists for morphology. Moreover, uncertainty in phylogeny reconstruction cannot be solved by continuously adding more characters, regardless of quality, to overwhelm the noise to signal ratio and increase clade resolution, particularly in morphologic systematics (Wagner 2000b; Bapst 2012). Ideally, all “errors” discovered among primary homology statements through phylogenetic analysis arise for

23 biologic reasons and not from poor interpretations of morphology. Therefore, it is critical when making phylogenetic inferences that all a priori assumptions of primary homology have a logical, biologic basis supported by empirical observations via character analysis.

Character analysis: Remane’s criteria and developmental biology—Hypotheses of primary homology arise from character analysis, not phylogenetic analysis. These hypotheses are critical to accurately reconstruct evolutionary relationships because the outcome of a phylogenetic analysis is determined by the matrix analyzed (Bryant 1989).

Given their fundamental importance, assessments of primary homology should be critically evaluated and only proposed after careful analysis and argumentation. Remane

(1952) outlined three principles useful for recognizing potentially homologous features among organisms: (1) similarity in position, (2) similarity in structure, and (3) the existence of transitional forms.

Criterion (1) corresponds with Geoffroy Saint-Hilaire’s (1830) “principe des connexions” and refers to similarities in the topological position of a character and its relation to other characters; whereas criterion (2) refers to an “intrinsic” similarity where two or more features match in their structural details and complexity without reference to topological position (Wiley and Lieberman 2001). For a given character, an unambiguous match between criteria (1) and (2) would support the proposition of a primary homology statement. Unfortunately, it is not always clear what constitutes a “match”. Multiple hypotheses of primary homologies can arise when evolution has transformed one structure into another and/or a character has shifted in topological position. What if the evidence supporting alternative interpretations is equivocal? Criterion (3) offers a

24 solution to this dilemma by incorporating information on the transitional forms of fossils and development.

Observations of the embryologic stages of development combined with the geologic succession of fossilized morphologies have long helped guide the recognition of homologous characters (von Baer 1828; Darwin 1859; Hall 2002). Homologous characters need not be similar in structure or position if it can be shown they have common genealogical origins via the existence of transitional forms exhibited in developmental patterns or the fossil record. Moreover, the existence of intermediate forms in fossils and embryos provides a temporal axis to character transformations.

Gilbert and Bolker (2001) pointed out that a significant feature of embryological development is not necessarily the appearance (or disappearance) of individual transient morphologic structures, it is the temporal sequence of changes the embryo undergoes and their underlying genetic mechanisms. In other words, these temporal sequences

(paleontologic or developmental) themselves can provide a logical basis for proposing primary homologies.

Conflictions between homology, terminology, and phylogenetics: examples from the Echinodermata

“I salute the echinoderms as a noble group especially designed to puzzle the zoologist.”

–Libbie H. Hyman (1955: p. vi)

25

A major source of character data in systematic studies come from taxonomic descriptions of morphology. This is particularly true in studies utilizing paleontological data because most fossils are limited to providing only morphologic information about extinct organisms. However, taxonomic descriptions and the terminology employed therein do not express unbiased observations of nature. All descriptive observations, including those in this paper, are colored by theories and expectations (Eldredge and

Gould 1972). Notably, detailed taxonomic descriptions are often entrenched in theoretical considerations of homology and rich in predictions of character evolution.

When theories of homology and character evolution change, the terminology used to name and/or identify a character may or may not. Obviously, problems arise in downstream comparative analyses if the descriptive terminology used to describe a morphologic feature does not reflect its evolutionary history. The existence (and persistence) of a poor descriptive nomenclature in taxonomic literature obfuscates hypotheses of primary homology when building a character matrix and results in specious topologies when such a matrix is analyzed using phylogenetic methods. The examples below taken from echinoderm studies demonstrate that hypotheses of primary homology should not be taken from taxonomic literature uncritically.

Echinoderm homologies and phylogeny—Echinoderms are a phylum of marine organisms represented by more than 7,000 living species (Brusca and Brusca 2003) distributed among five classes: Crinoidea (sea lilies and feather stars), Ophiuroidea

(brittle stars), Asteroidea (sea stars), Echinoidea (urchins and sand dollars), and

26

Holothuroidea (sea cucumbers). The apparent diversity of extant echinoderms masks

their more prodigious geologic history. The half-billion-year echinoderm fossil record is spectacularly complete and reveals approximately 30 clades distributed among 21 taxonomic classes spanning the entire Eon (Sprinkle and Kier 1987).

Moreover, the calcitic endoskeletons of fossil and living echinoderms showcase a bewildering array of disparate morphologies making them ideal for studying large-scale evolutionary patterns (Foote 1992; Sprinkle and Guensburg 1997).

Yet the phylum’s extreme morphologic disparity also presents difficulties for determining homologies between and among clades and obstructs accurate phylogenetic inferences (Paul and Smith 1984; Sumrall 1997). Research assembling a “complete” echinoderm phylogeny has been stymied for decades in part due to the lack of a unified set of morphologic terms representing homologous skeletal structures and much effort has recently been applied to the problem (Mooi et al. 1994; Mooi and David 1997; Mooi et al. 2005; David et al. 2000; Sumrall 2008; 2010; Sumrall and Waters 2012; Zamora et al. 2012; Kammer et al. 2013). For example, Sumrall and Waters (2012) examined thecal plate elements among four clades of fossil blastozoan (i.e., stalked, non-crinoid) echinoderms and discovered that homologous plates in closely related clades often had different names and that some nonhomologous plates had the same name. Egregious terminology is not limited to fossil echinoderms. Mooi and David (1997) pointed out that the five living classes also have major terminological problems that obfuscate common features across clades, such as the name applied to the various expansions of the oral ring

(Hyman 1955). Thus, using the traditional names of plates to construct hypotheses of

27

primary homology would produce spurious topologies in phylogenetic analysis because

these plates are only “homologous” in the sense that they share the same name but do not

share evolutionary origins. Clearly, character analyses and revisions of morphologic terms must accompany efforts to reconstruct echinoderm phylogeny.

Crinoids as a fractal analog of a phylum—Difficulties determining skeletal

homologies are pervasive within as well as between echinoderm clades. Within the

Crinoidea, the 600 or so living species constitute a fractal analog of homology problems

outlined above characteristic of the phylum and offer the opportunity for a smaller-scale case-study detailing a Phylogenetic Paleo-ontogenetic approach to discovering homology

(Simms 1993; Ausich 1996). Below, I resolve contention surrounding primary homologies for posterior plates of “pan-cladid” crinoids by combining data from fossil morphology and developmental patterns in living crinoids. Although conducting a comprehensive phylogenetic analysis on pan-cladid crinoids is beyond the scope of this paper, the analysis presented here provides a logical basis for choosing hypotheses of primary homologies for posterior plate characters in future phylogenetic analyses.

Moreover, it is hoped that the argumentation used herein will serve as general framework for others seeking a logical basis for proposing primary homologies to test phylogenetic hypotheses.

28

Terminological antinomies and transitional homologies: parallels

between crinoid ontogeny and fossilized evolutionary history

“The mystery of this controversy is curiously full of misunderstandings and

misrepresentations.” –Bather (1891: p. 480)

Natural history of the pan-cladid Crinoidea—The Pan-Cladida are a long-lived clade of crinoids spanning the Ordovician Period (~ 485.4 Ma to 443.4 Ma) (Cohen et al.

2013) to the present and include all extant species of Crinoidea as well as fossil forms.

Together, the pan-cladids are the most diverse clade of crinoids and comprise three of the

fived named subclasses: Cladida, Flexibilia, and (Ausich et al. in press; Wright

and Ausich in press). The subclass Cladida is paraphyletic and gave rise to both the

Flexibilia and Articulata (Springer 1920; Simms and Sevastopulo 1993). Thus, the name

“Pan-Cladida” used herein refers to the common ancestor of all species included within

the subclass Cladida and all of its descendants, regardless of taxonomic rank in the

Linnaean hierarchy (Wright and Ausich in press). The flexible crinoids split from the

cladids during the Late Ordovician, diversified, and went extinct at the end of the

Paleozoic . The Articulata, of which all living crinoid species are grouped, originated

from cladid ancestors either during the Late Paleozoic or earliest post-Paleozoic (Simms

and Sevastopulo 1993; Webster and Jell 1999) (Fig. 2.1). Recent phylogenies of living

crinoids indicate a monophyletic Crinoidea, but the of Articulata has been

questioned and awaits further analysis (Rouse et al. 2013; Roux et al. 2013). Because the

29 greatest diversity of pan-cladid crinoids are known only as fossils, tracing the ancestry of living crinoids and discovering their phylogenetic affinities with extinct fossil lineages requires a detailed understanding of crinoid comparative morphology to generate hypotheses of homology.

The skeletal morphology of a crinoid is highly complex and consists of several morphologic modules: the size and shape of the calyx, number and arrangement of posterior plates, arm morphology and branching pattern, and stem (Fig. 2.2). The combinatorial nature of module configurations has given rise to a staggering diversity of pan-cladid morphologies and nearly 1,000 named genera. In Paleozoic cladids, the pentaradial symmetry of the calyx is interrupted by one to three plates located in the posterior CD interray of the cup. These additional, so-called “anal” (sensu Ubaghs 1978) plates, in the posterior region are referred as the radianal, anal X, and right tube plate

(listed from the aboral to oral direction) (Fig. 2.2 B-C). Fossil forms display temporal variation in the number, shape, and positional relations of posterior plates (Moore et al.

1978; Webster and Maples 2006). In extant crinoids, a posterior plate is present in juvenile stages but absent in adults (Clark 1915; Amemiya et al. 2014). Because posterior plating patterns are an important module of morphologic differentiation, they have been extensively used as taxonomically significant characters for delimiting crinoid species and higher taxa (Moore et al., 1978; Webster and Maples 2006). Thus, previous attempts at testing phylogenetic hypotheses and evolutionary patterns among pan-cladids included posterior plate characters despite substantial uncertainty underlying their primary

30

homology statements (Brower 1995; Ausich 1998; Gahn and Kammer 2002; Kammer

2008).

Posterior plates: temporal trends and contentious homologies—Patterns of

posterior plate evolution in fossil pan-cladids have been characterized as exhibiting a

“progressive change toward increased simplicity” (Moore and Laudon 1943: p. 34). The

oldest known pan-cladids have multi-plated posterior interrays (Sprinkle and Wahlman

1994; Guensburg and Sprinkle 2009) and many lineages subsequently exhibit a general

trend to reduce the number of posterior plates in the cup throughout the Paleozoic (Moore

and Teichert 1978). Temporal changes in the number of posterior plates broadly

correspond with concomitant shifts in ecologic abundance and taxonomic diversity

(Webster and Maples 2006). Complex, multi-plated morphologies were the most diverse during the Ordovician and subsequently disappear from the fossil record; whereas crinoids with three posterior plates were most dominant throughout the to

Pennsylvanian (Webster and Maples 2006). Crinoids with a single posterior plate rapidly diversified during the Pennsylvanian and increased in frequency to become the most common morphology during the (Webster and Maples 2006). Older taxonomic literature capture these so-called progressive changes by describing a plate arrangement as having either a “primitive” or “advanced” condition, where primitive refers to stratigraphically older multi-plated morphologies and advanced refers to younger two or single-plated forms (Moore et al. 1978).

Homology schemes for posterior plates among fossil lineages and between fossil and living crinoids have been debated, somewhat inimically, for more than a century (P.

31

H. Carpenter 1882; Wachsmuth and Springer 1879; Bather 1980; 1891; 1918; Clark

1915; Mortensen 1920; Springer 1920; Ubaghs 1953; Moore 1962; Phillip 1964; Moore and Teichert 1978; Webster and Maples 2006). The overwhelming majority of named fossil genera have three posterior plates in the cup, including putative Paleozoic ancestors of living crinoids and fossil flexibles (Webster and Jell 1999; Webster and Maples 2006).

Using Remane’s (1) criterion described above, proposing primary plate homologies for crinoids with three posterior plates is straightforward. The radianal is the most proximal posterior plate to (and always in contact with) the C radial, typically occupying a position beneath or to the left of the C radial. The anal X is interradial in position and may be in lateral contact with the radianal, C radial, BC and CD basals, or the CD , and occupies a position above and/or to the left of the radianal (Ubaghs 1978). The right-tube plate rests either above the radianal or both the radianal and anal X and typically provides support for other plates in the anal sac (Fig. 2.2). Where only two plates are present in the cup, Remane’s (1) criterion can once again be used to propose primary homologies for the two remaining more proximal plates: the radianal and anal X (Moore and Teichert

1978). However, any further reduction in the number of posterior plates renders

Remane’s (1) criterion inapplicable because the single posterior plate does not occupy a position more similar to either the radianal or anal X where two or more posterior plates are present. In other words, one of the posterior plates migrated from its ancestral position and the other is absent.

An antinomy can be characterized as a kind of paradox that describes two equally compelling but mutually incompatible explanations and is a useful term to describe the

32 contention that arises when only a single posterior plate is in the cup. Choosing whether a single plate in the posterior interradius is the radianal or anal X presents an antinomy of alternative homology schemes (Fig. 2.3). If there is only one posterior plate in a fossil pan-cladid, is it the radianal or anal X? Is this fossilized single posterior plate homologous with the single plate present in the juvenile stages of extant crinoids? If so, which posterior plate is it? Unfortunately, Remane’s (2) criterion cannot help because the shape of a single posterior plate is constrained to accommodate changes in the size and shape of the calyx and therefore does not retain the shape of either the radianal or anal X when alone in the cup.

The problem is further confounded by a history of problematic terminology favoring plate topologies over plate homologies. Remarkably, a perusal of the taxonomic literature of fossil crinoids reveals the presence of a single posterior plate in a fossil cladid is frequently termed an anal X in taxonomic descriptions and figured specimens even when an author considered the plate homologous to the ancestral radianal or the evidence equivocal (cf. Moore and Laudon 1943; Ubaghs 1953; 1978; Moore 1962;

Philip 1964). Kirk’s (1944: p. 234) description of the single posterior plate in the

Mississippian genus Cymbiocrinus epitomizes this dubious practice: “it is doubtful if this plate is homologous to [the] anal X, but we may so denominate it for convenience”.

These misnomers obfuscate any notion of evolutionary continuity among characters. It is not surprising there has been much confusion given that different sections of the crinoid

Treatise of Paleontology (Moore and Teichert 1978) disagree with one another with respect to posterior plate homologies and terms for fossil and living pan-

33

cladids (cf. Breimer 1978, Brower 1978, Moore et al. 1978; Strimple 1978; and Ubaghs

1978). Thus, crinoid paleobiologists should not take Treatise descriptions of posterior

plate characters at face value when proposing primary homologies to make phylogenetic

inferences.

Potentially more confusing for a phylogeneticist is the terminological scheme

proposed by Webster and Maples (2006). In an attempt to rectify terminological

misnomers, Webster and Maples (2006) proposed to abolish all implications of homology

from morphologic nomenclature by renaming the posterior plates the primanal,

secundanal, and tertanal. Under this terminology, the primanal is always the most

proximal plate in the cup, regardless of whether it is the radianal or anal X. Thus,

Webster and Maples (2006) “solved” the epistemological dilemma of proposing

homology statements by avoiding them altogether. This so-called solution is problematic for two reasons. First, because the terminology of Webster and Maples (2006) is based purely on topology without reference to any ontological theory of homology, it is useless for proposing primary homologies to test phylogenetic hypotheses. Second, the proposed terms are already in use to describe an unrelated, non-homologous set of plates in a phylogenetically distant clade of crinoids, the subclass (Moore and Teichert

1978; Ausich 1998). For these reasons, Webster and Maples’ (2006) terminology should not be followed. Instead, we should confront the problem directly by seeking a logical basis for choosing between alternative primary homology schemes. This approach has the advantage of combining posterior plate characters with other modular complexes of crinoid morphology in a phylogenetic analysis. Thus, characters can be tested against one

34 another when inferring phylogenetic relationships and any parallel instances of plate reduction (or addition) can be determined empirically.

In the next two sections, I apply Remane’s (3) criterion to examine potential transitional forms in two independent sources of data: the ontogeny of extant crinoids and paleontologic studies on individual lineages. The recognition of intermediate morphologies present in developmental patterns of extant representatives and/or high- resolution paleontologic sequences may provide paleo-ontogenetic evidence supporting one set of primary homologies over another and dissolve antinomies arising from examining comparative morphology alone.

Developmental patterns in living crinoids—Numerous observations of embryologic stages in extant crinoids have been described, particularly for the stalkless comatulid crinoids (Thomson, 1865; Carpenter 1866; Clark 1915; Springer 1920;

Mortensen 1920; Mladenov and Chia 1983; Lahaye and Jangoux 1987; Shibata et al.

2008). Developmental patterns in living species of stalked crinoids were largely unknown until recently (Nakano et al. 2003; Amemiya et al. 2014). Because stalked crinoids and comatulids share many similarities with respect to the skeletal development of the calyx and posterior plates, the distinction between the two adult forms is not pertinent for the purposes of this paper.

Kammer (2008) lucidly described common themes of crinoid development, from which I base the following conspectus. Crinoid ontogeny consists of five successive life stages: the embryo, doliolaria, cystidean, pentacrinoid, and (in comatulids) the comatulid stage (Mladenov and Chia 1983; Lahaye and Jangoux 1987). The doliolaria is an

35

endotrophic, free-swimming larva that emerges from the embryonic membrane. Once the

doliolaria settles on a suitable substrate, it metamorphoses into the stalked cystidean

stage. The skeleton of a cystidean stage consists of paired, interradially-oriented basals and primary peristomial cover plates (i.e., oral plates, see Kammer et al. 2013), a terminal stem plate, and a few columnals. Infrabasal plates, common in fossil pan-cladids, do not usually develop in living articulates but are known to occur in some species (Rasmussen

1978). Next, more skeletal plates are added including an “anal” plate, radials, and numerous additional plates relating to the construction of the arms and pinnules initiating the exotrophic pentacrinoid stage. Comatulids become adults when they excise the stem; whereas stalked crinoids grow into larger “adult” pentacrinoids. Although variation in the details of crinoid ontogeny exists, all crinoid species exhibit the general growth sequence and patterns of skeletal plate addition outlined above.

This single posterior plate was originally called an “anal” plate in living species because it was first discovered in the pentacrinoid stage of comatulids to directly overlie the BC basal between the radials, and therefore assumed on the basis of topological position to be homologous with the anal X in fossil crinoids (Thomson 1865; Carpenter

1866; Bather 1890). However, development is not a static process and characters cannot be accurately traced by giving special consideration to a single ontogenetic stage.

Skeletal plates within the calyx grow at different rates and change in size and shape over the course of ontogeny (Lahaye and Jangoux 1987). In addition, the topological arrangement of plates may vary in different growth stages. These changes must be accounted for when tracing parallels between development and evolution. Comparisons

36 between a single ontogenetic stage and adult morphology can lead to specious proposals of primary homology because the developmental origin of a character may not correspond to its position during a later time in ontogeny. Instead, the entire sequence of changes must be considered when connecting the developmental origin of a morphologic character with its ontogenetic history and ultimate fate in the adult.

The developmental origin, migration, and eventual resorption of the single posterior plate in juvenile crinoids have been well characterized throughout the succession of ontogenetic stages by Clark (1915), Springer (1920), and Lahaye and

Jangoux (1987) (Fig. 2.4). Detailed observations by these authors demonstrate that the development of the posterior “anal” plate originates in a radial position (with respect to the basals and primary peristomial cover plates) prior to the radials during the cystidean stage, and occurs within the same as the C radial. When the radial plates begin to grow, the larger posterior plate occupies a position beneath and to the left of the C radial, maintaining a close affinity with the developing gut tract (Springer 1920). As the radials grow larger, the posterior plate occupies a position on the right- side of the CD interray and is accommodated within a concavity in the C radial plate. Eventually the radials push the posterior plate into an inter-radial position (i.e., the CD interray) where it supports a lappet that protects the developing anal cone (Lahaye and Jangoux 1987). The posterior plate subsequently migrates out of the cup and is resorbed once the anal cone has formed. In summary, the posterior plate originates in a radial position, maintains lateral continuity with the C radial, and later moves into the CD interray and out of the cup. In fossil pan-cladids with multiple posterior plates, it is the radianal that has

37

affinities with the C radial and maintains lateral contact with it; whereas the anal X is an

interradial plate (Ubaghs 1978). Thus, the evidence from crinoid ontogeny indicates that

the single, prominent posterior plate found in living crinoids is homologous with the

radianal in fossil pan-cladids, not the anal X.

If the single posterior plate in the juvenile stages of living species is the radianal,

is there any ontogenetic evidence for the existence of an anal X? Recall that the anal X is

an interradial plate within the CD interray. Interradial plates are known to occur in

several species of living comatulids and appear late in development during the

pentacrinoid stage (Clark 1915; Breimer 1978). They either occur in all five interrays

with the same degree of development or they are absent entirely (Breimer 1978). The

early-mid pentacrinoid stage of Comactinia meridionalis displays posterior plating

strikingly similar to the arrangement in many fossil cladid and flexible species, with the

radianal side by side with the additional posterior plate (Springer 1920, plate B, fig. 5a).

Later, the interradials are resorbed along with the radianal leading to their absence in the

adult phenotype (Springer 1920; Breimer 1978). If the posterior interradial is homologous

with the anal X in fossil crinoids, than the right tube plate (and any other posterior plates)

in fossil pan-cladids may have developed from additional posterior interray plates. In this scenario, the interradial plates outside the CD interray either did not develop or were resorbed prior to adulthood, as they are unknown in fossil species. Alternatively, the

posterior interradial plate may not be homologous with the anal X. The insertion of

interradial plates may instead be a novel feature present among a subset of living

comatulids. Interradial plates are presently unknown in juvenile stalked crinoids, but this

38

may simply reflect the paucity of developmental studies on stalked crinoids (Amemiya et

al. 2014).

These observations have significant implications for ascertaining evidence-based primary homology statements if one is willing to assume geology’s useful aphorism regarding the present as the key to the past: the morphologic transitions unfolding during crinoid ontogeny are developmental relics of their Paleozoic precursors. The migration pathway of the radianal during crinoid development (viz, originating radially in the C ray followed by movement to the posterior interradius and out of the cup) parallels temporal trends in the paleontologic succession of posterior plate morphologies (Fig. 2.4).

Therefore, a uniformitarian perspective supports the hypothesis that the radianal plate, not the anal X, is homologous with the single posterior plate in Paleozoic pan-cladids.

Thus, the single posterior plate in the cup of a fossil pan-cladid should be coded as a primary homolog of the radianal plate occurring in crinoids with multi-plated posterior interrays. The implications of this hypothesis are not conditional on the monophyly of the

Articulata because the pageant of ancestral morphologies echoed in crinoid development empirically demonstrates that the single posterior plate in at least one lineage is equivalent to the radianal, not the anal X. Given the similarities in posterior plate development among extant crinoids, a polyphyletic Articulata would only strengthen the argument above because it would suggest that the same pattern occurred independently among multiple Paleozoic ancestors.

Phyletic evolution in Paleozoic pan-cladids: a test of relative frequencies—The argumentation above concerning posterior plate homologies rests upon the assumption

39

that patterns of character change exhibited by extant crinoids can be extended to apply to

all pan-cladid lineages known only as fossils. When conducting a phylogenetic analysis

incorporating extinct pan-cladid lineages, is such an extrapolation from living

representatives justified? Large-scale temporal trends in evolution need not be invariant

among lineages and do not necessarily correspond to phylogenetic trends. Individual

clades may exhibit idiosyncratic trends to reduce (and/or subsequently add) posterior

plates iteratively over time. Nevertheless, character coding decisions must be made if one

seeks to conduct a phylogenetic analysis. Moreover, the question is not whether or not

extant crinoids are closely related to Paleozoic pan-cladids. All evidence indicates they

are closely related and share an overlapping distribution of taxonomically significant

traits (Moore and Teichert 1978; Roux et al. 2013). In the absence of any additional

information, basing primary homologies on ontogenetic sequence data present in living

representatives is justified because empirical science must proceed in the direction of

available evidence. However, one is faced with a seemingly impossible problem of

assessing how often such a hypothesis is objectively accurate.

Null hypotheses in phylogenetics are constructed by making a priori assumptions

of homology among comparable characters and using Hennig’s (1966) auxiliary

principle. What if a mistake is made in determining what constitutes a “comparable

character” (or character state) in the first place? In other words, if the presence of a single

posterior plate is determined to be equivalent to a radianal in pan-cladids during character analysis (H0 = a single posterior plates is homologous with the radianal), how often has a researcher committed a type 2 error? Although such mistakes may, in principle, be

40

unknowable; it is humbling to consider that a type 2 error during character analysis may

result in type 1 errors during phylogenetic analysis and obstruct the recognition of “true”

(i.e., empirical) evolutionary homologies and patterns of character evolution.

Central tendencies in historical science must be determined by examining relative

frequencies (Gould 1989). Although single or isolated occurrences of a phenomenon may

be interesting and/or otherwise worthy of study, they do not provide insight into building

general expectations of a theory. Luckily, the pan-cladid fossil record supplies several

key examples relevant to assessing the relative frequency in which the fossil record supports or refutes available evidence from ontogeny. If examples of morphologic transitions from the fossil record predominantly corroborate the ontogeny-based solution to the posterior plate antinomy, then one can be more confident in the results of a phylogenetic analysis assuming those primary homologies to reconstruct evolutionary relationships and patterns of character evolution.

Although the fossil record of crinoid genera is well-sampled (Foote and Raup

1996), many pan-cladid species are based on only one or a few specimens (Webster and

Maples 2006). All macroevolutionary studies must consider the species-level, even when

using higher taxa as proxies, because species are the fundamental units that preserve

phenotypic change among populations (Hendricks et al. 2014). Unfortunately, there is a

dearth of pan-cladid species-level phylogenies available (Kammer and Ausich 2007;

Gahn and Kammer 2012), and none contain a set of species relevant to the problem

addressed here. Moreover, conducting such a species-level analysis itself may require

addressing the antinomies this paper is attempting to resolve: is the single posterior plate

41

homologous with the radianal or anal X? A comparison of fossilized ontogenetic

sequences among pan-cladids might be helpful, but there are currently no known fossils preserving the larval and early developmental stages of posterior plates. At this point, an interesting light may be thrown on the problem by examining variations in posterior plate

conditions among species with transitional morphologies. The collection of a large

number of specimens allows one to examine a distribution of transitional morphologies

and make comparisons combining all three of Remane’s (1952) criteria.

Webster and Maples (2006) noted numerous instances where intraspecific

variations in posterior plate conditions occur when a large number of specimens were

collected. These examples are particularly informative because they come from

paleontologically well-sampled stratigraphic sections representing short time intervals

(Webster and Maples 2006). Thus, it is likely they represent paleontologic sampling taking place over short enough time scales to sample transitional morphologies. Any bed- scale time averaging and/or taphonomic discrepancies among first and last appearances between stratigraphic sections do not affect the inferences obtained herein because it is the overall distribution of sampled morphologies, not their sampled paleontologic sequence, that are used to make comparisons and examine intermediate forms.

For example, Wanner (1916) recognized the presence of a single, large interradial posterior plate in his original description of the Permian pan-cladid Hydreionocrinus variabilis (= Cadocrinus variabilis), but called this plate the radianal rather than anal X because in some specimens the position of the plate was lying at an angle to the CD basal and in a more proximal position with the C radial. Wanner (1916) noted that among the

42

197 specimens in his collection, a subset possessed proximal tips of one to two small

plates above the radial summit, which he termed the anal X and right tube plate. Thus,

Wanner (1916) concluded that the large posterior plate must actually be a large radianal

that migrated to an interradial position, supporting the anal X and right tube plate above.

Webster and Maples (2006) suggested that if Wanner had a smaller sample size typical of

fossil pan-cladid species, he would likely have committed a misnomer by calling it an

anal X. Given the inconsistent treatment in the Treatise and the standard portrayal of

posterior plates in crinoid plate diagrams (Moore and Teichert 1978), it is probable that

any crinoid worker since the 1800s would have nearly committed the same misnomer.

Similarly, Wright (1920; 1926; 1927) examined posterior plate conditions in a

large number of Upper Paleozoic specimens of Eupachycrinus calyx (n = 1,000),

Zeacrinus? konincki (n = 342), Ulocrinus globosus (n = 480), and Hydreionocrinus sp.

(n = 130) (= Phanocrinus calyx, Parazeacrinites konicki, and Ureocrinus globosus

respectively). Wright (1926, p. 149) noted that “extreme” variations in posterior plates

are apparent in these species when samples become sufficiently large. With such a large

sample size displaying a semi-continuous distribution of intermediate posterior plate conditions, Wright (1920; 1926; 1927) applied the logic of Remane’s criteria to the distribution of transitional morphologies. In all cases, the radianal can be seen to have increased in size and subsequently moved into an interradial position, pushing the other posterior plates out of the cup (e.g., Wright 1926, figs. 1-59). Intermediate morphologies

display small remnants of an anal X and right tube plate above the interradially

positioned radianal plate. Thus, the paleontologic evidence indicates that the posterior

43

plate most proximal to the C ray is the last to migrate from the cup in a phyletic

sequence. Another example supporting the conclusions of Wanner (1916) and Wright

(1926) was discovered by Webster and Lane (1967) for Arroyocrinus popenoei (n = 83).

The examples above should not be interpreted as suggesting that all (or most) species of pan-cladids exhibit variation in posterior plate position and arrangement when large samples are collected. Webster and Lane (1967) also examined Moapacrinus rotundus (n = 137) and Erisocrinus longwelli (n = 59) and found no variation in the posterior plating in either species, therefore supporting the conclusion that the above examples represent special cases where an exceptionally abundant number of specimens in an evolving lineage were sampled over a short temporal duration. Further work is needed to address whether any of the observed posterior plate variations corresponded with other characters and/or were related to speciation.

Webster and Maples (2006) compiled a reference list for the number and arrangement of posterior plates for the type species of 378 pan-cladid genera from which relative frequency of the radianal as the single posterior plate may be inferred. Their dataset includes 152 genera with three posterior plates and therefore can be unambiguously homologized between species using procedures outlined earlier in this paper. Of these, 96% have the radianal “moderately or mostly underlying the right side” of the anal X (Webster and Maples 2006: p. 200). This condition was also found to be the most common arrangement for pan-cladids with two plates in the posterior interray

(Webster and Maples 2006). These paleontological observations support the hypothesis that the single posterior plate in fossil pan-cladids is predominantly the radianal, not the

44

anal X. Therefore, the relative frequency of morphologic transitions recovered from the

fossil record overwhelmingly (≥ 96%) agrees with and corroborates the ontogenetic evidence provided by extant crinoids. This discovery is significant because it overturns more than 120 years of the status quo in crinoid paleontology.

Of course, it is possible that a small number of lineages with no living

representatives lost the radianal plate yet retained the anal X. Both Kammer and Ausich

(1996) and Webster and Maples (2006) suggested that this may have happened among

species of the Mississippian genus Barycrinus. Some species of Barycrinus possess small radianal plates below and to the right of the much larger anal X, where others have only a single, large visible plate interpreted as an anal X. However, as noted by Gahn and

Kammer (2002), it is possible that the anal X has overgrown the radianal in these species and is lamentably invisible on the calyx exterior. Although this condition is presently unknown in Barycrinus, it is known to occur among other Late Paleozoic pan-cladids.

For example, the Pennsylvanian pan-cladid Perimestocrinus calyculus (= Vertigocrinus calyculus) described by Moore and Plummer (1940) has only two plates visible on the exterior surface of the calyx. However, three posterior plates are visible when examined from inside of the cup. When viewed from the inside of the calyx, these plates are readily interpreted as the radianal, anal X, and right tube plate occupying an otherwise ordinary arrangement (Moore and Plummer 1940). Without examining the calyx interior, one may have inferred from topology that the radianal was absent and/or incorrectly coded the exterior plates in a character matrix. Additional taxa with cryptic plates within the calyx interior include species placed in the Pennsylvanian crinoid genera Arkacrinus,

45

Paradelocrinus, Plaxocrinus, Vertigocrinus, and the Mississippian to Permian genus

Erisocrinus (Strimple, 1978; Webster and Maples 2006). Based on these considerations, further collection, preparation, and careful inspection of aberrant fossil specimens is necessary to better understand morphologies that fall outside of general expectations (see

Rozhnov and Mirantsev 2014).

Crinoid ontogeny and phylogeny: implications for evolutionary studies

“Ontogeny does not recapitulate phylogeny: it creates it” –Walter Garstang (1921: p. 82)

The existence of parallels between ontogeny and phylogeny is one of the most pervasive and influential concepts in evolutionary biology and relate directly to the discovery of homologous characters (Gould 1977). Evolutionary developmental biology

(EDB) has recently refurbished von Baer’s law to study parallels between evolution and development (Abzhanov 2013). Von Baer’s law states that more generalized characters appear earlier in ontogeny than specialized characters, with specialized characters developing from generalized characters (Gould 1977). For example, ontogenetic sequences for species within a clade may share generalized developmental features, but subclades may share additional more specialized features that reflect more recent evolutionary changes not shared with other subclades. Because developmental programs in the gene regulatory network (GRN) have phylogenetic memory, they retain aspects of

46

phylogenetic history and may “recapitulate” sensu von Baer 1828 (non Haeckel 1866) ancestral morphologies in the juvenile forms of descendants (Abzhanov 2013). Parallels between evolution and development predict a significant degree of recapitulation between cladistically significant traits nested at different phylogenetic levels (Abzhanov 2013).

Kammer (2008) suggested that the appearance and position of plates during crinoid development are likely controlled by genetic switches, such as Hox genes, that regulate proximal-distal morphogenesis and other aspects of skeletal development.

Although developmental genetic studies on living crinoids have discovered the expression of Hox genes in the earliest developmental stages of the sea lily rotundus (Hara et al. 2006), it is for the moment an understatement to suggest that much work needs to be done to discover how these genes interact with other aspects of the

GRN (such as cis-regulatory elements) to direct downstream development and morphogenesis in crinoids. Nevertheless, the near uniform predictability in the developmental timing of skeletal elements among extant comatulid and stalked species suggests they are controlled by highly conserved developmental modules.

Such evo-devo perspectives serve to strengthen the paleo-ontogenetic character analysis above. In living crinoids, the radianal appears much earlier in ontogeny than other additional plates in the posterior interray (i.e., as in Comactinia meridionalis) and exhibits the same degree of developmental canalization as other calyx plates (Lahaye and

Jangoux 1987). Parallels between crinoid development and fossilized morphology, as interpreted using von Baer’s refurbished law in EDB, indicate that living crinoids recapitulate the morphologies of Paleozoic ancestors because they inherited a common

47 pathway of development (Abzhanov 2013). Thus, the regulatory machinery in the GRN controlling posterior plate development may not have substantively changed since the

Late Paleozoic. This result corroborates studies by Foote (1995; 1999) examining morphologic disparity in Paleozoic and post-Paleozoic crinoids. Foote (1995) found that pan-cladids steadily increase in disparity throughout the Paleozoic despite considerable volatility in generic richness. However, pan-cladid disparity dropped after the end-

Permian extinction and post-Paleozoic forms do not broadly overlap with their Paleozoic representatives in morphospace (Foote 1999). Differences in both overall disparity and morphospace occupation led Foote (1999) to conclude that genetic and/or developmental constraints may have been responsible for substantive differences between Paleozoic and post-Paleozoic crinoids. Given that posterior plate characters show considerable variation throughout the Paleozoic and are correlated with other evolutionary changes occurring in the size and shape of the calyx (Moore et al. 1978; Webster and Maples 2006), it is likely that developmental modules within the GRN of the lineage(s) that survived the end-

Permian extinction became rigidly constrained. Such constraints would limit developmental variation in calyx design and decrease the propensity for lineages to expand into unoccupied regions of crinoid morphospace.

The notion of deep homology provides a potential explanation for putative parallel (i.e., homoplasious) trends in posterior plate characters occurring in different lineages of Late Paleozoic cladids (Moore et al. 1978; Webster and Maples 2006). Given the numerous paleontologic species that display a net reduction in the number of posterior plates, it is possible that different lineages lost posterior plates independently

48

through instances of . Parallel morphologic evolution can occur among

distantly related lineages arising from selection acting on variation within shared

developmental toolkits (Hall 2003). Because most Paleozoic pan-cladids have three

posterior plates, the tendency to evolve less complex morphologies via plate reduction is

interpreted as a paedomorphic trend reflecting arrested development or progenesis

(Kammer 2008). Precocious maturation can result from an adaptive response to pressures

for small body size and/or as an r selection strategy for rapid growth rates for taxa

inhabiting unstable environments (Gould 1977). All putative ancestors of extant crinoids

are members of the Late Paleozoic Crinoid Macroevolutionary Fauna (LPCMF) and

broadly overlap in niche space (Kammer and Ausich 1987; Ausich et al. 1994). The

transition to a cladid-dominated LPCMF was concomitant with considerable

environmental changes, such as an increase in the abundance and distribution of

siliciclastic habitats (relative to carbonate platforms) preferred by Late Paleozoic cladids

(Kammer and Ausich 2006). This increase in siliciclastic environments is related to

orogenic activity and an increase in the frequency of sediment disturbance (Walker et al.

2002). Kammer (2008) noted that pan-cladids in unstable environments commonly have

smaller body sizes. Thus, decreased environmental stability in substrate conditions may

have led to an increase in frequency of pan-cladids with rapid growth rates and small

body sizes exhibiting paedomorphic morphologies. Given that Late Paleozoic pan-cladids likely shared many developmental modules at “deep” genetic levels, different species facing similar selection pressures may have independently evolved paedomorphic morphologies as a response to similar environmental changes. Such selection could arise

49

either through adaptive evolution within populations or at the species level if posterior

plate characters are strongly correlated with diversification rates (Rabosky and McCune

2010). Thus, a combination of selective trends and developmental constraints are

hypothesized herein to have produced parallel evolution of paedomorphic morphologies

in different lineages of Late Paleozoic pan-cladids. Further exploration of these observations await the construction of model phylogenies to test the relationship between the evolutionary acquisition and environmental context of paedomorphic morphologies in

Late Paleozoic pan-cladid crinoids.

The discussion above suggests homoplasy may be common for posterior plate characters in pan-cladid crinoids. An a priori belief of homoplasy is not grounds for excluding posterior plate characters from future phylogenetic analyses because not all instances of homoplasy are artifactual results of poor character state interpretations

(Wagner 2000b). Character reversals and instances of parallel evolution are real events in life’s history and must be mapped onto a model phylogeny to gain insight into historical patterns of character evolution. Moreover, numerous other morphologic characters are correlated with changes in posterior plates including the presence of muscular articulations, zyzygial sutures, structure and branching of the arms, and the development of pinnules (Kammer 2008; Webster and Maples 2008). Moreover, if one were to a priori disregard from phylogenetic analysis any character with the propensity for homoplasy, there would be no characters left to study or much point to conduct a quantitative phylogenetic analysis because such a proposal assumes one already knows the “true” phylogeny and is merely cutting out “noise” among characters to generate a

50

more well-resolved and/or well-supported tree. To be clear, I am not advocating that

homoplasious characters are desirable in phylogenetic analysis. I am only advocating that

features revealed to be comparable elements through paleo-ontogenetic character analysis, as described in this paper, constitute ‘real’, empirical data that need to be rigorously tested against other characters in a phylogenetic context. Given the finite set of morphologic characters available and the important role of posterior plate characters in pan-cladid morphology, I suggest researchers employ a limited “prior belief” in the available paleo-ontogenetic evidence and therefore propose that future phylogenetic analyses consider the single posterior plate in fossil pan-cladids as a primary homolog with the radianal in multi-plated taxa.

Phylogenetic Paleo-Ontogeny: a multidisciplinary approach to

discovering homology

Paleobiology, phylogenetic systematics, and EDB play complementary roles in

evolutionary biology because they provide the theoretical basis and basic historical data

for understanding patterns and processes of macroevolution. Phylogenetic Paleo-

ontogeny constitutes a total evidence approach to homology recognition by unifying

these seemingly disparate fields to propose, test, and empirically “discover” homologous

characters in fossil taxa. A case study in homologizing posterior plates among pan-cladid crinoids was presented to illustrate how a research program in Phylogenetic Paleo- ontogeny can provide insight into how developmental and fossil data can be combined to

51

dissolve morphologic antinomies and ameliorate terminological difficulties obstructing

clarity in homology schemes. It is hoped that this paper will serve as a template for other

researchers seeking an ontological basis and epistemological framework for discovering

homologies of fossil and living species.

For example, within the Echinodermata two models of character analysis have been proposed to resolve homologies among classes: Extraxial-Axial theory (EAT)

(Mooi et al. 1994) and Universal Element Homology (UEH) (Sumrall 2010). The EAT model uses embryological and ontogenetic criteria and designates different kinds of skeletal regions in the echinoderm body as homologous based on developmental and growth patterns; whereas the UEH model attempts to identify individual skeletal plates across clades without reference to regional skeletal patterns. Interestingly, EAT and UEH make different predictions regarding the branching of major clades in echinoderm phylogeny (cf. David et al. 2000; Kammer et al. 2014) and are sometimes contrasted as alternatives in the literature (Zamora and Rahman 2014). Under the umbrella of

Phylogenetic Paleo-ontogeny, these two competing approaches to echinoderm homology become unified. Data supporting EAT and UEH are not mutually exclusive and likely reflect phylogenetic information nested within different hierarchical levels of body plan organization. Paleo-ontogenetic hypotheses for both skeletal regions and individual elemental homologies can be combined into a set of primary homology statements and tested against one another in a future phylogenetic analysis of fossil and living

Echinodermata using all available evidence: , ontogeny, genes, and morphology.

52

I do not propose that combining paleo-ontogenetic character analysis with

phylogenetic systematics leads to a more “objective” framework for discovering

homology. All propositions of primary homology are necessarily subjective. Indeed,

there is no such thing as an assumption free phylogenetic analysis. In science, theories

explaining the natural world arise from hypotheses that have survived repeated subjection

to a battery of rigorous testing and empirical corroboration. Should we not subject our a

priori assumptions behind such “testing” to even greater scrutiny? Phylogenetic

empiricism is the basis of evolutionary inference in (Wiley and

Lieberman 2011). Phylogenetic information is necessarily obscured when a theoretical

concept (such as homology) is empirically approximated (phylogenetic analysis), yet how

else should science proceed towards asymptotically ascertaining “truth” if not by

successive approximation? A research program in Phylogenetic Paleo-ontogeny may help pave the way towards such a future.

53

References

Abzhanov, A. 2013. von Baer’s law for the ages: lost and found principles of

developmental evolution. Trends in Genetics 29:712-722.

Alroy, J. 2010. Geographical, environmental and intrinsic biotic controls on Phanerozoic

marine diversification. Palaeontology 53:1211-1235.

Amemiya, S. A. Omori, T. Tsurugaya, T. Hibino, M. Yamaguchi, R. Kuraishi, M.

Kiyomoto, and T. Minokawa. 2014. Early stalked stages in ontogeny of the living

isocrinid sea lily . Acta Zoologica doi: 10.1111/azo.12109.

Arnone, M. I. and D. H. Davidson. 1997. The hardwiring of development: organization

and function of genomic regulatory systems. Development 124:1851-1864.

Ausich, W. I. 1996. Crinoid plate circlet homologies. Journal of Paleontology 70:955-

964.

—1998. Phylogeny of Arenig to Caradoc crinoids (Phylum Echinodermata) and

suprageneric classification of the Crinoidea. University of Kansas Paleontological

Contributions 9:1-36.

Ausich, W. I., T. W. Kammer, and T. K. Baumiller. 1994. Demise of the Middle

Paleozoic Crinoid Fauna: a single or rapid faunal turnover?

Paleobiology 20:345-361.

Ausich, W. I., T. W. Kammer, D. F. Wright , S. R. Cole, E. Rhenberg, and M. Peter.

2015. Phylogenetic classification of the Crinoidea (Echinodermata). Cuadernos

del Museo Geominero.

54

Bapst, D. W. 2012. When can clades be potentially resolved with morphology? PLoS

One e62312. doi:10.1371/journal.pone.0062312

—2014. Assessing the effect of time-scaling on phylogeny-based analyses in the fossil

record. Paleobiology 40:331-351.

Bateson, W. 1894. Materials for the study of variation. MacMillan, London.

Bather, F. A. 1890. British fossil crinoids: The classification of the Inadunata Fistulata.

Annals and Magazine of Natural History 5:373-788.

—1891. Some alleged cases of misrepresentation. Annals and Magazine of Natural

History 6:480-489.

—1918. The homologies of the anal plate in Atedon. Annals and Magazine of Natural

History 9:294-302.

Bolker, J. A. and R. A. Raff. 1996. Developmental genetics and traditional homology.

Bioessays 18:489-491.

Breimer, A. 1978. General morphology, recent crinoids. Pp. T9-T58 in R. C. Moore and

Teichert, C. (eds.), Treatise on Invertebrate Paleontology. Part T, Echinodermata

2. Geological Society of America and University of Kansas Press, Lawrence.

Brigandt, I. 2003. Homology in comparative, molecular, and evolutionary developmental

biology: the radiation of a concept. Journal of Experimental 299b:9-17.

Brower, J. C. 1978. Camerates. Pp. T244-T263 in R. C. Moore and Teichert, C. (eds.),

Treatise on Invertebrate Paleontology. Part T, Echinodermata 2. Geological

Society of America and University of Kansas Press, Lawrence.

55

—1995. Dendrocrinid crinoids from the Ordovician of northern and southern

Minnesota. Journal of Paleontology 69:939-960.

Brusca, R. C. and G. J. Brusca. 2003. Invertebrates (2nd edition). Sinauer Associates,

Sunderland.

Bryant, H. N. 1989. An evaluation of and character analyses as hypothetico-

deductive procedures and the consequences for character weighting. Systematic

Biology 38:314-227.

Carpenter, P. H. 1882. On the relations of Hybocrinus, Baerocrinus, and Hybocystites.

Quarterly Journal of the Geological Society of London 38:298-212.

—1884. Report on the Crinoidea—the stalked crinoids. Report on the scientific results of

the H. M. S. Challenger. Zoology 11:1-440.

Carpenter, W. B. 1866. Researches on the structure, physiology and development of

Antedon (Comatula) rosaceus. Philosophical Transactions of the Royal Society of

London 156:671-756.

Carroll, S. B. 2008. Evo-Devo and an expanding evolutionary synthesis: a genetic theory

of morphological evolution. Cell 134:25-36.

Clark, A. H. 1915. A monograph of the existing crinoids. Bulletin of the United States

National Museum 32, part 2.

Cohen, K. M., S. C. Finney, P. L. Gibbard, and J. X. Fan. 2013. The ICS International

Chronostratigraphic Chart. Episodes, 36:199-204.

Darwin, C. 1859. by means of . Murray,

London.

56

David, B. B. Lefebvre, R. Mooi, and R. L. Parsley. 2000. Are homalozoans echinoderms?

An answer from extraxial-axial theory. Paleobiology 26:529-555.

Erwin, D. E. 2007. Disparity: morphological pattern and development. Palaeontology

50:57-73.

Erwin, D. E. and E. H. Davidson 2009. The evolution of hierarchical gene regulatory

networks. Nature Reviews Genetics 10:141-148. de Pinna, M. G. G., 1991. Concepts and tests of homology in the cladistics paradigm.

Cladistics 7:387-394.

Eldredge, N. and S. J. Gould. 1972. Punctuated equilibria: an alternative to phyletic

gradualism. Pp. 82-115 in T. J. M. Schopf (ed.), Models in Paleobiology.

Freeman, Cooper and Company, San Fransisco.

Foote, M. 1992. Paleozoic record of morphological diversity in blastozoan echinoderms.

Proceedings of the National Academy of Sciences 89:7325-7329.

—1995. Morphological diversification of Paleozoic crinoids. Paleobiology 21:273-299.

—1999. Morphological diversity in the evolutionary radiation of Paleozoic and post-

Paleozoic crinoids. Paleobiology 25:1-115.

Foote, M. and D. M. Raup. 1996. Fossil preservation and the stratigraphic ranges of taxa.

Paleobiology, 22:121-140.

Gahn, F. J. and T. W. Kammer. 2002. The cladid crinoid Barycrinus from the Burlington

Limestone (early Osagean) and the phylogenetics of Mississippian botryocrinids.

Journal of Paleontology 76:123-133.

57

Garstang, W. 1921. The theory of Recapitulation: a critical re-statement of the biogenic

law. Journal of the Linnaean Society of London 35:81-101.

Gilbert, S. F. and J. A. Bolker. 2001. Homologies of process and modular elements of

embryonic construction Pp. 437-456 in G. P. Wagner (ed.), The Character

Concept in Evolutionary Biology. Academic Press, London.

Gould, S. J. 1973. Systematic pluralism and the uses of history. Systematic Zoology

22:322-324.

—1977. Ontogeny and Phylogeny. Belknap Press of Harvard University Press,

Cambridge.

—1989. Wonderful Life: the Burgess Shale and the nature of history. Norton, New York.

Guensburg, T. E. and J. Sprinkle. 2009. Solving the mystery of crinoid ancestry: new

fossil evidence of arm origin and development. Journal of Paleontology 83:350-

364.

Haeckel, E. 1866. Generelle Morphologie der Organismen. Allgemeine Grundzüge der

organischen Formen-Wissenschaft, mechanisch begründet durch die von Charles

Darwin reformirte Descendenz-Theorie. Georg Reimer, Berlin.

Hall, B. K. 2002. Palaeontology and evolutionary developmental biology: a science of the

nineteenth and twenty-first centuries. Palaeontology 45:647-669.

—2003. Descent with modification: the unity underlying homology and homoplasy as

seen through an analysis of development and evolution. Biology Reviews 78:409-

433.

58

Hara, Y., M. Yamaguchi, K. Akasaka, H. Nakano, M. Nonaka, and S. Amemiya. 2006.

Expression patterns of Hox genes in larvae of the sea lily Metacrinus rotundus.

Development Genes and Evolution 216:797-809.

Harnik, P. G., P. C. Fitzgerald, J. L. Payne, and S. J. Carson 2014. Phylogenetic signal in

extinction selectivity in Devonian terebratulide . Paleobiology

40:679-692.

Hendricks, J. R., E. E. Saupe, C. E. Myers, E. J. Hermsen, and W. D. Allmon. 2014. The

generification of the fossil record. Paleobiology 40:511-528.

Hennig, W. 1966. Phylogenetic systematics. University of Press, Urbana.

Hyman, L. 1955. The Invertebrates, Vol. 5. Phylum Echinodermata. McGraw-Hill, New

York.

Kammer, T. W. 2008. Paedomorphosis as an adaptive response in pinnulate cladid

crinoids from the Burlington (Mississippian, Osagean) of the

Mississippi Valley. Pp. 177-195 in W. I. Ausich and G. D. Webster (eds.),

Echinoderm Paleobiology, University of Indiana Press, Bloomington.

Kammer, T. W. and W. I. Ausich. 1996. Primitive cladid crinoids from Upper Osagean-

Lower Meramecian (Mississippian) rocks of east-central United States. Journal of

Paleontology 70:835-866.

—2006. The “age of crinoids”: a Mississippian biodiversity spike coincident with

widespread carbonate ramps. Palaios 21:238-248.

—2007. New cladid and flexible crinoids from the Mississippian (, Ivorian)

of England and Wales. Palaeontology 50:1039-1050.

59

Kammer, T. W. C. D. Sumrall, S. Zamora, W. I. Ausich, and B. Deline. 2013. Oral region

homologies in Paleozoic crinoids and other plesiomorphic pentaradial

echinoderms. PLoS ONE 8e77989. doi:10.1371/journal.pone.0077989

Kirk, E. 1944. Cymbiocrinus, a new inaduate crinoid genus from the Upper

Mississippian. American Journal of Science 242:233-245.

Lahaye, M. C. and M. Jangoux. 1987. The skeleton of the stalked stages of the comatulid

crinoid Atedon bifida (Echinodermata). Zoomorphology 107:58-65.

Laubichler, M. D. 2000. Homology in development and the development of the

homology concept. American Zoologist 40:777-788.

Lefebvre, B. G. J. Eble, N. Navarro, and B. David. 2006. Diversification of atypical

Paleozoic echinoderms: a quantitative survey of patterns of stylophoran disparity,

diversity, and geography. Paleobiology 32:483-510.

Lieberman, B. S. 2000. Paleobiogeography: Using Fossils to Study Global Change, Plate

Tectonics, and Evolution. New York: Kluwer Academic/Plenum Publishers.

Mladenov, P. V. and F. S. Chia. 1983. Development, settling behavior, metamorphosis,

and pentacrinoid feeding and growth of the feather star .

Marine Biology 73:309-323.

Mooi, R. B., B. David, and D. Marchland. 1994. Echinoderm skeletal homologies:

classical morphology meets modern phylogenetics. Pp. 87-95 in B. David, A.

Guille, J. Féral, and M Roux (eds.), Echinoderms Through Time. A. A. Balkema,

Rotterdam.

60

Mooi, R. B. and B. David. 1997. Skeletal homologies of echinoderms. Pp. 305-335 in J.

A. Waters and C. G. Maples (eds.), of Echinoderms, Paleontological

Society Papers 3.

Mooi, R., B. David, and G. A. Wray. 2005. Arrays in rays: terminal addition in

echinoderms and its correlation with . Evolution and

Development 7:542-555.

Moore, R. C. 1962. Ray structures of some inadunate crinoids. University of Kansas

Paleontological Contributions 5:1-47.

Moore, R. C. and Laudon, L. R. 1943. Evolution and Classification of Paleozoic crinoids.

Geological Society of America Special Paper, 46:1-154.

Moore, R. C. and F. B. Plummer. 1940. Crinoids from the Upper and

Permian strata in Texas. University of Texas Publication 3945.

Moore, R. C., and Teichert, C. (eds.) 1978. Treatise on Invertebrate Paleontology, Part T,

Echinodermata 2. Geological Society of America and University of Kansas Press,

Lawrence, 1027 p.

Moore, R. C., N. G. Lane, and H. L. Strimple. 1978. Order Cladida Moore and Laudon,

1943, Pp. 578-759 in R. C. Moore and Teichert, C. (eds.), Treatise on Invertebrate

Paleontology. Part T, Echinodermata 2. Geological Society of America and

University of Kansas Press, Lawrence.

Mortensen, T. 1920. Studies in the development of crinoids. Papers from the Department

of Marine Biology, Carnegie Institution of Washing 16.

61

Nakano, H., T. Hibino, T. Oji, Y. Hara, and S. Amemiya. 2003. Larval stages of a living

sea lily (stalked crinoid echinoderm). Nature 421:158-160.

Patterson, C. 1982. Morphological characters and homology. Pp. 21-74 in K. A. Joysey

and A. E. Friday (eds.), Problems of Phylogenetic Reconstruction. Academic

Press, New York.

Paul, C. R. C. and A. B. Smith. 1984. The early radiation and phylogeny of echinoderms.

Biological Reviews 59:443-481.

Philip, G. M. 1964. Australian fossil crinoids. I, Introduction and terminology for the anal

plates of crinoids. Proceedings of the Linnaean Society of New South Wales

88:259-272.

Purvis, A. 2008. Phylogenetic approaches to the study of extinction. Annual Review of

Ecology, Evolution, and Systematics 39:301-319.

Rabosky, D. L. and A. R. McCune 2010. Reinventing species selection with molecular

phylogenies. Trends in Ecology and Evolution 25:68-74.

Raff, R. A. 2007. Written in stone: fossils, genes, and evo-devo. Nature Reviews

Genetics 8:911-920.

Rasmussen, H. W. Evolution of articulate crinoids. Pp T302-T316 in R. C. Moore and

Teichert, C. (eds.), Treatise on Invertebrate Paleontology. Part T, Echinodermata

2. Geological Society of America and University of Kansas Press, Lawrence.

Remane, A. 1952. Die Grundlagen des Naturlichen Systems der Vergleichenden

Anatomie und der Phylogenetik. Geest und Portig, Leipzig.

62

Rouse, G. W., L. S. Jermiin, N. G. Wilson, I. Eeckhaut, D. Lanterbecq, T. Oji, C. M.

Young, T. Browning, P. Cisternas, L. E. Helgen, M. Stuckey, and C. G. Messing.

2013. Fixed, free, and fixed: the fickle phylogeny of extant Crinoidea

(Echinodermata) and their Permian– origin. Molecular Phylogenetics and

Evolution 66:161-181.

Roux, M. M. Eleaume, L. G. Hemery, and N. Ameziane. 2013. When morphology meets

molecular data in crinoid phylogeny: a challenge. Cahiers de Biologie Marine

54:541-548.

Rozhnov S. V. and G. V. Mirantsev. 2014. Structural aberrations in the cup in cladid

crinoids from the Carboniferous of the Moscow region. Paleontological Journal

48:1243-1257.

Saint-Hilaire, E. G. 1830. Principes de Philosophie Zoologique, discutés en Mars 1830,

au Sein de l’Académie Royale des Sciences. Pichon et Dider, Paris.

Sereno, P. C. 2007. Logical basis for morphological characters in phylogenetics.

Cladistics, 23:565-587.

Sepkoski, J. J., Jr. 1981. A factor analytic description of the Phanerozoic marine fossil

record. Paleobiology 7:36-53.

Shibata, T. F. A. Sato, T. Oji, and K. Akasaka. 2008. Development and growth of the

feather star Oxyamanthus japonicas to sexual maturity. Zoological Science

25:1075-1083.

Shubin, N. H. and C. R. Marshall 2000. Fossils, genes, and the origin of novelty.

Paleobiology 26:324-340. 63

Simms, M. J. 1993. Reinterpretation of thecal plate homology and phylogeny in the

Crinoidea. Lethaia 26: 303-312.

Simms, M. J. and G. D. Sevastopulo. 1993. The origin of articulate crinoids.

Palaeontology, 36:91-109.

Smith, A. B. 2001. Large-scale heterogeneity of the fossil record: implications for

Phanerozoic biodiversity studies. Philosophical Transactions of the Royal Society,

London 356:351-368.

Springer, F. 1920. The Crinoidea Flexibilia. Smithsonian Institution, Publication 2501.

Sprinkle, J. and T. E. Guensburg. 1997. Early radiation of echinoderms. Pp. 205-224 in J.

A. Waters and C. G. Maples (eds.), Geobiology of Echinoderms, Paleontological

Society Papers 3.

Sprinkle, J. and P. M. Kier. Phylum Echinodermata. 1987. Pp. 550-611 in R. S.

Boardman, A. H. Cheetham, and A. J. Rowell (eds.), Fossil Invertebrates,

Blackwell Scientific, Palo Alto.

Sprinkle, J. and G. P. Wahlman. 1994. New echinoderms from the Early Ordovician of

West Texas. Journal of Paleontology 68:324-338.

Strimple, H. L. 1948. Notes on Phanocrinus from the Fayetteville Formation of

Northeastern Oklahoma. Journal of Paleontology 22:490-493.

—1978. Evolutionary trends among Poteriocrinina. Pp. T298-T301 in R. C. Moore and

Teichert, C. (eds.), Treatise on Invertebrate Paleontology. Part T, Echinodermata

2. Geological Society of America and University of Kansas Press, Lawrence.

64

Sumrall, C. D. 1997. The role of fossil in the phylogenetic reconstruction of

Echinodermata. Pp. 267-288 in J. A. Waters and C. G. Maples (eds.), Geobiology

of Echinoderms, Paleontological Society Papers 3.

—2008. The origin of Lovén’s Law in glyptocystitoid rhombiferans and its bearing on

the plate homology and the heterochronic evolution of the hemicosmitid

peristomal border. Pp. 228-241 in W. I. Ausich and G. D. Webster (eds.),

Echinoderm Paleobiology, University of Indiana Press, Bloomington.

—2010. A model for elemental homology for the peristome and ambulacra in blastozoan

echinoderms. Pp. 269-276 in Harris, L. G., Böttger, S. A., Walker, C. W., and

Lesser, M. P. (eds.) Echinoderms: Durham. Taylor and Francis Group, London,

UK.

Sumrall, C. D. and J. A. Waters. 2012. Universal elemental homology in glyptocystitoids,

hemicosmitoids, coronoids and : steps toward echinoderm phylogenetic

reconstruction in derived blastozoa. Journal of Paleontology 86:956-972.

Thomson, W. C. 1865. On the embryogeny of rosaceus (Lmk) (Comatula

rosacea of Lamark). Philosophical Transactions of the Royal Society of London

155:513-545.

Ubaghs, G. 1953. Classe des crinoides. In J. Piveteau (ed.). Traité de Paléontologie. 3:

658-773.

—1978. Skeletal morphology of fossil crinoids. T58-T216 in R. C. Moore and C.

Teirchert (eds.), Treatise on invertebrate paleontology. Part T, Echinodermata 2,

65

Crinoidea, Geological Society of America, Boulder and University of Kansas

Press, Lawrence.

van Valen, L. M. 1982. Homology and causes. Journal of Morphology. 173:305-312.

von Baer, K. E. 1828. Uber Entwickelungsgeschichte der Thiere: Beobachtung und

Reflektion, Bornträger.

Wagner, G. P. 2006. Homologues, natural kinds and the evolution of .

American Zoologist 36:36-43.

—2007. The developmental genetics of homology. Nature Reviews Genetics 8:473-479.

Wagner, P. J. 1995. Testing evolutionary constraint hypotheses with Early Paleozoic

gastropods. Paleobiology 21:248-272.

—2000a. Phylogenetic analyses and the fossil record: tests, inferences, hypotheses, and

models. Paleobiology 26:341-371.

—2000b. Exhaustion of morphologic character states among fossil taxa. Evolution

54:365-386.

Walker, L. J., B. H. Wilkinson, and L. C. Ivany. 2002. Continental drift and Phanerozoic

carbonate accumulation in shallow-shelf and deep-marine settings. The Journal of

Geology 110:75-87.

Wanner, J. 1916. Dei Permischen echinodermen von Timor, I. Teil. Paläeontology von

Timor 11:1-329.

Wheeler, W. C. 2012. Systematics: a course of lectures. John Wiley and Sons. 426 pp.

Wiley, E. O. 1975. Karl R. Popper, systematics and classification: a reply to Walter Bock

and other evolutionary taxonomists. Systematic Zoology 24:233-243.

66

Wiley, E. O. and B. S. Lieberman. 2011. Phylogenetics: theory and practice of

phylogenetic systematics. John Wiley and Sons.

Webster, G. D. and P. A. Jell. 1999. New Permian crinoids from Australia. Memoirs of

the Queensland Museum 33:349-359.

Webster, G. D. and N. G. Lane. 1967. Additional Permian crinoids from southern

Nevada. University of Kansas Paleontological Contributions 27:1-32.

Webster, G. D. and C. G. Maples. 2006. Cladid crinoid (Echinodermata) anal conditions:

a terminology problem and proposed solution. Palaeontology. 49: 187-212.

Wright, A. M. and D. M. Hillis 2014. Bayesian analysis using a simple likelihood model

outperforms parsimony for estimation of phylogeny from discrete morphological

data. PLoS ONE 9e109210. doi:10.1371/journal.pone.0109210

Wright, D. F. and W. I. Ausich. 2015. From the stem to the crown: phylogeny and

diversification of pan-cladid crinoids. Cuadernos del Museo Geominero.

Wright, D. F. and A. L. Stigall 2014. Geologic drivers of Late Ordovician faunal change

in Laurentia: investigating links between tectonic, speciation, and biotic

invasions. PLoS ONE 8e68353. doi:10.1371/journal.pone.0068353

Wright, J. 1920. On Carboniferous crinoids from Fife; with notes on some localities, and

provisional lists of species. Transactions of the Geological Society of Glasgow

16:363-392.

—1926. Notes on the anal plates of Epachycrinus calyx and Zeacrinus konincki.

Geological Magazine 64:352-373.

67

—1927. Some variations in Ulocrinus and Hydreionocrinus. Geological Magazine

71:241-268.

Zamora, S., I. A. Rahman, and A. B. Smith. 2012. Plated bilaterians reveal the

earliest stages of echinoderm evolution. PLoS One e38296.

doi:10.1371/journal.pone.0038296

Zamora, S. and I. A. Rahman. 2014. Deciphering the early evolution of echinoderms with

Cambrian fossils. Palaeontology 57:1105-1119.

68

Figure 2.1 Summary of previously proposed phylogenetic relationships of pan- cladid crinoids. A, Phylogeny and classification from the Treatise of Invertebrate

Paleontology based on Moore and Teichert (1978) and Moore et al. (1978). B, A revised phylogeny and classification based on Simms and Sevastopulo (1993), Ausich (1998). C,

Phylogenetic relationships according to Webster and Jell (1999).

69

Fig. 2.1

70

Figure 2.2 Generalized morphology of cladid crinoids. A, Reconstruction of the

Silurian crinoid Dictenocrinus decadactylus depicting the morphologic features described

in the text. Note that a modern reconstruction would place the crown in a down-current

position with the arms in an outstretched position to form a rheophilic filtration fan

(modified from Bather 1900, fig. 3). B, Plate diagram of Dendrocrinus longidactylus

showing the orientation rays and interrays in pan-cladid crinoids. The crinoidal plane of symmetry is interrupted in the posterior region by the addition of plates in the CD interray (modified from Moore et al. 1978, fig. 395). C, The CD interray of three genera depicting common positions and arrangements of posterior plates (modified from Moore et al. 1978, fig. 394). (A-E ray designations in Carpenter’s [1884] system, radials black, radianal cross ruled, anal X and right tube plate [rt] stippled.)

71

Fig. 2.2

72

Figure 2.3. An antinomy of alternative primary homology schemes for a single posterior plate in the Mississippian genus Phanocrinus. A, Possible homology scheme depicting an evolutionary trend in posterior plate reduction leaving the anal X in the cup.

B, An alternative homology scheme depicting the radianal as the single posterior plate.

(Redrawn from Strimple [1948]. Radials black, radianal cross ruled, anal X and right tube plate [rt] stippled.)

73

Fig. 2.3

74

Figure 2.4 Pentacrinoid stage development and morphogenesis of the extant comatulid crinoid Comactinia meridionalis. A, Early pentacrinoid stage showing the radianal in the C ray with close affinities to the gut tract. B, The radianal is larger than the newly formed C radial and lying obliquely below it. C, Radial plates have increased in size. The radianal has migrated toward the middle of the underlying CD basal while still occupying the inner margin of the C radial plate. D, Radial plates are in in lateral contact except in the CD interray, where the radianal occupies a medial position. E, Radial plates are now in complete lateral contact. The radianal has been lifted up within the cup with the growth of the anal tube. F, The radianal rests on the shoulders of the C and D radial plates and continues its upward migration with the anal tube. The radianal is resorbed shortly after this stage. (Stages redrawn from Springer [1920: plate B]. Radials black, radianal cross ruled, AN—anus, B—basal plate, PPCP—primary peristomial cover plate,

RA—radianal plate).

75

Fig. 2.4

76

Chapter 3: Bayesian estimation of fossil phylogenies and the evolution of early to middle

Paleozoic crinoids (Echinodermata)

Abstract.—Knowledge of phylogenetic relationships among species is fundamental to understanding basic patterns in evolution and underpins nearly all research programs in biology and paleontology. However, most methods of phylogenetic inference typically used by paleontologists do not accommodate the idiosyncrasies of fossil data and therefore do not take full advantage of the information provided by the fossil record. The advent of Bayesian “tip-dating” approaches to phylogeny estimation is especially promising for paleo-systematists because time-stamped comparative data can be combined with probabilistic models tailored to accommodate the study of fossil taxa. Under a Bayesian framework, the recently developed fossilized birth‒death process (FBD) provides a more realistic tree prior model for paleontological data that accounts for macroevolutionary dynamics, preservation, and sampling when inferring phylogenetic trees containing fossils. In addition, FBD tree prior also allows for the possibility of sampling ancestral morphotaxa. Although paleontologists are increasingly embracing probabilistic phylogenetic methods, these recent developments have not previously been applied to the deep-time invertebrate fossil record. Here, I examine phylogenetic relationships among Ordovician through Devonian crinoids using a 77

Bayesian tip-dating approach. Results support several clades recognized in previous

analyses sampling only Ordovician taxa, but also reveal instances where

phylogenetic affinities are more complex and extensive revisions are necessary,

particularly among the Cladida. The name Porocrinoidea is proposed for a well- supported clade of Ordovician ‘cyathocrine’ cladids and hybocrinids. The Eucladida is proposed as a clade name for the of the Flexibilia herein comprised of cladids variously considered ‘cyathocrines’, ‘dendrocrines’, and/or ‘poteriocrines’ by other authors.

Introduction

Modern macroevolutionary research resides at the nexus of paleontology and

phylogenetic comparative biology. The fossil record provides a spectacular temporal

window into the vicissitudes of life’s history and paleontologists have long used its

patterns to investigate large-scale trends in diversification dynamics and morphologic

evolution over timescales inaccessible to experimental manipulation or field-based investigation (Simpson, 1944; Sepkoski, 1981; Hunt et al., 2008; Alroy, 2010). Similarly, biologists armed with molecular phylogenies of extant species and tree-based statistical techniques have increasingly become interested in addressing macroevolutionary questions traditionally studied by paleontologists (e.g., O’Meara et al., 2006; Bokma,

2008; Rabosky, 2009; Rabosky and McCune, 2009; Harmon et al., 2010; Pennell et al.,

2014). Although differences between paleontologic and biologic perspectives remain,

78

attempts to bridge disciplinary gaps between fields have wide-reaching implications for

assembling a more synthetic macroevolutionary theory (Jablonski, 2008; Slater and

Harmon, 2013; Hunt and Slater, 2016).

Instances of integration between fields, such as paleontology and molecular phylogenetics, often provide opportunities for reciprocal illumination. For example, fossils play a major role in dating divergences among extant species. Without external information to constrain absolute ages, branch length estimation is confounded by the fact that both rates of molecular sequence evolution and elapsed time contribute to observed distances among species. Thus, the construction of a time-calibrated molecular

phylogeny requires information on fossil morphologies and their temporal distributions to

provide a numerical timescale for testing alternative models of macroevolutionary

dynamics (Donoghue and Benton, 2007; Ksepka et al., 2015; dos Reis et al., 2015).

Equally illuminating for paleontologists, many probabilistic methods originally

developed by molecular phylogeneticists can be modified and applied to paleontologic

data (Wagner, 2000a; Wagner and Marcot, 2010; Lee and Palci, 2015; but see Spencer

and Wilberg, 2013). For example, Lewis (2001) developed a k-state Markov model for

calculating likelihoods of discrete, morphologic characters based on a generalization of

the Jukes-Cantor model of molecular sequence evolution. Although simplistic, Lewis’

(2001) “Mk” model has recently been demonstrated in a Bayesian context to outperform

other phylogenetic methods under a range of conditions present in real datasets, including

missing character data, high rates of character evolution (and therefore homoplasy), and

rate heterogeneity among characters (Wright and Hillis, 2014; O’Reilly et al., 2016). The

79

recent resurgence of “total-evidence” (Pyron, 2011; Ronquist et al., 2012) approaches in

phylogenetics coincides with a renewed interest among biologists in phenotypic evolution

and the utility of morphologic phylogenetics in an age of “post-molecular systematics”

(Lee and Palci, 2015; Pyron, 2015). This revival of interest in morphologic

phylogenetics is good news for paleontologists because nearly all phylogenies of fossil

species are inferred using only morphologic character data. Indeed, there has been an

increasing number of studies employing probabilistic approaches to estimate phylogenies

with morphologic data, especially in paleontology (e.g., Wagner, 1998, 1999; Snively et

al., 2004; Pollitt et al., 2005; Clarke and Middleton, 2008; Pyron, 2011; Ronquist et al.,

2012; Wright and Stigall, 2013; Lee et al., 2014; Slater, 2013, 2015; Close et al., 2015;

Gorscak and O’Connor, 2016; Bapst et al., 2016).

Particularly promising for systematic paleontology is the advent of tip-dating

approaches for inferring phylogenies containing non-contemporaneous taxa (Pyron,

2011; Ronquist et al., 2012; Gavryushkina et al., 2014). Bayesian “total-evidence” tip-

dating combines molecular sequences, morphologic character data, and temporal

information on fossil distributions to simultaneously estimate the best tree topologies,

branch lengths, and divergence times among extinct and extant lineages (Ronquist et al.,

2012; Lee and Palci, 2015). Tip-dating approaches operate on the simple assumption that

evolution can be modeled as a function of time, with either a strict or relaxed clock-like

model of character change. Although most tip-dating studies combine fossil and living species (e.g., Pyron, 2011; Ronquist et al., 2012; Slater, 2013), tip-dating approaches

equally apply to character matrices containing only morphologic data (Slater, 2015)

80

and/or with only fossil taxa (Lee et al., 2014; Gorscak and O’Connor, 2016; Bapst et al.,

2016). Moreover, mathematical models originally developed for studying the spread of

viruses in epidemiology have found applications in fossil tip-dating (Stadler et al., 2012;

Stadler and Yang, 2013; Gavryushkina et al., 2014). The “fossilized birth‒death” process

(Stadler, 2010; Heath et al., 2014) has recently been applied within a Bayesian context as

a more realistic tree prior distribution that accounts for macroevolutionary and sampling

processes (Gavryushkina et al., 2014).

This paper presents the first application of Bayesian tip-dating methods to a

fossil-only dataset of invertebrate animals. Here, I examine phylogenetic relationships

among early to middle Paleozoic crinoids (Echinodermata). Crinoids are particularly

amenable for the purposes herein because (1) they have a well-sampled fossil record

(Foote and Raup, 1996), (2) their skeletal morphology is highly complex and character-

rich (Ubaghs, 1978; Foote, 1994; Ausich et al., 2015), and (3) testing phylogenetic

hypotheses among crinoid higher taxa requires sampling non-contemporaneous taxa over

long timescales (> 107 years), making them an ideal system for implementing a tip-dating approach (Ronquist et al., 2012). Because the approach taken herein is novel to the invertebrate fossil record, I provide a brief discussion on Bayesian tip-dating and the fossilized birth‒death process tree prior to familiarize the reader with these emerging methods. Although this makes the paper necessarily technical in places, it is hoped those sections will provide a useful resource for other systematic paleontologists interested in probabilistic approaches to fossil phylogenetics.

81

Previous work on crinoid phylogeny

The Crinoidea form the sister group to all other extant echinoderm classes (Asteroidea,

Echinoidea, Holothruoidea, and Ophiuroidea) and have an extensive geologic history spanning the Lower Ordovician (~480 Ma) to the present day. Ever since Bather (1899) published his seminal work A Phylogenetic Classification of the Pelmatozoa, crinoid systematists have sought a robust evolutionary template for understanding the phylogenetic distribution of fossil and living species (Ausich and Kammer, 2001). Other than a few isolated studies conducted at low taxonomic levels (e.g., Kammer, 2001; Gahn and Kammer, 2002), most phylogenetic research using computational methods have focused on inferring relationships within two key time intervals: the Ordovician and the

Recent (Ausich, 1998; Guensburg, 2012; Hemery et al., 2013; Rouse et al., 2013; Ausich et al., 2015; Summers et al., 2014; Cole, in press). These intervals are significant because they bookend the evolutionary history of crinoids into their early diversification during the Ordovician Period and their present day diversity in marine . However, these intervals are separated by ~480 million years and phylogenetic research linking post-Ordovician stem taxa with the crown Crinoidea remains a largely unexplored area of research (Simms, 1988; Simms and Sevastopulo, 1993; Webster and Jell, 1999).

Crinoids are traditionally divided into several higher taxa, including the Camerata,

Disparida, Hybocrinida, Cladida, Flexibilia, and the Articulata (Moore and Teichert,

1978). Except for articulate crinoids, these groups first appear in Ordovician rocks.

Despite more than a century of controversy, phylogenetic relationships among

82

Ordovician taxa are reaching a consensus. For example, all recent analyses of Ordovician

crinoids strongly support an early divergence between camerate and non-camerate crinoids (Guensburg, 2012; Ausich et al., 2015; Cole, in press). Thus, the Camerata is the sister clade to all non-camerate crinoids. Similarly, both Guensburg (2012) and Ausich et al. (2015) recovered a monophyletic Hybocrinida as the sister clade to a subset of cladid taxa. Ordovician analyses also recover a monophyletic Disparida as sister to the clade of cyathocrine cladids and hybocrinids (Guensburg, 2012; Ausich et al., 2015). However, relationships among taxa placed currently within the Cladida, and their relationships to other higher taxa, have been a long-standing problem in crinoid systematics (McIntosh,

1986, 2001; Sevastopulo and Lane, 1988; Simms and Sevastopulo, 1993; Kammer and

Ausich, 1992, 1996).

In his review of echinoderm phylogeny and classification, Smith (1984) lamented the cladid portion of the crinoid tree was one of the “outstanding areas of ignorance in echinoderm phylogeny” (Smith, 1984, p. 456). Indeed, the Cladida (sensu Moore and

Laudon, 1943) have long been considered a paraphyletic group because some nominal cladids are hypothesized to be more closely related to flexible and/or articulate crinoids than other cladids (Springer, 1920; Simms and Sevastopulo, 1993; Brower, 1995; Ausich,

1998; Wright, 2015). Unfortunately, recent phylogenetic analyses not only confirm that

Ordovician cladids are a paraphyletic assemblage (Guensburg, 2012; Ausich et al., 2015), but also that the validity of the Cladida and their constituent higher taxa (i.e.,

Dendrocrinida and Cyathocrinida) cannot be fully remedied by simply adopting Simms and Sevastopulo’s (1993) recommendation to place the Flexibilia and Articulata within

83 the Cladida. In addition, because the monophyletic status of a taxon requires a temporal reference frame (conventionally taken as the present day), it is unknown whether some recovered “clades” in Ordovician analyses retain their monophyletic status when younger taxa are considered. For example, Ordovician cyathocrine cladids are typically recovered as a clade (Ausich, 1998; Ausich et al., 2015), but are sometimes nested within a more inclusive clade of cyathocrines and hybocrinids when hybocrinids are sampled in the same analysis. Testing whether or not the other cyathocrine cladids belong to this clade remains an open question and requires sampling younger species. Similarly, Ordovician cladids placed within the Dendrocrinida (sensu Moore and Laudon, 1943) are paraphyletic, but there may nevertheless be latent phylogenetic structure present among subsets of post-Ordovician dendrocrines that could inform taxonomic revisions.

The analyses conducted herein build on and further test recently proposed phylogenetic hypotheses that sample only Ordovician crinoids (Guensburg, 2012; Ausich et al., 2015).

Thus, this analysis includes a broad sample of early to middle Paleozoic (Ordovician‒

Devonian) non-camerate crinoids and primarily focuses on resolving relationships among the problematic Cladida.

Bayesian phylogenetics and the fossilized birth-death process

Bayesian phylogenetic methods combine a likelihood model of evolution with a set of prior probabilities to generate a posterior probability distribution of phylogenetic trees 84 and their associated parameters. The Bayesian framework used herein is adapted from

Gavryushkina et al. (2015). This model uses time-stamped morphologic character data to simultaneously estimate a posterior distribution of phylogenetic trees, divergences times, and other macroevolutionary and sampling parameters.

Let be a phylogeny (i.e., tree topology with branch lengths in units of time), be a vector 𝛹𝛹of parameters describing morphologic evolution, and be the tree prior (andδ its associated hyper parameters). Using Bayes theorem, the posteriorπ probability distribution is,

[ | , ] [ | ] [ | ] [ ] [ ] [ , , | , ] = , [ , ] 𝑓𝑓 𝑋𝑋 𝛹𝛹 δ 𝑓𝑓 𝑑𝑑 𝛹𝛹 𝑓𝑓 𝛹𝛹 π 𝑓𝑓 π 𝑓𝑓 δ 𝑓𝑓 𝛹𝛹 π δ 𝑋𝑋 𝑑𝑑 𝑓𝑓 𝑋𝑋 𝑑𝑑 (1)

where X is a character by taxon matrix of morphologic character data and d is a vector of age ranges for each fossil taxon. The numerator on the right hand side of the equation can be separated into the tree likelihood function, [ | , ], and the remaining terms, which comprises the prior. Equations for calculating𝑓𝑓 𝑋𝑋 𝛹𝛹[ |δ , ] are well described in literature and therefore a disquisition on tree likelihoods𝑓𝑓 𝑋𝑋 is𝛹𝛹 notδ presented here. Interested readers are advised to see summaries in Swofford et al. (1996), Lewis (2001), Felsenstein

(2004), and Yang (2014). The density [ | ] describes the tree prior (see below) and

[ | ] is the density of obtaining fossil𝑓𝑓 𝛹𝛹occurrenceπ ranges given (this term is treated herein𝑓𝑓 𝑑𝑑 𝛹𝛹 as a constant, see Gavryushkina et al., 2015). The denominator𝛹𝛹 [ , ] is a

𝑓𝑓 𝑋𝑋 𝑑𝑑 85

normalizing constant and is equal to the marginal probability of the data. Given the

necessary inputs, the posterior distribution of trees is estimated using a numerical

technique called Markov chain Monte Carlo (MCMC) that eliminates the need to

calculate [ , ] when estimating the posterior distribution of trees.

A𝑓𝑓 great𝑋𝑋 𝑑𝑑 strength of the Bayesian paradigm is that sources of uncertainty can be

explicitly incorporated into the evolutionary model via the use of prior distributions on

pertinent parameters (Heath and Moore, 2014). For example, numerous factors can

influence and potentially distort the accuracy of reconstructed evolutionary trees—

arguably the most important parameter in phylogenetic inference. Even in cases where

these other factors are not of primary interest, acknowledging and estimating “nuisance

parameters” is nevertheless important because it reduces the chance that any particular

incorrect assumption will lead to the recovery of specious tree topologies (Huelsenbeck et

al., 2002; Wagner and Marcot, 2010; Gavryushkina et al., 2014). Potential biasing factors

may include variation in rates of morphologic evolution, taxonomic diversification rates,

ancestor-descendant relationships, and (incompletely) sampling taxa over time rather than

from a single time slice (Smith, 1994; Wagner, 2000b, 2000c; Wagner and Marcot, 2010;

Bapst, 2012). Variability in evolutionary rates can be modeled with prior distributions to

describe rate variation among characters. Similarly, rate variation among lineages can be

modeled using uncorrelated ‘relaxed clock’ models where branch-specific rates are independently drawn from the same underlying parametric distribution (Lepage et al.,

2007; Heath and Moore, 2014).

86

Bayesian inference weights the likelihood of a tree by its prior probability. The

fossilized birth-death (FBD) process (Stadler, 2010; Didier et al., 2012; Heath et al.,

2014) is an extension of the constant rate birth-death models commonly used in paleontology (e.g., Raup et al., 1973; Raup, 1985) and considers fossil preservation in addition to diversification dynamics. Below, I briefly describe the FBD process as a tree prior and argue it is well-suited to accommodate these additional sources of concern.

Tree prior.—The fossilized birth‒death (FBD) process is a stochastic branching model for describing macroevolutionary dynamics, fossil preservation, and sampling (Stadler,

2010; Heath et al., 2014). The FBD processes begins at some time to > 0 in the past and

ends when t = 0. As time moves forward (i.e., decreasing toward the “present”), each

lineage may probabilistically undergo one of three process-based events, each according to a distinct constant rate Poisson process: branching (i.e., lineage splitting via speciation)

with rate p, extinction with rate q, or fossil preservation and sampling with rate r (Stadler,

2010; Heath et al., 2014). Lineages alive at the end of the process are sampled with

probability ε corresponding to the sampling fraction of “extant” taxa. Importantly, the

start and end times for the FBD process are arbitrary and time can therefore be shifted to

accommodate temporal frameworks more commonly used in paleontology. For example,

paleontological systematists working on entirely extinct groups (e.g., ) or

sampling taxa from a restricted temporal interval (e.g., Paleozoic crinoids) can shift time

such that t = 0 corresponds to the age of the youngest species sampled. The FBD process

represents a major advance over other birth-death models in paleontology (e.g., Raup,

87

1985) because fossil preservation and sampling issues are modeled in addition to clade

diversification. In Gavryushkina et al.’s (2014) implementation, a lineage may be

sampled more than once, thereby producing an internal node connected to only two

(rather than three) branches. A two-degree internal node in a implies a

hypothesized ancestor-descendent relationship, via direct or indirect ancestry (Fig. 3.1)

(Foote, 1996; Gavryushkina et al., 2014, 2015).

The set of stochastic branching, extinction, and sampling events for a single FBD

process gives rise to a “complete” phylogeny with generating parameters π = (p, q, r, ε, to). The “sampled” FBD phylogeny is obtained when all lineages with unsampled descendants produced by the process are pruned from the tree and therefore represents

the reconstructed tree topology and divergence times implied by the sampled taxa (Fig.

3.1). Trees sampled from the FBD process are called sampled ancestor phylogenetic trees

(even if no ancestors were sampled) and their nodes can be labeled to summarize their

unique history of macroevolutionary and sampling events (Gavryushkina et al., 2014).

Equations for calculating the probability density of [ | ] given the FBD parameters p,

q, r, and ε can be derived by modifying birth-death sampling𝑓𝑓 𝛹𝛹 π models used to study virus

transmissions in epidemiology (Stadler, 2010; Stadler et al., 2012; Gavryushkina et al.,

2014; Zhang et al., 2016) and their details are discussed in supplemental information at

the end of this chapter.

88

Taxon sampling, characters analyzed, and specimens examined

The character matrix analyzed herein was constructed as part of a larger project resolving phylogenetic relationships among Paleozoic crinoids. I updated, modified, and expanded the character list of Ausich et al. (2015) to include an ensemble of new characters to better capture variation among post-Ordovician taxa, particularly among ‘cladids’

(Ausich et al., 2015; Wright and Ausich, 2015). Because the taxonomic diversity of fossil crinoids is formidably high for comprehensive analysis, taxon sampling was restricted to

Ordovician through Devonian species and multiple exemplars were sampled at pertinent taxonomic scales appropriate to the present analysis (Brusatte, 2010). The matrix was constructed in an attempt to maximize sampling across the broad spectrum of taxonomic, morphologic, and preservation gradients while keeping rigorous analysis tractable

(Wagner, 2000b; Carlson and Fitzgerald, 2007; Heath et al., 2008).

The dataset contains representative species from Ordovician, Silurian, and

Devonian families of nominal ‘cladids’ (including cyathocrines and dendrocrines), disparids, hybocrinids, and flexibles (all taxa sensu Moore and Laudon, 1943). Species chosen as exemplars were typically the type species of a type genus that well-characterize the distribution of morphologic traits for each higher taxon, but sometimes geologically older species and/or more complete specimens were sampled instead (Table 1).

Characters, plate homologies, and terminology are after Ubaghs (1978) and Ausich et al.

(2015), with updates from Webster and Maples (2006) and Wright (2015). All but four traits were treated as unordered binary or multistate characters (Supplemental Data 1).

89

These four characters were ordered based on known patterns of crinoid development and

arguments for ordering these traits are discussed in Wright (2015) and Webster and

Maples (2006, 2008). Unknown and inapplicable character states were coded as missing.

This new compilation of more than 3,000 specimen-based observations is the largest and most comprehensive morphologic data matrix ever constructed sampling Ordovician and post-Ordovician fossil crinoids (Supplemental Data 2).

In the final matrix, a total of 87 discrete morphologic characters comprising more than 300 character states were sampled across 42 species of non-camerate crinoids

(Supplemental Data 2). Camerates were not included in the analysis because they diverged from non-camerate crinoids by at least the earliest Ordovician (Guensburg and

Sprinkle, 2003; Guensburg, 2012; Ausich et al., 2015; Cole, in press). Although tip- dating analyses do not per se require use of an outgroup (Ronquist et al., 2012), there are several reasons I used the Tremodocian species Apektocrinus ubaghsi Guensburg and

Sprinkle, 2009 to assist rooting the tree. Apektocrinus was originally described as a tentative cladid that featured traits intermediate between protocrinoids and nominal cladids (Guensburg and Sprinkle, 2009). Protocrinoids were originally considered basal crinoids stemward of the divergence between camerates and non-camerates (Guensburg and Sprinkle, 2003). However, Guensburg (2012) subsequently placed protocrinoids within the Camerata and later analyses by Ausich et al. (2015) recovered protocrinoids to be closer to non-camerates than camerates. Regardless of the labile phylogenetic position of protocrinids, both Guensburg (2012) and Ausich et al. (2015) recovered tree topologies with Apektocrinus as the sister taxon to the clade comprised of non-camerate crinoids.

90

Thus, the early stratigraphic position, mosaic distribution of plesiomorphic and

apomorphic traits, and strong support from previous phylogenetic analyses all indicate

Apektocrinus occupies a position near the base of the non-camerate tree (Guensburg and

Sprinkle, 2009; Guensburg, 2012; Ausich et al., 2015).

Matrix construction required extensive first-hand examination of well-preserved specimens housed in museum collections and the published taxonomic literature. When possible, I coded characters from direct observations of type-series specimens for each species. Although emphasis was placed on observing characters from type specimens, non-type specimens were also examined to ensure the character distributions for each species were coded as completely as possible. Specimens were examined from collections within the United States National Museum of Natural History; the Field

Museum of Natural History, the Lapworth Museum of Geology, and the Natural History

Museum (London).

Phylogenetic analyses

Bayesian phylogenetic analyses were conducted using Markov chain Monte Carlo

(MCMC) sampling in the MPI-version of MrBayes 3.2.5 (Ronquist et al., 2012), which

implements MCMC proposals for FBD trees (Zhang et al., 2016). To account for

differences among alternative model configurations, multiple phylogenetic analyses were

conducted and Bayes Factors (BF) were calculated to statistically compare models.

91

Bayes Factors are used in Bayesian model selection to determine which parameter

configurations provide the best fit to the data and are equal to twice the difference in

marginal log-likelihoods between models (Kass and Raftery, 1995). Following

phylogenetic analyses, I then estimated the marginal log-likelihood of each model using

the stepping-stone sampling method (Xie et al., 2010) with 50 steps and powers of β

corresponding to quantiles of a Beta(0.5, 1.0) distribution. Parsimony-based calculations

were performed using PAUP* 4.0a147 (Swofford, 2002). All additional analyses were

conducted using custom scripts written in the R statistical computing environment

making use of functions from the packages (Paradis et al., 2004), and STRAP (Bell

and Lloyd, 2015). Details regarding choices of prior distributions, constraints, and

MCMC convergence are discussed below.

Morphologic character evolution was modeled using the Mk model with equal transition frequencies among character states and a correction for ascertainment bias in character acquisition (Lewis, 2001). The distribution of rates among characters can assume either a uniform “equal rates” model or explicitly account for rate heterogeneity using a skewed parametric distribution. A preliminary parsimony-based estimation of rate

variation in the crinoid character matrix depicts a highly skewed distribution (Fig. 3.2);

strongly suggesting it is unwise to assume a model of equal rates of change among

characters. This is particularly striking given that parsimony-based rate distributions tend

to underestimate morphologic changes and are therefore slightly biased toward equal

rates (Harrison and Larsson, 2015). To further test this hypothesis, I conducted separate

analyses assuming equal, lognormal, and gamma distributed rates of character change.

92

Following Harrison and Larsson’s (2015) recommendation, analyses with gamma or lognormal variation used eight instead of four discrete rate categories (commonly applied to molecular data).

Fossil tip-dating requires a model characterizing the distribution of evolutionary rates throughout the tree. The case of a strict-morphologic clock assumes rates are constant among lineages. However, the assumption of the strict-clock can be “relaxed” by allowing rates to vary among branches throughout the tree (Lepage et al., 2007; Heath and Moore, 2014). To test whether evolutionary rates vary among lineages, analyses using strict and relaxed-clock analyses were conducted. The independent gamma rates

(IGR) relaxed-clock model was applied to account for variation in rates among branches.

The IGR model facilitates episodic “white noise” variation in rates across the tree and is appropriate because large-scale morphologic evolution is a function of both waiting times and stochastic selective forces (Wagner, 2012; Heath and Moore, 2014). A lognormal distribution was placed on the base-rate of the clock using methods outlined in Ronquist et al. (2012).

A key assumption of tip-dating is that evolutionary change is a function of time.

In other words, geologically younger species are expected to have undergone a greater amount of within-lineage evolution (i.e., ) than older species because more time has elapsed for changes to occur (Smith et al., 1992; Wagner, 2000a). To test whether this assumption holds (and therefore whether or not the tip-dating method is valid for these data), the parsimony-based -to-tip pathlength of each species from a

93

non-clock analysis was regressed against median age dates from the IGR analysis using

both phylogenetically corrected and uncorrected methods (Lee et al., 2014).

The FBD process was used as a tree prior (i.e., “samplestrat = random” in

MrBayes 3.2.5). The implementation of the FBD in MrBayes reparameterizes the FBD

process in terms of net diversification (= p – q), turnover (= q / p), and sampling

probability (= r / (q + r). I placed an Exp(1) prior on net diversification, a Beta(1,1) uniform prior on turnover, and a Beta(2,2) prior on the sampling probability. To account for uncertainty in divergence time estimation, age ranges for fossil species were given broad uniform distributions typically corresponding to the stratigraphic range of their higher taxon and were taken from an updated version of Webster’s (2003) index of

Paleozoic crinoids (Supplemental Data 2). Because the age of the most recent common ancestor of all species in the analysis is well-constrained by fossil evidence to be near the

base of the Ordovician, the tree age prior was fixed to correspond to the earliest

Tremadocian.

Tip-dating is a computationally demanding phylogenetic method that requires a time-

consuming exploration of parameter space. To assist the analysis, several topological

constraints were applied to reduce MCMC exploration of very unlikely trees and to test

more specific phylogenetic hypotheses (see Guillerme and Cooper, 2016). For example,

the monophyly of disparids and flexibles are well-supported by other studies (Brower,

1995; Ausich, 1998; Guensburg, 2012; Ausich et al., 2015) and preliminary analyses.

Thus, their status as clades is not in question. However, the branching positions of these

clades within the larger crinoid tree remain an open question and their phylogenetic

94 placement is evaluated herein. A partial constraint was placed on Eustenocrinus,

Iocrinus, Ibexocrinus, and Heviacrinus that allowed for either Merocrinus and/or

Alphacrinus to be included within the disparid clade if the data support that hypothesis.

This was done because whether Merocrinus is closer to cladids or disparids requires additional testing (cf. Ausich, 1998; Guensburg, 2012). Alphacrinus is a stratigraphically old taxon with a combination of unique and disparid-like traits that may or may not be stemward to ingroup Disparida (Guensburg, 2010). A hard constraint was placed on a flexible clade comprised of Homalocrinus, Icthyocrinus, Lecanocrinus, Protaxocrinus, and Sagenocrinites.

Markov chain Monte Carlo analyses consisted of two independent runs of four chains sampling every 4000 generations for 40 million generations per run with a burn-in percentage of 35%. Convergence was assessed using multiple criteria: average standard deviation of split frequencies among chains were below 0.01 (< 0.05 for some IGR analyses) (Gelman and Rubin, 1992), potential scale reduction factors of ~1.0 (Lakner et al., 2008), effective sample sizes greater than 300 (with many > 1000), and visual inspection of log-likelihoods plots among runs using Tracer v.1.6 (Drummond and

Rambaut, 2007). Finally, the analysis with the best fit parameter settings was repeated three times to ensure estimates of optimal tree topologies were robust across runs.

Together, these diagnostics indicate convergence among tree topologies and parameter estimates.

To summarize the posterior distribution of tree topologies, I generated a maximum clade credibility tree (MCC) using TreeAnnotator (Rambaut and Drummond,

95

2015). Although there is no single agreed upon method for summarizing Bayesian

posterior distributions of phylogenetic trees (Heled and Bouckaert, 2013), posterior

probability can be viewed as an optimality criterion in phylogenetic inference (Rannala

and Yang, 1996; Huelsenbeck et al., 2002; Wheeler and Pickett, 2007; Wheeler, 2012;

Rambaut, 2014). The MCC tree is the tree in the posterior distribution with the maximum

product of clade posterior probabilities and represents a Bayesian point estimate of

phylogeny (Rambaut, 2014).

I also ran a series of sensitivity analyses (n > 5) to explore the effects of choosing different priors, including the prior placed (1) on the variance of the gamma distribution in the IGR model, and (2) on the FBD turnover (= q / p) parameter. In all cases, statistically indistinguishable median estimates were obtained for node ages, branch lengths, and FDB parameters. In addition, I calculated the pairwise Robinson-Foulds

(RF) distance (Robinson and Foulds, 1981) within and among tree distributions from separate analyses and ordinated the resulting RF matrix using principal coordinate analysis. A visual inspection of RF distances in principal coordinate space reveals substantial overlap between distributions, with no obvious gradient or isolated “islands”

(analogous to Maddison, 1991) of trees. Thus, the analysis presented herein is considered robust across a range of possible prior configurations.

96

Results

The relationship between within-lineage morphologic evolution and IGR age estimates indicate early to middle Paleozoic crinoids conform well to the assumptions of tip-dating.

Parsimony-based branch lengths from a non-clock analysis reveals geologically younger taxa have higher amounts of anagenesis compared to older taxa (p = 0.017) (Fig. 3.3).

This relationship holds even when accounting for phylogenetic non-independence among comparisons, as phylogenetic independent contrasts (Felsenstein, 1985) also reveal a strong statistical association between branch durations (in the best-fit IGR analysis) and morphologic divergence (measured in parsimony steps from an undated analysis) (p =

0.005) (see Lee et al., 2014).

Bayes Factors provide evidence for heterotachy in crinoid evolution throughout this interval (Supplemental Data 3). The equal rates model of character evolution was strongly rejected in favor of models incorporating rate variation (BF = 166.42 for lognormal, BF = 162.04 for gamma), with the lognormal slightly outperforming a gamma distribution (BF = 4.38). Similarly, the strict-morphologic clock was strongly rejected in favor of the IGR relaxed-clock model accounting rate variation among lineages (BF =

54.56).

The MCC tree from the best fit model is presented in Figure 3.4. Although posterior probabilities for some clades are low, they are comparable to other tip-dating studies using morphologic characters (Lee et al., 2014; Gavryushkina et al., 2015;

Gorscak and O’Connor, 2016). Because Bayesian analyses account for uncertainty in the

97

phylogenetic placement of taxa, it is important to stress that relationships depicted in

Figure 3.4 are not the only ones supported in the posterior distribution. However, MCC

tree topologies were broadly consistent across all analyses, indicating topological results

are robust across different model configurations. Keeping uncertainty in mind, there are

many salient features of the MCC tree that support previous phylogenetic hypotheses and

throw light on several taxonomic questions.

The base of the MCC tree is characterized by a basal divergence between

disparids (and disparid-like taxa) and all other non-camerate crinoids. Neither

Merocrinus nor Alphacrinus were placed within the clade comprised of Iocrinus,

Ibexocrinus, Eustenocrinus, and Heviacrinus. However, Merocrinus is placed as the sister taxon to Metabolocrinus, which together form a sister to the above mentioned disparid clade. Interestingly, Alphacrinus is placed as the sister taxon to the clade comprised of Merocrinus, Metabolocrinus, and disparids, which supports Guensburg’s

(2010) original hypothesis of Alphacrinus being a basal disparid-like taxon and contrasts with Guensburg’s (2012) subsequent analysis recovering Alphacrinus as nested within the disparid clade.

Similar to other studies (Ausich, 1998; Guensburg, 2012; Ausich et al., 2015), a clade comprised of cladids and hybocrinids was recovered. This clade is strongly supported with posterior probability 0.96. The hybocrinids Hybocrinus and Hybocystites are sister taxa and occupy a nested position within a clade of dicyclic ‘cyathocrine’ cladids (i.e., Carabocrinus and Porocrinus), suggesting these taxa are pseudomonocyclic

(Sprinkle, 1982; Guensubrg, 2012; Ausich et al., 2015).

98

A clade comprised of Cupulocrinus and flexible crinoids was recovered as sister

to a clade comprised of only cladids, with the inclusive clade supported with posterior

probability 0.65. Although Cupulocrinus is a nominal cladid (sensu Moore and Laudon,

1943), it is placed closer to flexibles than other cladids in 96% of trees in the posterior

distribution.

The large clade comprising the most recent common ancestor of

Plicodendrocrinus and Corematocrinus and its descendants contains a scattering of taxa traditionally placed within the cladid orders Cyathocrinida and Dendrocrinida (sensu

Moore and Laudon, 1943). Thus, the traditionally recognized ‘cyathocrine’ and

‘dendrocrine’ cladids represent evolutionary grades of body plan organization rather than clades. For example a clade of cyathocrine-grade crinoids containing the most recent common of Thalamocrinus and Gasterocoma is supported with posterior probability

0.80, with subclades supported by posterior probabilities between 0.53‒0.75. Similarly, a different clade containing the most recent common ancestor of Lecythocrinus and

Petalocrinus also features ‘cyathocrine’ morphologies. These results support (perhaps unfortunately) long-held suspicions of taxonomic anarchy among cladids recognized by

previous authors (McIntosh, 1986; 2001; Simms and Sevastopulo, 1993; Kammer and

Ausich, 1992, 1996; Webster and Maples, 2006).

The clade of dendrocrine-grade crinoids containing the most recent common

ancestor of Thenarocrinus and Corematocrinus, and its constituent subclades, are

supported with posterior probabilities between 0.68‒0.84. This clade contains a subclade

comprised of members of the Rutkowskicrinidae, Glossocrinidae, Corematocrinidae, and

99

Amabilicrinidae. This clade is supported with posterior probability 0.84 and is equivalent to the Superfamily Glossocrinacea originally recognized by Webster et al. (2003).

Discussion

It is generally appreciated that quantitative phylogenetic methods do not typically take full advantage of the complete spectrum of information supplied by the fossil record

(Wagner and Marcot, 2010). However, recently developed probabilistic macroevolutionary models and powerful computational tools have provided major advancements for estimating phylogenies containing fossil taxa and accommodating paleontologic idiosyncrasies, such as sampling taxa (incompletely) over time (Stadler,

2010; Ronquist et al., 2012; Gavryushkina et al., 2014; Lee and Palci, 2015). This paper builds on these advances by implementing a Bayesian framework to estimate time-scaled phylogenetic hypotheses of early to middle Paleozoic fossil crinoids. The resulting phylogeny indicates extensive taxonomic revisions are necessary, especially among the

‘cladid’ crinoids, and points to several areas where further analysis at lower taxonomic levels is needed. In addition to evolutionary implications for crinoids, the results also raise several key issues regarding probabilistic approaches to phylogenetic inference in the fossil record and suggest possible directions for future research.

100

Implications for crinoid evolution and systematics.—The phylogenetic analysis presented herein offers several insights into early to middle Paleozoic crinoid evolution and provides a basis for requisite taxonomic revisions. Although I provide suggestions for revisions below, attempting to more fully resolve outstanding problems in crinoid systematics and classification is beyond the scope of this paper. Instead, that topic is addressed in a companion paper (see Wright et al., in press).

A basal divergence between disparids and most other non-camerate crinoids has been recovered in a number of recent phylogenetic studies (cf. Guensburg, 2012; Ausich et al., 2015), including the analysis herein. Although several nominal ‘cladids’ are stemward of this split (see Ausich et al., 2015, fig. 5), the overwhelming majority of nominal cladids (sensu Moore and Laudon, 1943) are not. Thus, disparids are nested within a clade comprised of the common ancestor of all nominal ‘cladids’ (sensu Moore and Laudon, 1943) and all of its descendants. Following Simms and Sevastopulo (1993), the Flexibilia and Articulata are placed within the Cladida, but no previous phylogenetic hypothesis has considered the Disparida a subclade within the Cladida. In an effort to retain as much of the original intent and traditional use of taxonomic names as possible, a re-definition of the Cladida is necessary to prevent the Disparida from being considered a subclade of cladids, particularly since the Cladida is already in need of extensive revision for other reasons. A simple solution to remedy the problem could be obtained using phylogeny-based clade definitions that recognize the Disparida and Cladida (sensu

Simms and Sevastopulo, 1993) as sister clades. This would require orphaning only a small number of so-called ‘cladids’ (sensu Moore and Laudon, 1943) as stem taxa to the

101

Disparida + Cladida clade and retain the majority of cladids (sensu Moore and Laudon,

1943) within the more inclusive Cladida (sensu Simms and Sevastopulo, 1993). The

Disparida is considered herein to include the most recent common ancestor of

Alphacrinus and Eustenocrinus. Thus, Merocrinus and Metabolocrinus (typically

considered cladids [Ausich, 1998], but see Guensburg, 2012) are tentatively placed

within the disparid clade (Fig. 3.4), but further work and character-based analyses are needed to confirm the precise phylogenetic position of these problematic taxa. A more comprehensive discussion with rigorous phylogenetic definitions and a revised classification for cladid and disparid clades is provided in Wright et al. (in press).

Previous analyses of Ordovician taxa have recovered monophyletic groups, such as a clade comprised of cyathocrine cladids and hybocrinids (Guensburg, 2012; Ausich et al., 2015). This offered some hope that potentially the Cyathocrinida might be monophyletic (Ausich, 1998; Guensburg, 2012; Ausich et al., 2015). However, the present analysis rejects the monophyly of the Cyathocrinida because some nominal cyathocrines are more closely related to nominal dendrocrines than to other cyathocrines.

If these results are taken seriously, then an extensive revision of higher taxa within the

Cladida is needed. To further test this issue, I conducted an additional analysis placing a hard topological constraint on all cyathocrine cladids (see Bergsten et al., 2013).

Comparing this analysis to the best fit model described above, Bayes Factors strongly reject a model where cyathocrines are forced to be monophyletic (BF = 9.64). Thus, caution should be exercised when extrapolating results from an analysis considering one timeslice to subsequent time intervals. Nevertheless, there is strong support for a clade of

102

cyathocrine-grade cladids and hybocrinids (Ausich, 1998; Guensburg, 2012; Ausich et

al., 2015). This clade is characterized by a number of morphologic features convergent

with blastozoan echinoderms, such as thecal respiratory structures, calyx and/or arm plate

reduction, and recumbent ambulacra. Given the strong statistical support and

morphologic distinctness of this clade, I propose the name Porocrinoidea to represent this

idiosyncratic group of crinoids. The Hybocrinida is considered herein a subclade within

the Porocrinoidea (Fig. 3.4).

The position of Cupulocrinus at the base of the flexible clade has strong statistical

support and corroborates earlier studies linking Cupulocrinus with flexible crinoids

(Springer, 1911, 1920; Brower, 1995; Ausich, 1998) (Fig. 3.4). For example, Springer

(1920) considered Cupulocrinus to have traits intermediate between cladids and flexibles and hypothesized a species of Cupulocrinus was ancestral to the Flexibilia. Because phylogenetic relationships were estimated using methods that include the possibility of potentially sampling ancestral morphotaxa, Springer’s (1920) hypothesis can be quantitatively addressed. The posterior probability of a taxon being a sampled ancestor can be estimated as the frequency in which it was recovered to have a zero-length branch

in the posterior distribution of trees (Matzke, 2015). Examining the posterior distribution

of trees from the best-fit model, the probability of Cupulocrinus humilis (Billings, 1857) being an ancestral morphotaxon is 0.99. Given that a clade is defined to be an ancestor and all of its descendants, Cupulocrinus is removed from the Cladida and placed within the Flexibilia. Additional analyses with more comprehensive sampling of Cupulocrinus

103

and flexible species (including a broader sample of taxon-specific characters) are needed

to further test this hypothesis at finer taxonomic scales.

The sister clade to the Flexibilia contains the majority of all nominal taxa

currently placed within the Cladida (sensu Moore and Laudon, 1943). This clade originated prior to the close of the Ordovician and contains most taxa traditionally placed within the orders Dendrocrinida and Cyathocrinida. Notably, all species in this clade share a more recent common ancestor with an extant crinoid than with flexible crinoids

(Simms and Sevastopulo, 1993). Thus, I propose the name Eucladida to distinguish this important group of crinoids from their sister clade.

The recovery of the Glossocrinacea as a clade provides quantitative support for

evolutionary inferences discussed in Webster et al. (2003). These “transitional

dendrocrinids” (McIntosh, 2001) are among the first cladids to evolve pinnules and are

traditionally placed within the Order Poteriocrinida. Most crinoid workers since

publication of the Treatise of Invertebrate Paleontology (Moore and Teichert, 1978) have

hesitated to recognize the Poteriocrinida because they are widely considered to be

polyphyletic (Kammer and Ausich, 1992; McIntosh, 2001). Regardless of their status as a

clade or a grade, the crinoids considered poteriocrines in the Treatise are the most

dominant and ecologically abundant group of crinoids throughout the middle to late

Paleozoic (Ausich et al., 1994). Notably, the ancestor of extant articulate crinoids is

widely considered to be placed among a paraphyletic group of ‘poteriocrine’ crinoids

(Simms and Sevastopulo, 1993; Webster and Jell, 1999; Rouse et al., 2013). The recovery

of a clade of poteriocrines herein suggests there may be some phylogenetic structure

104 present among Paleozoic poteriorcrine taxa. However, future analyses sampling younger taxa and a broader sample of Treatise (Moore and Teichert, 1978) poteriocrines are needed to test whether this is the case.

Probabilistic approaches to fossil phylogenies.—Tree-based comparative methods are becoming commonplace in paleontology for testing macroevolutionary patterns and processes within a fully phylogenetic context. To date, most of these studies apply an a posteriori timescaling algorithm to an unscaled cladogram (e.g., Brusatte et al., 2008;

Hunt and Carrano, 2010; Lloyd et al., 2012; Hopkins and Smith, 2015). Although useful for removing the zero-length branches that arise from polytomies in cladistic hypotheses, many of these a posteriori timescaling methods are problematic because they make ad- hoc and unrealistic assumptions regarding node ages and/or ancestor-descendant relationships (Bapst and Hopkins, in press). The cal3 method developed by Bapst (2013) is a promising a posteriori approach that overcomes many of these problems via a model of branching, extinction, and sampling similar to the FBD process. However, this technique requires a priori estimates of these parameters and can only be applied to unscaled . The Bayesian tip-dating approach advocated herein simultaneously estimates tree topologies and divergence dates using time-stamped comparative data.

Thus, a sample of trees from the posterior distribution of a tip-dated analysis provides a more natural framework for testing macroevolutionary patterns using the fossil record while accounting for uncertainty in tree topology and node ages (Close et al., 2015;

Gorscak and O’Connor, 2016).

105

Evaluating the efficacy of competing phylogenetic methods is a contentious (and

sometimes acrimonious) debate, yet inferences using simple probabilistic methods

perform well when inferring trees from paleontologic data and can explicitly consider

different evolutionary and sampling parameters potentially influencing recovered

topologies (Wagner, 1998; Wagner and Marcot, 2010; Wright and Hillis, 2014). Thus, it

seems that in the future more phylogenetic analyses will take advantage of fully

probabilistic frameworks such as the one presented herein. However, researchers

conducting tree-based comparative analyses on older (or otherwise unscaled) cladograms

require an a posteriori time-scaling approach, which might take the form of applying

FBD-like divergence dating methods to a fixed cladistic topology (Bapst and Hopkins, in

press). Regardless of whether Bayesian tip-dating or a model-based a posteriori method

like cal3 becomes the dominant approach to timescaling trees in the future, it is apparent that both of these approaches recover more accurate estimates and are strongly recommended over ad-hoc methods when conducting downstream macroevolutionary

analyses.

Evolutionary patterns among early to middle Paleozoic crinoids strongly favor

models incorporating rate heterogeneity among characters and among lineages. This not

only supports previous investigations demonstrating differential disparity patterns among

crinoid clades (Foote, 1994; Deline and Ausich, 2011), but may also be a more general

feature of morphologic evolution. Probabilistic models of morphologic evolution

commonly assume either uniform or gamma distributed rates among characters. Models

of rate variation predict some characters evolve at higher rates than others (and therefore

106

anticipate a degree of homoplasy in the data). Variable rate distributions are potentially

more realistic than an equal rates model because morphologic characters commonly

experience different selective pressures and/or developmental constraints. Moreover,

accounting for rate variation has a practical value because it may help resolve branches at different levels in a phylogenetic tree (Wright and Hillis, 2014).

Until recently, only uniform and gamma distributions were available to model

among character rate variation in common software packages for Bayesian inference

(Huelsenbeck et al., 2015). However, Wagner (2012) found that fossil datasets commonly

favor lognormal rather than gamma distributed rates, particularly for echinoderm and

mollusc character matrices. Interestingly, gamma and lognormal rate distributions arise

from different underlying processes of character evolution. Gamma rate distributions

assume rates are Poisson processes, whereas lognormal rate distributions suggest

morphologic evolution results from multiple probabilistic processes and/or hierarchical

integration among characters (Wagner, 2012). Thus, the relative fit of the data to these

two distributions gives some insight into the dynamics of character change. The analysis

herein provides positive evidence supporting lognormal over gamma-distributed rates, but only modestly (BF = 4.38). Although the choice of rate distribution had no obvious effect on recovered topologies herein, other workers should nevertheless test alternative distributions and choose the best fit model for their data (Harrison and Larsson, 2015).

The FBD process provides paleontologists with a far more realistic tree prior model than others previously available. For example, the FBD tree prior has recently been demonstrated to outperform a uniform prior (Matzke and Wright, 2016). Other

107

models, such as the Yule process or simple birth-death process are strongly violated when making phylogenetic inferences from paleontologic data. Moreover, the FBD tree prior has a high level of internal consistency for estimating age dates and a good fit to morphologic and geologic data in empirical, well-characterized data sets (e.g., penguins and canids, see Drummond and Stadler, 2016). Analyses of the crinoid data set herein assumed the simple case of constant rates for macroevolutionary and sampling parameters. However, the FBD process can be extended to a more sophisticated time- varying (i.e., piecewise-constant) model that may be useful for other data sets. Similarly, models could be developed to account for geographic variation in sampling probabilities

(Wagner and Marcot, 2013). Such models may be especially beneficial for studies with larger, more comprehensive character matrices spanning similar to longer time intervals than considered herein (Gavryushkina et al., 2014; Zhang et al., 2016).

A major innovation of the FBD process as a tree prior is the ability to account for sampling ancestor-descendant relationships in phylogenetic analysis. Although the notion of discovering “true” ancestors is somewhat contentious (see Smith, 1994; Foote, 1996), modeling studies suggest ancestral morphotaxa are likely present in the fossil record of many paleontologically important groups (Foote, 1996). Gavryushkina et al. (2014) demonstrated that sampled ancestors should be accounted for when estimating phylogenies, even when ancestral morphotaxa are not of specific interest in the analysis, because not including them introduces biases in parameter estimation. Thus, even if sampled ancestors are considered nuisance parameters (Close et al., 2015; Gorscak and

108

O’Connor, 2016), they may nevertheless be important for accurately estimating more accurate tree topologies and node ages.

Parameter estimation under the FBD process does not require exhaustive sampling of fossil taxa, but it does require a representative random sample of species

(e.g., the sampling strategy used herein) (Didier et al., 2012). However, the application of sampled-ancestor tip-dating methods to exhaustively sampled species-level (sensu Smith,

1994) or specimen-level data represents an important (but unexplored) frontier in phylogeny-based analyses of macroevolution. For example, coding multiple fossil specimens of species-level morphotaxa from different time horizons and/or geographic localities may provide a means for testing whether speciation events occur primarily though budding or bifurcating (Gavryushkina et al., 2014; Hunt and Slater,

2016). In addition, paleontologists are commonly interested in whether morphologic change is punctuated or gradual (Eldredge and Gould, 1972; Hunt, 2008). Testing alternative probabilistic models of character evolution within a similar phylogenetic and sampling framework as described above, where each model makes different assumptions about punctuated vs. gradual rates of change (as well as durations of stasis), would provide additional insight into the dynamics of morphologic evolution during speciation events (see Wagner and Marcot, 2010).

The emerging synthesis between paleontology and model-based phylogenetics contributes to the growing consensus that research programs in systematic paleontology are greatly enhanced when grounded in rigorous analytical approaches (Smith, 1994;

Wagner, 2000a; Wagner and Marcot, 2010; Slater and Harmon, 2013; Hunt and Slater,

109

2016). Many of the analytical tools discussed in this paper were originally developed for non-paleontologic purposes. Thus, it is perhaps not surprising there is plenty of room for future modifications and refinement of these techniques to better test paleontologic patterns. Nevertheless, the development and application of these methods has already expanded our ability to quantitatively address macroevolutionary questions. Continued research implementing probabilistic approaches in phylogeny-based paleontology will likely return the favor and provide neontologists with evolutionary insights and unique perspectives only accessible to paleontologists.

110

Supplemental information: Calculating the probability density of FBD

phylogenetic trees

This section provides equations for calculating the probability density of a sampled ancestor phylogenetic tree generated by the fossilized birth‒death process. These trees are

used as Bayesian priors for phylogenies (i.e., topology and divergence times) in the tip-

dating analysis presented in the main text. The equations presented here follow Stadler

(2010), Stadler et al. (2012), and Didier et al. (2012), with subsequent modifications by

Gavryushkina et al. (2014). I encourage readers to refer to these references for additional

details on birth-death-sampling models and their applications. Interested readers are also

encouraged to see Gavryushkina et al. (2014) and Zhang et al. (2016) for a detailed

description of a birth-death-sampling process with time heterogeneous (i.e., piecewise-

constant) rates, as well as Bapst’s (2013) probabilistic method to a posteriori timescale

un-dated cladograms. Please note that the notation used in this paper differs in places from that of the references mentioned above to be more consistent with the paleontological literature (e.g., Foote, 1997, 2000; Wagner and Marcot, 2010; Bapst,

2012).

For simplicity, ordered trees are used in the derivation below. However, ordered trees are subsequently modified to labeled trees (using a conversion factor) to calculate likelihoods (Stadler, 2010). Ordered trees distinguish between left and right branches and identify nodes via a unique left-right path from the root, whereas labeled trees instead give sampled nodes distinct labels to describe branching patterns and divergences

111

(Stadler, 2010; Gavryushikina et al., 2014). Each labeled phylogenetic tree (Ψ) consists of two elements: a discrete, ranked tree topology Ψ and a continuous time vector τ. The ranked tree topology Ψ with ordered branching events is simply the branching pattern of common ancestry among m taxa distributed at the tips of the tree (i.e., an unscaled cladogram). The time vector τ assigns an unambiguous time to each node, such that it contains the following elements arranged in the exact descending temporal order in which they occur in the tree: the m – 1 lineage splitting times (x1…xm-1,where xm-1 < … < x1); the

m tip-taxon sampling times (y1…ym, where ym < … < y1); and k sampling times of two- degree nodes (z1…zk, where zk < … < z1) (Gavryushkina et al., 2014). The density of an

ordered tree can be converted to a labeled tree by multiplying by the conversion factor [2n

+ m – 1 / n!(m + k)!] (see Stadler, 2010, p. 402).

To obtain the probability density of a given sampled ancestor phylogenetic tree

(Ψ), the likelihood is calculated along each branch b in Ψ moving backward in time. The

probability of any one event (i.e., branching with rate p, extinction with rate q, or

sampling with rate r) occurring during a very small time step Δt is the product of its

Poisson rate and Δt. For example, the probability of a single branching event over a small

time interval is pΔt. The probabilities corresponding to any event happening more than

once during Δt is summarized by an order term, O(Δt2) (Feller, 1968). If Δt is chosen to

be very small, then the probability of more than one event happening during the time

interval can be ignored (see below). Because time is measured from the tips backward in

time toward the root, each small time step Δt moves further into the past.

112

Let Ψb(t) be the probability density that a lineage corresponding to branch b at

time t produced the observed Ψ between time t and t = 0. Thus,

Ψb(to) = f [Ψ | π = (p, q, r, ε, to)].

To obtain the probability density for Ψb(t), I follow Stadler et al. (2012) and first consider

describing Ψb(t + Δt) and assume Ψb(t) is known. After deriving an equation for how the

probability density changes over a small time interval Δt, the definition of a derivative

can be used to obtain a differential equation describing the change in probability toward

the root. This differential equation is then integrated along branches to solve for Ψb(t)

across the phylogeny.

Consider the possible evolutionary histories for single lineage along b from time t

to Δt in the past. Moving down the branch, a lineage may either have undergone at least

one branching event during Δt or experienced no event at all (note that extinction is not

considered here because a lineage cannot logically have gone extinct prior to its

subsequent sampling). Thus, the equation for Ψb(t + Δt) is

2 2 Ψb(t + Δt) = (1 – (p + q + r)Δt – O(Δt )) Ψb(t) + pΔt 2S0 (t) Ψb(t) + O(Δt ),

(A1) where S0 (t) is the probability that a lineage at time t has no sampled descendants (Stadler,

2010). The quantity S0 (t) is given by Gavryushkina et al. (2014) as:

113

(1 ) (1 + ) + + + −𝑐𝑐1(1 ) + (1 + ) ( ) = 2 2 , 𝑒𝑒 1 − 𝑐𝑐 − 𝑐𝑐 𝑝𝑝 𝑞𝑞 𝑟𝑟 𝑐𝑐1 −𝑐𝑐2 𝑒𝑒 − 𝑐𝑐2 𝑐𝑐2 𝑆𝑆0 𝑡𝑡 where 𝑝𝑝

2 = ( + + ) + 4 and = . 2 𝑝𝑝 − 𝑞𝑞 − 𝑝𝑝𝑝𝑝 − 𝑟𝑟 𝑐𝑐1 �� 𝑝𝑝 𝑞𝑞 𝑟𝑟 𝑝𝑝𝑝𝑝� 𝑐𝑐2 𝑐𝑐1

The logic behind Equation A1 is straightforward to interpret if each term on the right hand side is considered in isolation. The first term, (1 – (p + q + r)Δt – O(Δt2)), is the probability that no event takes place along b during Δt. The second term reflects the probability of lineage splitting, where pΔt is the probability that one branching event takes place and 2S0 (t) accounts for the probability that one of the two descendant lineages (either the left or right descendant) left no sampled descendants in the observed tree. The final term, O(Δt2), is a summary term corresponding to the probabilities for multiple events during Δt.

Now that we have an expression for Ψb(t + Δt), if we modify Equation A1 to consider how the probability density changes from time t to time t + Δt, we get:

( + ) ( ) = – ( + + ) ( ) + 2 ( ) ( ) + ( ). 𝜳𝜳𝑏𝑏 𝑡𝑡 Δ𝑡𝑡 − 𝜳𝜳𝑏𝑏 𝑡𝑡 𝑝𝑝 𝑞𝑞 𝑟𝑟 𝜳𝜳𝑏𝑏 𝑡𝑡 𝑝𝑝𝑆𝑆0 𝑡𝑡 𝜳𝜳𝑏𝑏 𝑡𝑡 𝑂𝑂 Δ𝑡𝑡 Δ𝑡𝑡

Finally, taking the limit as Δt → 0 results in the differential equation:

114

( ) = – ( + + ) ( ) + 2 ( ) ( ). 𝑑𝑑 𝜳𝜳𝑏𝑏 𝑡𝑡 𝑝𝑝 𝑞𝑞 𝑟𝑟 𝜳𝜳𝑏𝑏 𝑡𝑡 𝑝𝑝𝑆𝑆0 𝑡𝑡 𝜳𝜳𝑏𝑏 𝑡𝑡 𝑑𝑑𝑑𝑑 (A2)

Letting Te represent the terminal end of a branch b, then the initial values at t = Te are:

( ) ( ) if has two descendant branches, and ( ) if has one descendant branch b ( ) = 𝑏𝑏1 𝑒𝑒 𝑏𝑏2 𝑒𝑒 1 2 𝑝𝑝 𝜳𝜳( )𝑇𝑇 𝜳𝜳 𝑇𝑇 if 𝑏𝑏 has no descendant branches and𝑏𝑏 >𝑏𝑏0 𝑏𝑏1 𝑒𝑒 1 𝑏𝑏 𝑒𝑒 𝑟𝑟 𝜳𝜳 𝑇𝑇 if 𝑏𝑏 has no descendant branches and = 0. 𝜳𝜳 𝑇𝑇 � 𝑒𝑒 𝑟𝑟𝑆𝑆0 𝑇𝑇𝑒𝑒 𝑏𝑏 𝑇𝑇 𝑒𝑒 𝜀𝜀 𝑏𝑏 𝑇𝑇

Conditioning on the clade’s origin (i.e., the age of the root) and the event Sε, where Sε

denotes that at least one lineage was sampled at the end of the process, the closed form

solution for the probability density of Ψ accounting for labeled trees is (Stadler, 2010, p.

400; Gavryushkina et al., 2014, equation 3):

( ) = [ | = ( , , , , , )]

𝑏𝑏 𝑜𝑜 𝑜𝑜 𝜀𝜀 𝜳𝜳 𝑡𝑡 𝑓𝑓 𝛹𝛹 π 1 𝑝𝑝 𝑞𝑞 𝑟𝑟 𝜀𝜀 𝑡𝑡 (𝑆𝑆 ) ( ) = 𝑚𝑚+𝑛𝑛−1 2 ( ) 𝑚𝑚 , ( + )! ( (1𝑘𝑘 𝑛𝑛 ( )) ( ) 𝑟𝑟 𝜀𝜀 𝜑𝜑 𝑡𝑡𝑜𝑜 𝑟𝑟𝑆𝑆0 𝑦𝑦𝑖𝑖 2 𝑖𝑖 � 𝑝𝑝𝑝𝑝 𝑥𝑥 � 𝑖𝑖 𝑚𝑚 𝑘𝑘 ̂0 𝑜𝑜 𝑖𝑖=1 𝑖𝑖=1 𝜑𝜑 𝑦𝑦 𝑝𝑝 − 𝑆𝑆 𝑡𝑡 (A3)

where n is the number of ε sampled lineages, S0, c1, and c2 are defined as above, with

4 ( ) = , 2(1 ) + (1 ) + (1 + ) 2 −𝑐𝑐1𝑥𝑥 2 𝑐𝑐1𝑥𝑥 2 𝜑𝜑 𝑥𝑥 2 2 − 𝑐𝑐2 𝑒𝑒 − 𝑐𝑐 𝑒𝑒 𝑐𝑐

and 115

( ) ( ) = 1 . + ( (1 ) ) ( ) 𝜀𝜀 𝑝𝑝 − 𝑞𝑞 𝑆𝑆̂0 𝑡𝑡𝑜𝑜 − − 𝑝𝑝−𝑞𝑞 𝑡𝑡 𝑝𝑝𝑝𝑝 𝑝𝑝 − 𝜀𝜀 − 𝑞𝑞 𝑒𝑒

Accessibility of supplemental data

Supplemental data deposited in Dryad data package http://datadryad.org/resource/doi:10.5061/dryad.6hb7j

116

References

Angelin, N.P., 1878, Iconographia Crinoideorum: in stratis Sueciae Siluricis fossilium. Samson

and Wallin, Holmiae, 62 p.

Alroy, J., 2010, Geographical, environmental and intrinsic biotic controls on Phanerozoic

marine diversification: Palaeontology, v. 53, p. 1211–1235.

Ausich, W.I., 1998, Phylogeny of Arenig to Caradoc Crinoids (Phylum Echinodermata)

and suprageneric classification of the Crinoidea: The University of Kansas

Paleontological Contributions Papers, New Series, no. 9, 36 p.

Ausich, W.I., and Kammer, T.W., 2001, The study of crinoids during the 20th century

and the challenges of the 21st century: Journal of Paleontology, v. 75, p. 1161–

1173.

Ausich, W.I., Kammer, T.W., and Baumiller, T.K., 1994, Demise of the Middle

Paleozoic crinoid fauna: a single extinction event or rapid faunal turnover?:

Paleobiology, v. 20, p. 345–361.

Ausich, W.I., Kammer, T.W., Rhenberg, E.C., and Wright. D.F., 2015, Early phylogeny

of crinoids within the Pelmatozoan clade: Palaeontology, v. 58, p. 937–952.

Bapst, D.W. 2012, When can clades be potentially resolved with morphology?: PLoS

One e62312. doi:10.1371/journal.pone.0062312

Bapst, D.W., 2013, A stochastic rate‐calibrated method for time‐scaling phylogenies of

fossil taxa: Methods in Ecology and Evolution, v. 4, p.724–733.

117

Bapst, D.W., and Hopkins, M.J., in press, Comparing cal3 and other a posteriori time-

scaling approaches in a case study with the Pterocephaliid trilobites:

Paleobiology.

Bapst, D.W., Wright, A.M., Matzke, N.J., and Lloyd, G.T., 2016, Topology, divergence

dates, and macroevolutionary inferences vary between different tip-dating

approaches applied to fossil theropods (Dinosauria): Biology Letters, 12,

20160237, doi: 10.1098/rsbl.2016.0237.

Bather, F.A., 1890, British fossil crinoids. I. Historical introduction: Annals and Magazine of

Natural History, ser. 6, v. 5, p. 306‒310.

Bather, F.A., 1899, A phylogenetic classification of the Pelmatozoa: British Association

for the Advancement of Science (1898), p. 916–923.

Bather, F.A., 1900, Part III The Echinoderma. The Pelmatozoa, in Lankester, E.R., ed., A

Treatise on Zoology: Adam and Charles Black, London, p. 94–204.

Bell, M.A., and Lloyd, G.T., 2015, strap: an R package for plotting phylogenies against

stratigraphy and assessing their stratigraphic congruence: Palaeontology, v. 58, p.

379–389.

Bergsten, J., Nilsson, A.N., and Ronquist, F., 2013, Bayesian tests of topology

hypotheses with an example from diving : Systematic Biology, v. 62, p. 660‒

673.

Billings, E., 1857, New species of fossils from Silurian rocks of . Canada Geological

Survey, Report of Progress 1853-1856: Report for the year 1856, p. 247‒345.

118

Billings, W.R., 1887. A new genus and three new species of crinoids from the Trenton

Formation with notes on a large specimen of Dendrocrinus proboscidiatus: The Ottawa

Naturalist (Ottawa Field Naturalists' Club Transactions), v. 111, p. 49‒54.

Bokma, F., 2008, Detection of "" by Bayesian estimation of

speciation and extinction rates, ancestral character states, and rates of anagenetic

and cladogenetic evolution on a molecular phylogeny: Evolution, v. 62, p. 2718–

26.

Brower, J.C., 1995, Dendrocrinid crinoids from the Ordovician of northern Iowa and

southern Minnesota: Journal of Paleontology, v.69, p. 939–960.

Brusatte, S.L., 2010, Representing supraspecific taxa in higher‐level phylogenetic

analyses: guidelines for palaeontologists: Palaeontology, v. 53, p. 1–9.

Brusatte, S.L., Benton, M.J., Ruta, M. and Lloyd, G.T., 2008, Superiority, competition,

and opportunism in the evolutionary radiation of : Science, v., 321,

p.1485–1488.

Clarke, J.A., and Middleton, K.M., 2008, Mosaicism, modules, and the evolution of

: results from a Bayesian approach to the study of morphological evolution

using discrete character data: Systematic Biology, v. 57, p. 185–201.

Cole, S.R., in press, Phylogeny and morphologic evolution of the Ordovician Camerata

(Class Crinoidea, Phylum Echinodermata): Journal of Paleontology.

119

Conrad, T.A., 1842, Descriptions of new species of organic remains belonging to the Silurian,

Devonian and Carboniferous Systems of the U. S.: Philadelphia Academy Natural

Sciences, Journal, v. 8, old ser. no. 2, p. 228‒280.

Close, R.A., Friedman, M., Lloyd, G.T. and Benson, R.B., 2015, Evidence for a mid-

Jurassic in : Current Biology, v. 25, p. 2137‒2142.

Deline, B., and Ausich, W.I., 2011, Testing the plateau: a reexamination of disparity and

morphologic constraints in early Paleozoic crinoids: Paleobiology, v. 37, p.214‒

236.

Didier, G., Royer-Carenzi, M. and Laurin, M., 2012, The reconstructed evolutionary

process with the fossil record: Journal of Theoretical Biology, v. 315, p.26‒37.

Donoghue, P.C.J., and Benton, M.J., 2007, Rocks and clocks: calibrating the Tree of Life using

fossils and molecules: Trends in Ecology and Evolution, v. 22, p. 424‒431.

Drummond, A.J., and Rambaut, A. 2007, BEAST: Bayesian evolutionary analysis by sampling

trees: BMC Evolutionary Biology: v. 7, doi:10.1186/1471-2148-7-214.

Drummond, A. J., and Stadler, T., 2016, Bayesian phylogenetic estimation of fossil ages:

arXiv, preprint 1601.07447v1 dos Reis, M., Donoghue, P.C., and Yang, Z., 2016. Bayesian dating of

species divergences in the genomics era: Nature Reviews Genetics, v. 17, p. 71‒

80.

Eldredge, N., and Gould, S.J., 1972, Punctuated equilibria: an alternative to phyletic

gradualism. in. Schopf, T.J.M., (ed.), Models in Paleobiology. Freeman, Cooper

and Company, San Fransisco, p. 82‒115.

120

Feller, W., 1968, An Introduction to Probability Theory and its Applications, v. 1 (third

edition): New York, Wiley, 509 p.

Felsenstein, J., 1985. Phylogenies and the comparative method. American Naturalist, v.

126, p.1‒15.

Felsenstein, J., 2004, Inferring phylogenies, vol 2: Sunderland, Sinauer Associates, 664 p.

Foote, M., 1994, Morphological disparity in Ordovician-Devonian crinoids and the early

saturation of morphological space: Paleobiology, v. 20, p.320–344.

Foote, M., 1996, On the probability of ancestors in the fossil record: Paleobiology, v. 22,

p. 141–151.

Foote, M., 1997, Estimating taxonomic durations and preservation probability:

Paleobiology, v. 23, p. 278‒300.

Foote, M., 2000, Origination and extinction components of taxonomic diversity: general

problems: Paleobiology, v. 26, p.74‒102.

Foote, M., and Raup, D.M., 1996, Fossil preservation and the stratigraphic ranges of taxa:

Paleobiology, v. 22, p. 121‒140.

Frest, T.J., and Strimple, H.L., 1978, Manicrinus (Nov.), a cladid evolutionary homeomorph of

the bottom dwelling Hybocrinus, Brownsport (Silurian. Ludlow) of Tennessee:

Southeastern Geology, v. 19, p. 157‒175.

Gahn, F.J., and Kammer, T.W., 2002, The cladid crinoid Barycrinus from the Burlington

Limestone (Early Osagean) and the phylogenetics of Mississippian botryocrinids:

Journal of Paleontology, v. 76, p. 123–133.

121

Gavryshkina, A., Welch, D., Stadler, T., and Drummond, A. J., 2014, Bayesian inference

of sampled ancestor trees for epidemiology and fossil calibration: PLoS

Computational Biology, v. 10, e1003919.

Gavryshkina, A., Heath, T.A., Ksepka, D.T., Stadler, T., Welch, D., and Drummond,

A.J., 2015, Bayesian total evidence dating reveals the recent crown radiation of

penguins: arXiv, preprint 1506.04797

Gelman, A. and Rubin, D. B., 1992, Inference from iterative simulation using multiple

sequences: Statistical Science, v. 7, p. 457–472.

Gil Cid, M.D., Alonso, P.D, and Pobes, E.S., 1996, Reconstrucción y modo de videa de

Heviacrinus melendezi nov. gen. nov. sp. (Disparida, Iocrinidae), primer crinoide descrito

del Ordovícico medio de los Montes de Toledo (España): Revista de la Sociedad

Geológica España, v. 9, no. 1-2, p. 19‒27.

Goldfuss, G. A., 1826‒44. Petrefacta Germaniae, tam ea, Quae in Museo Universitatis Regiae

Borussicae Fridericiae Wilhelmiae Rhenanea, serventur, quam alia quaecunque in Museis

Hoeninghusiano Muensteriano aliisque, extant, iconibus et descriiptionns illustrata. --

Abbildungen und Beschreibungen der Petrefacten Deutschlands und der Angränzende

Länder, unter Mitwirkung des Hern Grafen Georg zu Münster, herausgegeben von

August Goldfuss. v. 1 (1826‒1833), Divisio prima. Zoophytorum reliquiae, p. 1‒114;

Divisio secunda. Radiariorum reliquiae, p. 115‒221 [Echinodermata]; Divisio tertia.

Annulatorium reliquiae, p. 222-242; v. 2 (1834-40), Divisio quarta. Molluscorum

acephalicorum reliquiae. I. , p. 65‒286; II. Brachiopoda, p. 287-303; III. (1841‒

122

44), Divisio quinta. Molluscorum gasteropodum reliquiae, p. 1-121: Düsseldorf, Arnz &

Co.

Goldring, W., 1923, The Devonian crinoids of the state of New York: New York State Museum,

Memoir 16, p. 1‒670.

Gorscak, E. and O‘Connor, P.M., 2016, Time-calibrated models support congruency

between continental rifting and titanosaurian evolutionary history:

Biology Letters, v. 12, p.20151047.

Guensburg, T.E., 2010, Alphacrinus new genus and origin of the disparid clade: Journal

of Paleontology, v. 84, p. 1209–1216.

Guensburg, T.E., 2012, Phylogenetic implications of the oldest crinoids: Journal of

Paleontology, v. 86, p. 455–461.

Guensburg, T.E., and Sprinkle, J., 2003, The oldest known crinoids (Early Ordovician,

Utah) and a new crinoid plate homology system: Bulletins of American

Paleontology, v. 364, 43 p.

Guensburg, T.E., and Sprinkle, J., 2009. Solving the mystery of crinoid ancestry: new fossil

evidence of arm origin and development: Journal of Paleontology, v. 83, p. 350‒364.

Guillerme, T. and Cooper, N., 2016, Effects of missing data on topological inference

using a Total Evidence approach: Molecular Phylogenetics and Evolution, v. 94,

p.146–158.

Hall, J., 1852, Palaeontology of New York, v. 2, Containing descriptions of the organic remains

of the lower middle division of the New-York system. Natural History of New York,

123

New York: Boston, D. Appleton & Co. and Wiley & Putnam; Gould, Kendall, & Lincoln,

v. 6, 362 p.

Harmon, L.J., Losos, J.B., Davies, T.J., Gillespie R.G., Gittleman J.L., et al. 2010. Early

bursts of body size and shape evolution are rare in comparative data: Evolution, v.

64, p. 2385–96.

Harrison, L. B. and Larsson, H.C.E., 2015, Among-character rate variation distributions

in phylogenetic analysis of discrete morphologic characters: Systematic Biology, v.

64, p. 307–324.

Heath, T.A., and Moore, B.R., 2014, Bayesian inference of species divergence times, in

Chen, M-H., Kuo, L., and Lewis, P.O., Bayesian Phylogenetics: Methods,

Algorithms, and Applications: CRC Press, Oxfordshire, p. 277–318.

Heath, T.A., Hedtke, S.M., and Hillis, D.M., 2008, Taxon sampling and the accuracy of

phylogenetic analyses: Journal of Systematics and Evolution, v. 46, 239–257.

Heath, T.A., Huelsenbeck, J.P. and Stadler, T., 2014, The fossilized birth–death process

for coherent calibration of divergence-time estimates: Proceedings of the National

Academy of Sciences, v. 111, p. E2957–E2966.

Heled, J., and Bouckaert, R.R., 2013, Looking for trees in the forest: summary tree from

posterior samples: BMC Evolutionary Biology, v. 13, 221.

Hemery, L.G., Roux, M., Ameziane, N., and Eleaume, M., 2013, High-resolution crinoid

phyletic inter-relationships derived from molecular data: Cahiers De Biologie

Marine, v. 54, p. 511–523.

124

Hopkins, M.J. and Smith, A.B., 2015, Dynamic evolutionary change in post-Paleozoic

echinoids and the importance of scale when interpreting changes in rates of

evolution: Proceedings of the National Academy of Sciences, v. 112, p.3758–

3763.

Huelsenbeck, J.P., Ronquist, F., and Teslenko, M., 2015, Command Reference for

MrBayes ver. 3.2.5: http://mrbayes.sourceforge.net/manual.php (accessed June

2016).

Huelsenbeck, J.P., Larget, B., Miller, R.E. and Ronquist, F., 2002, Potential applications

and pitfalls of Bayesian inference of phylogeny: Systematic Biology, v. 51),

p.673–688.

Hunt, G., 2008, Gradual or pulsed evolution: when should punctuational explanations be

preferred?: Paleobiology, v. 34, p.360–377.

Hunt, G., and Carrano, M.T, 2010, Models and methods for analyzing phenotypic

evolution in lineages and clades, in Alroy, J., and Hunt, G., (eds.), Short Course

on Quantitative Methods in Paleobiology. Paleontological Society, New Haven,

Connecticut, p. 245–269.

Hunt, G., and Slater, G., 2016, Integrating paleontological and phylogenetic approaches

to macroevolution: Annual Review of Ecology, Evolution, and Systematics, v. 47.

P. 189–213.

Hunt, G., Bell, M.A., and Travis, M.P., 2008, Evolution toward a new adaptive optimum:

phenotypic evolution in a fossil stickleback lineage: Evolution, v. 62, p. 700–710.

125

Jablonski, D., 2008, Biotic interactions and macroevolution: extensions and mismatches across

scales and levels: Evolution, v. 62, 715–739.

Jaekel, O., 1902, Über verschiedene Wege phylogenetischer Entwicklung: 5th Verhandlungen

der International Zoological-Congress Berlin, 1901, p. 1058‒1117.

Jaekel, O., 1906, Der oberste Lenneschiefer zwischen Letmathe und Iserlohn in Schmidt, W.E.,

Der oberste Lenneschiefer zwischen Letmathe und Iserlohn: Zeitschrift der Deutschen

Geologischen Geselleschaft, v. 57, p. 544.

Kammer, T.W., 2001, Phenotypic bradytely in the Costalocrinus-Barycrinus lineage of

Paleozoic cladid crinoids: Journal of Paleontology, v. 75, p. 383‒389.

Kammer, T.W. and Ausich, W.I., 1992, Advanced cladid crinoids from the middle

Mississippian of the east-central United States: primitive-grade calyces: Journal

of Paleontology, v. 66, p. 461‒480.

Kammer, T.W. and Ausich, W.I., 1996, Primitive cladid crinoids from Upper Osagean-

Lower Meramecian (Mississippian) rocks of east-central United States: Journal of

Paleontology, v. 70, p. 835‒866.

Kass, R.E., and Raftery, A.E., 1995, Bayes factors. Journal of the American Statistical

Association, v. 90, p. 773‒795.

Kier, P.M., 1952, Echinoderms of the Middle Devonian Silica Formation of Ohio: University of

Michigan Contributions from Museum of Paleontology, v. 10, p. 59‒81.

Ksepka, D.T., Parham, J.F., Allman, J.F., Benton, M.J., Carrano, M.T., Cranston, K.A.,

Donoghue, P.C., Head, J.J., Hermsen, E.J., Irmis, R.B. and Joyce, W.G., 2015. The Fossil

126

Calibration Database—A new resource for divergence dating: Systematic biology, v. 64,

p.853‒859.

Koenig, J.W., and Meyer, D.L., 1965, Two new crinoids from the Devonian of New York:

Journal of Paleontology, v. 39, p. 391‒397.

Lakner, C., van der Mark, P., Huelsenbeck, J. P., Larget, B., and Ronquist, F., 2008,

Efficiency of Markov chain Monte Carlo tree proposals in Bayesian phylogenetics:

Systematic Biology, v. 57, 86‒106.

Lane, N.G., 1970, Lower and Middle Ordovician crinoids from west-central Utah: Brigham

Young University Geology Studies, v. 17, p. 3‒17.

Lepage, T., Bryant, D., Philippe, H. and Lartillot, N., 2007, A general comparison of

relaxed molecular clock models: and Evolution, v. 24, p.

2669‒2680.

Lee, M.S.Y., and Palci, A., 2015, Morphological phylogenetics in the genomic age: Current

Biology, v. 25, p. R922‒R929.

Lee, M.S., Cau, A., Naish, D. and Dyke, G.J., 2014, Morphological clocks in

paleontology, and a mid-Cretaceous origin of crown Aves: Systematic Biology, v.

63, p. 442‒449.

Lewis, P.O., 2001, A likelihood approach to estimating phylogeny from discrete

morphological character data: Systematic biology, v. 50, p. 913‒925.

Lloyd, G.T., Wang, S.C. and Brusatte, S.L., 2012, Identifying heterogeneity in rates of

morphological evolution: discrete character change in the evolution of

(; Dipnoi): Evolution, v. 66, p.330‒348.

127

Maddison, D.R., 1991, The discovery and importance of multiple islands of most-

parsimonious trees: Systematic Zoology, v. 40, p. 315‒328.

Matzke, N.J., 2015, The evolution of antievolution policies after Kitzmiller v. Dover.

Science, p.aad4057.

Matzke, N.J. and Wright, A., 2016, Ground truthing tip-dating methods using fossil

Canidae reveals major differences in performance: Biology Letters, v. 12,

2010328, doi:10.1098/rsbl.2016.0328.

McIntosh, G.C. 1986, Phylogeny of the dicyclic inadunate order Cladida: Fourth North

American Paleontological Convention Abstracts. University of Colorado,

Boulder, p. A31.

McIntosh, G.C., 2001, Devonian cladid crinoids: Families Glossocrinidae Goldring, 1923, and

Rutkowskicrinidae new : Journal of Paleontology, v. 75, p. 783‒807.

Meek, F.B., 1871, On some new Silurian (Ordovician) crinoids and shells: American

Journal of Science, ser. 3, v. 1, p. 295‒299.

Meek, F.B., and Worthen, A.H., 1865, Descriptions of new species of crinoidea, etc. from the

Paleozoic rocks of Illinois and some of the adjoining states: Proceedings of the Academy

of Natural Sciences of Philadelphia, v. 17, p. 143‒155.

Miller, S.A., and Gurley, W.F.E., 1895, New and interesting species of Palaeozoic

fossils: Illinois State Museum, Bulletin 7, 89, p.

Moore, R.C., and Laudon, L.R., 1943, Evolution and classification of Paleozoic crinoids:

Geological Society of America Special Paper, no. 46, 151 p.

128

Moore, R.C., and Teichert, C., eds., 1978, Treatise on Invertebrate Paleontology, Part T,

Echinodermata 2: Geological Society of America and University of Kansas Press,

Lawrence, Kansas, 1027 p.

Müller, J., 1859, Bericht über die zur Bekanntmachung geeigneten Verhandlungen der

Königlich Preussischen Akademie der Wissenschaften zu Berlin: Monatsberichte

der Königlich Preussischen Akademie der Wissenschaften zu Berlin (1858), p.

185-198.

O’Meara, B.C., Ané, C., Sanderson, M.J., Wainwright P. C., and Hansen, T., 2006,

Testing for Different Rates of Continuous Trait Evolution Using Likelihood:

Evolution, v. 60, p. 922–933.

O’Reilly, J.E., Puttick, M.N., Parry, L., Tanner, A.R., Tarver, J.E., Fleming, J., Pisani, D.,

and Donoghue, P.C.J., 2016, Bayesian methods outperform parsimony but at the

expense of precision in the estimation of phylogeny from discrete morphological

data: Biology Letters, v. 12, 20160081, http://dx.doi.org/10.1098/rsbl.2016.0081

Paradis, E., Claude, J. and Strimmer, K., 2004, APE: analyses of phylogenetics and

evolution in R language: , v. 20, p. 289–290.

Pennell, M.W., Harmon, L.J., Uyeda, J.C., 2014, Is there room for punctuated

equilibrium in macroevolution?: Trends in Ecology and Evolution, v. 29, p. 23‒

32.

Phillips, J., 1839, Encrinites and zoophytes of the Silurian System, in Murchison, R.T.,

ed., The Silurian System, p. 670‒675.

129

Pollitt, J.R., Fortey, R.A. and Wills, M.A., 2005. Systematics of the families

Lichidae Hawle and Corda, 1847 and Lichakephalidae Tripp, 1957: The

application of Bayesian inference to morphological data: Journal of Systematic

Palaeontology, v. 3, p. 225‒241.

Pyron, R.A., 2011, Divergence Time Estimation Using Fossils as Terminal Taxa and the

Origins of Lissamphibia: Systematic Biology, v. 60, p. 466‒481.

Pyron, R.A., 2015, Post-molecular systematics and the future of phylogenetics: Trends in

ecology and evolution, v. 30, p. 384‒389.

Rabosky, D.L., 2009, Ecological limits and diversification rate: alternative paradigms to

explain the variation in species richness among clades and regions: Ecology

Letters, v. 12, p. 735–743.

Rabosky, D.L., and McCune, A.R., 2009, Reinventing species selection with molecular

phylogenies: Trends in Ecology and Evolution, v. 25, p. 68–74.

Rannala, B., and Yang, Z., 1996, Probability distribution of molecular evolutionary trees:

a new method of phylogenetic inference: Journal of , v. 43, p.

304–311.

Rambaut, A., and Drummond, A.J., 2015, TreeAnnotator v2 2.1: MCMC ouput analysis.

Rambaut, A. 2014, Summarizing posterior trees: http://beast.bio.ed.ac.uk/summarizing-

posterior-trees (accessed June 2016).

Raup, D. M., 1985, Mathematical models of cladogenesis: Paleobiology, v. 11, p. 42–52.

130

Raup, D. M., Gould, S. J., Schopf, T. J. M., and Simberloff, D. S., 1973, Stochastic

models of phylogeny and the evolution of diversity: Journal of Geology, v. 81, p.

525–542.

Robinson, D.F., and Foulds, L.R., 1981, Comparison of phylogenetic trees: Mathematical

Biosciences, v. 53, p. 131–147.

Ronquist, F., Klopfstein, S., Vilhelmsen, L., Schulmeister, S., Murray, D.L. and

Rasnitsyn, A.P., 2012, A total-evidence approach to dating with fossils, applied to

the early radiation of the : Systematic Biology, v. 61, p. 973–999.

Rouse, G.W., Jermiin, L.S., Wilson, N.G., Eeckhaut, I., Lanterbecq, D., Oji, T., Young,

C.M., Browning, T., Cisternas, P., Helgen, L.E., Stuckey, M., and Messing, C.G.,

2013, Fixed, free, and fixed: the fickle phylogeny of extant Crinoidea

(Echinodermata) and their Permian–Triassic origin: Molecular Phylogenetics and

Evolution v. 66, p. 161–181.

Salter, J.W., 1873, A catalogue of the collection of Cambrian and Silurian fossils

contained in the geological museum of the University of Cambridge: Cambridge,

Cambridge University Press, 204 p.

Sepkoski, J. J., Jr. 1981, A factor analytic description of the Phanerozoic marine fossil

record: Paleobiology, v. 7, p. 36–53.

Simpson, G.G., 1944, Tempo and Mode in Evolution: Columbia University Press, New

York.

131

Simms, M.J., 1988, The phylogeny of post-Palaeozoic crinoids, in Burke, R. D., P. V.

Mladenov, P. Lambert, and R. L. Parsley, eds, Echinoderm Biology: Balkema,

Rotterdam, p. 97–102.

Simms, M.J., and Sevastopulo, G.D., 1993, The origin of articulate crinoids:

Palaeontology, v. 36, p. 91–109.

Schlotheim, E.F. von, 1820, Die Petrefactenkunde auf ihrem jetzigen Standpunkte durch

die Beschreibung seiner Sammlung versteinerter und fossiler Überreste des Thier-

und Pflanzenreichs der Vorwelt erläutert: Gotha, Beckersche Buchhandlung, p.

437 p.

Schultze, L., 1867, Monographie der Echinodermen des Eifler Kalkes: Denkschriften der

Kaiserlich Akademie der Wissenschaften Mathematisch-Naturwissenschaftlichen

Classe, Wien, v. 26, no. 2, p. 113‒230.

Sevastopulo, G.D., and Lane, N.G., 1988, Ontogeny and phylogeny of disparid crinoids,

in Paul, C.R.C., and Smith, A.B., eds, Echinoderm phylogeny and evolutionary

biology: Oxford, Clarendon Press, p. 245–253.

Slater, G.J., 2013, Phylogenetic evidence for a shift in the mode of mammalian body size

evolution at the Cretaceous-Palaeogene boundary: Methods in Ecology and

Evolution, v. 4, p. 734‒744.

Slater, G.J., 2015, Iterative adaptive radiations of fossil canids show no evidence for

diversity-dependent trait evolution: Proceedings of the National Academy of

Sciences, v. 112, p. 4897‒4902.

132

Slater, G.J., and Harmon, L.J., 2013, Unifying fossils and phylogenies for comparative

analyses of diversification and trait evolution: Methods in Ecology and Evolution,

v. 4, p. 699‒702.

Smith, A.B., 1984, Classification of the Echinodermata: Palaeontology, v. 27, p. 431‒

459.

Smith, A.B., 1994, Systematics and the Fossil Record: Documenting Evolutionary

Patterns: Blackwell Science, Oxford, 223 p.

Smith, A.B., Lafay, B., and Christen, R., 1992, Comparative variation of morphological

and molecular evolution through geologic time: 28S ribosomal RNAversus

morphology in echinoids: Philosophical Transactions: Biological Sciences, v.

338, p. 365‒382.

Snively, E., Russell, A.P. and Powell, G.L., 2004, Evolutionary morphology of the

coelurosaurian arctometatarsus: descriptive, morphometric and phylogenetic

approaches: Zoological Journal of the Linnean Society, v. 142, p. 525‒553.

Spencer, M.R., and Wilberg, E.W., 2013. Efficacy or convenience? Model‐based

approaches to phylogeny estimation using morphological data: Cladistics, v. 29,

p. 663‒671.

Springer, F., 1911, On a Trenton echinoderm fauna: Canada Department Mines Memoir

no. 15-P, 70 p.

Springer, F., 1920, The Crinoidea Flexibilia: Smithsonian Institution Publication no.

2501, 486 p.

133

Sprinkle, J., 1982, Hybocrinus, in Sprinkle, J., ed., Echinoderm from the Bromide

Formation (Middle Ordovician) of Oklahoma: The University of Kansas Paleontological

Contributions, Monograph 1, p. 119–128.

Sprinkle, J., and Kolata, D. R., 1982, "Rhomb-bearing" camerate, in Sprinkle, J., ed.,

Echinoderm faunas from the Bromide Formation (Middle Ordovician) of

Oklahoma: University of Kansas Paleontological Contributions, Monograph 1, p.

206‒211.

Sprinkle, J., and Moore, 1978, Hybocrinida, in Moore, R. C. and Teichert, C., eds:

Treatise on Invertebrate Paleontology, Pt. T, Echinodermata 2: Geological Society

of America and University of Kansas Press, Kansas, p. T564–T574.

Stadler, T., 2010, Sampling-through-time in birth-death trees: Journal of Theoretical

Biology, vol. 267, p. 396‒404.

Stadler, T. and Yang, Z., 2013, Dating phylogenies with sequentially sampled tips:

Systematic Biology, v. 62, p. 674‒688.

Stadler, T., Kouyos, R., Wyl, V. von, Yerly, S., Böni, J., Bürgisser, P., Klimkait, T., Joos,

B., Rieder, P., Xie, D., Günthard, H. F., Drummond, A. J., Bonhoeffer, S., the Swiss

HIV Cohort Study, 2012, Estimating the basic reproductive number from viral

sequence data: Molecular Biology and Evolution, v. 29, p. 347‒357.

Summers, M.M., Messing, C.G., and Rouse, G.W., 2014, Phylogeny of

(Echinodermata: Crinoidea: ): a new classification and an assessment

of morphological characters for crinoid taxonomy: Molecular Phylogenetics and

Evolution, v. 80, p. 319–339. 134

Swofford, D.L., 2002, PAUP*: Phylogenetic Analysis Using Parsimony (*and Other

Methods): Version 4. Massachusetts, Sinauer Associates.

Swofford, D.L., Olsen, G.J., Waddell, P.J., and Hillis, D.M., 1996, Phylogenetic

inference, in Hillis, D.M., Moritz, C., and Mable, B.K., ed., Molecular Systematics

(second edition), Massachusetts, Sinauer, p. 407‒514.

Ubaghs, G., 1969, Aethocrinus moorei Ubaghs N. Gen., N. sp., le Plus ancien Crinoide

dicyclique Connu: University of Kansas Paleontological Contributions, Paper 38, p. 1‒

25.

Ulrich, E.O., 1925, The lead, zinc, and fluorspar deposits of Western Kentucky; Chapter

2, Stratigraphic geology: U. S. Geological Survey, Professional Paper 36, p. 22‒

71.

Ubaghs, G., 1978, Origin of crinoids, in Moore, R. C. and Teichert, C., eds, Treatise on

Invertebrate Paleontology, Pt. T, Echinodermata, v. 2 (2): Geological Society of

America and University of Kansas Press, Lawrence, p. T275–T281.

Wagner, PJ. 1998, A likelihood approach for evaluating estimates of phylogenetic

relationships among fossil taxa: Paleobiology, v. 24, p. 430‒49.

Wagner, P.J., 1999, The utility of fossil data in phylogenetic analyses: a likelihood

example using Ordovician-Silurian species of the Lophospiridae (:

Murchisoniina): American Malacological Bulletin, v. 15, p. 1‒32.

Wagner, P.J., 2000a, Phylogenetic analyses and the fossil record: tests and inferences,

hypotheses, and models: Paleobiology, v. 26, p. 341‒371.

135

Wagner, P.J., 2000b, The quality of the fossil record and the accuracy of phylogenetic

inferences about sampling and diversity: Systematic Biology, v. 49, p. 65‒86.

Wagner, P.J., 2000c, Exhaustion of morphologic character states among fossil taxa:

Evolution, v. 54, 365‒386.

Wagner, P.J., 2012, Modelling rate distributions using character compatibility:

implications for morphological evolution among fossil invertebrates: Biology

Letters, v. 8, p. 143‒146.

Wagner, P.J. and Marcot, J.D., 2010, Probabilistic phylogenetic inference in the fossil

record: current and future applications, in Alroy, J. and Hunt, G., eds.,

Quantitative Methods in Paleobiology. The Paleontological Society Papers, v. 16,

p. 189‒211

Wagner, P.J. and Marcot, J.D., 2013, Modelling distributions of fossil sampling rates

over time, space and taxa: assessment and implications for macroevolutionary

studies: Methods in Ecology and Evolution, v. 4, p.703‒713.

Walcott, C.D., 1884. Descriptions of new species of fossils from the Trenton Group of

New York: New York State Museum of Natural History, Annual Report, v. 35, p.

207‒214.

Webster, G.D., 2003, Bibliography and index of Paleozoic crinoids, coronates, and

hemistreptocrinoids: Geological Society of America Special Paper 363.

Webster, G.D., and Jell, P.A., 1999, New Permian crinoids from Australia: Memoirs of

the Queensland Museum, v. 33, p. 349–359.

136

Webster, G.D. and Maples, C.G., 2006, Cladid crinoid (Echinodermata) anal conditions:

a terminology problem and proposed solution: Palaeontology, v. 49, p. 187–212.

Webster, G.D. and Maples, C.G., 2008, Cladid crinoid radial facets, brachials, and arm

: a terminology solution for studies of lineage, classification, and

paleoenvironment, in Ausich, W.I., and Webster, G.D., eds., Echinoderm

Paleobiology. Indiana University Press, Bloomington, p.197–226.

Webster, G.D., Maples, C.G., Mawson, R., and Dastanpour, M., 2003, A cladid-dominated Early

Mississippian crinoid and fauna from Kerman Province, Iran and revision of

the glossocrinids and rhenocrinids: Journal of Paleontology, Memoir 60, v. 77, suppl. to

no. 3, 35 p.

Weller, S. and Davidson, A.D., 1896. Petalocrinus mirabilis (n. sp.) and a new American

fauna: Journal of Geology, v. 4, p. 166‒173.

Wetherby, A.G., 1880, Remarks on the Trenton Limestone of Kentucky, with descriptions of

new fossils from that formation and the Kaskaskia (Chester) Group, Sub-carboniferous:

Journal of the Cincinnati Society of Natural History, v. 3, p. 144‒160.

Wheeler, W.C., 2012, Systematics: a course of lectures: Chichester, John Wiley and

Sons. 426 p.

Wheeler, W.C., and Pickett, K.M., 2007, Topology-Bayes versus Clade-Bayes in

phylogenetic analysis: Molecular Biology and Evolution, v. 25, p. 447‒453.

Wright, A.M., and Hillis, D.M., 2014, Bayesian analysis using a simple likelihood model

outperforms parsimony for estimation of phylogeny from discrete morphological

data: PloS One, v. 9, e109210.

137

Wright, D. F., 2015, Fossils, homology, and “Phylogenetic Paleo-ontogeny”: a

reassessment of primary posterior plate homologies among fossil and living

crinoids with insights from developmental biology: Paleobiology, v. 41, p. 570‒

591.

Wright, D.F., and Ausich, 2015, W.I., From the stem to the crown: phylogeny and

diversification of pan-cladid crinoids, in Zamora, S. and Rábano, I., eds., Progress

in Echinoderm Paleobiology: Cuadernos del museo Geominero, 19, Instituto

Geológico y Minero de España, p. 199–202.

Wright, D. F. and Stigall, A. L., 2013, Phylogenetic revision of the Late Ordovician

orthid brachiopod genera Plaesiomys and Hebertella from Laurentia: Journal of

Paleontology, v. 87, p. 1107‒1128.

Wright, D.F., Ausich, W.I., Cole, S.R., Peter, M.E., Rhenburg, E.C., in press,

Phylogenetic taxonomy and classification of the Crinoidea: Journal of

Paleontology.

Xie, W., Lewis, P.O., Fan, Y., Kuo, L., and Chen, M.–H., 2010, Improving marginal

likelihood estimation for Bayesian phylogenetic model selection: Systematic Biology,

v. 60, p. 150‒160.

Yang, Z., 2014, Molecular Evolution: a Statistical Approach: Oxford University Press,

Oxford, 492 p.

Zhang, C., Stadler, T., Klopfstein, S., Heath, T.A. and Ronquist, F., 2016, Total-Evidence

Dating under the Fossilized Birth–Death Process: Systematic Biology, v. 65, p.,

228‒249.

138

Table 3.1 Species sampled for phylogenetic analysis.

139

Species Author Aethocrinus moorei Ubaghs, 1969 Alphacrinus mansfieldi Guensburg, 2010 Webster, Maples, Mawson, and Dastanpour, Amabilicrinus iranensis 2003 Apektocrinus ubaghsi Guensburg and Sprinkle, 2009 Botryocrinus ramosissimus Angelin, 1878 Carabocrinus radiatus Billings, 1857 Codiacrinus granulatus Schultze, 1867 Colpodecrinus quadrifidus Sprinkle and Kolata, 1982 Corematocrinus plumosus Goldring, 1923 Crotalocrinites verucosus (Schlotheim, 1820) Cupulocrinus humilis (Billings, 1857) Dendrocrinus longidactylus Hall, 1852 Euspirocrinus spiralis Angelin, 1878 Eustenocrinus springeri Ulrich, 1925 Gasterocoma antiqua Goldfuss, 1839 Glossocrinus naplesensis Goldring, 1923 Heviacrinus melendezi Gil Cid, Alonso, and Pobes, 1996 Homalocrinus nanus (Salter, 1873) Hybocrinus conicus Billings, 1857 Hybocystites problematicus Wetherby, 1880 Ibexocrinus lepton Lane, 1970 Icthyocrinus laevis Conrad, 1842 Iocrinus subcrassus (Meek and Worthen, 1865) Lecanocrinus macropetalus Hall, 1852 Lecythocrinus eifelianus Müller, 1859 Manicrinus hybocriniformis Frest and Strimple, 1978 Mastigocrinus arboreus (Salter, 1873) Merocrinus typus Walcott, 1884 Metabolocrinus rossicus Jaekel, 1902 Ottawacrinus typus Billings, 1887 Petalocrinus mirabilis Weller and Davidson, 1896 Plicodendrocrinus casei Meek, 1871 Porocrinus conicus Billings, 1857 Proctothylacocrinus longus Kier, 1952

Continued

140

Table 3.1 continued

Protaxocrinus ovalis (Angelin, 1878) Rhenocrinus ramoisissimus Jaekel, 1906 Rutkowskicrinus patriciae McIntosh, 2001 Sagenocrinites expansus (Phillips, 1839) Sphaerocrinus geometricus (Goldfuss, 1831) Streblocrinus brachiatus Koenig and Meyer, 1965 Thalamocrinus ovatus Miller and Gurley, 1895 Thenarocrinus callipygus Bather, 1890

141

Figure 3.1 Illustration of the fossilized birth‒death process. (1) A full realization of the FBD process from time t in the past to the end of the process. Diversification produces a tree with branching and extinction events and random sampling of nodes.

Sampling events are indicated by dots. (2) A “reconstructed” phylogenetic produced by pruning all of the unsampled lineages in 1.1. Thus, only the observed portion of samples participating in the macroevolutionary process is observed. Note that some sampled nodes represent ancestors.

142

Fig. 3.1

143

Figure 3.2 Parsimony-based estimate of rate variation among characters. This distribution suggests that many characters evolve slowly, whereas a small number of characters evolve at much higher rates.

144

Fig. 3.2

145

Figure 3.3 Testing statistical associations between the amount of morphologic

evolution inferred from an undated analysis and divergence times estimated under a

relaxed morphologic-clock model: (1) node age in millions of years and anagenesis

(measured as the total root-to-tip distance in parsimony steps from an undated analysis),

(2) phylogenetic independent contrasts between branch durations (from time-scaled

analysis) and anagenesis (from an undated analysis). Both regressions are statistically

significant.

146

Fig. 3.3

147

Figure 3.4 Maximum Clade Credibility tree of early to middle Paleozoic crinoids.

Posterior probabilities (> 0.50) are located next to nodes and expressed in percent, blue node bars represent the 95% highest posterior density age estimates, thick black bars represent genus-level stratigraphic ranges. Note the inclusion of stratigraphic ranges give the appearance of sister taxon relationships where zero-length branches were sampled

(e.g., Cupulocrinus). The Cladida (sensu Simms and Sevastopulo, 1993) are sister to the

Disparida. Crinoid taxa depicted represent the major clades discussed in the text: (from top to bottom), the disparid Eustenocrinus springeri Ulrich, 1925, redrawn from Ubaghs

(1978); representative porocrinoid and hybocrinid (see text), Hybocrinus conicus

Billings, 1857, redrawn from Sprinkle and Moore (1978); the flexible Protaxocrinus laevis (Billings, 1857), redrawn from Springer (1911); representative eucladid

Dictenocrinus, redrawn from Bather (1900).

148

Fig. 3.4

149

Chapter 4: Phylogenetic taxonomy and classification of the Crinoidea (Echinodermata)

Abstract.—A major goal of biological classification is to provide a system that conveys phylogenetic relationships while facilitating lucid communication among researchers. Phylogenetic taxonomy is a useful framework for defining clades and delineating their taxonomic content based on well-supported phylogenetic hypotheses.

The Crinoidea (Echinodermata) is one of the five major clades of living echinoderms and has a rich fossil record spanning nearly a half billion years. Using principles of phylogenetic taxonomy and recent phylogenetic analyses, we provide the first phylogeny- based definition for the Clade Crinoidea and its constituent subclades. A series of stem- and node-based definitions are provided for all major taxa traditionally recognized within the Crinoidea, including the Camerata, Disparida, Hybocrinida, Cladida, Flexibilia, and the Articulata. Following recommendations proposed in recent revisions, we recognize several new clades, including the Eucamerata Cole (in press), Porocrinoidea Wright (in press), and Eucladida Wright (in press). In addition, recent phylogenetic analyses support the resurrection of two names previously abandoned in the crinoid taxonomic literature: the Pentacrinoidea Jaekel, 1918 and Inadunata Wachsmuth and Springer, 1885. Lastly, a phylogenetic perspective is used to inform a comprehensive revision of the traditional rank-based classification. Although an attempt was made to minimize changes to the rank-based system, numerous changes were necessary in some cases to achieve 150

monophyly. These phylogeny-based classifications provide a useful template for

paleontologists, biologists, and non-experts alike to better explore evolutionary patterns and processes with fossil and living crinoids.

Introduction

Crinoids are a diverse, long-lived clade of echinoderms with a fossil record spanning

nearly half a billion years and are represented by more than 600 species living in marine

ecosystems today (Hess et al., 1999). The geologic history of crinoids is revealed through

a highly complete, well-sampled fossil record (Foote and Raup, 1996; Foote and

Sepkoski, 1999) displaying a complex pageant of evolutionary radiation, extinction,

ecologic innovation, and morphologic diversification (Ausich and Bottjer, 1982; Ausich

et al., 1994; Foote, 1999; Peters and Ausich, 2008; Deline and Ausich, 2011; Gorzelak et

al, 2015). The spectacular fossil record of crinoids is greatly enriched and complemented

by detailed biologic studies on living species. These studies facilitate opportunities to

synthesize information from fossil and extant forms. For example, comparative studies

between fossil and living crinoid species have provided insight into species ecology and

niche dynamics (Meyer and Macurda, 1977; Ausich, 1980; Roux, 1987; Kitazawa et al.,

2007; Baumiller 2008), established developmental bases for morphologic homologies

(Shibata et al., 2008; Wright, 2015a), and informed phylogenetic hypotheses (Simms and

151

Sevastopulo, 1993; Rouse et al., 2013). Thus, crinoids form a data-rich model system for exploring major questions in the .

Given their general significance and broad scientific utility across multiple disciplines of inquiry, it is paramount that the biological classifications of crinoids reflect their evolutionary heritage. Numerous emendations and informal suggestions for major taxonomic revisions have been opined over the last few decades (e.g., Kelly, 1986;

Simms and Sevastopulo, 1993; Ausich, 1998a, 1998b; Webster and Jell, 1999; Hess and

Messing, 2011), but the most recent comprehensive revision to crinoid classification is the 1978 Treatise on Invertebrate Paleontology (Moore and Teichert, 1978). Since publication of the Treatise, the value of revising rank-based systematic classifications to be consistent with phylogenetic hypotheses and/or the explicit use of phylogenetic taxonomy (sensu de Quieroz and Gauthier, 1990; Sereno, 1999, 2005) has become increasingly common in paleontology (e.g., Smith, 1984, 1994; Holtz, 1996, 1998;

Sereno, 1997; Padian et al., 1999; Brochu and Sumrall, 2001; Carlson, 2001; Carlson and

Leighton, 2001; Brochu, 2003; Forey et al., 2004; Sereno et al., 2005; Butler et al., 2008;

Kelley et al., 2013). We agree with these authors that all named taxa in a biological classification system should ideally represent clades (i.e., monophyletic groups). The development of phylogeny-based classifications is not without difficulties or criticism

(e.g., Benton, 2000; 2007). However, we advocate that recent advances in understanding the phylogenetic relationships of major crinoid lineages make the biological classification of the Crinoidea ripe for revision.

152

A great strength of so-called “phylogenetic taxonomy” is its potential for

increasing nomenclatural stability (de Quieroz and Gauthier, 1994; Brochu and Sumrall,

2001). Under a phylogeny-based system of classification, groups of taxa are organized by

their patterns of shared common ancestry rather than diagnostic traits. This is a

particularly useful aspect of phylogenetic taxonomy: if named evolutionary units are

defined by their history of common ancestry, they do not change if new information

comes to light that necessitates modification of taxonomic diagnoses. For example, new

fossil discoveries and/or more nuanced understandings of phylogenetic relationships may

alter the distribution of synapomorphies among members of a clade but do not alter the

definition of the clade. Moreover, by naming taxa based on cladogram topologies,

phylogenetic taxonomy can provide a precise definition for groups previously difficult to

diagnose by a unique combination of synapomorphies, such as the Articulata (Simms,

1988; Webster and Jell, 1999; Rouse et al., 2013). To avoid potential instability in

taxonomic nomenclature and/or the proliferation of clade names, we advocate that major

changes in crinoid systematics should (1) be based on well-supported phylogenetic

hypotheses inferred using rigorous and repeatable quantitative techniques and (2) employ

widely used names and/or names with historical precedence if available.

In this paper, we propose a series of stem-based and node-based clade definitions to help standardize nomenclature for crinoid higher taxa. The clade definitions proposed herein are informed by a series of recent phylogenetic analyses (Ausich et al., 2015; Cole, in press; Wright, in press) and represent the first attempt to classify crinoids using the principles of phylogenetic taxonomy (de Queiroz and Gauthier 1992, 1994).

153

Although Linnaean classifications lack rigorous criteria for assigning ranks, they

can nevertheless provide useful (if coarse) reflections of phylogenetic relatedness and

divergence among taxa, particularly in paleontology (Smith, 1984; Potter and

Freudenstein, 2005; Jablonski and Finarelli, 2009; Soul and Friedman, 2015). Given the

widespread use of rank-based classifications among invertebrate paleontologists in both

alpha taxonomy and paleobiological studies, it is prudent to present a phylogenetically

informed revision of the rank-based classification of the Crinoidea. These revisions

modify the existing Linnaean classification of crinoids to better represent the set of

nested hierarchies implied by phylogenetic trees (Ausich et al., 2015; Cole, in press;

Wright, in press).

In their review of progress made in crinoid research during the 20th century,

Ausich and Kammer (2001, p. 1167) stated the “immediate challenge for the [21st

century] study of crinoids is to establish a phylogenetic classification for the entire class.”

It is our hope that the dual classification systems presented herein will provide a

foundation for future studies employing phylogenetic nomenclature in crinoid research

and also promote the use of an improved classification system among researchers who

choose to work with the Linnaean system.

154

The dredge and the hammer: a brief history of crinoid classification

“The whole history of the attempts to classify the Crinoidea shows…the gradual emancipation from the older habit of lumping forms together because they are alike in structure without considering how the likeness arose” –F.A. Bather (1898 , p. 339)

Formal scientific description and classification of crinoids began in 1821 when

J.S. Miller recognized fossilized stalked echinoderms from the “environs of Bristol” as a distinct group. Although he did not include comatulids in his original conception of the

Crinoidea, he anticipated that they were crinoids: “The combination of these results with those from the Crinoidea made me anxious to examine the Comatulae…an which would be defined with sufficient precision as a Pentacrinus destitute of the column

(Miller, 1821, p. 127).” Further, he judged Marsupites ornatus Miller, 1821 (an unstalked crinoid of Cretaceous age) to be the “link” between comatulids and his Crinoidea (Miller,

1821, p. 139). Extant stalked crinoids were unknown until the mid- to late 1860s, when their discovery during oceanic dredging expeditions provided fodder for early debates regarding the efficacy of Darwin’s (1859) then recently proposed theory of natural selection (see Alaniz, 2014; Etter and Hess, 2015). Thus, the original description, definition, and diagnosis of the Crinoidea relied entirely on fossil remains. Despite the morphological diversity and deep phylogenetic relationships among groups of extant species, the inclusion of living crinoids with fossil forms has not fundamentally altered

Miller’s (1821) concept. Following subsequent inclusion of the comatulids and extant

155

stalked crinoids with fossil forms, the Crinoidea has withstood nearly 200 years of

scrutiny as a distinct group within the Echinodermata.

In contrast with their long-term recognition as a clade, the classification of taxa

within the Crinoidea has been widely debated since the 19th century (Müller, 1841;

Angelin, 1878; Wachsmuth and Springer, 1897; Bather, 1899; Springer, 1913; Jaekel,

1918). With few exceptions, debates on crinoid classification have primarily been based on disagreements over phylogenetic affinities among taxa rather than systematic practices among researchers (see Bather, 1899 for a counter example). The intensity of early debates over crinoid classification is best epitomized by the frequent yet acrimonious exchanges between Wachsmuth and Springer (e.g., 1885, 1891, 1897) and Bather (e.g.,

1898, 1899, 1900). Attempts to resolve these debates among 19th century systematists

have largely shaped the last ~70 years of crinoid research (Ausich and Kammer, 2001).

In their seminal work Evolution and Classification of Paleozoic Crinoids, Moore

and Laudon (1943) presented a classification that incorporated aspects of both Frank

Springer’s and Francis Bather’s ideas (see discussion in Ausich and Kammer, 2001).

With few modifications, Moore and Laudon’s (1943) publication formed the basis for the

Treatise on Invertebrate Paleontology (Moore and Teichert, 1978). Following

publication of the 1978 Treatise, the classification of crinoids entered a protracted yet

frail era of nomenclatural stability. Although few authors have advanced major revisions

or comprehensive modifications, many have voiced contention with the Treatise

classification (Kelly, 1982, 1986; Kolata, 1982; McIntosh, 1984, 1986, 2001; Ausich,

1986, 1998a, 1998b; Donovan, 1988; Simms, 1988; Simms and Sevastopulo, 1993;

156

Brower, 1995; Webster and Jell, 1999; Guensburg and Sprinkle, 2003; Hess and Messing,

2011; Guensburg, 2012; Ausich et al., 2015). With the exception of Simms and

Sevastopulo (1993), all of these studies have basically been readjustments of the Moore

and Teichert (1978) classification to accommodate rank changes, the addition of new

groups, and delineation of clade membership defined by phylogenetic studies of extant

species.

The study of extant crinoids remains in the shadow of A. H. Clark, who published

more than 100 publications on their morphology, taxonomy, and classification during the

early to middle 20th century (e.g., A. H. Clark, 1915, 1921; A. H. Clark and A. M. Clark,

1967). The advent and application of molecular phylogenetic methods to crinoid

phylogeny has recently thrown light on relationships among extant species (Cohen et al.,

2004; Hemery et al., 2013; Rouse et al., 2013; Summers et al., 2014). However, these analyses also point toward the need for extensive taxonomic revisions and an improved understanding of morphologic traits among living species (Messing and White, 2001;

David et al., 2006; Roux et al., 2013; Summers et al., 2014; Hays et al., 2015).

Remarkably, there has been little previous work to combine molecular phylogenetic studies of extant crinoids with paleontologic data to assemble a more complete picture of post-Paleozoic crinoid evolutionary history. Efforts to integrate these rich sources of information present both challenges and opportunities for future researchers to resolve patterns and processes shaping the crinoid tree of life (Lee and Palci, 2015; Pyron, 2015).

157

Crinoid origins and classification

Extant echinoderms include the Crinoidea, Echinoidea, Ophiuroidea, Asteroidea, and the

Holothuroidea, with the latter four comprising the . Although it has been

long established that crinoids form the sister group to the Eleutherozoa, the relationships

among many fossil and extant echinoderm groups are controversial (Paul and Smith,

1984; Sumrall 1997; David et al., 2000; Smith, 2005; Pisani et al., 2012; Telford et al.,

2014; Zamora and Rahman, 2014; Feuda and Smith, 2015; Reich et al., 2015). The phylogenetic position of crinoids within the Echinodermata was contested throughout the late 20th century, with a focal question whether the Pelmatozoa (i.e., stalked echinoderms including blastozoans and crinoids) and/or the Blastozoa are monophyletic groups or a

“grade” of body plan organization. This is a fundamental question not only for

understanding the origin of crinoids but also for resolving phylogenetic relationships

among clades within the Echinodermata. One hypothesis of crinoid origins postulates that

crinoids and blastozoan echinoderms independently evolved pelmatozoan-grade body

plans (e.g., Sprinkle, 1973, 1976; Mooi and David, 1998, 2008; Guensburg and Sprinkle,

2003; David et al., 2000; Guensburg, 2012). This hypothesis proposes that blastozoans

and crinoids each comprise distinct monophyletic groups. In contrast, an alternative

hypothesis postulates that blastozoans and crinoids are members of an inclusive

pelmatozoan clade, with crinoids nested within a paraphyletic Blastozoa (Leuckart, 1848;

Bather, 1899, 1900; Smith, 1984; Paul and Smith, 1984; Paul, 1988; Smith and Jell,

1990; Smith, 1994; Sumrall, 1997; Ausich, 1998a, 1998b; Clausen et al., 2009; Zamora

158

and Smith, 2011; Kammer et al., 2013; O’Malley, 2016). In this hypothesis, the

blastozoan body plan represents a grade of organization within the more inclusive

Pelmatozoa, a clade comprising all blastozoan grade echinoderms and crinoids (including

the ). Although the inclusive group of nominal ‘blastozoan’ taxa is not

monophyletic, there are undoubtedly assemblages of ‘blastozoan’ taxa that do correspond

to monophyletic groups (Smith, 1984; Sumrall and Wray, 2007; Zamora and Smith,

2011; Sumrall and Waters, 2012; Zamora et al., 2016).

Important to this debate are the differences among researchers with respect to their underlying taxonomic principles and systematic practices (see Smith, 1988). Those who support the monophyly of the Blastozoa and Crinoidea embrace systematic practices that emphasize differences (rather than similarities) among taxa, recognize plesiomorphic traits as taxonomically informative characters, exclude character data from consideration of relationships because of a priori beliefs regarding the distribution of homoplastic traits, and conflate sister group hypotheses with ancestor-descendant relationships (e.g.,

Guensburg and Sprinkle, 2003, 2007; Guensburg, 2012; Guensburg et al., 2016). These practices differ considerably from those who infer the Pelmatozoa as a clade. These workers tend to emphasize similarities (rather than differences) among taxa, minimize a priori assumptions regarding hypotheses of character evolution, and utilize the principles of phylogenetic systematics to rigorously test whether apparent similarities in form reflect synapomorphies or homoplasy (e.g., Sumrall and Waters, 2012; Sumrall, 2014;

Ausich et al., 2015). Given the recent advances in homology assessment among pentaradiate echinoderms (e.g., Sumrall, 1997, 2008, 2010, 2014; Sumrall and Waters,

159

2012; Kammer et al., 2013) and computational phylogenetic analyses of echinoderm taxa

based on a large ensemble of characters, it is becoming increasingly clear that a

blastozoan-grade taxon likely forms the closest immediate outgroup to the Crinoidea

(Kammer et al., 2013; Sumrall, 2014). In the future, new developments in phylogenetic research along with a continued search for the oldest “crinoid” fossils will continue to play a role in uncovering the sequence of morphologic transitions behind the assembly of the crinoid body plan.

Despite desultory disagreements regarding crinoid origins (Sprinkle, 1973;

Ubaghs, 1978; Donovan, 1988; Ausich, 1998a, 1998b; Ausich and Babcock, 1998;

Guensburg and Sprinkle, 2007, 2009; Kammer et al., 2013; Guensburg, 2012; Ausich et al., 2015; Guensburg et al., 2016), there is nevertheless considerable agreement among workers regarding the pattern of branching relationships within the crinoid ingroup. For example, the recent phylogenetic analyses of Guensburg (2012) and Ausich et al. (2015) reveal highly congruent patterns of branching relationships among crinoid higher taxa despite the use of alternative outgroups, different datasets, and alternative interpretations of homologous morphologic characters. We surmise this growing consensus stems from the improved taxonomic sampling of the oldest known crinoids (Guensburg and Sprinkle,

2003, 2009; Guensburg, 2010) and implementation of more rigorous quantitative approaches to testing phylogenetic hypotheses (Guensburg, 2012; Ausich et al., 2015;

Cole, in press; Wright, in press).

We conclude that congruence observed among tree topologies obtained from researchers with different perspectives indicates strong support for these patterns.

160

Although questions surrounding crinoid origins remain, this debate is moot with respect

to the phylogeny-based definitions and classification presented herein and ultimately has no bearing on the focus and conclusions of this paper.

Toward a phylogenetic classification of the Crinoidea

From the perspective of their geologic history, crinoids are a bottom-heavy clade (Gould

et al., 1987). In contrast to the tremendously diverse assemblage of stem lineages,

comparatively few species are encompassed within the crown group (Fig. 4.1). Because

of the enormous diversity of the stem group relative to the crown group, fossil crinoids

have received much systematic attention compared to their extant representatives (but see

A. H. Clark, 1915; David et al., 2006; Hess and Messing, 2011; Rouse et al., 2013;

Hemery et al., 2013). Aside from a number of smaller studies examining relationships

among species of middle to late Paleozoic genera (e.g., Ausich and Kammer, 2008; Gahn

and Kammer, 2002; Kammer and Gahn, 2003), most investigations of crinoid phylogeny

have focused on discerning relationships among Ordovician taxa (Brower, 1995; Ausich

1998b; Guensburg 2012; Ausich et al. 2015; Cole, in press). The Ordovician Period

represents a key interval in crinoid evolution because species belonging to various groups

of traditionally named taxa first appear in rocks of the Lower Ordovician ()

(Guensburg and Sprinkle, 2003, 2009; Guensburg, 2010) and the majority of well-studied

groups had originated prior to its close.

161

The divergence between camerate and non-camerate lineages forms a fundamental, early split in the history of crinoid evolution (Jaekel, 1918; Donovan, 1988;

Guensburg, 2012; Ausich et al., 2015; Cole in press; Wright, in press) (Fig. 1). For example, in the recent phylogeny of Ausich et al. (2015), taxa belonging to the Camerata

(sensu Moore and Teichert, 1978) form the sister clade to all other crinoids, including the protocrinoids (Guensburg and Sprinkle, 2003). Disparids were recovered as sister to a clade comprised of most ‘cladid’ taxa and hybocrinids were recovered as sister to a group of ‘cyathocrine’ cladids (sensu Moore and Teichert, 1978). A similar pattern was recovered by Guensburg (2012, fig. 2).

Building on these studies, Cole (in press) further assessed the basal split between camerates and non-camerates and tested the taxonomic status of the and

Diplobathrida (Fig. 1). Wright’s (in press) analysis of relationships among non-camerate crinoids offers a more nuanced perspective of this portion of the crinoid tree than previously recovered. Notably, many so-called Ordovician “clades” of Guensburg (2012) and Ausich et al. (2015) do not retain their status of monophyly when post-Ordovician taxa are considered (Wright, in press).

Recent molecular phylogenetic studies indicate broad relationships among major clades of extant crinoids are also reaching a consensus, with the representing the sister clade to all other extant crinoids (Rouse et al., 2013, 2015). Interestingly, divergence time estimation based on relaxed molecular clock models suggests the split between isocrinids and other extant groups took place some 231–252 million years ago

162

(Rouse et al., 2013). Thus, molecular phylogenetic analyses and paleontological evidence

are in general agreement regarding an ancient origin of the crinoid crown group.

A summary tree based on results presented in Ausich et al. (2015), Cole (in

press), Wright (in press), and Rouse et al. (2013) is depicted in the form of a simplified

cladogram in Figure 4.2. This cladogram is annotated with the clade names we propose

below. Terminal taxa in the cladogram were carefully chosen to maximize stability in

phylogenetic nomenclature (Table 3.1). Sereno (2005) listed numerous criteria for

choosing the taxon specifiers in clade definitions. These recommendations include choosing specifiers that are nested rather than basal (if possible), represented by well-

known or readily available material, and using multiple specifiers where necessary to

accommodate phylogenetic uncertainty and/or alternative hypotheses. We have carefully chosen our clade definitions to not hinge on labile phylogenetic hypotheses or specific interpretations of unusual and/or problematic taxa.

Classes of clade definitions used in phylogenetic taxonomy and their graphical representations used herein closely follow Sereno (1999, 2005). Node-based clade definitions circumscribe the most recent common ancestor of at least two taxa and all of its descendants. Thus, node-based definitions form the least inclusive clade containing a minimum of two specifiers. In contrast, stem-based definitions circumscribe the most inclusive clade containing at least one internal specifier. In both cases, additional precision is obtained by identifying external specifiers falling outside the clade (i.e., the outgroup). For example, a stem-based definition for hypothetical Clade A with two internal and one external taxon specifiers can be stated as “all species sharing a more

163

recent common ancestor with species X and Y than Z”, where X and Y are internal taxon specifiers and Z is an external specifier. In other words, Clade A is stem-defined as the most inclusive clade containing X and Y but not Z. Note the presence of one species as an external specifier effectively eliminates the entire clade to which it belongs. By definition, a clade cannot contain an ancestor of its sister group.

In phylogenetic taxonomy, clade membership is not determined by the presence or absence of a “key” morphologic feature unless that apomorphy (or set of apomorphies) is listed in the definition as a qualifying clause (Sereno, 2005). We avoid apomorphic qualifiers in our definitions for several reasons. First, incomplete preservation may lead to cases where it is unknown whether a fossil species has the “key” feature diagnostic of the clade in question. Thus, the inclusion or exclusion of a fossil species depends on character state optimizations rather than direct data. Second, a trait may be “absent” in a taxon either because it was truly absent or because it was secondarily lost. Similarly, a trait may be “present” because of . Moreover, stem group taxa commonly have highly heterogeneous distributions of apomorphic traits, which may lead to instability when new taxa are sampled and/or alternative topologies are equally likely.

Lastly, the timing of a divergence event may not correspond with the acquisition of a diagnostic apomorphy. For example, the blastozoan Macrocystella is widely recognized as a basal glyptocystitoid rhombiferan even though it lacks the respiratory structures traditionally “diagnostic” of the Glyptocystida (Paul, 1968; Sprinkle, 1973; Zamora et al.,

2016). All of these considerations are highly important when considering patterns of

164

character evolution but may lead to nomenclatural instability if incorporated into clade

definitions.

Although we avoid the use of apomorphies to define clades, we do discuss

morphological traits potentially useful for taxonomic diagnoses. In some cases, our

proposed clade definitions retain much of their traditional meaning and taxonomic

content, with constituent taxa sharing numerous synapomorphies that form unambiguous

taxonomic boundaries (e.g., the Flexibilia). However, in other cases, either substantial

revision was necessary and/or a list of unambiguous diagnostic characters were difficult

or impossible to obtain (e.g., the Articulata). These challenges highlight the utility of

phylogenetic taxonomy. For example, many authors have remarked that the Articulata

has lacked a concise, unambiguous definition since it was first erected by J. S. Miller

nearly 200 years ago (Simms, 1988; Webster and Jell, 1999; Hess and Messing, 2011;

Rouse et al., 2013). A phylogenetic definition of the Articulata provides a clearer

criterion for clade membership and results in a framework for future phylogenetic

research assessing relationships among hypothesized stem clades, crown group

synapomorphies, and subsequent morphologic transitions among crown group subclades.

The clade definitions and revised classification proposed herein represent the present state of knowledge, but systematics is a dynamic science and taxonomic theories are commonly reinterpreted in light of new discoveries. We fully expect our definitions to be refined and/or modified as more information becomes available. Some places of the crinoid tree still require extensive taxonomic revisions, such as Upper Paleozoic ‘cladids’

(sensu Moore and Laudon, 1943) and stem articulates (Wright, 2015b). Despite these

165

potential vicissitudes in the taxonomic content and/or definitions within our proposed

classification, we agree with G.G. Simpson’s sentiment: “It is pusillanimous to avoid

making our best efforts today because they may appear inadequate tomorrow” (1944, p.

xxx [sic]).

Systematic paleontology

Crinoidea Miller, 1821

Definition.—The Crinoidea is stem-defined as the most inclusive clade containing

Rhodocrinites verus Miller, 1821, triacontadactylus Miller, 1821, and

Pentacrinites fossilis Blumenbach, 1804 but not Rhopalocystis detombesi Ubaghs, 1963,

Echinosphaerites aurantium (Gyllenhaal, 1772), Eumorphocystis multiporate Branson and Peck, 1940, Protocrinites ouiformis Eichwald, 1840, Cheirocystis antiqua Paul,

1972, Glyptocystella loeblichi (Bassler, 1943), Cambraster cannati Miquel, 1894, and

Cambroblastus enubilatus Smith and Jell, 1990.

Remarks.—This definition captures J. S. Miller’s (1821) original concept based on fossil specimens and retains the name Crinoidea as the clade comprising the crown group plus all extinct species sharing a more recent common ancestor with a living crinoid than any echinoderm taxon listed above as external specifiers (Fig. 4.2). Further, this definition

166 closely resembles the traditional use and taxonomic content of the Crinoidea as used by both biologists and paleontologists (Bather, 1899; Clark, 1915; Jaekel, 1918; Moore and

Teichert, 1978; Hess et al., 1999; Rouse et al., 2013) and accommodates the current state of uncertainty regarding their nearest extinct sister group. In the interest of preserving the taxonomic content and common meaning of a widely used name, our Clade Crinoidea is preferred over Sumrall’s (1997) similarly defined Crinoidoformes (see Cantino and de

Queiroz, 2010, p. 42). The Crinoidea is comprised of two major clades, the Camerata and the Pentacrinoidea, reflecting the early divergence between camerate and non-camerate crinoids (Jaekel, 1918; Donovan, 1988; Guensburg, 2012; Ausich et al., 2015). Because we provide the Crinoidea with a stem-based definition, the discovery of stemward fossils is accommodated within this definition.

Internal taxon specifiers were chosen because they were included in J. S. Miller’s

(1821) original description and represent well known, well preserved, and highly-nested members of their respective subclades. In contrast to the internal taxon specifiers, the choice of external specifiers is more complex. The use of external specifiers in this definition spanning various ‘blastozoan’ and edrioasteroid-grade groups reflects the current difficulty involved in postulating the nearest definitive sister group as well as the uncertain state of relationships among extinct stemmed echinoderms (Smith, 1984;

Sumrall, 1997, 2014; Ausich, 1998a, 1998b; Guensburg and Sprinkle, 2009; Kammer et al., 2013; Ausich et al., 2015; O’Malley et al., 2016; Guensburg et al., 2016).

Ausich et al.’s (2015) analysis of Ordovician crinoids took a conservative approach to outgroup selection by sampling broadly across taxa nested within the Clade

167

Pelmatozoa (Kammer et al., 2013; Sumrall, 2014). Similarly, we have chosen species from multiple pelmatozoan groups as external specifiers to help provide nomenclatural stability in the presence of phylogenetic uncertainty. Other taxa hypothesized to represent the crinoid sister group include the stylophorans (David et al., 2000) and edrioasteroids

(Guensburg and Sprinkle, 2009; Guensburg et al., 2016). Stylophorans have long been considered non-radiate stem group echinoderms (e.g., Paul and Smith, 1984; Smith,

1984; Smith, 2008) and have been cogently demonstrated to lack crown group synapomorphies (Smith, 2005). Thus, we do not consider the stylophoran hypothesis further. Guensburg and Sprinkle (2009) and Guensburg et al. (2016) regard edrioasteroid echinoderms, such as the stromatocystidid Cambraster or the edrioblastoid

Cambroblastus, to possess apomorphies indicating they share a more recent common ancestor with crinoids than with other echinoderms. Although this hypothesis contrasts with previous studies regarding edrioasteroids as stem group eleutherozoans (Paul and

Smith, 1984; Smith 1984, 1985, 1990; Smith and Zamora, 2013), recent investigations suggest that edrioasteroids may comprise a para- or polyphyletic group (Kammer et al.,

2013; Zamora 2013; Zamora and Rahman, 2014). Some edrioasteroids, such as the isorophids, may be closely related to gogiid eocrinoids, whereas other edrioasteroids, such as Cambraster, may be closer to glyptocystitoid blastozoans and crinoids (Kammer et al., 2013; Zamora et al., 2013; Zamora and Rahman, 2014). Because a comprehensive, up-to-date phylogeny of pentaradiate echinoderm lineages is currently lacking, we tentatively follow Guensburg and Sprinkle (2009) and Guensburg et al. (2016) by

168 including both Cambraster and the edrioblastoid Cambroblastus as additional external taxon specifiers.

Identifying synapomorphies of the Clade Crinoidea requires a phylogenetic hypothesis of their position within the broader echinoderm clade. As discussed above, this remains an open question. Basal members of both the Camerata and Pentacrinoidea have a dicyclic calyx with an irregular field of plates intercalating between fixed proximal brachials, suggesting these may be plesiomorphic traits (cf. Apektocrinus,

Cnemecrinus, Glenocrinus) (Guensburg 2012, Ausich et al., 2015; Cole, in press; Wright, in press), but a definitive list of shared derived traits cannot be provided here. Moreover, it is challenging to propose a list of apomorphies that unambiguously differentiate crinoids from other echinoderm taxa because many traits are not exclusive to crinoids.

Crinoids have been traditionally recognized as distinct from blastozoan-grade echinoderms in having true “arms”, where arms are defined as coelomic extensions of the body cavity (Sprinkle, 1973). However, morphologic observations of solute and diploporitan echinoderms such as Eumorphocystis, as well as the discovery of various

Cambrian ‘blastozoans’ with arm-like appendages strongly suggest that arms may not be an apomorphy unique to crinoids (Clausen et al., 2009; Zamora and Smith, 2011;

Sumrall, 2014; Zamora and Rahman, 2014).

We anticipate future phylogenetic research will help resolve these broader issues in echinoderm phylogeny and evolution. Improved knowledge of relationships among extinct pentaradiate echinoderms may also help refine our definition of the Clade

Crinoidea by removing pleonastic external specifiers. We await its refinement.

169

Camerata Wachsmuth and Springer, 1885

Definition.—The Camerata is stem-defined as the most inclusive clade containing

Actinocrinites triacontadactylus Miller, 1821 and Rhodocrinites verus Miller, 1821 but

not Pentacrinites fossilis Blumenbach, 1804.

Remarks.—Camerate crinoids represent a diverse, morphologically distinct “stem clade”

(sensu Sereno, 1999; 2005) ranging from the Lower Ordovician to Permian and contain all taxa traditionally placed within the Diplobathrida and Monobathrida (Moore and

Teichert, 1978; Cole, in press). Camerates are most easily differentiated from pentacrinoids in having calyx plates united by rigid sutures, a heavily plated tegmen surface covering the mouth, and a medial plate (or series of plates) in the posterior (i.e.,

CD) interray. Unlike pentacrinoids, the camerate posterior plate series has no proximal topographic affinity with the C ray, although some camerate posterior plates may be homologous with those of pentacrinoids (see Jaekel, 1918, p. 46; Moore and Laudon,

1943; Brower, 1973, p. 301–304; Guensburg and Sprinkle, 2003). In addition, typical camerate species have fixed proximal brachials, interradials, and sometimes intrabrachials, whereas most derived pentacrinoid clades lack these features.

Multiple studies indicate strong support for camerate monophyly (Ausich, 1998b;

Ausich et al., 2015; Cole, in press). However, Cole’s (in press) analysis of Ordovician camerates did not find support for a strict division between monocyclic and dicyclic forms. Cole’s (in press) phylogenetic revision proposed narrower restrictions on clade

170

membership to render these taxa monophyletic. Following revision, the Monobathrida

and Diplobathrida are sister clades that together comprise the more inclusive Eucamerata

(Cole, in press). Thus, the stem-based definition of the Camerata contains the Clade

Eucamerata and their stem taxa, including representatives of the oldest known crinoid fossils (e.g., Eknomocrinus, Cnemecrinus), genera placed within the problematic

Reteocrinitidae (see Cole, in press), and may or may not contain the protocrinoids (see

Guensburg and Sprinkle, 2003; Guensburg, 2012; Ausich et al., 2015; Cole, in press).

Eucamerata Cole, in press

Definition.— The Eucamerata is node-defined as the least inclusive clade containing

Actinocrinites triacontadactylus Miller, 1821, Rhodocrinites verus Miller, 1821, and

Rosfacrinus robustus LeMenn and Spjeldnaes, 1996.

Remarks.—Cole (in press) revised the Monobathrida and Diplobathrida to represent monophyletic groups while attempting to preserve the greatest number of taxa traditionally included within each (Moore and Teichert, 1978). The name Eucamerata was proposed to identify the clade of camerates comprised of the sister groups

Monobathrida and Diplobathrida, which necessarily excludes stem taxa such as

Cnemecrinus and Reteocrinus (Cole, in press). The Eucamerata comprise the vast majority of camerate taxa and span the Ordovician through Permian. Eucamerates are characterized generally by the traits listed above for the Camerata, but differ in typically

171

having more strongly ankylosed calyx plate sutures, primaxils on the second

primibrachial, holomeric stems, and pinnulate arms (cf. Actinocrinites and Rhodocrinites

with Eknomocrinus and Reteocrinus).

In an attempt to preserve the stability of sister group relationships between

monobathrid and diplobathrid clades, we provide a node-based definition for the

Eucamerata and stem-based definitions for the Monobathrida and Diplobathrida. The

internal taxon specifiers Actinocrinites and Rhodocrinites are highly-nested constituents

of their respective monobathrid and diplobathrid subclades (Moore and Laudon, 1943;

Cole, in press). Rosfacrinus is cautiously included as an additional external specifier because it occupies a somewhat uncertain position at the base of the eucamerate tree (see discussion in Cole, in press).

Monobathrida Moore and Laudon, 1943

Definition.— The Monobathrida is stem-defined as the most inclusive clade containing

Glyptocrinus decadactylus Hall, 1847 and Actinocrinites triacontadactylus Miller, 1821 but not Rhodocrinites verus Miller, 1821 and Archaeocrinus lacunosus (Billings, 1857).

Remarks.—When revising Bather’s (1899) polyphyletic division of crinoids into the

Monocyclica and Dicyclica, Moore and Laudon (1943) placed all camerates with monocyclic calyces into the Monobathrida. Cole’s (in press) phylogenetic analysis of

Ordovician camerate crinoids indicates a strict adherence to Moore and Laudon’s (1943)

172

concept of the Monobathrida is not monophyletic. However, removal of the basal

camerates Eknomocrinus and Adelphicrinus renders the Monobathrida a clade (Cole, in press). The internal and external specifiers defining this stem-based clade ensure the taxonomic content closely matches Moore and Laudon (1943).

Monobathrids are a taxonomically diverse group of camerates ranging from the

Ordovician to Permian and are traditionally diagnosed as monocyclic camerates.

Although other clades had similar trends in circlet reduction (e.g., the Disparida), the transformation from a dicyclic to monocyclic calyx likely represents a veritable synapomorphy of monobathrid camerates, as the dicyclic crinoid Gaurocrinus was

recovered as the sister taxon to the Monobathrida by Cole (in press). Additional features diagnostic of a typical monobathrid species include having radial plates larger than other calyx plates, an upright basal circlet, an uninterrupted radial circlet (except in the posterior interray), and a posterior interray with anitaxis plating and an anitaxial ridge.

Diplobathrida Moore and Laudon, 1943

Definition.—The Diplobathrida is stem-defined as the most inclusive clade containing

Archaeocrinus lacunosus (Billings, 1857) and Rhodocrinites verus Miller, 1821 but not

Actinocrinites triacontadactylus Miller, 1821 and decadactylus Hall, 1847.

Remarks.—Similar to the discussion above, Moore and Laudon (1943) placed all of

Bather’s (1899) dicyclic camerate crinoids within the Diplobathrida. As with the

173 monobathrids, Cole’s (in press) phylogenetic analysis of Ordovician camerates revealed

Moore and Laudon’s (1943) Diplobathrida required revision. To achieve monophyly of diplobathrids while retaining much of Moore and Laudon’s (1943) taxonomic content, all dicyclic stem taxa equally related to both monobathrid and diplobathrid camerates sensu

Cole (in press) are removed from the Diplobathrida (e.g., Eknomocrinus, Reteocrinids, etc.). Following Cole’s (in press) suggested revision, our stem-based definition stabilizes the long-held hypothesis that monobathrids and diplobathrids represent sister clades

(Moore and Laudon, 1943; Cole, in press).

Diplobathrids range from the Ordovician through lower Carboniferous

(). Cole’s (in press) discussion on the taxonomic distribution of diplobathrid morphologies suggests they are generally characterized by a combination of character states, including a dicyclic calyx, a concave calyx base either concealing or partially concealing the infrabasal plates, and the presence of additional plates interrupting the radial circlet in all interrays (e.g., Rhodocrinites). Some diplobathrids sensu Cole (in press), such as the Dimerocrinitidae, are similar to monobathrids in having their radial circlet interrupted only in the posterior interray but can easily be distinguished by their dicyclic calyx. A closer examination of post-Ordovician species indicates a substantial revision of subclades within the Diplobathrida is needed and additional research is currently underway (Cole, 2015)

174

Pentacrinoidea Jaekel, 1918

Definition.—The Pentacrinoidea is stem-defined and as the most inclusive clade containing Apektocrinus ubaghsi Guensburg and Sprinkle, 2009 and Pentacrinites fossilis

Blumenbach, 1804 but not Rhodocrinites verus Miller, 1821 and Actinocrinites triacontadactylus Miller, 1821.

Remarks.—The name Pentacrinoidea originates from Jaekel’s (1894; 1918) prescient observation that camerate and non-camerate crinoid form distinct clades. Although authors after Jaekel (1918) did not adopt this name in subsequent classifications (see

Lane, 1978; Ausich and Kammer, 2001), Jaekel’s usage coincides with this strongly supported clade (Guensburg, 2012; Ausich et al., 2015; Cole, in press; Wright, in press).

Thus, we propose to reinstate the name Pentacrinoidea with the definition above.

We have chosen two phylogenetically distant non-camerate species as internal specifiers. Pentacrinites fossilis is a well-known fossil species from rocks of age and is closely related to extant isocrinid crinoids (David et al., 2006), placing it within the

Crown Crinoidea (see Articulata below). The species Apektocrinus ubaghsi is a Lower

Ordovician fossil and ranks among the stratigraphically oldest known crinoids

(Guensburg and Sprinkle, 2009). However, all phylogenetic research indicates it is closer to non-camerates than camerates and diverges stemward of other basal ‘cladid’ (sensu

Moore and Teichert, 1978) taxa such as Aethocrinus (Guensburg and Sprinkle, 2009;

Guensburg 2012; Ausich et al., 2015; Wright, in press). Our stem-based definition

175

recognizes Jaekel’s (1918) priority of this concept and effectively places all known non-

camerate species within the Pentacrinoidea.

Pentacrinoids are a spectacularly diverse and morphologically heterogeneous clade ranging from the Early Ordovician to present day marine communities. The primary apomorphies differentiating pentacrinoids from camerates relate to their distinctive posterior plating patterns, the degree of calyx plate suturing, and oral region rigidity (‘tegmen’ terminology here is from Ausich and Kammer, 2016). Posterior plates among pentacrinoids display a proximal relationship with the C-ray radial plate

(Guensburg 2010; Wright, 2015a). Subclades within the Pentacrinoidea express this affinity differently (cf. Cladida and Disparida), and extant crinoids do not retain posterior plates as adults. However, the ontogenetic trajectory of posterior plate development in extant crinoids is tightly linked with morphologic patterns among their Paleozoic precursors (Wright, 2015a). Pentacrinoid calyx plates are less closely sutured (i.e., ankylosed) than camerates and typically have a non-rigid to flexible oral region. In many pentacrinoids, the mouth is directly exposed on the oral surface rather than beneath a tegmen (Ausich and Kammer, 2016).

There are several other morphologic features less diagnostic than those described above but still useful for distinguishing most pentacrinoid species from camerates. For example, some basal pentacrinoids such as Apektocrinus, Aethocrinus, and Alphacrinus incorporate additional plates within the calyx (similar to camerates). However, the overwhelming majority of pentacrinoid clades do not. A major exception occurs among flexible crinoids, but flexibles are a derived group of pentacrinoids and can be

176 differentiated from camerates by other apomorphies (see Flexibilia below). Similarly, eucamerate crinoids have pinnules but most early to middle Paleozoic pentacrinoids do not. Pinnulation evolved at least once (and probably several times) during the middle to late Paleozoic among the subclade Cladida (Wright, 2015b), but these taxa can readily be distinguished from eucamerates in having a pentacrinoid-like posterior plating pattern and free arms above the radials.

Inadunata Wachsmuth and Springer, 1885

Definition.—The Inadunata is node-defined and is the least inclusive clade containing

Synbathocrinus conicus Phillips, 1836 and Dendrocrinus longidactylus Hall, 1852.

Remarks.—Wachsmuth and Springer (1885) placed non-articulate fossil crinoids with free arms above the radial plates within the Inadunata. Subsequent classifications divided the Inadunata into the Cladida and Disparida based on the number of circlets in the calyx

(Moore and Laudon, 1943; Moore and Teichert, 1978). In a pioneering study on phylogenetic approaches to crinoid classification, Simms and Sevastopulo (1993) pointed out the Inadunata of Moore and Teichert (1978) was paraphyletic and recommended the name be abandoned. In addition, Simms and Sevastopulo’s (1993) revision resolved the paraphyly of cladid inadunates by including the Flexibilia and Articulata within the

Cladida.

177

The division between the Camerata and Pentacrinoidea (discussed above)

indicates disparids and cladids are more closely related to one another than to camerates

(Fig 4.2). Indeed, recent phylogenetic analyses of Ordovician crinoids recover a sister

group relationship between disparids and cladids (sensu Moore and Laudon, 1943), with

hybocrinids nested within the Cladida (Fig. 4.2) (Guensburg, 2012; Ausich et al., 2015;

Wright, in press). Our definition of the Inadunata combines Wachsmuth and Springer’s

(1885) original concept with Simms and Sevastopulo’s (1993) revision of the Cladida to

include flexibles and articulates. Note that this definition places basal pentacrinoids, such

as Apektocrinus, outside the Inadunata. We combine a node-based definition of the

Inadunata with stem-based definitions for the subclades Disparida and Cladida to form a node-stem triplet to increase the stability of sister relationships between these taxa

(Sereno, 1999).

The Clade Inadunata ranges from the Early Ordovician to the present and are as a whole well-characterized by Wachsmuth and Springer’s (1885) general concept of crinoids with free arms above the radial plates. Exceptions to this diagnosis occur but are mostly restricted to a few possibly basal taxa and the Flexibilia, which represent a

derived group of inadunates (Springer, 1920).

Disparida Moore and Laudon, 1943

Definition.—The Disparida is stem-defined as the most inclusive clade containing

Synbathocrinus conicus Phillips, 1836 but not Dendrocrinus longidactylus Hall, 1852.

178

Remarks.—Disparids comprise a diminutive but morphologically and taxonomically

diverse clade of fossil crinoids ranging from the Ordovician through Permian. Moore and

Laudon (1943) erected the Disparida to include all monocyclic inadunates. Disparid

monophyly is well-supported by phylogenetic analyses of Ordovician crinoids

(Guensburg, 2012; Ausich et al., 2015; Wright, in press) and contains all species closer to

Synbathocrinus than the cladid Dendrocrinus. Given the similar topologies across these studies, the Clade Disparida retains taxa traditionally placed within disparids (sensu

Moore and Laudon, 1943) except for the hybocrinids.

A major synapomorphy and useful diagnostic trait of disparid crinoids is the presence of a single circlet of plates below the radials. All other pentacrinoids are either dicyclic (cladids), pseudomonocyclic (hybocrinids) (see Sprinkle, 1982a), or otherwise

phylogenetically distant from disparids (some derived articulates may not develop

infrabasals, see Lahaye and Jangoux, 1987). Disparids also have simple or compound

radial plates, typically lack pinnules, and approximate bilateral symmetry between rays

oriented in one of several possible planes (see Moore et al, 1978). As pentacrinoids,

disparids have posterior plates in a proximal position to the C ray but differ from cladids

in having plates positioned above rather than below or in-line with the C-ray radial plate.

However, posterior plate homologies among disparids and between inadunate clades are

presently obscured by a set of descriptive terms opaque to homology. Whether the

proximal C-ray posterior plate is an “anibrachial”, “radianal”, “anal X”, “superradial”, or

a “radial” is uncertain (Moore, 1962; Moore and Teichert, 1978; Ausich, 1996). Future

179

work is needed to help clarify primary posterior plate homologies among disparids and

between cladids and disparids. The results of Wright’s (in press) analysis of Ordovician

through Devonian pentacrinoid taxa support Guensburg’s (2010) assessment of

Alphacrinus as a lower Tremadocian crinoid phylogenetically close to the base of the

disparid clade. Guensburg (2010) considered the posterior of Alphacrinus to express a

transitional form between “typical” pentacrinoid posterior plates and the ray-like

extensions common among disparid taxa. A re-examination of the posterior interray of

basal taxa combined with studies on disparid ontogeny may help resolve this issue.

Cladida Moore and Laudon, 1943

Definition.—The Cladida is stem-defined as the most inclusive clade containing

Dendrocrinus longidactylus Hall, 1852 but not Synbathocrinus conicus Phillips, 1836.

Remarks.—The Cladida were originally defined by Moore and Laudon (1943) to comprise a tremendously diverse and long-ranging (Ordovician–Triassic) assemblage of dicyclic inadunates with their mouths covered with primary peristomial cover plates

(Ausich and Kammer, 2016). Moore and Laudon’s (1943) original concept and taxonomic content of the Cladida is paraphyletic, as they agreed with Springer’s (1920) earlier assessment that flexible crinoids were more closely related to some cladids than others but did not place the Flexibilia within the Cladida. Moreover, post-Paleozoic crinoids within Miller’s (1821) Articulata have long been considered descendants of

180

Paleozoic cladids (Jaekel, 1918; Moore et al., 1952; Rassmussen, 1978; Simms, 1988).

Simms and Sevastopulo (1993) conducted a cladistic analysis of Paleozoic cladids,

flexibles, and articulate crinoids and subsequently remedied cladid paraphyly by placing

the Flexibilia and the Articulata within the Cladida (sensu Moore and Laudon, 1943).

Although many authors have followed Simms and Sevastopulo’s (1993) interpretations of

relationships among these taxa, only a few authors have since followed their revised

rank-based classification (e.g., Brower, 2001, 2002; Donovan and Harper, 2003).

Our stem-based definition of the Cladida is similar in taxonomic content to

Simms and Sevastopulo (1993) because it includes all species closer to Dendrocrinus

than to the disparid Synbathocrinus. Thus, the Cladida spans the Ordovician to the Recent

and contains the major subclades Porocrinoidea, Flexibilia, and Articulata. Cladids are

most easily distinguished from their sister group, the Disparida, in typically having a

dicyclic calyx and posterior plates (as adults or during development) located below

and/or in-line with the radial plate circlet (Wright, 2015a). Lastly, many middle Paleozoic

to Recent cladids have pinnules, whereas most disparids do not (Frest et al., 1979).

Porocrinoidea Wright, in press

Definition.—The Porocrinoidea is node-defined and is the least inclusive clade containing

Carabocrinus radiatus Billings, 1857 and Hybocrinus conicus Billings, 1857.

181

Remarks.—In their description of crinoids belonging to Bather’s (1899) ‘Cyathocrinina’,

Moore and Laudon (1943) speculated that “primitive” cyathocrinoids such as

Carabocrinus might be closely related to the enigmatic taxon Hybocrinus. Sprinkle

(1982a) argued the stem and calyx morphology of Hybocrinus suggested hybocrinids

were “pseudomonocyclic” and listed a number of characters linking hybocrinids with

cladids. Although hybocrinids have not traditionally been classified within the Cladida,

many phylogenetic analyses of Ordovician crinoids have recovered a clade of

‘cyathocrine’ grade cladids and hybocrinids (Guensburg, 2012; Ausich et al., 2015;

Wright, in press). Wright’s (in press) phylogenetic analysis of Ordovician through

Devonian pentacrinoids recovered a clade comprised of Porocrinus, Carabocrinus, and

the hybocrinids Hybocrinus and Hybocystites. Notably, this clade is stemward of the split

between flexible and other cladid crinoids. Thus, Wright (in press) proposed the name

Porocrinoidea to encompass this early diverging and morphologically unique clade of

Ordovician crinoids.

Our node-based definition of the Porocrinoidea sets up a node-stem triplet that

stabilizes the sister clade relationship among the Porocrinida and Hybocrinida recovered by Ausich et al. (2015), which had denser taxon sampling of Ordovician crinoids than

Wright (in press). The Clade Porocrinoidea is likely limited to the Ordovician Period, but

additional analyses sampling younger species are needed to test the extent of their

geologic duration. Porocrinoids are a subclade of cladids characterized by globose,

conical, or ovate calyces that possess a number of apomorphies convergent with

blastozoan echinoderms, such as having thecal respiratory structures, reduction in arm

182

number and calyx plates, and/or recumbent ambulacra (see Moore and Teichert, 1978;

Sprinkle, 1982a, 1982b).

Porocrinida Miller and Gurley, 1894

Definition.—The Porocrinida is stem-defined as the most inclusive clade containing

Porocrinus conicus Billings, 1857 and Carabocrinus radiatus Billings, 1857 but not

Hybocrinus conicus Billings, 1857.

Remarks.—The Porocrinida comprise a small clade of Ordovician porocrinoids with

apomorphic endothecal and/or exothecal respiratory structures. Sprinkle (1982b) pointed

to many similarities among Carabocrinus, Palaeocrinus, and the Porocrinidae and

hypothesized they may be closely related. Ausich et al. (2015) recovered a topology

supporting this hypothesis with the Euspirocrinid Illemocrinus as their sister taxon.

However, Wright (in press) recovered Euspirocrinus outside the porocrinid clade within a

different clade of ‘cyathocrine’ grade cladids. Thus, Illemocrinus is tentatively placed

within the Porocrinida, but other taxa within the Euspriocrinidae should not be placed

within the Porocrinida at this time as additional revisions are necessary. Guensburg

(2012) recovered a similar tree to Ausich et al. (2015) that suggested Perittocrinus may

be also be a porocrinid.

The stem-based definition of the Porocrinida makes them sister to the

Hybocrinida and retains the taxonomic membership of this clade recovered in Ausich et

183

al. (2015) and Guensburg (2012). Porocrinids can easily be distinguished from

hybocrinids in having a dicyclic calyx and the presence of thecal respiratory structures

(Kesling and Paul, 1968; Sprinkle, 1982b).

Hybocrinida Jaekel, 1918

Definition.—The Hybocrinida is stem-defined as the most inclusive clade containing

Hybocrinus conicus Billings, 1857 and Hybocystites problematicus Wetherby, 1880 but not Porocrinus conicus Billings, 1857 and Carabocrinus radiatus Billings, 1857.

Remarks.—Hybocrinids comprise a small yet morphologically disparate clade of

Ordovician crinoids. Although the monocyclic hybocrinids have either been considered disparids or classified outside the Inadunata (Moore and Laudon, 1943; Moore and

Teichert, 1978; Ausich 1998b), Sprinkle (1982a) suspected hybocrinids might be

“pseudomonocyclic” and potentially related to ‘cyathocrine’ cladids (see Sprinkle, 1982a,

1982b). Phylogenetic analyses by Guensburg (2012), Ausich et al. (2015), and Wright (in

press) all support the monophyly of the Hybocrinida and their sister group relationship

with taxa placed in the Porocrinida (see Sprinkle, 1982a).

In addition to having a “pseudomonocyclic” calyx (infrabasals absent),

hybocrinids are characterized by a number of unusual apomorphies that distinguish them

from Porocrinids (and all other crinoids). Many of these traits are similar to those

typically present in blastozoan echinoderms, including reduction in the number of arms,

184

modification of food-gathering appendages to be recumbent (sometimes extending

downward over calyx plates), and reduction in the number of calyx plates (Sprinkle and

Moore, 1978).

Flexibilia Zittel, 1895

Definition.—The Flexibilia is stem-defined as the most inclusive clade containing

Taxocrinus macrodactylus (Phillips, 1841) but not Dendrocrinus longidactylus Hall,

1852.

Remarks.—Flexible crinoids are a morphologically homogeneous clade that originated sometime during the Middle to Late Ordovician and range through the Permian. Frank

Springer (1911, 1920) was the first to recognize that flexible crinoids were closely related to inadunates. In his comprehensive 1920 monograph, The Crinoidea Flexibilia, Springer

compared morphologic characteristics of the inadunate Cupulocrinus with the earliest

known flexible Protaxocrinus, citing numerous similarities in calyx plating, interradial

areas, and the arrangement of posterior plates. Springer (1920) concluded Cupulocrinus

was potentially a that linked inadunates with flexibles, stating, “there is

clearly an intermingling of the characters…and it is evident that in Cupulocrinus we have

to deal with a transition [sic] form whose exact status is difficult to decide” (Springer,

1920, p. 89). Subsequent taxonomic treatments have also recognized Cupulocrinus as

185 occupying a proximal position to the base of the flexible tree (Moore and Laudon, 1943;

Moore and Teichert, 1978).

Phylogenetic analyses sampling flexible and other crinoid taxa have invariably recovered tree topologies supporting Springer’s (1911, 1920) hypothesis, with

Cupulocrinus recovered as the sister taxon to Flexibilia (Brower, 1995, 2001; Ausich,

1998b; Ausich et al., 2015; Wright, in press). Wright’s (in press) analysis used Bayesian methods to estimate the probability of Cupulocrinus being ancestral (sensu Foote, 1996) to the flexible clade. Results strongly support Cupulocrinus as occupying an ancestral position (posterior probability = 0.99) (Wright, in press). Given these results and our stem-based definition of the Flexibilia, species of Cupulocrinus are now placed within the flexibles.

Flexible crinoids have loosely-sutured calyx plating and a remarkably uniform set of apomorphies relative to other crinoid clades. For example, flexibles differ from cladids in having interradial and intrabrachial plates and differ from other dicyclic crinoids in typically having their lowermost circlet comprised of three (rather than five) infrabasal plates. One infrabasal plate, the “azygous”, is smaller than the other two that formed by fusion and is located in the C ray (except for the derived Forbesiocrinus). Many flexibles retain posterior plate arrangements similar to other cladids, but posterior plates are sometimes absent in more derived flexibles. In contrast with cladids, arms of flexible crinoids are universally uniserial and lack pinnules, and the stem is nearly always transversely circular (Springer, 1920).

186

Eucladida Wright, in press

Definition.—The Eucladida is stem-defined as the most inclusive clade containing

Dendrocrinus longidactylus Hall, 1952 and Pentacrinites fossilis Blumenbach, 1804 but

not Taxocrinus macrodactylus (Phillips, 1841).

Remarks.—The revision of the Cladida to be monophyletic requires placing the subclades

Porocrinoidea, Flexibilia, and Articulata within a more inclusively defined Clade Cladida

(Simms and Sevastopulo, 1993; Wright, in press). However, the Cladida (sensu Moore

and Laudon, 1943) is traditionally conceived as a Paleozoic-age paraphyletic group that

excludes the Flexibilia. The Eucladida was proposed by Wright (in press) to comprise all

species within the Clade Cladida sharing a more recent common ancestor with

Dendrocrinus and Pentacrinites than with Taxocrinus. Thus, the stem-based clades

Flexibilia and Eucladida are sister to one another and articulates are nested within the

Eucladida. This Eucladida retains much of the meaning and taxonomic content of Moore and Laudon’s (1943) concept for Paleozoic cladids while eschewing paraphyly.

In the Treatise on Invertebrate Paleontology, Moore et al. (1978) recognized three rank-based taxa within the Cladida: the Dendrocrinida, the Cyathocrinida, and the

Poteriocrinida. However, it has long been questioned whether or not these taxa represent monophyletic groups (McIntosh, 1986; 2001; Sevastopulo and Lane, 1988; Simms and

Sevastopulo, 1993; Kammer and Ausich, 1992, 1996; Wright, 2015a, 2015b; Wright and

Ausich, 2015). Indeed, the Poteriocrinida is depicted in the Treatise as a polyphyletic

187

group (Moore et al., 1978, fig. 412). A phylogenetic analysis of Ordovician through

Devonian pentacrinoids by Wright (in press) has confirmed the doubts over the monophyly of these taxa. Much of the problem arises from ambiguous and/or uninformative apomorphies chosen for these taxa that perpetuate taxonomic anarchy via

“undiagnostic diagnoses” (Wright, 2015b, see Lane, 1978, p. T295). Although much revision is needed, recent analyses indicate there is nevertheless considerable phylogenetic structure among subclades of Paleozoic cladids and additional work is underway to revise this diverse group (Wright, 2015b).

Articulata Miller, 1821

Definition.—The Articulata is node-defined as the least inclusive clade containing

Endoxocrinus parrae (Gervais, 1835) and (Pennant, 1777).

Remarks.—The Articulata was proposed by Miller (1821) and has since developed a longstanding reputation as a problematic group that lacks a concise and unambiguous definition (Rasmussen, 1978; Simms, 1988; Simms and Sevastopulo, 1993; Webster and

Jell, 1999; Rouse et al., 2013). Although all extant crinoids are invariably recognized as articulates, much confusion surrounds the recognition of fossil articulates and the timing of their origin. The primary difficulties surround which apomorphy (or combination of apomorphies) is useful for diagnosing the Articulata. For example, it is widely appreciated that no apomorphy or unique set of apomorphies can presently diagnose

188 fossil articulates without ambiguity (Simms, 1988; Simms and Sevastopulo, 1993;

Webster and Jell, 1999; Rouse et al, 2013). Most crinoid workers have obviated this problem by simply treating the Articulata as synonymous with post-Paleozoic crinoids

(see Simms and Sevastopulo, 1993). However, this usage is problematic because this definition is not based on any explicit phylogenetic hypothesis. Moreover, many

Paleozoic groups of fossil cladids share different combinations of traits typically listed as

“diagnostic” for the Articulata (Webster and Jell, 1999; Webster and Lane, 2007). If the concept of what defines the Articulata depends on the choice of a particular combination of apomorphies alone, then questions regarding the “origin of the Articulata” will always depend on which specific combination was chosen a priori to be diagnostic. Without a phylogenetic definition, it is impossible to objectively specify a precise set of synapomorphies for the Articulata. Thus, we propose herein to define the Articulata as the crinoid crown group containing the last common ancestor of the extant isocrinid

Endoxocrinus parrae and the comatulid Antedon bifida, and all of its descendants.

As discussed by Ruta et al. (2003), the concepts of stem groups and crown groups are sometimes misinterpreted or misused in the paleontological literature. Used properly, crown groups are defined by extant taxon specifiers. Notably, crown groups may be comprised of many (or mostly) extinct fossil species. For example, if a fossil crinoid is more closely related to some extant species than others, it is a member of the crown group. According to Rouse et al. (2013), the most recent common ancestor of all extant crinoids lived sometime during the Middle to Upper Triassic. Thus, our node-based definition eliminates the non-phylogenetic concept of “post-Paleozoic Crinoidea” while

189 retaining the majority of post-Paleozoic crinoids traditionally included within the

Articulata. The Clade Articulata is synonymous with the Crown Crinoidea (Sumrall,

2014), and we advocate workers use these terms interchangeably depending on context

(e.g., discussing relationships among crinoids or between crinoids and non-crinoids).

Traits that may be present in the Articulate ancestor are listed in Simms (1988), Simms and Sevastopulo (1993), Webster and Jell (1999), and Rouse et al. (2013).

The Articulata likely contains most post-Paleozoic taxa traditionally considered articulates, including the ~600 or so extant species. Although our definition of the

Articulata is defined with precision and phylogenetic stability (Rouse et al., 2013; Rouse et al., 2015), it remains difficult in practice to unambiguously identify fossil articulates, particularly among specimens near the base of the articulate tree. However, such difficulties are already present and have long obfuscated the origin of the crinoid crown group. The more important problem is resolving the phylogenetic position of the common ancestor of extant crinoids within the myriad of fossil lineages. Our definition provides a useful framework for future phylogenetic research to uncover relationships between potential stem articulates, extinct crown group lineages, and extant species.

A revised rank-based classification of the Crinoidea

Crinoid clades identified herein confirm many long-held views on the major divisions among crinoids from both the foundational work of Moore and Laudon (1943) and

190

Moore and Teichert (1978) to more recent analyses (i.e., Ausich, 1998a, 1998b;

Guensburg and Sprinkle, 2003; Guensburg, 2012; Ausich et al., 2015). Results from all of

these studies recognized the Camerata, Diplobathrida, Monobathrida, Hybocrinida,

Disparida, Cladida, and Flexibilia. The challenge is to represent these widely recognized

clades in a rank-based Linnaean classification scheme that maximizes common usages of names for crinoid lineages (Moore and Teichert, 1978) and is consistent with a phylogenetic understanding of relationships (Wiley and Lieberman, 2011). In our revision, the Crinoidea remain a class and every attempt is made to retain orders as

recognized in Moore and Teichert (1978). Unfortunately, the tree topology of Figure 4.2

prevented the attainment of the latter in all instances, but the addition of intermediate

Linnaean ranks makes it easier to apply a phylogenetic perspective to rank-based crinoid

classification. The use of intermediate ranks (e.g., Parvclass) follows traditional use in

pre-existing taxonomic literature (see Carroll 1988; Sibley, 1994; Benton, 2005). Two older taxonomic names, the Pentacrinoidea Jaekel, 1918 and Inadunata Wachsmuth and

Springer, 1885, are formally reinstated herein because they represent meaningful clades as described above.

Post-Ordovician cladids (sensu Moore and Laudon, 1943) and the Protocrinoida

(Guensburg and Sprinkle, 2003) remain problematic groups. Because the rank for a monophyletic Cladida must be above flexibles and articulates (Simms and Sevastopulo,

1993), we propose the name Cyathoformes to contain taxa traditionally placed within the

Cladida that are sister to the Articulata. Relationships among these taxa are the subject of future phylogenetic research (Wright, 2015b) and are not treated further here. From their

191

initial description (Guensburg and Sprinkle, 2003), the protocrinoids have been an

important but confounding group of crinoids that display characteristics of both crinoids

and other stalked echinoderms. Guensburg and Sprinkle (2003) regarded the protocrinoid

as an “order (plesion)”. The validity of the protocrinoids was later questioned by

Guensburg and Sprinkle (2009) and led Guensburg (2012) to formally place them within

the Camerata. However, Ausich et al. (2015) recovered a sister group relationship

between Titanocrinus and Glenocrinus, but with protocrinoids more closely related to

non-camerates than camerates. In contrast, Cole’s (in press) analysis of Ordovician crinoids recovered the protocrinoids as more closely related to camerates than non- camerates. Thus, we have carefully chosen our clade definitions above to not depend on a particular phylogenetic hypothesis or morphologic interpretation of these significant but problematic taxa. For the moment, we tentatively place both protocrinoid taxa as

Crinoidea incertae sedis subclass Protocrinoida.

In our present understanding of crinoid evolution, the first major divergence

occurs between camerates and all other crinoids (Fig. 4.2). The subclass rank is retained

for the Camerata; and the subclass Pentacrinoidea Jaekel, 1918 is proposed for its sister

group (Table 3.2). Within the Camerata, the orders Diplobathrida and Monobathrida are

retained as sister groups, and the infraclass Eucamerata Cole (in press) unites these two

orders. In phylogenetic analyses of camerates, several taxa are not placed within the

Monobathrida and Diplobathrida (sensu Cole, in press). Thus, they are considered here to be stem eucamerates (see remarks for Eucamerata above). The subclass Camerata unites these stem taxa with eucamerates.

192

In terms of species richness, the subclass Pentacrinoidea is the largest crinoid

clade. This includes the Disparida, Cladida, Hybocrinida, and Articulata of Moore and

Teichert (1978), which coincides exactly with the Jaekel’s (1918) concept of the

Pentacrinoidea (see Lane, 1978). Hence, we have proposed the reinstatement of this

name. The Pentacrinoidea is comprised of the infraclass Inadunata and their stem taxa

(e.g., Apektocrinus). The concept for the Inadunata in Moore and Teichert (1978) united the Disparida and the Cladida. Here, the infraclass Inadunata unites the Disparida,

Cladida, and all of their descendants, the latter of which were traditionally placed outside the Cladida. This usage circumvents the non-phylogenetic usage of the Inadunata (sensu

Moore and Teichert, 1978) and is consistent with the phylogenetic conclusions of Simms and Sevastopulo (1993). With the Inadunata an infraclass, the Disparida and Cladida are both parvclasses. Taxa placed within the Disparida are in need of revision and current work is underway to establish relationships among sublcades (Ausich and Donovan,

2015).

Within the Cladida, the Hybocrinida, Porocrinida, Taxocrinida Springer, 1913; and Sagenocrinida Springer, 1913 are orders (Table 2). The orders Hybocrinida and

Porocrinida are sister groups forming the superorder Porocrinoidea Wright (in press). The

Flexibilia are transferred to the superorder rank which is comprised of the two sister groups, order Taxocrinida and order Sagenocrinida. The Cyathoformes and Articulata comprise the magnorder Eucladida (Wright, in press). The Eucladida retains most cyathocrinids, dendrocrinids, and poteriocrinids of Moore and Teichert (1978). As discussed above, phylogenetic relationships within this clade await further study (Wright,

193

2015b). Analyses detailing the late Paleozoic and early crinoid phylogeny are needed to understand this crucial period of crinoid evolution. Lastly, the Articulata is considered a superorder within the Cladida.

Conclusions

A phylogeny-based revision of crinoid systematics is proposed to clarify the definition of

clades and inform a major revision of the rank-based Linnaean classification. These

revisions are based on recent computational phylogenetic analyses that build on the

historic subdivision of crinoids into major lineages. It is hoped that the phylogenetic

classification schemes presented herein will help provide a framework for future research

on crinoid phylogeny and offer guidance to crinoid workers and non-specialists alike

interested in using this fascinating group of echinoderms to study evolutionary patterns

and processes.

194

References

Alaniz, R.J., 2014, Dredging evolutionary theory: the emergence of the deep sea as a

transatlantic site for evolution 1853–1876 [Ph.D. dissertation]: University of

California, San Diego, 360 p.

Angelin, N. P., 1878, Iconographia Crinoideorum. in stratis Sueciae Siluricis fossilium:

Samson and Wallin, Holmiae, 62 p.

Ausich, W.I., 1980, A model for niche differentiation in lower Mississippian crinoid

communities: Journal of Palontology, v. 54, p. 273–288.

Ausich, W.I., 1986, Early Silurian rhodocrinitacean crinoids (Brassfield Formation,

Ohio): Journal of Paleontology, v. 60, p. 84–106.

Ausich, W.I., 1996, Crinoid plate circlet homologies: Journal of Paleontology, v. 70, p.

955–964.

Ausich, W.I., 1998a, Early phylogeny and subclass division of the Crinoidea (phylum

Echinodermata): Journal of Paleontology, v. 72, p. 499–510.

Ausich, W.I., 1998b, Phylogeny of Arenig to Caradoc Crinoids (Phylum Echinodermata)

and suprageneric classification of the Crinoidea: The University of Kansas

Paleontological Contributions Papers, New Series, no. 9, 36 p.

Ausich, W.I., and Babcock, L.E., 1998, The phylogenetic position of Echmatocrinus

brachiatus, a probable octocoral from the Burgess Shale: Palaeontology, v. 41, p.

193–202.

195

Ausich, W.I., and Bottjer, D.J., 1982, Tiering in suspension-feeding communities on soft

substrata throughout the Phanerozoic: Science, v. 216, p. 173–174.

Ausich, W.I., and Copper, P., 2010, The Crinoidea of Anticosti Island, Québec: (Late

Ordovician to Early Silurian): Palaeontographica Canadiana, vol. 29, 157 p.

Ausich, W.I., and Donovan, S.K., 2015, Phylogeny of disparid crinoids: Geological

Society of America Abstracts with Programs, v. 47, p. 855.

Ausich, W.I., and Kammer, T.W., 2001, The study of crinoids during the 20th century

and the challenges of the 21st century: Journal of Paleontology, v. 75, p. 1161–

1173.

Ausich, W.I., and Kammer, T.W., 2008, Generic concepts in the Amphoracrinidae

Bather, 1899 (Class Crinoidea) and evaluation of generic assignments of North

American species: Journal of Paleontology, v. 82, p. 1139–1149.

Ausich, W.I., and Kammer, T.W., 2016, Exaptation of pelmatozoan oral surfaces:

constructional pathways in tegmen evolution: Journal of Paleontology,

doi:10.1017/jpa.2016.73.

Ausich, W.I., Kammer, T.W., and Baumiller, T.K., 1994, Demise of the Middle

Paleozoic crinoid fauna: a single extinction event or rapid faunal turnover?:

Paleobiology, v. 20, p. 345–361.

Ausich, W.I., Kammer, T.W., Rhenberg, E.C., and Wright. D.F., 2015, Early phylogeny

of crinoids within the Pelmatozoan clade: Palaeontology, v. 58, p. 937–952.

196

Bassler, R. S., 1943, New Ordovician cystidean echinoderm from Oklahoma: American

Journal of Science, v. 241, p. 694–703.

Bather, F.A., 1898, Wachsmuth and Springer’s classification of crinoids: Natural

Science, v. 12, p. 337–345.

Bather, F.A., 1899, A phylogenetic classification of the Pelmatozoa: British Association

for the Advancement of Science (1898), p. 916–923.

Bather, F.A., 1900, Part III The Echinoderma. The Pelmatozoa, in Lankester, E.R., ed., A

Treatise on Zoology: Adam and Charles Black, London, p. 94–204.

Baumiller, T.K., 2008, Crinoid ecological morphology: Annual Review of Earth and

Planetary Sciences, v. 36, p. 221–249.

Benton, M.J., 2000, Stems, nodes, crown clades, and rank-free lists: is Linnaeus dead?:

Biological Reviews of the Cambridge Philosophical Society, v. 75, p. 633–648.

Benton, M.J., 2005, Paleontology, 3rd edition: Blackwell Publishing, Oxford.

Benton, M.J., 2007, The Phylocode: Beating a dead ?: Acta Palaeontologica

Polonica, v. 52, p. 651–655.

Billings, E., 1857, New species of fossils from Silurian rocks of Canada: Canada Geological

Survey, Report of Progress 1853-1856, Report for the year 1856, p. 247–345.

Blumenbach, J. F., 1802–1804, Abbildungen naturhistorischer Gegenstiinde: Gottingen,

pt. 70, no. 70, 4 p.

Branson, E.B., and Peck, R.E., 1940, A new cystoid from the Ordovician of Oklahoma: Journal

of Paleontology, v. 14, p. 89–92. 197

Brochu, C.A., 2003, Phylogenetic approaches toward crocodylian history: Annual

Review of Earth and Planetary Sciences, v. 31, p. 357–397.

Brochu, C.A., and Sumrall, C.D., 2001, Phylogenetic nomenclature and paleontology:

Journal of Paleontology, v. 75, p. 754–757.

Brower, J.C., 1973, Crinoids from the Girardeau Limestone (Ordovician): Palaeontographica

Americana, v. 7, p. 263–499.

Brower, J.C., 1995, Dendrocrinid crinoids from the Ordovician of northern Iowa and

southern Minnesota: Journal of Paleontology, v.69, p. 939–960.

Brower, J. C., 2001, Flexible crinoids from the Upper Ordovician

of the Northern Midcontinent and the evolution of early flexible crinoids: Journal

of Paleontology, v. 75, p. 370–382.

Brower, J.C., 2002, Cupulocrinus angustatus (Meek and Worthen, 1870), A cladid

crinoid from the Upper Ordovician Maquoketa Formation of the Northern

Midcontinent of the United States: Journal of Paleontology, v. 76, p. 109–122.

Butler, R.J., Upchurch, P., and Norman, D.B., 2008, The phylogeny of the ornithischian

dinosaurs: Journal of Systematic Palaeontology, v. 6, p. 1–40.

Cantino, P.D., and de Queiroz, K., 2010, PhyloCode: International code of phylogenetic

nomenclature, version 4c, 90 p.

Carlson, S. J., 2001, Ghosts of the past, present, and future in brachiopod systematics:

Journal of Paleontology, v. 75, p. 1109–1118.

198

Carlson, S.J., and Leighton, L.R., 2001, The phylogeny and classification of

Rhynchonelliformea: The Paleontological Society Special Papers, v. 7, p. 27–51.

Carroll, R. L., 1988, and Evolution: WH Freeman, New York,

698 p.

Clark, A. H. 1908. Description of new species of crinoids, chiefly from the collections

made by U.S. Fisheries steamer ‘‘Albatross’’ at the Hawaiian Islands in 1902,

with remarks on the classification of the Comatulida: Proceedings of the U.S.

National Museum, v. 34, p. 209–239.

Clark, A.H., 1915, A monograph of the existing crinoids, part 1: Bulletin of the United

States National Museum 82, p. 1–406.

Clark, A.H., 1921, A monograph of existing crinoids, part 2: Bulletin of the United States

National Museum 82, p. 1–795.

Clark, A.H., and Clark, A.M., 1967, A monograph of existing crinoids, part 5: Bulletin of

the United States National Museum 82, p. 1–806.

Clausen, S., Jell, P.A., Legrain, X., and Smith, A.B., 2009, Pelmatozoan arms from the

Middle Cambrian of Australia: bridging the gap between brachioles and

brachials?: Lethaia, v. 42, p. 283–296.

Cohen, B.L., Ameziane, N., Eleaume, M., and de Forges, B.R., 2004, Crinoid phylogeny:

a preliminary analysis (Echinodermata: Crinoidea): Marine Biology, v. 144, p.

605–617.

199

Cole, S.R., 2015, A phylogenetic test of the suprageneric classification of diplobathrid

crinoids: Geological Society of America Abstracts with Programs, v. 47, p. 853.

Cole, S.R., in press, Phylogeny and morphologic evolution of the Ordovician Camerata

(Class Crinoidea, Phylum Echinodermata): Journal of Paleontology.

Darwin, C., 1859, On the origin of species by means of natural selection: Murray,

London, 502 p.

David, B., Lefebvre B., Mooi, R., and Parsley, R., 2000, Are homalozoans echinoderms?

An answer from extraxial-axial theory: Paleobiology, v. 26, p. 529–555.

David, B., Roux, M., Messing, C.G., and Ameziane, N., 2006, Revision of the

pentacrinid stalked crinoids of the genus Endoxocrinus (Echinodermata,

Crinoidea), with a study of environmental control characters and its consequences

for taxonomy: Zootaxa, v. 1156, p. 1–50.

Deline, B., and Ausich, W.I., 2011, Testing the plateau: a reexamination of disparity and

morphologic constraints in early Paleozoic crinoids: Paleobiology, v. 37, p. 214–

236. de Queiroz, K., and Gauthier, J., 1990, Phylogeny as a central principle in taxonomy:

phylogenetic definitions of taxon names: Systematic Biology, v. 39, p. 307–322. de Queiroz, K., and Gauthier, J., 1992, Phylogenetic taxonomy: Annual Review of

Ecology and Systematics, v. 23, p. 449–480. de Queiroz, K., and Gauthier, J., 1994, Toward a phylogenetic system of biological

nomenclature: Trends in Ecology and Evolution, v. 9, p. 27–31.

200

Donovan, S.K., 1988, The early evolution of the Crinoidea, in Paul, C.R.C., and Smith, A.B.,

eds, Echinoderm phylogeny and evolutionary biology: Oxford, Clarendon Press, p. 235–

244.

Donovan, S.K., and Harper, D.A.T., 2003, Llandovery crinoidea of the British Isles, including

description of a new species from the Kilbride Formation () of western Ireland:

Geological Journal, v. 38, p. 85–97.

Eichwald, E., von, 1840, Ueber das Silurische Schichtensystem in Ethlands, St. Petersburg, 200

p.

Etter, W., and Hess, H., 2015, Reviews and syntheses: the first records of deep-sea fauna

– a correction and discussion: Biogeosciences, v. 12, p. 6453–6462.

Feuda, R., and Smith, A.B., 2015, Phylogenetic signal dissection identifies the root of

: PLoS ONE, v. 10.5, e0123331.

Foote, M., 1996, On the probability of ancestors in the fossil record: Paleobiology, v. 22,

p. 141–151.

Foote, M., 1999, Morphological diversity in the evolutionary radiation of Paleozoic and

post-Paleozoic crinoids: Paleobiology, v. 25, p. 1–115.

Foote, M., and Raup, D.M., 1996, Fossil preservation and the stratigraphic ranges of taxa:

Paleobiology, v. 22, p. 121–140.

Foote, M., and Sepkoski, J.J., 1999, Absolute measures of the completeness of the fossil

record: Science, v. 398, p. 415–417.

201

Forey, P.L., Fortey, R.A., Kenrick, P., and Smith, A.B., 2004, Taxonomy and fossils: a

critical appraisal: Philosophical Transactions of the Royal Society of London B:

Biological Sciences, v. 359, p. 639–653.

Frest, T.J., Strimple, H.L., and McGinnis, M.R., 1979, Two new crinoids from the Ordovician of

Virginia and Oklahoma, with notes on pinnulation in the Disparida: Journal of

Paleontology, v. 53, p. 399–415.

Gahn, F.J., and Kammer, T.W., 2002, The cladid crinoid Barycrinus from the Burlington

Limestone (Early Osagean) and the phylogenetics of Mississippian botryocrinids:

Journal of Paleontology, v. 76, p. 123–133.

Gervais, F.L.P., 1835, In Guérin, ed., Dictionnaire pittoresque d’histoire naturelle et des

phénomènes de la nature, vol. 3, p. 48–49, Lenormand, Paris.

Goldfuss, G.A., 1826-44, Petrefacta Germaniae, tam ea, Quae in Museo Universitatis Regiae

Borussicae Fridericiae Wilhelmiae Rhenanea, serventur, quam alia quaecunque in Museis

Hoeninghusiano Muensteriano aliisque, extant, iconibus et descriiptionns illustrata. --

Abbildungen und Beschreibungen der Petrefacten Deutschlands und der Angränzende

Länder, unter Mitwirkung des Hern Grafen Georg zu Münster, herausgegeben von August

Goldfuss. v. 1 (1826-1833), Divisio prima. Zoophytorum reliquiae, p. 1-114; Divisio

secunda. Radiariorum reliquiae, p. 115-221 [Echinodermata]; Divisio tertia. Annulatorium

reliquiae, p. 222-242; v. 2 (1834-40), Divisio quarta. Molluscorum acephalicorum reliquiae.

I. Bivalvia, p. 65-286; II. Brachiopoda, p. 287-303; III. (1841-44), Divisio quinta.

202

Molluscorum gasteropodum reliquiae, p. 1-121; atlas of plates, 1–199, Düsseldorf, Arnz &

Co.

Gorzelak, P., Salamon, M.A., Trzęsiok, D., Lach, R., and Baumiller, T.K., 2015,

Diversity dynamics of post‐Palaeozoic crinoids–in quest of the factors affecting

crinoid macroevolution: Lethaia, v. 49, p. 231–244.

Gould, S.J., Gilinsky, N.L, and German, R.Z., 1987, Asymmetry of lineages and the

direction of evolutionary time: Science, v. 236, p. 1437–1441.

Guensburg, T.E., 2010, Alphacrinus new genus and origin of the disparid clade: Journal

of Paleontology, v. 84, p. 1209–1216.

Guensburg, T.E., 2012, Phylogenetic implications of the oldest crinoids: Journal of

Paleontology, v. 86, p. 455–461.

Guensburg, T.E., and Sprinkle, J., 2003, The oldest known crinoids (Early Ordovician,

Utah) and a new crinoid plate homology system: Bulletins of American

Paleontology, v. 364, 43 p.

Guensburg, T.E., and Sprinkle, J., 2007, Phylogenetic implications of the Protocrinoida:

Blastozoans are not ancestral to crinoids: Annals de Paleontologie, v. 93, p. 277–

290.

Guensburg, T.E., and Sprinkle, J., 2009, Solving the mystery of crinoid ancestry: new

fossil evidence of arm origin and development: Journal of Paleontology, v. 83, p.

350–364.

203

Guensburg, T.E., Blake, D.B., Sprinkle, J., and Mooi, R., 2016, Crinoid ancestry without

blastozoans: Acta Palaeontological Polonica, v. 61, p. 253–266.

Gyllenhaal, J.A., 1772, Beskrifning på de så kallade Crystall-äplen och kalkbollar, såsom

petreficerade Djur af Echini genus, eller dess närmaste slägtingar. Kongl. Svenska

Vetenskaps Academiens Handlingar: v. 33, p. 239–261.

Hall, J., 1847., Palaeontology of New York Volume I. Containing descriptions of the organic

remains of the lower division of the New-York system (equivalent of the Lower Silurian

rocks of Europe), Natural History of New York: New York, C. Van Benthuysen, Albany,

v. 6, 338 p.

Hall, J., 1852, Palaeontology of New York Volume II. Containing descriptions of the organic

remains of the lower middle division of the New-York system, Natural History of New

York: New York, D. Appleton & Co. and Wiley & Putnam; Boston, Gould, Kendall, &

Lincoln, v. 6, 362 p.

Hall, J., 1858, Chapter 8. Palaeontology of Iowa. in Hall, J. and Whitney, J.D., Report of the

Geological Survey of the state of Iowa. Embracing the results of investigations made during

portions of the years 1855, 56 &. 57, v. 1, part II; Palaeontology, p. 473–724.

Hall, J., 1859, Contributions to the palaeontology of Iowa, being descriptions of new species of

Crinoidea and other fossils: Geological Report of Iowa Supplement to vol. I, part II, p. 1-92.

Hall, J., 1866, Descriptions of new species of Crinoidea and other fossils from the Lower

Silurian strata of the age of the Hudson-River Group and Trenton Limestone: Albany,

privately distributed reprint,17 p.

204

Hays, B., Rouse, G., Thomas, J., and Messing, C., 2015. Can morphology support new

molecular phylogenies of (Crinoidea: Comatulida)?: 15th

International Echinoderm Conference Abstracts, p. 41–42.

Hemery, L.G., Roux, M., Ameziane, N., and Eleaume, M., 2013, High-resolution crinoid

phyletic inter-relationships derived from molecular data: Cahiers De Biologie

Marine, v. 54, p. 511–523.

Hess H., and Messing, C.G., 2011, Treatise on Invertebrate Paleontology Part T,

Echinodermata 2, Revised, Crinoidea 3, in Ausich, W.I., coordinating author,

Seldon, P., ed., University of Kansas and Paleontological Institute, Lawrence,

Kansas, 216 p.

Hess, H., Ausich, W.I., Brett, C.E., and Simms, M.J., 1999, Fossil crinoids:

Cambridge University Press, 275 p.

Holtz, T.R., 1996, Phylogenetic taxonomy of the Coelurosauria (Dinosauria: ):

Journal of Paleontology, v. 70, p. 536–538.

Holtz, T.R., 1998, A new phylogeny of the carnivorous dinosaurs: Gaia, v. 15, p. 5–61.

Jablonski, D., and Finarelli, J.A., 2009, Congruence of morphologically-defined genera

with molecular phylogenies: Proceedings of the National Academy of Sciences, v.

106, p. 8262–8266.

Jaekel, O., 1894, Über die Morphogenie und Phylogenic der Crinoiden: Sitzungsberichten der

Gesellschaft Naturforschender Freunde, Jahrgang 1894, v. 4, p. 101–121.

205

Jaekel, O., 1918, Phylogenie und System der Pelmatozoen: Paläeontologische Zeitschrift,

v. 3, p. 1–128.

Kammer, T.W., and Ausich, W.I., 1992, Advanced cladid crinoids from the Middle

Mississippian of the east-central United States. primitive-grade calyces: Journal of

Paleontology, v. 66, p. 461–480.

Kammer, T.W., and Ausich. W.I., 1996, Primitive cladid crinoids from upper Osagean-lower

Meramecian (Mississippian) rocks of east-central United States: Journal of Paleontology, v.

70, p. 835–866.

Kammer, T.W., and Gahn, F.J., 2003, Primitive cladid crinoids from the early Osagean

Burlington Limestone and the phylogenetics of Mississippian species of

Cyathocrinites: Journal of Paleontology, v. 77, p. 121–138.

Kammer, T.W., Sumrall, C.D., Zamora, S., Ausich, W.I., and Deline, B., 2013, Oral

region homologies in Paleozoic crinoids and other plesiomorphic pentaradiate

echinoderms: PLoS ONE, v. 8.11, e77989.

Kelley, P.H., Fastovsky, D.E., Wilson, M.A., Laws, R.A., and Raymond, A., 2013, From

paleontology to paleobiology: a half-century of progress in understanding life

history: Geological Society of America Special Paper, v. 500, p. 191–232.

Kelly, S.M., 1982, Origin of the crinoid orders Disparida and Cladida: possible inadunate

cup plate homologies: Third North American Paleontological Convention

(Montreal), Proceedings Abstracts, Montreal, Canada, p. 285–290.

206

Kelly, S.M., 1986, Classification and evolution of class Crinoidea: Fourth North

American Paleontological Convention, Abstracts, p. A23.

Kesling, R.V., and Paul, C.R.C., 1968, New species of Porocrinidae and brief remarks upon

these unusual crinoids: University of Michigan Contributions from Museum of

Paleontology, v. 22, p. 1–32.

Kirk, E., 1914, Notes on the fossil crinoid genus Homocrinus Hall. United States

National Museum Proceedings: v. 46, p. 473–483.

Kitazawa, K., Oji, T., and Sunamura, M., 2007, Food composition of crinoids (Crinoidea:

Echinodermata) in relation to stalk length and fan density: their paleoecological

implications: Marine Biology, v. 152, p. 959–968.

Kolata, D.R., 1982, Camerates, in Sprinkle, J., ed., Echinoderm faunas from the Bromide

Formation (Middle Ordovician) of Oklahoma: The University of Kansas

Paleontological Contributions, Monograph, v. 1, p. 170–205.

Lahaye, M.C., and Jangoux, M., 1987, The skeleton of the stalked stages of the comatulid

crinoid Antedon bifida (Echinodermata): Zoomorphology, v. 107, p. 58–65.

Lane, N.G., 1978, Historical review of classification of Crinoidea, in Moore, R.C., and

Teichert, C., eds., Treatise on Invertebrate Paleontology, Part T, Echinodermata 2,

Volume 2: Geological Society of America and University of Kansas Press,

Lawrence, p. T348–T358.

Lee, M.S., and Palci, A., 2015, Morphological phylogenetics in the genomic age: Current

Biology, v. 25, p. R922–R929.

207

Le Menn, J., and Spjeldnaes, N., 1996, Un nouveau crinoïde Dimerocrinitidae (Camerata,

Diplobathrida) de l'Ordovicien supérieur du Maroc: Rosfacrinus robustus nov. gen., nov.

sp.: Geobios, v. 29, p. 341–351.

Leuckart, R. von, 1848, Uber de Morphologie und die Verandtschaftsverhältnisse der

wibellosen Thiere: Friedrich ,Vieweg and Sohn, Braunschweig, 180 p.

Matsumoto, H., 1929, Outline of a classification of Echinodermata: Science Reports of

the Tohoku Imperial University, Sendai, Japan, Second Series (Geology): v. 8, p.

27–33.

McIntosh, G.C., 1984, Devonian cladid inadunate crinoids: family Botryocrinidae Bather,

1899: Journal of Paleontology, v. 58, p. 1260–1281.

McIntosh, G.C., 1986, Phylogeny of the dicyclic inadunate crinoid order Cladida: Fourth

North American Paleontological Convention Abstracts A, v. 31.

McIntosh, G.C., 2001, Devonian cladid crinoids: families Glossocrinidae Goldring, 1923,

and Rutkowskicrinidae new family: Journal of Paleontology, v. 75, p. 783–807.

Meek, F.B., 1873, Descriptions of invertebrate fossils of the Silurian and Devonian

systems: Ohio Geological Survey, v. 1, pt. 2, p. 1–243.

Meek, F., and Worthen, A.H., 1869, Descriptions of New Crinoidea and Echinoidea,

from the Carboniferous Rocks of the Western States, with a Note on the Genns

Onychaster: Proceedings of the Academy of Natural Sciences of Philadelphia, v.

21, p. 67–83.

Messing, C.G., and White, C.M., 2001, A revision of the Zenometridae (new rank)

(Echinodermata, Crinoidea, Comatulidina): Zoologica Scripta, v. 30, p. 159–180. 208

Meyer, D. L., and Macurda, D.B., 1977, Adaptive radiation of the comatulid crinoids:

Paleobiology, v. 3, p. 74–82.

Miller, J.S., 1821, A Natural History of the Crinoidea, or Lily-Shaped Animals; with

Observations on the Genera, Asteria, Euryale, Comatula and Marsupites: Bristol,

C. Frost, 150 p.

Miller, S.A., 1883, The American Palaeozoic fossils: a catalogue of the genera and

species, with names of authors, dates, places of publication, groups of books in

which found, and the etymology and significance of the words, and an

introduction devoted to the stratigraphical geology of the Palaeozoic rocks, 2nd

edition, Echinodermata, Cincinnati, Ohio, p. 247-334.

Miller, S.A. and Gurley, W.F.E., 1894, New genera and species of Echinodermata:

Illinois State Museum of Natural History, v. 5, 53 p.

Miquel, J. 1894. Note sur la géologie des terrains primaires du département de l’Hérault.

Le Cambrien et l’Arenig: Bulletin de la Société d’étude des Sciences naturelles de

Béziers v. 17, p. 5–36.

Mooi, R., and David, B., 1998, Evolution within a bizarre phylum; homologies of the

first echinoderms: American Zoologist, v. 38, p. 965–974.

Mooi, R., and David, B., 2008, Radial symmetry, the anterior/posterior axis, and

echinoderm Hox genes: Annual Review of Ecology, Evolution, and Systematics,

v. 39, p. 43–62.

209

Moore, R.C., 1962, Ray structures of some inadunate crinoids. University of Kansas

Paleontological Contributions, v. 5, 1–47.

Moore, R.C., and Laudon, L.R., 1943, Evolution and classification of Paleozoic crinoids:

Geological Society of America Special Paper, no. 46, 151 p.

Moore, R.C., and Teichert, C., eds., 1978, Treatise on Invertebrate Paleontology, Part T,

Echinodermata 2: Geological Society of America and University of Kansas Press,

Lawrence, Kansas, 1027 p.

Moore, R. C., Lalicker, C.G., and Fischer, A.G., 1952, Invertebrate Fossils: McGraw-Hill

Book Company, New York, 766 p.

Moore, R.C., Lane, N.G., and Strimple, H.L., 1978, Order Cladida Moore and Laudon,

1943, in Moore, R. C. and Teichert, C., eds: Treatise on Invertebrate

Paleontology, Pt. T, Echinodermata 2: Geological Society of America and

University of Kansas Press, Kansas, p. T578–T759.

Moore, R.C., Lane, N.G., Strimple, H.L., and Sprinkle, J., 1978, Systemetic Descriptions,

Crinoidea, Order Disparida, in Moore, R. C. and Teichert, C., eds: Treatise on

Invertebrate Paleontology, Pt. T, Echinodermata 2: Geological Society of

America and University of Kansas Press, Kansas, p. T520–T564.

Müller, J., 1841, Über die Gattungen und Arten der Comatulen: Archiv für

Naturgeschichte, v. 7, p. 179–189.

210

O’Malley, C.E., Ausich, W.I., and Chin, Y., 2016, Deep echinoderm phylogeny

preserved in organic molecules from Paleozoic fossils: Geology, v. 44, p. 379–

382.

Orbigny, A.D. d', 1849, Cours élémentaire de paléontologie et géologie Stratigraphiques:

v. 1, 299 p. Paris, Victor Masson.

Padian, K., Hutchinson, J.R., and Holtz, T.R., 1999, Phylogenetic definitions and

nomenclature of the major taxonomic categories of the carnivorous Dinosauria

(Theropoda): Journal of Vertebrate Paleontology, v. 19, p. 69–80.

Parks, W.A., 1908, On an occurrence of Hybocystites in Ontario: Ottawa Naturalist, v.

21, p. 232– 236.

Paul, C.R.C., 1968, Macrocystella Callaway, the earliest glyptocystitid cystoid:

Palaeontology, v. 11, p. 580–600.

Paul, C.R.C., 1972, Cheirocystella antiqua gen. et sp. nov. from the Lower Ordovician of

western Utah and its bearing on the evolution of the Cheirocrinidae (Rhombifera:

Glyptocystitida): Brigham Young University Geology Studies, v. 19, p. 15–63.

Paul, C.R.C., 1988, The phylogeny of the cystoids, in Paul, C.R.C., and Smith, A.B., eds.,

Echinoderm phylogeny and evolutionary biology: Clarendon Press, Oxford, p.

199–213.

Paul, C.R.C., and Smith, A.B., 1984, The early radiation and phylogeny of echinoderms:

Biological Reviews v. 59, p. 443–481.

Pennant, T., 1777, The British Zoology (4th edition, volume 4): London, Warrington.

211

Peters, S.E., and Ausich, W.I., 2008, A sampling-adjusted macroevolutionary history for

Ordovician-Early Silurian crinoids: Paleobiology, v. 34, p. 104–116.

Phillips, J., 1836, Illustrations of the geology of Yorkshire, or a description of the strata

and organic remains, Pt. 2, The Mountain Limestone districts, 2nd edition:

London, John Murray, p. 203–208.

Phillips, J., 1841, Figures and descriptions of the Palaeozoic fossils of Cornwall, Devon,

and West Somerset; observed in the course of the ordinance geological survey of

that district: London, Longmans, Brown, Green, and Longmans, 232 p.

Pisani, D., Feuda, R., Peterson, and Smith, A.B., 2012, Resolving phylogenetic signal

from noise when divergence is rapid: A new look at the old problem of

echinoderm class relationships: Molecular Phylogenetics and Evolution, v. 62, p.

27–34.

Potter, D., and Freudenstein, J.V., 2005, Character-based phylogenetic Linnaean

classification: taxa should be both ranked and monophyletic: Taxon, v. 54, p.

1033–1035.

Pyron, R.A., 2015, Post-molecular systematics and the future of phylogenetics: Trends in

Ecology and Evolution, v. 30, p. 394–389.

Rasmussen, H.W., 1978, Articulata, in Moore, R.C., and Teichert, K. eds., Treatise on

Invertebrate Paleontology. Part T, Echinodermata 2, vol. 2: The Geological

Society of America, and University of Kansas Press, Boulder and Lawrence, p.

T813–T997.

212

Reich, A., Dunn, C., Akasaka, K., and Wessel, G., 2015, Phylogenomic analyses of

Echinodermata support the sister groups of and : PLoS ONE,

v. 10, e0119627.

Rouse, G.W., Jermiin, L.S., Wilson, N.G., Eeckhaut, I., Lanterbecq, D., Oji, T., Young,

C.M., Browning, T., Cisternas, P., Helgen, L.E., Stuckey, M., and Messing, C.G.,

2013, Fixed, free, and fixed: the fickle phylogeny of extant Crinoidea

(Echinodermata) and their Permian–Triassic origin: Molecular Phylogenetics and

Evolution v. 66, p. 161–181.

Rouse, G.W, Carvajal, J., Oji, T., and Messing, C.G., 2015, Further insights into extant

crinoid phylogeny from molecular sequence data: 15th International Echinoderm

Conference Abstracts, p. 67–68.

Roux, M., 1987, and of recent stalked crinoids as a

model for the fossil record, in Jangoux, M., and Lawrence, J.M., eds., Echinoderm

studies, vol. 2: A. A. Balkma, Rotterdam, the Netherlands, p. 1–53.

Roux, M., Eleaume, M., Hemery, L.G., and Ameziane, N., 2013, When morphology

meets molecular data in crinoid phylogeny: a challenge: Cahiers de Biologie

Marine, v. 54, p. 541–548.

Ruta, M., Coates, M.I., and Quicke, D.L.J., 2003, Early relationships revisited:

Biological Reviews, v. 78, p. 351–345.

Schuchert, C., 1900, On the Lower Silurian (Trenton) fauna of Baffin Land: Proceedings

of the U. S. National Museum, v. 22, p. 143–178.

213

Sereno, P.C., 1997, The origin and evolution of dinosaurs: Annual Review of Earth and

Planetary Sciences, v. 25, p. 235–489.

Sereno, P.C., 1999, Definitions in phylogenetic taxonomy: critique and rationale:

Systematic Biology, v. 48, p. 329–351.

Sereno, P.C., 2005, The logical basis of phylogenetic taxonomy: Systematic Biology, v.

54, p. 595–619.

Sereno, P.C., McAllister, S., and Brusatte, S.L., 2005, TaxonSearch: a relational database

for suprageneric taxa and phylogenetic definitions: PhyloInformatics, v. 8, p. 1–

21.

Sevastopulo, G.D., and Lane, N.G., 1988, Ontogeny and phylogeny of disparid crinoids,

in Paul, C.R.C., and Smith, A.B., eds, Echinoderm phylogeny and evolutionary

biology: Oxford, Clarendon Press, p. 245–253.

Shibata, T.F., Sato, A., Oji, T., and Akasaka, K., 2008, Development and growth of the

feather star Oxycomantus japonicus to sexual maturity: Zoological Science, v. 25,

p. 1075–1083.

Sibley, C.G., 1994, On the phylogeny and classification of living birds: Journal of Avian

Biology, v. 25, p. 87–92.

Simms, M.J., 1988, The phylogeny of post-Palaeozoic crinoids, in Burke, R. D., P. V.

Mladenov, P. Lambert, and R. L. Parsley, eds, Echinoderm Biology: Balkema,

Rotterdam, p. 97–102.

214

Simms, M.J., and Sevastopulo, G.D., 1993, The origin of articulate crinoids:

Palaeontology, v. 36, p. 91–109.

Simpson, G.G., 1944, Tempo and Mode in Evolution: Columbia University Press, New

York.

Siverts-Doreck, H. 1953. Sous-classe 4. Articulata, p. 756–765. In J. Priveteau (ed.),

Traité de paléontologie. Volume 3. Masson and Cie, Paris.

Smith, A.B., 1984, Classification of the Echinodermata: Palaeontology, v. 27, p. 431–

459.

Smith, A.B., 1985, Cambrian eleutherozoan echinoderms and the early diversification of

edrioasteroids: Palaeontology, v. 28, p. 87–89.

Smith, A.B., 1988, Patterns of diversification and extinction in early Palaeozoic

echinoderms: Palaeontology, v. 31, p. 799–828.

Smith, A.B., 1990, Evolutionary diversification of the echinoderms during the Early

Paleozoic: Systematics Association, Special Volume 42, p. 256–286.

Smith, A.B., 1994, Systematics and the Fossil Record: Documenting Evolutionary

Patterns: Blackwell Science, Oxford, 223 p.

Smith, A.B., 2005, The pre‐radial history of echinoderms: Geological Journal, v. 40, p.

255-280.

Smith, A.B., 2008, Duterostomes in a twist: the origins of a radical new body plan:

Evolution and Development, v. 10, p. 493–503.

215

Smith, A.B., and Jell, P.A., 1990, Cambrian edrioasteroids from Australia and the origin

of starfishes: Queensland Museum Memoir, v. 28, p. 715–778.

Smith, A.B., and Zamora, S., 2013, Cambrian spiral-plated echinoderms from

reveal the earliest pentaradial body plan: Proceedings of the Royal Society B, v.

280:20131197, doi.org/10.1098/rspb.2013.1197

Soul, L.C., and Friedman, M., 2015, Taxonomy and phylogeny can yield comparable

results in comparative paleontological analyses: Systematic Biology, v. 64, p.

608–620.

Springer, F., 1911, On a Trenton echinoderm fauna: Canada Department Mines Memoir

no. 15-P, 70 p.

Springer, F., 1913, Crinoidea, in Zittel, K.A. von, ed., Text-book of paleontology

(translated and edited by C. R. Eastman), 2nd edit: London, Macmillan & Co.,

Ltd., v. 1, p. 173–243.

Springer, F., 1920, The Crinoidea Flexibilia: Smithsonian Institution Publication no.

2501, 486 p.

Sprinkle, J., 1973, Morphology and evolution of blastozoan echinoderms: Museum of

Comparative Zoology Special Publication, Harvard University, 283 p.

Sprinkle, J., 1976, Classification and phylogeny of “pelmatozoan” echinoderms:

Systematic Zoology, v. 25, p. 83–91.

216

Sprinkle, J., 1982a, Hybocrinus, in Sprinkle, J., ed., Echinoderm faunas from the Bromide

Formation (Middle Ordovician) of Oklahoma: The University of Kansas Paleontological

Contributions, Monograph 1, p. 119–128.

Sprinkle, J., 1982b, Cylindrical and globular rhombiferans, in Sprinkle, J., ed., Echinoderm

faunas from the Bromide Formation (Middle Ordovician) of Oklahoma: The University

of Kansas Paleontological Contributions, Monograph 1, p. 231–273.

Sprinkle, J., and Moore, 1978, Hybocrinida, in Moore, R. C. and Teichert, C., eds:

Treatise on Invertebrate Paleontology, Pt. T, Echinodermata 2: Geological Society

of America and University of Kansas Press, Kansas, p. T564–T574.

Stukalina, G.A., 1980, New species of quadrilaterial from the Ordovician of Kazakhstan,

Urals, and eastern European platform, in Stukalina, G.A., ed., New species of

ancient and invertebrates of the USSR, 5: Akademiia Natik SSSR,

Paleontologischeskii Institut. Moscow. p. 88-95.

Summers, M.M., Messing, C.G., and Rouse, G.W., 2014, Phylogeny of Comatulidae

(Echinodermata: Crinoidea: Comatulida): a new classification and an assessment

of morphological characters for crinoid taxonomy: Molecular Phylogenetics and

Evolution, v. 80, p. 319–339.

Sumrall, C.D., 1997, The role of fossils in phylogenetic reconstructions of the

Echinodermata: Paleontological Society Papers, v. 3, p. 267–288.

Sumrall, C.D., 2008, The origin of Lovéns Law in glyptocystitoid rhombiferans and its

bearing on plate homology and heterochronic evolution of the hemicosmitoid

217

peristomial border, in Ausich, W.I., and Webster, G.D., eds., Echinoderm

Paleobiology: Indiana University Press, Bloomington, p. 28–41.

Sumrall, C.D., 2010, A model for elemental homology for the peristome and ambulacra

in blastozoan echinoderms, in Harris, L.G., Bottger, S.A., Walker, C.W., and

Lesser, M.P., eds., Echinoderms: Durham, CRC Press, New York, p. 269–276.

Sumrall, C.D., 2014, Echinoderm phylogeny – the path forward: Geological Society of

America Abstracts with Programs, v. 46, p. 78.

Sumrall, C.D., and Waters, J.A., 2012, Universal Elemental Homology in

glyptocystitoids, hemicosmitoids, coronoids, and blastoids: steps toward

echinoderm phylogenetic reconstruction in derived blastozoans: Journal of

Paleontology, v, 68, p. 956–972.

Sumrall, C.D., and Wray, G.A., 2007, Ontogeny in the fossil record: diversification of

body plans and the evolution of "abberant" symmetry in Paleozoic echinoderms:

Paleobiology, v. 33, p. 149–163.

Telford, M.J., Low, C.J., Cameron, C.B., Ortega-Martinez, O., Aronowicz, J., Oliveri, P.,

and Copley, R.R., 2014, Phylogenetic analysis of echinoderm class relationships

supports Asterozoa: Proceedings of the Royal Society B, v. 281, 20140479.

Ubaghs, G., 1963, Rhopalocystis destombesi n. g., n. sp. Eocrinoide de l’Ordovicien

inferieur (Tremadocien superieur) du Sud marocain: Notes du service geologic du

Maroc, v. 23, p. 25–44.

218

Ubaghs, G., 1978, Origin of crinoids, in Moore, R. C. and Teichert, C., eds, Treatise on

Invertebrate Paleontology, Pt. T, Echinodermata, v. 2 (2): Geological Society of

America and University of Kansas Press, Lawrence, p. T275–T281.

Ulrich, E.O., 1925, The lead, zinc, and fluorspar deposits of Western Kentucky; Chapter

2, Stratigraphic geology: U. S. Geological Survey, Professional Paper 36, p. 22‒

71.

Wachsmuth, C., and Springer, F., 1880–1886, Revision of the Palaeocrinoidea.

Proceedings of the Academy of Natural Sciences of Philadelphia Pt. I. The

families Ichthyocrinidae and Cyathocrinidae (1880), p. 226‒378 (separate repaged

p. 1–153). Pt. II. Family Sphaeroidocrinidae, with the sub-families Platycrinidae,

Rhodocrinidae, and Actinocrinidae (1881), p. 177‒411 (separate repaged, p. 1‒

237). Pt. III, Sec. 1. Discussion of the classification and relations of the brachiate

crinoids, and conclusion of the generic descriptions (1885), p. 225‒364 (separate

repaged, 1‒138). Pt. III, Sec. 2. Discussion of the classification and relations of

the brachiate crinoids, and conclusion of the generic descriptions (1886), p. 64‒

226 (separate repaged to continue with section 1, 139‒302).

Wachsmuth, C., and Springer, F., 1891, The perisomic plates of the crinoids: Proceedings

of the Academy of Natural Sciences of Philadelphia, v. 41, p. 345–39.

Wachsmuth, C., and Springer, F., 1897, The North American Crinoidea Camerata:

Harvard College Museum of Comparative Zoology Memoir, v. 20 and 21, 897 p.

219

Webster, G.D., and Jell, P.A., 1999, New Permian crinoids from Australia: Memoirs of

the Queensland Museum, v. 33, p. 349–359.

Webster, G.D., and Lane, N.G., 2007, New Permian crinoids from the Battleship Wash patch

reef in southern Nevada: Journal of Paleontology, v. 81, p. 951–965.

Wetherby, A.G., 1880, Remarks on the Trenton Limestone of Kentucky, with

descriptions of new fossils from that formation and the Kaskaskia (Chester)

Group, Sub-carboniferous: Journal of the Cincinnati Society of Natural History, v.

3, p. 144–160.

White, C. A., 1881, Fossils of the Indiana rocks: Indiana Department of Statistics and

Geology, Annual Report, v. 2, p. 471–522.

Wiley, E.O., and Lieberman, B.S., 2011, Phylogenetics: theory and practice of

phylogenetic systematics: John Wiley & Sons, 406 p.

Wright, D.F., 2015a, Fossils, homology, and “Phylogenetic Paleo-ontogeny”: a

reassessment of primary posterior plate homologies among fossil and living

crinoids with insight from developmental biology: Paleobiology, v. 41, p. 570–

591.

Wright, D.F., 2015b, Testing the taxonomic structure of Paleozoic pan-cladid crinoids: a

statistical approach using the fossilized birth-death process and Bayesian

phylogenetic inference: Geological Society of America Abstracts with Programs,

v. 7, p. 854.

220

Wright, D.F., in press, Bayesian estimation of fossil phylogenies and the evolution of

early to middle Paleozoic crinoids (Echinodermata): Journal of Paleontology.

Wright, D.F., and Ausich, 2015, W.I., From the stem to the crown: phylogeny and

diversification of pan-cladid crinoids, in Zamora, S. and Rábano, I., eds., Progress

in Echinoderm Paleobiology: Cuadernos del museo Geominero, 19, Instituto

Geológico y Minero de España, p. 199–202.

Zamora, S., 2013, Morphology and phylogenetic interpretation of a new Cambrian

edrioasteroid (Echinodermata) from Spain: Palaeontology, v. 56, p. 421–431.

Zamora, S., and Rahman, I.A., 2014, Deciphering the early evolution of echinoderms

with Cambrian fossils: Palaeontology, v. 57, p. 1105–1119.

Zamora, S., and Smith, A.B., 2011, Cambrian stalked echinoderms show unexpected

plasticity of arm construction: Proceedings of the Royal Society of London B:

Biological Sciences, rspb20110777.

Zamora, S., Sumrall, C.D., and Vizcaino, D., 2013, Morphology and ontogeny of the

Cambrian edrioasteroid echinoderm Cambraster cannati from western

Gondwana: Acta Palaeontologica Polonica, v. 58, p. 545-559.

Zamora, S., Sumrall, C.D., Zhu, X., Lefebvre, B., 2016, A new stemmed echinoderm

from the of China and the origin of Glyptocystitida (Blastozoa,

Echinodermata): Geological Magazine, doi:10.1017/S001675681600011X.

Zittel, K.A., von, 1895, Grundzüge der Palaeontologie (Palaeozoologie), 1st edition: R.

Oldenbourg, München, 971 p.

221

Table 4.1 Species name, least inclusive clade, and first appearances interval for each taxon depicted in Figure 4.1.

222

Least Inclusive Species First Occurrence of Species Clade Actinocrinites triacontadactylus Miller, 1821 Monobathrida Mississippian (Tournaisian) Adelphicrinus fortuitus Guensburg and Sprinkle, 2003 Camerata Ordovician (Tremadocian) Alphacrinus mansfieldi Guensburg, 2010 Disparida Ordovician (Tremadocian) Antedon bifida (Pennant, 1777) Articulata Recent Apektocrinus ubaghsi Guensburg and Sprinkle, 2009 Pentacrinoidea Ordovician (Tremadocian) Archaeocrinus lacunosus (Billings, 1857) Diplobathrida Ordovician (Katian) Carabocrinus radiatus Billings, 1857 Porocrinida Ordovician (Sanbian) Cupulocrinus heterocostalis (Hall, 1847) Flexibilia Ordovician (Katian) Dendrocrinus longidactylus Hall, 1852 Eucladida Silurian (Wenlockian) Endoxocrinus parrae (Gervais, 1835) Articulata Recent Eknomocrinus wahwahensis Guensburg and Sprinkle, 2003 Camerata Ordovician (Tremadocian) Glyptocrinus decadactylus Hall, 1847 Monobathrida Ordovician (Katian) Hybocrinus conicus Billings, 1857 Hybocrinida Ordovician (Sanbian) Hybocystites problematicus Wetherby, 1880 Hybocrinida Ordovician (Katian) Pentacrinites fossilis Blumenbach, 1804 Articulata Triassic () Porocrinus conicus Billings, 1857 Porocrinida Ordovician (Katian) Rhodocrinites verus Miller, 1821 Diplobathrida Mississippian (Tournaisian) Rosfacrinus robustus LeMenn and Spjeldnaes, 1996 Eucamerata Ordovician (Katian) Synbathocrinus conicus Phillips, 1836 Disparida Mississippian (Tournaisian) Taxocrinus macrodactylus (Phillips, 1841) Flexibilia Devonian ()

Table 4.1

223

Table 4.2 Revised rank-based classification of the Crinoidea. † indicates an extinct taxon.

224

Class Crinoidea Miller, 1821

†Subclass Camerata Wachsmuth and Springer, 1885

‘Stem eucamerates’ (e.g., Eknomocrinus)

Infraclass Eucamerata Cole, in press

Order Diplobathrida Moore and Laudon, 1943

Order Monobathrida Moore and Laudon, 1943

Crinoidea incertae sedis: †Protocrinoidea Guensburg and Sprinkle, 2003

Subclass Pentacrinoidea Jaekel, 1894

†‘Stem inadunates’ (e.g., Apektocrinus)

Infraclass Inadunata Wachsmuth and Springer, 1885

†Parvclass Disparida Moore and Laudon, 1943

Order Eustenocrinida Ulrich, 1925

Order Maennilicrinida Ausich, 1998b

Order Tetragonocrinida Stukalina, 1980

Order Calceocrinida Meek and Worthen, 1869

Disparida incertae sedis: “Homocrinida” Kirk, 1914

Disparida incertae sedis: “Myelodactyla” S. A. Miller, 1883

Disparida incertae sedis: “Pisocrinoidea” Ausich and Copper, 2010

Parvclass Cladida Moore and Laudon, 1943

†Superorder Porocrinoidea Wright, in press

Order Porocrinida Miller and Gurley, 1894

Continued

225

Table 4.2 continued

Order Hybocrinida Jaekel, 1918

†Superorder Flexibilia Zittel, 1895 (Cupulocrinus d’Orbigny, 1849)

Order Taxocrinida Springer, 1913

Order Sagenocrinida Springer, 1913

Magnorder Eucladida Wright, in press

†Superorder Cyathoformes new superorder

Cyathoformes incertae sedis: “Cyathocrinida” Bather, 1899

Cyathoformes incertae sedis: “Dendrocrinida” Bather, 1899

Cyathoformes incertae sedis: “Poteriocrinida” Jaekel, 1918

Eucladida incertae sedis: †“Ampelocrinida” Webster and Jell, 1999

Superorder Articulata Miller, 1821

†Order Holocrinida Jaekel, 1918 Rasmussen, 1978

†Order Encrinida Matsumoto, 1929

†Order Sieverts-Doreck, 1953

†Order Uintacrinida Zittel 1879

†Order Roveacrinida Sieverts-Doreck, 1953

Order Sieverts-Doreck, 1953

Order Rasmussen, 1978

Order Isocrinida Sieverts-Doreck, 1953

Order Comatulida A. H. Clark, 1908

226

Figure 4.1 Taxa representing major crinoid clades: (1) Pentacrinites fossilis

Blumenbach, 1804, articulate, from Goldfuss (1831); (2) Taxocrinus colletti White, 1881, flexible, from Springer (1920); (3) Actinocrinites jugosus (Hall, 1859), monobathrid camerate, from Wachsmuth and Springer, 1897; (4) Synbathocrinus swallovi Hall, 1858, disparid, from Wachsmuth and Springer (1897); (5) Dendrocrinus caduceus Hall, 1866, eucladid, from Meek (1873); (6) Hybocystites eldonensis Parks, 1908, hybocrinid, from

Springer (1911); (7) Porocrinus shawi Schuchert, 1900, porocrinid, from Kesling and

Paul (1968); (8) Archaeocrinus microbasalis (Billings, 1857), diplobathrid camerate, from Wachsmuth and Springer (1897). All scale bars 0.5 cm and applicable as indicated.

227

Fig. 4.1

228

Figure 4.2 Cladogram depicting phylogenetic relationships among species used to define major clades within the Crinoidea. Terminal tips correspond to species listed in

Table 1. Clades given stem-based definitions are indicated with a downward facing arrow, whereas clades given node-based definitions are indicated with a circle. Note that many clades named are nested inside other more inclusive clades. Graphical notation of stem- and node-defined clades follows Sereno (2005).

229

Fig. 4.2

230

Chapter 5: Ecologic innovation and phenotypic constraint in the evolutionary radiation of

Paleozoic crinoids

Abstract.—Evolutionary radiations are one of the most discussed features in the history of life. The patterns and processes underlying phenotypic diversification during such radiations have been much debated by both paleontologists and comparative biologists. In this chapter, I take advantage of recently developed methods to test the relationships between phylogeny-based rates of morphologic evolution, taxonomic diversification, and morphospace occupation in the ~200 million year evolutionary radiation of Paleozoic crinoids. Focusing on the Eucladida,

I find evidence for highly heterogeneous evolutionary dynamics, including variation in morphologic rates and shifts in evolutionary mode. The early history of eucladid diversification is characterized by “early burst” dynamics, but a late Paleozoic peak in morphologic rates is not associated with the diversification of a rapidly radiating subclade. Instead, the late peak in rates likely involved multiple smaller radiations linked to ecologic change. Results indicate phenotypic diversification is more complex than models commonly assumed in comparative biology, at least over geologic timescales—highlighting the need for continued synthesis between fossil and phylogenetic approaches to macroevolution.

231

Introduction

“How fast, as a matter of fact, do animals evolve in nature?” –G.G. Simpson (1944)

It is often stated that the major features of Earth’s biodiversity are the result of successive evolutionary radiations spanning the geologic history of life (Simpson, 1953). Adaptive radiation, a process whereby species rapidly diverge from a common ancestor and increase in phenotypic disparity as a result of ecologic opportunity, is the most commonly invoked type of evolutionary radiation and is backed by a rich theory and wealth of empirical case studies (Simpson, 1953; Schluter, 2000; Erwin, 2007; Losos, 2010).

Adaptive radiation theory predicts that rates of phenotypic evolution are initially high during the early stages of radiation, such as when a lineage enters a new adaptive zone, but subsequently decline as a clade ages because niche space becomes saturated

(Simpson, 1953; Schluter, 2000). A corollary of this prediction is that variation in morphologic disparity should peak early in a clade’s history. A possible alternative explanation to adaptive radiation is that a clade may achieve early maximum disparity for reasons not related to ecologic opportunities, but because phenotypic diversification was later limited by developmental constraints or time-heterogeneous selective pressures

(Wagner, 1996, 2000; Erwin, 2007). Nevertheless, patterns of so-called “early burst”

(Harmon et al., 2010) diversifications are commonly observed in paleontologic data

(Foote, 1994, 1999; Erwin, 2007; Wagner, 2010; Hughes et al., 2013; Oyston et al.,

2015), and recent radiations of well-studied living clades are broadly consistent with 232

predictions from adaptive radiation theory (Baldwin and Sanderson, 1998; Grant and

Grant, 2002). In contrast with the fossil record, phylogenetic comparative data rarely

support the early burst model and instead suggest that instances of evolutionary radiation

are more frequently characterized either by patterns of constrained diversification around

an adaptive peak or random diffusion through morphospace (Harmon et al., 2010).

Although the lack of support for adaptive radiation in comparative data may reflect

difficulties in testing early burst models (Harmon et al., 2010; Slater et al., 2010; Slater

and Pennell, 2014), it nevertheless highlights the need to consider alternative models of

phenotypic diversification associated with evolutionary radiations (Erwin, 1992; Simões

et al., 2015) (Fig. 1).

For the last several decades, paleobiologists have tested conceptual models of radiation patterns by measuring temporal variation in morphologic disparity in fossil lineages (Foote, 1992, 1997; Erwin, 2007; Wagner, 2010). Disparity profiles are then compared with concomitant trends in taxonomic diversity to determine which alternative model best fits the observed trends (Foote, 1997; Jablonski, in press). A common assumption is that fluctuations in disparity primarily reflect underlying rates of morphologic change. Although this prodigious research program has generated important

insights into patterns of morphologic diversification (Wagner, 2010), such as whether the

high early disparity patterns observed in the fossil record are consistent with ecologic

processes or other mechanisms (Foote, 1997; Erwin, 2007), there are several causes for

concern as to whether or not disparity measures provide a general context for inferring

phenotypic rates of evolution. For example, alternative evolutionary processes with

233

strikingly different rates of morphologic evolution can yield equivocal disparity profiles

(Foote, 1996; Slater, 2015). Moreover, even under simple models of trait evolution, such

as Brownian motion, sister clades with different rate dynamics may theoretically obtain

identical disparities because of the way trait variation is partitioned among subclades

(O’Meara et al., 2006). Thus, assessing the rate of morphologic evolution requires the

explicit use of phylogenetic trees (Foote, 1996). When combined with time-calibrated

phylogenies, disparity estimates can more explicitly be used to test relationships between

rates of change and morphospace occupation (Wagner, 1997).

This chapter takes advantage of recently developed phylogeny-based methods to test two questions pertinent to how diversification trajectories unfold over geologic time.

Using the fossil record of Paleozoic eucladid crinoids (Echinodermata), I test whether their ~200 million year radiation exhibits temporal dynamics predicted by adaptive radiation theory. I then compare per-lineage-million-year rates of character change with patterns of morphospace occupation to test whether morphologic evolution was driven primarily by ecologic opportunities, innovation, or phenotypic constraints. The Eucladida

(Wright, in press) are a major clade of crinoids and provide a model system for addressing these questions. Eucladids are both ecologically and taxonomically diverse

(Webster, 2013), and are characterized by a well sampled fossil record (Foote and Raup,

1996; Foote and Sepkoski, 1999). Importantly, fossil crinoids preserve morphologic features useful for inferring evolutionary relationships from phenotypic data (Wright et al., in press) and their can be compared with living species to infer feeding ecology and life history (Ausich, 1980; Kitazawa et al., 2007; Kammer, 1985, 2008).

234

These features make fossil crinoids ideal for testing relationships between patterns of

morphologic diversification and adaptive zone occupation, as well as evaluating ecologic

differences among species shape evolutionary radiations over geologic timescales.

Results and Discussion

Rates of morphologic evolution for Ordovician through Permian eucladids were

calculated using 92 discrete morphologic traits spanning the entire eucladid body plan.

Two phylogeny-based methods were used. The first method divides the number of

character changes occurring within ~7 million year bins by the sum of within-bin branch

length durations and uses likelihood ratio tests to infer whether intervals are characterized

by statistically elevated or depressed rates. The second method estimates the median percent amount of character change per million years using a Bayesian relaxed morphologic clock model. Although recent work has illuminated phylogenetic relationships among crinoids (Ausich et al., 2015; Cole, in press; Wright, in press; Wright et al., in press), the phylogeny of Paleozoic eucladids remains elusive, and they have not received comprehensive systematic treatment since publication of the crinoid Treatise

(Moore et al., 1978). A phylogenetic analysis of a morphologic character matrix coded for 81 species did not recover any of the order-level taxa listed in the crinoid Treatise

(Moore et al., 1978). However, the validity of these orders has long been questioned and

results are broadly consistent with previous taxonomic recommendations (McIntosh,

235

1986, 2001; Sevastopulo and Lane, 1988; Simms and Sevastopulo, 1993; Kammer and

Ausich, 1992, 1993, 1996). The maximum clade credibility (MCC) tree resulting from of

a Bayesian analysis using a relaxed morphologic clock and a time-varying fossilized

birth-death process (FBD) tree prior is well resolved near the base of the tree, but

elsewhere characterized by low support (see Materials and Methods section). To account

for uncertainties in evolutionary relationships and divergence times, all rate analyses

were performed over random samples of time-calibrated trees from the Bayesian

posterior distribution.

Both maximum likelihood and morphologic-clock methods reveal major

fluctuations in the rate of morphologic evolution for eucladid crinoids across the

Paleozoic (Fig. 5.2). Morphologic rates for Paleozoic eucladids were at their highest early

in the clade’s history and exhibit a long-term secular decrease in mean rates of evolution

through time (Spearman’s Rho = -0.46, P = 0.022). The Late Ordovician to middle

Silurian is characterized by elevated rates of character change that are approximately

twice as high as most subsequent intervals of the Paleozoic. Rates significantly drop

during the Devonian, reaching their Paleozoic minimum during the -Famennian stages. Although rates modestly increased during the early Carboniferous, they do not rise above background levels and are comparable to those of the late Silurian to early

Devonian. Rates were elevated throughout much of the late Carboniferous (-

Kasimovian), with a burst of morphologic evolution occurring during the Moscovian stage that resulted in the only significant post-Silurian peak in morphologic rates (Fig.

5.2). Although the early to middle Paleozoic radiation of eucladids is consistent with an

236

“early burst” type pattern, the episodic peak during the late Carboniferous suggests

morphologic diversification during the late Paleozoic was driven by different

evolutionary dynamics.

It is important to rule out possible biases that may influence these patterns

because variation in rate estimates can arise even when underlying rates of evolution are

constant. For example, elevated rates of evolution may be inferred simply because taxa

were more densely sampled and therefore more changes can be recorded. However, a

phylogeny-based diversity curve (Fig. 5.3, [lower figure]) shows a major peak during the

early Carboniferous, which is not associated with high rates of change. Similarly,

diversity was much lower during the Ordovician to middle Silurian than most of the

Paleozoic. A comparison of the phylogeny-based diversity of sampled lineages with the

number of all known valid eucladid genera reveals late Carboniferous eucladids are

proportionally undersampled in the dataset (Webster, 2013) (Fig. 5.3 [cf. middle and

lower figures]). Thus, the late Carboniferous peak in morphologic rates would be

expected to be even higher if species had been sampled in proportion to taxonomic

richness.

Sampling biases arising from incompleteness of the fossil record could also affect

rate estimates. Paleontologic metrics assuming uniform preservation (Foote and Raup,

1996) indicate sampling probabilities for Paleozoic eucladids are slightly higher than

estimates for all Ordovician to Devonian crinoid genera (per-interval preservation probability = 0.60, sampling rate = 0.12) (Foote and Raup, 1996) suggesting the eucladid record is ~92% complete when evaluated at the family level. Crinoids as a whole exhibit

237

different post-mortem disarticulation rates (Meyer et al., 1989), but patterns of

taphonomic degradation are similar within major clades such as the Eucladida (Ausich et

al., 1999). Thus, there is no a priori reason to suspect substantial variation in preservation

potential among taxa analyzed. However, non-uniform preservation of the rock record

itself may also influence paleontologic patterns, especially over geologically long

timescales (107-108 years) (Holland, 2016). A stochastic model of non-uniform sampling

was incorporated directly into the phylogenetic analysis and used to estimate time-

varying rates of fossil sampling (see Materials and Methods). Although sampling

fluctuated through time (Fig. 5.4C), there is a nonsignificant correlation between

sampling rate and inferred rates of character change (Spearman’s Rho = 0.155, P =

0.459). In summary, the inferred rates of morphologic evolution are robust to a number of

potential biases that may affect rate inferences, including variation in tree topology,

branch lengths, divergences times, and incomplete or non-uniform fossil sampling.

Other phylogeny-based disparity studies with paleontologic perspectives have suggested that bursts of morphologic change occurring after a clade’s early history are associated with major ecologic shifts and diversification of one or more subclades (Slater,

2012, 2015; Close et al., 2015; Hopkins and Smith, 2015). Feeding ecology is the primary control on niche differentiation among crinoids (Ausich, 1980; Kitazawa et al.,

2007). Crinoids are passive suspension feeders that depend on the current-driven nutrient supply to feed and have evolved a variety of ecomorphologic traits to help capture food particles and reduce competition among species (Ausich, 1980; Kitazawa et al., 2007;

Baumiller, 2008). Two eucladid subclades, cyathoform and cladoform crinoids (sensu

238

Kammer and Ausich, 1992, 1996; Webster, 2012), represent different functional groups and are each strongly supported as monophyletic (posterior probability = 1) (see

Materials and Methods) (Fig. 5.2). Although cyathoform and cladoform species often co- occur in the same paleocommunties (e.g., on the same bedding plane), they are characterized by traits reflecting to different ecologic and environmental conditions (Ausich, 1980; Kammer, 1985, 2008). Cladoform crinoids have terminal arm appendages called pinnules that serve to increase filtration fan density, which allows for feeding to take place at higher current velocities (Kammer, 1985). In addition, cladoforms have muscular appendages in the arms that increase motility and enable shifts in feeding posture (Kammer, 1985; Baumiller and Messing, 2007). In contrast with cladoforms, cyathoform crinoids almost universally lack pinnules and have ligamentary articulations with limited motility.

The early history of eucladid evolution predominantly reflects patterns of cyathoform diversification. The cyathoform subclade is characterized by deaccelerating rates through time (Spearman’s Rho = -0.930, P < 0.001) (Fig. 5.5), consistent with expectations of an early burst model. In contrast, variation in rates after the Devonian is most strongly associated with cladoforms. A small rise in rates is coincident with the gradual assembly of the cladoform body plan during the Devonian (Fig. 5.2), and rates of change within the subclade are higher early in their history than in subsequent intervals

(Fig. 5.5). However, long-term rates of change among cladoforms show no evidence for a secular trend (Spearman’s Rho = 0.042, P = 0.873), and the highest rates of evolution occur much later in their history (Fig. 5.5). Thus, rates of morphologic evolution

239 characterizing cladoform diversification do not support an early burst pattern of phenotypic diversification despite substantial ecomorphologic innovation.

Major trends in crinoid evolution have been attributed to long-term effects of biotic interactions and/or species ecology (Meyer and Macurda, 1977; Baumiller, 1993;

Ausich et al., 1994; Kammer and Ausich, 2006; Sallan et al., 2011; Gorzelak et al.,

2015). The pattern of shifting subclade contributions to the broader eucladid radiation is consistent with paleoecologic observations regarding ecologic abundance, taxonomic diversity, and turnover between these groups (Ausich et al., 1994). Thus, the ecologic differences between cladoform and cyathoform eucladids may have resulted in different macroevolutionary dynamics. Indeed, background rates of morphologic evolution are higher among cladoforms than cyathoforms (Mann-Whitney U test, W = 231, P < 0.001).

It is possible that higher rates in cladoform eucladids may have resulted from increased competition with other pinnulate crinoid clades (e.g., camerates) that share similar functional ecologic traits. However, eucladids and camerate crinoids broadly differ in their habit preferences (Kammer and Ausich, 2006), which suggests elevated rates of morphologic evolution in cladoform eucladids were driven by processes other than (or in addition to) competition. Similarly, the rate of adaptive phenotypic evolution is expected to be greater in clades experiencing increased predation pressure (Vermeij,

1987). Increased predator-prey dynamics have previously been linked with macroevolutionary patterns in Paleozoic crinoids (Signor and Brett, 1984; Sallan et al.,

2011), but a comparison of diversification trajectories found no significant relationship between eucladids and their probable vertebrate predators (Sallan et al., 2011). Lastly,

240

higher rates for cladoform eucladids may have resulted from environmental factors that

influence ecologic specialization, which has been linked with higher rates of evolution

(Vrba, 1987). Baumiller (1993) found evidence for higher rates of taxonomic

diversification for camerate crinoids with more finely-filtered fan morphologies and attributed the pattern to long-term stenotopic effects on diversification. Although cladoforms also have dense filtration fans, the evolution of muscular articulations in their arms facilitated a greater range of feeding postures that led to cladoforms being able to inhabit a broader range of environments than other crinoid lineages (Baumiller, 1993;

Holterhoff, 1997). Thus, stenotopy in eucladids is not easily equated with stenotopy in other crinoids. The concept of evolution occurring within adaptive zones (Simpson,

1953) provides a possible explanation. If cladoforms occupied a broader adaptive zone than cyathoforms, then evolution within subzones (Simpson, 1953) would lead to ecologic specialization and increased rates of phenotypic change in cladoform subclades.

A morphospace analysis of the complete character matrix shows disparity within cladoform eucladids is much greater than among cyathoforms (Fig. 5.6), especially when patterns of morphospace occupation are evaluated over time (Fig. 5.7). Cyathoform eucladids rapidly expanded in morphospace during their explosive Ordovician-Silurian radiation and reached their maximal disparity during the Silurian. Although cyathoforms continued to diversify throughout the Paleozoic, they never expanded into new regions of morphospace outside those that originally evolved during the early Paleozoic. Thus, the early history of cyathoform evolution is consistent with an early burst-type pattern and with expectations from adaptive radiation theory. However, their subsequent

241

diversification during the middle to late Paleozoic is more consistent with increased

constraints on morphologic evolution that limited expansion into previously unexplored

regions of morphospace.

In contrast, cladoform eucladids expanded in disparity during the Devonian and

reached peak morphologic diversity during the Mississippian, occupying vast regions of

morphospace spanning multiple Simpsonian subzones (Figs. 5.6-5.7). The distribution of

eucladids in morphospace was highly asymmetric during the Pennsylvanian, as the

advanced cladoform subzone became saturated and very few regions outside this zone

were occupied. However, the advanced cladoform cluster diminished in importance by

the Permian. Cyathoforms and intermediate zone cladoforms re-radiated into pre-

explored regions of morphospace, whereas extinction of closely spaced taxa occurred

within the advanced cladoform zone. The broad distribution in eucladids in morphospace

coupled with extinction of intermediate forms ultimately led to the Permian being the

time of maximal morphologic disparity in the Paleozoic radiation of eucladids (Fig. 5.3).

The preceeding discussion has focused on rates of morphologic evolution and

overall trends in morphospace occupation through time. However, most theoretical

models describing the dynamics of evolutionary radiations also make predictions about

patterns of taxonomic diversification (Rabosky and Adams, 2012; Jablonski, in press).

Taxonomic diversification is an exponential process, whereas morphologic

diversification is approximately linear under constant rate Brownian motion assumptions

(Foote, 1996). If taxonomic diversity is plotted on a log scale in diversity-disparity space, then exponential taxonomic diversification coupled within a Brownian motion-like

242

diffusion process for morphologic evolution would predict a linear fit between diversity

and disparity measures (Jablonski, in press).

The relationship between morphologic disparity and taxonomic diversity in the

radiation of Paleozoic eucladids is depicted in Figure 5.8. Diversity and disparity both

rapidly increase from the Ordovician through the Devonian and rise significantly above

the 1:1 diagonal (Fig. 5.3 (upper), Fig. 5.9). Likelihood ratio tests confirm that rates of

morphologic evolution were high throughout the Ordovician to early Silurian (Fig. 5.2),

with high rates accompanying the rapid expansion of cyathoform eucladids in

morphospace. When coupled with the concomitant rise in taxonomic diversity during this

interval, these patterns suggest the early history of eucladid radiation was characterized

by an early burst-like model of morphologic evolution and is consistent with adaptive

radiation theory. The end-Ordovician extinction dramatically altered the structure of

crinoid communities (Ausich et al., 2004), but it is not clear how the extinction may have

affected eucladid diversification. Net taxonomic diversification fell dramatically during

the end Ordovician (Fig. 5.4), but the overall diversity trajectory for eucladids shows a

nearly constant rate increase from the Ordovician through the end of the Silurian (Fig.

5.3). It is possible the end-Ordovician extinction played a role in sustaining rates of

morphologic evolution for such geologically long intervals. For example, the opening of

niche space that followed the end-Ordovician may have allowed their early Paleozoic radiation to continue in a protracted diversification. Alternatively, the extinction event may have led eucladids to undergo a post-extinction adaptive radiation and the observed

geologically sustained rates may reflect two contiguous diversification events.

243

Post-Silurian dynamics reveal more complex patterns. Despite significant expansion into new regions of morphospace that correspond to the initial radiation of cladoforms during the Devonian (Fig. 5.7), the rise in disparity was not accompanied by a significant increase in morphologic rates. Instead, rates of morphologic evolution do not differ significantly from background for much of the Devonian. Further, the Late

Devonian marks the only interval in eucladid diversification with evidence for significantly low rates of character change (Fig. 5.2). Thus, morphologic innovation associated with the assembly of the cladoform body plan did not result in an “early burst”-like peak corresponding with the origination of the clade (Hopkins and Smith,

2015). Instead, morphologic diversification from the Devonian to early Carboniferous is more similar to a Brownian motion-like process. Brownian motion diffusion through morphospace would be expected to result in increased disparity even if morphologic change was occurring at a low, constant rate. This suggests the Devonian increase in morphologic disparity should not be interpreted as high underlying rates of change, but instead as a fundamental shift in the underlying mode of evolution.

The early Carboniferous (Mississippian) features a dramatic increase in taxonomic diversity, marking the so-called “Age of Crinoids” (Kammer and Ausich,

2006) (Figs. 5.3, 5.8). Notably, taxonomic diversification among eucladids was most dramatic among cladoforms during this interval (Kammer and Ausich, 2006). The macroevolutionary lag (Jablonski, 2008) between the assembly of the cladoform body plan and subsequent taxonomic diversification may be explained, in part, by the late

Devonian global collapse of reefs. The demise of -stromatoporoid dominated reefs

244 facilitated a major environmental transition from abundant, more restricted rimmed carbonate platforms to widespread open ramp settings in epicontinental seas, resulting in a substantial increase in habitat space favorable for crinoids (Kammer and Ausich, 2006).

Global geographic expansion of eucladids during the Mississippian may have led to increased speciation propensity (Simões et al., 2015), resulting in a substantial net increase in taxonomic diversification (Kammer and Ausich, 2006). Although disparity slightly increased during the Mississippian, morphospace became more tightly clustered in cladoform zones. Rates of evolution may have continued to follow a Brownian motion- like process, but with an increasing role for constraint in morphologic diversification rather than unbounded diffusion in morphospace.

The rise in morphologic rates of evolution during the late Carboniferous

(Pennsylvanian) is coincident with a similar rise in taxonomic diversity (Figs. 5.2-5.3), culminating in both maximal genus-level diversity and the only post-Silurian rate peak supported by likelihood ratio tests during the Middle Pennsylvanian. However, the

Pennsylvanian did not exceed the Mississippian in net taxonomic diversification, which suggests the diversity peak is a result of increased taxonomic longevity rather than elevated origination (Fig. 5.4). Episodic, elevated rates can result from early burst like patterns of adaptive radiation within a single lineage (Hopkins and Smith, 2015; Slater,

2015). Examination of rates inferred on the MCC from the Bayesian morphologic clock analysis tree instead reveals multiple lineages were evolving at high rates during the

Pennsylvanian (Fig. 5.2). Interestingly, the Pennsylvanian is associated with an overall decrease in disparity despite the elevated rates of character change (Fig. 5.8).

245

Diversification during the Pennsylvanian was dominated by taxa within a narrow region

of morphospace contained within the advanced cladoform subzone (Fig. 5.7). The

centripetal direction of changes implied by the MCC phylomorphospace indicates

morphologic diversification was highly constrained within the cladoform subzone.

That multiple clades were rapidly evolving similar morphologies is suggestive of a shared response to similar ecologic or selective pressures. For example, Paleozoic eucladids exhibit a temporal trend in the calyx plates in pinnulate eucladids. This trend is best expressed in Carboniferous eucladids and is correlated with a reduction in body size

(Kammer, 2008). These trends may reflect an adaptive paedomorphic response to rapidly fluctuating environmental conditions and increased sediment disturbance (Kammer,

2008; Wright, 2015). Thus, morphologic evolution during the Pennsylvanian is suggestive of another major shift in underlying evolutionary mode. Constrained diversification within a limited, circumscribed region of morphospace is consistent with a macroevolutionary Ornstein-Uhlenbeck (OU) model of morphologic evolution. Although phenotypic optima for an OU-like dynamic may have theoretically existed any time after

the advanced cladoform zone became available, the evolutionary attractor driving

bounded morphologic evolution during this interval was much stronger than any time

before or after the Pennsylvanian. This and other instances of replicate radiations within

Simpsonian subzones, such in fossil canids (Slater, 2015b), strongly suggests that

adaptive evolution may result in patterns of constrained phenotypic diversification during

instances of ecologic radiation (Erwin, 1992).

246

If these results can be generalized to apply to other clades, then the extreme heterogeneity observed in Paleozoic eucladid evolution strongly suggests that commonly used models of phenotypic diversification (Fig. 5.1) do not adequately capture the dynamics of evolutionary radiations, at least those occurring over long timescales. In addition, paleontologic inferences from studies of disparity through time may be misled if a strong relationship between rates of evolution and morphospace occupation is assumed.

Rather than dismay over challenges presented, I advocate that where possible, the data and methods of both fields be combined and synthesized to help better understand the pattern and process in macroevolutionary biology. This study highlights how integrating phylogeny-based estimates of morphologic evolution with traditional paleontologic measures of disparity provides a more nuanced understanding of phenotypic diversification, especially during major radiations spanning vast amounts of geologic time.

Materials and Methods

Morphologic and stratigraphic data

I compiled a data matrix of 92 morphologic characters for 81 species of Paleozoic eucladid crinoids and one outgroup (Appendix A). The majority of characters included overlap with Wright (in press), but I excluded any characters with invariant distributions among sampled taxa and included new characters (and character states) to accommodate

247 morphologic variation in middle to upper Paleozoic eucladids. The Eucladida (Wright, in press) range from the Ordovician (Katian) to the present (Wright et al., in press), and their Paleozoic fossil record includes more than 500 genera (Webster, 2013). To obtain an analytically tractable sample of species for analysis, representative species were chosen from nominal families (Moore et al., 1978; Webster, 2013) to capture broad taxonomic, morphologic, and preservational gradients among taxa. In most cases, species sampled were type species of a type genus. However, some geologically older species were sampled, especially when an older species was represented by more completely preserved specimens and was available for coding. All missing or inapplicable characters were coded as “?” (~20% of the cells). The mean number of unscored cells per taxon

(18.2%) and per character (16.5%) are both relatively low and similar to one another.

Thus, taxonomic and character-based biases introduced by missing data are likely minimal for this dataset. Notably, this dataset is more complete than comparable studies testing phylogeny-based rates of morphologic evolution over geologic timescales (e.g.

Close et al., 2015; Hopkins and Smith, 2015).

Although the character list was constructed primarily for use in phylogenetic analysis, it is also amenable for studying character-based morphologic disparity. A comparison of disparity metrics using cladistic matrices of discrete characters indicates cladistic data are often good proxies for summarizing morphologic variation

(Hetherington et al., 2015), and a number of empirical studies use cladistic matrices for testing general hypotheses of morphologic evolution (Close et al., 2015; Hopkins and

Smith, 2015; Lloyd, 2016). are often excluded from phylogenetic

248

datasets because they are not parsimony-informative (i.e., they have the same length on

all trees), but are included in disparity studies because they contribute to the total

variation in morphologic form. However, this study includes parsimony-informative characters and a small number of autapomorphies. The inclusion of autapomorphies in model-based (i.e., probabilistic) phylogenetic analyses may improve phylogenetic inferences via more accurate estimation of branch lengths and rates of change (Lewis,

2001). Characters were coded by examining specimens housed in museum collections, including the Smithsonian Institution, Field Museum of Natural History, the Natural

History Museum (London), and the Lapworth Museum of Geology. Most species were coded using multiple specimens, such as types and paratypes in their type-series.

Additional data was collected by examining the primary taxonomic literature.

All numerical dates and names for geologic intervals are taken from the 2016

International Chronostratigraphic Chart (Cohen et al., 2013, updated). Stratigraphic data constraining species temporal ranges were obtained from an updated version of Webster

(2013). Most analyses consist of data binned to the stage-level and range from the

Ordovician (Katian) to Permian (either the Wordian or Wuchiapingian depending on the analysis). However, Silurian data was binned according to the series-level in an attempt to make the temporal duration between time bins more comparable and reduce difficulties in fine-scale stratigraphic correlation when assigning fossil occurrence dates (see

Webster, 2013). Following Cohen et al. (2013), the Carboniferous Period was subdivided into subperiods: the early Carboniferous (Mississippian) and late Carboniferous

(Pennsylvanian).

249

Phylogenetic analysis

To obtain a distribution of time-calibrated phylogenies, I used a Bayesian fossil tip-dating

approach as implemented in MrBayes 3.2.6 (Ronquist et al., 2012). A Markov model of

character evolution (Lewis, 2001) was used with character coding set to variable and a 4-

category gamma distribution was applied to account for rate variation among characters.

All characters were treated as unordered. Rate variation among lineages was estimated

using the uncorrelated, independent gamma rates (IGR) relaxed morphologic clock model

(Lepage et al., 2007), with a broad exponential hyperprior (λ = 10) on the variance of the

gamma distribution. A truncated Gaussian prior was placed on the base rate of the clock

(mean = 0.01, standard deviation = 0.1).

The time-varying (piecewise-constant) skyline implementation of the sampled

ancestor fossilized birth-death process (SA-FBD) was used as a prior distribution on

branch lengths (Heath et al., 2014; Gavryushkina et al., 2014; Zhang et al., 2016; Wright,

in press).Thus, SA-FBD parameters were estimated across broad geologic intervals,

corresponding to the Late Ordovician-Silurian, Devonian, Mississippian, and the

Pennsylvanian-mid Permian. For each interval, SA-FBD parameters were assigned an exponential prior (λ = 10) on net diversification, a flat (Beta[1,1]) prior on relative extinction, and a Beta(2,2) distribution on fossil sampling. Because taxon sampling is highly incomplete at the species-level but broadly covers all major eucladid higher taxa, the diversified sampling correction of Zhang et al. (2016) was applied. The proportion of

‘extant’ species sampled at the end of the SA-FBD process was set to 0.008, which is

250 based on the assumption that middle to late Permian crinoid species diversity was at least as high as that of today (Donovan et al., 2016).

Bayesian fossil tip-dating requires information on the temporal ages and/or ranges of fossil samples (Gavryushkina et al., 2014; Wright, in press). Preliminary tip-dating analyses had difficulty converging (according to MrBayes diagnostics) when species occurrences were assigned uniform distributions according to the duration of their first appearing stages in Webster (2013). However, convergence was more easily reached when point occurrences were used. Because data is lacking for precise numerical estimates of species first-appearances (Webster, 2013), the mid-point of the time bin for each species’ first appearance was used as a point constraint on occurrence times. Based on both stratigraphic and phylogenetic evidence, the split between Eucladida and their nearest outgroup occurred during the earliest part of the Katian stage of the Ordovician

(Webster, 2013; Wright, in press). Thus, the tree age prior was set to correspond to an early Katian divergence.

To assist the analysis, a series of partial and hard topological constraints based on results of Wright (in press) were applied. Maximum parsimony and topology-only

Bayesian analyses (using PAUP* [Swofford, 2002] and MrBayes) were used to identify additional sets of taxa that could be constrained, and clades recovered by both analyses were given partial constraints. Two Markov chain Monte-Carlo runs with four chains were run for 30 million generations, sampling trees every 3000 steps. MCMC convergence was assessed using methods described in chapter 3 and 50% of the samples were discarded as burn-in, resulting in a posterior distribution of 5,000 time-calibrated

251

phylogenetic trees. The maximum clade credibility tree (MCC) depicting the posterior

probability of nodes and 95% highest density probability for divergences are presented in

Figure 5.2.

Rates of morphologic evolution

Rates of character change were inferred using two methods: a maximum-likelihood approach that estimates rates of character changes occurring along branches of time- calibrated trees (Lloyd et al., 2012), and a Bayesian relaxed morphologic clock analysis in MrBayes (Ronquist et al., 2012).

Lloyd et al. (2012) developed a novel approach to estimated rates of morphologic evolution in a likelihood framework that accounts for the effect of missing data, which was later refined by Brusatte et al. (2014) and Lloyd, 2016. This method models character changes as a Poisson process with a rate parameter (λi) and uses likelihood ratio

tests to identify intervals with significantly high or low rates of evolution. Ancestral

states at internal nodes are first estimated for all comparable characters from time-

calibrated phylogenies using maximum likelihood (Lloyd, 2016), and then rates of

morphologic evolution are calculated as the number of changes occurring along a branch

(i.e., number of parsimony steps) divided by the product of time (measured in millions of

years) and completeness (measured as the proportion of the observed comparable

characters). Following Lloyd et al. (2012), let Xi denote the number of character changes

th occurring in the i interval, ci is completeness, and ti represents the temporal duration of a

252

time bin in millions of years. Thus, the probability of observing x character changes is

given as (Lloyd et al., 2012)

( ) P(X = x|λ ) = . i i −𝜆𝜆 𝑡𝑡 𝑐𝑐 ! 𝑥𝑥 𝑒𝑒 𝑖𝑖 𝑖𝑖 𝑖𝑖 𝜆𝜆𝑖𝑖𝑡𝑡𝑖𝑖𝑐𝑐𝑖𝑖 𝑥𝑥

Likelihood ratio tests (LRT) are used to test whether λ for given time bin is significantly

high or lower than λ estimated from the remaining pooled bins. The significance level α

for LRTs was set to 0.01. The Benjamini-Hochberg test was used to correct for multiple comparisons and their associated type I errors (Benjamini and Hochberg, 1995; Lloyd et al., 2012). Because the rates inferred from this method are sensitive to the tree topology and branch lengths, I calculated rates of evolution over a random sample of 100 trees from the Bayesian posterior distribution of time-calibrated trees. Thus, the results are robust across many alternative hypotheses of evolutionary relationships and node age estimates. If a time bin contained >55% of rate estimates with statistically significant

(i.e., P < 0.01) high or low values across the posterior sample of trees, it was regarded as having rates significantly different from background.

Rates inferred using a Bayesian relaxed morphologic clock were estimated in

MrBayes 3.2.6 (Ronquist et al., 2012) using the same prior distribution settings described above for phylogenetic analysis. To account for uncertainty, percent character changes per million years were obtained by taking the median values of rates sampled in the

Bayesian posterior distribution of trees. Time-calibrated branch lengths and rates of evolution are estimated simultaneously. It should be noted that recovery of time- 253

heterogeneous rate dynamics using this approach runs opposite of methodological biases

because it assumes morphologic evolution can be fit to a clock-like model (Lloyd, 2016).

Standing diversity and diversification rates

Taxonomic diversity and taxonomic rates of origination and extinction were

estimated from data contained within a recently updated compilation of 505 genus-level eucladid occurrences spanning the Paleozoic (Webster, 2013). Where appropriate, I modified the data listed in Webster (2013) to update stratigraphic correlations and alpha taxonomy. Generic diversity was estimated using the first and last appearances recorded by Webster (2013) with the common range-through assumption (Foote and Miller, 2007).

Uncertainty in the number of genera occurring per bin was estimated using a Monte-

Carlo analysis consisting of 1,000 bootstrap replicates.

Per-capita rates of taxonomic diversification were estimated using equations

presented in Foote (2000) for tabulations of biostratigraphic ranges and is based on the

assumption of exponential survivorship (Raup, 1985). The extinction rate is calculated as

the proportion of taxa alive at the start of an interval that survived at least to the end of

the interval. Thus, the extinction rate is q = -ln[(Nbt) / (Nb)] / Δt, where q is the extinction

rate per-lineage-million years, Nbt is the number of taxa that cross both the beginning and end of the interval, Nb is the number of taxa that cross the bottom boundary, and Δt is the

temporal duration of the interval in units of 106 years. Similarly, the origination rate is

based on the proportion of taxa present at the end of an interval that already existed prior

to the beginning of the interval. The per-lineage-million year origination rate is calculated

254

as p = -ln[(Nbt) / (Nt)] / Δt, where p is the origination rate, Nt is the number of taxa

crossing the top boundary, and other terms are the same as described above. Net

taxonomic diversification per interval is simply p – q, and taxonomic turnover is q/p (Fig.

5.4). Although improved analytical methods exist to estimate standing diversity and

taxonomic diversification (e.g., Alroy, 2010, 2016), these approaches require detailed

information on fossil occurrences rather than tabulation of first and last appearances.

Discipline-based efforts to compile such data exist (e.g., paleobiodb.org). However, at

present these databases do not have sufficient taxonomic and temporal coverage to satisfy

the requisite data needed to implement occurrence-based approaches. Nevertheless, the

taxonomic rates presented herein represent estimates based on the most comprehensive,

taxonomically vetted dataset of eucladid crinoids ever compiled and stratigraphic range-

based methods often provide a good proxy for diversification dynamics (Foote, 2000;

Peters and Ausich, 2008).

In addition to taxonomic rates, the time-series of SA-FBD parameters estimated during phylogenetic analysis can also be used to infer diversification dynamics and compared with paleontologic estimates. These parameters are summarized in Figure 5.4.

Although differences exist between taxonomic and SA-FBD estimates (e.g., the SA-FBD represent coarse scale averages whereas the finer resolution taxonomic rates are subject to sampling biases [see Alroy, 2010]), broad pattern trajectories are similar when paleontologic estimates are averaged over timescales comparable with FBD estimates

(Fig. 5.4). For example, the median rate of taxonomic turnover for the Devonian Period is

0.73, which is close to the SA-FBD estimate of 0.71.

255

Phylogenetic standing diversity was calculated as the median diversity across 500

randomly sampled time-calibrated trees from the posterior distribution (Bapst, 2012).

Uncertainty around the median was characterized using two-tailed 95% upper and lower

quantiles (see Bapst, 2012).

Morphologic disparity and morphospace occupation

Morphologic disparity was calculated as the average squared distance between species in

morphospace, which is proportional to the multivariate variance for discrete characters

(Foote and Miller, 2007). All characters were unordered and equally weighted. Phenetic

distances between taxa (Sij) were calculated using Gower’s coefficient (Gower, 1971),

which rescales distances by the dividing by the number of comparable characters:

Sij = 𝑣𝑣 ∑𝑘𝑘=1 𝑆𝑆𝑖𝑖𝑖𝑖𝑖𝑖 𝑣𝑣 ∑𝑘𝑘=1 𝛿𝛿𝑖𝑖𝑖𝑖𝑖𝑖

where δijk is equal to one if a character k can be coded for species i and j and otherwise equal to zero (Lloyd, 2016). Temporal variation in morphospace occupation was assessed by assigning taxa to period-level bins and calculating disparity for each interval.

Uncertainty in disparity estimates was evaluated using 1,000 bootstrap replicates.

To visually inspect morphospace distances between taxa in the complete dataset, a principal coordinate analysis (PCO) was conducted on the distance matrix. A phylomorphospace of the MCC tree was projected on species PCO scores distributed along the first two PCO axes, with ancestral state values inferred using maximum 256 likelihood (Revell, 2012) (Fig. 5.6). The first two PCO axes comprise 34.1% of the variance. Lower PCO 1 scores are associated with higher filtration fan densities, whereas higher scores are associated with lower filtration fan densities (Spearman’s Rho = -0.772,

P < 0.001). Fan density reflects the presence of pinnules, number of distal arms, and arm type (e.g., uniserial/biserial). PCO 2 is inversely proportional to calyx complexity, where calyx complexity is defined as the number of primary calyx plates in the cup. Taxa with fewer calyx plates have higher PCO 2 scores (Spearman’s Rho = -0.713, P < 0.001). The spatial distribution of taxonomic clusters in principal coordinate space (Fig. 5.6) closely corresponds with previously identified grades of body plan organization in eucladids

(Kammer and Ausich, 1992, 1993, 1996; Webster and Maples, 2006; Wright, 2015).

257

References

Alroy, J., 2010, Geographical, environmental and intrinsic biotic controls on Phanerozoic

marine diversification. Palaeontology, v. 53, p. 1211-1235.

Alroy, J., 2016, On a conservative Bayesian method of inferring extinction: Paleobiology,

v. 42, p. 670-679.

Ausich, W.I., 1980, A model for niche differentiation in Lower Mississippian crinoid

communities: Journal of Paleontology, v. 54, p. 273-288.

Ausich, W.I., Kammer, T.W., and Baumiller, T.K., 1994, Demise of the Middle

Paleozoic crinoid fauna: A single extinction event or rapid faunal turnover?:

Paleobiology, v. 20, p. 345‒361.

Ausich, W.I., Kammer, T.W., Rhenberg, E.C., and Wright, D.F., 2015, Early phylogeny

of crinoids within the pelmatozoan clade: Palaeontology, v. 58, p. 937–952.

Baldwin, B.G., and Sanderson, M.J., 1998, Age and rate of diversification of the

Hawaiian silversword alliance (Compostiae): Proceedings of the National

Academy of Sciences, v. 95, p. 9402-9406.

Bapst, D.W., 2012, paleotree: an R package for paleontological and phylogenetic

analyses of evolution. Methods in Ecology and Evolution: v. 3, p. 803-807.

Baumiller, T.K., 1993, Survivorship analysis of Paleozoic Crinoidea: effect of filter

morphology on evolutionary rates: Paleobiology, v. 19, p. 304-321.

Baumiller, T.K., 2008, Crinoid ecological morphology: Annual Review of Earth and

Planetary Sciences, v. 36, p. 221-249.

258

Baumiller, T.K., and Messing, C.G., 2007, Stalked crinoid locomotion and its ecological

and evolutionary implications: Palaeontologica Electronica, v. 10, 2A:10p.

Benjamini, Y., and Hochberg, Y., 1995, Controlling the false discovery rate: a practical

and powerful approach to multiple test: Journal of the Royal Statistical Society,

Series B, v. 57, p. 289-300.

Brusatte, S.L., Lloyd, G.T., Wang, S.C., and Norell, M.A., 2014, Gradual assembly of

avian body plan culminated in rapid rates of evolution across the -

transition: Current Biology, v. 24, p. 2386-2392.

Butler, M. A., and King, A.A., 2004, Phylogenetic Comparative Analysis: A Modeling

Approach for Adaptive Evolution: The American Naturalist, v. 164, p. 683-695.

Close, R.A., Friedman, M., Lloyd, G.T. and Benson, R.B., 2015, Evidence for a mid-

Jurassic adaptive radiation in mammals: Current Biology, v. 25, p. 2137‒2142.

Cohen, B.L., Ameziane, N., Eleaume, M., and de Forges, B.R., 2004, Crinoid phylogeny:

a preliminary analysis (Echinodermata: Crinoidea): Marine Biology, v. 144, p.

605–617.

Cole, S.R., in press, Phylogeny and morphologic evolution of the Ordovician Camerata

(Class Crinoidea, Phylum Echinodermata): Journal of Paleontology.

Donovan, S.K., Webster, G.D., and Waters, J.A., 2016, A last peak in diversity: the

stalked echinoderms of the Permian of Timor: Geology Today, v. 32, p. 179-185.

Erwin, D.H., 1992, A preliminary classification of evolutionary radiations: Historical

Biology, v. 6, p. 133-147.

259

Erwin, D.H., 2007, Disparity: morphological pattern and developmental context:

Palaeontology, v. 50, p. 57-73.

Foote, M. 1992, Paleozoic record of morphological diversity in blastozoan echinoderms:

Proceedings of the National Academy of Sciences: v. 89, p. 7325-7329.

Foote, M., 1994, Morphological disparity in Ordovician–Devonian crinoids and the early

saturation of morphological space: Paleobiology, v. 20, p. 320–344.

Foote, M., 1995, Morphological diversification of paleozoic crinoids: Paleobiology, v.

21, p. 273-299.

Foote, M., 1996, Ecological controls on the evolutionary recovery of post-Paleozoic

crinoids: Science, v. 274, p. 1492-1495.

Foote, M., 1997, Estimating taxonomic durations and preservation probability:

Paleobiology, v. 23, p. 278–300.

Foote, M., 1999, Morphological diversity in the evolutionary radiation of Paleozoic and

post-Paleozoic crinoids: Paleobiology, v. 25, p. 1–115.

Foote, M., 2000, Origination and extinction components of taxonomic diversity: general

problems: Paleobiology, v. 26, p. 74-102.

Foote and Miller, 2007, Principles of Paleontology. Macmillian.

Foote, M., and Raup, D.M., 1996, Fossil preservation and the stratigraphic ranges of taxa:

Paleobiology, v. 22, p. 121–140.

Foote, M., and Sepkoski, J.J., 1999, Absolute measures of the completeness of the fossil

record: Nature, v. 398, p. 415–417.

260

Gavryushkina, A., Welch, D., Stadler, T., and Drummond, A. J., 2014, Bayesian

inference of sampled ancestor trees for epidemiology and fossil calibration: PLoS

Computational Biology, v. 10, e1003919.

Gorzelak, P., Salamon, M.A., Trzęsiok, D., Lach, R., and Baumiller, T.K., 2015,

Diversity dynamics of post‐Palaeozoic crinoids–in quest of the factors affecting

crinoid macroevolution: Lethaia, v. 49, p. 231–244.

Gower, J.C., 1971, A general coefficient of similarity and some of its properties:

Biometrics, v. 27, p. 857-871.

Grant, P.R., and Grant,B.R., 2002, Adaptive radiation of Darwin’s finches: American

Journal of Science, v. 90, p. 130-139.

Harmon, L. J., J. B. Losos, T. J. Davies, R. G. Gillespie, J. L. Gittleman, W. B. Jennings,

K. H. Kozak, M. A. McPeek, F. Morena-Roark, T. J. Near, A. Purvis, R. E.

Ricklefs, D. Schluter, J. A. Schulte II, O. Seehausen, B. L. Sidlauskas, O. Torres-

Carvajal, J. T. Weir, and A. Ø. Mooers. 2010. Early bursts of body size and shape

evolution are rate in comparative data: Evolution, v. 64, p. 2385-2396.

Heath, T.A., Huelsenbeck, J.P. and Stadler, T., 2014, The fossilized birth–death process

for coherent calibration of divergence-time estimates: Proceedings of the National

Academy of Sciences, v. 111, p. E2957–E2966.

Hetherington, A.J., Sherratt, E., Ruta, M., Wilkinson, M., Deline, B., and Donoghue,

P.C., 2015, Do cladistic and morphometric data capture common patterns of

morphological disparity?: Palaeontology, v. 58, p. 393-399

261

Holland, S.M., 2016, The non-uniformity of fossil preservation: Philosophical

Transactions of the Royal Society B, v. 371, 20150130.

Holterhoff, P.F., 1997, Filtration models, guilds, and biofacies: Crinoid of

the Stanton Formation (Upper Pennsylvanian), midcontinent, North America:

Palaeogeography, Palaeoclimatology, Palaeoecology, v. 130, p. 177-208.

Hopkins, M.J. and Smith, A.B., 2015, Dynamic evolutionary change in post-Paleozoic

echinoids and the importance of scale when interpreting changes in rates of

evolution: Proceedings of the National Academy of Sciences, v. 112, p.3758–

3763.

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, v. 110, p. 13875-13879.

Jablonski, D., 2008, Biotic interactions and macroevolution: extensions and mismatches

across scales and levels: Evolution, v. 62, p. 715-739.

Jablonski, in press, Macroevolutionary theory, in Scheiner, S.M., and Mindell, D.P.,

(eds.), The theory of evolution. Chicago: Press.

Kammer, T.W., Aerosol filtration theory applied to Mississippian deltaic crinoids:

Journal of Paleontology, v. 59, p. 551-560.

Kammer, T.W. 2008, Paedomorphosis as an adaptive response in pinnulate cladid

crinoids from the Burlington Limestone (Mississippian, Osagean) of the

Mississippi Valley. Pp. 177-195 in W. I. Ausich and G. D. Webster (eds.),

Echinoderm Paleobiology, University of Indiana Press, Bloomington.

262

Kammer, T.W., and Ausich, W.I., 1992, Advanced cladid crinoids from the Middle

Mississippian of the east-central United States. primitive-grade calyces: Journal of

Paleontology, v. 66, p. 461-480.Kammer and Ausich, 1993

Kammer, T.W., and Ausich. W.I., 1996, Primitive cladid crinoids from upper Osagean-lower

Meramecian (Mississippian) rocks of east-central United States: Journal of Paleontology, v.

70, p. 835-866.

Kammer, T.W., and Ausich, W.I., 1992, The “age of crinoids”: a Mississippian

biodiversity spike coincident with widespread carbonate ramps. Palaios, v. 2, p. 238-

248.

Kitazawa, K., Oji, T., and Sunamura, M., 2007, Food composition of crinoids (Crinoidea:

Echinodermata) in relation to stalk length and fan density: their paleoecological

implications: Marine Biology, v. 152, p. 959-968.

Lepage, T., Bryant, D., Philippe, H. and Lartillot, N., 2007, A general comparison of

relaxed molecular clock models: Molecular Biology and Evolution, v. 24, p.

2669‒2680.

Lewis, P.O., 2001, A likelihood approach to estimating phylogeny from discrete

morphological character data: Systematic biology, v. 50, p. 913‒925.

Lloyd, G.T., 2016, Estimating morphological diversity and tempo with discrete character-

taxon matrices: implementation, challenges, progress, and future directions:

Biological Journal of the Linnaean Society, v. 118, p. 131-151.

263

Lloyd, G.T., Wang, S.C. and Brusatte, S.L., 2012, Identifying heterogeneity in rates of

morphological evolution: discrete character change in the evolution of lungfish

(Sarcopterygii; Dipnoi): Evolution, v. 66, p.330‒348.

Losos, J.B., 2010, Adaptive radiation, ecological opportunity, and evolutionary

determinism: The American Naturalist, v. 175, p. 623-639.

McIntosh, G.C., 1986, Phylogeny of the dicyclic inadunate crinoid order Cladida: Fourth

North American Paleontological Convention Abstracts A, v. 31.

McIntosh, G.C., 2001, Devonian cladid crinoids: families Glossocrinidae Goldring, 1923,

and Rutkowskicrinidae new family: Journal of Paleontology, v. 75, p. 783–807.

Meyer, D.L., and Macurda, B.D., 1977, Adaptive radiation of the comatulid crinoids:

Paleobiology, v. 3, p. 74–82.

Meyer, D.L., Ausich, W.I., and Terry, R.E., 1989, Comparative of

echinoderms in carbonate facies: (Lower Mississippian) of

Kentucky and Tennessee: Palaios, v. 4, p. 533-552.

Moore, R. C., Lane, N. G., and Strimple, H. L., 1978, Order Cladida Moore and Laudon,

1943, in Moore, R. C. and Teichert, C., eds: Treatise on Invertebrate

Paleontology, Pt. T, Echinodermata 2: Geological Society of America and

University of Kansas Press, Kansas, p. T578–T759.

O'Meara, B.C., C. Ané, M. J. Sanderson, P. C. Wainwright, and T. Hansen. 2006, Testing

for Different Rates of Continuous Trait Evolution Using Likelihood. Evolution, v.

60, p. 922-933.

264

Oyston, J.W., Hughes, M., Wagner, P.J., Gerber, S., and Wills, M.A., 2015, What limits

the morphological disparity of clades?: Interface Focus, v. 5, 20150042, 18 p.

Peters, S.E., and Ausich, W.I., 2008, A sampling-adjusted macroevolutionary history for

Ordovician-Early Silurian crinoids: Paleobiology, v. 34, p. 104‒116.

Rabosky, D.L., and Adams, D.C., 2012, Rates of morphological evolution are correlated

with species richness in salamanders: Evolution, v. 66, p. 1807-1818.

Raup, D. M., 1985, Mathematical models of cladogenesis: Paleobiology, v. 11, p. 42–52.

Revell, Liam J. "phytools: an R package for phylogenetic comparative biology (and other

things)." Methods in Ecology and Evolution 3, no. 2 (2012): 217-223.

Ronquist, F., Klopfstein, S., Vilhelmsen, L., Schulmeister, S., Murray, D.L. and

Rasnitsyn, A.P., 2012, A total-evidence approach to dating with fossils, applied to

the early radiation of the Hymenoptera: Systematic Biology, v. 61, p. 973–999.

Sallan, L.C., Kammer, T.W., Ausich, W.I., and Cook, L.A., 2011, Persistent predator-

prey dynamics revealed by mass extinction: Proceedings of the National Academy

of Science, v. 108, p. 8335–8338.

Schluter, D., 2000, The ecology of adaptive radiation. OUP Oxford.

Sevastopulo, G.D., and Lane, N.G., 1988, Ontogeny and phylogeny of disparid crinoids,

in Paul, C.R.C., and Smith, A.B., eds, Echinoderm phylogeny and evolutionary

biology: Oxford, Clarendon Press, p. 245–253.

Signor P.W., and Brett, C.E., 1984, The mid-Paleozoic precursor to the Mesozoic marine

revolution: Paleobiology, v. 10, p. 229-245.

265

Simms, M.J., and Sevastopulo, G.D., 1993, The origin of articulate crinoids:

Palaeontology, v. 36, p. 91–109.

Simões, M., Breitkreuz, L., Alvarado, M., Baca, S., Cooper, J.C., Heins, L., Herzog, K.

and Lieberman, B.S., 2016, The Evolving Theory of Evolutionary Radiations.

Trends in ecology & evolution, v. 31, p.27-34.

Simpson, G.G., 1955. Major features of evolution. Columbia University Press, 434 p.

Slater, G.J., 2013, Phylogenetic evidence for a shift in the mode of mammalian body size

evolution at the Cretaceous-Palaeogene boundary: Methods in Ecology and

Evolution, v. 4, p. 734‒744.

Slater, G.J., 2015, Not so early bursts and the dynamic nature of morphological

diversification: Proceedings of the National Academy of Sciences, v. 112, p.

3595-3596.

Slater, G.J., 2015, Iterative adaptive radiations of fossil canids show no evidence for

diversity-dependent trait evolution: Proceedings of the National Academy of

Sciences, v. 112, p. 4897‒4902.

Slater, G.J., and Harmon, L.J., 2013, Unifying fossils and phylogenies for comparative

analyses of diversification and trait evolution: Methods in Ecology and Evolution,

v. 4, p. 699‒702.

Slater, G.J., Price, S.A., Santini, F. and Alfaro, M.E., 2010, Diversity versus disparity and

the radiation of modern cetaceans. Proceedings of the Royal Society of London B:

Biological Sciences, v. 277, p. 3097-3104.

266

Swofford, D., 2002, PAUP*: Phylogenetic Analysis Using Parsimony (*and Other

Methods): Version 4. Massachusetts, Sinauer Associates.

Vermeij, G.J., 1987, Evolution and escalation: an ecological history of life. Princeton

University Press.

Vrba, E.S., 1987, Ecology in relation to speciation rates: some case histories of -

Recent clades. Evolutionary Ecology, v. 1, p.283-300.

Wagner, P.J., 1997, Patterns of morphologic diversification among the Rostroconchia:

Paleobiology, v. 23, p. 115-150.

Wagner, P.J., 2000, Exhaustion of morphologic character states among fossil taxa:

Evolution, v. 54, 365‒386.

Wagner, P.J., 2010. Paleontological perspectives on morphological evolution. in

Evolution since Darwin: the first, 150 years, pp.451-478.

Wagner, P. J., 2012, Modelling rate distributions using character compatibility:

implications for morphological evolution among fossil invertebrates: Biology Letters,

v. 8, p. 143—146.

Webster, G.D., 2013, Bibliography and index of Paleozoic crinoids, coronates, and

hemistreptocrinoids, 1758–2012: [http://crinoids.azurewebsites.net/]

Webster, G.D. and Maples, C.G., 2006, Cladid crinoid (Echinodermata) anal conditions:

a terminology problem and proposed solution: Palaeontology, v. 49, p. 187–212.

Wright, D.F., 2015, Fossils, homology, and “Phylogenetic Paleo-ontogeny”: a

reassessment of primary posterior plate homologies among fossil and living

267

crinoids with insight from developmental biology: Paleobiology, v. 41, p. 570–

591.

Wright, D.F., in press, Bayesian estimation of fossil phylogenies and the evolution of

early to middle Paleozoic crinoids (Echinodermata): Journal of Paleontology.

Wright, D.F., Ausich, W.I., Cole, S.R., Rhenberg, E.C., and Peter, M.E., in press,

Phylogenetic taxonomy and classification of the Crinoidea (Echinodermata):

Journal of Paleontology.

Zhang, C., Stadler, T., Klopfstein, S., Heath, T.A. and Ronquist, F., 2016, Total-Evidence

Dating under the Fossilized Birth–Death Process: Systematic Biology, v. 65, p.,

228‒249.

268

Figure 5.1 Macroevolutionary models of trait evolution. Commonly used quantitative models of trait evolution among lineages are depicted as traitgrams. Models differ both conceptually and mathematically, especially how trait variation is expected to accumulate over time. Each represents adifferent ‘mode’ of evolution (Hunt and Carrano, 2010).

Comparative data can be fitted to alternative models using likelihood or Bayesian frameworks, and parameter estimates can be used to infer evolutionary tempo and mode

(but see Hunt, 2012). All models shown expect a degree of phylogenetic autocorrelation among species except “white noise”, which is equivalent to Brownian motion on a star phylogeny. Equations describing the models shown are described in Hansen (1997),

Butler and King (2004), Felsenstein (2004), Harmon et al. (2010), and Paradis (2012).

269

Fig. 5.1

270

Fig. 5.2 Rates of phenotypic evolution in eucladid crinoids during the Paleozoic. The upper figure is the Maximum Clade Credibility tree from the Bayesian phylogenetic analysis. Major clades are identified at nodes. Median rates of morphologic evolution are shown along branches. Red branches indicate elevated rates and black branches indicate lower rates. Node bars are 95% highest probability densities for divergences.The lower figure are results from the maximum-likelihood analysis over a random sample of 100 time-calibrated trees from the Bayesian posterior distribution.Colored circles are mean rates and open circles represent medians. Red circles indicate intervals with >55% of sampled phylogenies were characterized by statistically high rates, whereas blule circles represent intervals where >55% of sampled phylogenies had statistically low rates (See

Materials and Methods). Error bars represent 95% confidence intervals.

271

Fig. 5.2

272

Figure 5.3 Disparity and diversity profiles for Paleozoic eucladid crinoids. (Upper)

Squares indicate the average pairwise distance among species placed in period-level bins.

Note the Ordovician point is offset because eucladids originated during the Late

Ordovician. Errors bars are 1 standard error based on bootstrap resampling. Although

based on different taxon sampling strategies, discrete characters, and homologies, the

major features and overall trajectory of this curve are strikingly similar to previous

disparity profiles generated for Paleozoic eucladids (Foote, 1995 [fig. 15], 1999 [fig.

39]). (Middle) Circles represent taxonomic diversity, calculated as the number of genera

per time bin using taxonomic first and last appearances recorded in Webster (2013).

Errors bars are 1 standard error based on 1,000 bootstrap replicates. (Lower) Median

diversity of taxa sampled for phylogenetic analysis calculated across 500 randomly

sampled time-calibrated trees from the posterior distribution. Uncertainty around the

median curve is indicated by the shaded region bounded by two-tailed 95% upper and lower quantiles.

273

Fig. 5.3 274

Figure 5.4 Time-varying diversification and sampling parameters estimated from

phylogenetic analysis and stratigraphic ranges of genera. Net diversification (A),

relative extinction (B), and sampling rates (C) from the Bayesian posterior distribution

obtained from phylogenetic analysis. Points indicate median values for all parameter

estimates in the posterior distribution. Error bars reflect 95% condifence intervals. The

circles (D) represent net taxonomic diversification calculated at the genus-level, whereas the X’s indicate per-interval rates of turnover among genera (see Materials and Methods).

275

Fig. 5.4

276

Figure 5.5 Rates of morphologic evolution among eucladid subclades. Rates of morphologic evolution among the cyathoform (left) and cladoform (right) subclades of eucladid crinoids. Rates were inferred using the maximum-likelihood method over a random sample of 100 time-calibrated trees from the posterior distribution. Colored circles are mean rates and open circles represent medians. Error bars represent 95% confidence intervals.

277

Fig. 5.5

278

Figure 5.6 Identification of adaptive zones based on principal coordinate analysis of

the morphologic character matrix. Adaptive zones are identified based on the clustering of taxa in morphospace. Zones were identified by rank correlation of PCO axes and morphologic traits (see Materials and Methods) and correspond with previously identified grades of body plan organization in eucladids (Kammer and Ausich, 1992,

1993, 1996; Webster and Maples, 2006; Wright, 2015).

279

Fig. 5.6

280

Figure 5.7 Phylomorphospace and adaptive zone occupation throughout the history of Paleozoic radiation. Results of principal coordinate analysis conducted on the entire matrix with taxa joined according to phylogenetic relationships implied by the MCC tree.

The time-series shows morphologic disparity within eucladids through time and among clades. Taxa occurring within a timeslice are indicated by filled circles, whereas taxa not occurring within a timeslice are removed. Cyathoform taxa are distributed on the right hand side of PCO1, whereas taxa within cladoform subzones are in the center and left.

See Figure 5.6 and the Materials and Methods section for delimitation and interpretation of eucladid adaptive zones.

281

Fig. 5.7

282

Figure 5.8 The relationship between taxonomic and morphologic diversification in

Paleozoic eucladids. Evolutionary radiation in Paleozoic eucladid crinoids shown in a

diversity-disparity plot. Taxonomic data were transformed to a log scale, generating a

linear expectation in taxonomic diversification. Disparity and taxonomic diversity were

range standardized using [(xi – min[x] ) / (max[x] – min[x])]. If data fall above the red line, disparification is outpacing diversification. If data instead fall below the red line, diversification is outpacing the rate at which morphologic innovation is generated. Data falling on the red line imply concordance between diversity and disparity. See Jablonski

(in press) for additional information on types of evolutionary radiations in diversity- disparity space. Error bars are not shown (see Figure 5.9), but the Silurian and Devonian points are significantly above the 1:1 line and Pennsylvanian point falls significantly below it. Ord = Ordovician, Sil = Silurian, Dev = Devonian, Miss = Mississippian (early

Carboniferous), Penn = Pennsylvanian (late Carboniferous), and Perm = Permian.

283

Fig. 5.8

284

Figure 5.9 Additional plots depicting the radiation of eucladids in diversity-disparity space. These plots are included to show uncertainty in taxonomic diversity and morphologic disparity plotted in diversity-disparity space. Error bars are 1 standard error from 1,000 bootstrap replicates. Unlike Figure 5.9, data were not range standardized.

Note that taxonomic diversity is plotted on a log scale in the upper figure, but not in the lower figure.

285

Fig. 5.9

286

Figure 5.10 Maximum Clade Credibility tree from Bayesian analysis of 81 Paleozoic

eucladid crinoid species. Node support is indicated by posterior probabilities. Basal

nodes are well supported, but many more nested nodes are not. However, cyathoform and cladoform clades (see main text and Figure 5.3) are supported with PP = 1.

287

Fig. 5.10

288

Chapter 6: Crinoids from the Bobcaygeon and Verulam Formations (Upper Ordovician)

near Brechin, Ontario: new taxa and emended descriptions

Abstract.— A survey of newly collected material from the Bobcaygeon and Verulam

Formations reveals a large number of exceptionally preserved crinoid specimens.

Focusing on taxa placed within the Cladida, this chapter contains taxonomic descriptions for eight species across five genera. The genus Konieckicrinus n. gen. n. sp. is proposed, comprised of two previously unknown species. In addition, descriptions for other species within the Brechin fauna are included, with comments on their taxonomic membership and recent changes in crinoid classification.

Introduction

The Ordovician is a key interval for understanding the evolutionary history and diversification of crinoids (Echinodermata, Crinoidea) because nearly all major clades first appear in Ordovician strata. After their initial diversification during the Early

Ordovician, crinoids subsequently underwent a taxonomic and morphologic radiation coincident with the Great Ordovician Biodiversification Event (GOBE) and reached peak

Ordovician diversity during the Sandbian. Following the GOBE, crinoid taxonomic

289 diversity dramatically declined and was decimated by the end-Ordovician mass extinction

(Peters and Ausich, 2008).

The Upper Ordovician (lower Katian) Bobcaygeon and Verulam Formations near

Brechin, Ontario are comprised of a highly diverse, well-preserved crinoid fauna. These

“Brechin Fauna” crinoids are preserved alongside a rich Late Ordovician echinoderm community including representative ophiuroids, paracrinoids, cystoids, edrioasteroids, edrioblastoids, cyclocystoids, and homalozoans. This fauna provides an exceptional window into a taxonomically diverse Late Ordovician crinoid community from the interval between the Sandbian diversity peak and the end-Ordovician extinction.

Although the sedimentology, stratigraphy and paleoecology of these Upper Ordovician strata have been studied in detail, the Brechin Fauna crinoids have not had a comprehensive taxonomic evaluation since Frank Springer’s (1911) classic monograph.

A survey of newly collected material from the Bobcaygeon and Verulam

Formations reveals a large number of exceptionally preserved crinoid specimens with arms, stems, and attachment structures intact. Most importantly, this chapter describes two new, previously undescribed species and places them in a new genus. Many previously described species from this locality were based on incomplete material; thus, this new collection provides an opportunity to more fully describe morphologic details of known Ordovician crinoid taxa and to conduct a taxonomic re-evaluation.

290

Geographic and stratigraphic occurrences and geologic setting

Fossil material was collected from the Bobcaygeon and Verulam Formations (Upper

Ordovician) exposed in the Lake Simcoe region near Brechin, Ontario (Fig. 6.1).

Laurentia was located ~20o south of the equator during the Late Ordovician (Scotese and

McKerrow, 1991) and sediments comprising both formations were deposited in warm water subtropical environments (Brett and Taylor, 1999, but see Brookfield, 1988 for an alternative interpretation). The Bobcaygeon Formation ranges in age from the Sandbian to Katian, whereas the Verulam is early Katian (Fig. 6.2). These strata consist of bioclastic wackestones, grainstones, and packstones interbedded with calcareous shales.

The lithology of the Bobcaygeon Formation ranges from lime mudstone to grainstone dominant lithologies and are frequently interbedded with shale, suggesting deposition in lagoonal and shoaling settings to shallow shelf environments (Brookfield and Brett, 1988;

Brett and Taylor, 1999). In contrast, in the Verulam are somewhat thinner bedded and have greater proportion of interbedded shale, indicated deposition in a distal- shelf environment (Brett and Taylor, 1999). Both formations contain exceptionally preserved echinoderm faunas, including a diverse assemblage of crinoids. Although all major clades of Ordovician crinoids are represented in the Brechin fauna, cladids are the most abundant in the collections, especially cladids belonging to the Eucladida (Wright, in press) (Fig. 6.3).

291

Systematic paleontology

The classification system used here is from Wright et al. (in press). Morphologic terms follow Ubaghs (1978), Webster and Maples (2008), and Wright (2015). All measurements are in mm. References listed in the synonomies below can be found in

Webster (2013). Specimen numbers refer to the reference numbers given by Joe Koniecki during collection and curation. Following publication of in prep manuscripts describing

Brechin crinoids (Wright, in prep, Cole, in prep, Ausich et al., in prep), all fossils will be donated to the University of Michigan Museum of Paleontology and will be given museum new numbers when reposited. Although new species and genera are provided with taxonomic names, this dissertation chapter represents a nascent version of research currently in preparation for publication. All names, diagnoses, descriptions, and identifications are preliminary and should not be cited or used formally until results are published in a peer-reviewed journal.

Class Crinoidea Miller, 1821

Subclass Pentacrinoidea Jaekel, 1918

Infraclass Inadunata Wachsmuth and Springer, 1885

Inadunata incertae sedis: Family Metabolocrindae Jaekel, 1902

Remarks.—The Metabolocrindae are traditionally placed within the Cladida as defined by

Moore and Laudon (1943). However, the phylogenetic position of the Metabolocrinidae

292 within the broader Pentacrinoidea clade is currently in question. Phylogenetic analysis of early to middle Paleozoic non-camerate taxa by Wright (in press) recovered

Metabolocrinus as more closely related to the disparid clade than cladids, suggesting the metabolocrinids are in need of taxonomic revision. The Metabolocrindae is therefore considered herein as Inadunata incertae sedis.

Genus Konieckicrinus new genus

Type species.— Konieckicrinus brechinensis new species.

Diagnosis.—Low bowl-shaped calyx, dicyclic, stellate plate sculpturing, second primibrachial axillary, pinnulate, arms branch isotomously on the second primibrachial, asymmetric endotomous to distal isotomous branching, brachials moderately cuneate, radianial in cup, located below and to the left of the C-ray radial, anal X above radianal, in contact the C-ray radial and supporting a median anal sac comprised of numerous tiny plates, column round, lumen round.

Etymology.—Konieckicrinus is named after Joseph Koniecki, who collected specimens of

Konieckicrinus from quarries near Brechin.

Remarks.— Konieckicrinus is assigned to the Metabolocrinidae based on the presence of pinnules arms with endotomous branching. The presence of cuneate arms is

293

Konieckicrinus is most similar to the Ordovician genus Eopinnacrinus Brower and

Veinus, 1982 which was placed in the Metabolocrinidae by Ausich (1998).

Konieckicrinus differs from Eopinnacrinus in having aymmetric endotomous to isotomous branching above secundibrachials, stellate plate ornamentation, and an axillary on the second primibrachial.

Konieckicrinus brechinensis n. gen n. sp.

Figures 6.4-6.5

Holotype.—5097.2JK, paratype 5097.3JK.

Diagnosis.—Species of Konieckicrinus with asymmetric endotomous branching above secundibrachials.

Occurrence.—Bobcaygeon Formation, (Sandbian-Katian, Upper Ordovician), Ontario,

Canada.

Description.—Calyx medium size, low bowl shape, base convex, stellate ornamentation on plates; infrabasals 5, visible from the side, basals 5, higher than wide, radials 5, plates larger than other calyx plates, subequal to slightly wider than high.

Brachials moderately cuneate; pinnulate arms; first primibrachial wider than higher, axillary on the second primibrachial, triangular, branching with asymmetric

294

endotomy on secundibrachial 4-8, distal branches endotomous, with subequal rami

arising from each axillary, three to six distal arms in each ray.

Radianal beneath and to the left of the C ray radial plate, approximate size of basals; anal X above radianal, in contact with C ray radial, supporting a median column of anal sac plates.

Column circular, lumen round.

Etymology.— Named for the town of Brechin, Ontario, which is located near where the specimens were collected.

Konieckicrinus josephi n. gen n. sp.

Figures 6.6-6.8

Holotype.—5097.7JK, paratypes 5097.1JK, 5097.6JK.

Diagnosis.—Species of Konieckicrinus with distal isotomous branching above

secundibrachials.

Occurrence.—Bobcaygeon Formation, (Sandbian-Katian, Upper Ordovician), Ontario,

Canada.

295

Description.—Calyx small to medium size, low to moderate bowl shape, base convex, stellate ornamentation on plates; infrabasals 5, visible from the side, wider than high; basals 5, higher than wide, radials 5, larger than other calyx plates, wider than high.

Brachials moderately cuneate; pinnulate arms; first branching on the second

primibrachial, branching isotomously above secundibrachials, typically at or above

secundibrachial 8, distal branches endotomous to isotomous, with subequal rami arising

from each axillary, three to six distal arms in each ray.

Radianal beneath and to the left of the C ray radial plate, approximate size of

basals, anal X above radianal, in contact with C ray radial and CD basal, supporting a

median column of anal sac plates, anal sac plates with stellate ornamentation.

Column circular, lumen round, attachment structure unknown, stem length is at

least more twice that of crown.

Etymology.— Named for the town of Brechin, Ontario, which is located near where the specimens were collected.

.

Parvclass Cladida Moore and Laudon, 1943

Superorder Porocrinoidea, Wright, in press

Order Porocrinida Miller and Gurley, 1894

Genus Illemocrinus Eckert, 1987

Type species.—Illemocrinus amphiatus Eckert, 1987; by monotypy.

296

Other species.—Monospecific.

Illemocrinus amphiatus Eckert, 1987

Fig. 6.12

1987 Illemocrinus amphiatus Eckert, 1987, p. 862.

1993 Illemocrinus amphiatus; Webster, p. 72.

1988 Illemocrinus amphiatus; Ausich, p. 34.

Ocurrence.—Late Ordovician (Katian), Bobcaygeon Formation; Ontario, Canada.

Description.—Calyx small, medium bowl to globe shape; slightly higher than wide, infrabasals small, pentagonal; basal large, much higher than wide, with 5 to 6 ridges radiating from the center; radials 5, curved inward, slightly wider than high.

Radianal positioned below and to the left of the C ray radial, quadrangular; anal X above and to the right of the radianal, positioned in between the C and D ray radials, slightly larger than the radianal, supporting three additional anal plates above; anal sac massive, high, approximately twice the length of the calyx, comprised of stacked plates, becoming more irregular near the summit; anal sac plates with stellate ornamentation.

297

Brachials uniserial, wider than high; radial facets angustary, sloping down and inward; arms branch isotomously on the second primibrachial, higher branching isotomous up to at least four bifurcations.

Column pentagonal.

Materials.—5140.1JK.

Family Porocrinidae Miller and Gurley, 1894

Genus Carabocrinus

Type species.—Carabocrinus radiatus Billings, 1857.

Other species.—Carabocrinus boltoni Ausich and Copper, 2010; C. crateriformis (Hall,

1866); C. conoideus Sardeson, 1925; C. dicyclicus (Sardeson, 1899); C. esthonus Jaekel,

1918; C. geometricus Hudson, 1905; C. granulosus Evans, 1926; C. huronensis Foerste,

1924; C. magnificus Sardeson, 1939; C. micropunctatus Brower and Veinus, 1974; C. oogyi Kolata, 1975; C. ovalis Miller and Gurley, 1894; C. stellifer Brower and Veinus,

1974; C. treadwelli Sinclair, 1945; C. vancortlandti.

Carabocrinus radiatus Billings, 1857

Fig. 6.9

298

1857 Carabocrinus radiatus Billings, 1857, p. 276.

1859 Carabocrinus radiatus; Billings p. 31, pl. 2, figs. 3a-3e.

1868 Carabocrinus radiatus; Shumard, p. 357.

1868 Carabocrinus radiatus; Bigsby, p. 18.

1889 Carabocrinus radiatus S.A. Miller, p. 182.

1910 Carabocrinus radiatus; Grabau and Shimer, p. 504, fig. 1817.

1912 Carabocrinus cf. radiatus; Ruedemann, p. 86, pl., 3, fig. 6.

1915 Carabocrinus radiatus; Basser, p. 183.

1915 Carabocrinus cf. radiatus; Basser, p. 183.

1938 Carabocrinus radiatus; Bassler, p. 60.

1943 Carabocrinus radiatus; Bassler and Moodey, p. 356.

1943 Carabocrinus radiatus; Moore and Laudon, p. 153, pl., 53, fig. 5.

1946, Carabocrinus radiatus; Wilson, p. 38, pl., 5, fig. 5.

1973 Carabocrinus radiatus; Webster, p. 78.

1996 Carabocrinus radiatus; Brower, 1996, p. 615, figs. 1-2.

Ocurrence.—Late Ordovician (Katian), New York, Iowa; Bobcaygeon Formation;

Ontario, Canada.

Description.—Calyx low, rounded base; relatively wide; bearing multiple stellate ridges;

exothecal respiratory canals; infrabasals 5, visible in side view, slightly wider than high,

299

pentagonal; basals large, hexagonal to septagonal, slightly higher than wide; radials 5,

wider than high.

Radianal split into infer- and superradianals, between CD and BC interray basals, higher than wide; anal X located above and to the left of the superradianal.

Uniserial brachials; narrow, angustary radial facets; isotomous branching on the second primibrachial; higher branching isotomous to somewhat heterotomous.

Proximal column circular.

Materials.—5080.2bJK and 5080.2cJK.

Carabocrinus vancortlandti Billings, 1859

Figs. 6.10

1859 Carabocrinus vancortlandti Billings, p. 52, pl, 2, fig. 4.

1868 Carabocrinus vancortlandti; Shumard, p. 357.

1868 Carabocrinus vancortlandti; Bigsby, p. 18.

1889 Carabocrinus vancortlandti; Miller, S.A. p. 231.

1900 Carabocrinus vancortlandti; Weller, p. 28, fig. 9.

1915, Carabocrinus vancortlandti; Bassler, p. 183.

1943 Carabocrinus vancortlandti; Bassler and Moodey, p. 356.

1943 Carabocrinus vancortlandti; Moore and Laudon, p. 153, pl., 53, fig. 6.

1973 Carabocrinus vancortlandti Webster, p. 78.

300

1996 Carabocrinus vancortlandti Brower, p. 617, figs. 4-5.

Ocurrence.—Middle to Late Ordovician (-Katian), Iowa, Kentucky, United

States; Ontario, Canada.

Description.— Calyx high, rounded base and side; oval to egg shaped; bearing multiple,

well-defined stellate riches; exothecal respiratory canals; infrabasals 5, visible in side view, slightly wider than high, pentagonal; basals large, hexagonal to septagonal, slightly higher than wide; radials 5, wider than high.

Radianal split into infer- and superradianals, between CD and BC interray basals, both plates higher than wide; anal X located above and to the left of the superradianal.

Uniserial brachials; narrow facets, angustary; isotomous branching on the second primibrachial; higher branching isotomous, higher axillaries on different secundibrachials within each ray.

Proximal column circular, at least three time the length of the crown.

Materials.—5143JK.

Magnorder Eucladida Wright, in press

Superorder Cyathoformes Wright et al., in press

Cyathoformes incertae sedis: “Dendrocrinida” Bather, 1899

Family Dendrocrinidae Wachsmuth and Springer, 1886

301

Genus Grenprisia Moore, 1962

Type species.—Grenprisia billingsi Moore, 1962.

Other species.—G. springeri Moore, 1962.

Grenprisia billingsi (Springer, 1911)

1911 Ottawacrinus billingsi Springer; p. 40, pl. 4, figs. 1-4.

1915 Ottawacrinus billingsi; Bassler, p. 926.

1943 Ottawacrinus billingsi; Bassler and Moodey, p. 577.

1944 Ottawacrinus billingsi; Moore and Laudon, p. 158, pl. 53, fig. 23.

1962 Grenprisia billingsi; Moore, p. 38.

1973 Grenprisia billingsi; Webster, p. 143.

1978 Grenprisia billingsi; Brower and Veinus, p. 443, pl, 13, figs. 1-4.

1978 Grenprisia billingsi; More and Lane, p. T612, figs. 397, nos. 2a-2g.

1986,Grenprisia billingsi; Webster, p. 162.

Ocurrence.—Late Ordovician (Katian), Hull Limestone, Bobcaygeon Formation;

Ontario, Canada.

Materials.—5110.2JK.

302

Grenprisia springeri (Springer, 1911)

Fig. 6.11

1911 Ottawacrinus typus Springer 1911 (non Billings, 1887), p. 37, pl. 4, figs. 5-7.

1962 Grenprisia springeri; Moore, p. 38.

1973 Grenprisia springeri; Webster, p. 143.

1978 Grenprisia springeri; Moore and Lane, p. T612, fig. 397, no. 2h.

1986 Grenprisia springeri; Webster, p. 162.

Ocurrence.— Late Ordovician (Katian), Hull Limestone, Bobcaygeon Formation;

Ontario, Canada.

Description.—Calyx medium to small in size, high bowl shape, flat base; infrabasals large, visible from the side, subequal; basals large, equal in size or larger than radials; radial plates wider than high.

Radianal large, directly above posterior basal plate, supporting an anal X and right tube plate above leading to an anal sac; cylindrical, large anal sac, comprised of small plates with stellate ornamentation.

Arms uniserial; heterotomous, branching isotomously on primibrachial 3, multiple bifurcations above secundibrachials.

Pentemeric stem, transversely pentagonal.

303

Materials.—5110.11JK.

Remarks.—Grenprisia billingsi is distinguished from G. springeri by its much larger size

and heterotomous arms with secondary ramules that only branch on one side, whereas G.

springeri is much smaller and typically has isotomous branching arms with one or two

bifurcations (Springer, 1911).

Superorder Flexibilia Zittel, 1895

Remarks.—The Flexibilia have traditionally been considered a subclass of crinoids (e.g.,

Moore and Teichert, 1978). However, it has been hypothesized since Springer (1911,

1920) that flexible crinoids descended from ancestors that were placed within the

Cladida, rendering the cladids paraphyletic. Simms and Sevastopulo (1993) recommended the Flexibilia be placed within the Cladida to eschew paraphyly in the

classification of crinoid higher taa. Wright et al.’s (in press) revised classification of the

Crinoidea follows Simms and Sevastopulo (1993) and placed the Flexibilia as a

superorder within the Cladida.

304

Family Cupulocrinidae Moore and Laudon, 1943

Remarks.—The Cupulocrinidae was originally erected by Moore and Laudon (1943) as a monogeneric family of Ordovician crinoids from North America. However,

Morenacrinus Ausich, Gil Cid, and Alonso, 2002 from the Darriwilian of Spain and

Stewbrecrinus Jell, 1999 from the Devonian of Australia have since been placed within the family. Building on the work of Springer, 1920, a recent phylogenetic analysis of

Wright (in press) found evidence to support Cupulocrinus heterocostalis as potentially ancestral to flexible crinoids. Wright et al. (in press) presented a phylogeny-based revision to crinoid classification and suggested Cupulocrinus be included within the definition of the flexible clade. Thus, the monophyly of the family is in question.

However, a comprehensive revision is beyond the scope of this chapter.

Genus Cupulocrinus d’Orbigny, 1849

Type species.—Cupulocrinus heterocostalis (Hall, 1847).

Other species.—Cupulocrinus angustatus (Meek and Worthen, 1870); C. austrogracilis

Jell, 1999; C. canaliculatus Brower and Veinus, 1978; C. crossmani Brower, 1992; C. heterobrachialis Rambsbottom, 1961; C. humilis (Billings, 1857); C. jewetti (Billings,

1859); C. kentuckiensis Springer, 1911; C. latibrachiatus (Billings, 1857); C. leversoni

Kolata, 1986; C. minimus Springer, 1911; C. minimus Springer 1911; C. molanderi

305

Kolata, 1975; C. plattesvillensis Kolata, 1975; C. sepulchrum Ramsbottom, 1961; C. tuberculatus d’Orbigny, 1847.

Cupulocrinus humilis (Billings, 1857)

Fig. 6.13

1857 Dendrocrinus humilis Billings, 1857, p. 270.

1859 Dendrocrinus humilis; Billings, p. 39.pl. 3, fig. 4.

1868 Dendrocrinus humilis; Shumard, p. 365.

1868 Dendrocrinus humilis Bigsby, p. 19.

1889 Dendrocrinus humilis Miller, 1889, p. 238.

1911 Cupulocrinus humilis; Springer, 1911, p. 28, pl. 1, figs. 8-9.

1915 Cupulocrinus humilis; Bassler, p. 315.

1943 Cupulocrinus humilis; Bassler and Moodey, p. 387.

1943 Cupulocrinus humilis; Moore and Laudon, p. 155, pl., 53, figs. 7, 25.

1953 Cupulocrinus humilis; Ubaghs, p. 751, figs. 14, no. b.

1973 Cupulocrinus humilis; Webster, p. 92.

1975 Cupulocrinus humilis; Kolata, p. 38, pl. 6, fig. 7.

1978 Cupulocrinus humilis; Ubaghs, p. T125, fig. 94, no. 2.

1981 Cupulocrinus humilis; Lewis, p. 227.

1986 Cupulocrinus humilis; Webster, p. 111.

1988 Cupulocrinus humilis; Webster, p. 63.

2005 Cupulocrinus humilis Sloan, p. 153, fig. 4, no. 6.

306

Ocurrence.—Late Ordovician (Katian), Kentucky, New York, Minnesota, United Staes;

Ontario, Canada.

Description.—Calyx conical, flat based, much higher than wide, smooth; infrabasals 5, pentagonal, higher than wide; basals large, hexagonal, comprising >50% of calyx height, radials subequal in height:width, closely spaced, abutting one another.

Radianal large, approximate in size to radials, hexagonal, located directly beneath the C-ray radial plate, in contact with CD and BC basal plates; anal X similar in size to

other calyx plates, higher than wide, above and to the left of radianal, situated between

the C and D radials, supporting sac plates above; anal sac tube-like, erect, consisting of stacked, rounded plates, distally tapering.

Brachials uniseral, much wider than high, narrowing distally; radial facets wide, extending across the full width of the radial facets; weakly developed “petelloid” process, number of primibrachials variable, branching isotomously on primibrachials 4-12; higher bifurcations isotomous to heterotomous;

Column circular; proximal portion wider than the rest of the stem; large nudinodals, regularly spaced.

Materials.—5115JK and 5123.2JK.

307

Cupulocrinus jewetti (Billings, 1859)

Fig. 6.14

1859 Dendrocrinus jewetti Billings, E., p. 43, fig. 13.

1868 Dendrocrinus jewetti; Shumard, p. 356.

1868 Dendrocrinus jewetti; Bigsby, p. 19.

1883 Dendrocrinus jewetti; Billings W.R., p. 51.

1889 Dendrocrinus jewetti; Miller, p. 238, fig. 283.

1911Cupulocrinus jewetti; Springer, p. 28, pl. 1, figs. 10-12, pl. 3, figs. 5-7c.

1915 Cupulocrinus jewetti; Bassler, p. 315.

1915 Dendrocrinus jewetti, Bassler, p. 315.

1920 Cupulocrinus jewetti; Springer, p. 88, pl. 75, figs. 2-4.

1943 Cupulocrinus jewetti; Bassler and Moodey, p. 387.

1943 Cupulocrinus jewetti; Moore and Laudon, p. 155, pl 53, fig. 12.

1970 Cupulocrinus jewetti; Norford et al., p. 606, pl 4, fig. 12.

1970 Cupulocrinus jewetti; Bolton, p. 63, pl. 13, fig. 12.

1972 Cupulocrinus jewetti; Bolton and Copeland, p. 54, pl. A, fig. 22.

1977 Cupulocrinus jewetti; Webster, p. 62.

1978 Cupulocrinus jewetti; Frest and Strimple, p. 688, fig. 7, no. A.

1978 Cupulocrinus jewetti; Brower and Veinus, p. 430, pl. 14, figs. 1-4, pl. 15, figs. 4, 7.

1978 Cupulocrinus jewetti; Ubaghs, p. T189, fig. 160, no. 8.

1978 Cupulocrinus jewetti; Moore, p. T627, fig. 409, no. 2a-2b.

308

1981 Cupulocrinus jewetti; Lewis, p. 227.

1986 Cupulocrinus jewetti; Kolata, p. 714, fig. 2, no. 1.

1986 Cupulocrinus jewetti; Webster, p. 111.

1988 Cupulocrinus jewetti; Webster, p. 63.

1993 Cupulocrinus jewetti; Webster, p. 45.

1999 Cupulocrinus jewetti; Brett and Taylor, p. 69, fig. 80, no. e, Fig. 83.

Ocurrence.—Late Ordovician (Katian), Kentucky, Wisconsin, Minnesota, Illinois, United

States; Ontario, Canada.

Description.— Calyx conical to medium bowl shape, flat based, widest at the summit, smooth; infrabasals 5, pentagonal, higher than wide; basals large, hexagonal, comprising

>50% of calyx height, radials subequal in height:width, closely spaced, abutting one

another.

Radianal large, approximate in size to radials, hexagonal, located directly beneath

the C-ray radial plate, in contact with CD and BC basal plates; Anal X similar

approximate size of radials, subequal, above and to the left of radianal, situated between

the C and D radials, supporting sac plates above; anal sac tube-like, erect, consisting of

stacked, rounded plates, distally tapering, with a median .

Brachials uniseral, much wider than high, narrowing distally; radial facets wide, extending across the full width of the radial facets; “petelloid” process, number of

309 primibrachials roughly constant, three in posterior rays, branching isotomously on primibrachials 4-12; higher bifurcations isotomous to heterotomous.

Column circular; proximal portion does not widen near calyx.

Materials.—5028.1.

310

References

Ausich, W.I., 1998, Phylogeny of Arenig to Caradoc Crinoids (Phylum Echinodermata)

and suprageneric classification of the Crinoidea: The University of Kansas

Paleontological Contributions Papers, New Series, no. 9, 36 p.

Bather, F.A., 1899, A phylogenetic classification of the Pelmatozoa: British Association

for the Advancement of Science (1898), p. 916–923.

Billings, E., 1857, New species of fossils from Silurian rocks of Canada: Canada Geological

Survey, Report of Progress 1853-1856, Report for the year 1856, p. 247–345.

Brett, C.E., and Taylor, W.L., 1999, Middle Ordovician of the Lake Simcoe area of

Ontario, Canada, in Hess, H., Ausich, W.I., Brett, C.E., and Simms, M.H., Fossil

Crinoids. Cambridge University Press, p. 63-74.

Brookfield, M. E. 1988, A mid-Ordovician temperate carbonate shelf—the Black River

and Trenton Limestone Groups of southern Ontario, Canada. Sedimentary

Geology, v., 60, p. 137-153.

Brookfield, M. E. and C. E. Brett. 1988, Paleoenvironments of the mid-Ordovician

(Upper Caradocian) Trenton Limestones of southern Ontario, Canada: storm

sedimentation on a shoal-basin shelf model. Sedimentary Geology, v. 57, p. 75-

105.

Eckert, J.D., 1987, Illemocrinus amphiatus, a new cladid inadunate crinoid from the

Middle Ordovician of Ontario: Canadian Journal of Earth Sciences, v. 24, p. 860-

865. 311

Jaekel, O., 1918, Phylogenie und System der Pelmatozoen: Paläeontologische Zeitschrift,

v. 3, p. 1–128.

Orbigny, A.D. d', 1849, Cours élémentaire de paléontologie et géologie Stratigraphiques:

v. 1, 299 p. Paris, Victor Masson.

McKerrow, W.S., Dewey, J.F. and Scotese, C.R., 1991, The Ordovician and Silurian

development of the Iapetus ocea: Special Papers in Palaeontology, v. 44, p.165-

178.

Miller, J.S., 1821, A Natural History of the Crinoidea, or Lily-Shaped Animals; with

Observations on the Genera, Asteria, Euryale, Comatula and Marsupites: Bristol,

C. Frost, 150 p.

Miller, S.A. and Gurley, W.F.E., 1894, New genera and species of Echinodermata:

Illinois State Museum of Natural History, v. 5, 53 p.

Moore, R.C., 1962, Ray structures of some inadunate crinoids. University of Kansas

Paleontological Contributions, v. 5, 1–47.

Moore, R.C., and Laudon, L.R., 1943, Evolution and classification of Paleozoic crinoids:

Geological Society of America Special Paper, no. 46, 151 p.

Moore, R.C., and Teichert, C., eds., 1978, Treatise on Invertebrate Paleontology, Part T,

Echinodermata 2: Geological Society of America and University of Kansas Press,

Lawrence, Kansas, 1027 p.

Peters, S.E., and Ausich, W.I., 2008, A sampling-adjusted macroevolutionary history for

Ordovician-Early Silurian crinoids: Paleobiology, v. 34, p. 104‒116.

312

Springer, F., 1911, On a Trenton echinoderm fauna: Canada Department Mines Memoir

no. 15-P, 70 p.

Springer, F., 1920, The Crinoidea Flexibilia: Smithsonian Institution Publication no.

2501, 486 p.

Wachsmuth, C., and Springer, F., 1880–1886, Revision of the Palaeocrinoidea.

Proceedings of the Academy of Natural Sciences of Philadelphia Pt. I. The

families Ichthyocrinidae and Cyathocrinidae (1880), p. 226‒378 (separate repaged

p. 1–153). Pt. II. Family Sphaeroidocrinidae, with the sub-families Platycrinidae,

Rhodocrinidae, and Actinocrinidae (1881), p. 177‒411 (separate repaged, p. 1‒

237). Pt. III, Sec. 1. Discussion of the classification and relations of the brachiate

crinoids, and conclusion of the generic descriptions (1885), p. 225‒364 (separate

repaged, 1‒138). Pt. III, Sec. 2. Discussion of the classification and relations of

the brachiate crinoids, and conclusion of the generic descriptions (1886), p. 64‒

226 (separate repaged to continue with section 1, 139‒302).

Wright, D.F., 2015a, Fossils, homology, and “Phylogenetic Paleo-ontogeny”: a

reassessment of primary posterior plate homologies among fossil and living

crinoids with insight from developmental biology: Paleobiology, v. 41, p. 570–

591.

Wright, D.F., this issue, Bayesian estimation of fossil phylogenies and the evolution of

early to middle Paleozoic crinoids (Echinodermata): Journal of Paleontology.

313

Wright, D.F., Ausich, W.I., Cole, S.R., Peter, M.E., Rhenburg, E.C., in press,

Phylogenetic taxonomy and classification of the Crinoidea: Journal of

Paleontology.

314

Figure 6.1 Locality map of the Lake Simcoe region in Ontario.

315

Fig. 6.1

316

Figure 6.2 Stratigraphy of the Simcoe Group highlighting the position of the

Bobcaygeon and Verulam Formations. (Modified from Sproat et al., 2015).

317

Fig. 6.2

318

Figure 6.3 Abundance distribution of specimens represented in the Brechin collections. Number of specimens for in the collections for higher taxa. Note that cladids are far more abundant than camerates or disparids and most cladids belong to the

Eucladida rather than the Flexibilia or Hybocrinida (see Wright, in press).

319

Fig. 6.3

320

Figure 6.4 Konieckicrinus brechinensis n. gen. n. sp., specimen 5097.2. Bobcaygeon

Fm., Carden Quarry.

321

Fig. 6.4

322

Figure 6.5 Konieckicrinus brechinensis n. gen. n. sp., specimen 5097.3. Bobcaygeon

Fm., LaFarge Quarry.

323

Fig. 6.5

324

Figure 6.6 Konieckicrinus josephi n. gen. n. sp., specimen 5097.7JK. Bobcaygeon

Fm., Carden Quarry.

325

Fig. 6.6

326

Figure 6.7 Konieckicrinus josephi n. gen. n. sp., specimen 5097.1JK. Bobcageon Fm.,

LaFarge Quarry.

327

Fig. 6.7

328

Figure 6.8 Konieckicrinus josephi n. gen. n. sp., specimen 5097.6. Bobcaygeon Fm.,

Carden Quarry.

329

Fig. 6.8

330

Figure 6.9 Carabocrinus radiatus Billings, 1857. Specimens 5080.2b (left) and 5080.2c

(right), Bobcaygeon Fm., Carden Quarry.

331

Fig. 6.9

332

Figure 6.10 Carabocrinus vancortlandti Billings, 1859. Specimen 5413JK, preserved on a slab with the camerates Cleiocrinus and Periglyptocrinus; Bobcaygeon Fm., LaFarge

Quarry.

333

Fig. 6.10

334

Figure 6.11 Grenprisia springeri Moore 1962. Specimen 5110.11JK, Bobcaygeon Fm.

335

Fig. 6.11

336

Figure 6.12 Illemocrinus amphiatus Eckert, 1987. Specimens 5140.1JK (left) and

5140JK (right), Bobcaygeon Fm.

337

Fig. 6.12

338

Figure 6.13 Cupulocrinus humilis (Billings, 1857). Specimen 5115 (left), Bobcaygeon

Fm.; 5123.2 (right), Verulam Fm.

339

Fig. 6.13

340

Figure 6.14 Cupulocrinus jewetti (Billings, 1859). Specimen 5028.1, Bobcaygeon Fm.,

Carden Quarry. Note regenerating arms.

341

Fig. 6.14

342

References

Abzhanov, A. 2013, von Baer’s law for the ages: lost and found principles of

developmental evolution: Trends in Genetics, v. 29, p. 712-722.

Alaniz, R.J., 2014, Dredging evolutionary theory: the emergence of the deep sea as a

transatlantic site for evolution 1853–1876 [Ph.D. dissertation]: University of

California, San Diego, 360 p.

Alroy, J., 2010, Geographical, environmental and intrinsic biotic controls on Phanerozoic

marine diversification. Palaeontology, v. 53, p. 1211-1235.

Alroy, J., 2016, On a conservative Bayesian method of inferring extinction: Paleobiology,

v. 42, p. 670-679.

Amemiya, S. A. Omori, T. Tsurugaya, T. Hibino, M. Yamaguchi, R. Kuraishi, M.

Kiyomoto, and T. Minokawa. 2014, Early stalked stages in ontogeny of the living

isocrinid sea lily Metacrinus rotundus: Acta Zoologica doi: 10.1111/azo.12109.

Angelin, N. P., 1878, Iconographia Crinoideorum. in stratis Sueciae Siluricis fossilium:

Samson and Wallin, Holmiae, 62 p.

Arnone, M. I. and D. H. Davidson. 1997, The hardwiring of development: organization

and function of genomic regulatory systems: Development, v. 124, p. 1851-1864.

Ausich, W.I., 1980, A model for niche differentiation in Lower Mississippian crinoid

communities: Journal of Paleontology, v. 54, p. 273-288.

343

Ausich, W.I., 1986, Early Silurian rhodocrinitacean crinoids (Brassfield Formation,

Ohio): Journal of Paleontology, v. 60, p. 84–106.

Ausich, W.I., 1996, Crinoid plate circlet homologies: Journal of Paleontology, v. 70, p.

955–964.

Ausich, W.I., 1998, Early phylogeny and subclass division of the Crinoidea (phylum

Echinodermata): Journal of Paleontology, v. 72, p. 499–510.

Ausich, W.I., 1998, Phylogeny of Arenig to Caradoc Crinoids (Phylum Echinodermata)

and suprageneric classification of the Crinoidea: The University of Kansas

Paleontological Contributions Papers, New Series, no. 9, 36 p.

Ausich, W.I., and Babcock, L.E., 1998, The phylogenetic position of Echmatocrinus

brachiatus, a probable octocoral from the Burgess Shale: Palaeontology, v. 41, p.

193–202.

Ausich, W.I., and Bottjer, D.J., 1982, Tiering in suspension-feeding communities on soft

substrata throughout the Phanerozoic: Science, v. 216, p. 173–174.

Ausich, W.I., and Copper, P., 2010, The Crinoidea of Anticosti Island, Québec: (Late

Ordovician to Early Silurian): Palaeontographica Canadiana, vol. 29, 157 p.

Ausich, W.I., and Donovan, S.K., 2015, Phylogeny of disparid crinoids: Geological

Society of America Abstracts with Programs, v. 47, p. 855.

Ausich, W.I., Kammer, T.W., and Baumiller, T.K., 1994, Demise of the Middle

Paleozoic crinoid fauna: A single extinction event or rapid faunal turnover?:

Paleobiology, v. 20, p. 345‒361.

344

Ausich, W.I., and Kammer, T.W., 2001, The study of crinoids during the 20th century

and the challenges of the 21st century: Journal of Paleontology, v. 75, p. 1161–

1173.

Ausich, W.I., and Kammer, T.W., 2008, Generic concepts in the Amphoracrinidae

Bather, 1899 (Class Crinoidea) and evaluation of generic assignments of North

American species: Journal of Paleontology, v. 82, p. 1139–1149.

Ausich, W.I., and Kammer, T.W., 2016, Exaptation of pelmatozoan oral surfaces:

constructional pathways in tegmen evolution: Journal of Paleontology,

doi:10.1017/jpa.2016.73.

Ausich, W.I., Kammer, T.W., and Baumiller, T.K., 1994, Demise of the Middle

Paleozoic crinoid fauna: a single extinction event or rapid faunal turnover?:

Paleobiology, v. 20, p. 345–361.

Ausich, W.I., Kammer, T.W., Rhenberg, E.C., and Wright, D.F., 2015, Early phylogeny

of crinoids within the pelmatozoan clade: Palaeontology, v. 58, p. 937–952.

Baldwin, B.G., and Sanderson, M.J., 1998, Age and rate of diversification of the

Hawaiian silversword alliance (Compostiae): Proceedings of the National

Academy of Sciences, v. 95, p. 9402-9406.

Bapst, D.W., 2012, paleotree: an R package for paleontological and phylogenetic

analyses of evolution. Methods in Ecology and Evolution: v. 3, p. 803-807.

Bapst, D.W. 2012, When can clades be potentially resolved with morphology?: PLoS

One e62312. doi:10.1371/journal.pone.0062312

345

Bapst, D.W., 2013, A stochastic rate‐calibrated method for time‐scaling phylogenies of

fossil taxa: Methods in Ecology and Evolution, v. 4, p.724–733.

Bapst, D.W., 2014, Assessing the effect of time-scaling on phylogeny-based analyses in

the fossil record: Paleobiology, v. 40, p. 331-351.

Bapst, D.W., and Hopkins, M.J., in press, Comparing cal3 and other a posteriori time-

scaling approaches in a case study with the Pterocephaliid trilobites:

Paleobiology.

Bapst, D.W., Wright, A.M., Matzke, N.J., and Lloyd, G.T., 2016, Topology, divergence

dates, and macroevolutionary inferences vary between different tip-dating

approaches applied to fossil theropods (Dinosauria): Biology Letters, 12,

20160237, doi: 10.1098/rsbl.2016.0237.

Bassler, R. S., 1943, New Ordovician cystidean echinoderm from Oklahoma: American

Journal of Science, v. 241, p. 694–703.

Bassler, R. S., 1943, New Ordovician cystidean echinoderm from Oklahoma: American

Journal of Science, v. 241, p. 694–703.

Bateson, W. 1894. Materials for the study of variation. MacMillan, London.

Bather, F.A., 1890, British fossil crinoids. I. Historical introduction: Annals and Magazine of

Natural History, ser. 6, v. 5, p. 306‒310.

Bather, F. A. 1890, British fossil crinoids: The classification of the Inadunata Fistulata.

Annals and Magazine of Natural History, v. 5, p. 73-788.

Bather, F.A., 1891 Some alleged cases of misrepresentation: Annals and Magazine of

Natural History, v. 6, p. 480-489. 346

Bather, F.A., 1899, A phylogenetic classification of the Pelmatozoa: British Association

for the Advancement of Science (1898), p. 916–923.

Bather, F.A., 1900, Part III The Echinoderma. The Pelmatozoa, in Lankester, E.R., ed., A

Treatise on Zoology: Adam and Charles Black, London, p. 94–204.

Bather, F.A., 1918, The homologies of the anal plate in Atedon: Annals and Magazine of

Natural History, v. 9, p. 294-302.

Baumiller, T.K., 1993, Survivorship analysis of Paleozoic Crinoidea: effect of filter

morphology on evolutionary rates: Paleobiology, v. 19, p. 304-321.

Baumiller, T.K., 2008, Crinoid ecological morphology: Annual Review of Earth and

Planetary Sciences, v. 36, p. 221-249.

Baumiller, T.K., and Messing, C.G., 2007, Stalked crinoid locomotion and its ecological

and evolutionary implications: Palaeontologica Electronica, v. 10, 2A:10p.

Bell, M.A., and Lloyd, G.T., 2015, strap: an R package for plotting phylogenies against

stratigraphy and assessing their stratigraphic congruence: Palaeontology, v. 58, p.

379–389Benjamini, Y., and Hochberg, Y., 1995, Controlling the false discovery

rate: a practical and powerful approach to multiple test: Journal of the Royal

Statistical Society, Series B, v. 57, p. 289-300.

Benton, M.J., 2000, Stems, nodes, crown clades, and rank-free lists: is Linnaeus dead?:

Biological Reviews of the Cambridge Philosophical Society, v. 75, p. 633–648.

Benton, M.J., 2005, Vertebrate Paleontology, 3rd edition: Blackwell Publishing, Oxford.

347

Benton, M.J., 2007, The Phylocode: Beating a dead horse?: Acta Palaeontologica

Polonica, v. 52, p. 651–655.

Bergsten, J., Nilsson, A.N., and Ronquist, F., 2013, Bayesian tests of topology

hypotheses with an example from diving beetles: Systematic Biology, v. 62, p. 660‒

673.

Billings, E., 1857, New species of fossils from Silurian rocks of Canada: Canada Geological

Survey, Report of Progress 1853-1856, Report for the year 1856, p. 247–345.

Billings, W.R., 1887. A new genus and three new species of crinoids from the Trenton

Formation with notes on a large specimen of Dendrocrinus proboscidiatus: The Ottawa

Naturalist (Ottawa Field Naturalists' Club Transactions), v. 111, p. 49‒54.

Blumenbach, J. F., 1802–1804, Abbildungen naturhistorischer Gegenstiinde: Gottingen,

pt. 70, no. 70, 4 p.

Bokma, F., 2008, Detection of "punctuated equilibrium" by Bayesian estimation of

speciation and extinction rates, ancestral character states, and rates of anagenetic

and cladogenetic evolution on a molecular phylogeny: Evolution, v. 62, p. 2718–

26.

Bolker, J. A. and R. A. Raff. 1996, Developmental genetics and traditional homology:

Bioessays v. 18, p. 489-491.

Branson, E.B., and Peck, R.E., 1940, A new cystoid from the Ordovician of Oklahoma: Journal

of Paleontology, v. 14, p. 89–92.

348

Breimer, A. 1978, General morphology, recent crinoids. Pp. T9-T58 in R. C. Moore and

Teichert, C. (eds.), Treatise on Invertebrate Paleontology. Part T, Echinodermata

2. Geological Society of America and University of Kansas Press, Lawrence.

Brett, C.E., and Taylor, W.L., 1999, Middle Ordovician of the Lake Simcoe area of

Ontario, Canada, in Hess, H., Ausich, W.I., Brett, C.E., and Simms, M.H., Fossil

Crinoids. Cambridge University Press, p. 63-74.

Brigandt, I. 2003, Homology in comparative, molecular, and evolutionary developmental

biology: the radiation of a concept: Journal of Experimental Zoology, 299b:9-17.

Brochu, C.A., 2003, Phylogenetic approaches toward crocodylian history: Annual

Review of Earth and Planetary Sciences, v. 31, p. 357–397.

Brochu, C.A., and Sumrall, C.D., 2001, Phylogenetic nomenclature and paleontology:

Journal of Paleontology, v. 75, p. 754–757.

Brookfield, M. E. 1988, A mid-Ordovician temperate carbonate shelf—the Black River

and Trenton Limestone Groups of southern Ontario, Canada. Sedimentary

Geology, v., 60, p. 137-153.

Brookfield, M. E. and C. E. Brett. 1988, Paleoenvironments of the mid-Ordovician

(Upper Caradocian) Trenton Limestones of southern Ontario, Canada: storm

sedimentation on a shoal-basin shelf model. Sedimentary Geology, v. 57, p. 75-

105.

Brower, J.C., 1973, Crinoids from the Girardeau Limestone (Ordovician): Palaeontographica

Americana, v. 7, p. 263–499.

349

Brower, J. C. 1978, Camerates. Pp. T244-T263 in R. C. Moore and Teichert, C. (eds.),

Treatise on Invertebrate Paleontology. Part T, Echinodermata 2. Geological

Society of America and University of Kansas Press, Lawrence.

Brower, J.C., 1995, Dendrocrinid crinoids from the Ordovician of northern Iowa and

southern Minnesota: Journal of Paleontology, v.69, p. 939–960.

Brower, J. C., 2001, Flexible crinoids from the Upper Ordovician Maquoketa Formation

of the Northern Midcontinent and the evolution of early flexible crinoids: Journal

of Paleontology, v. 75, p. 370–382.

Brower, J.C., 2002, Cupulocrinus angustatus (Meek and Worthen, 1870), A cladid

crinoid from the Upper Ordovician Maquoketa Formation of the Northern

Midcontinent of the United States: Journal of Paleontology, v. 76, p. 109–122.

Brusatte, S.L., Benton, M.J., Ruta, M. and Lloyd, G.T., 2008, Superiority, competition,

and opportunism in the evolutionary radiation of dinosaurs: Science, v., 321,

p.1485–1488.

Brusatte, S.L., Lloyd, G.T., Wang, S.C., and Norell, M.A., 2014, Gradual assembly of

avian body plan culminated in rapid rates of evolution across the dinosaur-bird

transition: Current Biology, v. 24, p. 2386-2392.

Brusca, R. C. and G. J. Brusca. 2003, Invertebrates (2nd edition). Sinauer Associates,

Sunderland.

Bryant, H. N., 1989, An evaluation of cladistics and character analyses as hypothetico-

deductive procedures and the consequences for character weighting: Systematic

Biology, v. 38, p. 314-227. 350

Butler, M. A., and King, A.A., 2004, Phylogenetic Comparative Analysis: A Modeling

Approach for Adaptive Evolution: The American Naturalist, v. 164, p. 683-695.

Butler, R.J., Upchurch, P., and Norman, D.B., 2008, The phylogeny of the ornithischian

dinosaurs: Journal of Systematic Palaeontology, v. 6, p. 1–40.

Cantino, P.D., and de Queiroz, K., 2010, PhyloCode: International code of phylogenetic

nomenclature, version 4c, 90 p.

Carlson, S. J., 2001, Ghosts of the past, present, and future in brachiopod systematics:

Journal of Paleontology, v. 75, p. 1109–1118.

Carlson, S.J., and Leighton, L.R., 2001, The phylogeny and classification of

Rhynchonelliformea: The Paleontological Society Special Papers, v. 7, p. 27–51.

Carpenter, P. H., 1882, On the relations of Hybocrinus, Baerocrinus, and Hybocystites.

Quarterly Journal of the Geological Society of London, v. 38, p. 298-212.

Carpenter, P. H., 1884, Report on the Crinoidea—the stalked crinoids. Report on the

scientific results of the H. M. S. Challenger. Zoology, v. 11, p. 1-440.

Carpenter, W. B., 1866, Researches on the structure, physiology and development of

Antedon (Comatula) rosaceus: Philosophical Transactions of the Royal Society of

London, v. 156, p. 671-756.

Carroll, R. L., 1988, Vertebrate Paleontology and Evolution: WH Freeman, New York,

698 p.

Carroll, S. B. 2008, Evo-Devo and an expanding evolutionary synthesis: a genetic theory

of morphological evolution: Cell, v. 134, p. 25-36.

351

Clark, A. H. 1908. Description of new species of crinoids, chiefly from the collections

made by U.S. Fisheries steamer ‘‘Albatross’’ at the Hawaiian Islands in 1902,

with remarks on the classification of the Comatulida: Proceedings of the U.S.

National Museum, v. 34, p. 209–239.

Clark, A.H., 1915, A monograph of the existing crinoids, part 1: Bulletin of the United

States National Museum 82, p. 1–406.

Clark, A.H., 1921, A monograph of existing crinoids, part 2: Bulletin of the United States

National Museum 82, p. 1–795.

Clark, A.H., and Clark, A.M., 1967, A monograph of existing crinoids, part 5: Bulletin of

the United States National Museum 82, p. 1–806.

Clarke, J.A., and Middleton, K.M., 2008, Mosaicism, modules, and the evolution of

birds: results from a Bayesian approach to the study of morphological evolution

using discrete character data: Systematic Biology, v. 57, p. 185–201.

Clausen, S., Jell, P.A., Legrain, X., and Smith, A.B., 2009, Pelmatozoan arms from the

Middle Cambrian of Australia: bridging the gap between brachioles and

brachials?: Lethaia, v. 42, p. 283–296.

Close, R.A., Friedman, M., Lloyd, G.T. and Benson, R.B., 2015, Evidence for a mid-

Jurassic adaptive radiation in mammals: Current Biology, v. 25, p. 2137‒2142.

Cohen, B.L., Ameziane, N., Eleaume, M., and de Forges, B.R., 2004, Crinoid phylogeny:

a preliminary analysis (Echinodermata: Crinoidea): Marine Biology, v. 144, p.

605–617.

352

Cole, S.R., 2015, A phylogenetic test of the suprageneric classification of diplobathrid

crinoids: Geological Society of America Abstracts with Programs, v. 47, p. 853.

Cole, S.R., in press, Phylogeny and morphologic evolution of the Ordovician Camerata

(Class Crinoidea, Phylum Echinodermata): Journal of Paleontology.

Conrad, T.A., 1842, Descriptions of new species of organic remains belonging to the Silurian,

Devonian and Carboniferous Systems of the U. S.: Philadelphia Academy Natural

Sciences, Journal, v. 8, old ser. no. 2, p. 228‒280.

Darwin, C., 1859, On the origin of species by means of natural selection: Murray,

London, 502 p.

David, B., Lefebvre B., Mooi, R., and Parsley, R., 2000, Are homalozoans echinoderms?

An answer from extraxial-axial theory: Paleobiology, v. 26, p. 529–555.

David, B., Roux, M., Messing, C.G., and Ameziane, N., 2006, Revision of the

pentacrinid stalked crinoids of the genus Endoxocrinus (Echinodermata,

Crinoidea), with a study of environmental control characters and its consequences

for taxonomy: Zootaxa, v. 1156, p. 1–50.

Deline, B., and Ausich, W.I., 2011, Testing the plateau: a reexamination of disparity and

morphologic constraints in early Paleozoic crinoids: Paleobiology, v. 37, p. 214–

236.

de Queiroz, K., and Gauthier, J., 1990, Phylogeny as a central principle in taxonomy:

phylogenetic definitions of taxon names: Systematic Biology, v. 39, p. 307–322.

353

de Queiroz, K., and Gauthier, J., 1992, Phylogenetic taxonomy: Annual Review of

Ecology and Systematics, v. 23, p. 449–480.

de Queiroz, K., and Gauthier, J., 1994, Toward a phylogenetic system of biological

nomenclature: Trends in Ecology and Evolution, v. 9, p. 27–31.

de Pinna, M. G. G., 1991, Concepts and tests of homology in the cladistics paradigm:

Cladistics, v. 7, p. 387-394.

Didier, G., Royer-Carenzi, M. and Laurin, M., 2012, The reconstructed evolutionary

process with the fossil record: Journal of Theoretical Biology, v. 315, p.26‒37.

Donoghue, P.C.J., and Benton, M.J., 2007, Rocks and clocks: calibrating the Tree of Life using

fossils and molecules: Trends in Ecology and Evolution, v. 22, p. 424‒431.

Donovan, S.K., 1988, The early evolution of the Crinoidea, in Paul, C.R.C., and Smith, A.B.,

eds, Echinoderm phylogeny and evolutionary biology: Oxford, Clarendon Press, p. 235–

244.

Donovan, S.K., and Harper, D.A.T., 2003, Llandovery crinoidea of the British Isles, including

description of a new species from the Kilbride Formation (Telychian) of western Ireland:

Geological Journal, v. 38, p. 85–97.

Donovan, S.K., Webster, G.D., and Waters, J.A., 2016, A last peak in diversity: the

stalked echinoderms of the Permian of Timor: Geology Today, v. 32, p. 179-185.

dos Reis, M., Donoghue, P.C., and Yang, Z., 2016. Bayesian molecular clock dating of

species divergences in the genomics era: Nature Reviews Genetics, v. 17, p. 71‒

80.

354

Drummond, A.J., and Rambaut, A. 2007, BEAST: Bayesian evolutionary analysis by sampling

trees: BMC Evolutionary Biology: v. 7, doi:10.1186/1471-2148-7-214.

Drummond, A. J., and Stadler, T., 2016, Bayesian phylogenetic estimation of fossil ages:

arXiv, preprint 1601.07447v1

Eckert, J.D., 1987, Illemocrinus amphiatus, a new cladid inadunate crinoid from the

Middle Ordovician of Ontario: Canadian Journal of Earth Sciences, v. 24, p. 860-

865.

Eichwald, E., von, 1840, Ueber das Silurische Schichtensystem in Ethlands, St. Petersburg, 200

p.

Eldredge, N., and Gould, S.J., 1972, Punctuated equilibria: an alternative to phyletic

gradualism. in. Schopf, T.J.M., (ed.), Models in Paleobiology. Freeman, Cooper

and Company, San Fransisco, p. 82‒115.

Erwin, D.H., 1992, A preliminary classification of evolutionary radiations: Historical

Biology, v. 6, p. 133-147.

Erwin, D.H., 2007; Disparity: morphological pattern and developmental context:

Palaeontology, v. 50, p. 57-73.

Erwin, D. E., and E. H. Davidson, 2009, The evolution of hierarchical gene regulatory

networks. Nature Reviews Genetics, v. 10, p. 141-148.

Etter, W., and Hess, H., 2015, Reviews and syntheses: the first records of deep-sea fauna

– a correction and discussion: Biogeosciences, v. 12, p. 6453–6462.

Feller, W., 1968, An Introduction to Probability Theory and its Applications, v. 1 (third

edition): New York, Wiley, 509 p.

355

Feuda, R., and Smith, A.B., 2015, Phylogenetic signal dissection identifies the root of

starfishes: PLoS ONE, v. 10.5, e0123331.

Felsenstein, J., 1985. Phylogenies and the comparative method. American Naturalist, v.

126, p.1‒15.

Felsenstein, J., 2004, Inferring phylogenies, vol 2: Sunderland, Sinauer Associates, 664 p.

Foote, M. 1992, Paleozoic record of morphological diversity in blastozoan echinoderms:

Proceedings of the National Academy of Sciences: v. 89, p. 7325-7329.

Foote, M., 1994, Morphological disparity in Ordovician–Devonian crinoids and the early

saturation of morphological space: Paleobiology, v. 20, p. 320–344.

Foote, M., 1995, Morphological diversification of paleozoic crinoids: Paleobiology, v.

21, p. 273-299.

Foote, M., 1996, On the probability of ancestors in the fossil record: Paleobiology, v. 22,

p. 141–151.

Foote, M., 1996, Ecological controls on the evolutionary recovery of post-Paleozoic

crinoids: Science, v. 274, p. 1492-1495.

Foote, M., 1997, Estimating taxonomic durations and preservation probability:

Paleobiology, v. 23, p. 278–300.

Foote, M., 1999, Morphological diversity in the evolutionary radiation of Paleozoic and

post-Paleozoic crinoids: Paleobiology, v. 25, p. 1–115.

Foote, M., 2000, Origination and extinction components of taxonomic diversity: general

problems: Paleobiology, v. 26, p. 74-102.

Foote, M., and Miller, A., 2007, Principles of Paleontology. Macmillian.

356

Foote, M., and Raup, D.M., 1996, Fossil preservation and the stratigraphic ranges of taxa:

Paleobiology, v. 22, p. 121–140.

Foote, M., and Sepkoski, J.J., 1999, Absolute measures of the completeness of the fossil

record: Nature, v. 398, p. 415–417.

Forey, P.L., Fortey, R.A., Kenrick, P., and Smith, A.B., 2004, Taxonomy and fossils: a

critical appraisal: Philosophical Transactions of the Royal Society of London B:

Biological Sciences, v. 359, p. 639–653.

Frest, T.J., Strimple, H.L., and McGinnis, M.R., 1979, Two new crinoids from the Ordovician of

Virginia and Oklahoma, with notes on pinnulation in the Disparida: Journal of

Paleontology, v. 53, p. 399–415.

Gahn, F.J., and Kammer, T.W., 2002, The cladid crinoid Barycrinus from the Burlington

Limestone (Early Osagean) and the phylogenetics of Mississippian botryocrinids:

Journal of Paleontology, v. 76, p. 123–133.

Garstang, W., 1921.,The theory of Recapitulation: a critical re-statement of the biogenic

law: Journal of the Linnaean Society of London, v. 35, p. 81-101.

Gavryushkina, A., Welch, D., Stadler, T., and Drummond, A. J., 2014, Bayesian

inference of sampled ancestor trees for epidemiology and fossil calibration: PLoS

Computational Biology, v. 10, e1003919.

Gavryshkina, A., Heath, T.A., Ksepka, D.T., Stadler, T., Welch, D., and Drummond,

A.J., 2015, Bayesian total evidence dating reveals the recent crown radiation of

penguins: arXiv, preprint 1506.04797

357

Gelman, A. and Rubin, D. B., 1992, Inference from iterative simulation using multiple

sequences: Statistical Science, v. 7, p. 457–472.

Gervais, F.L.P., 1835, In Guérin, ed., Dictionnaire pittoresque d’histoire naturelle et des

phénomènes de la nature, vol. 3, p. 48–49, Lenormand, Paris.

Gilbert, S. F., and J. A., Bolker. 2001, Homologies of process and modular elements of

embryonic construction Pp. 437-456 in G. P. Wagner (ed.), The Character

Concept in Evolutionary Biology. Academic Press, London.

Gil Cid, M.D., Alonso, P.D, and Pobes, E.S., 1996, Reconstrucción y modo de videa de

Heviacrinus melendezi nov. gen. nov. sp. (Disparida, Iocrinidae), primer crinoide descrito

del Ordovícico medio de los Montes de Toledo (España): Revista de la Sociedad

Geológica España, v. 9, no. 1-2, p. 19‒27.

Goldfuss, G.A., 1826-44, Petrefacta Germaniae, tam ea, Quae in Museo Universitatis Regiae

Borussicae Fridericiae Wilhelmiae Rhenanea, serventur, quam alia quaecunque in Museis

Hoeninghusiano Muensteriano aliisque, extant, iconibus et descriiptionns illustrata. --

Abbildungen und Beschreibungen der Petrefacten Deutschlands und der Angränzende

Länder, unter Mitwirkung des Hern Grafen Georg zu Münster, herausgegeben von August

Goldfuss. v. 1 (1826-1833), Divisio prima. Zoophytorum reliquiae, p. 1-114; Divisio

secunda. Radiariorum reliquiae, p. 115-221 [Echinodermata]; Divisio tertia. Annulatorium

reliquiae, p. 222-242; v. 2 (1834-40), Divisio quarta. Molluscorum acephalicorum reliquiae.

I. Bivalvia, p. 65-286; II. Brachiopoda, p. 287-303; III. (1841-44), Divisio quinta.

Molluscorum gasteropodum reliquiae, p. 1-121; atlas of plates, 1–199, Düsseldorf, Arnz &

Co. 358

Goldring, W., 1923, The Devonian crinoids of the state of New York: New York State Museum,

Memoir 16, p. 1‒670.

Gorscak, E. and O‘Connor, P.M., 2016, Time-calibrated models support congruency

between Cretaceous continental rifting and titanosaurian evolutionary history:

Biology Letters, v. 12, p.20151047.

Gorzelak, P., Salamon, M.A., Trzęsiok, D., Lach, R., and Baumiller, T.K., 2015,

Diversity dynamics of post‐Palaeozoic crinoids–in quest of the factors affecting

crinoid macroevolution: Lethaia, v. 49, p. 231–244.

Gould, S. J., 1973, Systematic pluralism and the uses of history: Systematic Zoology, v.

22, p. 322-324.

Gould, S. J., 1977, Ontogeny and Phylogeny. Belknap Press of Harvard University Press,

Cambridge.

Gould, S. J., 1989, Wonderful Life: the Burgess Shale and the nature of history. Norton,

New York.

Gould, S.J., Gilinsky, N.L, and German, R.Z., 1987, Asymmetry of lineages and the

direction of evolutionary time: Science, v. 236, p. 1437–1441.

Gower, J.C., 1971, A general coefficient of similarity and some of its properties:

Biometrics, v. 27, p. 857-871.

Grant, P.R., and Grant,B.R., 2002, Adaptive radiation of Darwin’s finches: American

Journal of Science, v. 90, p. 130-139.

Guensburg, T.E., 2010, Alphacrinus new genus and origin of the disparid clade: Journal

of Paleontology, v. 84, p. 1209–1216. 359

Guensburg, T.E., 2012, Phylogenetic implications of the oldest crinoids: Journal of

Paleontology, v. 86, p. 455–461.

Guensburg, T.E., and Sprinkle, J., 2003, The oldest known crinoids (Early Ordovician,

Utah) and a new crinoid plate homology system: Bulletins of American

Paleontology, v. 364, 43 p.

Guensburg, T.E., and Sprinkle, J., 2007, Phylogenetic implications of the Protocrinoida:

Blastozoans are not ancestral to crinoids: Annals de Paleontologie, v. 93, p. 277–

290.

Guensburg, T.E., and Sprinkle, J., 2009, Solving the mystery of crinoid ancestry: new

fossil evidence of arm origin and development: Journal of Paleontology, v. 83, p.

350–364.

Guensburg, T.E., Blake, D.B., Sprinkle, J., and Mooi, R., 2016, Crinoid ancestry without

blastozoans: Acta Palaeontological Polonica, v. 61, p. 253–266.

Guillerme, T. and Cooper, N., 2016, Effects of missing data on topological inference

using a Total Evidence approach: Molecular Phylogenetics and Evolution, v. 94,

p.146–158.

Gyllenhaal, J.A., 1772, Beskrifning på de så kallade Crystall-äplen och kalkbollar, såsom

petreficerade Djur af Echini genus, eller dess närmaste slägtingar. Kongl. Svenska

Vetenskaps Academiens Handlingar: v. 33, p. 239–261.

360

Haeckel, E. 1866., Generelle Morphologie der Organismen. Allgemeine Grundzüge der

organischen Formen-Wissenschaft, mechanisch begründet durch die von Charles

Darwin reformirte Descendenz-Theorie. Georg Reimer, Berlin.

Hall, B. K., 2002, Palaeontology and evolutionary developmental biology: a science of

the nineteenth and twenty-first centuries: Palaeontology, v. 45, p. 647-669.

Hall, B. K , 2003, Descent with modification: the unity underlying homology and

homoplasy as seen through an analysis of development and evolution: Biology

Reviews, v. 78, p. 409-433.

Hall, J., 1847., Palaeontology of New York Volume I. Containing descriptions of the organic

remains of the lower division of the New-York system (equivalent of the Lower Silurian

rocks of Europe), Natural History of New York: New York, C. Van Benthuysen, Albany,

v. 6, 338 p.

Hall, J., 1852, Palaeontology of New York Volume II. Containing descriptions of the organic

remains of the lower middle division of the New-York system, Natural History of New

York: New York, D. Appleton & Co. and Wiley & Putnam; Boston, Gould, Kendall, &

Lincoln, v. 6, 362 p.

Hall, J., 1858, Chapter 8. Palaeontology of Iowa. in Hall, J. and Whitney, J.D., Report of the

Geological Survey of the state of Iowa. Embracing the results of investigations made during

portions of the years 1855, 56 &. 57, v. 1, part II; Palaeontology, p. 473–724.

Hall, J., 1859, Contributions to the palaeontology of Iowa, being descriptions of new species of

Crinoidea and other fossils: Geological Report of Iowa Supplement to vol. I, part II, p. 1-92.

361

Hall, J., 1866, Descriptions of new species of Crinoidea and other fossils from the Lower

Silurian strata of the age of the Hudson-River Group and Trenton Limestone: Albany,

privately distributed reprint,17 p.

Hara, Y., M. Yamaguchi, K. Akasaka, H. Nakano, M. Nonaka, and S. Amemiya. 2006,

Expression patterns of Hox genes in larvae of the sea lily Metacrinus rotundus:

Development Genes and Evolution , v. 216, p. 797-809.

Harmon, L. J., J. B. Losos, T. J. Davies, R. G. Gillespie, J. L. Gittleman, W. B. Jennings,

K. H. Kozak, M. A. McPeek, F. Morena-Roark, T. J. Near, A. Purvis, R. E.

Ricklefs, D. Schluter, J. A. Schulte II, O. Seehausen, B. L. Sidlauskas, O. Torres-

Carvajal, J. T. Weir, and A. Ø. Mooers. 2010. Early bursts of body size and shape

evolution are rate in comparative data: Evolution, v. 64, p. 2385-2396.

Harnik, P. G., P. C. Fitzgerald, J. L. Payne, and S. J. Carson 2014, Phylogenetic signal in

extinction selectivity in Devonian terebratulide brachiopods, Paleobiology, v. 40,

p. 679-692.

Harrison, L. B. and Larsson, H.C.E., 2015, Among-character rate variation distributions

in phylogenetic analysis of discrete morphologic characters: Systematic Biology, v.

64, p. 307–324.

Hays, B., Rouse, G., Thomas, J., and Messing, C., 2015. Can morphology support new

molecular phylogenies of Antedonidae (Crinoidea: Comatulida)?: 15th

International Echinoderm Conference Abstracts, p. 41–42.

362

Heath, T.A., and Moore, B.R., 2014, Bayesian inference of species divergence times, in

Chen, M-H., Kuo, L., and Lewis, P.O., Bayesian Phylogenetics: Methods,

Algorithms, and Applications: CRC Press, Oxfordshire, p. 277–318.

Heath, T.A., Hedtke, S.M., and Hillis, D.M., 2008, Taxon sampling and the accuracy of

phylogenetic analyses: Journal of Systematics and Evolution, v. 46, 239–257.

Heath, T.A., Huelsenbeck, J.P. and Stadler, T., 2014, The fossilized birth–death process

for coherent calibration of divergence-time estimates: Proceedings of the National

Academy of Sciences, v. 111, p. E2957–E2966.

Hennig, W., 1966, Phylogenetic systematics. University of Illinois Press, Urbana.

Heled, J., and Bouckaert, R.R., 2013, Looking for trees in the forest: summary tree from

posterior samples: BMC Evolutionary Biology, v. 13, 221.

Hemery, L.G., Roux, M., Ameziane, N., and Eleaume, M., 2013, High-resolution crinoid

phyletic inter-relationships derived from molecular data: Cahiers De Biologie

Marine, v. 54, p. 511–523.

Hess, H., and Messing, C.G., 2011, Treatise on Invertebrate Paleontology Part T,

Echinodermata 2, Revised, Crinoidea 3, in Ausich, W.I., coordinating author,

Seldon, P., ed., University of Kansas and Paleontological Institute, Lawrence,

Kansas, 216 p.

Hess, H., Ausich, W.I., Brett, C.E., and Simms, M.J., 1999, Fossil crinoids:

Cambridge University Press, 275 p.

363

Hetherington, A.J., Sherratt, E., Ruta, M., Wilkinson, M., Deline, B., and Donoghue,

P.C., 2015, Do cladistic and morphometric data capture common patterns of

morphological disparity?: Palaeontology, v. 58, p. 393-399

Holland, S.M., 2016, The non-uniformity of fossil preservation: Philosophical

Transactions of the Royal Society B, v. 371, 20150130.

Holterhoff, P.F., 1997, Filtration models, guilds, and biofacies: Crinoid paleoecology of

the Stanton Formation (Upper Pennsylvanian), midcontinent, North America:

Palaeogeography, Palaeoclimatology, Palaeoecology, v. 130, p. 177-208.

Holtz, T.R., 1996, Phylogenetic taxonomy of the Coelurosauria (Dinosauria: Theropoda):

Journal of Paleontology, v. 70, p. 536–538.

Holtz, T.R., 1998, A new phylogeny of the carnivorous dinosaurs: Gaia, v. 15, p. 5–61.

Hopkins, M.J. and Smith, A.B., 2015, Dynamic evolutionary change in post-Paleozoic

echinoids and the importance of scale when interpreting changes in rates of

evolution: Proceedings of the National Academy of Sciences, v. 112, p.3758–

3763.

Huelsenbeck, J.P., Ronquist, F., and Teslenko, M., 2015, Command Reference for

MrBayes ver. 3.2.5: http://mrbayes.sourceforge.net/manual.php (accessed June

2016).

Huelsenbeck, J.P., Larget, B., Miller, R.E. and Ronquist, F., 2002, Potential applications

and pitfalls of Bayesian inference of phylogeny: Systematic Biology, v. 51),

p.673–688.

364

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, v. 110, p. 13875-13879.

Hunt, G., 2008, Gradual or pulsed evolution: when should punctuational explanations be

preferred?: Paleobiology, v. 34, p.360–377.

Hunt, G., and Carrano, M.T, 2010, Models and methods for analyzing phenotypic

evolution in lineages and clades, in Alroy, J., and Hunt, G., (eds.), Short Course

on Quantitative Methods in Paleobiology. Paleontological Society, New Haven,

Connecticut, p. 245–269.

Hunt, G., and Slater, G., 2016, Integrating paleontological and phylogenetic approaches

to macroevolution: Annual Review of Ecology, Evolution, and Systematics, v. 47.

P. 189–213.

Hunt, G., Bell, M.A., and Travis, M.P., 2008, Evolution toward a new adaptive optimum:

phenotypic evolution in a fossil stickleback lineage: Evolution, v. 62, p. 700–710.

Hyman, L., 1955, The Invertebrates, Vol. 5. Phylum Echinodermata. McGraw-Hill, New

York.

Jablonski, D., 2008, Biotic interactions and macroevolution: extensions and mismatches

across scales and levels: Evolution, v. 62, p. 715-739.

Jablonski, D., and Finarelli, J.A., 2009, Congruence of morphologically-defined genera

with molecular phylogenies: Proceedings of the National Academy of Sciences, v.

106, p. 8262–8266.

365

Jablonski, in press, Macroevolutionary theory, in Scheiner, S.M., and Mindell, D.P.,

(eds.), The theory of evolution. Chicago: University of Chicago Press.

Jaekel, O., 1894, Über die Morphogenie und Phylogenic der Crinoiden: Sitzungsberichten der

Gesellschaft Naturforschender Freunde, Jahrgang 1894, v. 4, p. 101–121.

Jaekel, O., 1902, Über verschiedene Wege phylogenetischer Entwicklung: 5th Verhandlungen

der International Zoological-Congress Berlin, 1901, p. 1058‒1117.

Jaekel, O., 1906, Der oberste Lenneschiefer zwischen Letmathe und Iserlohn in Schmidt, W.E.,

Der oberste Lenneschiefer zwischen Letmathe und Iserlohn: Zeitschrift der Deutschen

Geologischen Geselleschaft, v. 57, p. 544.

Jaekel, O., 1918, Phylogenie und System der Pelmatozoen: Paläeontologische Zeitschrift,

v. 3, p. 1–128.

Kammer, T.W., Aerosol filtration theory applied to Mississippian deltaic crinoids:

Journal of Paleontology, v. 59, p. 551-560.

Kammer, T.W., 2001, Phenotypic bradytely in the Costalocrinus-Barycrinus lineage of

Paleozoic cladid crinoids: Journal of Paleontology, v. 75, p. 383‒389.

Kammer, T.W. 2008, Paedomorphosis as an adaptive response in pinnulate cladid

crinoids from the Burlington Limestone (Mississippian, Osagean) of the

Mississippi Valley. Pp. 177-195 in W. I. Ausich and G. D. Webster (eds.),

Echinoderm Paleobiology, University of Indiana Press, Bloomington.

Kammer, T.W., and Ausich, W.I., 1992, Advanced cladid crinoids from the Middle

Mississippian of the east-central United States. primitive-grade calyces: Journal of

Paleontology, v. 66, p. 461-480.Kammer and Ausich, 1993

366

Kammer, T.W., and Ausich. W.I., 1996, Primitive cladid crinoids from upper Osagean-lower

Meramecian (Mississippian) rocks of east-central United States: Journal of Paleontology, v.

70, p. 835-866.

Kammer, T.W., and Ausich, W.I., 2006, The “age of crinoids”: a Mississippian

biodiversity spike coincident with widespread carbonate ramps. Palaios, v. 2, p. 238-

248.

Kammer, T.W., and Ausich, W.I., 2007.,New cladid and flexible crinoids from the

Mississippian (Tournaisian, Ivorian) of England and Wales: Palaeontology, v. 50,

p. 1039-1050.

Kammer, T.W., and Gahn, F.J., 2003, Primitive cladid crinoids from the early Osagean

Burlington Limestone and the phylogenetics of Mississippian species of

Cyathocrinites: Journal of Paleontology, v. 77, p. 121–138.

Kammer, T.W., Sumrall, C.D., Zamora, S., Ausich, W.I., and Deline, B., 2013, Oral

region homologies in Paleozoic crinoids and other plesiomorphic pentaradiate

echinoderms: PLoS ONE, v. 8.11, e77989.

Kass, R.E., and Raftery, A.E., 1995, Bayes factors. Journal of the American Statistical

Association, v. 90, p. 773‒795.

Kelley, P.H., Fastovsky, D.E., Wilson, M.A., Laws, R.A., and Raymond, A., 2013, From

paleontology to paleobiology: a half-century of progress in understanding life

history: Geological Society of America Special Paper, v. 500, p. 191–232.

367

Kelly, S.M., 1982, Origin of the crinoid orders Disparida and Cladida: possible inadunate

cup plate homologies: Third North American Paleontological Convention

(Montreal), Proceedings Abstracts, Montreal, Canada, p. 285–290.

Kelly, S.M., 1986, Classification and evolution of class Crinoidea: Fourth North

American Paleontological Convention, Abstracts, p. A23.

Kesling, R.V., and Paul, C.R.C., 1968, New species of Porocrinidae and brief remarks upon

these unusual crinoids: University of Michigan Contributions from Museum of

Paleontology, v. 22, p. 1–32.

Kier, P.M., 1952, Echinoderms of the Middle Devonian Silica Formation of Ohio: University of

Michigan Contributions from Museum of Paleontology, v. 10, p. 59‒81.

Kirk, E., 1914, Notes on the fossil crinoid genus Homocrinus Hall. United States

National Museum Proceedings: v. 46, p. 473–483.

Kitazawa, K., Oji, T., and Sunamura, M., 2007, Food composition of crinoids (Crinoidea:

Echinodermata) in relation to stalk length and fan density: their paleoecological

implications: Marine Biology, v. 152, p. 959-968.

Kirk, E. 1944, Cymbiocrinus, a new inaduate crinoid genus from the Upper

Mississippian: American Journal of Science, v. 242, p. 233-245.

Kolata, D.R., 1982, Camerates, in Sprinkle, J., ed., Echinoderm faunas from the Bromide

Formation (Middle Ordovician) of Oklahoma: The University of Kansas

Paleontological Contributions, Monograph, v. 1, p. 170–205.

Ksepka, D.T., Parham, J.F., Allman, J.F., Benton, M.J., Carrano, M.T., Cranston, K.A.,

Donoghue, P.C., Head, J.J., Hermsen, E.J., Irmis, R.B. and Joyce, W.G., 2015. The Fossil 368

Calibration Database—A new resource for divergence dating: Systematic biology, v. 64,

p.853‒859.

Koenig, J.W., and Meyer, D.L., 1965, Two new crinoids from the Devonian of New York:

Journal of Paleontology, v. 39, p. 391‒397.

Lahaye, M.C., and Jangoux, M., 1987, The skeleton of the stalked stages of the comatulid

crinoid Antedon bifida (Echinodermata): Zoomorphology, v. 107, p. 58–65.

Lane, N.G., 1970, Lower and Middle Ordovician crinoids from west-central Utah: Brigham

Young University Geology Studies, v. 17, p. 3‒17.

Lane, N.G., 1978, Historical review of classification of Crinoidea, in Moore, R.C., and

Teichert, C., eds., Treatise on Invertebrate Paleontology, Part T, Echinodermata 2,

Volume 2: Geological Society of America and University of Kansas Press,

Lawrence, p. T348–T358.

Laubichler, M.D., 2000, Homology in development and the development of the

homology concept: American Zoologist, 40, p. 777-788.

Lee, M.S., and Palci, A., 2015, Morphological phylogenetics in the genomic age: Current

Biology, v. 25, p. R922–R929.

Lee, M.S., Cau, A., Naish, D. and Dyke, G.J., 2014, Morphological clocks in

paleontology, and a mid-Cretaceous origin of crown Aves: Systematic Biology, v.

63, p. 442‒449.

Lefebvre, B. G. J. Eble, N. Navarro, and B. David. 2006, Diversification of atypical

Paleozoic echinoderms: a quantitative survey of patterns of stylophoran disparity,

diversity, and geography: Paleobiology, v. 32, p. 483-510. 369

Lieberman, B. S., 2000, Paleobiogeography: Using Fossils to Study Global Change, Plate

Tectonics, and Evolution. New York: Kluwer Academic/Plenum Publishers.

Le Menn, J., and Spjeldnaes, N., 1996, Un nouveau crinoïde Dimerocrinitidae (Camerata,

Diplobathrida) de l'Ordovicien supérieur du Maroc: Rosfacrinus robustus nov. gen., nov.

sp.: Geobios, v. 29, p. 341–351.

Lepage, T., Bryant, D., Philippe, H. and Lartillot, N., 2007, A general comparison of

relaxed molecular clock models: Molecular Biology and Evolution, v. 24, p.

2669‒2680.

Leuckart, R. von, 1848, Uber de Morphologie und die Verandtschaftsverhältnisse der

wibellosen Thiere: Friedrich ,Vieweg and Sohn, Braunschweig, 180 p.

Lewis, P.O., 2001, A likelihood approach to estimating phylogeny from discrete

morphological character data: Systematic biology, v. 50, p. 913‒925.

Lloyd, G.T., 2016, Estimating morphological diversity and tempo with discrete character-

taxon matrices: implementation, challenges, progress, and future directions:

Biological Journal of the Linnaean Society, v. 118, p. 131-151.

Lloyd, G.T., Wang, S.C. and Brusatte, S.L., 2012, Identifying heterogeneity in rates of

morphological evolution: discrete character change in the evolution of lungfish

(Sarcopterygii; Dipnoi): Evolution, v. 66, p.330‒348.

Losos, J.B., 2010, Adaptive radiation, ecological opportunity, and evolutionary

determinism: The American Naturalist, v. 175, p. 623-639.

Maddison, D.R., 1991, The discovery and importance of multiple islands of most-

parsimonious trees: Systematic Zoology, v. 40, p. 315‒328.

370

Matsumoto, H., 1929, Outline of a classification of Echinodermata: Science Reports of

the Tohoku Imperial University, Sendai, Japan, Second Series (Geology): v. 8, p.

27–33.

Matzke, N.J., 2015, The evolution of antievolution policies after Kitzmiller v. Dover.

Science, p.aad4057.

Matzke, N.J. and Wright, A., 2016, Ground truthing tip-dating methods using fossil

Canidae reveals major differences in performance: Biology Letters, v. 12,

2010328, doi:10.1098/rsbl.2016.0328.

McIntosh, G.C., 1984, Devonian cladid inadunate crinoids: family Botryocrinidae Bather,

1899: Journal of Paleontology, v. 58, p. 1260–1281.

McIntosh, G.C., 1986, Phylogeny of the dicyclic inadunate crinoid order Cladida: Fourth

North American Paleontological Convention Abstracts A, v. 31.

McIntosh, G.C., 2001, Devonian cladid crinoids: families Glossocrinidae Goldring, 1923,

and Rutkowskicrinidae new family: Journal of Paleontology, v. 75, p. 783–807.

McKerrow, W.S., Dewey, J.F. and Scotese, C.R., 1991, The Ordovician and Silurian

development of the Iapetus ocea: Special Papers in Palaeontology, v. 44, p.165-

178.

Meek, F.B., 1871, On some new Silurian (Ordovician) crinoids and shells: American

Journal of Science, ser. 3, v. 1, p. 295‒299.

Meek, F.B., 1873, Descriptions of invertebrate fossils of the Silurian and Devonian

systems: Ohio Geological Survey, v. 1, pt. 2, p. 1–243.

371

Meek, F.B., and Worthen, A.H., 1865, Descriptions of new species of crinoidea, etc. from the

Paleozoic rocks of Illinois and some of the adjoining states: Proceedings of the Academy

of Natural Sciences of Philadelphia, v. 17, p. 143‒155.

Meek, F.B., and Worthen, A.H., 1869, Descriptions of New Crinoidea and Echinoidea,

from the Carboniferous Rocks of the Western States, with a Note on the Genns

Onychaster: Proceedings of the Academy of Natural Sciences of Philadelphia, v.

21, p. 67–83.

Messing, C.G., and White, C.M., 2001, A revision of the Zenometridae (new rank)

(Echinodermata, Crinoidea, Comatulidina): Zoologica Scripta, v. 30, p. 159–180.

Meyer, D.L., and Macurda, B.D., 1977, Adaptive radiation of the comatulid crinoids:

Paleobiology, v. 3, p. 74–82.

Meyer, D.L., Ausich, W.I., and Terry, R.E., 1989, Comparative taphonomy of

echinoderms in carbonate facies: Fort Payne Formation (Lower Mississippian) of

Kentucky and Tennessee: Palaios, v. 4, p. 533-552.

Miller, J.S., 1821, A Natural History of the Crinoidea, or Lily-Shaped Animals; with

Observations on the Genera, Asteria, Euryale, Comatula and Marsupites: Bristol,

C. Frost, 150 p.

Miller, S.A., 1883, The American Palaeozoic fossils: a catalogue of the genera and

species, with names of authors, dates, places of publication, groups of books in

which found, and the etymology and significance of the words, and an

introduction devoted to the stratigraphical geology of the Palaeozoic rocks, 2nd

edition, Echinodermata, Cincinnati, Ohio, p. 247-334.

372

Miller, S.A. and Gurley, W.F.E., 1894, New genera and species of Echinodermata:

Illinois State Museum of Natural History, v. 5, 53 p.

Miller, S.A., and Gurley, W.F.E., 1895, New and interesting species of Palaeozoic

fossils: Illinois State Museum, Bulletin 7, 89, p.

Miquel, J. 1894. Note sur la géologie des terrains primaires du département de l’Hérault.

Le Cambrien et l’Arenig: Bulletin de la Société d’étude des Sciences naturelles de

Béziers v. 17, p. 5–36.

Mladenov, P. V. and F. S. Chia, 1983, Development, settling behavior, metamorphosis,

and pentacrinoid feeding and growth of the feather star Florometra serratissima.

Marine Biology, v. 73, p. 309-323.

Mooi, R. B. and B. David. 1997, Skeletal homologies of echinoderms. Pp. 305-335 in J.

A. Waters and C. G. Maples (eds.), Geobiology of Echinoderms, Paleontological

Society Papers 3.

Mooi, R., and David, B., 1998, Evolution within a bizarre phylum; homologies of the

first echinoderms: American Zoologist, v. 38, p. 965–974.

Mooi, R., and David, B., 2008, Radial symmetry, the anterior/posterior axis, and

echinoderm Hox genes: Annual Review of Ecology, Evolution, and Systematics,

v. 39, p. 43–62.

Mooi, R., B. David, and G. A. Wray, 2005, Arrays in rays: terminal addition in

echinoderms and its correlation with gene expression: Evolution and

Development, v. 7, p. 542-555.

373

Mooi, R. B., David, B., and D., Marchland. 1994, Echinoderm skeletal homologies:

classical morphology meets modern phylogenetics. Pp. 87-95 in B. David, A.

Guille, J. Féral, and M Roux (eds.), Echinoderms Through Time. A. A. Balkema,

Rotterdam.

Moore, R.C., 1962, Ray structures of some inadunate crinoids. University of Kansas

Paleontological Contributions, v. 5, 1–47.

Moore, R.C., and Laudon, L.R., 1943, Evolution and classification of Paleozoic crinoids:

Geological Society of America Special Paper, no. 46, 151 p.

Moore, R.C., and Teichert, C., eds., 1978, Treatise on Invertebrate Paleontology, Part T,

Echinodermata 2: Geological Society of America and University of Kansas Press,

Lawrence, Kansas, 1027 p.

Moore, R. C., Lalicker, C.G., and Fischer, A.G., 1952, Invertebrate Fossils: McGraw-Hill

Book Company, New York, 766 p.

Moore, R.C., Lane, N.G., and Strimple, H.L., 1978, Order Cladida Moore and Laudon,

1943, in Moore, R. C. and Teichert, C., eds: Treatise on Invertebrate

Paleontology, Pt. T, Echinodermata 2: Geological Society of America and

University of Kansas Press, Kansas, p. T578–T759.

Moore, R.C., Lane, N.G., Strimple, H.L., and Sprinkle, J., 1978, Systemetic Descriptions,

Crinoidea, Order Disparida, in Moore, R. C. and Teichert, C., eds: Treatise on

Invertebrate Paleontology, Pt. T, Echinodermata 2: Geological Society of

America and University of Kansas Press, Kansas, p. T520–T564.

374

Mortensen, T. 1920, Studies in the development of crinoids. Papers from the Department

of Marine Biology, Carnegie Institution of Washing 16.

Müller, J., 1841, Über die Gattungen und Arten der Comatulen: Archiv für

Naturgeschichte, v. 7, p. 179–189.

Nakano, H., T. Hibino, T. Oji, Y. Hara, and S. Amemiya. 2003, Larval stages of a living

sea lily (stalked crinoid echinoderm): Nature, v. 421, p. 158-160.

O’Malley, C.E., Ausich, W.I., and Chin, Y., 2016, Deep echinoderm phylogeny

preserved in organic molecules from Paleozoic fossils: Geology, v. 44, p. 379–

382.

O'Meara, B.C., C. Ané, M. J. Sanderson, P. C. Wainwright, and T. Hansen. 2006, Testing

for Different Rates of Continuous Trait Evolution Using Likelihood. Evolution, v.

60, p. 922-933.

Orbigny, A.D. d', 1849, Cours élémentaire de paléontologie et géologie Stratigraphiques:

v. 1, 299 p. Paris, Victor Masson.

O’Reilly, J.E., Puttick, M.N., Parry, L., Tanner, A.R., Tarver, J.E., Fleming, J., Pisani, D.,

and Donoghue, P.C.J., 2016, Bayesian methods outperform parsimony but at the

expense of precision in the estimation of phylogeny from discrete morphological

data: Biology Letters, v. 12, 20160081, http://dx.doi.org/10.1098/rsbl.2016.0081

Oyston, J.W., Hughes, M., Wagner, P.J., Gerber, S., and Wills, M.A., 2015, What limits

the morphological disparity of clades?: Interface Focus, v. 5, 20150042, 18 p.

375

Padian, K., Hutchinson, J.R., and Holtz, T.R., 1999, Phylogenetic definitions and

nomenclature of the major taxonomic categories of the carnivorous Dinosauria

(Theropoda): Journal of Vertebrate Paleontology, v. 19, p. 69–80.

Paradis, E., Claude, J. and Strimmer, K., 2004, APE: analyses of phylogenetics and

evolution in R language: Bioinformatics, v. 20, p. 289–290.

Parks, W.A., 1908, On an occurrence of Hybocystites in Ontario: Ottawa Naturalist, v.

21, p. 232– 236.

Patterson, C., 1982, Morphological characters and homology. Pp. 21-74 in K. A. Joysey

and A. E. Friday (eds.), Problems of Phylogenetic Reconstruction. Academic

Press, New York.

Paul, C.R.C., 1968, Macrocystella Callaway, the earliest glyptocystitid cystoid:

Palaeontology, v. 11, p. 580–600.

Paul, C.R.C., 1972, Cheirocystella antiqua gen. et sp. nov. from the Lower Ordovician of

western Utah and its bearing on the evolution of the Cheirocrinidae (Rhombifera:

Glyptocystitida): Brigham Young University Geology Studies, v. 19, p. 15–63.

Paul, C.R.C., 1988, The phylogeny of the cystoids, in Paul, C.R.C., and Smith, A.B., eds.,

Echinoderm phylogeny and evolutionary biology: Clarendon Press, Oxford, p.

199–213.

Paul, C.R.C., and Smith, A.B., 1984, The early radiation and phylogeny of echinoderms:

Biological Reviews v. 59, p. 443–481.

Pennant, T., 1777, The British Zoology (4th edition, volume 4): London, Warrington.

376

Pennell, M.W., Harmon, L.J., Uyeda, J.C., 2014, Is there room for punctuated

equilibrium in macroevolution?: Trends in Ecology and Evolution, v. 29, p. 23‒

32.

Peters, S.E., and Ausich, W.I., 2008, A sampling-adjusted macroevolutionary history for

Ordovician-Early Silurian crinoids: Paleobiology, v. 34, p. 104‒116.

Phillips, J., 1836, Illustrations of the geology of Yorkshire, or a description of the strata

and organic remains, Pt. 2, The Mountain Limestone districts, 2nd edition:

London, John Murray, p. 203–208.

Phillips, J., 1839, Encrinites and zoophytes of the Silurian System, in Murchison, R.T.,

ed., The Silurian System, p. 670‒675Phillips, J., 1841, Figures and descriptions of

the Palaeozoic fossils of Cornwall, Devon, and West Somerset; observed in the

course of the ordinance geological survey of that district: London, Longmans,

Brown, Green, and Longmans, 232 p.

Pisani, D., Feuda, R., Peterson, and Smith, A.B., 2012, Resolving phylogenetic signal

from noise when divergence is rapid: A new look at the old problem of

echinoderm class relationships: Molecular Phylogenetics and Evolution, v. 62, p.

27–34.

Pollitt, J.R., Fortey, R.A. and Wills, M.A., 2005. Systematics of the trilobite families

Lichidae Hawle and Corda, 1847 and Lichakephalidae Tripp, 1957: The

application of Bayesian inference to morphological data: Journal of Systematic

Palaeontology, v. 3, p. 225‒241.

377

Potter, D., and Freudenstein, J.V., 2005, Character-based phylogenetic Linnaean

classification: taxa should be both ranked and monophyletic: Taxon, v. 54, p.

1033–1035.

Purvis, A. 2008, Phylogenetic approaches to the study of extinction: Annual Review of

Ecology, Evolution, and Systematics, v. 39, p. 301-319.

Pyron, R.A., 2011, Divergence Time Estimation Using Fossils as Terminal Taxa and the

Origins of Lissamphibia: Systematic Biology, v. 60, p. 466‒481.

Pyron, R.A., 2015, Post-molecular systematics and the future of phylogenetics: Trends in

Ecology and Evolution, v. 30, p. 394–389.

Rabosky, D.L., 2009, Ecological limits and diversification rate: alternative paradigms to

explain the variation in species richness among clades and regions: Ecology

Letters, v. 12, p. 735–743.

Rabosky, D.L., and McCune, A.R., 2009, Reinventing species selection with molecular

phylogenies: Trends in Ecology and Evolution, v. 25, p. 68–74.

Rabosky, D.L., and Adams, D.C., 2012, Rates of morphological evolution are correlated

with species richness in salamanders: Evolution, v. 66, p. 1807-1818.

Raff, R. A., 2007, Written in stone: fossils, genes, and evo-devo: Nature Reviews

Genetics v. 8, p. 911-920.

Rannala, B., and Yang, Z., 1996, Probability distribution of molecular evolutionary trees:

a new method of phylogenetic inference: Journal of Molecular Evolution, v. 43, p.

304–311.

378

Rasmussen, H. W., 1978, Evolution of articulate crinoids. Pp T302-T316 in R. C. Moore

and Teichert, C. (eds.), Treatise on Invertebrate Paleontology. Part T,

Echinodermata 2. Geological Society of America and University of Kansas Press,

Lawrence.

Rasmussen, H.W., 1978, Articulata, in Moore, R.C., and Teichert, K. eds., Treatise on

Invertebrate Paleontology. Part T, Echinodermata 2, vol. 2: The Geological

Society of America, and University of Kansas Press, Boulder and Lawrence, p.

T813–T997.

Rambaut, A., and Drummond, A.J., 2015, TreeAnnotator v2 2.1: MCMC ouput analysis.

Rambaut, A. 2014, Summarizing posterior trees: http://beast.bio.ed.ac.uk/summarizing-

posterior-trees (accessed June 2016).

Raup, D. M., 1985, Mathematical models of cladogenesis: Paleobiology, v. 11, p. 42–52.

Raup, D. M., Gould, S. J., Schopf, T. J. M., and Simberloff, D. S., 1973, Stochastic

models of phylogeny and the evolution of diversity: Journal of Geology, v. 81, p.

525–542.

Reich, A., Dunn, C., Akasaka, K., and Wessel, G., 2015, Phylogenomic analyses of

Echinodermata support the sister groups of Asterozoa and Echinozoa: PLoS ONE,

v. 10, e0119627.

Remane, A., 1952, Die Grundlagen des Naturlichen Systems der Vergleichenden

Anatomie und der Phylogenetik. Geest und Portig, Leipzig.

Revell, L. J., 2012, phytools: an R package for phylogenetic comparative biology (and

other things): Methods in Ecology and Evolution, v. 3, p. 217-223.

379

Robinson, D.F., and Foulds, L.R., 1981, Comparison of phylogenetic trees: Mathematical

Biosciences, v. 53, p. 131–147.

Ronquist, F., Klopfstein, S., Vilhelmsen, L., Schulmeister, S., Murray, D.L. and

Rasnitsyn, A.P., 2012, A total-evidence approach to dating with fossils, applied to

the early radiation of the Hymenoptera: Systematic Biology, v. 61, p. 973–999.

Rouse, G.W., Jermiin, L.S., Wilson, N.G., Eeckhaut, I., Lanterbecq, D., Oji, T., Young,

C.M., Browning, T., Cisternas, P., Helgen, L.E., Stuckey, M., and Messing, C.G.,

2013, Fixed, free, and fixed: the fickle phylogeny of extant Crinoidea

(Echinodermata) and their Permian–Triassic origin: Molecular Phylogenetics and

Evolution v. 66, p. 161–181.

Rouse, G.W, Carvajal, J., Oji, T., and Messing, C.G., 2015, Further insights into extant

crinoid phylogeny from molecular sequence data: 15th International Echinoderm

Conference Abstracts, p. 67–68.

Roux, M., 1987, Evolutionary ecology and biogeography of recent stalked crinoids as a

model for the fossil record, in Jangoux, M., and Lawrence, J.M., eds., Echinoderm

studies, vol. 2: A. A. Balkma, Rotterdam, the Netherlands, p. 1–53.

Roux, M., Eleaume, M., Hemery, L.G., and Ameziane, N., 2013, When morphology

meets molecular data in crinoid phylogeny: a challenge: Cahiers de Biologie

Marine, v. 54, p. 541–548.

380

Rozhnov S. V. and G. V. Mirantsev. 2014, Structural aberrations in the cup in cladid

crinoids from the Carboniferous of the Moscow region: Paleontological Journal,

v. 48, 1243-1257.

Ruta, M., Coates, M.I., and Quicke, D.L.J., 2003, Early tetrapod relationships revisited:

Biological Reviews, v. 78, p. 351–345.

Saint-Hilaire, E. G., 1830, Principes de Philosophie Zoologique, discutés en Mars 1830,

au Sein de l’Académie Royale des Sciences. Pichon et Dider, Paris

Sallan, L.C., Kammer, T.W., Ausich, W.I., and Cook, L.A., 2011, Persistent predator-

prey dynamics revealed by mass extinction: Proceedings of the National Academy

of Science, v. 108, p. 8335–8338.

Salter, J.W., 1873, A catalogue of the collection of Cambrian and Silurian fossils

contained in the geological museum of the University of Cambridge: Cambridge,

Cambridge University Press, 204 p.

Schuchert, C., 1900, On the Lower Silurian (Trenton) fauna of Baffin Land: Proceedings

of the U. S. National Museum, v. 22, p. 143–178.

Schlotheim, E.F. von, 1820, Die Petrefactenkunde auf ihrem jetzigen Standpunkte durch

die Beschreibung seiner Sammlung versteinerter und fossiler Überreste des Thier-

und Pflanzenreichs der Vorwelt erläutert: Gotha, Beckersche Buchhandlung, p.

437 p.

Schluter, D., 2000, The ecology of adaptive radiation. OUP Oxford.

381

Schultze, L., 1867, Monographie der Echinodermen des Eifler Kalkes: Denkschriften der

Kaiserlich Akademie der Wissenschaften Mathematisch-Naturwissenschaftlichen

Classe, Wien, v. 26, no. 2, p. 113‒230.

Sepkoski, J. J., Jr. 1981, A factor analytic description of the Phanerozoic marine fossil

record: Paleobiology, v. 7, p. 36–53.

Sereno, P.C., 1997, The origin and evolution of dinosaurs: Annual Review of Earth and

Planetary Sciences, v. 25, p. 235–489.

Sereno, P.C., 1999, Definitions in phylogenetic taxonomy: critique and rationale:

Systematic Biology, v. 48, p. 329–351.

Sereno, P.C., 2005, The logical basis of phylogenetic taxonomy: Systematic Biology, v.

54, p. 595–619.

Sereno, P.C., McAllister, S., and Brusatte, S.L., 2005, TaxonSearch: a relational database

for suprageneric taxa and phylogenetic definitions: PhyloInformatics, v. 8, p. 1–

21.

Sevastopulo, G.D., and Lane, N.G., 1988, Ontogeny and phylogeny of disparid crinoids,

in Paul, C.R.C., and Smith, A.B., eds, Echinoderm phylogeny and evolutionary

biology: Oxford, Clarendon Press, p. 245–253.

Signor P.W., and Brett, C.E., 1984, The mid-Paleozoic precursor to the Mesozoic marine

revolution: Paleobiology, v. 10, p. 229-245.

382

Shibata, T.F., Sato, A., Oji, T., and Akasaka, K., 2008, Development and growth of the

feather star Oxycomantus japonicus to sexual maturity: Zoological Science, v. 25,

p. 1075–1083.

Sibley, C.G., 1994, On the phylogeny and classification of living birds: Journal of Avian

Biology, v. 25, p. 87–92.

Shubin, N. H., and C. R., Marshall, 2000, Fossils, genes, and the origin of novelty:

Paleobiology v. 26, p. 324-340.

Simms, M.J., 1988, The phylogeny of post-Palaeozoic crinoids, in Burke, R. D., P. V.

Mladenov, P. Lambert, and R. L. Parsley, eds, Echinoderm Biology: Balkema,

Rotterdam, p. 97–102.

Simms, M. J. 1993, Reinterpretation of thecal plate homology and phylogeny in the Class

Crinoidea: Lethaia, v. 26, p. 303-312.

Simms, M.J., and Sevastopulo, G.D., 1993, The origin of articulate crinoids:

Palaeontology, v. 36, p. 91–109.

Simões, M., Breitkreuz, L., Alvarado, M., Baca, S., Cooper, J.C., Heins, L., Herzog, K.

and Lieberman, B.S., 2016, The Evolving Theory of Evolutionary Radiations.

Trends in ecology & evolution, v. 31, p.27-34.

Simpson, G.G., 1944, Tempo and Mode in Evolution: Columbia University Press, New

York.

Simpson, G.G., 1955. Major features of evolution. Columbia University Press, 434 p.

Siverts-Doreck, H. 1953. Sous-classe 4. Articulata, p. 756–765. In J. Priveteau (ed.),

Traité de paléontologie. Volume 3. Masson and Cie, Paris. 383

Slater, G.J., 2013, Phylogenetic evidence for a shift in the mode of mammalian body size

evolution at the Cretaceous-Palaeogene boundary: Methods in Ecology and

Evolution, v. 4, p. 734‒744.

Slater, G.J., 2015, Not so early bursts and the dynamic nature of morphological

diversification: Proceedings of the National Academy of Sciences, v. 112, p.

3595-3596.

Slater, G.J., 2015, Iterative adaptive radiations of fossil canids show no evidence for

diversity-dependent trait evolution: Proceedings of the National Academy of

Sciences, v. 112, p. 4897‒4902.

Slater, G.J., and Harmon, L.J., 2013, Unifying fossils and phylogenies for comparative

analyses of diversification and trait evolution: Methods in Ecology and Evolution,

v. 4, p. 699‒702.

Slater, G.J., Price, S.A., Santini, F. and Alfaro, M.E., 2010, Diversity versus disparity and

the radiation of modern cetaceans. Proceedings of the Royal Society of London B:

Biological Sciences, v. 277, p. 3097-3104.

Smith, A.B., 1984, Classification of the Echinodermata: Palaeontology, v. 27, p. 431–

459.

Smith, A.B., 1985, Cambrian eleutherozoan echinoderms and the early diversification of

edrioasteroids: Palaeontology, v. 28, p. 87–89.

Smith, A.B., 1988, Patterns of diversification and extinction in early Palaeozoic

echinoderms: Palaeontology, v. 31, p. 799–828.

384

Smith, A.B., 1990, Evolutionary diversification of the echinoderms during the Early

Paleozoic: Systematics Association, Special Volume 42, p. 256–286.

Smith, A.B., 1994, Systematics and the Fossil Record: Documenting Evolutionary

Patterns: Blackwell Science, Oxford, 223 p.

Smith, A.B., 2001, Large-scale heterogeneity of the fossil record: implications for

Phanerozoic biodiversity studies: Philosophical Transactions of the Royal

Society, London: v. 356, p. 351-368.

Smith, A.B., 2005, The pre‐radial history of echinoderms: Geological Journal, v. 40, p.

255-280.

Smith, A.B., 2008, Duterostomes in a twist: the origins of a radical new body plan:

Evolution and Development, v. 10, p. 493–503.

Smith, A.B., and Jell, P.A., 1990, Cambrian edrioasteroids from Australia and the origin

of starfishes: Queensland Museum Memoir, v. 28, p. 715–778.

Smith, A.B., and Zamora, S., 2013, Cambrian spiral-plated echinoderms from Gondwana

reveal the earliest pentaradial body plan: Proceedings of the Royal Society B, v.

280:20131197, doi.org/10.1098/rspb.2013.1197

Smith, A.B., Lafay, B., and Christen, R., 1992, Comparative variation of morphological

and molecular evolution through geologic time: 28S ribosomal RNAversus

morphology in echinoids: Philosophical Transactions: Biological Sciences, v.

338, p. 365‒382.

Snively, E., Russell, A.P. and Powell, G.L., 2004, Evolutionary morphology of the

coelurosaurian arctometatarsus: descriptive, morphometric and phylogenetic 385

approaches: Zoological Journal of the Linnean Society, v. 142, p. 525‒553Soul,

L.C., and Friedman, M., 2015, Taxonomy and phylogeny can yield comparable

results in comparative paleontological analyses: Systematic Biology, v. 64, p.

608–620.

Spencer, M.R., and Wilberg, E.W., 2013. Efficacy or convenience? Model‐based

approaches to phylogeny estimation using morphological data: Cladistics, v. 29,

p. 663‒671.

Springer, F., 1911, On a Trenton echinoderm fauna: Canada Department Mines Memoir

no. 15-P, 70 p.

Springer, F., 1913, Crinoidea, in Zittel, K.A. von, ed., Text-book of paleontology

(translated and edited by C. R. Eastman), 2nd edit: London, Macmillan & Co.,

Ltd., v. 1, p. 173–243.

Springer, F., 1920, The Crinoidea Flexibilia: Smithsonian Institution Publication no.

2501, 486 p.

Sprinkle, J., 1982, Hybocrinus, in Sprinkle, J., ed., Echinoderm faunas from the Bromide

Formation (Middle Ordovician) of Oklahoma: The University of Kansas Paleontological

Contributions, Monograph 1, p. 119–128.

Sprinkle, J., 1973, Morphology and evolution of blastozoan echinoderms: Museum of

Comparative Zoology Special Publication, Harvard University, 283 p.

Sprinkle, J., 1976, Classification and phylogeny of “pelmatozoan” echinoderms:

Systematic Zoology, v. 25, p. 83–91.

386

Sprinkle, J., 1982, Cylindrical and globular rhombiferans, in Sprinkle, J., ed., Echinoderm faunas

from the Bromide Formation (Middle Ordovician) of Oklahoma: The University of

Kansas Paleontological Contributions, Monograph 1, p. 231–273.

Sprinkle, J., and Guensburg, T.E., 1997, Early radiation of echinoderms. Pp. 205-224 in

J. A. Waters and C. G. Maples (eds.), Geobiology of Echinoderms,

Paleontological Society Papers 3.

Sprinkle, J., and Kolata, D. R., 1982, "Rhomb-bearing" camerate, in Sprinkle, J., ed.,

Echinoderm faunas from the Bromide Formation (Middle Ordovician) of

Oklahoma: University of Kansas Paleontological Contributions, Monograph 1, p.

206‒211.

Sprinkle, J., and Moore, 1978, Hybocrinida, in Moore, R. C. and Teichert, C., eds:

Treatise on Invertebrate Paleontology, Pt. T, Echinodermata 2: Geological Society

of America and University of Kansas Press, Kansas, p. T564–T574.

Sprinkle, J., and Wahlman, G.P., 1994, New echinoderms from the Early Ordovician of

West Texas. Journal of Paleontology, v. 68, p. 324-338.

Stadler, T., 2010, Sampling-through-time in birth-death trees: Journal of Theoretical

Biology, vol. 267, p. 396‒404.

Stadler, T. and Yang, Z., 2013, Dating phylogenies with sequentially sampled tips:

Systematic Biology, v. 62, p. 674‒688.

Stadler, T., Kouyos, R., Wyl, V. von, Yerly, S., Böni, J., Bürgisser, P., Klimkait, T., Joos,

B., Rieder, P., Xie, D., Günthard, H. F., Drummond, A. J., Bonhoeffer, S., the Swiss

387

HIV Cohort Study, 2012, Estimating the basic reproductive number from viral

sequence data: Molecular Biology and Evolution, v. 29, p. 347‒357.

Strimple, H.L., 1948, Notes on Phanocrinus from the Fayetteville Formation of

Northeastern Oklahoma: Journal of Paleontology, v. 22, p. 490-493.

Strimple, H.L., 1978, Evolutionary trends among Poteriocrinina. Pp. T298-T301 in R. C.

Moore and Teichert, C. (eds.), Treatise on Invertebrate Paleontology. Part T,

Echinodermata 2. Geological Society of America and University of Kansas Press,

Lawrence.

Stukalina, G.A., 1980, New species of quadrilaterial from the Ordovician of Kazakhstan,

Urals, and eastern European platform, in Stukalina, G.A., ed., New species of

ancient plants and invertebrates of the USSR, 5: Akademiia Natik SSSR,

Paleontologischeskii Institut. Moscow. p. 88-95.

Summers, M.M., Messing, C.G., and Rouse, G.W., 2014, Phylogeny of Comatulidae

(Echinodermata: Crinoidea: Comatulida): a new classification and an assessment

of morphological characters for crinoid taxonomy: Molecular Phylogenetics and

Evolution, v. 80, p. 319–339.

Sumrall, C.D., 1997, The role of fossils in phylogenetic reconstructions of the

Echinodermata: Paleontological Society Papers, v. 3, p. 267–288.

Sumrall, C.D., 2008, The origin of Lovéns Law in glyptocystitoid rhombiferans and its

bearing on plate homology and heterochronic evolution of the hemicosmitoid

peristomial border, in Ausich, W.I., and Webster, G.D., eds., Echinoderm

Paleobiology: Indiana University Press, Bloomington, p. 28–41. 388

Sumrall, C.D., 2010, A model for elemental homology for the peristome and ambulacra

in blastozoan echinoderms, in Harris, L.G., Bottger, S.A., Walker, C.W., and

Lesser, M.P., eds., Echinoderms: Durham, CRC Press, New York, p. 269–276.

Sumrall, C.D., 2014, Echinoderm phylogeny – the path forward: Geological Society of

America Abstracts with Programs, v. 46, p. 78.

Sumrall, C.D., and Waters, J.A., 2012, Universal Elemental Homology in

glyptocystitoids, hemicosmitoids, coronoids, and blastoids: steps toward

echinoderm phylogenetic reconstruction in derived blastozoans: Journal of

Paleontology, v, 68, p. 956–972.

Sumrall, C.D., and Wray, G.A., 2007, Ontogeny in the fossil record: diversification of

body plans and the evolution of "abberant" symmetry in Paleozoic echinoderms:

Paleobiology, v. 33, p. 149–163.

Swofford, D.L., Olsen, G.J., Waddell, P.J., and Hillis, D.M., 1996, Phylogenetic

inference, in Hillis, D.M., Moritz, C., and Mable, B.K., ed., Molecular

Systematics (second edition), Massachusetts, Sinauer, p. 407‒514.

Swofford, D., 2002, PAUP*: Phylogenetic Analysis Using Parsimony (*and Other

Methods): Version 4. Massachusetts, Sinauer Associates.

Telford, M.J., Low, C.J., Cameron, C.B., Ortega-Martinez, O., Aronowicz, J., Oliveri, P.,

and Copley, R.R., 2014, Phylogenetic analysis of echinoderm class relationships

supports Asterozoa: Proceedings of the Royal Society B, v. 281, 20140479.

389

Thomson, W. C. 1865, On the embryogeny of Antedon rosaceus (Lmk) (Comatula

rosacea of Lamark). Philosophical Transactions of the Royal Society of London,

v. 155, p. 513-545.

Ubaghs, G., 1953, Classe des crinoides. in J. Piveteau (ed.). Traité de Paléontologie. 3:

658-773.

Ubaghs, G., 1963, Rhopalocystis destombesi n. g., n. sp. Eocrinoide de l’Ordovicien

inferieur (Tremadocien superieur) du Sud marocain: Notes du service geologic du

Maroc, v. 23, p. 25–44.

Ubaghs, G., 1969, Aethocrinus moorei Ubaghs N. Gen., N. sp., le Plus ancien Crinoide

dicyclique Connu: University of Kansas Paleontological Contributions, Paper 38, p. 1‒

25.

Ubaghs, G., 1978, Origin of crinoids, in Moore, R. C. and Teichert, C., eds, Treatise on

Invertebrate Paleontology, Pt. T, Echinodermata, v. 2 (2): Geological Society of

America and University of Kansas Press, Lawrence, p. T275–T281.

Ulrich, E.O., 1925, The lead, zinc, and fluorspar deposits of Western Kentucky; Chapter

2, Stratigraphic geology: U. S. Geological Survey, Professional Paper 36, p. 22‒

71. van Valen, L. M., 1982, Homology and causes. Journal of Morphology, v. 173, p. 305-

312.

Vermeij, G.J., 1987, Evolution and escalation: an ecological history of life. Princeton

University Press.

390 von Baer, K. E., 1828, Uber Entwickelungsgeschichte der Thiere: Beobachtung und

Reflektion, Bornträger.

Vrba, E.S., 1987, Ecology in relation to speciation rates: some case histories of Miocene-

Recent mammal clades. Evolutionary Ecology, v. 1, p.283-300.

Wachsmuth, C., and Springer, F., 1880–1886, Revision of the Palaeocrinoidea.

Proceedings of the Academy of Natural Sciences of Philadelphia Pt. I. The

families Ichthyocrinidae and Cyathocrinidae (1880), p. 226‒378 (separate repaged

p. 1–153). Pt. II. Family Sphaeroidocrinidae, with the sub-families Platycrinidae,

Rhodocrinidae, and Actinocrinidae (1881), p. 177‒411 (separate repaged, p. 1‒

237). Pt. III, Sec. 1. Discussion of the classification and relations of the brachiate

crinoids, and conclusion of the generic descriptions (1885), p. 225‒364 (separate

repaged, 1‒138). Pt. III, Sec. 2. Discussion of the classification and relations of

the brachiate crinoids, and conclusion of the generic descriptions (1886), p. 64‒

226 (separate repaged to continue with section 1, 139‒302).

Wachsmuth, C., and Springer, F., 1891, The perisomic plates of the crinoids: Proceedings

of the Academy of Natural Sciences of Philadelphia, v. 41, p. 345–39.

Wachsmuth, C., and Springer, F., 1897, The North American Crinoidea Camerata:

Harvard College Museum of Comparative Zoology Memoir, v. 20 and 21, 897 p.

Wagner, G.P., 2006. Homologues, natural kinds and the evolution of modularity:

American Zoologist, v. 36, p. 36-43.

Wagner, G.P., 2007. The developmental genetics of homology: Nature Reviews Genetics,

v. 8, p. 473-479. 391

Wagner, P.J., 1995,Testing evolutionary constraint hypotheses with Early Paleozoic

gastropods: Paleobiology, v. 21, p. 248-272.

Wagner, P.J., 1997, Patterns of morphologic diversification among the Rostroconchia:

Paleobiology, v. 23, p. 115-150.

Wagner, PJ. 1998, A likelihood approach for evaluating estimates of phylogenetic

relationships among fossil taxa: Paleobiology, v. 24, p. 430‒49.

Wagner, P.J., 1999, The utility of fossil data in phylogenetic analyses: a likelihood

example using Ordovician-Silurian species of the Lophospiridae (Gastropoda:

Murchisoniina): American Malacological Bulletin, v. 15, p. 1‒32.

Wagner, P.J., 2000, Phylogenetic analyses and the fossil record: tests and inferences,

hypotheses, and models: Paleobiology, v. 26, p. 341‒371.

Wagner, P.J., 2000, The quality of the fossil record and the accuracy of phylogenetic

inferences about sampling and diversity: Systematic Biology, v. 49, p. 65‒86.

Wagner, P.J., 2000, Exhaustion of morphologic character states among fossil taxa:

Evolution, v. 54, 365‒386.

Wagner, P.J., 2010. Paleontological perspectives on morphological evolution. in

Evolution since Darwin: the first, 150 years, pp.451-478.

Wagner, P. J., 2012, Modelling rate distributions using character compatibility:

implications for morphological evolution among fossil invertebrates: Biology Letters,

v. 8, p. 143—146.

392

Wagner, P.J., 2012, Modelling rate distributions using character compatibility:

implications for morphological evolution among fossil invertebrates: Biology

Letters, v. 8, p. 143‒146.

Wagner, P.J. and Marcot, J.D., 2010, Probabilistic phylogenetic inference in the fossil

record: current and future applications, in Alroy, J. and Hunt, G., eds.,

Quantitative Methods in Paleobiology. The Paleontological Society Papers, v. 16,

p. 189‒211

Wagner, P.J. and Marcot, J.D., 2013, Modelling distributions of fossil sampling rates

over time, space and taxa: assessment and implications for macroevolutionary

studies: Methods in Ecology and Evolution, v. 4, p.703‒713.

Walcott, C.D., 1884. Descriptions of new species of fossils from the Trenton Group of

New York: New York State Museum of Natural History, Annual Report, v. 35, p.

207‒214.

Walker, L. J., Wilkinson, B.H., and Ivany, L.C., 2002, Continental drift and Phanerozoic

carbonate accumulation in shallow-shelf and deep-marine settings: The Journal of

Geology, v. 110, p.75-87.

Wanner, J., 1916, Dei Permischen echinodermen von Timor, I. Teil. Paläeontology von

Timor, v. 11, p. 1-329.

Webster, G.D., 2003, Bibliography and index of Paleozoic crinoids, coronates, and

hemistreptocrinoids: Geological Society of America Special Paper 363.

Webster, G.D., 2013, Bibliography and index of Paleozoic crinoids, coronates, and

hemistreptocrinoids, 1758–2012: [http://crinoids.azurewebsites.net/]

393

Webster, G.D., and Jell, P.A., 1999, New Permian crinoids from Australia: Memoirs of

the Queensland Museum, v. 33, p. 349–359.

Webster, G.D., and Lane, N.G.,1967, Additional Permian crinoids from southern Nevada:

University of Kansas Paleontological Contributions, v. 27, p. 1-32.

Webster, G.D., and Lane, N.G., 2007, New Permian crinoids from the Battleship Wash patch

reef in southern Nevada: Journal of Paleontology, v. 81, p. 951–965.

Webster, G.D. and Maples, C.G., 2006, Cladid crinoid (Echinodermata) anal conditions:

a terminology problem and proposed solution: Palaeontology, v. 49, p. 187–212.

Webster, G.D. and Maples, C.G., 2008, Cladid crinoid radial facets, brachials, and arm

appendages: a terminology solution for studies of lineage, classification, and

paleoenvironment, in Ausich, W.I., and Webster, G.D., eds., Echinoderm

Paleobiology. Indiana University Press, Bloomington, p.197–226.

Webster, G.D., Maples, C.G., Mawson, R., and Dastanpour, M., 2003, A cladid-dominated Early

Mississippian crinoid and conodont fauna from Kerman Province, Iran and revision of

the glossocrinids and rhenocrinids: Journal of Paleontology, Memoir 60, v. 77, suppl. to

no. 3, 35 p.

Weller, S. and Davidson, A.D., 1896. Petalocrinus mirabilis (n. sp.) and a new American

fauna: Journal of Geology, v. 4, p. 166‒173.

Wetherby, A.G., 1880, Remarks on the Trenton Limestone of Kentucky, with

descriptions of new fossils from that formation and the Kaskaskia (Chester)

394

Group, Sub-carboniferous: Journal of the Cincinnati Society of Natural History, v.

3, p. 144–160.

Wheeler, W.C., 2012, Systematics: a course of lectures: Chichester, John Wiley and

Sons. 426 p.

Wheeler, W.C., and Pickett, K.M., 2007, Topology-Bayes versus Clade-Bayes in

phylogenetic analysis: Molecular Biology and Evolution, v. 25, p. 447‒453.

White, C. A., 1881, Fossils of the Indiana rocks: Indiana Department of Statistics and

Geology, Annual Report, v. 2, p. 471–522.

Wiley, E.O.,1975, Karl R. Popper, systematics and classification: a reply to Walter Bock

and other evolutionary taxonomists: Systematic Zoology, v. 24, p. 33-243.

Wiley, E.O., and Lieberman, B.S., 2011, Phylogenetics: theory and practice of

phylogenetic systematics: John Wiley & Sons, 406 p.

Wright, A.M., and Hillis, D.M., 2014, Bayesian analysis using a simple likelihood model

outperforms parsimony for estimation of phylogeny from discrete morphological

data: PloS One, v. 9, e109210.

Wright, D.F., 2015, Fossils, homology, and “Phylogenetic Paleo-ontogeny”: a

reassessment of primary posterior plate homologies among fossil and living

crinoids with insight from developmental biology: Paleobiology, v. 41, p. 570–

591.

Wright, D.F., 2015, Testing the taxonomic structure of Paleozoic pan-cladid crinoids: a

statistical approach using the fossilized birth-death process and Bayesian

395

phylogenetic inference: Geological Society of America Abstracts with Programs,

v. 7, p. 854.

Wright, D.F., in press, Bayesian estimation of fossil phylogenies and the evolution of

early to middle Paleozoic crinoids (Echinodermata): Journal of Paleontology.

Wright, D.F., and Ausich, 2015, W.I., From the stem to the crown: phylogeny and

diversification of pan-cladid crinoids, in Zamora, S. and Rábano, I., eds., Progress

in Echinoderm Paleobiology: Cuadernos del museo Geominero, 19, Instituto

Geológico y Minero de España, p. 199–202.

Wright, D. F. and Stigall, A. L., 2013, Phylogenetic revision of the Late Ordovician

orthid brachiopod genera Plaesiomys and Hebertella from Laurentia: Journal of

Paleontology, v. 87, p. 1107‒1128.

Wright, D.F., Ausich, W.I., Cole, S.R., Rhenberg, E.C., and Peter, M.E., in press,

Phylogenetic taxonomy and classification of the Crinoidea (Echinodermata):

Journal of Paleontology.

Wright, J., 1920, On Carboniferous crinoids from Fife; with notes on some localities, and

provisional lists of specie:. Transactions of the Geological Society of Glasgow, v.

16, p. 363-392.

Wright, J., 1926, Notes on the anal plates of Epachycrinus calyx and Zeacrinus konincki.:

Geological Magazine, v. 64, p. 352-373.

Wright, J., 1927, Some variations in Ulocrinus and Hydreionocrinus. Geological

Magazine 71:241-268.

396

Xie, W., Lewis, P.O., Fan, Y., Kuo, L., and Chen, M.–H., 2010, Improving marginal

likelihood estimation for Bayesian phylogenetic model selection: Systematic Biology,

v. 60, p. 150‒160.

Yang, Z., 2014, Molecular Evolution: a Statistical Approach: Oxford University Press,

Oxford, 492 p.

Zamora, S., 2013, Morphology and phylogenetic interpretation of a new Cambrian

edrioasteroid (Echinodermata) from Spain: Palaeontology, v. 56, p. 421–431.

Zamora, S., and Rahman, I.A., 2014, Deciphering the early evolution of echinoderms

with Cambrian fossils: Palaeontology, v. 57, p. 1105–1119.

Zamora, S., and Smith, A.B., 2011, Cambrian stalked echinoderms show unexpected

plasticity of arm construction: Proceedings of the Royal Society of London B:

Biological Sciences, rspb20110777.

Zamora, S., I. A. Rahman, and A. B. Smith. 2012, Plated Cambrian bilaterians reveal the

earliest stages of echinoderm evolution: PLoS One e38296.

doi:10.1371/journal.pone.0038296

Zamora, S., Sumrall, C.D., and Vizcaino, D., 2013, Morphology and ontogeny of the

Cambrian edrioasteroid echinoderm Cambraster cannati from western

Gondwana: Acta Palaeontologica Polonica, v. 58, p. 545-559.

Zamora, S., Sumrall, C.D., Zhu, X., Lefebvre, B., 2016, A new stemmed echinoderm

from the Furongian of China and the origin of Glyptocystitida (Blastozoa,

Echinodermata): Geological Magazine, doi:10.1017/S001675681600011X.

397

Zhang, C., Stadler, T., Klopfstein, S., Heath, T.A. and Ronquist, F., 2016, Total-Evidence

Dating under the Fossilized Birth–Death Process: Systematic Biology, v. 65, p.,

228‒249.

Zittel, K.A., von, 1895, Grundzüge der Palaeontologie (Palaeozoologie), 1st edition: R.

Oldenbourg, München, 971 p.

398

Appendix A: Character list for phylogenetic analysis in Chapter 5.

1. Calyx shape ratio (height /width): very high, >1.5 (0); high, 1.5-1.0 (1); intermediate

1.0-0.5 (3); low, 0.5-0.25 (3); very low, < 0.25 (4).

2. Calyx profile: straight sided, cone shape (0); rounded base with widest point at the

summit of the calyx (1); rounded base with widest point below the summit of the

cup (2).

3. Calyx shape in transverse section: round (0); subpentagonal (1).

4. Calyx radial plate thickness: thin, <25% height or width of plate (0); thick, >25%

height or width of plate (1).

5. Calyx plate sculpturing: absent (0); present (1).

6. Type of calyx suturing: non-stellate ridges (0); stellate ridges (1); spinose (2); nodose

(3); granulose (4).

7. Outline shape of calyx base: upright (0); broad and flat (1); steeply concave (2).

8. Basal invagination of calyx: absent (0); present (1).

9. Attitude of the infrabasal circlet: all plates visible in side view (0); along base of calyx,

neither entirely in a basal concavity nor visible in side view (1); not visible in side

view (2).

10. Number of infrabasal plates: five (0); three (1); one (2).

399

11. Infrabasal plate dimensions: width > height (0); width ~ height (1); width < height

(2).

12. Cross sectional shape of the infrabasal plates: straight upwards/flared (0);

curved/downflared (1).

13. Attitude of the basal circlet: all plates visible in side view (0); along base of calyx,

neither entirely in a basal concavity nor visible in side view (1); not visible in side

view (2).

14. Basal plate dimensions: width > height (0); width ~ height (1); width < height (2).

15. Relative sizes of plates in the basal circlet: equal/subequal (0); unequal (1).

16. Cross sectional shape of the basal plates: straight (0); curved (1); bulbous/inflated (2).

17. Interruption in the radial circlet CD interray: absent (0); present (1).

18. Radial plate dimensions: width > height (0); width ~ height (1); width < height (2).

19. Attitude of the radial circlet: (0); along base of calyx, neither entirely in a basal

concavity nor visible in side view (1).

20. Cross sectional shape of the radial plates: straight (0); curved (1); bulbous/inflated

(2).

21. Largest plate in the calyx: below the radials circlet (0); in the radial circlet (1).

22. Compound/ bi-radial: absent (0); present (1).

23. C-ray radial plate: similar size to other radial plates (0); smaller than other radial

plates (1).

24. Radial facet width (sensu Webster and Maples 2008): angustary (0); peneplenary (1);

plenary (2).

400

25. Contact type when radial facets extend across the radial plate (Webster 2007): plenary

(0); explenary (1); inplenary (2).

26. Radial facet dimensions: width > height (0); width ~ height (1); width < height (2).

27. Adaxial notch on radial facets: absent (0); present (1).

28. Orientation of the radial facets: planate (0); declivate (1); sursumate (2).

29. Radial facet type (sensu Webster and Maples, 2008): unifascial (0); bifascial (1);

multifascial ‘muscular’ (2); trifascial (3).

30. Transverse ridge on radial facet: absent (0); present (1).

31. Thick entoneural canal on facet: absent (0); present (1).

32. Ligament pit on facet: absent (0); present (1).

33. Crenulated surface proximal to aboral ligamentary fossae: absent (0); present (1).

34. Number of posterior plates partially or completely within the cup: three (0); ≤ three

(1).

35. Radianal plate: absent (0); present (1).

36. Visibility of the radianal plate: visible on the exterior (0); cryptic (1).

37. Shape of the radianal plate (upper plate when subdivided): pentagonal (0); tetragonal

(1); hexagonal (2).

38. Proximal position of the radianal plate: full width beneath the C-ray radial (0); to the

left and below the C radial plate (1); within the radial circlet above the CD basal

plate (2).

39. Circlet in most proximal position with the radianal plate: basals (0); infrabasals (1);

radials (2).

401

40. Anal X: absent (0); present (1).

41. Shape of the anal X plate: pentagonal (0); tetragonal (1); hexagonal (2);

42. Position of the anal X plate within the calyx: entirely in the cup (0); partially in the

cup (1); entirely above the cup (2).

43. Position of the anal X plate relative to the plates below: above and to the left of the

radianal and in contact with the CD basal (0); above and to the left of the radianal

and in contact with the D radial but not the CD basal (1); directly above the

radianal and not in proximal contact with the CD basal or the D radial (2).

44. Plates right-lateral to the anal X: in contact with the C-ray radial (0); in contact with

other anal plates (1).

45. Right tube plate: absent (0); present (1).

46. Shape of the right tube plate: pentagonal (0); tetragonal (1); hexagonal (2).

47. Position of the right tube plate within the calyx: entirely in the cup (0); partially in the

cup (1); entirely above the cup (2).

48. Position of the right tube plate relative to the plates below: above and to the right of

the anal X and in contact with the radianal (0); directly above the anal X and not

in contact with the radianal (1); above and to the right of the anal X and not in

contact with the radianal (2).

49. Plates left-lateral to the right tube plate: anal X and additional anal plates (0); anal X

only (1); D-ray radial plate (2).

50. Anal sac: absent (0); present (1).

402

51. Height of the anal opening above the cup: ≤ half the height of the arms (0); > half the

height of the arms and ≤ maximum arm height (1); > than maximum arm height

(2).

52. Anus position relative to the crown: near the summit (0); below the summit (1).

53. Anal sac plating: vertical columns, regular plating (0); irregular plating (1).

54. Medial column supporting the anal sac: absent (0); present (1).

55. Anal sac plate sculpturing: absent (0); present (1).

56. Anal sac sculpturing pattern: radiating ridges (0); vertical grooves and ridges (1).

57. Anal sac plate cross section: flat (0); convex (1); spinose (2); plicated (3).

58. Shape of the anal sac: cylindrical (0); tapers distally (1).

59. Pore structures on the anal sac: absent (0); present (1).

60. Anal sac spines: absent (0); present (1).

60. Number of arm openings originating from the calyx: five (0); three (1); one (2).

61. Arrarngement of spines along the anal sac summit: single spine (0); multiple spines

(1); roof-forming (e.g., “umbrella” of spines).

62. Arm openings into the calyx: five (0); three (1); one (2).

63. Proximal free arm projection: outward and upward (0); directed vertically (1).

64. Fixed lower brachial plates: absent (0); present (1).

65. Primaxil branching on B-E rays: absent (0); present (1).

66. Branching in the A ray: same as B-E rays (0); different from B-E rays (1).

67. Highest position of an axillary on B-E ray primibrachials: IPr1 (0); IPr2 (1); ≥ IPr3

(2).

403

68. Bifurcation pattern on B-E ray primaxils: atomous (0); similar spacing (1); irregular

spacing (2); plates fuse to form a mesh/fan shape (3).

69. Spacing between axillaries at and above the second brachitaxes in B-E rays: regular

(0); irregular (1).

70. Secundaxil branching in the B-E rays: absent (0); present (1).

71. Branching above the secundaxil in B-E rays: absent (0); present (1).

72. Largest number of axillaries per ray: zero (0); one (1); two (2); ≥ 3 (3).

73. Outline pattern of bifurcation (Ubaghs, 1978, fig. 115): isotomous (0); heterotomous

(# 2 and 3 in Ubaghs, 1978, fig 115) (1); endotomous (2); exotomous (3).

74. Pinnulation: absent (0); present (1).

75. Branching complexity of minor arm appendages on brachials distributed along the

main ray axis: absent (0); ramulate (i.e., terminal appendages are separated by at

least one brachial, branched or unbranched) (1); pinnulate (i.e., unbranched

terminal appendages distributed along successive brachial plates) (2).

76. Mature (i.e., distal) brachial type: uniserial (0); biserial (1).

77. Transition from proximal uniserial to distal biseral brachials: absent (0); present (1).

78. Mature shape of uniserial brachials (sensu Webster 2007): rectilinear (0); weakly

cuneate (1); moderately cuneate (2); strongly cuneate (3).

79. Distal biserial brachial shape (sensu Webster 2007): wedge (0); round (1); flat chisel

(2).

80. Size of the primibrachials: smaller than radials (0); subequal to radials (1).

404

81. Number of axillaries in the A ray: same as other rays (0); first branching is higher

than other rays (1).

82. Shape of the first primibrachial: tetragonal (0); pentagonal (1); triangular (2).

83. Dimensions of the first primibrachial: width > height (0); width ~ height (1); width <

height (2).

84. Primaxil ornamentation: absent (0); present (1).

85. Maximum number of secundibrachials: zero (0); one (1); two (2); ≥ 3 (3).

86. Secundaxil spines: absent (0); present (1).

87. Syzygial sutures on brachials: absent (0); present (1).

88. Column: absent (0); present (1).

89. Proximal shape of the stalk: circular (0); pentagonal (1).

90. Lumen shape in proxistele columnals: circular (0); pentalobate (1); pentastellate (2).

91. Branching appendages on proximal stem: absent (0); present (1).

92. Stem appendages: absent (0); cirri (1); highly cirriferous (2).

405