+ All Categories
Home > Documents > Ciliate Biodiversity and Phylogenetic Reconstruction ...

Ciliate Biodiversity and Phylogenetic Reconstruction ...

Date post: 03-Feb-2023
Category:
Upload: khangminh22
View: 0 times
Download: 0 times
Share this document with a friend
237
University of Massachusetts Amherst University of Massachusetts Amherst ScholarWorks@UMass Amherst ScholarWorks@UMass Amherst Open Access Dissertations 9-2009 Ciliate Biodiversity and Phylogenetic Reconstruction Assessed by Ciliate Biodiversity and Phylogenetic Reconstruction Assessed by Multiple Molecular Markers Multiple Molecular Markers Micah Dunthorn University of Massachusetts Amherst Follow this and additional works at: https://scholarworks.umass.edu/open_access_dissertations Part of the Life Sciences Commons Recommended Citation Recommended Citation Dunthorn, Micah, "Ciliate Biodiversity and Phylogenetic Reconstruction Assessed by Multiple Molecular Markers" (2009). Open Access Dissertations. 95. https://doi.org/10.7275/fyvd-rr19 https://scholarworks.umass.edu/open_access_dissertations/95 This Open Access Dissertation is brought to you for free and open access by ScholarWorks@UMass Amherst. It has been accepted for inclusion in Open Access Dissertations by an authorized administrator of ScholarWorks@UMass Amherst. For more information, please contact [email protected].
Transcript

University of Massachusetts Amherst University of Massachusetts Amherst

ScholarWorks@UMass Amherst ScholarWorks@UMass Amherst

Open Access Dissertations

9-2009

Ciliate Biodiversity and Phylogenetic Reconstruction Assessed by Ciliate Biodiversity and Phylogenetic Reconstruction Assessed by

Multiple Molecular Markers Multiple Molecular Markers

Micah Dunthorn University of Massachusetts Amherst

Follow this and additional works at: https://scholarworks.umass.edu/open_access_dissertations

Part of the Life Sciences Commons

Recommended Citation Recommended Citation Dunthorn, Micah, "Ciliate Biodiversity and Phylogenetic Reconstruction Assessed by Multiple Molecular Markers" (2009). Open Access Dissertations. 95. https://doi.org/10.7275/fyvd-rr19 https://scholarworks.umass.edu/open_access_dissertations/95

This Open Access Dissertation is brought to you for free and open access by ScholarWorks@UMass Amherst. It has been accepted for inclusion in Open Access Dissertations by an authorized administrator of ScholarWorks@UMass Amherst. For more information, please contact [email protected].

CILIATE BIODIVERSITY AND PHYLOGENETIC RECONSTRUCTION

ASSESSED BY MULTIPLE MOLECULAR MARKERS

A Dissertation Presented

by

MICAH DUNTHORN

Submitted to the Graduate School of the University of Massachusetts Amherst in partial fulfillment

of the requirements for the degree of

Doctor of Philosophy

September 2009

Organismic and Evolutionary Biology

© Copyright by Micah Dunthorn 2009

All Rights Reserved

CILIATE BIODIVERSITY AND PHYLOGENETIC RECONSTRUCTION

ASSESSED BY MULTIPLE MOLECULAR MARKERS

A Dissertation Presented

By

MICAH DUNTHORN

Approved as to style and content by:

_______________________________________ Laura A. Katz, chair _______________________________________ Robert L. Dorit, member _______________________________________ George B. McManus, member _______________________________________ Benjamin B. Normark, member

_______________________________________ Joseph S. Elkinton, Graduate Program Director

Program in Organismic & Evolutionary Biology

To Frank Buchholz

13 February 1952—15 July 1991

v

ACKNOWLEDGEMENTS

Much thanks goes to my dissertation committee. Laura A. Katz, my advisor,

provided her support throughout this dissertation. I thank Laura for allowing me to work

in her lab and for teaching me many molecular methods. Rob Dorit was able to point out

the bigger picture and context. George McManus was great to talk about ciliate evolution

and the intricacies of ciliate taxonomy. And Ben Normark was always there to push me

towards our current theoretical gaps.

Much thanks also goes to Professor Dr. Wilhelm Foissner, who isolated and

identified most of the ciliates that were used in this dissertation, and is a co-author on

many of the chapters. I am not sure how he was able to put up with me while I visited his

laboratory. Chapter 2 was a collaboration with Thorsten Stoeck, Marion Eppinger, M. V.

Julian Schwarz, Michael Schweikert, and Jens Boenigk. Denis Lynn collaborated on the

PhyloCode classification in Chapter 7. Phil Cantino, the editor for the Companion

Volume to which the PhyloCode definitions will be published, was extremely helpful and

patient in his edits.

This dissertation would also not be possible without the many folks in the Katz

lab. I will miss Mary Doherty’s jokes and her insights into ciliate biogeography. The

following directly or indirectly also made an impact: Becky Zufall, Sofia Annis, Erica

Barbero, Jennifer DeBarardinis, Jessica Grant, Meaghan Hall, Hannah Jaris, Allie

Kovner, Dan Lahr, Christina Lyman, Bob Merritt, Laura Parfrey, Maiko Tamura, Yonas

Tekle, Sam Torquato, and Laura Vann. Help also came from members of the Organismic

and Evolutionary Biology program: Penny Jaques, Norman Johnson, David Lahti, and

vi

Lynn Margulis. Caffeine and delicious vegan snacks were made possible by Melissa

Krueger at the Elbow Room Café.

Partial funding for this dissertation and for presenting at conferences came from:

International Travel Award for Young Scholars, Society for the Study of Evolution;

Holz-Connor Travel Grant, International Society of Protozoologists; Mini-PEET Award,

Society of Systematic Biologists; Graduate Research Award, Society of Systematic

Biologists; Travel Grant, 56th Annual Meeting, Society of Protozoologists; Graduate

Fellowship, Massachusetts Space Grant Consortium. Funding was also provided by Laura

A. Katz through her U.S. National Science Foundation grant #636498.

I would also like to thank my mother and father—Lani and Tom—for their

patience and help throughout my time in graduate school. Finally, I want to thank my

son—(captain) Aureliano—for thinking that working in the lab consists of watching

movies in my office, since that is what he did the so many times I dragged him to the lab

so I can perform just more PCR and cloning on a Saturday afternoon.

vii

ABSTRACT

CILIATE BIODIVERSITY AND PHYLOGENETIC RECONSTRUCTION

ASSESSED BY MULTIPLE MOLECULAR MARKERS

SEPTEMBER 2009

MICAH DUNTHORN, B.A., GEORGE MASON UNIVERSITY

M.A., UNIVERSITY OF MISSOURI ST. LOUIS

PH.D., UNIVERSITY OF MASSACHUSETTS AMHERST

Directed by: Laura A. Katz

Ciliates provide a powerful system within microbial eukaryotes in which

molecular genealogies can be compared to detailed morphological taxonomies. Two

groups with such detailed taxonomies are the Colpodea and the Halteriidae. There are

about 200 described Colpodea species that are found primarily in terrestrial habitats. In

Chapters 1 and 2, taxon sampling is increased to include exemplars from all major

subclades using nuclear small subunit rDNA (nSSU-rDNA) sequencing. Much of the

morphological taxonomy is supported, but extensive non-monophyly is found

throughout. The conflict between some nodes of the nSSU-rDNA genealogy and

morphology-based taxonomy suggests the need for additional molecular marker. In

Chapter 3, character sampling is increased using mitochondrial small subunit rDNA

(mtSSU-rDNA) sequencing. The nSSU-rDNA and mtSSU-rDNA topologies for the

Colpodea are largely congruent for well-supported nodes, suggesting that nSSU-rDNA

work in other ciliate clades will be supported by mtSSU-rDNA as well. Chapter 4

compares the underlying genetic variation within two closely related species in the

viii

Halteriidae with increased taxon and molecular sampling using nSSU-rDNA and

internally-transcribed spacer (ITS) region sequencing. The morphospecies Halteria

grandinella shows extensive genetic variation that is consistent with either a large

effective population size or the existence of multiple cryptic species. This pattern

contrasts with the minimal of genetic variation in the morphospecies Meseres corlissi.

Chapter 5 discusses the congruence and incongruence among morphological and

molecular data in ciliates. Most of the incongruence occurs where there is little statistical

support for the molecules, or where molecular data is consistent with alternative

morphological hypotheses. Chapter 6 reviews the data for sex, or lack thereof, in the

Colpodea, a potentially ancient asexual group where sex was regained in a derived

species. In Chapter 7, four ciliate clades are redefined using the PhyloCode.

ix

CONTENTS Page

ACKNOWLEDGEMENTS……………………………………………………………… v ABSTRACT…………………………………………………………………………..…vii LIST OF TABLES………………………………………………………………………xiv LIST OF FIGURES……………………………………………………………………...xv CHAPTERS 1. MOLECULAR PHYLOGENETIC ANALYSIS OF CLASS COLPODEA (PHYLUM CILIOPHORA) USING BROAD TAXON SAMPLING………………….……………..1 1.1. Abstract……………………………………………………………………….2

1.2. Introduction……………………….…………………………………………..3 1.3. Materials and methods…………………….……………………….................7 1.3.1. Taxon sampling and collection……………………………..………7 1.3.2. Identification………………………………………………………..7 1.3.3. DNA extraction, amplification, cloning and sequencing…………...7 1.3.4. Genealogical analyses………………………………………………9

1.4. Results……………………………………………………………………….10 1.4.1. Pairwise distances within collections……………………………...10 1.4.2. Deletion within one SSU rDNA copy in Bryometopus

pseudochilodon…………………………………………………..11 1.4.3. Intron in Cyrtolophosis mucicola…………………………………12 1.4.4. SSU rDNA genealogy of the Colpodea…………………………...12

1.5. Discussion…………………………………………………………………...16 1.5.1. Comparisons between morphology and molecules………………..16 1.5.1.1. The class…………………………………………………16 1.5.1.2. The orders……………………………………………….16 1.5.1.3. The genus Colpoda……………………………………...20 1.5.2. Open questions with some species designations………………..…21 1.5.3. Evidence for sex…………………………………………………...22 1.5.4. Group I intron in Cyrtolophosis mucicola………………………...23 1.5.5. Reconciling morphology and molecules in the Colpodea………...24

1.6. References………………………………………………………………...…25

x

2. PHYLOGENETIC PLACEMENT OF THE CYRTOLOPHOSIDIDAE STOKES, 1888 (CILIOPHORA; COLPODEA) AND NEOTYPIFICATION OF ARISTEROSTOMA MARINUM KAHL, 1931…………………………………………...40 2.1. Abstract……...………………………………………………………………41 2.2. Introduction………………………………………………………………….42 2.3. Methods……………………………………………………………………...43 2.3.1. Taxon sampling……………………………………………………43 2.3.2. Light and electron microscopy…………………………………….44 2.3.3. Terminology……………………………………………………….45 2.3.4. Autecology……………………………………………………...…45 2.3.5. Amplification and sequencing ……………………………………47 2.3.6. Genealogical analyses……………………………………………..48 2.3.7. Constrained analysis……………………………………………....49 2.3.8. Rate class analysis…………………………………………………49 2.4. Results……………………………………………………………………….49 2.4.1. Description of the neotype of Aristerostoma marinum

Kahl, 1931………………………………………………………..49 2.4.2. Electron microscopy………………………………………………51 2.4.3. Autecology………………………………………………………..52 2.4.4. Pairwise SSU rDNA sequence differences within the

Cyrtolophosididae………………………………………………..52 2.4.5. Genealogical analysis……………………………………………...53 2.4.6. Comparisons of hypotheses……………………………………….54 2.4.7. Long-branch attraction…………………………………………….54

2.5. Discussion…..……………………………………………………………….56

2.5.1. Neotypification and emended diagnosis of Aristerostoma marinum Kahl 1931………………………....…………………...55

2.5.2. Emended diagnosis………………………………………………..56 2.5.3. Occurrence and ecology……………………………………….…..57 2.5.4. Comparison with original descriptions and related species……….58 2.5.5. Neotype specimens…………………………………………..……61 2.5.6. Pairwise distance between Aristerostoma morphospecies………...61 2.5.7. Phylogenetic placement of the order Cyrtolophosidida…………...61 2.5.8. Reconciling morphology and molecules…………………………..62 2.5.9. Assignment of GenBank environmental SSU-rDNA

sequences to the Cyrtolophosididae…………………………...…64 2.6. Conclusions………………………………………………………………….65 2.7. References…………………………………………………………………...67

xi

3. EXPANDING CHARACTER SAMPLING IN CILIATE PHYLOGENETIC RECONSTRUCTION: MITOCHONDRIAL SSU-rDNA AS A MOLECULAR MARKER………………………………………………………………………………..86 3.1. Abstract……...………………………………………………………………87

3.2. Introduction………………………………………………………………….88 3.3. Methods……………………………………………………………………...90 3.3.1. Taxon sampling and terminology…………………………………90 3.3.2. DNA amplification and sequencing……………………………….91 3.3.3. Genealogical analyses……………………………………………..91 3.3.4. Data partitioning and congruence testing…………………………92

3.4. Results……………………………………………………………………….93

3.4.1. Characteristics of gene sequences…………………………………93 3.4.2. MtSSU-rDNA analyses……………………………………………94 3.4.3. Nuclear SSU-rDNA analyses and topology congruence………….96

3.4.4. Concatenated analyses…………………………………………….97 3.5. Discussion……………………………………………………………….…..98

3.5.1. Mitochondrial SSU-rDNA as a ciliate molecular marker…………98 3.5.2. Morphology vs. molecules in ciliates……………………………100 3.5.3. Systematic implications………………………………………….101

3.6. Conclusions………………………………………………………………...101 3.7. References………………………………………………………………….102

4. EXTENSIVE GENETIC DIVERSITY WITHIN HALTERIID CILIATES………...115 4.1. Abstract……...……………………………………………………………..116

4.2. Introduction………………………………………………………………...117 4.3. Methods…………………………………………………………………….118 4.3.1. Taxon sampling and identification……………………………….118 4.3.2. Amplification, cloning, and sequencing…………………………119 4.3.3. Genealogical analyses……………………………………………120

4.4. Results……………………………………………………………………...121

4.4.1. Intra-isolate pairwise distances…………………………………..121 4.4.2. Genealogies………………………………………………………122

xii

4.5. Discussion………………………………………………………………….124

4.4.1. Genetic variation underlying morphospecies…………………….124 4.6. References………………………………………………………………….126 5. RICHNESS OF MORPHOLOGICAL HYPOTHESES IN CILIATE SYSTEMATICS ALLOWS FOR DETAILED ASSESSMENT OF HOMOLOGY AND COMPARISONS WITH GENE TREES………………………………………...138 5.1. Abstract……...……………………………………………………………..139

5.2. Introduction………………………………………………………………...140 5.3. Strengths of morphology…………………………………………………...140 5.3.1. Species delimitations…………………………………………….140 5.3.2. Ciliate tree of life………………………………………………...141

5.4. Challenges to Morphological analyses…………………………………….142

5.4.1. Decline in number of trained taxonomists……………………….142 5.4.2. Number of characters…………………………………………….142 5.4.3. Homology assessment……………………………………………143 5.5. Concordance with molecular hypotheses …………………………………..144

5.6. Open Questions…………………………………………………………….145

5.6.1. Lack of morphological variation when there is genetic diversity..145 5.6.2. Lack of genetic diversity when there is morphological variation..146

5.7. References………………………………………………………………….148 6. ANCIENT ASEXUAL LINEAGE OF CILIATES THAT REGAINED SEX?..........157 6.1. Abstract……...……………………………………………………………..158

6.2. Introduction………………………………………………………………...159 6.3. Empirical evidence…………………………………………………………160

6.4. Are colpodeans ancient asexuals? …………………………………………162 6.5. Did colpodeans reverse the loss of sex? …………………………………..163 6.6. Conclusion…………………………………………………………………165 6.7. References………………………………………………………………….166

xiii

7. PHYLOCODE DEFINITIONS FOR FOUR CILIATE CLADES..…………………173 7.1. Abstract……...………………………………………………………….….174

7.2. Introduction………………………………………………………………...175 7.3. The PhyloCode……………………………………………………………..176 7.4. Why the PhyloCode should be applied to ciliate…………………………..177 7.5. Application of the PhyloCode to four ciliate clades……………………….179 7.5.1. Ciliophora………………………………………………………..180 7.5.2. Postciliodesmatophora…………………………………………..182 7.5.3. Intramacronucleata………………………………………………184 7.5.4. Colpodea…………………………………………………………186 7.6. References…………………………………………………………………188 BIBLIOGRAPHY………………………………………………………………………195

xiv

LIST OF TABLES

Figure Page

1.1 Taxon sampling within the Colpodea……………………………………………31 1.2 Pairwise distance between collections for species sampled more than once…….32 2.1 Taxon sampling and GenBank numbers used in this study……………………...71 2.2 Morphometric data for Aristertostoma marinum population Framvaren Fjord….72 2.3 Comparative morphology of described representatives of four genera in

The Cyrtolophosididae…………………………………………………………...73 3.1 Taxon sampling for mtSSU…………………..………………………………...108 4.1 Isolates newly collected for this study………………………………………….130 4.2 Newly sampled loci……………………………………………………………..131

xv

LIST OF FIGURES

Figure Page

1.1 Evolution among some morphological characters within the class Colpodea…...33 1.2 SSU rDNA genealogy of the class Colpodea and potential sister classes……….35 2.1 Light microscopy of living (a, b) and protargol impregnated (c) Aristerostoma

marinum cells....………………………………………………………………….74 2.2 Scanning and transmission electron microscopy of Aristerostoma marinum……76 2.3 Schematic drawing of Aristerostoma marinum (right lateral view) that combines

diagnostic features as revealed by live cell imaging, protargol staining and scanning electron microscopy.…………………………………………………...78

2.4 Ecophysiological tolerance limits of Aristerostoma marinum population

Framvaren Fjord………………….………………………….…………………...80 2.5 Schematic drawing of the idealized oral structures (paroral membranes and

adoral organelles) of the Cyrtolophosididae……………………………………..82 2.6 Most likely Bayesian SSU-rDNA genealogy of the class Colpodea…………….84 3.1 Mitochondrial SSU-rDNA genealogy of the Colpodea………………………...109 3.2 Concatenated nuclear and mitochondrial SSU-rDNA genealogy of the

Colpodea………………………………………………………………………..112 4.1 Concatenated SSU-rDNA/ITS genealogy of the halteriids……………….....…132 6.1 Phylogeny and distribution of sex within the Colpodea…….………………….171 Supplementary Figures 1.1 SSU-rDNA genealogy of the class Colpodea and potential outgroups with support from all nodes indicated…………………………………………………37 3.1 Nuclear SSU-rDNA genealogy of the Colpodea…………..…………………...113 4.1 SSU-rDNA genealogy of the halteriids………………………………………...134 4.2 ITS genealogy of the halteriids…………………………………………………136

xvi

Appendix 1.A GenBank accessions used in analyses for both previously sequences Colpodea taxa and outgroups…………………………………………………….39

1

CHAPTER 1

MOLECULAR PHYLOGENETIC ANALYSIS OF CLASS COLPODEA

(PHYLUM CILIOPHORA) USING BROAD TAXON SAMPLING

Micah Dunthorn,a Wilhelm Foissnerb and Laura A. Katza,c

aGraduate Program in Organismic and Evolutionary Biology, University of

Massachusetts Amherst, USA

bFB Organismische Biologie, Universität Salzburg, Austria

cDepartment of Biological Sciences, Smith College, USA

2

1.1. Abstract

The ciliate class Colpodea provides a powerful case in which a molecular� genealogy can

be compared to a detailed morphological taxonomy of a microbial �group. Previous

analyses of the class using the small-subunit rDNA are based on� sparse taxon sampling,

and are therefore of limited use in �comparisons with morphologically-based

classifications. Taxon sampling is increased here to �include all orders within the class,

and more species within previously sampled� orders and in the genus Colpoda. Results

indicate that the Colpodea may be �paraphyletic, although there is no support for deep

nodes. The orders �Bursariomorphida, �Grossglockneriida, and Sorogenida are

monophyletic. The orders Bryometopida,� Colpodida and Cyrtolophosidida, and the genus

Colpoda, are not monophyletic. �Although congruent in many aspects, the conflict

between some nodes on this single �genealogy and morphology-based taxonomy suggests

the need for additional �markers as well as a reassessment of the Colpodea taxonomy.

3

1.2. Introduction

Assessment of phylogeny based on morphological characters is limited in many

microbial eukaryotes. In most amoebae and flagellates morphology provides little

guidance and taxonomic resolution is not rich below the class and ordinal levels. In

contrast, phylum Ciliophora Doflein, 1901 is relatively morphologically rich and has a

well-described taxonomy (Lynn, 2003; Lynn and Small, 2002). Class Colpodea Small

and Lynn, 1981 provides a particularly good opportunity to compare the power of

morphology and molecular analyses in reconstructing the phylogeny of ciliates. The

Colpodea is monographed and contains a number of somatic and oral characteristics that

were used to establish an extensive classification (Foisser, 1993a). Because previous

molecular investigations of the class are based on sparse taxon sampling (Lasek-

Nesselquist and Katz, 2001; Lynn et al., 1999; Stechmann et al., 1998), molecular

support for the groups established by Foissner (1993a) remains to be evaluated.

The Colpodea is one of eleven ciliate classes (Adl et al., 2005; Lynn, 2003).

Although its position in the subphylum Intramacronucleata is established (Lynn, 2003),

well-supported evidence for the sister class of the Colpodea remains elusive. With current

taxonomic sampling, neither morphology nor molecules give convincing or consistent

arguments because of homoplasy, low bootstrap support, and problems from both

paralogy and rate heterogeneity in protein-coding genes (Lasek-Nesselquist and Katz,

2001; Lynn et al., 1999; Stechmann et al., 1998). The classes Nassophorea,

Oligohymenophorea, Plagiopylea, and Prostomatea are the likely sister-group candidates.

Historically, members of the class Colpodea were placed in disparate groups

based on oral structure differences (Foissner, 1993a; Lynn et al., 1999). With Lynn’s

4

(1976; 1981) structural conservatism hypothesis, somatic kinety (kinetosomes and

associated fibers) differences were found to be more a appropriate guide to the deep

divisions within the ciliates. The Colpodea were united because of their left kinetodesmal

fiber (LKm fiber) (Foissner, 1993a). This fiber extends posteriorly and to the left of the

posterior kinetosome of their somatic dikinetids. In contrast, Bardele (1981; 1989) argues

against the monophyly of the class because of differences in the presence or absence of

particles (ciliary plaques) in the membrane of the somatic cilia: they are present only in

one order in the Colpodea (Colpodida) and are absent in the rest of the class.

The Colpodea are a group of primarily terrestrial ciliates (Foissner, 1993a).

Besides the unique LKm fiber, the class Colpodea contains distinctive silverline patterns

of regular meshes: ‘colpodid’, with large, rectangular meshes; ‘platyophryid’, meshes

divided by median silverline between the kineties, or ‘kreyellid’, with minute irregular

meshes (Foissner, 1993a). Members of the Colpodea also have somatic stomatogenesis,

where parental oral structures are partially or completely reorganized before new oral

structures develop during cell division (Foissner, 1993a; Foissner, 1996). In general, a

single ‘germline’ micronucleus is close to the single ‘somatic’ macronucleus; in at least

some taxa in order Cyrtolophosidida the micronucleus and macronucleus share an outer

membrane of the nuclear envelope (Foissner, 1993a). Sex has only been demonstrated in

Bursaria truncatella and is unreported in the rest of the class (Foissner, 1993a; Raikov,

1982).

Foissner (1993) monographed about 170 species and established an extensive

higher-level classification for the Colpodea. Subsequently, new genera and species have

been described (Foissner, 1993b; Foissner, 1993c; Foissner, 1994; Foissner, 1995;

5

Foissner, 1999; Foissner, 2003; Foissner et al., 2002; Foissner et al., 2003). Foissner’s

(1993a) split the Colpodea into two subclasses: one with the order Bryometopida based

on a ‘kreyellid’ silverline pattern; with the rest of the orders in another subclass, based on

‘colpodid’ and ‘platyophryid’ silverline patterns. These silverline patterns were later

argued to be misleading, as SSU rDNA places the Bryometopida next to order

Bursariomorphida (Lynn et al., 1999). In large part there is agreement over Foissner’s

(1993a) orders and families among other classifications (e.g., Puytorac, 1994), except

order Grossglockneriida is lumped with order Colpodida in Lynn and Small (1997;

2002).

Based on morphological characters, Foissner (1993a) offers several hypotheses

for relationships among these Colpodea orders (Figure 1.1, Table 1.1). First, Colpodida

and Grossglockneriida are sister taxa since they share merotelokinetal stomatogenesis

(complete reorganization of parental oral structures), which is probably the derived

condition (Figure 1.1A, character 12). In contrast, the other orders have the possibly

plesiomorphic state of pleurotelokinetal stomatogenesis (partial reorganization of parental

oral structures) (Figure 1.1A, character 1); this hypothesis is supported in previous

molecular analyses (Lynn et al., 1999; Lasek-Nesselquist and Katz, 2001). Second,

Bursariomorphida, Colpodida and Grossglockneriidae form a clade because of the

possibly apomorphic equally-spaced rows of oral polykinetids (Figure 1.1A, characters

9), as opposed to the possibly plesiomorphic brick-shaped adoral organelles (Figure

1.1A, character 1); this hypothesis is not supported in previous molecular and

morphological analyses (Foissner and Kreutz, 1998; Lasek-Nesselquist and Katz, 2001;

Lynn et al., 1999). Third, Bryophryida, Bursariomorphida, Colpodida and

6

Grossglockneriida form a clade because of the possibly apomorphic deep vestibulum

(depression in the cell with oral structures) (Figure 1.1A, character 7), as opposed to the

possibly plesiomorphic flat vestibulum (Figure 1.1A, character 1).

Using limited taxon sampling with small-subunit rDNA (SSU rDNA), monophyly

of the class Colpodea is strongly supported by least-squares (LS) and neighbor-joining

(NJ), and weakly supported by maximum parsimony (MP) analyses in Lynn et al. (1999).

In contrast, with just one additional taxon sampled, Lasek-Nesselquist and Katz (2001)

find the class to be paraphyletic, with Nassophorea embedded within it—although

support is weak from NJ, MP, and maximum likelihood (ML).

Here we increase taxon sampling of the Colpodea using SSU rDNA sequences,

including morphospecies from all seven orders, multiple morphospecies within most

orders, and multiple morphospecies in the genus Colpoda. Our aim is to assess the

following hypotheses: (1) the class Colpodea is monophyletic, (2) orders within the class

Colpodea are monophyletic, and (3) the genus Colpoda is monophyletic. We will also

discuss other features uncovered during characterization of SSU rDNA sequences: two

distinct copies of SSU rDNA in one taxon, a group I intron in another, and evidence for

sex in the Colpodea taxa sampled here. Furthermore, we discuss alternative hypotheses of

morphological evolution based on the SSU rDNA topology. Results from these analyses

will further development of a predictive, tree-based framework for the taxonomy of the

Colpodea.

7

1.3. Materials and methods

1.3.1. Taxon sampling and collection

To reconstruct an SSU rDNA genealogy of the Colpodea 27 collections

representing 22 species were sampled for this study (Table 1.1). Most species sequenced

were collected from soil, i.e., from non-flooded Petri dish cultures as described in

Foissner et al. (2002). Some were from the water and mud occurring in the tanks of

bromeliad plants (Foissner et al., 2003). Cells were either collected from the raw culture

(with other species in the dish) or were isolated into clonal culture (with one to few

starter cells). With the addition of GenBank accessions from previous studies (Appendix

1.A), the current sampling includes exemplars from all orders, 15 families, 18 genera, and

seven morphospecies in the genus Colpoda. Outgroup selection is based partially on

previous analyses.

1.3.2. Identification

Species were identified according to the monograph of Foissner (1993), using live

observations and various silver impregnation techniques. The new species collected

here—Bursaria sp., Platyophrya-like, Platyophrya sp., Rostrophyra sp., and Sagittaria

sp.—will be described in separate papers.

1.3.3. DNA extraction, amplification, cloning and sequencing

Between 10 and 10,000 cells were picked with a micropipette, washed, and placed

into DNA lysis buffer. Genomic DNA was extracted using phenol/chloroform following

standard protocols (Ausubel et al., 1993) or with a DNeasy Tissue kit (Qiagen, CA).

8

Genomic DNA was amplified using universal 5’ and 3’ prime SSU rDNA primers

(Medlin et al., 1998) with one of two polymerases. For some species Vent polymerase

(New England BioLabs, MA) was used with the following cycling conditions: 4:00 at

950; 32 cycles of 0:30 at 950, 0:30 at 540, and 2:00 at 720; 10:00 extension at 720. For

others Phusion polymerase (New England BioLabs, MA) was used with the following

cycling conditions: 0:30 at 980; 36 cycles of 0:30 at 980, 0:15 at 680, 1:30 at 720; 10:00

extension at 720. Amplified products were cleaned with a low-melt gel and Ultrafree-Da

columns (Millipore, MA), or with microCLEAN (The Gel Company, CA).

To assess within-sample variation, amplified products were cloned with the PCR-

SMART Cloning kit (Lucigen, WI), or Zero Blunt TOPO kit (Invitrogen, CA). Positive

clones were identified by PCR screening with AmpliTag Gold polymerase (Applied

Biosystems, CA), and minipreped using Qiaprep Spin Miniprep kit (Qiagen, CA). Clones

were sequenced with the Big Dye terminator kit (Applied Biosystems, CA), using 5’ and

3’ primers as well as two internal primers (Snoeyenbos-West et al., 2002). All sequences

were run on an ABI 3100 automated sequencer.

Three samples required further methods. For the Bryometopus pseudochilodon

indel found in this study, a 5’ primer (AAA CAG TTA TAG GCA GGC AAT TG) was

designed that spanned both sides of the deletion to make sure the sequences containing

the deletion were not an amplification artifact. Genomic DNA was amplified with this

primer along with the universal 3’ primer, following the above protocol. For Colpoda

aspera and Cyrtolophosis mucicola (from Austria), algal contaminant SSU rDNA

sequences were removed by enzymatic digestion. Amplified products were cleaned with

microCLEAN. Re-suspended DNA was incubated at 370 for three hours with BamH1

9

(New England BioLabs, MA). The reaction was stopped with microCLEAN, and the

DNA was cloned with the Zero Blunt TOPO kit and sequenced following the above

protocol.

1.3.4. Genealogical analyses

Phylotypes were constructed from the consensus of the multiple sequence reads of

the cloned products and edited in SeqMan (DNAStar). Pairwise distances for within

samples were calculated as uncorrected distances in PAUP* v4.0b8 (Swofford, 2002).

Phylotypes were aligned using Hmmer v2.1.4 (Eddy, 2001), with default settings. The

training alignment for model building was all available ciliate SSU rDNA sequences

downloaded from the European Ribosomal Database (Wuyts et al., 2004) and aligned

according to their secondary structure. The alignment was further edited by eye in

MacClade v4.05 (Maddison and Maddison, 2005), with ambiguously aligned regions and

base-pair positions with more than five taxa having a gap masked. Remaining gaps were

treated as missing data.

The GTR+I+G evolutionary model was estimated using hLTR in MrModeltest v2

(Nylander, 2004). Maximum parsimony (MP) analyses were carried out in PAUP*

v4.0b8 (Swofford, 2002), with all characters equally weighted and unordered. The TBR

heuristic search option was used, running ten random additions with MulTree option on.

Maximum likelihood (ML) analyses were carried out in RAxML v2.2.0 (Stamatakis,

2006) running 100 replicates. Support for MP and ML analyses came from 1000

bootstrap replicates using heuristic searches. Bayesian analyses was carried out using

MrBayes v3.2.1 (Huelsenbeck and Ronquist, 2003) with support coming from posterior

10

probability using four chains and running 10 million generations. Trees were sampled

every 100 generations. The first 25% of sampled trees were considered ‘burnin’ trees and

were discarded prior to tree reconstruction. A 50% majority rule consensus of the

remaining trees was used to calculate posterior probability. Trees were imaged with

TreeView v1.6.6 (Page, 1996).

1.4. Results

1.4.1. Pairwise distances within collections

SSU rDNA sequences from twenty-seven collections representing 22

morphospecies show, for the most part, less than 0.50% average pairwise difference

among clones within samples (Table 1.1). Sequences are deposited in GenBank, numbers

EU039884-EU039908. Six clones from Bardeliella pulchra show more variation with an

average pairwise difference of 0.94%. Clones from the Bursaria sp. 2 collection contain

two different phylotypes that are 0.94% different. The phylotypes of the Ilsiella palustris

collection from Brazil are 0.57% different, while in the Hawaiian collection phylotypes

are 0.54% different. The levels of within-collection variation are assumed to be a

combination of intraspecific variation and experimental error. Contaminant SSU rDNA

sequences were found in a few cases; for example: algae in B. pulchra, Bresslauides

discoideus, Colpoda aspera, Cyrtolophosis mucicola from Austria, and I. palustris from

Hawaii; fungi in Bryometopus pseudochilodon, and Mykophagophrys terricola; and a

tetrahymenid ciliate from Hausmanniella discoidea.

11

Five species were collected more than once, allowing for some within species

comparison (Table 1.2). There is no variation between the two B. discoideus collected

from Dominican Republic. Between the Malaysian and Niger Colpoda cucullus

collections the average pairwise difference among collections was 0.47%, with no

phylotype shared between the sites. Although there is no difference between the two

Brazilian collections of C. mucicola, the Brazilian and the Austrian collections are 1.71%

different with no phylotype shared between the countries. The I. palustris collections are

0.64% different and likewise do not share phylotypes between the sites.

1.4.2. Deletion within one SSU rDNA copy in Bryometopus pseudochilodon

Two distinct SSU rDNA sequences were characterized from the B.

pseudochilodon collection (Table 1.1). One sequence corresponds to the other full-length

Colpodea sequences found here and from GenBank accessions. The second sequence is

almost identical to the first except there is a 642 bp deletion towards the 5’ prime end of

the SSU rDNA sequence and there are two nucleotide differences on the 5’ end. The

deletion starts at nucleotide position 129 in E. coli (GB# J01695) (Cannone et al., 2002).

There is no evidence of elevated substitutions in the sequence with the deletion. This

shorter sequence was uncovered in two separate amplifications using universal SSU

rDNA primers, as well as from amplifications using a 5’ primer (see methods) that was

designed to span either side of the deletion (data not shown). The deletion spanned

multiple regions of the SSU rDNA molecule that are conserved in all extant organisms

(Mears et al., 2002). We hypothesize that the deletion sequence is a macronuclear variant,

12

which occurred in the process of macronuclear development and has been perpetuated

during asexual divisions.

1.4.3. Intron in Cyrtolophosis mucicola

A 427 bp intron was found in all SSU rDNA clones from C. mucicola collected

from Austria but not the C. mucicola collected from Brazil. The start residue is T and the

ending residue is G, which is consistent with group I introns. Blast results also point to

this sequence being a group I intron (E value= 1e-16 with the group I intron in Fulgio

septica, GB# AJ555452.1; E value= 5e-13 with the group I intron in Acanthamoeba sp.,

GB# EF140633.1). There is no evidence for a homing endoculease gene in the intron.

The insertion position of this intron in the SSU rDNA molecule corresponds to nucleotide

516 in E. coli (GB# J01695) (Cannone et al., 2002), which is a hotspot for group I intron

insertions (Jackson et al., 2002).

1.4.4. SSU rDNA genealogy of the Colpodea

After a preliminary analysis using multiple exemplars from all eleven ciliate

classes, only Colpodea sequences and close outgroups were chosen for more detailed

analyses. The potential sister classes in this analysis as determined in the preliminary

global ciliate analysis are the same as in previous studies: Nassophorea, Plagiopylea,

Prostomatea, and Oligohymenophorea (data not shown). One phylotype from each

sampled species was used in the alignment, except two representatives of C. mucicola

(because they may underlie two species, see below) and Bursaria sp. 2 (because the other

Bursaria sequences are relatively close).

13

The final SSU rDNA alignment used for comparing the morphological

hypotheses of the Colpodea and its subgroups includes 59 sequences and has a length of

1582 unmasked nucleotides, of which 219 are parsimoniously informative. The most

parsimonious tree from the MP analysis is 3349 in length, with a Consistency index of

0.3842, and a Homoplasy index of 0.6157. The most likely tree from the ML analysis has

a log likelihood of -17450.098, while the most likely tree from the Bayesian analysis has

a log likelihood of -17445.169.

Here we present only the most likely Bayesian tree with node support from all

three methods (Figure 1.2, see Supplementary Figure 1.1 for all node support values).

The topologies of the MP- and ML-derived genealogies are mostly congruent with the

Bayesian topology, except in three places. First, in the MP and ML analyses

Cyrtolophodidia II (see below) is basal to the rest of the

Colpodea+Oligohymenophorea+Plagiopylea+Prostomatea with no bootstrap support, and

a paraphyletic Nassophorea is basal to this group with no bootstrap. Second, in the MP

and ML analyses Bryometopus pseudochilodon is basal to the rest of its order with no

bootstrap support, while in the Bayesian tree B. sphagni is basal. Third, in the MP and

ML analyses the order Grossglockneriida forms an unsupported clade with Colpoda

aspera, C. steinii, Chain-forming colpodid, and Hausmaniella discoidea, while in the

Bayesian tree it does not.

In our analyses, there is no support for the monophyly of the Colpodea:

monophyly of the class is weakly rejected by all three methods based on tree topologies

and support values. Furgasonia and Obertrumia (both in the class Nassophorea) fall out

sister to part of the order Cyrtolophosidida with no support from all three methods (- MP

14

bootstrap/- ML bootstrap/- Bayesian posterior probability; support >50% or 0.5 is shown

as ‘-’). Changing the number of outgroup classes does not significantly alter this

nonmonophyletic topology as the deep nodes are not well resolved anyways and there is

no support for any class to be sister to the Colpodea. The rest of the Colpodea forms a

monophyletic clade with weak support from MP and ML but with high support from

Bayesian analysis (53/56/0.99).

Support for relationships among the outgroups varied by method. The class

Prostomatea is paraphyletic with only moderate support from Bayesian analysis (-/-

/0.90), with the genus Coleps sister to a well-supported monophyletic class Plagiopylea

(100/100/1.00). The monophyly of Oligohymenophorea is moderately to highly

supported in ML and Bayesian analyses (59/71/1.00). The clade containing Prostomatea,

Plagiopylea and Oligohymenophorea is moderately to highly supported by ML and

Bayesian analysis (53/-/1.00).

Monophyly of the morphologically defined groups could be assessed with our

single gene tree for every order within the Colpodea, except Bryophryida as only one

morphospecies was sampled for this order. Sorogenida, with two genera, is monophyletic

with full support (100/100/1.00). Bursariomorphida, with four taxa, is monophyletic also

with high support (93/89/1.00). Grossglockneriida, with two genera, is likewise

monophyletic with moderate to high support (90/82/1.00).

Order Cyrtolophosidida, with six sampled genera, is not monophyletic. The genus

Cyrtolophosis falls sister to the order Colpodida with moderate to high support

(86/96/1.00). The remaining Cyrtolophosidida genera (Ottowphrya, Platyophrya-like,

15

Platyophrya, Rostrophrya, and Sagittaria) form a paraphyletic group, with order

Sorogenida nested within it, at the base of the class with high support (100/99/1.00).

Order Bryometopida, with one genus and three species sampled, is not

monophyletic with full support from all three methods (100/100/1.00). Order

Bursariomorphida is nested within this order, and it is sister to Bryometopus

pseudochilodon with full support (100/100/1.00). To determine whether this topology is

spurious due to the GenBank accession for Bryometopus sphagni missing about 500 bp

from the 5’ end, the Bryometopida and Bursariomorphida sequences were realigned (with

C. magna and C. mucicola as outgroups) minus the 5’ end; the same topology was found

with this alignment (data not shown). Order Colpodida is not monophyletic with high

support (97/93/1.00), containing orders Grossglockneriida and Bryophryida. B. pulchra is

sister to Notoxoma parabryophryides (order Bryophryida) with moderate to high support

(87/94/0.98).

Monophyly of the genus Colpoda was assessed in this molecular analysis using

eight morphospecies within the genus and numerous close outgroups. Colpoda is not

monophyletic in the SSU rDNA genealogy with moderate to full support (73/72/1.00).

Most Colpoda species form a sister group to the Grossglockneriida with no support from

any method (-/-/-). B. vorax, B. discoideus, and Colpoda henneguyi form a clade with

moderate to high support (89/84/1.00). This clade is in turn sister to most of the

remaining Colpoda species with moderate to high support (72/80/0.97). To determine if

the topology of the Colpoda phylotypes is robust, only Colpodida, Grossglockneriida,

and Bryophryida phylotypes were realigned and remasked for a separate analysis; overall,

the resulting ingroup topology is concordant with the full class analysis (data not shown).

16

1.5. Discussion

1.5.1. Comparisons between morphology and molecules

Here we compare the morphologically-based classification and well-supported

SSU rDNA nodes. Furthermore, we evaluate the possible evolution of morphological

characters in light of the SSU rDNA genealogy.

1.5.1.1. The class: We find no molecular support for the monophyly of the class

Colpodea based on analyses of SSU rDNA sequences (Figure 1.2). Conversely, the

nonmonophyly of the class (with part of the class Nassophorea being sister to part of the

order Cyrtolophosidida) is not well supported either. Similarly, the Nassophorea is also

not monophyletic with respect to the Colpodea, though with no support. The

nonmonophyletic relationships of the Colpodea with respect to the Nassophorea should

not be given much weight, as there is neither support for this relationship nor for the

Nassophorea even being sister to the Colpodea. The SSU rDNA genealogy here provides

little support for class-level relationships within the subphylum Intramacronucleata in

general, as seen elsewhere (Lynn, 2003).

These results do not pose a serious challenge to Lynn’s (1976; 1981) structural

conservatism hypothesis given the limited support at deep nodes. On the other hand,

these results do challenge Bardele’s (1981; 1989) use of ciliary plaques in his argument

that the members in the Colpodea are not closely related.

1.5.1.2. The orders: Molecular support for monophyly could be assessed for all orders

within the Colpodea except the Bryophryida. The SSU rDNA genealogy presented here

17

does support much of the morphologically-based classification of the Colpodea, although

there is some discordance at the ordinal level between morphology and molecules

(Figure 1.2).

The order Cyrtolophosidida is polyphyletic. Cyrtolophosidida I, containing the

genus Cyrtolophosis, falls away from Cyrtolophosidida II, containing the most recent

common ancestor of Sagittaria and Platyophrya and all of its descendants plus the order

Sorogenida. This nonmonophyly of the Cyrtolophosidida suggests the need for a

reevaluation of the character that was used to establish this group. Cyrtolophosidida was

circumscribed based on the shared outer membrane of the nuclear envelope of the

micronucleus and macronucleus (Foissner, 1985; Foissner, 1993a). This character,

however, has only been confirmed with transmission electron microscopy for six species:

Aristerostoma marinum (Detcheva and Puytorac, 1979), Cyrtolophosis mucicola

(Detcheva, 1976; Didier et al., 1980), Platyophrya sphagni (Kawakami, 1991),

Platyophrya spumacola (Dragesco et al., 1977), Pseudocyrtolophosis alpestris (Foissner,

1993a), and Woodruffides metabolicus (Golder, 1976). Njine (1979) states that nuclei in

Kuklikophrya ougandae share an outer membrane (and presents a drawing of a stained

cell showing this), but does not present an electron micrograph. Platyophryides latus is

drawn with a shared outer membrane by Dragesco and Dragesco-Kernéis (1979), but

Puytorac et al. (1992) show that the membranes are separate with their transmission

electron micrographs. Foissner (1993a) argues that two taxa, Sagittaria australis and

Woodruffia australis, have the shared outer membrane because of their thick silver-

stained membranes. On the other hand, Díaz et al. (2000) show separate outer nuclear

membranes in Cyrtolophosis elongata. Hence, the shared outer membrane of the nuclear

18

envelope of the micronucleus and macronucleus is not only a weak character for the

Cyrtolophosidida, but also one whose distribution is neither well known nor confirmed

(Figure 1.1B, character 4). Future transmission electron microscopy studies are much

needed to confirm the presence or absence of this character in other species. Foissner et

al. (2002) suggest those species with a separate outer micronucleus and macronucleus

membrane can be transferred to the clade Plesiocaryon or into the order Sorogenida (as

was done with Ottowphrya).

There are morphological differences between the two Cyrtolophosidida groups. In

Cyrtolophosidida I, there are two segments in the paroral (right oral) membranes, the

anterior bearing tuft-like cilia (the unique feature of its family) (Figure 1.1B, character

15). Only one paroral segment is present in taxa in Cyrtolophosidida II (Figure 1.1B,

character 14). These groups also differ in the presence of non-ciliated kinety on the right

margin of the adoral organelles in Cyrtolophosidida I (and its family), which is absent in

Cyrtolophosidida II (Figure 1.1B, character 16).

Although originally placed with the haptorid ciliates (Bradbury and Olive, 1980),

the close relationship between the Sorogenida and the Cyrtolophosidida was soon

recognized morphologically (Bardele et al., 1991; Foissner, 1985; Small and Lynn,

1981). This relationship was confirmed in a previous SSU rDNA analysis (Lasek-

Nesselquist and Katz, 2001) and the SSU rDNA topology presented here. The Sorogenida

was originally separated from the Cyrtolophosidida because it lacked the shared outer

membrane of the nuclear envelope of the micronucleus and macronucleus (Foissner,

1985; Foissner, 1993a)—although this character maybe is weak (see above)—and

because of its slime mold-like aerial sorocarp in one life history stage (Fig. 1.1, character

19

5). Like the Cyrtolophosidida, the Sorogenida has brick-shaped organelles on the left

slope of the vestibulum and pleurotelokinetal stomatogenesis (Figure 1.1B, character 1)

(Foissner, 1993a). The SSU rDNA genealogy suggests that the aerial sorocarp of

Sorogena may represent a complex apomorphy arising from within Cyrtolophosidida II.

The order Bryometopida is paraphyletic in relation to the monophyletic

Bursariomorphida in the SSU rDNA genealogy. The close relationship between

Bryometopida and Bursariomorphida was also found by Foissner and Kreutz (1998) and

Lynn et al. (1999). Although these two orders differ in their silverline pattern

(Bryometopida having ‘kryellid’ to ‘platyophryid,’ Bursariomorphida having ‘colpodid’),

taxa in these two orders share an apical oral opening, a ventral cleft, conspicuous adoral

organelles, and an emergence pore in their cysts (Foissner and Kreutz, 1998; Foissner,

pers. obs.).

The order Colpodida is paraphyletic in our molecular analyses, though support is

limited at many nodes. That the Grossglockneriida was close to the Colpodida has been

proposed as they share the unique (in the Colpodea) merotelokinetal stomatogenesis

(Figure 1.1B, character 12), colpodid silverline pattern (Figure 1.1B, character 6), and a

simple oral polykinetid (Aescht et al., 1991; Foissner, 1993a). These two orders are even

lumped together in some classifications (Lynn and Small, 1997; Lynn and Small, 2002).

The question remained just how they were related: the SSU rDNA genealogy here

suggests that the Grossglockneriida falls within the Colpodida, not sister to it. The

position of Bryophryida within the Colpodida has not been hypothesized as the

Bryophryida has a platyophryid silverline pattern and brick-shaped organelles on the left

20

vestibulum (Figure 1.1B, character 1). The Bryophryida and Colpodida do, though, share

a deep vestibulum (Foissner, 1993a).

The use of differences in the type of division seems to be helpful at the ordinal

level. As suggested by Foissner (1993a): pleurotelekinetal stomatogenesis is probably

plesiomorphic within the Colpodea. Only orders Colpodida and the Grossglockneriida

have merotelokinetal division (Foissner, 1993a). Stomatogenesis is undescribed in

Bryophryida; assuming that its phylogenetic position found here is confirmed in future

studies, then it is predicted that its division type should be merotelokinetal. On the other

hand, the power of the silverline pattern for use in the systematics of the Colpodea at the

ordinal level is debatable. While Foissner (1993a) uses differences in silverlines to help

construct a higher-level classification, Foissner and Kreutz (1998) Lynn et al. (1999)

argue that this character is sometimes misleading. The results presented here are in

agreement with Lynn et al. (1999) on the limitations of the use of silverline patterns at the

ordinal level.

1.5.1.3. The genus Colpoda: In our molecular analyses the large genus Colpoda is

paraphyletic not only in relation to genera within its own family, but also to other

families in its order (Figure 1.2). Most of the relationships among the Colpoda

morphospecies in the SSU rDNA tree are not well supported; there is support for

Bresslaua and Bresslauides nesting within the Colpoda. Bresslaua was originally

separated from Colpoda based on a difference in vestibulum size (Kahl, 1931). However,

Claff et al. (1941), Foissner (1985; 1993a), and Lynn (1979) find that Bresslaua’s

voracious feeding behavior and its left-projecting vestibular wall (as opposed to right-

21

projecting in Colpoda) are probably better characters to separate the genus from Colpoda.

The SSU rDNA topology suggests that these characters may represent apomorphies

arising from within a Colpoda clade. Bresslauides (and its family Hausmanniellidae) also

falls within the Colpoda in the SSU rDNA tree. This genus was circumscribed based on

the unique semicircular right oral polykinetid that was longer than the left as opposed to

being equal in the Colpodidae (Foissner, 1987; Foissner, 1993a). Because Bresslauides is

not falling out with the other member of its family (Hausmanniella) sampled here, the

character of a semicircularly curved right oral polykinetid may have evolved more than

once.

1.5.2. Open questions with some species designations

The level of diversity among some SSU rDNA sequences from the morphospecies

collected here suggests possible problems with some circumscriptions. Colpoda magna

and C. minima differ little in the SSU rDNA phylotypes, indicating a need for further

genetic studies. These morphospecies species differ in size and kinety number, as well as

the number of micronuclei, with one in C. minima and 2-16 in C. magna (Foissner,

1993a). The low genetic distance between these two species and the lack of much

morphological differences could point to these being nascent but “biological” species.

Alternatively, C. minima and C. magna may represent morphological variation within a

single species where a change in micronuclei and kinety number is correlated with size.

The C. mucicola morphospecies may represent two genetic species: there is a

putative group I intron in the Austrian collection that is absent from the Brazilian

collections, and there is greater than 1% pairwise distance between the Austrian and

22

Brazilian collections. In contrast, the diversity in SSU rDNA phylotypes for the genus

Bursaria from this study and GenBank accessions (1.31% average pairwise distance)

supports the view that there are more than one species in the genus, although some argue

for there being only one species.

We do not find the large sequence diversity in our Colpoda morphospecies as

does Nanney et al. (1998). In our analysis we find a 2.79% average pairwise distance

among the Colpoda morphospecies sampled here and from GenBank accessions, while

Nanney et al. (1998) find an average “slack” value of 31.5% among their Colpoda. There

are at least two reasons for this difference. First, our analyses were based on SSU rDNA,

while theirs is based on 190 bp of the hyper-variable D2 region of the large subunit rRNA

(LSU rDNA). Second, our analyses of distance used the uncorrected distance method in

PAUP*, while theirs use string analyses in the program PHYLOGEN. Using our distance

method, Nanney et al.’s (1998) data show an average pairwise distance of 20.73% for the

D2 region of the LSU rDNA (data not shown). Despite the difference in levels of

variation between the SSU rDNA and the short variable region of the LSU rDNA, the

topology found by Nanney et al. (1998) among their five Colpoda morphospecies is

congruent with our analyses (data not shown).

1.5.3. Evidence for sex

Conjugation (ciliate sex) is documented in all ciliate classes (Bell, 1988; Dini and

Nyberg, 1993; Miyake, 1996; Sonneborn, 1957). In the Colpodea conjugation is only

known in B. truncatella even though over the decades researchers have looked for

conjugation in other species but have yet to observe it (Foissner, 1993a; Raikov, 1982).

23

There are a few reports of possible conjugation in some species of Colpoda; because

nuclear division or exchange was not shown these observations are possibly of

“pseudoconjugation,” where exchange of nuclei does not occur (Foissner, 1993a).

Assuming that Colpodea species behave genetically in a way similar to other

eukaryotes, we could predict that if the Colpodea were asexual, allelic variation would be

high within species (Mark Welch and Meselson, 2000; Normark, et al. 2003). The low

allelic values within most collections sampled here suggest that the Colpodea species are

indeed having sex albeit covertly. There is an important caveat in this statement in that

the number of clones sequenced per morphospecies in this study is relatively low (1-7

clones) and we could have missed some variation. The results here are in opposition to

Bowers et al. (1998), who present isozyme evidence for asexuality for three Colpoda

species. Although cryptic sex is consistent with the low allelic values found here, further

evidence of conjugation is much needed to confirm sex within the Colpodea beyond B.

truncatella.

1.5.4. Group I intron in Cyrtolophosis mucicola

While group I introns are widespread in microbial eukaryotes (Bhattacharya et al.,

1996; Haugen et al., 2003; Haugen et al., 2005; Snoeyenbos-West et al., 2004; Wikmark

et al., 2007), the putative group I intron found in the Austrian C. mucicola morphospecies

is the fourth identification of this type of intron in ciliates. The other known species with

group I introns are: Tetrahymena thermophila (Grabowski et al., 1981), Acineta sp.

(Snoeyenbos-West et al., 2004), and Tokophrya lemnarum (Snoeyenbos-West et al.,

2004). Undoubtedly there remain more of these introns to be uncovered in future

24

sequencing projects of the various ciliate groups. We suggest that the intron in the

Austrian Cyrtolophosis mucicola is a product of a recent horizontal transfer into the SSU

rDNA locus, as group I introns are known to be mobile over relatively short evolutionary

time scales (Haugen et al., 2005; Simon et al., 2005) and because it was not found in

other isolates of the species or other Colpodea sequences.

1.5.5. Reconciling morphology and molecules in the Colpodea

In large part morphology and the SSU rDNA genealogy agree in the hypothesized

relationships within the ciliate class Colpodea, although the paraphyletic relationships

among previously hypothesized closely related taxa was unexpected (Figure 1.1). The

SSU rDNA genealogy is based on a single gene and may not follow the actual species

phylogeny (e.g., Doyle, 1992; Maddison, 1997). Further tests using other loci are needed

to confirm the areas where there is discordance between morphology and molecules.

25

1.6. References

Adl, S. M., A. G. B. Simpson, M. A. Farmer, R. A. Andersen, O. R. Anderson, J. R. Barta, S. S. Browser, G. Brugerolle, R. A. Fensome, S. Fredericq, T. Y. James, S. Karpov, P. Kugrens, J. Krug, C. E. Lane, L. A. Lewis, J. Lodge, D. H. Lynn, D. G. Mann, R. M. McCourt, L. Mendoza, O. Moestrup, S. E. Mozley-Standridge, T. A. Nerad, C. A. Shearer, A. V. Smirnov, F. W. Spiegel, and M. Taylor. 2005. The new higher level classification of eukaryotes with emphasis on the taxonomy of protists. J. Eukaryot. Microbiol. 52:399-451.

Aescht, E., W. Foissner, and M. Mulisch. 1991. Ultrastructure of the mycophagous

ciliate Grossglockneria acuta (Ciliophora, Colpodea) and phylogenetic affinities of colpodid ciliates. Eur. J. Protistol 26:350-364.

Ausubel, F. M., R. Brent, R. E. Kingston, D. D. Moore, J. G. Seidman, J. A. Smith, and

K. Struhl. 1993. Current Protocols in Molecular Biology. Wiley-Liss, New York. Bardele, C. F. 1981. Functional and phylogenetic aspects of the ciliary membrane: a

comparative freeze-fracture study. Biosystems 14:403-421. Bardele, C. F. 1989. From ciliate ontogeny to ciliate phylogeny: a program. Boll. Zoo.

56:235-243. Bardele, C. F., W. Foissner, and R. L. Blanton. 1991. Morphology, morphogenesis and

systematic position of the sorocarp forming ciliate Sorogena stoianovitchae Bardbury & Olive 1980. J. Protozool 38:7-17.

Bell, G. 1988. Sex and Death in Protozoa: The History of an Obsession. Cambridge

University Press, Cambridge. Bhattacharya, D., T. Friedl, and S. Damberger. 1996. Nuclear-encoded rDNA group I

introns: Origin and phylogenetic relationships of insertion site lineages in the green algae. Mol. Biol. Evol. 13:978-989.

Bowers, N., T. T. Kroll, and J. R. Pratt. 1998. Diversity and geographic distribution of

riboprints from three cosmopolitan species of Colpoda Müller (Ciliophora: Colpodea). Eur. J. Protistol. 34:341-347.

Bradbury, P. C., and L. S. Olive. 1980. Fine structure of the feeding stage of a sorogenic

ciliate, Sorogena stoianoitchae gen. n. sp. n. J. Protozool. 27:267-277. Cannone, J. J., S. Subramanian, M. N. Schnare, J. R. Collett, L. M. D'Souza, Y. S. Du, B.

Feng, N. Lin, L. V. Madabusi, K. M. Muller, N. Pande, Z. D. Shang, N. Yu, and R. R. Gutell. 2002. The Comparative RNA Web (CRW) Site: an online database of comparative sequence and structure information for ribosomal, intron, and other RNAs. BMC Bioinformatics 3:2.

26

Claff, C. L., V. C. Dewey, and G. W. Kidder. 1941. Feeding mechanisms and nutrition in three species of Bresslaua. Biol. Bull. 81:221-234.

Detcheva, R. B. 1976. Particularités ultracrturales du cilié Cyrtolophosis mucicola

Stokes, 1885. C. r. Séanc. Soc. Biol. Clermon-Ferrand 170:112-114. Detcheva, R. B., and P. d. Puytorac. 1979. Un nouvel de cilie a micronoyau inclus dans

l'enveloppe macronucleaire: le genera Aristerostoma Kahl, 1926. Annls. Stn. Biolog. de Besse 13:247-251.

Didier, P., P. Depuytorac, N. Wilbert, and R. Detcheva. 1980. A propos d'observations

sur l'ultrastructure de cilié Cyrtolophosis mucicola Stokes, 1885. J. Protozool. 27:72-79.

Dini, F., and D. Nyberg. 1993. Sex in ciliates. Adv. Microb. Ecol. 13:85-153. Doyle, J. J. 1992. Gene trees and species trees: molecular systematics as one-character

taxonomy. Syst. Bot. 17:144-163. Dragesco, J., and A. Dragesco-Kernéis. 1979. Cilies muscicoles nouveaux ou peu

connus. Acta Protozool. 18:401-416. Dragesco, J., G. Fryd-versavel, F. Iftode, and P. Didier. 1977. Le cilie Platyophrya

spumacola Kahl, 1926: morphologie, stomatogenèse et ultrastructure. Protistologica 13:419-434.

Díaz, S., A. Martin-González, S. Borniquel, and J. C. Gutiérrez. 2000. Cyrtolophosis

elongata (Colpodea, Ciliophora): Some aspects of ciliary pattern, division, cortical and nuclear changes during encystment and resting cyst ultrastructure. Eur. J. Protistol. 36:367-378.

Eddy, S. R. 2001. HMMER: Profile hidden markov models for biological sequence

analysis. http://hmmer.wustl.edu. Foissner, W. 1985. Klassifikation und phylogenie der Colpodea (Protozoa: Ciliophora).

Arch. Protistenk. 129:239-290. Foissner, W. 1987. Neue und wenig bekannte hypotriche und colpodide Ciliaten

(Protozoa: Ciliophora) aus Böden und Moosen. Zool. Beitr. (N. F.) 31:187-282. Foissner, W. 1993a. Colpodea (Ciliophora). Protozoenfauna 4/1:i-x, 1-798. Foissner, W. 1993b. Corticocolpoda kaneshiroae n. g., n. sp., a new colpodid ciliate

(Protozoa, Ciliophora) from the bark of ohia trees in Hawaii. J. Eukaryot. Microbiol. 40:764-775.

27

Foissner, W. 1993c. Idiocolpoda pelobia gen. n., sp. n., a new colpodid ciliate (Protozoa, Ciliophora) from an ephemeral stream in Hawaii. Acta Protozool. 32:175-182.

Foissner, W. 1994. Pentahymena Corticicola nov. gen., nov. spec., a new colpodid

ciliate (Protozoa, Ciliophora) from bark of acacia trees in Costa Rica. Arch. Protistenk. 144:289-295.

Foissner, W. 1995. Tropical protozoan diversity: 80 ciliate species (Protozoa,

Ciliophora) in a soil sample from a tropical dry forest of Costa Rica, with description of four new genera and seven new species. Arch. Protistenk. 145:37-79.

Foissner, W. 1996. Ontogenesis of ciliated protozoa with emphasis on stomatogenesis.

Pp. 95-178 in K. Hausmann and P. C. Bradbury, eds. Ciliates: Cells as Organisms. Gustav Fischer, Stuttgart.

Foissner, W. 1999. Description of two new, mycophagous soil ciliates (Ciliophora,

Colpodea): Fungiphrya strobli n. g., n. sp. and Grossglockneria ovata n. sp. J. Eukaryot. Microbiol. 46:34-42.

Foissner, W. 2003. Pseudomaryna australiensis nov gen., nov spec. and Colpoda

brasiliensis nov spec., two new colpodids (Ciliophora, Colpodea) with a mineral envelope. Eur. J. Protistol. 39:199-212.

Foissner, W., S. Agatha, and B. H. 2002. Soil ciliates (Protozoa, Ciliophora) from

Namibia (Southwest Africa), with emphasis on two contrasting environments, the Etosha Region and the Namib Desert: Part I. Densia 5:1-1063.

Foissner, W., and M. Kreutz. 1998. Systematic position and phylogenetic relationships

of the genera Bursaridium, Paracondylostoma, Thylakidium, Bryometopus, and Bursaria (Ciliophora: Colpodea). Acta Protozool. 37:227-240.

Foissner, W., M. Struder-Kypke, G. W. M. van der Staay, S. Y. Moon-van der Staay, and

J. H. P. Hackstein. 2003. Endemic ciliates (Protozoa, Ciliophora) from tank bromeliads (Bromeliaceae): a combined morphological, molecular, and ecological study. Eur. J. Protistol. 39:365-372.

Golder, T. K. 1976. The macro-micronuclear complex of Woodruffia metabolica. J.

Ultrastruct. Res. 54:169-175. Grabowski, P. J., A. J. Zaug, and T. R. Cech. 1981. The intervening sequence of the

ribosomal RNA precursor is converted to a circular RNA in isolated nuclei of Tetrahymena. Cell 23:467-476.

28

Haugen, P., D. H. Coucheron, S. B. Ronning, K. Haugli, and S. Johansen. 2003. The molecular evolution and structural organization of self-splicing group I introns at position 516 in nuclear SSU rDNA of myxomycetes. J. Eukaryot. Microbiol. 50:283-292.

Haugen, P., D. M. Simon, and D. Bhattacharya. 2005. The natural history of group I

introns. Trends Genet. 21:111-119. Huelsenbeck, J. P., and F. R. Ronquist. 2003. MrBayes 3: Bayesian phylogenetic

inference under mixed models. Bioinformatics 19:1572-1574. Jackson, S. A., J. J. Cannone, J. C. Lee, R. R. Gutell, and S. A. Woodson. 2002.

Distribution of rRNA introns in the three-dimensional structure of the ribosome. J. Mol. Biol. 323:35-52.

Kahl, A. 1931. Urtiere oder Protozoa I: Wimpertiere oder Ciliate (Infusoria) 2.

Holotricha außer den im 1. Teil behandelten Prostomata. Tierwelt Dtl. 21:181-398.

Kawakami, H. 1991. An endosymbiotic Chlorella-bearing ciliate: Platyophrya

chlorelligera Kawakami 1989. Eur. J. Protistol. 26:245-255. Lasek-Nesselquist, E., and L. A. Katz. 2001. Phylogenetic position of Sorogena

stoianovitchae and relationships within the class Colpodea (Ciliophora) based on SSU rDNA sequences. J. Eukaryot. Microbiol. 48:604-607.

Lynn, D. H. 1976. Comparative ultrastructure and systematics of colpodida. Structural

conservatism hypothesis and a description of Colpoda steinii Maupas. J. Protozool. 23:302-314.

Lynn, D. H. 1979. Fine structural specializations and evolution of carnivory in

Bresslaua (Ciliophora, Colpodida). Trans. Am. Microsc. Soc. 98:353-368. Lynn, D. H. 1981. The organization and evolution of microtubular organelles in ciliated

protozoa. Biol. Rev. 56:243-292. Lynn, D. H. 2003. Morphology or molecules: How do we identify the major lineages of

ciliates (Phylum Ciliophora)? Eur. J. Protistol. 39:356-364. Lynn, D. H., and E. B. Small. 1997. A revised classification of the phylum Ciliophora

Doflein, 1901. Rev. Soc. Mex. Hist. Nat. 47:65-78. Lynn, D. H., and E. B. Small. 2002. Phylum Ciliophora Doflein, 1901. Pp. 371-656 in J.

J. Lee, G. F. Leedale and P. C. Bradbury, eds. An Illustrated Guide to the Protozoa, 2nd ed. Society of Protozoologists, Lawrence, Kansas.

29

Lynn, D. H., A.-D. G. Wright, M. Schlegel, and W. Foissner. 1999. Phylogenetic relationships of orders within the class Colpodea (Phylum Ciliophora) inferred from small subunit rRNA gene sequences. J. Mol. Evol. 48:605-614.

Maddison, W. P. 1997. Gene trees in species trees. Syst. Biol. 46:523-536. Maddison, W. P., and D. R. Maddison. 2005. MacClade v. 4.0.8. Sinauer Assoc. Mark Welch, D., and M. Meselson. 2000. Evidence for the evolution of bdelloid rotifers

without sexual reproduction or genetic exchange. Science 288:1211-1215. Mears, J. A., J. J. Cannone, S. M. Stagg, R. R. Gutell, R. K. Agrawal, and S. C. Harvey.

2002. Modeling a minimal ribosome based on comparative sequence analysis. J. Mol. Biol. 321:215-234.

Miyake, A. 1996. Fertilization and sexuality in ciliates. Pp. 243-290 in K. Hausmann

and P. C. Bradbury, eds. Ciliates: Cells as Organisms. Gustav Fischer, Stuttgart. Nanney, D. L., C. Park, R. Preparata, and E. M. Simon. 1998. Comparison of sequence

differences in a variable 23S rRNA domain among sets of cryptic species of ciliated protozoa. J. Euk. Microbiol. 45:91-100.

Njine, T. 1979. Etude ultrastructurale de cilié Kuklikophyra dragescoi gen. n., sp. n. J.

Protozool. 26:589-598. Normark, B. B., O. P. Judson, and N. A. Moran. 2003. Genomic signatures of ancient

asexual lineages. Biol. J. Linn. Soc. 79:69-84. Nylander, J. A. 2004. MrModeltest v2. Program distributed by the author. Evolutionary

Biology Center, Uppsala University Page, R. D. 1996. TREEVIEW: An application to display phylogenetic tress on personal

computers. Comput. Appl. Biosci. 12:357-358. Puytorac, P. d. 1994. Phylum Ciliophora Doflein, 1901. Pp. 1-15 in P. P. de, ed. Traité

de Zoologie, Tome II, Infusoires Ciliés, Fasc. 2, Systématique. Masson, Paris. Puytorac, P. d., M. R. Kattar, C. A. Grolière, and I. d. Silva Neto. 1992. Polymorphism

and ultrastructure of the colpodean ciliate of the genus Platyophryides Foissner, 1987. J. Protozool. 39:154-159.

Raikov, I. B. 1982. The Protozoan Nucleus: Morphology and Evolution. Springer-

Verlag, Wien.

30

Simon, D., J. Moline, G. Helms, T. Friedl, and D. Bhattacharya. 2005. Divergent histories of rDNA group I introns in the lichen family Physciaceae. J. Mol. Evol. 60:434-446.

Small, E. B., and D. H. Lynn. 1981. A new macrosystem for the phylum Ciliophora

Doflein, 1901. Biosystems 14:387-401. Snoeyenbos-West, O. L. O., J. Cole, A. Campbell, D. W. Coats, and L. A. Katz. 2004.

Molecular phylogeny of phyllopharyngean ciliates and their group I introns. J. Eukaryot. Microbiol. 51:441-450.

Snoeyenbos-West, O. L. O., T. Salcedo, G. B. McManus, and L. A. Katz. 2002. Insights

into the diversity of choreotrich and oligotrich ciliates (Class: Spirotrichea) based on genealogical analyses of multiple loci. Int. J. Syst. Evol. Microbiol. 52:1901-1913.

Sonneborn, T. M. 1957. Breeding systems, reproductive methods, and species problems

in protozoa. Pp. 155-324 in E. Mayr, ed. The Species Problem. American Association for the Advancement of Science, Washington D. C.

Stamatakis, A. 2006. RAxML-VI-HPC: Maximum likelihood-based phylogenetic

analyses with thousands of taxa and mixed models. Bioinformatics 22:2688-2690. Stechmann, A., M. Schlegel, and D. H. Lynn. 1998. Phylogenetic relationships between

Prostome and Colpodean ciliates tested by small subunit rRNA sequences. Mol. Phyl. Evol. 9:48-54.

Swofford, D. 2002. PAUP*. Phylogenetic Analysis Using Parsimony (*and Other

Methods). Sinauer Assoc, Sunderland MA. Wikmark, O.-G., P. Haugen, E. W. Lundblad, K. Haugli, and S. D. Johansen. 2007. The

molecular evolution and structural organization of group I introns at position 1389 in nuclear small subunit rDNA of myxomycetes. J. Eukaryot. Microbiol. 54:49-56.

31

Table 1.1. Taxon sampling within the Colpodea. Species were identified using silver

impregnation by W. Foissner. Type and voucher material of the new species and

the newly investigated populations are deposited at the Oberoesterreichische

Landesmuseum in Linz (LI), Austria. nc - non-clonal culture, c- clonal culture,

npc - non-pure culture, pc - pure culture.

32

Table 1.2. Pairwise distance between collections for species sampled more than once.

PairwiseTaxon Collection site distance (%)Bresslauides discoideus Dominican Republic 1 and 2 0Colpoda cucullus Malaysia and Niger 0.47Cyrtolophosis mucicola Brazil 1 and 2 0Cyrtolophosis mucicola Brazil 1 and Austria 1.71Ilsiella palustris Brazil and Hawaii 0.64

33

Figure 1.1. Evolution among some morphological characters within the class Colpodea.

(A) Hypotheses of relationships among orders and morphological character evolution

modified from Foissner (1993a), where some characters are removed. (B) Possible

alternative evolution of characters mapped out on the SSU rDNA gene tree found here;

the deeper nodes in the Colpodea are not well supported and are thus shown as a

polytomy. The character are: 1) Lkm fiber, pleurotelokinetal stomatogenesis, brick-

shaped adoral organelles, flat vesitibulum, and ‘kreyellid,’ ‘platyophryid,’ or ‘colpodid’

silverline pattern; 2) ‘kreyellid’ silverline pattern; 3) ‘platyophrid’ or ‘colpodid’ silverline

pattern; 4) shared micronuclear and macronuclear outer membrane of the nuclear

envelope; 5) aerial sorocarps; 6) ‘colpodid’ silverline pattern; 7) deep vestibulum, 8)

paroral formation with radial ciliary fields; 9) equidistantly spaced adoral organelles; 10)

conjugation; 11) emergence pore in resting cysts; 12) merotelokinetal stomatogenesis;

13) feeding tube; 14) one paroral membrane segment; 15) two paroral membrane

segments; 16) postoral pseudomembrane. See text for explanations of characters.

34

35

Figure 1.2. SSU rDNA genealogy of the class Colpodea and potential sister classes. The

most likely Bayesian tree is shown. Bayesian posterior probability support is shown by

differences in thickness of branches. Numerical values from bootstrap support is shown

next to the branches as: MP bootstrap/ML bootstrap. Values <50% are shown as ‘-’.

Monophyletic classes and orders are labeled with a solid line, while nonmonophyletic

ones labeled with a dashed line. All support values for all nodes are given in

Supplementary Figure 1.

36

37

Supplementary Figure 1.1: SSU rDNA genealogy of the class Colpodea and potential

outgroups with support for all nodes indicated. The most likely Bayesian tree is shown.

Node support is as follows: MP bootstrap/ML bootstrap/Bayesian posterior probability.

Values <50% are shown as ‘-’. Monophyletic classes and orders are shown with a solid

line, while nonmonophyletic ones shown with a dashed line.

38

39

Appendix 1.A: GenBank accessions used in analyses for both previously sequenced Colpodea taxa and outgroups.Colpodea: GB# Glaucoma chattoni X56533Bresslaua vorax AF060453 Glauconema trihymene AY169274Bryometopus sphagni AF060455 Gruberia sp. L31517Bursaria truncatella U82204 Haleria grandinella AY007443Chain-forming colpodid AY398684 Heliophrya erhardi AY007445Colpoda inflata M97908 Isotricha intestinalis U57770Colpoda steinii DQ388599 Loxodes magnus L31519Platyophrya vorax AF060454 Loxophyllum utriculariae L26448Pseudoplatyophrya nana AF060452 Metopus contortus Z29516Sorogena stoianovitchae AF300285 Metopus palaeformis AY007450

Nyctotherus ovalis AY007454Obertrumia aurea* X65149

Outgroups: GB# Ophrydium versatile AF401526Anophryoides haemophila U51554 Ophryoglena catenula U17355Anoplophrya marylandensis AY547546 Orthodonella apohamatus DQ232761Apofrontonia dohrni AM072621 Oxytricha nova X03948Blepharisma americanum M97909 Paramecium tetraurelia X03772Caenomorpha uniserialis U97108 Parduczia orbis AY187924Cardiostomatella vermiforme AY881632 Pleuronema coronatum AY103188Chilodonella uncinata AF300281 Prorodon teres X71140Climacostomum virens X65152 Prorodon viridis U97111Coleps hirtus U97109 Protocruzia sp. AF194409Coleps sp. X76646 Pseudomicrothorax dubius X65151Didinium nasutum U57771 Schizocaryum dogieli AF527756Diplodinium dentatum U57764 Spirostomum ambiguum L31518Discophrya collini L26446 Stentor roeseli AF357913Ephelota sp. AF326357 Strombidium purpureum U97112Epidinium caudatum U57763 Stylonychia lemnae AF164124Epistylis chrysemydis AF335514 Tetrahymena thermophila X56165Eufolliculina uhligi U47620 Tokophrya lemnarum AY332720Euplotes crassus AY007437 Tracheloraphis sp. L31520Frontonia lynni DQ190463 Trithigmostoma steini X71134Furgasonia blochmanni X65150 Uronema elegans AY103190Geleia simplex AY187927 Vorticella campanula AF335518*In GenBank as Obertrumia georgiana, which is a junior synonym.

40

CHAPTER 2

PHYLOGENETIC PLACEMENT OF THE CYRTOLOPHOSIDIDAE STOKES,

1888 (CILIOPHORA; COLPODEA) AND NEOTYPIFICATION OF

ARISTEROSTOMA MARINUM KAHL, 1931

Micah Dunthorna, Marion Eppingerb, M. V. Julian Schwarzb, Michael Schweikertc, Jens

Boenigkd, Laura A. Katza,e, and Thorsten Stoeckb

aGraduate Program in Organismic and Evolutionary Biology, University of

Massachusetts, Amherst, Massachusetts, 01003, USA

bDepartment of Ecology, University of Kaiserslautern, Erwin-Schroedinger-Strasse 14, D-

67773, Kaiserslautern, Germany

cDepartment of Zoology, Biological Institute, University of Stuttgart, Pfaffenwaldring

57, D-67633 Germany

dAustrian Academy of Sciences, Institute for Limnology, Mondseestr. 9, A-5310

Mondsee, Austria

eDepartment of Biological Sciences, Smith Collage, College Road, Northampton,

Massachusetts, 01063, USA

41

2.1. Abstract

The ciliate family Cyrtolophosididae Stokes, 1888 contains species that are poorly known

from both the morphological and molecular perspectives. To further our understanding of

this family we redescribe one species, Aristerostoma marinum Kahl, 1931. Cells in our

population have an average in vivo size of 15 x 8 µm. There are six rows of somatic

kineties, as well as six dorsal kinetids belonging to sparsely ciliated somatic kineties. The

oral apparatus is comprised of a bipartite paroral membrane and four adoral organelles.

The optimal ecological tolerances match those of the environment in which it was

collected for pH and O2, but not for salinity and temperature. To further test the

phylogenetic placement of the Cyrtolophosididae with increased taxon sampling, we

characterize the small subunit rDNA of three morphospecies: A. marinum, Aristerostoma

sp. ATCC® Number 50986™, and Pseudocyrtolophopsis alpestris. Unconstrained and

constrained molecular analyses support the non-monophyly of the order

Cyrtolophosidida. The family Cyrtolophosididae falls out separately from the rest of its

order. We also place haplotypes from previous environmental studies in a phylogenetic

context within the class Colpodea.

42

2.2. Introduction

Like other taxa in the class Colpodea Lynn and Small, 1981, the taxonomic

history of the family Cyrtolophosididae Stokes, 1888 is one of shifting classifications.

Members of the group were originally placed in the Frontoniidae by Kahl (1931), and

later in the Tetrahymenidae by Corliss (1961). Foissner (1978) used silverline patterns,

and Didier et al. (1980), Lynn (1981), and Puytorac et al. (1979) used transmission

electron micrographs of kinetid ultrastructure to link this family with other Colpodea.

The Cyrtolophosididae was then placed in the order Cyrtolophosidida with other

genera—such as Platyophrya and Sagittaria—in which micronuclei and macronuclei

share an outer membrane of the nuclear envelope (Foissner 1985; Foissner 1993).

The Cyrtolophosididae is currently diagnosed with a number of morphological

characters: species have a bipartite paroral membrane, “colpodid” silverline pattern, and a

non-ciliated kinety on the right margin of their adoral organelles (Foissner 1993).

Morphological variation among the four described genera in the family—Aristerostoma,

Cyrtolophosis, Plesiocaryon, Pseudocyrtolophosis—is not as distinct; Foissner (1993)

even suggests that they might need to be synonymized. Most species are relatively small,

20-35 x 15 µm, with a few living in presumably mucocyst-derived tubes (Foissner 1993).

Some of these, like Aristerostoma marinum (part of the focus of this manuscript), lack

modern descriptions and silver impregnations.

The phylogenetic placement of the Cyrtolophosididae has recently been

questioned. In a molecular analysis of all orders within the Colpodea using small subunit

rDNA (SSU-rDNA) sequences, two Cyrtolophosis mucicola sequences branched

separately from the rest of its morphologically-defined order (Dunthorn et al. 2008). This

43

result challenges the use of the shared outer membranes of the nuclear envelope to unite

the order Cyrtolophosidida. The non-monophyletic SSU-rDNA topology of the order,

though, requires further evaluation with increased taxon sampling.

Here we redescribe A. marinum and a new name-bearing type is designated. We

also move sampling beyond the only one sequenced species, C. mucicola, and further test

the monophyly of the order Cyrtolophosidida with three previously uncharacterized

morphospecies in two genera using SSU-rDNA phylogenetic analyses. Morphological

and molecular hypotheses are compared with constrained analyses, and possible issues

leading to differences between the morphological and molecular hypotheses are

examined. Furthermore, we place GenBank accessions from previous SSU-rDNA

environmental surveys in the context of our increased taxon sampling within the

Cyrtolophosididae.

2.3. Methods

2.3.1. Taxon sampling

Three morphospecies in the family Cyrtolophosididae were isolated for this study

and SSU-rDNA was sequenced from them (Table 2.1). A. marinum was collected from

surface waters of the Framvaren Fjord in southwest Norway (58°09’N, 06°55’E). Pure

cultures for a redescription of this species were established using Schmaltz-Pratt medium

(0.01 g K2HPO4 * 3H2O l-1, 0.1 g KNO3 l-1, 1.45 g CaCl2 * 2H2O l-1, 6.92 g MgSO4 * 7H2O

l-1, 5.51 g MgCl2 * 6H2O l-1, 0.67 g KCl l-1 and 28.15 g NaCl l-1) with heat-inactivated

Klebsiella minuta as a food source. Aristerostoma sp. ATCC® Number 50986™ was

44

originally collected from a marine environment in the Great Marsh, Delware, USA,

where there was gray mud mixed with roots, sand, and clay. This isolate will be

examined morphologically elsewhere. Pseudocyrtolophopsis alpestris was collected from

litter of a spruce forest in Lambrechtshausen near Salzburg, Austria, by W. Foissner.

With the addition of the two GenBank accessions for C. mucicola and sequences

from environmental studies (Table 1), we now have eight exemplars in the

Cyrtolophosididae. Sequences from the rest of the Cyrtolophosidida and from the other

Colpodea orders are from GenBank (Table 1). Outgroup selection was based on previous

studies.

2.3.2. Light and electron microscopy

For light microscopy of living and stained cells, we used a Zeiss Axioplan 2.

Protargol impregnation followed Foissner et al. (1999), with the cells fixed in 1 ml

aqueous saturated HgCl2 with 100 µl Bouin’s fluid (cells vol/fixant vol 1:1, 30 min, RT).

Due to cell sensitivity, a high salt concentration in the medium (causing precipitates

during processing), and a mucus shell covering the organisms (mucocysts, see Results)

other staining and impregnation methods failed (e.g. silver nitrate impregnation with the

Chatton-Lwoff technique and silver carbonate impregnation with the Fernandes-Galiano

technique for saltwater ciliates). All images from living, fixed, and stained cells were

taken by a QImager® Microcam (Intas, Göttingen, Germany) and QCapture© software

(http://www.qimaging.com). For further image processing we used Adobe Photoshop©

7.0 and ImageJ 1.32 (http://rsb.info.nih.gov/ij/).

45

For scanning electron microscopy (SEM), cells were processed following Stoeck

et al. (2005) with slight modifications for fixation: cells/fixans, 1/1; fixans of osmium

tetroxide (2% in artificial seawater [36‰] and ASW following Stetter et al. (1983) for 60

minutes at room temperature). Stubs with fixed and dehydrated cells were coated with

gold (Edwards E306) and observed with a Zeiss DSM940A (Carl Zeiss GmbH, Jena,

Germany).

Preparation of cells for transmission electron microscopy (TEM) followed Stoeck

et al. (2005) with slight modifications in the fixation procedure: a culture aliquot was first

fixed with glutaraldehyde (2.5% final) for 60 min at 4°C. Cells were pelleted and

embedded in a 4% low-melt sea-prep agarose (Roth, Karlsruhe, Germany) (Reize and

Melkonian 1989) with 4% osmium tetroxide in ASW for 60 min in order to concentrate

and handle the small target cells. Ultrathin sections were investigated with a Zeiss EM10

(Oberkochen, Germany) and documented on a Kodak 4489 film (Eastman Kodak, NY).

2.3.3. Terminology

Terminology follows Corliss (1979), Foissner (1993), and the International Code

of Zoological Nomenclature (ICZN 1999). In designating a new type specimen, we

follow Foissner (2002) by allowing the new name bearer to be from a different location

than the specimen originally described by Kahl (1931).

2.3.4. Autecology

The ecological tolerances of A. marinum towards four parameters (% O2 in the

headspace gas, pH, ‰ salinity, and temperature) were experimentally tested in 2-ml

46

batch incubations. Cell activity was measured by counting moving and/or dividing cells

under a dissection microscope within 24, 48, 72, 96 and 168 h periods after inoculation.

The cultures were gradually adapted to higher/lower pH, salinity, and temperature as

outlined by Stoeck et al. (2005).

All incubations were inoculated in six 2-ml parallels (six wells on a 24-wellplate,

Greiner, Germany) in chemically adjusted Schmaltz-Pratt medium at room temperature

(with the exception of the temperature experiment). Salinity was changed by the addition

of 1 M NaCl or Volvic™ water. Changes in pH were adjusted by addition of 1 M NaCO3

or 1 M KH2PO4. The cells were inoculated after pH stabilization (24 h). Heat-deactivated

K. minuta were added as a food source at saturated concentration (108-109 cells/ml).

The preferred oxygen regime was tested by incubation of seven 2-ml parallels in

10-ml injection bottles (Ochs GmbH, Bovenden-Lenglern, Germany). These were stored

inside a gas-tight 1000 ml glass chamber containing a defined headspace gas composition

(0, 1, 2 or 21% oxygen in N2, 4.0 calibration gas qualities, AirLiquide, Darmstadt,

Germany). Final O2-concentration in the medium was reached after 24 h (t0). Anoxic

conditions were established by flushing the medium and incubation vessel with N2

(AirLiquide, Darmstadt, Germany) and by the addition of anaerocult-plates (Merck AG,

Darmstadt, Germany) as an oxygen scavenger. Suboxic and oxic incubations were

flushed twice daily with the appropriate calibrated gas. For each gas concentration, we

prepared four replicates, one of which was sacrificed after each of the testing periods (24,

48, 72, 96 and 168 h) to count moving and/or dividing cells. All experiments were

incubated at room temperature in the dark.

47

2.3.5. Amplification and sequencing

To extract genomic DNA, 0.5-ml aliquots of a culture or 5-10 individually picked

cells were picked with a micropipette, washed, and processed using the protocol for

cultured animal cells of the DNEasy Tissue Kit (Qiagen, Hildesheim, Germany). SSU-

rDNA was amplified using the universal eukaryotic primers EukA and EukB (Medlin et

al. 1988). For A. marinum and P. alpestris, each amplification contained 10-20 ng of

DNA template, 2.5 U HotStar Taq DNA polymerase (Qiagen) in the manufacturer-

provided reaction buffer, 1.5 mM MgCl2, 200 µM of each dNTP, and 0.5 µM of each

oligonucleotide primer. The final volume was adjusted to 50 µl with sterile distilled

water. The PCR protocol for SSU-rDNA gene amplification consisted of an initial hot

start incubation of 15 min at 95 ˚C followed by 30 identical amplification cycles (i.e.,

denaturing at 95 ˚C for 45 s, annealing at 55 ˚C for 1 min, and extension at 72 ˚C for 2.5

min), and a final extension at 72 ˚C for 7 min. Negative control reactions included

Escherichia coli DNA as a template. The resulting PCR products were cleaned with the

PCR MinElute Kit (Qiagen) and cloned into a vector using the TA-Cloning kit

(Invitrogen, Carlsbad, CA). Plasmids were isolated with Qiaprep Spin Miniprep Kit

(Qiagen) from overnight cultures and PCR-reamplified using M13F and M13R primers to

screen for inserts of the expected size (ca. 1.8 kb in case of the SSu-rDNA fragment). For

Aristeristoma sp., Phusion polymerase (New England BioLabs, MA) was used for

amplification and the products were cloned following Dunthorn et al. (2008). All clones

were sequenced bidirectionally (M13 sequence primers) with the Big Dye terminator kit

(Applied Biosystems, Foster City, CA) on either an ABI 3100 or 3730 automated

sequencer.

48

2.3.6. Genealogical analyses

We determined and edited haplotypes from overlapping sequence reads in

SeqMan (DNAStar, Inc., Madison, WI) or CodonCode Aligner v1.2.4 (CodonCode

Corporation, Dedham, MA). Pairwise distances for within and among samples were

calculated as uncorrected “p” distances in PAUP* v4.0b8 (Swofford, 2002). Haplotypes

generated here and the environmental sequences from GenBank were placed into the

alignment used in Dunthorn et al. (2008), but with most non-Colpodea outgroups

removed. The GTR+I+G evolutionary model was selected using hLTR in MrModeltest

v2 (Nylander 2004).

Maximum parsimony (MP) and maximum likelihood (ML) analyses were carried

out in PAUP* v4.0b8 (Swofford 2002), with all characters equally weighted and

unordered. The TBR heuristic option was used to search trees, running ten random

additions with MulTree option on. Support for MP and ML analyses came from 100

bootstrap replicates using heuristic searches. ML bootstraps were run on the Beowulf

cluster at the University of Missouri St. Louis. Bayesian analyses was carried out using

MrBayes v3.2.1 (Huelsenbeck and Ronquist 2003) with support coming from posterior

probability using four chains and running 10 million generations. Trees were sampled

every 1000 generations. The first 25% of sampled trees were considered ‘burnin’ trees

and were discarded prior to tree reconstruction. A 50% majority rule consensus of the

remaining trees was used to calculate posterior probability.

49

2.3.7. Constrained analysis

In addition to the genealogical analysis above, a ML analysis was carried out with

all exemplars in the Cyrtolophosidida constrained to be monophyletic in PAUP* v4.0b8

(Swofford 2002); the particular relationship within the Cyrtolophosidida, though, were

not specified. The resulting tree was compared to the unconstrained ML tree using a one-

tailed KA test (Kishino and Hasegawa 1989) as implemented in PAUP* v4.0b8

(Swofford 2002).

2.3.8. Rate class analyse

Using the full unconstrained alignment, nucleotide positions were partitioned into

eight rate classes using HYPY v0.9b (Kosakovsky Pond et al. 2005). The fastest rate

class was removed from the alignment, and explored using Bayesian analysis as above,

except running 3,000,000 generations. The second fastest rate class was then also

removed and examined likewise.

2.4. Results

2.4.1. Description of the neotype of Aristerostoma marinum Kahl, 1931

While free-swimming, the fast cells spirals, rotating around their longitudinal

axis. Cells have no tendency to clump together either while swimming or in the resting

state. After a few minutes under the microscope numerous cells attached either to the

water surface or to the cover slip with their posterior end. In vivo, the ciliate is 9-23 µm

in length (mean 15 µm, n = 29; measurements rounded to the nearest digit) and 4-11 µm

50

wide (mean 8 µm, n = 29) (Table 2.2). The neotype population of A. marinum has an

oval shape tapering towards the anterior end. An oral structure is visible in the anterior

third of the cell (Figs. 2b, 2.3), while the posterior end displays a pulsating vacuole (Fig.

2.1A). While swimming it becomes visible that the left dorsal side is flattened. The cell

surface displays prominent longitudinal ribs, and on the right lateral side the cortex

carries easily recognizable rows of cilia. Only a few irregularly distributed cilia are

recognizable on the dorsal side. In vivo the somatic cilia are ca. 4 µm in length. We did

not observe any resting stages in any of our cultures. Reproduction occurs by

symmetrogenic binary fission (=perkinetal, Fig. 2.1B).

All protargol impregnations are suboptimal (Fig. 2.1C) and only the examination

of numerous cells enabled a schematic drawing of protargol impregnated structures (Fig.

2.3). Cells are 10-20 µm in length (mean 15 µm, n = 47) and 6-10 µm wide (mean = 8

µm, n = 47). The oval-shaped macronucleus is located submedian (in the posterior half of

the cell) and has a mean diameter of 3 µm. In protargol preparations, the micronucleus is

sometimes delimited from the macronucleus as a lighter-colored structure with a mean

diameter of 1 µm (Fig. 2.1C). Protargol impregnation does not reveal whether the

micronucleus lies within the nuclear membrane. The distribution of extrusomes

(mucocysts) in the cell’s cortex becomes visible (Fig. 2.1C). The oral structure appears

subapical, but details cannot be resolved using this impregnation. Protargol impregnated

kinetids are displayed in the schematic drawing (Fig. 2.3). The cell is characterized by six

rows of somatic kineties. Kinety 1 consists of six dikinetids extending from the anterior

end of the cell 2/3 towards the posterior along the longitudinal axis (Fig. 2.3); it is located

right laterally. Also the right lateral kineties 2 and 3 consist of eight dikinetids each that

51

extend along the complete longitudinal axis from anterior to posterior poles (Fig. 2.3).

Kinety 4 (right lateral-dorsal) is composed of four dikinetids and three uniciliated

kinetids (which are likely to be dikinetids), running from the anterior end of the cell 2/3

towards the posterior along the longitudinal axis (Fig. 2.3). Kinety 5 (dorsal-left lateral)

comprises only two dikinetids located at the anterior end of the cell (Fig. 2.3). Kinety 6

(left lateral - ventral) also consists exclusively of dikinetids (n = 8), which extend along

the whole length of the longitudinal axis and abut left lateral the oral apparatus (Fig. 2.3).

Six kinetids are on the dorsal side. They cannot be assigned to any of the six longitudinal

kineties but most likely belong to sparsely ciliated somatic kineties (Fig. 2.2A). As we

did not succeed in obtaining appropriate transmission electron micrographs of these

kineties, we are not able to define if we are dealing with mono- or di-kinetids—further

TEM work is needed.

2.4.2. Electron microscopy

Scanning electron microscopy confirms protargol impregnation results (Fig.

2.2A-C) and reveals details of the subapical oral structure of A. marinum (Fig. 2.2B). A

gapless paroral membrane surrounds the triangular oral structure on the right side

consisting of 5 anterior dikinetids and 3 adjacent posterior monokinetids (bipartite) (Fig.

2.5). The oral structure’s right margin is bulged and separates the paroral membrane from

the oral structure (Figs. 2.2B, 2.3). Four adoral organelles (membranelle 1 to 4) originate

from the vestibulum. Membranelle 1 consists of four kinetids and emanates in the upper

anterior end of the slightly depressed vestibulum. Three more adoral organelles

(membranelles 2 to 4), each consisting of four kinetids derived from the vestibulum,

52

comprising two cilia rows each. Two additional dikinetids were located right below the

posterior end of the oral structure, which could not be assigned to any of the six somatic

kineties (Fig. 2.2A). All dikinetids possess two cilia.

TEM observations (twelve individual cells were analyzed) of our collection of A.

marinum show numerous extrusomes (mucocysts) located below the pellicle. These

mucocysts are highly sensitive and partly discharged during cell fixation. Mitochondria

are characterized by tubular cristae (Fig. 2.2D-2.E). The micronucleus and the

macronucleus share an outer membrane of the nuclear envelope; we were not able to

clarify the exact organization of this membrane. The nucleolus is located peripherally and

clearly visible as a dense round structure.

2.4.3. Autecology

A. marinum is a bacterivore with a preference for smaller bacteria (<1 µm, data

not shown). In mixed cultures we did not observe smaller flagellates in the food vacuoles.

Laboratory autecological experiments show that the optimal salinity for cell growth is

between 35 and 40‰, with growth ceasing below 17.5 and above 45‰ salinity. Cell

growth is highest between pH 7 and 8, with growth ceasing below pH 4 and above pH 10.

At a temperature of 28 °C cell growth is highest, but at temperatures below 12 °C and at

37 °C cell growth stops. A. marinum is an obligate aerobe with highest growth rate when

the level of oxygen is 21% in the headspace, with growth ceasing below 1%.

53

2.4.4. Pairwise SSU rDNA sequence differences within the Cyrtolophosididae

The average pairwise distance among the eight Cyrtolophosididae sequences

generated here and from GenBank is 5.413%. The pairwise distance between the two

Aristerostoma spp. collections is 3.99%. The distance between P. alpestris and

HAVOmat-euk43 and LKM63 is 0.65 and 0.889%, respectively.

2.4.5. Genealogical analysis

The SSU-rDNA alignment used for testing the phylogenetic placement of the

Cyrtolophodididae contains 43 sequences, including three new morphospecies sequenced

here: Pseudocyrtolophopsis alpestris and two in the genus Aristerostoma. The alignment

has a length of 1623 unmasked nucleotides, 347 of which are parsimony-informative. The

most parsimonious tree from the MP analysis is 1580 steps in length, with a Consistency

index of 0.48, and a Homoplasy index of 0.52. The most likely tree from the ML analysis

has a lnL of -10080.85. The most likely tree from the Bayesian analysis has a lnL of -

10109.04 (Fig. 2.6).

With the limited outgroups used for this study, the class Colpodea is

monophyletic with weak support from all methods of analysis (57 MP bootstrap/- ML

bootstrap/0.88 Bayesian posterior probability; support less than 50% or 0.5 is shown as

‘-’)—but see Dunthorn et al. (2008). Trees generated from each method have largely

congruent topologies within the class. In the MP tree, Bryometopus pseudochilodon is

basal to its order (plus the order Bursariomorphida), while in the ML and Bayesian trees

Bryometopus sphagni is basal. In the Bayesian tree the chain-forming colpodid+Colpoda

steinii and the order Grossglockneriida form a clade, although not supported, while in the

54

MP and ML analysis these taxa form a polytomy with Colpoda aspera and

Hausmanniela discoidea.

The order Cyrtolophosidida falls out in two groups with moderate to well

supported intervening nodes between them. Cyrtolophosidida I contains all exemplars

from the family Cyrtolophosididae and the included environmental samples available

from GenBank, with strong to full support from all methods (97/100/1.00).

Cyrtolophosidida II contains the remaining exemplars from the order Cyrtolophosidida

with the order Sorogenida embedded within them, and receives full support from all

methods (100/100/1.00).

Relationships among the sampled Cyrtolophosididae are for the most part well

resolved. Both Aristerostoma sequences form a clade with full support from all methods

(100/100/1.00). The genus Aristerostoma is sister the rest of the Cyrtolophosididae with

high to full support from all methods (92/87/1.00). Cyrtolophosis mucicola is in turn

sister to the Pseudocyrtolophosis nana and the environmental samples with low support

from all methods (-/64/0.67).

2.4.6. Comparisons of hypotheses

To compare the morphological hypothesis of Foissner (1993) (where the

Cyrtolophosidida is monophyletic) with that of the SSU-rDNA gene tree estimated here

(where it is not) the likelihood between the alternative hypotheses was examined. The

ML genealogy from the constrained analysis where the Cyrtolophosidida was forced to

me monophyletic has a lnL of -10221.46 (tree not shown). This likelihood values is

140.61 less the non-constrained ML genealogy, and is rejected by the KA test (P value <

55

0.000) in favor of the non-constrained tree. This indicates further support for the non-

monophyly of the Cyrtolophosidida

2.4.7. Long-branch attraction

Nucleotides were partitioned into eight rate classes to test the possibility that the

non-monophyletic topology of the Cyrtolophosidida is a spurious result due to unequal

rates of mutation. If long-branch attraction is effecting the full dataset, then subtracting

the fastest evolving nucleotide sites should remove its effect.

The SSU-rDNA alignment with the fastest rate class removed has 1462 unmasked

nucleotides, 134 of which are parsimony-informative; the most likely tree from the

Bayesian analysis has a lnL of -5717.45 (data not shown). The SSU-rDNA alignment

with the fastest and second fastest rate classes removed has 1256 unmasked nucleotides,

35 of which are parsimony-informative; the most likely tree from the Bayesian analysis

has a lnL of -2769.44 (data not shown). In both of these trees much of the structure of the

topology is lost among and within clades; however, the Cyrtolophosididae does not

branch next to, nor nests within, the other Cyrtolophosidida taxa within either one. These

results support the view that the non-monophyletic topology of the Cyrtolophosidida is

not the result of long-branch attraction.

56

2.5. Discussion

2.5.1. Neotypification and emended diagnosis of Aristerostoma marinum Kahl 1931

1931 Aristerostoma marinum Kahl, Tierwelt Dtl., 21:340

1979 Aristerostoma marinum - Detcheva & Puytorac, Annls. Stn. Limnol. De Besse,

13:247

1993 Aristerostoma marinum Kahl 1931 – Foissner, Protozoenfauna, 4/1:557 (revision)

Reference to the Neotype: Individual specimen marked with a circle on slide 1 at

the collection of microscopic slides of the Biology Center at the Upper Austrian Museum

of Natural History (Linz, Austria), storage code 2007/580-582.

Neotype material: Neotypified from brackish surface waters of the Framvaren

Fjord in southwestern Norway (58°09’ N, 06°55’ E) for the following reasons: (i) no type

material is available and no type location has been defined; (ii) the existing descriptions

are decisively incomplete; (iii) the genus has a proposed subjective junior synonym

(Foissner 1993); (iv) there are several similar species whose identity is threatened by the

species to be neotypified; (v) new preparations (“neotype slides”) are of a quality

allowing the specific features to be clearly recognized. The sample from which we

isolated A. marinum was taken with a Niskin bottle in a depth of ca. 3 m. At the time of

sampling temperature of the surface water was 16 °C, salinity was 15‰ and the water

was saturated with oxygen. Nutrient measurements indicated oligotrophic conditions. The

species was isolated directly from the natural sample without prior enrichment.

57

2.5.2. Emended diagnosis

Size in vivo 9 – 23 x 4 – 11 µm, not contractile, oval shape, tapering anteriorly, 1

spherical-elliposid macronucleus located in the centre of the cell, 1 smaller spherical-

ellipsoid micronucleus, both nuclei share an outer membrane of the nuclear envelope,

extrusomes (mucocysts) hardly recognizable in vivo but impregnate well with protargol,

6 rows of somatic kineties, oral apparatus subapical located anteriorly, oral aperature

triangular, four oral membranelles, bipartite paroral membrane on right slope of

vestibulum, consists of five anterior dikinetids and three uniciliated kinetids (although

they may be dikinetids), 1 contractile vacuole is located posteriorly, no resting stages

observed.

2.5.3. Occurrence and Ecology

The type location is not known, but is assumed to be the German Sea coast near

Hamburg (Foissner 1993). Kahl (1931) found A. marinum in infusions (location not

defined) and only mentions that it is fairly common and sometimes occurs in high

abundances. Detcheva (1982) reported A. marinum from the Bulgarian coast of the Black

Sea where it frequently appeared in high numbers. We here define the Framvaren Fjord

in Norway as neotype location. The salinity range of the locations where the Detcheva

(1982) found A. marinum was 1 to 18‰. The salinity of the surface waters in the

Framvaren Fjord, where we found A. marinum, was 15‰. Interestingly, we did not

observe growth of our A. marinum population in laboratory experiments below 17‰. At

this point we lack an explanation for the discrepancy of the salinity of the populations

natural habitat and the laboratory experiments. In contrast, optimal growth at pH 7-8

58

under laboratory conditions matches the pH of the natural habitat (pH 7.2). Optimum

growth temperature under laboratory conditions was between 20 and 31 °C.

Temperatures above 20 °C are only found in the summer month in the surface waters of

the Framvaren Fjord. It is interesting that the laboratory cultures did not survive

temperatures below 12 °C as the natural habitats temperature may drop below 5 °C in the

winter months (data unpublished). As we did not observe that this ciliate is capable to

form resting stages; thus, it remains an open question how and where these organisms

survive winter temperatures in Norwegian waters.

2.5.4. Comparison with original descriptions and related species

The genus Aristerostoma Kahl, 1926 (order Cyrtolophosidida Foissner, 1978,

family Cyrtolophosididae Stokes, 1888) is diagnosed as very small, laterally flattened and

completely ciliated (Foissner 1993). As pointed out by Foissner (1993), the currently

applied diagnostic characters are based on the incomplete description by Kahl (1931).

The genus is hardly distinguishable by light microscopy from Cyrtolophosis and

Pseudocyrtolophosis, the other two genera within the Cyrtolophosididae. It was

suggested that Pseudocyrtolophosis may in fact be a junior synonym of Aristerostoma, or

both may in turn be a junior synonym of Cyrtolophosis (Foissner 1993). A major reason

for an uncertain species identification is a lack of information about the infraciliature

leading to an obscured general morphology of Aristerostoma (Foissner 1993).

One marine and one freshwater species have been described in Aristerostoma by

Kahl (1926; 1931). Both of the taxa lack type material and modern silver stains. While

our findings contrast the description of A. minutum (habitat, body shape, longitudinal

59

rows, see Foissner 1993) they agree with the rudimentary morphological description of A.

marinum by Kahl (1931) (Table 2.3) and the ultrastructural description (nuclear and oral

structures) of Detcheva and Puytorac (1979) as summarized in Foissner (1993). This is

the position and shape of the macronucleus (spherical, in the centre of the cell) and the

position of the contractile vacuole (near the posterior end). The oral apparatus is in the

anterior third and bordered on right by a paroral membrane. On the opposite side of the

oral apparatus four adoral organelles that are composed of two ciliary rows each.

Our analyses revealed the following additional characters of the oral apparatus: (i)

the first oral membranelle (membranelle 1) consists only of four monokinetids (Fig. 2.4);

(ii) in contrast to former descriptions (see Foissner 1993) the paroral membrane is not

composed of only dikinetids, but of five anterior dikinetids and three posterior

monokinetids and thus, is bipartite (Fig. 2.2b, 2.3) (iii) the bipartite paroral membrane

does not display a gap. Thus, it is very similar but not identical to other

Cyrtolophosididae genera (Fig. 2.4). For example, the gap between the anterior and

posterior segment of the paroral membrane is highly variable in Cyrtolophosis (Fig. 2.4,

see Fig. 215b in Foissner 1993) and in Pseudocyrtolophosis about two thirds of the

paroral membrane is non-ciliated (Fig. 4, Foissner 1993). In Plesiocaryon elongatum (=

Balantiophorus elongatus Schewiakoff, 1892, = Cyrtolophosis elongata (Schewiakoff,

1892) Kahl, 1931) the bipartite paroral membrane consists of an anterior row of six or

seven pairs of kinetosomes whereas the posterior most is a row of five single kinetosomes

both being separated by a distinctive gap (see Fig. 4 and Fig. 1b in Diaz et al. 2002).

Furthermore, the posterior portion of Plesiocaryon terricola is composed of an average of

four widely spaced, barren monokinetids appearing as minute thickenings in vivo (Fig.

60

2.4, Foissner et al. 2002). Because of only rudimentary descriptions of A. minutum, at this

point we are not able to distinguish A. marinum from A. minutum based exclusively on

the oral structure. However, A. marinum is clearly distinct from other genera within the

Cyrtholophosididae regarding the structure of the oral apparatus.

In contrast to Kahl’s population (30 µm), our type material is in average half as

long (15 µm, Tables 2 and 3). Decheva and Puytorac (1979)—who did not provide a

species epithet—give a length of about 20 µm. A cell width is given in neither of these

earlier descriptions. Thus, we add this character to the species diagnosis (9-23 µm in

vivo, 10-20 µm after protargol). In his first description, Kahl pointed out that he was not

able to see the dorsal infraciliature under light microscopy. However, in a schematic

drawing the author shows six longitudinal rows on the right lateral side. Using protargol

impregnation and scanning electron microscopy, we can clearly identify six kinety rows

with their detailed infraciliature (see Figs. 2.2. 2.3). This serves as an additional criterion

for the identification of this species, as all other members of the family Cyrtolophosididae

possess 8-10 somatic kinety rows (Table 2.3). The detailed infraciliature of A. minutum is

unknown but seems to have more longitudinal kineties (at least eleven according to

Kahl’s (1931) schematic drawing).

We confirm the observation of Detcheva and Puytorac (1979) that the

macronucleus and the micronucleus share an outer membrane of the nuclear envelope in

the genus Aristerostoma. While they did not identify their isolate down to species, here

we show that this character state occurs in A. marinum. Since it is not know how

widespread the shared outer membrane is distributed in the Cyrtolophosidida, it is

important that each species be investigated using the SEM (Dunthorn et al. 2008). Like

61

the other species in the genus, we still lack detailed descriptions of division

morphogenesis in A. marinum, which will have to be investigated elsewhere.

2.5.5. Neotype specimens

Two neotype slides with protargol impregnated specimens have been deposited at

collection of microscopic slides of the Biology Center at the Upper Austrian Museum of

Natural History (Linz, Austria) (storage code 2007/580-582). Live cultures are available

from the authors. A circle on slide 1 marks an individual cell that designates the name-

bearing type.

2.5.6. Pairwise distance between Aristerostoma morphospecies

In this study we sequenced two different Aristerostoma populations: one from the

ATCC, which is an unidentified species; and one isolated from the Framvaren Fjord in

Norway, which we identified as A. marinum. As the SSU-rDNA sequences of both taxa

have a pairwise difference of 4.285% they may represent two different species. Thus,

future efforts are in order to characterize the ATCC population in detail. This ATCC

population is unlikely to be the other known species in the genus, A. minutum, which is

from freshwater.

2.5.7. Phylogenetic placement of the order Cyrtolophosidida

Increased taxon sampling using three new morphospecies sequenced here does

not support the monophyly of the order Cyrtolophosidida (Fig. 2.6). Instead, the order

divides into two groups that have distinctive oral morphologies. Genera in the clade

62

Cyrtolophosidida I, containing all Cyrtolophosididae sequences, have a non-ciliated

kinety on the right margin of the adoral organelles and the oral apparatus is located sub-

apically. Genera in the group Cyrtolophosidida II, which contains the remainder of the

order plus the order Sorogenida, lack a non-ciliated kinety and the oral apparatus is

located at the apical pole (Dunthorn et al. 2008; Foissner 1993).

To further test the monophyly of the Cyrtolophosidida, we compared Foissner’s

(1993) morphologically-based hypothesis of the Cyrtolophosidida with the topology of

the SSU-rDNA maximum likelihood tree found here. An ML analysis of the SSU-rDNA

sequences with the non-constrained topology was significantly more likely than the

topology constrained to be monophyletic according the KA test. Like the increased taxon

sampling discussed above, this comparison also does not support the monophyly of the

Cyrtolophosidida.

2.5.8. Reconciling morphology and molecules

Here, we focus on arguments presented by Foissner et al. (2004) to discuss why

there is disagreement between the morphological classification of the Cyrtolophosidida

by Foissner (1993) with that of the SSU-rDNA topology found here.

SSU-rDNA sequences are misidentified. The sequences for Cyrtolophosis from

Dunthorn et al. (2008) could be from another ciliate. However, we argue that we can

reject this argument as the other sequences from the family Cyrtolophosididae here more

closely match the Cyrtolophosis sequences than any other ciliate.

SSU-rDNA sequences are paralogs. Paralogs for protein-coding loci seem to be

rampant within ciliates (e.g., Israel et al. 2002; Katz et al. 2004). However, we argue

63

against this as divergent SSU-rDNA paralogs have yet to be reported within ciliates (or at

least ones in which there has not been concerted evolution).

Insufficient taxon sampling. Insufficient taxon sampling can lead to spurious

results (Cummings and Meyer 2005; Graybeal 1998; Hedtke et al. 2006; Hillis 1998;

Hillis et al. 2003; Poe and Swofford 1999). We argue that we can reject this argument as

we now have eight sequences from the Cyrtolophosididae (representing three of the four

genera in the family) and the same non-monophyletic topology is recovered as in

Dunthorn et al. (2008), which only had two sequences from the same morphospecies.

Long-branch attraction. Heterogeneous rates of evolution among branches can

lead to spurious results (Felsenstein 1978). All methods of analyses suffer from this

problem to one degree or another (Hendy and Penny 1989; Huelsenbeck and Hillis 1993;

Kolaczkowski and Thornton 2004), although the extent of statistical inconsistency in real

datasets is questionable (Anderson and Swofford 2004; Bergsten 2005; Siddall and

Whiting 1999). We argue that we can reject long-branch attraction for our SSU-rDNA

dataset for two reasons. First, a visual inspection of the genealogy does not show any

protruding single or paired long-branches. Second, successively removing the fastest and

second fastest nucleotide sites still produced the same non-monophyletic topology.

Gene tree vs. species tree. The topology of any gene genealogy may not

accurately reflect the actual species phylogeny (Doyle 1992; Doyle 1997; Maddison

1997). Since here we only have a single-gene genealogy, we cannot rule out problems

caused by incomplete lineage sorting of ancestral alleles.

Cyrtolophosidida is truly not monophyletic. The Cyrtolophosidida may actually

be a non-monophyletic group brought together based on the combination of possibly

64

homoplastic characters. If this is indeed the case then we can suggest two scenarios of

morphological evolution: either the members of the Cyrtolophosidida may contain

pleseiomorphic characters (e.g., oral morphology), or there was convergent evolution

along the two branches leading to Cyrtolophosidida I and Cyrtolophosidida II (e.g.,

shared outer membrane of the nuclear envelope). Further analyses of other loci are

needed to test these scenarios.

2.5.9. Assignment of GenBank environmental SSU-rDNA sequences to the

Cyrtolophosididae

In this study we also placed cloned haplotypes from previously published

molecular environmental diversity surveys into a phylogenetic context in the Colpodea. If

we consider the close sequence similarity between HAVOmat-euk43 and LKM63 to P.

alpestris, as well as the branching of these sequences inside a well-supported clade (Fig.

2.6), it is reasonable to assume that the respective organisms may indeed be tentatively

assigned to the genus Pseudocyrtolophosis. Assignment to the exactly which species in

Pseudocyrtolophosis, though, we cannot say since LKM63 originates from a freshwater

lake in the Netherlands (van Hannen et al. 1999) and HAVOmat-euk43 from a Hawaiian

lava cave microbial mat (Brown et al., unpublished), and the morphospecies in

Pseudocyrtolophosis have likewise been isolated from a number of terrestrial

environments (Foissner 1993).

Interestingly, the environmental sequence PAA10AU2004, retrieved from a

French freshwater lake (Lefèvre et al. 2007), cannot be phylogenetically assigned to any

65

of the sequenced genera within the family Cyrtolophosididae. Thus, it is likely that we

are still far away from knowing all genera within this family.

Our analyses of the Colpodea demonstrate that increasing sampling density of

known ciliate taxa in phylogenetic studies is needed in order to phylogenetically place

unidentified environmental clones to known ciliate lineages as seen previously in the

class Plagyopylea (Stoeck et al. 2007). By doing so, we can make predictive hypotheses

of the possible morphology of the ciliates from which the clones are derived, as well as

their possible metabolic and ecological roles in the environment from which they were

sampled. For example, the environmental clones from previous environmental surveys

placed here in the Cyrtolophosididae could be small herbivorous ciliates feeding

primarily on bacteria like other taxa in the family.

2.6. Conclusions

Based on our improved observations of A. marinum it became evident that the

genus Aristerostoma is a distinct taxon within the Cyrtolophosidida and probably not a

junior synonym of other genera it the family, although A. minutum needs to be sampled to

support this. A. marinum can be separated from other taxa in the family based on its

specific infraciliature and oral structure.

Increased taxon sampling within the family Cyrtolophosididae supports an earlier

analysis showing a non-monophyletic topology of the order Cyrtolophosidida.

Cyrtolophosididae again is on a separate branch from the rest of its order. Constrained

analyses comparing the likelihood of the morphological placement with that of the SSU-

66

rDNA placement, as well as removal of the fastest evolving sites, also supports this non-

monophyletic topology. However, sampling of other loci is needed to confirm these

results.

In the monograph of the class Colpodea, Foissner (1993) describes only one

marine species—with the remaining taxa being freshwater or terrestrial. This study,

which places two potentially different species cultured from separate marine

environments, points to the possibility of further Colpodea species waiting to be

uncovered in future observations of marine habitats.

67

2.7. References Anderson, F. E., and D. L. Swofford. 2004. Should we be worried about long-branch

attraction in real data sets? Investigations using metazoan 18S rDNA. Mol. Phylogenet. Evol. 33:440-451.

Bergsten, J. 2005. A review of long-branch attraction. Cladistics 21:163-193. Corliss, J. O. 1961. The ciliated protozoa: Characteriszation, classification, and guide to

the literature. Pergamon Press, London. Corliss, J. O. 1979. The ciliated protozoa: characterization, classification and guide to

the literature. Pergamon Press, Oxford. Cummings, M. P., and A. Meyer. 2005. Magic bullets and golden rules: data sampling

in molecular phylogenetics. Zoology 108:329-336. Detcheva, R. B. 1982. Recherches, en Bulgarie, sur les cilies de certaines plages de la

Mer Niore et des rivieres y aboutissant. Annls. Stn. limnol. Besse 15:231-253. Detcheva, R. B., and P. d. Puytorac. 1979. Un nouvel de cilie a micronoyau inclus dans

l'enveloppe macronucleaire: le genera Aristerostoma Kahl, 1926. Annls. Stn. limnol. Besse 13:247-251.

Didier, P., P. Depuytorac, N. Wilbert, and R. Detcheva. 1980. A propos d'observations

sur l'ultrastructure de cilié Cyrtolophosis mucicola Stokes, 1885. J. Protozool. 27:72-79.

Doyle, J. J. 1992. Gene trees and species trees: molecular systematics as one-character

taxonomy. Syst. Bot. 17:144-163. Doyle, J. J. 1997. Trees within trees: genes and species, molecules and morphology.

Syst. Biol. 46:537-553. Dunthorn, M., W. Foissner, and L. A. Katz. 2008. Molecular phylogenetic analysis of

class Colpodea (phylum Ciliophora) using broad taxon sampling. Mol. Phylogenet. Evol. 48:316-327.

Felsenstein, J. 1978. Cases in which parsimony or compatibility methods will be

positively misleading. Syst. Zool. 27:401-410. Foissner, W. 1978. Das Silberliniensystem und die Infraciliatur der Gattungen

Platyophrya Kahl, 1926, Cyrtolophosis Stokes, 1885 und Colpoda O.F.M., 1786: Ein Beitrag zur Systematik der Colpidida (Ciliata, Vestibulifera). Acta Protozool. 17:215-231.

68

Foissner, W. 1985. Klassifikation und phylogenie der Colpodea (Protozoa: Ciliophora). Arch. Protistenk. 129:239-290.

Foissner, W. 1993. Colpodea (Ciliophora). Protozoenfauna 4/1:i-x, 1-798. Foissner, W. 2002. Neotypification of protists, especially ciliates (Protozoa, Ciliophora).

Bull. Zool. Nom. 59:165-169. Foissner, W., H. Berger, and J. Schaumburg. 1999. Indentification and ecology of

limnetic plankton ciliates. Inf. Bayer. Landesamt Wasserwirtschaft 3/99:1-793. Foissner, W., S. Y. Moon-van der Staay, G. W. M. van der Staay, J. H. P. Hackstein, W.-

D. Krautgartner, and H. Berger. 2004. Reconciling classical and molecular phylogenies in the stichotrichines (Ciliophora, Spirotrichea), including new sequences from some rare species. Europ. J. Protistol. 40:265-281.

Graybeal, A. 1998. Is it better to add taxa or characters to a difficult phylogenetic

problem? Syst. Biol. 47:9-17. Hedtke, S. M., T. M. Townsend, and D. M. Hillis. 2006. Resolution of phylogenetic

conflict in large data sets by increased taxon sampling. Syst. Biol. 55:522-529. Hendy, M. D., and D. Penny. 1989. A framework for the quantitative study of evolution.

Syst. Zool. 38:297-309. Hillis, D. M. 1998. Taxonomic sampling, phylogenetic accuracy, and investigator bias.

Syst. Biol. 47:3-8. Hillis, D. M., D. D. Pollock, J. A. McGuire, and D. J. Zwickl. 2003. Is sparse taxon

sampling a problem for phylogenetic inference? Syst. Biol. 52:124-126. Huelsenbeck, J. P., and D. M. Hillis. 1993. Success of phylogenetic methods in the four-

taxon case. Syst. Biol. 42:247-264. Huelsenbeck, J. P., and F. R. Ronquist. 2003. MrBayes 3: Bayesian phylogenetic

inference under mixed models. Bioinformatics 19:1572-1574. ICZN. 1999. International Code of Zoological Nomenclature, 4th ed. The International

Trust for Zoological Nomenclature, London, U.K. Israel, R. L., S. L. Kosakovsky Pond, S. V. Muse, and L. A. Katz. 2002. Evolution of

duplicated alpha-tubulin genes in ciliates. Evolution 56:1110-1122. Kahl, A. 1926. Neue und wenig bekannte Formen der holotrichen und heterotrichen

Ciliaten. Arch. Protistenk. 55:197-438.

69

Kahl, A. 1931. Urtiere oder Protozoa I: Wimpertiere oder Ciliate (Infusoria) 2. Holotricha außer den im 1. Teil behandelten Prostomata. Tierwelt Dtl. 21:181-398.

Katz, L. A., E. Lasek-Nesselquist, J. Bornstein, and S. V. Muse. 2004. Dramatic

diversity of ciliate histone H4 genes revealed by comparisons of patterns of substitutions and paralog divergences among eukaryotes. Mol. Biol. Evol. 21:555-562.

Kishino, H., and M. Hasegawa. 1989. Evaluation of the maximum likelihood estimate of

the evolutionary tree topologies from DNA sequence data and the branching order in Hominoidea. J. Mol. Evol. 29:170-179.

Kolaczkowski, B., and J. W. Thornton. 2004. Performance of maximum parsimony and

likelihood phylogenetics when evolution is heterogeneous. Nature 431:980-984. Kosakovsky Pond, S. L., S. D. W. Frost, and S. V. Muse. 2005. HyPhy: hypothesis

testing using phylogenies. Bioinformatics 21:676-679. Lefèvre, E., C. Bardot, C. Noël, J. F. Carrias, E. Viscogliosi, C. Amblard, and T. Sime-

Ngando. 2007. Unveiling fungal zooflagellates as members of freshwater picoeukaryotes: evidence from a molecular diversity study in a deep meromictic lake. Environ. Microbiol. 9:61-71.

Lynn, D. H. 1981. The organization and evolution of microtubular organelles in ciliated

protozoa. Biol. Rev. 56:243-292. Maddison, W. P. 1997. Gene trees in species trees. Syst. Biol. 46:523-536. Medlin, L., H. J. Elwood, S. Stickel, and M. L. Sogin. 1988. The characterization of

enzymatically amplified eukaryotes 16S-like ribosomal RNA coding regions. Gene 71:491-500.

Nylander, J. A. 2004. MrModeltest v2. Distrubuted by the author. Evolutionary Biology

Center, Uppsala University, Upsalla. Poe, S., and D. L. Swofford. 1999. Taxon sampling revisited. Nature 398:299-300. Puytorac, P. d., F. Perez-Paniagua, and J. Perez-Silva. 1979. A propos d'observations sur

la stomatogenèse et l'ultrastructure de cilié Woodruffia metabolica (Johnson et Larson, 1938). Protistologica 15:231-243.

Reize, I. B., and M. Melkonian. 1989. A new way to investigate living flagellate/ciliated

cells in the light microscopr: Immobilization of cells in agarose. Bot. Acta 102:145-151.

70

Siddall, M. E., and M. F. Whiting. 1999. Long-branch abstractions. Cladistics 15:9-24. Stetter, K. O., H. Köning, and E. Stachebrandt. 1983. Pyrodictium gen. nov., a new

genus of submarine disc-shaped sulphur reducing archaebacteria growing optimally at 1050C. Appl. Microbiol. 4:535-551.

Stoeck, T., W. Foissner, and D. H. Lynn. 2007. Small-subunit rDNA phylogenies sugest

that Epalxella antiquorum (Penard, 1922) Corliss, 1960 (Ciliophora, Odontostomatida) is a member of the Plagyopylea. J. Eukaryot. Microbiol. 54:436-442.

Stoeck, T., M. V. J. Schwarz, J. Boenigk, M. Schweikert, S. von der Heyden, and A.

Behnke. 2005. Cellular identity of an 18S rRNA gene sequence clade within the class Kinetoplastea: the novel genus Actuariola gen. nov. (Neobodonida) with description of the type species Actuariola framvarensis sp. nov. Int. J. Syst. Evol. Microbiol. 55:2623-2635.

Swofford, D. 2002. PAUP*. Phylogenetic Analysis Using Parsimony (*and Other

Methods). Sinauer Assoc, Sunderland MA. van Hannen, E. J., W. M. Mooij, M. P. van Agterveld, H. J. Gons, and H. J. Laanbroek.

1999. Detritus-dependent development of the microbial community in an experimental system: Qualitative analysis by denaturing gradient gel electrophoresis. Appl. Environ. Microbiol. 65:2478-2484.

71

Table 2.1. Taxon sampling and GenBank numbers used in this study. Newly sequenced taxa are in bold. * = sequences from environmental sampling. Taxon GB# Taxon GB#COLPODEA: Hausmanniella discoidea EU039900Aristerostoma marinum EU264562 Ilsiella palustris EU039901Aristerostoma sp. EU264563 Mykophagophrys terricola EU039902Bardeliella pulchra EU039884 Notoxoma parabryophryides EU039903Bresslaua vorax AF060453 Ottowphrya dragescoi EU039904Bresslauides discoideus EU039885 Platyophrya-like sp. EU039905Bryometopus atypicus EU039886 Platyophrya sp. EU039906Bryometopus pseudochilodon EU039887 Platyophrya vorax AF060454Bryometopus sphagni AF060455 Pseudoplatyophrya nana AF060452Bursaria sp. 1 EU039889 Pseudocyrtolophopsis alpestris EU264564Bursaria sp. 2 A EU039890 Rostrophrya sp. EU039907Bursaria sp. 2 B EU039891 Sagittaria sp. EU039908Bursaria truncatella U82204 Sorogena stoianovitchae AF300285Chain-forming colpodid AY398684 HAVOmet-euk43* EF032797Colpoda aspera EU039892 LKM63* AJ130861Colpoda cucullus EU039893 PAA10AU2004* DQ244029Colpoda inflata M97908Colpoda steinii DQ388599 OUTGROUPS:Colpoda henneguyi EU039894 Coleps hirtus U97109Colpoda lucida EU039895 Furgasonia blochmanni X65150Colpoda magna EU039896 Obertrumia georgiana X65149Colpoda minima EU039897 Orthodonella apohamatus DQ232761Cyrtolophosis mucicola Austria EU039899 Prorodon teres X71140Cyrtolophosis mucicola Brazil EU039898 Pseudomicrothorax dubius X65151

72

Table 2.2. Morphometric data of Atisterostoma marinum population Framvaren Fjord. Data are based on live observations, protargol impregnation, scanning (SEM) and transmission electron microscopy (TEM). AV=arithmetric mean; CV=coefficient of variation [%]; MA=macronucleus, max=maximum value; MI=micronucleus, min=minimum value; SE=standard error; STD=standard deviation. Morphometric

character min max AV STD SE CV n

observed individuals

method

Length [µm] 8.81 22.80 15.28 3.13 0.58 20.51 29 live 10.04 19.84 14.69 1.99 0.29 13.52 47 protargol 8.66 15.43 12.29 1.98 0.41 16.12 23 SEM

Width [µm] 3.62 10.72 7.59 1.79 0.33 23.59 29 live 5.53 10.19 7.80 1.15 0.17 14.78 47 protargol 4.98 9.02 7.14 0.93 0.19 12.96 23 SEM

number MA 1 1 n.a. n .a . n.a. n.a. 45/3 protargol/TEM

number MI 1 1 n .a . n.a. n.a. n.a. 11/4 protargol/TEM

diameter MA [µm]

2.18 4.51 3.28 0.63 0.09 19.09 45 protargol

diameter MI

[µm] 1.13 1.70 1.37 0.17 0.05 12.22 11 protargol

n somatic kineties

6 6 n .a . n.a. n.a. n.a. 23/15 SEM/protargol

n oral

membranelles 4 4 n .a . n.a. n.a. n.a. 5/15 SEM/protargol

n kineties of

paroral membrane

8 8 n .a . n.a. n.a. n.a. 7 SEM

distance MA-posterior end

4.10 9.11 6.70 1.36 0.20 20.37 45 protargol

n.a. = not applicable as characters are invariable

73

Table 2.3. Comparative morphology of described representatives of the four genera Cyrtolophosis, Pseudocyrtolophosis, Plesiocaryon und Aristerostoma within the family Cyrtolophosididae (Ciliophora: Colpodea). DK=dikinetids; MK=monokinetids.

74

Figure 2.1. Light microscopy of living (a, b) and protargol impregnated (c)

Aristerostoma marinum cells. (a) The contractile vacuole (CV) and the oral structure

(OS) are visible during live observation. The photograph shows the left lateral side of the

cell. (b) Cell(s) during perkinetal binary fission. (c) Due to poor impregnability (see text

and Foissner 1993) protargol microphotographs are suboptimal. Nevertheless, extrusomes

(mucocsyst) are clearly visible as dark-colored dots distributed evenly in the cell’s cortex

(arrows). Also the macronucleus (MA) and the micronucleus (MI) are visible. Scale bar

in all microphotographs= 5 µm

75

76

Figure 2.2. Scanning (a-c) and transmission (d, e) electron microscopy of Aristerostoma

marinum. (a) left lateral - dorsal view of the cell displaying the ellipsoid shape tapering

anteriorly, parts of kinety number 4 (white arrows, two dikinetids and two uniciliated

kinetids are visible—see text for discussion) and four of the six kinetids (black arrows)

that seem to belong to sparsely ciliated somatic kineties. Scale bar = 5 µm (b) Details of

the oral apparatus with the bulge (solid white arrow), oral kinetids (black arrow) and the

oral membranelle (dashed white arrow). Scale bar = 0.5 µm (c) Right lateral view with

kineties 1-4 (see schematic drawing Fig. 3). Scale bar = 5 µm. (d) The black arrows point

to discharged mucocysts (extrusomes) and the white arrows to mitochondria with tubular

cristae. Scale bar = 2.5 µm (e) Longitudinal section of the cell showing the size, position

and structure of the nuclear apparatus. The micronucleus (MIC) distinguishes from the

nucleolus (NU) by its higher electron density (darker color). MIC and macronucleus

(MAC) share an outer membrane of the nuclear envelope. However, we were not able to

reveal the detailed structure of the nuclear membranes (arrow). Scale bars a-c and d = 5

µm

77

78

Figure 2.3. Schematic drawing of Aristerostoma marinum (right lateral view) that

combines diagnostic features as revealed by live cell imaging, protargol staining and

scanning electron microscopy. For details see text.

79

80

Figure 2.4. Ecophysiological tolerance limits of Aristerostoma marinum population

Framvaren Fjord. We tested for salinity, pH, temperature and oxygen tolerance.

81

82

Figure 2.5. Schematic drawing of the idealized oral structures (paroral membranes and

adoral organelles) of genera in the family Cyrtolophosididae. Note that there can be much

variation among species within the genera, including the number and size of adoral

organelles. In some species there may be a short oblique kinety anteriorly to the

uppermost adoral organelle.

83

84

Figure 2.6. Most likely Bayesian SSU-rDNA genealogy of the class Colpodea. New

sequences are in bold. Numerical support values are shown as: MP bootstrap/ML

bootstrap/Bayesian posterior probability. Values <50% are shown as ‘-’. Monophyletic

classes and orders are labeled with a solid line, while non-monophyletic ones are labeled

with a dashed line.

85

86

CHAPTER 3

EXPANDING CHARACTER SAMPLING IN CILIATE PHYLOGENETIC

RECONSTRUCTION: MITOCHONDRIAL SSU-RDNA AS A

MOLECULAR MARKER

Micah Dunthorna, Wilhelm Foissnerb, Laura A. Katza,c

aGraduate Program in Organismic and Evolutionary Biology, University of

Massachusetts, Amherst, Massachusetts, USA

bFB Organismische Biologie, Universität Salzburg, Austria

cDepartment of Biological Sciences, Smith College, Northampton, Massachusetts, USA

87

3.1. Abstract

Ciliate molecular systematics has largely focused on increasing taxon sampling

using the nuclear small subunit rDNA (nSSU-rDNA) locus. Previous nSSU-rDNA

analyses have generally been congruent with the morphologically-based classification,

although there is extensive non-monophyly at many levels. Nuclear protein-coding loci

have been shown to be inadequate as independent phylogenetic markers because of

extensive paralogy and extremely heterogenous rates of substitution. Here the

mitochondrial small subunit rDNA (mtSSU-rDNA) is evaluated for deep ciliate nodes

using the Colpodea as an example. Overall, well-supported nodes in the mtSSU-rDNA

toplogy are congruent with well-supported nSSU-rDNA nodes within the Colpodea. The

one incongruence between the loci is the placement of the Sorogenida: nSSU-rDNA nests

the clade within Cyrtolophosidida II, while in the mtSSU-rDNA topology it is basal to

the Cyrtolophosidida II. The placement of the Sorogenida in the concatenated topology is

the same as that of mtSSU-rDNA topology. The mtSSU-rDNA and concatenated

topologies are generally concordant with the classifications in that most morphological

groups are supported. However, several proposed relationships are not supported by

molecular data. This indicates that the morphological characters used in taxonomic

circumscriptions of the Colpodea represent a mixture of ancestral and derived states. The

addition of mtSSU-rDNA as a marker enables phylogenetic reconstructions of the ciliate

tree of life to move from a single-locus effort to a multi-locus approach.

88

3.2. Introduction

Whether it is better to increase the number of sampled taxa or the number of

characters to increase the accuracy of phylogenetic inference is a central debate in

molecular systematics (Graybeal 1998; Hillis 1998; Rannala et al. 1998; Poe and

Swofford 1999; Hillis et al. 2003; Rokas et al. 2003; Cummings and Meyer 2005; Rokas

and Carroll 2005; Hedtke et al. 2006). In general both approaches have their strengths

and weaknesses, and it is advantageous to increase both when reconstructing the tree of

life of any group of organisms. But this has not always been possible in all clades—such

as in ciliates.

Since the discovery of ciliates (Ciliophora Doflein, 1901) by Leeuwenhoek

(Dobell 1932), our understanding of their evolutionary relationships has been improved

by new techniques in visualizing morphological and developmental characters (Lynn

2008). Analyses of these morphological characters have led to several classifications and

monographs for both deep and shallow relationships (Corliss 1979; Foissner 1993;

Puytorac 1994; Berger 1999; Lynn and Small 2002; Berger 2006; Foissner and Xu 2007;

Jankowski 2007; Lynn 2008). Molecular phylogenetic reconstructions to test these

hypothesized deep relationships have relied primarily on expanding taxon sampling while

using the nuclear small subunit ribosomal DNA (nSSU-rDNA) locus (Dunthorn and Katz

2008). Because of this single-locus approach, we do not know if the molecules are

elucidating ciliate morphological evolution or just misleading us.

For deep ciliate nodes, nSSU-rDNA gene trees are concordant with many

morphological hypotheses, but there are a number of discrepancies between (Lynn 2003;

Foissner et al. 2004; Dunthorn and Katz 2008; Lynn 2008). Other molecular markers

89

have been used for further testing of these deep ciliate nodes, but there are difficulties

with them. For instance, the internally transcribed spacer region (ITS) of the rDNA locus

is linked to nSSU-rDNA, and alignment is difficult because of high levels of insertions

and deletions (Snoeyenbos-West et al. 2002). Nuclear protein-coding loci are problematic

because of heterogeneous rates of evolution and extensive paralogy (Tourancheau et al.

1998; Israel et al. 2002; Katz et al. 2004). In contrast, there are more loci available as

molecular markers for shallower ciliate nodes (Morin and Cech 1988; Sadler and Brunk

1992; van Hoek et al. 2000a; Ye and Romero 2002; Barth et al. 2006; Hori et al. 2006;

Lynn and Strüder-Kypke 2006; Przyboś et al. 2006; Chantangsi et al. 2007; Barth et al.

2008; Catania et al. 2008).

One ciliate group in which nSSU-rDNA genealogies based on increased taxon

sampling have been compared to morphological hypotheses is the Colpodea Small &

Lynn, 1981. The Colpodea is diagnosed by a left kinetodesmal fiber (LKm) and unique

silverline patterns (Foissner 1993; Lynn 2008). This primarily terrestrial group contains

diverse oral morphologies and is potentially up to 900 million years old (Wright and

Lynn 1997; Lynn 2008). The almost 200 described species are monographed with an

extensive morphological classification (Foissner 1993). Previous analyses using nSSU-

rDNA have challenged some aspects of this morphologically-based classification (Lynn

et al. 1999; Lasek-Nesselquist and Katz 2001; Dunthorn et al. 2008; Dunthorn et al.

2009). In light of these molecular data, modified hypotheses of morphological evolution

have been proposed (Foissner and Kreutz 1998; Lynn et al. 1999; Lasek-Nesselquist and

Katz 2001; Dunthorn et al. 2008; Dunthorn et al. 2009).

90

Here we move ciliate systematics towards increasing character sampling by

sequencing a broad sample of the Colpodea for the mitochondrial small subunit rDNA

(mtSSU-rDNA). With this character-rich locus, our approach generates an additional 823

unmasked characters for analysis that is not only from an independent locus but also one

from a separate genome within ciliates. Hence, analyzing nSSU-rDNA and mtSSU-

rDNA has to potential to substantially increase our power for interpreting phylogenetic

relationships among ciliates and their morphological evolution.

3.3. Methods

3.3.1. Taxon sampling and terminology

Sequences were obtained from genomic DNA from earlier phylogenetic

reconstructions (Dunthorn et al. 2008; Dunthorn et al. 2009) or from newly isolated

material, as well as from GenBank. In total, our sampling includes 25 isolates from 24

morphospecies (Table 3.1). Of these, 20 are from the Colpodea. Exemplars from five of

the seven orders within the Colpodea as recognized by Foissner (1993) are in included.

Two Paramecium species, two Tetrahymena species, and Chilodonella uncinata are

included as outgroups. Initial analyses included mtSSU-rDNA hydrogenosome sequences

from Armorphorea accessions in GenBank; these were excluded from the final analyses

since they showed extreme rate heterogeneity compared to the rest of the sequences.

When possible, both nSSU-rDNA and mtSSU-rDNA were from the same source DNA.

Terminology follows Foissner (1993) and Lynn (2008). Classification follows Foissner

91

(1993), with the addition of the labeling of Cyrtolophosidida clades 1 and 2 in the tree

topologies following Dunthorn et al. (2008).

3.3.2. DNA amplification and sequencing

Genomic DNA was extracted using the DNeasy Tissue kit (Qiagen, CA).

Mitochondrial SSU-rDNA was amplified with the 5’ primer (TGT GCC AGC AGC CGC

GGT AA) and the 3’ primer (CCC MTA CCR GTA CCT TGT GT) from van Hoek et al.

(2000a). Phusion polymerase (New England BioLabs, MA) was used with the following

cycling conditions: 3:00 at 980; 40 cycles of 0:15 at 980, 0:15 at 670, 1:15 at 720; 10:00

extension at 720. Nuclear SSU-rDNA was amplified following Dunthorn et al. (2008).

Amplified products were cleaned with microCLEAN (The Gel Company, San

Francisco, CA), and cloned with the Zero Blunt TOPO kit (Invitrogen, Carlsbad, CA).

Positive clones were identified by PCR screening with AmpliTag Gold polymerase and

vector primers (Applied Biosystems, Foster City, CA), and minipreped using Qiaprep

Spin Miniprep kit (Qiagen). Clones were sequenced with the Big Dye terminator kit

(Applied Biosystems), using vector primers. Up to eight colonies were sequenced in the

forward direction, and up to five of these were also sequenced in the reverse direction.

Sequences were run on an ABI 3100 automated sequencer.

3.3.3. Genealogical analyses

Haplotypes were determined and edited from overlapping sequence reads in

SeqMan (DNAStar, Inc., Madison, WI). Vector and primer nucleotides were trimmed off.

Pairwise distances for within and among samples were calculated as uncorrected “p”

92

distances in PAUP* v4.0b8 (Swofford, 2002). Haplotypes were aligned using Clustal X

(Thompson et al. 1994), and further edited by eye in MacClade v4.05 (Maddison and

Maddison 2005). Ambiguously aligned regions were masked. For all datasets the GTR-I-

Γ evolutionary model was estimated using AIC in MrModeltest v2 (Nylander 2004).

Maximum parsimony (MP) analyses were carried out in PAUP* v4.0b8 (Swofford 2002),

with all characters equally weighted and unordered. The TBR heuristic search option was

used, running 100 random additions with MulTree option on. Maximum likelihood (ML)

analyses were carried out in RAxML v7.0.4 (Stamatakis et al. 2008). Support for MP and

ML analyses came from 1000 bootstrap replicates using heuristic searches. Bayesian

Inference (BI) analyses were carried out using MrBayes v3.2.1 (Huelsenbeck and

Ronquist 2003), with support coming from posterior probability using four chains and

running 10 million generations. Trees were sampled every 1000 generations. To

determine if the Bayesian analyses were run long enough, output files were examined

using AWTY (Nylander et al. 2008). The first 25% of sampled trees were considered

burn-in trees and were discarded prior to tree reconstruction. A 50% majority rule

consensus of the remaining trees was used to calculate posterior probability. Following

Holder et al. (Holder et al. 2008), this consensus tree is presented.

3.3.4. Data partitioning and congruence testing

Mitochondrial and nuclear datasets were first analyzed separately. To test if they

should be combined, constrained ML analyses were carried out forcing the

Cyrtolophosididae II topologies onto the other loci. For nSSU-rDNA dataset

Platyophrya, Platyophrya-like, Rostrophrya, and Sagittaria were constrained to be

93

monophyletic. In the mtSSU-rDNA dataset Ottowphrya, Platyophrya, and Sorogena were

constrained to be monophyletic. Resulting constrained topologies were then compared to

the non-constrained ML topologies using the AU test as implemented in CONSEL v0.1j

(Shimodaira and Hasegawa 2001).

3.4. Results

3.4.1. Characteristics of gene sequences

Among sequences new here and from GenBank, the mitochondrial SSU-rDNA

primers amplified fragments of variable size and G-C content (Table 3.1). For all

sequences, the average number of base pairs is 1070, with a minimum of 894 in

Chilodonella uncinata and a maximum of 1196 in Colpoda magna. The average G-C

content is 32.9%. Towards the five-prime end there is considerable variation in indel

length (not shown).

Genetic variation in the mtSSU-rDNA locus was not found within a single isolate

of the morphospecies, except in Colpoda henneguyi. The two C. henneguyi sequences are

2.69% different from each other; this same isolate had two different nSSU-rDNA

sequences differing by 0.12% (Dunthorn et al. 2008). The two Cyrtolophosis mucicola

isolates—one from Austria, the other from Brazil—differ by 10.053%; these two differ in

their nSSU-rDNA sequences by 1.71% (Dunthorn et al. 2008). The nSSU-rDNA data

from Dunthorn et al. (2008) and the mtSSU-rDNA here suggest that these two C.

mucicola isolates may represent cryptic species.

94

3.4.2. MtSSU-rDNA analyses

The mtSSU-rDNA alignment includes 823 characters, of which 491 are

parsimony-informative. The MP tree is 2334 in length. The ML tree has a -lnL of

11239.12302. There was little difference in topologies among the three methods for well-

supported nodes, and the ML and BI trees are identical. Here we present the most likely

ML tree with node support from all three methods (Figure 3.1).

In all analyses the Colpodea is monophyletic, with moderate to full node support.

Support for the monophyly of the Colpodea from nSSU-rDNA has been inconsistent in

previous analyses (Lynn et al. 1999; Lasek-Nesselquist and Katz 2001; Dunthorn et al.

2008). However, with the current limited taxon sampling from outgroup lineages,

mtSSU-rDNA does not provide a valid test of monophyly. For those Colpodea taxa

sampled here the mtSSU-rDNA topology for internal relationships is largely congruent

with previous molecular phylogenetic reconstructions using nSSU-rDNA (Lynn et al.

1999; Lasek-Nesselquist and Katz 2001; Dunthorn et al. 2008; Dunthorn et al. 2009).

Mitochondrial SSU-rDNA does not support the monophyly of the

Cyrtolophosidida (Figure 3.1). The Cyrtolophosidida falls into two clades with

moderately supported intervening nodes. Cyrtolophosidida I—including those taxa in the

Cyrtolophosididae—is sister to the Colpodida with no to low node support.

Cyrtolophosidida II—including the rest of the sampled Cyrtolophosidida—is sister to the

rest of the Colpodea with high to full node support. This same non-monophyletic

topology was found with nSSU-rDNA (Dunthorn et al. 2008; Dunthorn et al. 2009).

Using the nSSU-rDNA topology, Dunthorn et al. (2008) suggest that morphologically the

Cyrtolophosidida may represent the ancestral state within the Colpodea. These two

95

Cyrtolophosidida groups also differ in the number of right oral membranes: two in

Cyrtolophosidida I, and one in Cyrtolophosidida II (Dunthorn et al. 2008).

The Sorogenida is monophyletic and is sister to Cyrtolophosidida II with high to

full node support (Figure 3.1). While the monophyly of the Sorogenida is also supported

by nSSU-rDNA, nSSU-rDNA nests Sorogenida within the Cyrtolophosidida II clade

(Dunthorn et al. 2008; Dunthorn et al. 2009). Using the nSSU-rDNA topology, Dunthorn

et al. (2008) suggest that the aerial sorocarp of Sorogena may be a complex derived

character arising from within the Cyrtolophosidida II. The mtSSU-rDNA locus, while

likewise suggesting a close relationship between these two groups, indicates that this

complex feature arose outside the Cyrtolophosidia II. Although additional taxon sampling

of previously unsequenced Cyrtolophosidia II species is needed to determine the position

of these two groups, a close relationship between Sorogenida and Cyrtolophosidida II is

supported in that both having brick-shaped organelles on the left side of the oral structure

as well as pleurotelokinetal stomatogenesis (partial re-organization of parental oral

structures during cell division) (Dunthorn et al. 2008).

In the mtSSU-rDNA topology Bryometopida and the Bursariomorphida are sister

to each other with high to full node support (Figure 3.1), consistent with nSSU-rDNA

topology (Dunthorn et al. 2008; Dunthorn et al. 2009). Using nSSU-rDNA topologies,

Dunthorn et al. (2008), Foissner and Kreutz (1998) and Lynn et al. (1999) note that these

two groups fall sister and do share a number of morphological characters: apical oral

structures, ventral clefts, ardoral organelles that are conspicuous, and cysts with

emergence pores.

96

The Colpodida is monophyletic in the mtSSU-rDNA topology, though with no to

moderate node support (Figure 3.1). Within the Colpodida, Colpoda is not monophyletic

with high to full node support as the Bresslauides discoideus nested within it (Figure

3.1). The genus Colpoda was likewise not monophyletic in an earlier nSSU-rDNA

analysis (Dunthorn et al. 2008). Dunthorn et al. (2008) suggest that Bresslauideus, and

other groups, were split off from Colpoda because of apomorphies (e.g., semicircular

right oral polykinetids) that arose from within the Colpoda clade.

3.4.3. Nuclear SSU-rDNA analyses and topology congruence

To test whether truncated taxon sampling will affect the topology of the

Colpodea, a nSSU-rDNA alignment was made with the same taxon sampling as that in

the mtSSU-rDNA alignment here. This alignment includes 1631 characters, of which 438

are parsimony-informative. The two MP trees are1215 in length. The ML tree has a -lnL

of 8541.68589. The MP, ML and BI trees are identical, except that in the ML tree C.

magna and H. discoidea are sister to each other. Here we present the most likely ML tree

with node support from all three methods of analyses (Supplementary Figure 3.1).

The nSSU-rDNA topology for the Colpodea here is the same as those previously

published analyses based on greater taxon sampling (Dunthorn et al. 2008; Dunthorn et

al. 2009), except for the lower node support for the clade formed by Cyrtolophosidida I

and Colpodida (Supplementary Figure 3.1). The low support for this same clade in the

mtSSU topology (Figure 3.1), may likewise be due to the low taxon sampling here, and

may increase as more taxa are sampled for the mtSSU-rDNA.

97

3.4.4. Concatenated analyses

Overall, the nSSU- and the mtSSU-rDNA topologies are congruent, except in the

placement of the Sorogenida. This overall congruence supports the concatenation of the

two loci for a combined analysis. Furthermore, the nSSU-rDNA unconstrained topology

was not significantly better than the constrained topology in the AU test (p= 0.360).

However, the mtSSU-rDNA unconstrained topology was significantly better than the

constrained topology in the AU test (p= 0.046).

Given the overall congruence between the topologies, and the result of the AU

test for the nSSU-rDNA dataset, a concatenated alignment of nSSU and mtSSU-rDNA

was compiled. This alignment includes 2454 characters, of which 929 are parsimony-

informative. The MP tree is 3573 in length. The ML tree has a -lnL of 20998.62880.

There was little difference in topologies among the three methods for well-supported

nodes, and the ML and BI trees are identical. Here we present the most likely ML tree

with node support from all three methods (Figure 3.2).

The nSSU-rDNA, mtSSU-rDNA, and concatenated topologies are largely

congruent with each other almost all relationships. Unlike the nSSU-rDNA topology here

(Supplementary Figure 3.1) and from previous analyses (Dunthorn et al. 2008;

Dunthorn et al. 2009), the Sorogenida is sister to the Cyrtolophosidida II in both the

mtSSU-rDNA and concatenated topologies (Figures 3.1, 3.2). The clade formed by

Cyrtolophosidida I and Colpodida has low node support in the concatenated topology

(Figure 3.2). This low support may likewise be due to low taxon sampling (see above).

98

3.5. Discussion

3.5.1. Mitochondrial SSU-rDNA as a ciliate molecular marker

Nuclear SSU-rDNA has remained the primary locus for ciliate molecular

phylogenetic reconstructions since it was first sequenced by Sogin and Elwood (1986)

and Lynn and Sogin (1988). Although congruent in many aspects, nuclear SSU-rDNA

topologies have been used to break up or reshuffle large taxa, as well as recognize new

clades (Greenwood et al. 1991; van Hoek et al. 2000b; Lynn and Strüder-Kypke 2002;

Lynn 2003; Strüder-Kypke and Lynn 2003; Affa'a et al. 2004; Gong et al. 2006; Strüder-

Kypke et al. 2006; Stoeck et al. 2007; Lynn 2008; Yi et al. 2008). Nuclear protein-coding

loci have failed, so far, to provide an adequate and independent test of nSSU-rDNA

topologies because of heterogeneous rates of mutation and extensive paralogy

(Tourancheau et al. 1998; Israel et al. 2002; Katz et al. 2004). The resulting reliance on

just nSSU-rDNA locus to reconstruct the ciliate tree of life stands in contrast to the

increasing repertoire of both low- and high-copy loci available for many other microbial

and macrobial eukaryotic clades, as well as the number of loci used to reconstruct

shallow ciliate relationships.

Here we show that the mtSSU-rDNA locus can provide well-supported nodes for

the depths of the ciliate tree of life that were analyzed. Furthermore, most of these nodes

in the individual (Figure 3.1) and concatenated (Figure 3.2) analyses are congruent with

those that are well supported by the nSSU-rDNA locus in previous with previous

molecular phylogenetic reconstructions using nSSU-rDNA based on greater taxon

sampling (Lynn et al. 1999; Lasek-Nesselquist and Katz 2001; Dunthorn et al. 2008;

99

Dunthorn et al. 2009), as well as truncated taxon sampling here (Supplementary Figure

3.1).

The placement of the Sorogenida is the one well-supported incongruence between

the nuclear and mitochondrial loci. In a previous analysis using nSSU-rDNA, as well as

here (Supplementary Figure 3.1), the Sorogenida nests within the Cyrtolophosidida II.

Here, mtSSU-rDNA and the concatenated analyses place the Sorogenidas sister to the

Cyrtolophosidida II (Figures 3.1, 3.2). Effects from low taxon sampling may likely not

explain this congruence since the taxon sampling here is the same for both nSSU- and

mtSSU-rDNA. Differential rates of evolution between the loci or incomplete lineage

sorting are some possible explanations for this incongruence.

The rate of substitution appears to be faster in mtSSU-rDNA than in nSSU-rDNA,

such as in C. henneguyi and C. mucicola (see above). This discrepancy in rates can be

explained by a number of possible factors: effective population size of the mitochondria

vs. nuclei; homogenizing effects on nSSU-rDNA due to meiotic recombination, although

there is debate whether the Colpodea are sexual (Foissner 1993; Dunthorn et al. 2008);

and differential effects of selection among the genomes.

Further work using the mtSSU-rDNA locus is needed to test a number of

taxonomic hypotheses in the Colpodea and in other ciliate groups. One approach, as done

here, is to use genomic DNA previously extracted for nSSU-rDNA analyses, and reuse

them to amplify mtSSU-rDNA. The other is to amplify and sequence both loci each time

new genomic DNA is extracted from new isolates.

100

3.5.2. Morphology vs. molecules in ciliates

The potential problem that individual species trees may not necessarily be

following the species tree affects all organisms (Doyle 1992; Doyle 1997; Maddison

1997). So in ciliates, when there are discrepancies between evidence from morphology

and molecules, morphologists have rightly pointed out that nSSU-rDNA gene trees may

not be an accurate inference of phylogeny (Agatha 2004; Foissner et al. 2004; Schmidt et

al. 2007; Dunthorn et al. 2008). On the other hand, ciliate molecular systematists have

pointed out that they are analyzing not only more characters for phylogenetic

reconstruction, but they have also suggested alternative hypotheses, or re-interpretations,

of morphological evolution given the topology of gene trees (Lynn et al. 1999; Strüder-

Kypke and Lynn 2003; Dunthorn et al. 2008; Dunthorn and Katz 2008). In general, it

appears that morphologically circumscribed groups in ciliates can be based on a mixture

of ancestral, derived, and convergent states, and that molecules are needed to disentangle

them.

Now ciliate molecular systematists can also point to further support from analyses

of mtSSU-rDNA for their interpretations and hypotheses of morphological evolution.

From the work here in the Colpodea, when there are discrepancies between morphology

and nSSU-rDNA molecules, mSSU-rDNA comes down on the side of nSSU-rDNA

molecules and its reinterpretation of morphological evolution. It remains to be seen if

further mitochondrial analyses will likewise support re-interpretations of morphological

evolution within other ciliate clades.

101

3.5.3. Systematic implications

The first molecular phylogenetic reconstruction of the Colpodea by Lynn et al.

(1999) supported some aspects of the morphological classification, while others aspects

were challenged. Subsequent molecular phylogenetic analyses have upheld Lynn et al.

(1999), and challenged additional aspects of the classification (Lasek-Nesselquist and

Katz 2001; Dunthorn et al. 2008; Dunthorn et al. 2009). Nevertheless, the

morphologically-based taxa have remained unchanged (Puytorac 1994; Lynn and Small

2002; Adl et al. 2005; Jankowski 2007; Lynn 2008). The second-locus approach using

mtSSU-rDNA is generally concordant with previous nSSU-rDNA analyses of the

Colpodea, particularly at well-supported nodes. Where there are discrepancies between

morphology and molecules, this second line of evidence—from an independent locus in

another genome—can be used to support a reclassification the Colpodea to reflect

phylogenetic relationships.

3.6. Conclusions

Future molecular phylogenetic reconstructions of ciliate relationships can now use

a two-locus approach, with both nuclear and mitochondrial SSU-rDNA. This increasing

of character sampling will help bring ciliate molecular systematics up to current practices

in other eukaryotic clades where the use of multiple markers is standard. Mitochondrial

SSU-rDNA topologies support previous conclusions about morphological evolution made

in light of nSSU-rDNA studies.

102

3.7. References Adl, S. M., A. G. B. Simpson, M. A. Farmer, R. A. Andersen, O. R. Anderson, J. R.

Barta, S. S. Browser, G. Brugerolle, R. A. Fensome, S. Fredericq, T. Y. James, S. Karpov, P. Kugrens, J. Krug, C. E. Lane, L. A. Lewis, J. Lodge, D. H. Lynn, D. G. Mann, R. M. McCourt, L. Mendoza, Ø. Moestrup, S. E. Mozley-Standridge, T. A. Nerad, C. A. Shearer, A. V. Smirnov, F. W. Spiegel, and M. Taylor. 2005. The new higher level classification of eukaryotes with emphasis on the taxonomy of protists. J. Eukaryot. Microbiol. 52:399-451.

Affa'a, F. M., D. A. Hickey, M. Strüder-Kypke, and D. H. Lynn. 2004. Phylogenetic position of species in the genera Anoplophrya, Plagiotoma, and Nyctotheroides (Phylum Ciliophora), endosymbiotic ciliates of annelids and anurans. J. Eukaryot. Microbiol. 51:301-306.

Agatha, S. 2004. Evolution of ciliary patterns in the oligotrichida (Ciliophora, Spirotricha) and its taxonomic implications. Zoology 107:153-168.

Barth, D., S. Krenek, S. I. Fokin, and T. U. Berendonk. 2006. Intraspecific genetic variation in Paramecium revealed by cytochrome c oxidase I sequences. J. Eukaryot. Microbiol. 53:20-25.

Barth, D., K. Tischer, H. Berger, M. Schlegel, and T. U. Berendonk. 2008. High mitochondrial haplotype diversity of Coleps sp. (Ciliophora: Prostomatida). Environ. Microbiol. 10:626-634.

Berger, H. 1999. Monograph of the Oxytrichidae (Ciliophora, Hypotrichia). Monographiae Biol. 78:i-xii, 1-1080.

Berger, H. 2006. Monograph of the Urostyloidea (Ciliophora, Hypotrichia). Monographiae Biol. 85:i-xv, 1-1303.

Catania, F., F. Wurmser, A. A. Potekhin, E. Przyboś, and M. Lynch. 2008. Genetic diversity in the Paramecium aurelia species complex. Mol. Bio. Evol. 26:421-431.

Chantangsi, C., D. H. Lynn, M. T. Brandl, J. C. Cole, N. Hetrick, and P. Ikonomi. 2007. Barcoding ciliates: a comprehensive study of 75 isolates of the genus Tetrahymena. Int. J. Syst. Evol. Microbiol. 57:2412-2425.

Corliss, J. O. 1979. The ciliated protozoa: characterization, classification and guide to the literature. Pergamon Press, Oxford.

Cummings, M. P., and A. Meyer. 2005. Magic bullets and golden rules: data sampling in molecular phylogenetics. Zoology 108:329-336.

Dobell, C. 1932. Antony van Leeuwenhoek and his "little animals". Dover Publications, Inc., New York.

103

Doyle, J. J. 1992. Gene trees and species trees: molecular systematics as one-character taxonomy. Syst. Bot. 17:144-163.

Doyle, J. J. 1997. Trees within trees: genes and species, molecules and morphology. Syst. Biol. 46:537-553.

Dunthorn, M., M. Eppinger, M. V. J. Schwarz, M. Schweikert, J. Boenigk, L. A. Katz, and T. Stoeck. 2009. Phylogenetic placement of the Cyrtolophosididae Stokes, 1888 (Ciliophora; Colpodea) and neotypification of Aristerostoma marinum Kahl, 1931. Int. J. Syst. Evol. Microbiol. 59:167-180.

Dunthorn, M., W. Foissner, and L. A. Katz. 2008. Molecular phylogenetic analysis of class Colpodea (phylum Ciliophora) using broad taxon sampling. Mol. Phylogenet. Evol. 48:316-327.

Dunthorn, M., and L. A. Katz. 2008. Richness of morphological hypotheses in ciliate systematics allows for detailed assessment of homology and comparisons with gene trees. Denisia 23:389-394.

Foissner, W. 1993. Colpodea (Ciliophora). Protozoenfauna 4/1:i-x, 1-798.

Foissner, W., and M. Kreutz. 1998. Systematic position and phylogenetic relationships of the genera Bursaridium, Paracondylostoma, Thylakidium, Bryometopus, and Bursaria (Ciliophora: Colpodea). Acta Protozool. 37:227-240.

Foissner, W., S. Y. Moon-van der Staay, G. W. M. van der Staay, J. H. P. Hackstein, W.-D. Krautgartner, and H. Berger. 2004. Reconciling classical and molecular phylogenies in the stichotrichines (Ciliophora, Spirotrichea), including new sequences from some rare species. Europ. J. Protistol. 40:265-281.

Foissner, W., and K. Xu. 2007. Monograph of the Spathidiida (Ciliophora: Haptoria), Vol I: Protospathidiidae, Arcuospathidiidae, Apterospathulidae. Monographiae Biol. 81.

Gong, Y., Y. Yu, P. E. V., F. Zhu, and W. Miao. 2006. Reevaluation of he phylogenetic relationship between mobilid and sessilid peritrichs (Ciliophora, Oligohymenophorea) based on small subunit rRNA gene sequences. J. Eukaryot. Microbiol. 53:397-403.

Graybeal, A. 1998. Is it better to add taxa or characters to a difficult phylogenetic problem? Syst. Biol. 47:9-17.

Greenwood, S. J., M. L. Sogin, and D. H. Lynn. 1991. Phylogenetic relationships within the class oligohymenophorea, phylum Ciliophora, inferred from the complete small subunit rRNA gene sequences of Colpidium campylum, Glaucoma chattoni and Opisthonecta henneguyi. J. Mol. Evol. 33:163-174.

104

Hedtke, S. M., T. M. Townsend, and D. M. Hillis. 2006. Resolution of phylogenetic conflict in large data sets by increased taxon sampling. Syst. Biol. 55:522-529.

Hillis, D. M. 1998. Taxonomic sampling, phylogenetic accuracy, and investigator bias. Syst. Biol. 47:3-8.

Hillis, D. M., D. D. Pollock, J. A. McGuire, and D. J. Zwickl. 2003. Is sparse taxon sampling a problem for phylogenetic inference? Syst. Biol. 52:124-126.

Holder, M. T., J. Sukumaran, and P. O. Lewis. 2008. A justification for reporting the majority-rule consensus tree in Bayesian phylogenetics. Syst. Biol. 57:814-821.

Hori, M., I. Tomikawa, E. Przyboś, and M. Fugishima. 2006. Comparison of the evolutionary distances among syngens and sibling species of Paramecium. Molec. Phylogen. Evol. 38:697-704.

Huelsenbeck, J. P., and F. R. Ronquist. 2003. MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics 19:1572-1574.

Israel, R. L., S. L. Kosakovsky Pond, S. V. Muse, and L. A. Katz. 2002. Evolution of duplicated alpha-tubulin genes in ciliates. Evolution 56:1110-1122.

Jankowski, A. W. 2007. Phylum Ciliophora Doflein, 1901. Pp. 415-993 in A. F. Alimov, ed. Protista. Part 2, Handbook of Zoology. Russian Academy of Sciences, Zoological Institure, St. Petersburg.

Katz, L. A., E. Lasek-Nesselquist, J. Bornstein, and S. V. Muse. 2004. Dramatic diversity of ciliate histone H4 genes revealed by comparisons of patterns of substitutions and paralog divergences among eukaryotes. Mol. Biol. Evol. 21:555-562.

Lasek-Nesselquist, E., and L. A. Katz. 2001. Phylogenetic position of Sorogena stoianovitchae and relationships within the class Colpodea (Ciliophora) based on SSU rDNA sequences. J. Eukaryot. Microbiol. 48:604-607.

Lynn, D. H. 2003. Morphology or molecules: how do we identify the major lineages of ciliates (Phylum Ciliophora)? Europ. J. Protistol. 39:356-364.

Lynn, D. H. 2008. The ciliated protozoa: characterization, classification, and guide to the literature, 3rd edition. Springer, Dordrecht.

Lynn, D. H., and E. B. Small. 2002. Phylum Ciliophora Doflein, 1901. Pp. 371-656 in J. J. Lee, G. F. Leedale, and P. C. Bradbury, eds. An illustrated guide to the protozoa, 2nd ed. Society of Protozoologists, Lawrence, Kansas.

Lynn, D. H., and M. L. Sogin. 1988. Assessment of phylogenetic relationships among ciliated protists using partial ribosomal RNA sequences derived from reverse transcripts. Biosystems 21:249-254.

105

Lynn, D. H., and M. Strüder-Kypke. 2002. Phylogenetic position of Licnophora, Lechriopyla, and Schizocaryum, three unusual ciliates (Phylum Ciliophora) endosymbiotic in echinoderms (Phylum echinodermata). J. Eukaryot. Microbiol. 49:460-468.

Lynn, D. H., and M. Strüder-Kypke. 2006. Species of Tetrahymena indentical by small subunut rRNA gene sequences are discriminated by mitochondrial cytochrome c oxidase I gene sequences. J. Eukaryot. Microbiol. 53:385-387.

Lynn, D. H., A. D. G. Wright, M. Schlegel, and W. Foissner. 1999. Phylogenetic relationships of orders within the class Colpodea (phylum Ciliophora) inferred from small subunit rRNA gene sequences. J. Mol. Evol. 48:605-614.

Maddison, W. P. 1997. Gene trees in species trees. Syst. Biol. 46:523-536.

Maddison, W. P., and D. R. Maddison. 2005. MacClade v. 4.0.8. Sinauer Assoc.

Morin, G. B., and T. R. Cech. 1988. Phylogenetic relationships and altered genome structures among Tetrahymena mitochondrial DNAs. Nucleic Acids Res 16:327-346.

Nylander, J. A. 2004. MrModeltest v2. Distrubuted by the author. Evolutionary Biology Center, Uppsala University, Upsalla.

Nylander, J. A., J. C. Wilgenbusch, D. L. Warren, and D. L. Swofford. 2008. AWTY (are we there yet?): a system for graphical exploration of MCMC convergence in Bayesian phyologenetics. Bioinformatics 24:581-583.

Poe, S., and D. L. Swofford. 1999. Taxon sampling revisited. Nature 398:299-300.

Przyboś, E., A. Maciejewska, and B. Skotarczak. 2006. Relationship of species of the Paramecium aurelia complex (Protozoa, ph. Ciliophora, cl. Oligohymenophorea) based on sequences of the histone H4 gene fragment. Folia Biol. (Praha). 54:37-42.

Puytorac, P. d. 1994. Phylum Ciliophora Doflein, 1901. Pp. 1-15 in P. P. de, ed. Traité de Zoologie, Tome II, Infusoires Ciliés, Fasc. 2, Systématique. Masson, Paris.

Rannala, B., J. P. Huelsenbeck, Z. Yang, and R. Neilsen. 1998. Taxon sampling and the accuracy of large phylogenies. Syst. Biol. 47:702-710.

Rokas, A., and S. B. Carroll. 2005. More genes or more taxa? The relative contribution of gene number and taxon number to phylogenetic accuracy. Mol. Biol. Evol. 22:1337-1344.

Rokas, A., B. L. Williams, N. King, and S. B. Carroll. 2003. Genome-scale approaches to resolving incongruence in molecular phylogenies. Nature 425:798-804.

106

Sadler, L. A., and C. F. Brunk. 1992. Phylogenetic relationships and unusual diversity in histone H4 proteins within the Tetrahymena pyriformis complex. Mol. Biol. Evol. 9:70-84.

Schmidt, S. L., D. Bernard, M. Schlegel, and W. Foissner. 2007. Phylogeny of the Stichotrichea (Ciliophora: Spirotrichea) reconstructed with nulcear small subunit rRNA gene sequences: discrepancies and accordances with morphological data. J. Eukaryot. Microbiol. 54:201-209.

Shimodaira, H., and M. Hasegawa. 2001. CONSEL: for assessing the confidence of phylogenetic tree selection. Bioinformatics 17:1246-1247.

Snoeyenbos-West, O. L. O., T. Salcedo, G. B. McManus, and L. A. Katz. 2002. Insights into the diversity of choreotrich and oligotrich ciliates (Class : Spirotrichea) based on genealogical analyses of multiple loci. Int. J. Syst. Evol. Microbiol. 52:1901-1913.

Sogin, M. L., and H. J. Elwood. 1986. Primary structure of the Paramecium tetraurelia small-subunit rRNA coding region: phylogenetic relationships within the Ciliophora. J. Mol. Evol. 23:53-60.

Stamatakis, A., P. Hoover, and J. Rougemont. 2008. A fast boostrap algorithm for the RAxML web-servers. Syst. Biol. 57:758-771.

Stoeck, T., W. Foissner, and D. H. Lynn. 2007. Small-subunit rDNA phylogenies sugest that Epalxella antiquorum (Penard, 1922) Corliss, 1960 (Ciliophora, Odontostomatida) is a member of the Plagyopylea. J. Eukaryot. Microbiol. 54:436-442.

Strüder-Kypke, M., A.-D. G. Wright, W. Foissner, and D. H. Lynn. 2006. Molecular phylogeny of Litostome ciliates (Ciliophora, Litostomatea) with emphasis on free-living haptorian genera. Protist 157:261-278.

Strüder-Kypke, M. C., and D. H. Lynn. 2003. Sequence analyses of the small subunit rRNA gene confirm the paraphyly of oligotrich ciliates sensu lato and support the monophyly of the subclasses Oligotrichia and Choreotrichia (Ciliophora, Spirotrichea). J. Zool. 260:87-97.

Swofford, D. 2002. PAUP*. Phylogenetic Analysis Using Parsimony (*and Other Methods). Sinauer Assoc, Sunderland MA.

Thompson, J. D., D. G. Higgins, and T. J. Gibson. 1994. Clustal W: improving the sensitivity of progressive multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix choice. Nucl. Acids Res. 22:4673-4680.

107

Tourancheau, A. B., E. Villalobo, N. Tsao, A. Torres, and R. E. Pearlman. 1998. Protein coding gene trees in ciliates: comparisons with rRNA-based phylogenies. Mol. Phyl. Evol. 10:299-309.

van Hoek, A. H. A. M., A. S. Akhmanova, M. A. Huynen, and J. H. P. Hackstein. 2000a. A mitochondrial ancestry of the hydrogenosomes of Nyctotherus ovalis. Mol. Biol. Evol. 17:202-206.

van Hoek, A. H. A. M., T. A. van Alen, V. S. I. Sprakel, J. A. M. Leunissen, T. Brigge, G. D. Vogels, and J. H. P. Hackstein. 2000b. Multiple acquisitions of methanogenic Archaeal symbionts by anaerobic ciliates. Mol. Bio. Evol. 17:251-258.

Wright, A.-D. G., and D. H. Lynn. 1997. Maximum ages of ciliate lineages estimated using a small subunit rRNA molecular clock: crown eukaryotes date back to the paleoprotoerozoic. Arch. Protistenk. 148:329-341.

Ye, A., J., and D. P. Romero. 2002. Phylogenetic relationships amongst tetrahymenine ciliates inferred by a comparison of telomerase RNAs. Int. J. Syst. Evol. Microbiol. 52:2297-3202.

Yi, Z., W. Song, A. Warren, D. M. Roberts, K. A. S. Al-Rasheid, Z. Chen, S. A. Al-Farraj, and X. Hu. 2008. A molecular phylogenetic investigation of Pseudoamphisiellla and Parabirojima (Protozoa, Ciliophora, Spirotrichea), two genera with ambiguous systematic positions. Eur. J. Protistol. 44:45-53.

108

Table 4.1. Taxon sampling for mtSSU. GenBank numbers for new sequences are in bold. ForGenBank accessions, measurements were made for only the part of the sequence that is amplified by the primers used here.

mtSSU nucSSUsequence %

length GCTaxon (bp) content GenBank # GenBank #Aristerostoma sp. ATCC #50986 1076 36.95 X EU264563

Bardeliella pulchra 1183 34.5 X EU039884

Bresslauides discoideus 1108 34.71 X EU039885

Bryometopus atypicus 1082 35.49 X EU039886

Bursaria muco 1116 32.77 X EU039889

Bursaria truncatella 1144 32.45 X U82204

Chilodonella unicata 894 25.5 X X

Colpoda aspera 1149 31.77 X EU039892

Colpoda cucullus 1096 31.44 X EU039893

Colpoda henneguyi 1128 31.39 X EU0398941128 31.15 X1

Colpoda lucida 1170 30.46 X EU039895

Colpoda magna 1196 33.68 X EU039896

Cyrtolophosis mucicola Austria 1013 30.89 X EU039899

Cyrtolophosis mucicola Brazil 1015 33.33 X EU039898

Hausmaniella discoideus 1121 32.99 X EU039900

Ottowphrya dragescoi 964 31.33 X EU039904

Paramecium primaurelia 994 34.91 K01750 AF100315

Paramecium tetraurelia 992 34.98 X159172 X03772

Platyophrya sp. 1015 32.89 X EU039906

Platyophrya-like sp. 1048 34.2 X EU039905

Rostrophrya sp. 1055 34.54 X EU039907

Sagittaria sp. 1049 35.66 X EU039908

Sorogena stoianovitchae 1001 33.87 X AF300285

Tetrahymena pyriformis 1039 32.15 AF160864 M98021

Tetrahymena thermophila 1037 30.95 AF396436 X561651Not used in phylogenetic analyses. 2Labeled as Paramecium aurelia in GenBank.

109

Figure 3.1. Mitochondrial SSU-rDNA genealogy of the Colpodea. The most likely

Bayesian tree is shown. Support values are shown as: MP bootstrap/ML bootstrap/BI

posterior probability. Values <50% are shown as ‘-’. Monophyletic classes and orders are

labeled with a solid line, while non-monophyletic ones labeled with a dashed line.

110

111

Figure 3.2. Concatenated nuclear and mitochondrial SSU-rDNA genealogy of the

Colpodea. The most likely Bayesian tree is shown. Support values are shown as: MP

bootstrap/ML bootstrap/BI posterior probability. Values <50% are shown as ‘-’.

Monophyletic classes and orders are labeled with a solid line, while non-monophyletic

ones labeled with a dashed line.

112

113

Supplementary Figure 3.1. Nuclear SSU-rDNA genealogy of the Colpodea. The most

likely Bayesian tree is shown. Support values are shown as: MP bootstrap/ML

bootstrap/BI posterior probability. Values <50% are shown as ‘-’. Monophyletic classes

and orders are labeled with a solid line, while non-monophyletic ones labeled with a

dashed line.

114

115

CHAPTER 4

EXTENSIVE GENETIC DIVERSITY WITHIN HALTERIID CILIATES

Micah Dunthorna, Wilhelm Foissnerb, Laura A. Katza,c

aGraduate Program in Organismic and Evolutionary Biology, University of

Massachusetts, Amherst, Massachusetts, USA

bFB Organismische Biologie, Universität Salzburg, Austria

cDepartment of Biological Sciences, Smith College, Northampton, Massachusetts, USA

116

4.1. Abstract

Previous analyses of two morphospecies suggest that the underlying level of

genetic variation can vary among halteriid ciliates. Here sampling is increased to include

more worldwide isolates with nuclear SSU-rDNA and internally-transcribed spacer (ITS)

region of the SSU-rDNA locus. There is extensive genetic variation in the morphospecies

Halteria grandinella that is consistent with either a large effective population size or

multiple cryptic species. This extensive genetic variation is in contrast to the little genetic

variation in the close related morphospecies Meseres corlissi. Together these data point

out that different ciliate morphospecies can have different underlying genetic patterns and

may not be comparable in biodiversity studies.

117

4.2. Introduction

Most ciliate species are circumscribed using numerous morphological,

ultrastructural, and developmental characters (e.g., Berger 1999; Berger 2006; Foissner

1993; Foissner et al. 2002; Foissner et al. 2005; Lynn 2008). These morphospecies are

used in biodiversity studies analyzing the potential number of ciliates species and their

distributions (Finlay 2002; Finlay et al. 1996; Foissner 1999; Foissner et al. 2008).

Although morphologically circumscribed species are primarily used to understand

ciliate evolution and biodiversity, there is evidence for cryptic species as some

morphospecies that consist of numerous genetically distinct clades (reviewed in Foissner

et al. 2008; Lynn 2008). For example, in Tetrahymena and Paramecium, cryptic species

based on reproductive isolation have been described (Nanney and McCoy 1976;

Sonneborn 1937; Sonneborn 1957; Sonneborn 1975), and ecological differences within

cryptic species complexes have been demonstrated (Weisse 2002). The extent of the

taxonomic distribution of cryptic species in ciliates, though, is not known.

One clade that potentially has cryptic species is common, freshwater halteriid

ciliates. The Halteriidae Claparède and Lachmann, 1858 have two right oral membranes

(an endoral and a reduced paroral membrane), and more than three modified somatic

kineties or bristles (Agatha 2004; Foissner et al. 1991; Foissner et al. 2004; Lynn 2008;

Maeda 1986; Petz and Foissner 1992). Halteriids are model organisms used to test the

debate about microbial distributions because they can be easily found and cultivated

(Foissner et al. 2008; Katz et al. 2005; Weisse et al. 2008).

One recent analysis of halteriids found evidence for potential cryptic species

based on sequence divergences. Using the internally-transcribed spacer (ITS) region,

118

Katz et al. (2005) examined 15 populations of the morphospecies Halteria grandinella.

ITS sequences showed three distinct clades using a 2% cutoff value. In a different

morphospecies using both the small subunit rDNA (SSU-rDNA) and ITS, Weisse et al.

(2008) examined nine populations of Meseres corlissi. While their Chinese isolate is

genetically different by 1% from isolates collected in Australia, Austria, and the

Dominican Republic, no consistent pattern emerged from the morphological and

ecological variation among the populations.

Here, we expand on Katz et al.’s (2005) analysis of the halteriids using SSU-

rDNA and ITS sequencing with more exemplars from Halteria grandinella as well as

other taxa within the halteriids. We further investigate the relationships among the taxa

within the Halteriidae and evaluate the possibility of cryptic species.

4.3. Materials and methods

4.3.1. Taxon sampling and identification

Seven new halteriids were isolated from six countries (Table 4.1). These new

isolates were sequences for both SSU-rDNA and ITS (Table 4.2). Sequencing was also

performed to complement Katz et al. (2005) so that there are both SSU-rDNA and ITS

sequences for most isolates (Table 4.2). However, SSU-rDNA was not recovered from

Katz et al.’s (2005) Halteria sp. from Brazil, and hence its ITS sequence was not

included in the analyses here. Halteria sequences from Katz et al. (2005) used here are:

Halteria grandinella Massachusetts (DQ241751), Halteria grandinella Colorado

(AF508759), Halteria grandinella Florida (DQ241752), Halteria grandinella

119

Massachusetts (AY007444 and DQ241753), Halteria-like Dominican Republic

(DQ241757), Halteria-like Botswana (DQ241755), and Meseres corlissi Brazil

(DQ241754). Meseres sequences used here from Weisse et al. (2008) are: Austria 1

(EU339923), Austria 2 (EU399525), Austria 3 (EU339926), Austria 5 (EU399524),

Austria 6 (EU399527), Australia (EU3399528), China (EU399529), Dominican Republic

(EU399522). Species were identified using standard protocols (Foissner 1991).

4.3.2. Amplification, cloning, and sequencing

For new isolates, between 70 and 100 cells were picked with a micropipette,

washed, and placed into DNA lysis buffer. Genomic DNA was extracted using the

DNeasy Tissue kit (Qiagen, CA). Genomic amplifications used Phusion polymerase

following instructions (New England BioLabs, MA). Primers and cycling conditions to

generate SSU-rDNA sequences followed Dunthorn et al. (2008). Primers and cycling

conditions to generate ITS sequences followed Snoeyenbos-West et al. (2002).

For both Brazil samples, we also amplified SSU-rDNA through the ITS region in

a single product, using the SSU-rDNA 5’ primer of Medlin et al. (1988) and the ITS 3’

primer of Snoeyenbos-West et al. (2002), with the following cycling conditions: 3:00 at

980; 36 cycles of 0:15 at 980, 0:15 at 650, and 2:30 at 720; followed by a 10:00 extension

at 720. We also performed separate amplifications with an annealing temperature 700 to

make sure all different copies were found. To help distinguish between natural and PCR-

mediated chimeras for SSU-rDNA in the Brazil samples, we followed the

recommendations of Judo et al. (1998): during amplification of the genomic DNA we

120

used three times the amount of primers, and used an additional 30 seconds of extension

time.

Amplified products were cloned with Zero Blunt TOPO kit (Invitrogen, CA).

Clones were screened with AmpliTaq Gold polymerase (Applied Biosystems, CA), and

minipreped using Qiaprep Spin Miniprep kit (Qiagen, CA). Positive clones were

sequenced with the Big Dye terminator kit (Applied Biosystems, CA) using vector and

internal primers. Sequences were run on an ABI 3100 automated sequencer.

4.3.3. Genealogical analyses

Unique sequences were constructed from multiple sequence reads and edited in

SeqMan (DNAStar, Inc., Madison, WI). Vector and primer sequences were trimmed off

and polymorphisms confirmed by eye. Haplotypes were scanned for potential chimeras

using Chimeara (Maynard Smith 1992; Posada and Crandall 2001) as implemented in

RDP v2 (Martin and Rybicki 2000) and by eye. Chimeras were excluded from analyses.

Pairwise distances within samples were calculated as uncorrected “p” distances in

PAUP* v4.0b8 (Swofford, 2002). SSU-rDNA and ITS haplotypes were aligned in Clustal

W (Thompson et al. 1994) as implemented in DNAStar. Alignments were further edited

by eye in MacClade v4.05 (Maddison and Maddison 2005). Ambiguously aligned regions

were conservatively masked, and remaining gaps were treated as missing data.

Each locus was analyzed separately, and then concatenated. The GTR+I+G

evolutionary model was estimated for each alignment using AIC in MrModeltest v2

(Nylander 2004). Maximum parsimony (MP) analyses were carried out in PAUP* v4.0b8

(Swofford 2002) running 100 replicates with MulTrees on. Maximum likelihood (ML)

121

analyses were carried out in RAxML (Stamatakis et al. 2008). Support for MP and ML

analyses came from 1000 bootstrap replicates using heuristic searches. Bayesian

Inference (BI) were carried out in MrBayes v3.1 (Huelsenbeck and Ronquist 2003) with

support coming from posterior probability using four chains and running 5 million

generations and sampling every 100. The first 25% of sampled trees were considered

burn-in trees and were discarded prior to tree reconstruction.

Initial analyses including exemplars from all Spirotrichea clades (the potential

outgroups) showed halteriid SSU-rDNA sequences to be monophyletic and nesting

within stichotrich ciliates (data not shown). Furthermore, these broad initial analyses

showed Meseres sequences basal to all other halteriid ciliates (data not shown). Only

halteriida sequences were used in the analyses below with Meseres sequences rooting the

SSU-rDNA, ITS, and concatenated trees.

4.4. Results

4.4.1. Intra-isolate pairwise distances

The SSU-rDNA locus had no intra-isolate variation for any but the two new

Brazilian isolates (Table 4.1). The first Brazilian isolate contained two different

sequences with a pairwise distance of 1.82% (Brazil 1.1, Brazil 1.2). The second

Brazilian isolate also had two different sequences with a distance of 1.67% (Brazil 2.1,

Brazil 2.2). Within the amplifications of the second Brazilian isolate we also found

numerous chimeric sequences (data not shown). Each chimera sequence was found only

once in each of the twelve separate amplifications, while the two main sequences where

122

found in all amplifications. With an increase in primer concentration and extension time,

these chimeras were almost eliminated. Hence, we consider these chimeras to be PCR-

mediated and not included in the analyses.

The ITS locus had no intra-isolate variation for all but the new Brazilian isolates

and the Australian isolate (Table 4.1). The first Brazilian isolate contained two different

sequences with a pairwise distance of 2.9%. The second Brazilian isolate had two

different sequences with a distance of 3.56%. The Australian isolate also had two

different sequences with a distance of 0.19%. In all preliminary analyses these two

sequences formed a clade; therefore, only one is presented here.

4.4.2. Genealogies

The SSU-rDNA alignment includes 1705 characters, of which 56 are parsimony-

informative. The 80 MP trees are 168 in length. The ML tree has a –lnL of 3491.62804.

The BI tree has –lnL of 3487.65229. The ITS alignment for the Halteriidae includes 520

characters, of which 73 are parsimony-informative. The 45 MP trees are 156 in length.

The ML tree has a –lnL of 1631.19274. The BI tree has a –lnL of 1630.86044. The

concatenated SSU-rDNA and ITS alignment includes 2225 characters, of which 129 are

parsimoniously informative. The two MP trees are 330 in length. The ML tree has a –lnL

of 5246.37838. The BI tree has a –lnL of 3255.32580. The most likely BI tree from the

concatenated analysis is shown and discussed below, with node support from MP

bootstrap, ML bootstrap, and BI posterior probabilities (Fig. 4.1). Individual gene trees

are in the supplement (Supplements 4.1-4.2). Overall, well-supported nodes in the SSU-

rDNA, ITS, and combined analyses are congruent.

123

All isolates of M. corlissi are monophyletic with full node support (Fig. 4.1). The

Brazilian M. corlissi isolate characterized here is almost identical to the other M. corlissi

characterized by Weisse et al. (2008). The minimal overall genetic variation found by

Weisse et al. (2008) in the morphospecies M. corlissi is supported here.

In the concatenated tree (Fig. 4.1), the Botswana Halteria-like isolates forms

clade with the Dominican Republic Halteria-like isolate. This clade is in turn sister to the

Halteria grandinella morphospecies with no to low node support. In the SSU-rDNA tree

(Supplement 4.1), the Botswana isolate is sister to all other Halteria-like isolates with no

node support. In the ITS tree (Supplement 4.2), this sequence is sister to the clade that

includes some isolates of Halteria grandinella with moderate ML and BI node support.

The Dominican Republic isolate is sister to one of the sequences from the Brazil isolates

with high to full node support (Fig. 4.1).

The genomic DNA from the two Brazilian populations newly isolated for this

study each contained two different SSU-rDNA and ITS sequences. For both of these,

cells were taken from a communal culture and may have contained more than one

morphospecies. One sequence, Brazil 2.2, was almost exactly like that of the Dominican

Republic isolate. The other SSU-rDNA/ITS sequence (Brazil 2.1) nested within the core

Halteria grandinella clade with no node support (Figs. 4.1). The other Brazil isolate’s

two sequences (Brazil 1.1 and Brazil 1.2) formed a clade in the SSU-rDNA and

concatenated tree with no to moderate node support (Fig. 4.1, Supplement 4.1), but are

paraphyletic in the ITS tree with no node support (Supplement 4.2).

The Halteria grandinella morphospecies isolated by Katz et al. (2005) and those

from GenBank—Colorado, Ecuador, Florida, and Massachusetts—form a clade with the

124

new Peru isolate in all analyses with high to full node support (Fig. 4.1). Some of the

other morphologically similar Halteria grandinella specimens isolated for this study—

Africa, Australia, China, and Venezuela—formed another clade in all analyses with

variable support (Fig. 4.1). In the ITS and concatenated tree these two clades formed a

large clade with each other (and the Brazil 2.1 sequence) with low to no node support

(Fig. 4.1, Supplement 4.2).

4.5. Discussion

4.5.1. Genetic variation underlying morphospecies

In the SSU-rDNA, ITS, and concatenated analyses two distinct clades were

uncovered with full support from all methods: 1) the morphospecies Meseres corlissi

sequences; and 2) the other with Halteria sequences, which includes the morphospecies

Halteria grandinella and Halteria-like species (Fig. 4.1). These two clades are strikingly

different. There is minimal genetic variation within Meseres corlissi. In other words, if

you go out and collect Meseres corlissi from various locations you seem to get only one

genetic entity. In contrast, there is much genetic variation within Halteria grandinella.

Two possible reasons may underlie the extensive genetic variation within

Halteria grandinella morphospecies. The different genetic subclades may represent

cryptic sexual species as sex is known in Halteria grandinella (Agatha and Foissner

2009). Reproductive isolation experiments can be conducted to support this hypothesis.

Alternatively, the different genetic subclades may be the result of a very large effective

population size. Halteria grandinella is easily found in most freshwater environments

125

throughout the world, so census populations are large. Although the data here is

consistent with a large population, we see no evidence of recombination among

sequences as would be expected from a large interbreeding population. Although large

population sizes are often assumed for ciliates (Finlay 2002), molecular support for large

effective population sizes is conflicting (Catania et al. 2008; Katz et al. 2006; Lynch and

Conery 2003; Snoke et al. 2006). Further molecular investigations using protein-coding

loci are needed to test for effective population size.

126

4.6. References

Agatha, S. 2004. A cladistic approach for the classification of oligotrichid ciliates (Ciliophora: Spirotricha). Acta Protozool. 43:201-217.

Agatha, S., and W. Foissner. 2009. Conugation in the spirotrich ciliate Halteria grandinella (Müller, 1773) Dujardin, 1841 (Protozoa, Ciliophora) and its phylogenetic implications. Eur. J. Protistol. 45:51-63.

Berger, H. 1999. Monograph of the Oxytrichidae (Ciliophora, Hypotrichia). Monographiae Biol. 78:i-xii, 1-1080.

Berger, H. 2006. Monograph of the Urostyloidea (Ciliophora, Hypotrichia). Monographiae Biol. 85:i-xv, 1-1303.

Catania, F., F. Wurmser, A. A. Potekhin, E. Przyboś, and M. Lynch. 2008. Genetic diversity in the Paramecium aurelia species complex. Mol. Bio. Evol. 26:421-431.

Dunthorn, M., W. Foissner, and L. A. Katz. 2008. Molecular phylogenetic analysis of class Colpodea (phylum Ciliophora) using broad taxon sampling. Mol. Phylogenet. Evol. 48:316-327.

Finlay, B. J. 2002. Global dispersal of free-living microbial eukaryote species. Science 296:1061-1063.

Finlay, B. J., G. F. Esteban, and T. Fenchel. 1996. Global diversity and body size. Nature 383:132-133.

Foissner, W. 1991. Basic light and scanning electron microscopic methods for taxonomic studies of ciliated protozoa. Europ. J. Protistol. 27:313-330.

Foissner, W. 1993. Colpodea (Ciliophora). Protozoenfauna 4/1:i-x, 1-798.

Foissner, W. 1999. Protist diversity: estimates of the near-imponderable. Protist 150:363-368.

Foissner, W., S. Agatha, and B. H. 2002. Soil ciliates (Protozoa, Ciliophora) from Namibia (Southwest Africa), with emphasis on two contrasting environments, the Etosha Region and the Namib Desert: Part I. Densia 5:1-1063.

Foissner, W., H. Berger, K. Xu, and S. Zechmeister-Boltenstern. 2005. A huge, undescribed soil ciliate (Protozoa: Ciliophora) diversity in natural forest stands of Central Europe. Biodivers. Conserv. 14:617-701.

127

Foissner, W., H. Blatterer, H. Berger, and F. Kohmann. 1991. Taxonomische und ökologische Revision der Ciliaten des Saprobiensystems—Band I: Cyrtophorida, Oligotrichida, Hypotrichia, Colpodea. Informationsberichte Bayer. Landesmat für Wasserwirtschaft, München

Foissner, W., A. Chao, and L. A. Katz. 2008. Diversity and geographic distribution of ciliates (Protista: Ciliophora). Biodivers. Conserv. 17:345-363.

Foissner, W., S. Y. Moon-van der Staay, G. W. M. van der Staay, J. H. P. Hackstein, W.-D. Krautgartner, and H. Berger. 2004. Reconciling classical and molecular phylogenies in the stichotrichines (Ciliophora, Spirotrichea), including new sequences from some rare species. Europ. J. Protistol. 40:265-281.

Huelsenbeck, J. P., and F. R. Ronquist. 2003. MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics 19:1572-1574.

Judo, M. S. B., A. B. Wedel, and C. Wilson. 1998. Stimulation and suppression of PCR-mediated recombination. Nucleic Acids Res. 26:1819-1825.

Katz, L. A., G. B. McManus, O. L. O. Snoeyenbos-West, A. Griffin, K. Pirog, B. Costas, and W. Foissner. 2005. Reframing the 'Everything is everywhere' debate: Evidence for high gene flow and diversity in ciliate morphospecies. Aquat. Microb. Ecol. 41:55-65.

Katz, L. A., O. Snoeyenbos-West, and F. P. Doerder. 2006. Patterns of protein evolution in Tetrahymena thermophila: implications for estimates of effective population size. Mol. Biol. Evol. 23:608-614.

Lynch, M., and J. S. Conery. 2003. The origins of genome complexity. Science 302:1401-1404.

Lynn, D. H. 2008. The ciliated protozoa: characterization, classification, and guide to the literature, 3rd edition. Springer, Dordrecht.

Maddison, W. P., and D. R. Maddison. 2005. MacClade v. 4.0.8. Sinauer Assoc.

Maeda, M. 1986. An illustrated guide to the species of the families Halteriidae and Strobilidiidae (Oligotrichida, Ciliophora), free swimming protozoa common in the aquatic environment

Martin, D., and E. Rybicki. 2000. RDP: detection of recombination amongst aligned sequences. Bioinformatics 16:562-563.

Maynard Smith, J. 1992. Analyzing the mosaic structure of genes. J. Mol. Evol. 34:126-129.

128

Medlin, L., H. J. Elwood, S. Stickel, and M. L. Sogin. 1988. The characterization of enzymatically amplified eukaryotes 16S-like ribosomal RNA coding regions. Gene 71:491-500.

Nanney, D. L., and J. W. McCoy. 1976. Characterization of the species of the Tetrahymena pyriformis complex. Trans. Am. Microsc. Soc. 95:664-682.

Nylander, J. A. 2004. MrModeltest v2. Distrubuted by the author. Evolutionary Biology Center, Uppsala University, Upsalla.

Petz, W., and W. Foissner. 1992. Morphology and morphogenesis of Strobilidium caudatum (Fromentel), Meseres corlissi n. sp., Halteria grandinella (Muller), and Strombidium rehwaldi n. sp., and a proposed phylogenetic system for oligotrich ciliates (Protozoa, Ciliophora). J. Protozool. 39:159-176.

Posada, D., and K. A. Crandall. 2001. Evaluation of methods for detecting recombination from DNA sequences: computer simulation. Proc. Natl. Acad. Sci. USA 98:13757-13762.

Snoeyenbos-West, O. L. O., T. Salcedo, G. B. McManus, and L. A. Katz. 2002. Insights into the diversity of choreotrich and oligotrich ciliates (Class: Spirotrichea) based on genealogical analyses of multiple loci. Int. J. Syst. Evol. Microbiol. 52:1901-1913.

Snoke, M. S., T. U. Berendonk, D. Barth, and M. Lynch. 2006. Large global effective population sizes in Paramecium. Mol. Biol. Evol. 23:2474-2479.

Sonneborn, T. M. 1937. Sex, sex inheritance and sex determination in Paramecium aurelia. Proc. Natl. Acad. Sci. 23:378-383.

Sonneborn, T. M. 1957. Breeding systems, reproductive methods, and species problems in protozoa. Pp. 155-324 in E. Mayr, ed. The species problem. American Association for the Advancement of Science, Washington D. C.

Sonneborn, T. M. 1975. The Paramecium aurelia complex of fourteen sibiling species. Trans. Am. Microsc. Soc. 94:155-178.

Stamatakis, A., P. Hoover, and J. Rougemont. 2008. A fast boostrap algorithm for the RAxML web-servers. Syst. Biol. 57:758-771.

Swofford, D. 2002. PAUP*. Phylogenetic Analysis Using Parsimony (*and Other Methods). Sinauer Assoc, Sunderland MA.

Thompson, J. D., D. G. Higgins, and T. J. Gibson. 1994. Clustal W: improving the sensitivity of progressive multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix choice. Nucl. Acids Res. 22:4673-4680.

129

Weisse, T. 2002. The significance of inter- and intraspecific variation in bacterivorous and herbivorous protists. Antonie Van Leeuwenhoek 81:327-341.

Weisse, T., M. C. Strüder-Kypke, H. Berger, and W. Foissner. 2008. Genetic, morphological, and ecological diversity of spatially separated clones of Meseres corlissi Petz & Foissner, 1992 (Ciliophora, Spirotrichea). J. Eukaryot. Microbiol. 55:257-270.

130

Table 4.1. Isolates newly collected for this study.Designation

Taxon Place of collection, latitude/longitude in analysesHalteria grandinella Foissner 1: Venezuela VenezuelaHalteria grandinella Foissner 2: Brazil, Pantanal (Meseres site) Brazil 1Halteria grandinella Foissner 3: Peru, Late Titicaca PeruHalteria grandinella Foissner 4: China, Pearl River Floodplain ChinaHalteria grandinella Foissner 5: Africa, Kruger National Park AfricaHalteria grandinella Foissner 6: Brazil, Pantanal (Site 1 of Maria & Birgit) Brazil 2Halteria grandinella Foissner 7: Australia, site 1/2006 Australia

131

Table 4.2. Newly sampled loci. Sequencing of clone was performed with just the 5' primer (partial) or with the 3' and internal primers are well (full). Tbd = to be determined.

132

Figure 4.1. Concatenated SSU-rDNA/ITS genealogy of the halteriids. The most likely

Bayesian tree is shown. Support values are shown as: MP bootstrap/ML bootstrap/BI

posterior probability. Values <50% are shown as ‘-’. Monophyletic classes and orders are

labeled with a solid line, while non-monophyletic ones labeled with a dashed line.

133

134

Supplementary Figure 4.1. SSU-rDNA genealogy of the halteriids. The most likely

Bayesian tree is shown. Support values are shown as: MP bootstrap/ML bootstrap/BI

posterior probability. Values <50% are shown as ‘-’. Monophyletic classes and orders are

labeled with a solid line, while non-monophyletic ones labeled with a dashed line.

135

136

Supplementary Figure 4.2. ITS genealogy of the halteriids. The most likely Bayesian

tree is shown. Support values are shown as: MP bootstrap/ML bootstrap/BI posterior

probability. Values <50% are shown as ‘-’. Monophyletic classes and orders are labeled

with a solid line, while non-monophyletic ones labeled with a dashed line.

137

138

CHAPTER 5

RICHNESS OF MORPHOLOGICAL HYPOTHESES IN CILIATE

SYSTEMATICS ALLOWS FOR DETAILED ASSESSMENT OF HOMOLOGY

AND COMPARISONS WITHGENE TREES

Micah Dunthorna and Laura A. Katza,b

aGraduate Program in Organismic and Evolutionary Biology, University of

Massachusetts, Amherst, Massachusetts, 01003, USA

bDepartment of Biological Sciences, Smith Collage, College Road, Northampton,

Massachusetts, 01063, USA

139

5.1. Abstract

Morphological investigations are central to ciliate systematics. Morphology has provided

most species delimitations as well as almost all hypotheses on the ciliate tree of life.

Moreover, emerging analyses of molecular markers are generally concordant with

morphology-based ciliate taxonomies. Despite the richness of morphology-based

hypotheses, there are challenges to ciliate morphological systematics that include the

decreasing numbers of trained morphologists and the difficulty in establishing homology

for some morphological traits. There are also open questions about ciliate morphology,

such as the cause of morphological stasis in cryptic species, and the contrasting pattern of

considerable morphological variation with little underlying genetic variation.

140

5.2. Introduction

Ciliates are unique among microbial groups in that their diverse morphology,

abundance and relatively large sizes have enabled the creation of a comprehensive

morphology-based taxonomy. Analyses of morphological characters, first gathered using

light microscopy and more recently in analyses of electron micrographs, have led to

detailed hypotheses on the relationships among ciliates that extend across taxonomic

levels. Hence, those of us working on the systematics of ciliates find ourselves in the

enviable position of having numerous hypotheses that can be assessed through both

reexamination of morphology and characterization of molecular characters. Here, we

describe the strengths of morphological approaches to ciliate taxonomy, the challenges to

these types of analyses, the concordance between morphological and molecular

characters, and the nature of some of the open questions in ciliate systematics.

5.3. Strengths of Morphology

5.3.1. Species delimitations

The diverse morphology among ciliates has allowed for many in-depth studies

that have defined the limits of ciliate species. In general, the morphological species

concept is the standard for ciliates (e.g., Foissner et al. 2002), although species have been

named using other methods (e.g., Foissner and Berger 1999; Nanney and McCoy 1976;

e.g., Sonneborn 1975). These morphological investigations provide us with an estimate

of the minimal number of extant ciliate species. Although it is argued that the number is

around 3000 by ecologists (Finlay 2002; Finlay et al. 1996), the actual number may be

141

“near- imponderable” (Foissner 1999). Estimates of ciliate species numbers require

highly trained taxonomists exploring new environments and different parts of the Earth:

the more they look, the more they find (e.g., Berger 1999; Berger 2006; Berger et al.

2006; Foissner 1994; Foissner 1995; Foissner 1997a; Foissner 1998; Foissner 2003;

Foissner 2005a; Foissner et al. 2002; Foissner et al. 2005a; Foissner and Stoeck 2006;

Foissner et al. 2003; Foissner and Xu 2007; Foissner et al. 2005c; Kim et al. 2007; Lin et

al. 2007; Ma et al. 2006; Petz et al. 1995). Molecular investigations using environmental

sampling of SSU-rDNA haplotypes also points to such a high number (Doherty et al.

2007; Stoeck et al. 2006).Hence, there are likely many more ciliates that have yet to be

discovered, maybe even up to 30,000 to 40,000 (Chao et al. 2006; Foissner 1997b;

Foissner 1999; Foissner et al. 2008).

5.3.2. Ciliate tree of life

Analyses of morphological characters, including somatic and oral ciliature, and

ontogenesis have generated almost all hypotheses on the topology of the ciliate tree of

life for the most inclusive clades (Corliss 1979; Lynn and Small 1997; Lynn and Small

2002; Puytorac 1994; Small and Lynn 1981; Small and Lynn 1985). Recently,

morphological depictions, along with supporting molecular evidence, have divided

ciliates into two subphyla—the Postciliodematophora and Intramacronucleata—and

eleven classes (Adl et al. 2005; Lynn 2003). Detailed morphological hypotheses have

also generated almost all hypotheses of relationships within these eleven classes (e.g.,

Berger 1999; Berger 2006; Foissner 1993; Foissner and Xu 2007; Lynn and Small 2002;

Matthes et al. 1988).

142

5.4. Challenges to Morphological analyses

5.4.1. Decline in number of trained taxonomists

Like in all eukaryotic clades (Lee 2000; Wheeler 2008), one principal impediment

to our understanding of ciliate diversity is the lack of trained morphological taxonomists.

As much of science in the past century shifted to a focus on model organisms, fewer and

fewer students received training in collection, identification and analysis of diverse

lineages, particularly microorganisms. This problem is particularly acute today as there is

increasing interest in microbial diversity on Earth but few professors positioned to train

students in microbial morphological taxonomy.

5.4.2. Number of characters

While the number of characters needed for phylogenetic analyses is debated

(Gatesy et al. 2007; Rokas et al. 2003), morphological characters are limited, particularly

when compared to molecular characters (Givnish and Sytsma 1997; Hillis and Wiens

2000; Scotland et al. 2003). In ciliates this lack of numerous unambiguous

morphological characters remains problematic, particularly when compared to most

macrobes (e.g., Doyle and Endress 2000; Giribet and Wheeler 2002). In light of this, we

agree with Scotland et al. (2003) that it may be more critical for morphological studies to

investigate fewer characters in depth, such as in the studies of the cysts of Meseres

corlissi (Foissner 2005b; Foissner et al. 2005b; Foissner and Pichler 2006; Foissner et al.

2006).

143

5.4.3. Homology assessment

Likewise, homology assessment of morphological characters can be difficult in all

eukaryotic clades (Scotland et al. 2003), and it is not surprising that this problem is

amplified in microbial groups. We agree that homology is equated with synapomorphy

(de Pinna 1991; Patterson 1982; Stevens 1984) and hence, establishing homology is

essential for inferring evolutionary relationships. Establishing homology is a two-step

process. In a primary homology assessment, similarity among characters is initially

established and shared ancestry is hypothesized. In a secondary homology assessment,

the primary assessment is tested via congruence with other morphological or molecular

characters (de Pinna 1991). Hence, inferring robust phylogenies based requires

independent data sets and reassessment of primary homology statements.

While primary homology statements in ciliates can be relatively straight forward,

establishing secondary homology statements is problematic because of issues in

executing congruence tests. First, cladistic analyses using morphological characters in

ciliates are rare and often deal with few taxa (e.g., Agatha 2004; Agatha and Strüder-

Kypke 2007; Berger and Foissner 1997; Foissner et al. 2007; Puytorac et al. 1994). As a

result, most primary homology statements just have not been tested. Like in all

systematic analyses, there is a difficulty in these few ciliate examples of how many

taxa—both ingroup and outgroup—need to be sampled and coded for; the paper by

Foisser et al. (2007) stands out in increasing outgroups for the problem of placing the

Halteriids.

Second, the question of the level of generality of the homology of many ciliate

morphological characters remains unresolved because most molecular estimates of ciliate

144

relationships rely on single locus, SSU-rDNA (Dunthorn et al. 2008; Hewitt et al. 2003;

Schmidt et al. 2007a; Schmidt et al. 2007b; Shin et al. 2000; Snoeyenbos-West et al.

2004; Snoeyenbos-West et al. 2002; Strüder-Kypke et al. 2007; Strüder-Kypke et al.

2006; Williams and Clamp 2007). With the well-known gene tree versus species tree

problem (Doyle 1997; Maddison 1997), we do not know yet if the SSU-rDNA locus

accurately reflects species phylogeny. For example, the homology of Halteriid oral

membranes with other spirotrichs remains to be satisfactorily answered, although there

are numerous hypotheses (explicit or implicitly implied) and molecular tests (Agatha

2004; Agatha and Strüder-Kypke 2007; Foissner et al. 2004; Foissner et al. 2007; Petz

and Foissner 1992; Strüder-Kypke and Lynn 2003; Szabó 1935). Development of

additional molecular markers is essential for robustly testing homology.

5.5. Concordance with molecular hypotheses

Hypotheses of the ciliate tree of life are generally congruent with gene trees. For

example, the most inclusive clades proposed—ranked at the class level—have, for the

most part, either been supported or at least not refuted by molecules (Lynn 2003).

Molecules also generally support less inclusive ciliate clades. For example, much of

Foissner’s (1993) morphological classification of the class Colpodea is largely congruent

with SSU-rDNA gene trees (Dunthorn et al. 2008; Lasek-Nesselquist and Katz 2001;

Lynn et al. 1999).

One aspect of the incongruence between the morphological hypotheses and the

SSU-rDNA gene trees in the Colpodea centers upon paraphyletic groups; e.g., the

Sorogenida nesting within part of the Cyrtolophosidida, the Bursariomorphida nesting

145

within the Bryometopida, and the Grossglockneriida nesting within the Colpodida

(Dunthorn et al. 2008). In these three cases of paraphyly there are a number of

morphological characters that unite the respective groups within the gene trees (Dunthorn

et al. 2008). Another aspect of the incongruence is the challenge that plesiomorphic

characters pose when trying to uncover evolutionary relationships, where the ancestral

condition of the group remains in some taxa, causing them to be grouped together. An

example of this is in the possibility of the Cyrtolophosidida being polyphyletic (Dunthorn

et al. 2008).

5.6. Open Questions

5.6.1. Lack of morphological variation when there is genetic diversity

While morphology provides us with the minimal number of extant species, there

are undoubtedly many more. Like in other eukaryotic clades (Mayr 1963; Pfenninger and

Schwenk 2007), cryptic species are well-known in ciliates (Sonneborn 1937; Sonneborn

1957). Underlying these ciliate cryptic species there can be both a high genetic diversity

as well as ecological variation (Foissner et al. 2008; Katz et al. 2005; Nanney et al. 1998;

Simon et al. 2008; Weisse and Lettner 2002; Weisse and Rammer 2006).

Two main reasons why there are cryptic species have been postulated: the species

may be nascent, with little time to acquire morphological difference; or the conserved

morphology of the species may be of selective value (Mayr 1976). This second reason is

generally accepted for ciliates (Nanney 1977; Nanney 1982; Nanney 1999; Nanney et al.

1998). This selective reason is supported by the hypothesis that cryptic ciliates may be

146

ancient clades, although the actual age is debated (Nanney 1977; Nanney 1982; Nanney

1999; Van Bell 1985)—but the data on which these ages are set are weak to nonexistent.

There have been no tests, though, of the selective value of keeping the same

morphology among cryptic ciliate species. Equally plausible is that selection is not

occurring at all on morphology and that morphology remains in stasis for other reasons

such as constraints or canalization. Alternatively, the prevalence of cryptic species of

ciliates may be due to disparate rates of morphological and molecular evolution enabled

by the dual nature of ciliate genomes. The separation of genome function between the

unexpressed germline micronucleus and the expressed somatic macronucleus changes the

dynamics of molecular evolution in ciliates as compared to other eukaryotes. Based on

analyses of multiple molecular markers in diverse ciliate, the dual nature of genome

evolution has been shown to be related to elevated rates of protein evolution in this

lineage (McGrath et al. 2006; Zufall et al. 2006). This elevated rate of molecular

evolution, coupled with the prevalence of epigenetics in development (McGrath et al.

2006), may contribute to the generation of cryptic species.

5.6.2. Lack of genetic diversity when there is morphological variation

In contrast to cryptic species, there are also cases in ciliates in which there is only

limited genetic variation, at least as measured by SSU-rDNA divergence, in light of

considerable morphological variation. This phenomenon is best seen in comparisons

among SSU-rDNA gene trees of various clades in the class Spriotrichea. Morphological

and molecular changes are relatively concordant among members of the choreotrichs,

oligotrichs; in contrast, there is considerable discordance and very short SSU-rDNA

147

branches among stichotrich taxa (Agatha and Strüder-Kypke 2007; Snoeyenbos-West et

al. 2002; Strüder-Kypke and Lynn 2003). Intriguingly, it is only within the stichotrichs

that we have evidence of gene scrambling, a process whereby coding domains are

reshuffled in ciliate micronuclei (Ardell et al. 2003; Greslin et al. 1989; Prescott 1992).

We hypothesize that this type of heritable scrambling can cause instant, or at least rapid,

speciation as extensive gene scrambling will disrupt pairing of homologous

chromosomes. Under this scenario, accumulation of scrambled genes within populations

can lead to a barrier to gene exchange with other populations of the same species.

148

5.7. References

Adl, S. M., A. G. B. Simpson, M. A. Farmer, R. A. Andersen, O. R. Anderson, J. R. Barta, S. S. Browser, G. Brugerolle, R. A. Fensome, S. Fredericq, T. Y. James, S. Karpov, P. Kugrens, J. Krug, C. E. Lane, L. A. Lewis, J. Lodge, D. H. Lynn, D. G. Mann, R. M. McCourt, L. Mendoza, Ø. Moestrup, S. E. Mozley-Standridge, T. A. Nerad, C. A. Shearer, A. V. Smirnov, F. W. Spiegel, and M. Taylor. 2005. The new higher level classification of eukaryotes with emphasis on the taxonomy of protists. J. Eukaryot. Microbiol. 52:399-451.

Agatha, S. 2004. A cladistic approach for the classification of oligotrichid ciliates

(Ciliophora: Spirotricha). Acta Protozool. 43:201-217. Agatha, S., and M. Strüder-Kypke. 2007. Phylogeny of the order Choreotrichida

(Ciliophora, Spriotricha, Oligotrichea) as inferred from morphology, ultrastructure, ontogenesis, and SSrRNA sequences. Europ. J. Protistol. 43:37-63.

Ardell, D. H., C. A. Lozupone, and L. F. Landweber. 2003. Polymorphism,

recombination and alternative unscrambling in the DNA polymerase alpha gene of the ciliate Stylonychia lemnae (Alveolata, class Spirotrichea). Genetics 165:1761-1777.

Berger, H. 1999. Monograph of the Oxytrichidae (Ciliophora, Hypotrichia).

Monographiae Biol. 78:i-xii, 1-1080. Berger, H. 2006. Monograph of the Urostyloidea (Ciliophora, Hypotrichia).

Monographiae Biol. 85:i-xv, 1-1303. Berger, H., K. A. S. Al-Rasheid, and W. Foissner. 2006. Morphology and cell division

of Saudithrix terricola n. gen., n. sp., a large, stichotrich ciliate from Saudi Arabia. J. Eukaryot. Microbiol. 53:260-268.

Berger, H., and W. Foissner. 1997. Cladistic relationships and generic characterization

of oxytrichid hypotrichs (Protozoa, Ciliophora). Arch. Protistenk. 148:125-155. Chao, A., P. C. Li, S. Agatha, and W. Foissner. 2006. A statistical approach to estimate

soil ciliate diversity and distribution based on data from five continents. Oikos 114:479-493.

Corliss, J. O. 1979. The ciliated protozoa: characterization, classification and guide to

the literature. Pergamon Press, Oxford. de Pinna, M. G. G. 1991. Concepts and tests of homology in the cladistic paradigm.

Cladistics 7:367-394.

149

Doherty, M., B. Costas, G. B. McManus, and L. A. Katz. 2007. Culture-independent assessment of planktonic ciliate diversity in coastal northwest Atlantic water. Aquat. Microb. Ecol. 48:141-154.

Doyle, J. A., and P. K. Endress. 2000. Morphological phylogenetic analysis of basal

angiosperms: comparison and combination with molecular data. Int. J. Plant Sci. 161:S121-S153.

Doyle, J. J. 1997. Trees within trees: genes and species, molecules and morphology.

Syst. Biol. 46:537-553. Dunthorn, M., W. Foissner, and L. A. Katz. 2008. Molecular phylogenetic analysis of

class Colpodea (phylum Ciliophora) using broad taxon sampling. Mol. Phylogenet. Evol. 48:316-327.

Finlay, B. J. 2002. Global dispersal of free-living microbial eukaryote species. Science

296:1061-1063. Finlay, B. J., G. F. Esteban, and T. Fenchel. 1996. Global diversity and body size.

Nature 383:132-133. Foissner, W. 1993. Colpodea (Ciliophora). Protozoenfauna 4/1:i-x, 1-798. Foissner, W. 1994. Pentahymena corticicola nov. gen., nov. spec., a new colpodid

ciliate (Protozoa, Ciliophora) from bark of acacia trees in Costa Rica. Arch. Protistenk. 144:289-295.

Foissner, W. 1995. Tropical protozoan diversity: 80 ciliate species (Protozoa,

Ciliophora) in a soil sample from a tropical dry forest of Costa Rica, with description of four new genera and seven new species. Arch. Protistenk. 145:37-79.

Foissner, W. 1997a. Faunistic and taxonomic studies on ciliates (Protozoa, Ciliophora)

from clean rivers in Bavaria (Germany), with descriptions of new species and ecological notes. Limnologica 27:179-238.

Foissner, W. 1997b. Global soil ciliate (Protozoa, Ciliophora) diversity: a probability-

based approach using large sample collections from Africa, Australia and Antarctica. Biodivers. Conserv. 6:1627-1638.

Foissner, W. 1998. An updated compilation of world soil ciliates (Protozoa, Ciliophora),

with ecological notes, new recordsm and descriptions of new species. Europ. J. Protistol. 34:195-235.

Foissner, W. 1999. Protist diversity: estimates of the near-imponderable. Protist

150:363-368.

150

Foissner, W. 2003. Pseudomaryna australiensis nov gen., nov spec. and Colpoda brasiliensis nov spec., two new colpodids (Ciliophora, Colpodea) with a mineral envelope. Europ. J. Protistol. 39:199-212.

Foissner, W. 2005a. Two new "flagship" ciliates (Protozoa, Ciliophora) from

Venezuela: Sleighophrys pustulata and Luporinophrys micelae. Europ. J. Protistol. 41:99-117.

Foissner, W. 2005b. The unusual, lepidosome-coated resting cyst of Meseres corlissi

(Ciliophora: Oligotrichea): Transmission electron microscopy and phylogeny. Acta Protozool. 44:217-230.

Foissner, W., S. Agatha, and B. H. 2002. Soil ciliates (Protozoa, Ciliophora) from

Namibia (Southwest Africa), with emphasis on two contrasting environments, the Etosha Region and the Namib Desert: Part I. Densia 5:1-1063.

Foissner, W., and H. Berger. 1999. Identification and ontogenesis of the nomen nudum

hypotrichs (Protozoa: Ciliophora) Oxytricha nova (= Sterkiella nova sp. n.) and O. trifallax (= S. histriomuscorum). Acta Protozool. 38:215-248.

Foissner, W., H. Berger, K. Xu, and S. Zechmeister-Boltenstern. 2005a. A huge,

undescribed soil ciliate (Protozoa: Ciliophora) diversity in natural forest stands of Central Europe. Biodivers. Conserv. 14:617-701.

Foissner, W., A. Chao, and L. A. Katz. 2008. Diversity and geographic distribution of

ciliates (Protista: Ciliophora). Biodivers. Conserv. 17:345-363. Foissner, W., S. Y. Moon-van der Staay, G. W. M. van der Staay, J. H. P. Hackstein, W.-

D. Krautgartner, and H. Berger. 2004. Reconciling classical and molecular phylogenies in the stichotrichines (Ciliophora, Spirotrichea), including new sequences from some rare species. Europ. J. Protistol. 40:265-281.

Foissner, W., H. Müller, and S. Agatha. 2007. A comparative fine structural and

phylogenetic analysis of resting cysts in oligotrich and hypotrich Spirotrichea (Ciliophora). Europ. J. Protistol. 43:295-314.

Foissner, W., H. Müller, and T. Weisse. 2005b. The unusual, lepidosome-coated resting

cyst of Meseres corlissi (Ciliophora: Oligotrichea): Light and scanning electron microscopy, cytochemistry. Acta Protozool. 44:201-215.

Foissner, W., and M. Pichler. 2006. The unusual, lepidosome-coated resting cyst of

Meseres corlissi (Ciliophora: Oligotrichea): Genesis of four complex types of wall precursors and assemblage of the cyst wall. Acta Protozool. 45:339-366.

151

Foissner, W., M. Pichler, K. A. S. Al-Rasheid, and T. Weisse. 2006. The unusual, lepidosome-coated resting cyst of Meseres corlissi (Ciliophora: Oligotrichea): Encystment and genesis and release of the lepidosome. Acta Protozool. 45:323-338.

Foissner, W., and T. Stoeck. 2006. Rigidothrix goiseri nov. gen., nov. spec.

(Rigidotrichidae nov. fam.), a new "flagship" ciliate from the Niger floodplain breaks the flexibility-dogma in the classification of stichotrichine spirotrichs (Ciliophora, Spirotrichea). Europ. J. Protistol. 42:249-267.

Foissner, W., M. Struder-Kypke, G. W. M. van der Staay, S. Y. Moon-van der Staay, and

J. H. P. Hackstein. 2003. Endemic ciliates (Protozoa, Ciliophora) from tank bromeliads (Bromeliaceae): a combined morphological, molecular, and ecological study. Europ. J. Protistol. 39:365-372.

Foissner, W., and K. Xu. 2007. Monograph of the Spathidiida (Ciliophora: Haptoria),

Vol I: Protospathidiidae, Arcuospathidiidae, Apterospathulidae. Monographiae Biol. 81

Foissner, W., K. Xu, and M. Kreutz. 2005c. The Apertospathulidae, a new family of

Haptorid ciliates (Protozoa, Ciliophora). J. Eukaryot. Microbiol. 52:360-373. Gatesy, J., R. DeSalle, and N. Wahlberg. 2007. How many genes should a systematist

sample? Conflicting insights from a phylogenomic matrix characterized by replicated incongruence. Syst. Biol. 56:355-363.

Giribet, G., and W. Wheeler. 2002. On bivalve phylogeny: a high-level analysis of the

Bivalvia (Mollusca) based on combined morphology and DNA sequence data. Invert. Biol. 121:271-324.

Givnish, T. J., and K. J. Sytsma. 1997. Consistency, characters, and the likelihood of

correct phylogenetic inference. Mol. Phylogenet. Evol. 7:320-330. Greslin, A. F., D. M. Prescott, Y. Oka, S. H. Loukin, and J. C. Chappell. 1989.

Reordering of 9 exons is necessary to form a functional actin gene in Oxytricha nova. Proc. Natl. Acad. Sci. USA 86:6264-6268.

Hewitt, E. A., K. M. Müller, J. Cannone, D. J. Hoga, R. Gutell, and D. M. Prescott.

2003. Phylogenetic relationships among 28 spirotrichous ciliates documented by rDNA. Mol. Phylogenet. Evol. 29:258-267.

Hillis, D. M., and J. J. Wiens. 2000. Molecules versus morphology in systematics:

conflicts, artifacts, and misconceptions in J. J. Wiens, ed. Phylogenetic analysis of morphological data. Smithsonian Institution Press, Washington.

152

Katz, L. A., G. B. McManus, O. L. O. Snoeyenbos-West, A. Griffin, K. Pirog, B. Costas, and W. Foissner. 2005. Reframing the 'Everything is everywhere' debate: Evidence for high gene flow and diversity in ciliate morphospecies. Aquat. Microb. Ecol. 41:55-65.

Kim, J. S., H. J. Jeong, D. H. Lynn, J. Y. Park, Y. W. Lim, and W. Shin. 2007. Balanion

masanensis n. sp. (Ciliophora: Prostomatea) from the coastal waters of Korea: Morphology and small subunit ribosomal RNA gene sequence. J. Eukaryot. Microbiol. 54:482-494.

Lasek-Nesselquist, E., and L. A. Katz. 2001. Phylogenetic position of Sorogena

stoianovitchae and relationships within the class Colpodea (Ciliophora) based on SSU rDNA sequences. J. Eukaryot. Microbiol. 48:604-607.

Lee, M. S. Y. 2000. A worrying systematic decline. Trends Ecol. Evol. 15:346. Lin, X., W. Song, and J. Li. 2007. Description of two new marine pleurostomatid

ciliates, Loxophyllum choii nov. spec. and L. shini nov spec. (Ciliophora, Pleurostomatida) from China. Europ. J. Protistol. 43:131-139.

Lynn, D. H. 2003. Morphology or molecules: how do we identify the major lineages of

ciliates (Phylum Ciliophora)? Europ. J. Protistol. 39:356-364. Lynn, D. H., and E. B. Small. 1997. A revised classification of the phylum Ciliophora

Doflein, 1901. Rev. Soc. Mex. Hist. Nat. 47:65-78. Lynn, D. H., and E. B. Small. 2002. Phylum Ciliophora Doflein, 1901. Pp. 371-656 in J.

J. Lee, G. F. Leedale and P. C. Bradbury, eds. An illustrated guide to the protozoa, 2nd ed. Society of Protozoologists, Lawrence, Kansas.

Lynn, D. H., A. D. G. Wright, M. Schlegel, and W. Foissner. 1999. Phylogenetic

relationships of orders within the class Colpodea (phylum Ciliophora) inferred from small subunit rRNA gene sequences. J. Mol. Evol. 48:605-614.

Ma, H. W., R. M. Overstreet, J. H. Sniezek, M. Solangi, and D. W. Coats. 2006. Two

new species of symbiotic ciliates from the respiratory tract of cetaceans with establishment of the new genus Planilamina n. gen. (Dysteriida, Kyaroikeidae). J. Eukaryot. Microbiol. 53:407-419.

Maddison, W. P. 1997. Gene trees in species trees. Syst. Biol. 46:523-536. Matthes, D., W. Guhl, and G. Haider. 1988. Suctoria und Urceolariidae. Protozoenfauna

7/1:i-ix, 1-309. Mayr, E. 1963. Animal species and evolution. The Belknap Press of Harvard University

Press, Cambridge.

153

Mayr, E. 1976. Evolution and the diversity of life. The Belknap Press of Havard University Press, Cambridge.

McGrath, C. L., R. A. Zufall, and L. A. Katz. 2006. Genome evolution in ciliates. Pp.

64-77 in L. A. Katz and D. Bhattacharya, eds. Genomics and evolution of eukaryotic microbes. Oxford Uniuversity Press, New York.

Nanney, D. L. 1977. Molecules and morphospecies: the perpetuation of pattern in

ciliates protozoa. J. Protozool. 24:27-35. Nanney, D. L. 1982. Genes and phenes in Tetrahymena. Bioscience 32:783-788. Nanney, D. L. 1999. When is a rose? the kinds of Tetrahymena. Pp. 93-118 in R. A.

Wilson, ed. Specie: new interdisciplinary essays. The MIT Press, Cambridge. Nanney, D. L., and J. W. McCoy. 1976. Characterization of the species of the

Tetrahymena pyriformis complex. Trans. Am. Microsc. Soc. 95:664-682. Nanney, D. L., C. Park, R. Preparata, and E. M. Simon. 1998. Comparison of sequence

differences in a variable 23S rRNA domain among sets of cryptic species of ciliated protozoa. J. Euk. Microbiol. 45:91-100.

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, London.

Petz, W., and W. Foissner. 1992. Morphology and morphogenesis of Strobilidium

caudatum (Fromentel), Meseres corlissi n. sp., Halteria grandinella (Muller), and Strombidium rehwaldi n. sp., and a proposed phylogenetic system for oligotrich ciliates (Protozoa, Ciliophora). J. Protozool. 39:159-176.

Petz, W., W. Song, and N. Wilbert. 1995. Taxonomy and ecology of the ciliate fauna

(Protozoa: Ciliophora) in the endopagial and pelagial of the Weddell Sea, Antarctica. Staphia 4:1-223.

Pfenninger, M., and K. Schwenk. 2007. Cryptic animal species are homogeneously

distributed among taxa and biogeographical regions. BMC Evol. Biol. 7:121. Prescott, D. M. 1992. The unusual organization and processing of genomic DNA in

hypotrichous ciliates. Trends Genet. 8:439-445. Puytorac, P. d. 1994. Phylum Ciliophora Doflein, 1901. Pp. 1-15 in P. P. de, ed. Traité

de Zoologie, Tome II, Infusoires Ciliés, Fasc. 2, Systématique. Masson, Paris.

154

Puytorac, P. d., J. Grain, and P. Legendre. 1994. An attempt at reconstructing a phylogenetic tree of the ciliophora using parsimony methods. Europ. J. Protistol. 30:1-17.

Rokas, A., B. L. Williams, N. King, and S. B. Carroll. 2003. Genome-scale approaches

to resolving incongruence in molecular phylogenies. Nature 425:798-804. Schmidt, S. L., D. Bernard, M. Schlegel, and W. Foissner. 2007a. Phylogeny of the

Stichotrichea (Ciliophora: Spirotrichea) reconstructed with nulcear small subunit rRNA gene sequences: discrepancies and accordances with morphological data. J. Eukaryot. Microbiol. 54:201-209.

Schmidt, S. L., W. Foissner, M. Schlegel, and D. Bernhard. 2007b. Molecular

phylogeny of the heterotrichea (Ciliophora, postciliodesmatophora) based on small subunit rRNA gene sequences. J. Eukaryot. Microbiol. 54:358-363.

Scotland, R. W., R. G. Olmstead, and J. R. Bennett. 2003. Phylogeny reconstruction: the

role of morphology. Syst. Biol. 52:539-548. Shin, M. K., U. W. Hwang, W. Kim, A.-D. G. Wright, C. Krawczyk, and D. H. Lynn.

2000. Phylogenetic position of the ciliates Phacodinium (Order Phacodiniida) and Protocruzia (Subclass Protocruziidia) and systematics of the spirotrich ciliates examined by small subunit ribosomal RNA gene sequences. Europ. J. Protistol. 36:293-302.

Simon, E. M., D. L. Nanney, and F. P. Doerder. 2008. The "Tetrahymena pyformis"

complex of cryptic species. Biodivers. Conserv. 17:365-380. Small, E. B., and D. H. Lynn. 1981. A new macrosystem for the phylum Ciliophora

Doflein, 1901. Biosystems 14:387-401. Small, E. B., and D. H. Lynn. 1985. Phylum Ciliophora, Doflein 1901 in J. J. Lee, S. H.

Hutner and E. C. Bovee, eds. An illustrated guide to the protozoa. Society of Protozoologists, Lawrence, KS.

Snoeyenbos-West, O. L. O., J. Cole, A. Campbell, D. W. Coats, and L. A. Katz. 2004.

Molecular phylogeny of phyllopharyngean ciliates and their group I introns. J. Eukaryot. Microbiol. 51:441-450.

Snoeyenbos-West, O. L. O., T. Salcedo, G. B. McManus, and L. A. Katz. 2002. Insights

into the diversity of choreotrich and oligotrich ciliates (Class: Spirotrichea) based on genealogical analyses of multiple loci. Int. J. Syst. Evol. Microbiol. 52:1901-1913.

Sonneborn, T. M. 1937. Sex, sex inheritance and sex determination in Paramecium

aurelia. Proc. Natl. Acad. Sci. 23:378-383.

155

Sonneborn, T. M. 1957. Breeding systems, reproductive methods, and species problems in protozoa. Pp. 155-324 in E. Mayr, ed. The species problem. American Association for the Advancement of Science, Washington D. C.

Sonneborn, T. M. 1975. The Paramecium aurelia complex of fourteen sibiling species.

Trans. Am. Microsc. Soc. 94:155-178. Stevens, P. F. 1984. Homology and phylogeny: morphology and systematics. Syst. Bot.

9:395-409. Stoeck, T., B. Hayward, G. T. Taylor, R. Varela, and S. S. Epstein. 2006. A multiple

PCR-primer approach to access the microeukaryotic diversity in environmental samples. Protist 157:31-43.

Strüder-Kypke, M., O. A. Kornilova, and D. H. Lynn. 2007. Phylogeny of trichostome

ciliates (Ciliophora, Litostomatea) endosymbiotic in the Yakut horse (Equus caballus). Europ. J. Protistol. 43:319-328.

Strüder-Kypke, M., A.-D. G. Wright, W. Foissner, and D. H. Lynn. 2006. Molecular

phylogeny of Litostome ciliates (Ciliophora, Litostomatea) with emphasis on free-living haptorian genera. Protist 157:261-278.

Strüder-Kypke, M. C., and D. H. Lynn. 2003. Sequence analyses of the small subunit

rRNA gene confirm the paraphyly of oligotrich ciliates sensu lato and support the monophyly of the subclasses Oligotrichia and Choreotrichia (Ciliophora, Spirotrichea). J. Zool. 260:87-97.

Szabó, M. 1935. Neuere Beiträge zur Kenntnis der Gattung Halteria (Protozoa, Ciliata).

Arch. Protistenk. 86:307-317. Van Bell, C. T. 1985. The 5s and 5.8s ribosomal RNA sequences of Tetrahymena

thermophila and T. pyriformis. J. Protozool. 32:640-644. Weisse, T., and S. Lettner. 2002. The ecological significance of intraspecific variation

among freshwater ciliates. Verh. Internat. Verein. Limnol. 28:1880-1884. Weisse, T., and S. Rammer. 2006. Pronounced ecophysiological clonal differences of

two common freshwater ciliates, Coleps spetai (Prostomatida) and Rimostrombidium lacustris (Oligotrichida), challenge the morphospecies concept. J. Plankton Res. 28:55-63.

Wheeler, Q. D. 2008. Undisciplined thinking: morphology and Hennig's unfinished

revolution. Syst. Entomol. 33:2-7.

156

Williams, D., and J. C. Clamp. 2007. A molecular phylogenetic investigation of Opisthonecta and related genera (Ciliophora, Peritrichia, Sessilida). J. Eukaryot. Microbiol. 54:317-323.

Zufall, R. A., C. L. McGrath, S. V. Muse, and L. A. Katz. 2006. Genome architecture

drives protein evolution in ciliates. Mol. Biol. Evol. 23:1681-1687.

157

CHAPTER 6

ANCIENT ASEXUAL LINEAGES OF MICROBIAL EUKARYOTES THAT

REGAINED SEX?

Micah Dunthorna and Laura A. Katza,b

aGraduate Program in Organismic and Evolutionary Biology, University of

Massachusetts Amherst, USA

bDepartment of Biological Sciences, Smith College, USA

158

6.1. Abstract

Sex in ciliates occurs through mutual exchange of haploid nuclei during conjugation, a

process that is decoupled from cell division. Among ciliates in the clade Colpodea only

Bursaria truncatella is known to have sex. Here we review the evidence for and against

sexuality in the rest of the Colpodea. We discuss expectations of sexuality in light of the

ancient age of the Colpodea and the problem of reversing the loss of sex in B. truncatella.

Based on these arguments, we suggest that many, if not all, of the Colpodea may be

sexual. These expectations and arguments, though, are derived from theories and

observations from macrobes, and may not apply to microbial eukaryotes such as ciliates.

159

6.2. Introduction

Sex (amphimixis) in the ancestor of extant eukaryotes was likely facultative

(Dacks and Roger 1999). While most taxa have remained sexual, asexual lineages are

found scattered throughout the eukaryotic tree of life, primarily at the tips (Bell 1982).

The pattern where most eukaryotes are sexual has been explained by theories on the

advantages of the maintenance of sex; i.e., the Red Queen, Muller’s ratchet, and others

(Arkhipova and Meselson 2004; Bell 1982; Bell 1988; Burt 2000; Fischer 1930;

Hamilton 2001; Kondrashov 1982; Kondrashov 1993; Lynch et al. 1993; Maynard Smith

1978; Muller 1964; West et al. 1999; Williams 1975).

Ciliates—a clade of microbial eukaryotes—have remained facultatively sexual.

Sex in ciliates occurs during conjugation, where haploid nuclei are mutually exchanged

between complementary cells (Dini and Nyberg 1993; Lynn 2008; Sonneborn 1957).

These nuclei fuse to make a zygotic nucleus that mitotically divides to give rise to a

“germline” micronucleus and a “somatic” macronucleus. The micronucleus can divide

meiotically to produce the haploid nuclei that take part in conjugation (Lynn 2008;

Raikov 1996). Sex is assumed to occur in almost all ciliate clades, although details and

direct observations for most species is lacking. There are known derived asexual strains

that have lost their micronuclei and are thus unable to conjugate (Bell 1988; Lynn 2008).

The Colpodea—one of eleven major ciliate lineages—consists of about 200

species with similar somatic but diverse oral morphologies (Foissner 1993; Lynn 2008).

Colpodeans can be found in numerous habitats, some are fungivores, and at least one

species has a multicellular life stage (Foissner 1993). Much is known about colpodeans

and their evolution through morphological and molecular analyses (Dunthorn et al. 2009;

160

Dunthorn et al. 2008; Foissner 1985; Foissner 1993; Lasek-Nesselquist and Katz 2001;

Lynn 1976; Lynn 2008; Lynn et al. 1999). However, there is a lack of consensus on a

fundamental aspect of their biology—their sexuality. Although all known species have

micronuclei, conjugation has been observed only in Bursaria truncatella (Foissner 1993;

Poljansky 1934). The extent of sexuality in the rest of the colpodeans is debated: Foissner

(1993) proposes that they are asexual, while Dunthorn et al. (2008) suggest that they are

covertly sexual. It may not be surprising that we know little about sexuality in colpodean

ciliates. Even in known sexual ciliate species conjugation is not always easy to induce in

the laboratory (Sonneborn 1957). Ciliates also lack sex-specific morphologies or organs,

so you cannot look for morphologically different sexes and at least assume they are

sexual.

Here we review the evidence for and against sexuality in colpodeans. We also

review two reasons why we suspect the colpodeans to be sexual: the problems of ancient

asexuality and reversing the loss of sex.

6.3. Empirical evidence

While sex is well established in B. truncatella there is conflicting evidence for or against

sex in the rest of the colpodeans. Foissner (1993) proposes that they are asexual because

conjugation has not been observed even though numerous species have been examined

over many years. Likewise, darwinulid ostracods were long thought to be asexual

because males were never seen until they were found after a century of searching (Smith

161

et al. 2006). Asexuality of the colpodeans may represent a similar situation to these

ostracods: given enough time and requisite conditions, conjugation might be observed.

Two molecular studies of the colpodeans provide conflicting evidence. Bowers et

al. (1998) found that large subunit rDNA restriction digests of three species in the clade

Colpoda showed limited to no genetic variation within local populations. They suggest

that this data is consistent with asexuality. However, these molecular data are also

consistent with either sexuality (leading to homogenization of variation) or infrequent sex

(leading to clonality). This study also was not designed to test directly for the presence or

absence of sex.

In another molecular study not explicitly designed to test for sex, Dunthorn et al.

(2008) found little to no variation among small subunit rDNA sequences from a broad

sample of colpodeans. They suggest that this pattern is consistent with sexuality because

asexuality is hypothesized to lead to high intra-isolate allelic divergence because of the

absence of recombination (i.e., the Meselson effect; Birky 1996; Mark Welch and

Meselson 2000; Normark et al. 2003). However, testing for the Meselson effect is

problematic, as it can be masked or mimicked by a number of processes such as non-

meiotic recombination or rare sex (Ceplitis 2003; Normark et al. 2003; Omilian et al.

2006). Also ciliates, because of their sometimes non-canonical genetics and high levels of

paralogy (Katz et al. 2004; Nanney 1980; Snoeyenbos-West et al. 2002), pose additional

challenges to interpreting molecular signatures. Future studies could try to test for or

against sexuality by looking at other genomic signatures; e.g., the degree of decay of sex-

related loci, increased rate of deleterious mutations, number of retrotransposons, etc.

(Normark et al. 2003). However, ciliates might pose challenges to these tests are well.

162

6.4. Are colpodeans ancient asexuals?

If colpodeans are asexual (except B. truncatella) they would be the oldest putative

ancient asexual lineage. Molecular clock estimates place their age up to 900 million years

old (Wright and Lynn 1997). There are also fossils in 93 million year old amber (Martín-

González et al. 2008; Schmidt et al. 2001; Schönborn et al. 1999). Colpodeans are thus at

least as old, or perhaps even an order of magnitude older, than the putatively ancient

asexual bdelloid rotifers that date to about 130 million years old (Mark Welch et al.

2008).

While asexuals are known at the tips of the eukaryotic tree of life, ancient

asexuals are generally thought to be unlikely because asexuality is believed to lead to

rapid extinction (Lynch et al. 1993; Maynard Smith 1978; Normark et al. 2003)—but see

Schwander and Crespi (2009) for an alternative view. For example, without sex

eukaryotic lineages may be more susceptible to increased mutational load and

retrotransposons, may not be able to adapt to a changing environment, or may not be able

to escape predators and parasites over evolutionary time compared to those lineages that

remain sexual (Arkhipova and Meselson 2004; Bell 1982; Burt 2000; Hamilton 2001;

Kondrashov 1993; Maynard Smith 1978). Most claims of ancient asexuals have not been

supported (Judson and Normark 1996; Normark et al. 2003), except possibly the bdelloid

rotifers (Arkhipova and Meselson 2000; Mark Welch et al. 2004a; Mark Welch et al.

2008; Mark Welch and Meselson 2000; Mark Welch et al. 2004b).

This low expectation of ancient asexuality, though, derives from theory and

observations based on macrobes (Normark et al. 2003). Do these expectations apply to all

eukaryotes, including microbial lineages? We do not yet know. Little is even known

163

about the phylogenetic pattern of asexual microbial eukaryotes. Although many have

been postulated to be asexual (Sonneborn 1957), when they have been critically

examined, evidence for sex has been found; e.g., Giardia lambli (Birky 2005; Ramesh et

al. 2005), and Naegleria lovaniensis (Hurst et al. 1992; Pernin et al. 1992). There are

many ways in which microbial eukaryotes could pose challenges to macrobial

expectations. For example, many ciliates appear to have globally distributed populations

(Finlay 2002; Foissner et al. 2008), and many also have extremely large population sizes

(Lynch and Conery 2003; Snoke et al. 2006), but see Katz et al. (Foissner et al. 2008;

2006). Such population structures might allow microbial eukaryotes to defy predictions

that asexuality might lead to extinction.

If our macrobial theory of low expectations of ancient asexuals does not apply to

microbial eukaryotes, then colpodeans may very well be asexual. But if these

expectations do apply, then we would expect that colpodeans are covertly sexual since

ancient asexuals are not likely.

6.5. Did colpodeans reverse the loss of sex?

If colpodeans are asexual we would have to hypothesize a reverse of the loss of sex. This

is because the sexual B. truncatella is nested within putatively asexual clades (Fig. 1:

Dunthorn et al. 2009; Dunthorn et al. 2008; Lynn et al. 1999). Although there is debate

on the longevity of ciliate species (Dunthorn and Katz 2008; Nanney 1999), this reversal

back to sexuality in B. truncatella could have occurred many millions of years ago.

There is debate about the ability to reverse the loss of a complex character (Bull

and Charnov 1985; Collin and Miglietta 2008; Goldberg and Igic 2008; Gould 1970;

164

Porter and Crandall 2003; Simpson 1953; Teotónio and Rose 2001). Complex characters

can either be lost phenotypically or genotypically; the ability to regain depends on which

of these levels was originally involved and the amount of time intervening loss and

reversal (Collin and Miglietta 2008). If the genes remain, then regaining a character can

just be a process of shuffling around genotypes or turning back on suppressed loci. If the

genotype is lost, then it is much harder to regain the lost character.

Can the loss of sex be reversed? Sex may have been lost and regained along

lineages throughout the eukaryotic tree of life, though we would not know given the

current distribution of sex in extant species (Williams 1975). We do know of two putative

cases of regaining sexuality—both at relatively shallow nodes. In multiple populations of

the plant Hieracium pilosella (Chapman et al. 2003), reversal from asexuality to sexuality

entailed returning to homozygosity (and tetraploidy) for a recessive allele (Bicknell et al.

2000). In this case, although the phenotype of sex was lost, alleles encoding this

phenotype remained in the population. In oribatid mites, the case for reversal depends on

ancestral character state reconstructions (Domes et al. 2007). However, these ancestral

state reconstructions may be fundamentally flawed, leading to a false acceptance of

reversal (Goldberg and Igic 2008). The case of regaining sex in oribatid mites is thus

ambiguous, and may represent multiple, independent losses of sex. There is no current

evidence on whether the genotype was lost in oribatid mites. It should also be noted that

had Domes et al. (2007) applied their method to other macrobial taxa they may have

increased the number of putative cases of reversing the loss of sex, although these would

have been fundamentally flawed as well.

165

If sex-related loci are shown to be retained in the putative asexual relatives of B.

truncatella a strong case for reversing the loss of sex in colpodeans can be made.

However, the problem of reactivation of silenced loci increases over time as they may be

mutated or lost (Collin and Miglietta 2008; Marshall et al. 1994; Normark et al. 2003).

Adding the possible issues surrounding ancient asexuality, though, to the temporal issues

of reactivating loci would compound the problem of the colpodeans being asexual. If we

assume that both the phenotype and genotype of sex were lost, then our expectations for

reversing the loss of sex would be much smaller to none in the colpodeans.

6.6. Conclusion

We will probably not know if colpodeans are sexual until someone actually sees

conjugation in a Petri dish and reliably demonstrates nuclear exchange for additional

species. The lack of evidence for sex (e.g., no observation of conjugation) is not itself

evidence. Given the theories for the maintenance of sex, the combined problems of

ancient asexuals and reversing a complex trait, we suggest that many, if not all,

colpodeans are covertly sexual, not asexual. Nevertheless, it would be fruitful if it is

shown that macrobial theoretical expectations for ancient asexuality apply universally to

eukaryotes such ciliates and other microbial eukaryotes. It would also be fruitful to look

for loss or retention of sex-related loci in colpodean species.

166

6.7. References

Arkhipova, I., and M. Meselson. 2000. Transposable elements in sexual and ancient asexual taxa. Proc. Natl. Acad. Sci. USA 97:14473-14477.

Arkhipova, I., and M. Meselson. 2004. Deleterious transposable elements and the

extinction of asexuals. Bioessays 27:76-85. Bell, G. 1982. The masterpiece of nature: the evolution and genetics of sexuality.

University of California Press, Berkeley. Bell, G. 1988. Sex and death in protozoa: the history of an obsession. Cambridge

University Press, Cambridge. Bicknell, R. A., N. K. Borst, and A. M. Koltunow. 2000. Monogenic inheritance of

apomixis in two Hieracium species with distinct developmental mechanisms. Heredity 84:228-237.

Birky, C. W. 1996. Heterozygosity, heteromorphy, and phylogenetic trees in asexual

eukaryotes. Genetics 144:427-437. Birky, C. W. 2005. Sex: is Giardia doing in the dark? Curr. Biol. 15:R56-R58. Bowers, N., T. T. Kroll, and J. R. Pratt. 1998. Diversity and geographic distribution of

riboprints from three cosmopolitan species of Colpoda Müller (Ciliophora: Colpodea). Europ. J. Protistol. 34:341-347.

Bull, J. J., and E. L. Charnov. 1985. On irreversible evolution. 39:1149-1155. Burt, A. 2000. Sex, recombination, and the efficacy of natural selection — was

Weisman right? Evolution 54:337-351. Ceplitis, A. 2003. Coalescence times and the Meselson effect in asexual eukaryotes.

Genet. Res. 82:183-190. Chapman, H., G. J. Houliston, B. Robson, and I. Iline. 2003. A case of reversal: the

evolution and maintenance of sexuals from parthenogenetic clones in Hieracium pilosella. Int. J. Plant Sci. 164:719-728.

Collin, R., and M. P. Miglietta. 2008. Reversing opinions on Dollo's Law. Trends Ecol.

Evol. 23:602-609. Dacks, J., and A. J. Roger. 1999. The first sexual lineage and the relevance of

facultative sex. J. Mol. Evol. 48:779-783. Dini, F., and D. Nyberg. 1993. Sex in ciliates. Adv. Microb. Ecol. 13:85-153.

167

Domes, K., R. A. Norton, M. Maraun, and S. Scheu. 2007. Reevolution of sexuality breaks Dollo's law. Proc. Natl. Acad. Sci. USA 104:7139-7144.

Dunthorn, M., M. Eppinger, M. V. J. Schwarz, M. Schweikert, J. Boenigk, L. A. Katz,

and T. Stoeck. 2009. Phylogenetic placement of the Cyrtolophosididae Stokes, 1888 (Ciliophora; Colpodea) and neotypification of Aristerostoma marinum Kahl, 1931. Int. J. Syst. Evol. Microbiol. 59:167-180.

Dunthorn, M., W. Foissner, and L. A. Katz. 2008. Molecular phylogenetic analysis of

class Colpodea (phylum Ciliophora) using broad taxon sampling. Mol. Phylogenet. Evol. 48:316-327.

Dunthorn, M., and L. A. Katz. 2008. Richness of morphological hypotheses in ciliate

systematics allows for detailed assessment of homology and comparisons with gene trees. Denisia 23:389-394.

Felsenstein, J. 1974. The evolutionary advantage of recombination. Genetics 78:737-

756. Finlay, B. J. 2002. Global dispersal of free-living microbial eukaryote species. Science

296:1061-1063. Fischer, R. A. 1930. The genetical theory of natural selection. Oxford University Press,

Oxford. Foissner, W. 1985. Klassifikation und phylogenie der Colpodea (Protozoa: Ciliophora).

Arch. Protistenk. 129:239-290. Foissner, W. 1993. Colpodea (Ciliophora). Protozoenfauna 4/1:i-x, 1-798. Foissner, W., A. Chao, and L. A. Katz. 2008. Diversity and geographic distribution of

ciliates (Protista: Ciliophora). Biodivers. Conserv. 17:345-363. Goldberg, E. E., and B. Igic. 2008. On phylogenetic tests of irreversible evolution.

Evolution 62:2727-2741. Gould, S. J. 1970. Dollo on Dollo's Law: irreversibility and the status of evolutionary

laws. J. Hist. Biol. 3:189-212. Hamilton, W. D. 2001. Narrow roads of gene land, Vol. 2: the evolution of sex. Oxford

University Press, Oxford. Hur, J. H., K. V. Doninck, M. L. Mandino, and M. Meselson. 2008. Degenerate

tetraploidy was established before bdelloid rotifer families diverged. Mol. Bio. Evol. 26:375-383.

168

Hurst, L. D., W. D. Hamilton, and R. J. Ladle. 1992. Covert sex. Trends Ecol. Evol. 7:144-145.

Judson, O. P., and B. B. Normark. 1996. Ancient asexual scandals. Trends Ecol. Evol.

11:A41-A46. Katz, L. A., J. G. Bornstein, E. Lasek-Nesselquist, and S. V. Muse. 2004. Dramatic

diversity of ciliate histone H4 genes revealed by comparisons of patterns of substitutions and paralog divergences among eukaryotes. Mol. Biol. Evol. 21:555-562.

Katz, L. A., O. Snoeyenbos-West, and F. P. Doerder. 2006. Patterns of protein evolution

in Tetrahymena thermophila: implications for estimates of effective population size. Mol. Biol. Evol. 23:608-614.

Kondrashov, A. S. 1982. Selection against harmful mutations in large sexual and

asexual population. Genet. Res. 40:325-332. Kondrashov, A. S. 1993. Classification of hypotheses on the advantage of amphimixis.

J. Hered. 84:372-387. Lasek-Nesselquist, E., and L. A. Katz. 2001. Phylogenetic position of Sorogena

stoianovitchae and relationships within the class Colpodea (Ciliophora) based on SSU rDNA sequences. J. Eukaryot. Microbiol. 48:604-607.

Lynch, M., R. Bürger, D. Butcher, and W. Gabriel. 1993. Mutational meltdowns in

asexual populations. J. Hered. 84:339-344. Lynch, M., and J. S. Conery. 2003. The origins of genome complexity. Science

302:1401-1404. Lynn, D. H. 1976. Comparative ultrastructure and systematics of colpodida. Structural

conservatism hypothesis and a description of Colpoda steinii Maupas. J. Protozool. 23:302-314.

Lynn, D. H. 2008. The ciliated protozoa: characterization, classification, and guide to

the literature, 3rd edition. Springer, Dordrecht. Lynn, D. H., A. D. G. Wright, M. Schlegel, and W. Foissner. 1999. Phylogenetic

relationships of orders within the class Colpodea (phylum Ciliophora) inferred from small subunit rRNA gene sequences. J. Mol. Evol. 48:605-614.

Mark Welch, D., M. P. Cummings, D. M. Hillis, and M. Meselson. 2004a. Divergent

gene copies in the asexual class Bdelloidea (Rotifera) separated before the bdelloid radiation or within bdelloid families. Proc. Natl. Acad. Sci. USA 101:1622-1625.

169

Mark Welch, D., J. L. Mark Welch, and M. Meselson. 2008. Evidence for degenerate tetraploidy in bdelloid rotifers. Proc. Natl. Acad. Sci. USA 105:5145-5149.

Mark Welch, D., and M. Meselson. 2000. Evidence for the evolution of bdelloid rotifers

without sexual reproduction or genetic exchange. Science 288:1211-1215. Mark Welch, J. L., D. Mark Welch, and M. Meselson. 2004b. Cytogenetic evidence for

asexual evolution of bdelloid rotifers. Proc. Natl. Acad. Sci. USA 101:1618-1621. Marshall, C. R., E. C. Raff, and R. A. Raff. 1994. Dollo's law and the death and

resurrection of gene. Proc. Natl. Acad. Sci. USA 91:12283-12287. Martín-González, A., J. Wierzchos, J. C. Gutiérrez, J. Alonso, and C. Ascaso. 2008.

Morphological stasis of protists in lower Cretaceous amber. Protist 159:251-257. Maynard Smith, J. 1978. The evolution of sex. Cambridge University Press, Cambridge. Muller, H. J. 1964. The relation of recombination to mutational advance. Mutation Res.

1:2-9. Nanney, D. L. 1980. Experimental ciliatology: an introduction to genetic and

developmental analysis in ciliates. John Wiley & Sons, New York. Nanney, D. L. 1999. When is a rose? the kinds of Tetrahymena. Pp. 93-118 in R. A.

Wilson, ed. Species: new interdisciplinary essays. The MIT Press, Cambridge. Normark, B. B., O. P. Judson, and N. A. Moran. 2003. Genomic signatures of ancient

asexual lineages. Biol. J. Linn. Soc. 79:69-84. Omilian, A. R., M. E. A. Cristescu, J. L. Dudycha, and M. Lynch. 2006. Ameiotic

recombination in asexual lineages of Daphnia. Proc. Natl. Acad. Sci. USA 103:18638-18643.

Pernin, P., A. Ataya, and M. L. Cariou. 1992. Genetic structure of natural populations of

the free-living ameba, Naegleria lovaniensis. evidence for sexual reproduction. Heredity 68:173-181.

Poljansky, G. 1934. Geschlechtsprozesse bei Bursaria truncatella O.F. Müll. Arch.

Protistenk. 81:420-546. Porter, M. L., and K. A. Crandall. 2003. Lost along the way: significance of evolution in

reverse. Trends. Ecol. Evol. 18:541-547. Raikov, I. B. 1996. Nuclei of ciliates. Pp. 221-242 in K. Hausmann and P. C. Bradbury,

eds. Ciliates: cells as organisms. Gustav Fischer, Stuttgart.

170

Ramesh, M. A., S.-B. Malik, and J. M. Longsdon. 2005. A phylogenomic inventory of meiotic genes: evidence for sex in Giardia and an early eukaryotic origin of meiosis. Curr. Biol. 15:185-191.

Schmidt, A. R., H. von Eynatten, and M. Wagreich. 2001. The Mesozoic amber of

Schliersee (southern Germany) is Cretaceous in age. Cretaceous Res. 22:423-428. Schwander, T., and B. J. Crespi. 2009. Twigs on the tree of life? Neutral and selective

models for intergrating macroevolutionary patterns and microevolutionary processes in the analysis of asexuality. Molec. Ecol. 18:28-42.

Schönborn, W., H. Dörfelt, W. Foissner, L. Krientiz, and U. Schäfer. 1999. A fossilized

microcenosis in Triassic amber. J. Eukaryot. Microbiol. 46:571-584. Simpson, G. G. 1953. The major features of evolution. Columbia University Press, New

York. Smith, R. J., T. Kamiya, and D. J. Horne. 2006. Living males of the 'ancient asexual'

Darinulidae (Ostracoda: Crustacea). Proc. R. Soc. Lond. B 273:1569-1578. Snoeyenbos-West, O. L. O., T. Salcedo, G. B. McManus, and L. A. Katz. 2002. Insights

into the diversity of choreotrich and oligotrich ciliates (Class: Spirotrichea) based on genealogical analyses of multiple loci. Int. J. Syst. Evol. Microbiol. 52:1901-1913.

Snoke, M. S., T. U. Berendonk, D. Barth, and M. Lynch. 2006. Large global effective

population sizes in Paramecium. Mol. Biol. Evol. 23:2474-2479. Sonneborn, T. M. 1957. Breeding systems, reproductive methods, and species problems

in protozoa. Pp. 155-324 in E. Mayr, ed. The species problem. American Association for the Advancement of Science, Washington D. C.

Teotónio, H., and M. R. Rose. 2001. Perspective: reverse evolution. Evolution 55:653-

660. West, S. A., C. M. Lively, and A. F. Read. 1999. A pluralist approach to sex and

recombination. J. Evol. Biol. 12:1003-1012. Williams, G. C. 1975. Sex and evolution. Princeton University Press, Princeton. Wright, A.-D. G., and D. H. Lynn. 1997. Maximum ages of ciliate lineages estimated

using a small subunit rRNA molecular clock: crown eukaryotes date back to the paleoprotoerozoic. Arch. Protistenk. 148:329-341.

171

Figure 6.1. Phylogeny and distribution of sex within the Colpodea. Major clades are

labeled. Only Bursaria truncatella is known to have sex. The ancestral state (?) for

colpodeans is debated: either sexual or asexual. Modified from Dunthorn et al. (2009).

172

173

CHAPTER 7

PHYLOCODE DEFINITIONS FOR FOUR CILIATE CLADES

Micah Dunthorna and Denis H. Lynnb

aGraduate Program in Organismic and Evolutionary Biology, University of

Massachusetts, Amherst, Massachusetts, 01003, USA

bDepartment of Integrative Biology, University of Guelph, Guelph, Ontario, N1G 2W1,

Canada

174

7.1. Abstract

Ciliates clades have traditionally been named using the International Code of

Zoological Nomenclature, a rank-based system that governs names at or below the family

rank. Here we argue that the Internal Code of Phylogenetic Nomenclature should be used

to name ciliate. We apply this code to four ciliate clades above the rank of family:

Ciliophora, Postciliodesmatophra, Intramacronucleata, and Colpodea.

175

7.2. Introduction

Names for ciliate taxa are currently governed by the International Code of

Zoological Nomenclature (International Commission on Zoological Nomenclature 1999);

hereafter referred to as the ICZN. Here I will briefly argue that ciliate taxa should be

named using the PhyloCode, which is governed by the International Code of

Phylogenetic Nomenclature (Cantino and de Queiroz 2007); hereafter referred to as the

ICPN. As an example, four well-known ciliate names are converted using the ICPN.

Ciliates are a clade of microbial eukaryotes with dimorphic nuclei and cilia in at

least one life-cycle stage (Lynn 2008). The diploid ‘germline’ micronucleus is, for the

most part, transcriptionally inactive and is exchanged between ciliates during sex. The

usually polyploid ‘somatic’ macronucleus is transcriptionally active (Raikov 1996).

Ciliate morphospecies are described by the patterns formed by the kineties (arrays) of

kinetosomes (cilia basal bodies) and associated fibers in the somatic and oral membranes.

Analyses of these rich characters have led to several taxonomic schemes at a variety of

levels in the ciliate tree of life (Bütschli 1887-1889; Kahl 1926, 1930-1935; Kahl 1931;

Puytorac et al. 1974; Corliss 1979; Small and Lynn 1981; Foissner 1985; Matthes 1988;

Foissner 1993; Puytorac 1994; e.g., Berger 1999; Lynn and Small 2002; Berger 2006;

Lynn 2008).

176

7.3. The PhyloCode

The ICPN is a nomenclatural system that explicitly uses the theoretical approach

of phylogenetic systematics. In essence, this system uses hypotheses of relationships, or

distributions of apomorphies to name clades (monophyletic groups). To validly name a

taxon using the ICPN there must be a description of clade or species using internal and,

sometimes, external, specifiers that are either species or characters (Cantino and de

Queiroz 2007). Ideally there should be at least one reference phylogeny that includes the

specifiers that accompanies the description (Cantino and de Queiroz 2007). ICPN names

can be converted from the ICZN, or other rank-based codes, or new names can be coined.

Three main types of definitions are used in the ICPN: node-based, branch-based,

and apomorphy-based(Cantino and de Queiroz 2007). The exact wording for each

definition can take many forms, and node-based definitions can be branch- and

apomorphy-modified. Which type of definition to use depends on a number of issues;

e.g., the presence of absence of basal fossil lineages, the amount of support for the

monophyly of the clade, support for the relationships within the clade, or support for

sister groups (Bryant 1994; Lee 1996; Sereno 1999, 2005; Cantino and de Queiroz 2007;

Cantino et al. 2007).

A phylogenetically-based nomenclature system like the ICPN, as in any system,

has its strengths (de Queiroz and Gauthier 1990; de Queiroz and Gauthier 1992, 1994;

Cantino et al. 1997; de Queiroz 1997; Lee 2001; Bryant and Cantino 2002; Pleijel and

Rouse 2003; Sereno 2005; de Queiroz 2006; Laurin et al. 2006; de Queiroz 2007) and

weaknesses (Dominguez and Wheeler 1997; Moore 1998; Benton 2000; Nixon and

Carpenter 2000; Dyke 2002; Carpenter 2003; Keller et al. 2003; Kojima 2003; Moore

177

2003; Nixon et al. 2003; Schuh 2003; Barkley et al. 2004; Wenzel et al. 2004; Polaszek

and Wilson 2005). Only a few of these strengths will be briefly discussed below to show

why the ICPN should be used to name ciliate clades.

7.4. Why the PhyloCode should be applied to ciliates

The ICPN has been used in a number of terminal, macrobial eukaryotic taxa such

as animals, plants, and fungi (Hibbett and Donoghue 1998; Donoghue et al. 2001;

Stefanovic et al. 2003; Joyce et al. 2004; Modesto and Anderson 2004; Sangster 2005;

Taylor and Naish 2005; Cantino et al. 2007). Recently, Adl et al. (2007) argued that

microbial eukaryote species should be named using the ICPN or a similar nomenclatural

system, but they did not call for its application in naming deeper microbial eukaryotic

clades. Here we argue that the ICPN should be applied to deeper microbial eukaryotic

nodes, using ciliates as an example.

The ICPN applies at all phylogenetic depths. Deep nodes in the ciliate tree of life

are not covered by the ICZN because they are above the family rank. There are attempts

to make the ICZN and other rank-based codes apply at higher ranks (Hemming 1953;

Corliss 1972; Ghiselin 1977; Brothers 1983; Dundee 1989; Dubois 2005, 2006), but these

suggestions have yet to be codified. Those using the ICZN, or any rank-based system,

above the family rank are doing so with the consensus of the authors and users of the

classification, not with authority of these codes (Corliss 1983). On the other hand, the

ICPN governs the name of any taxon at any depth in the tree of life.

178

The ICPN allows for only names that designate taxa that are hypothesized to be

monophyletic. The ICZN does not say anything about what kind of groups

(monophyletic, paraphyletic, or polyphyletic) can or cannot be named. There is

opposition to strictly monophyletic taxonomies (Sosef 1997; Mayr 1998; Thorne 2000;

Mayr and Bock 2002; Wu et al. 2002; Brummitt 2003; Nordal and Stedje 2005; Heywood

et al. 2007; Hörandl 2007), but there still is no way to easily interpret paraphyletic or

polyphyletic taxa. The majority of macrobial classifications only recognize monyphyletic

taxa (e.g., APG II 2003; Faivovich et al. 2005), athough paraphyletic taxa are still

recognized researchers for microbial eukaryotes in general (e.g., Cavalier-Smith 1999;

Cavalier-Smith 2004, 2007) and ciliate in specific (e.g., Berger 1999, 2006).

Monophyletic taxa sensu Hennig (1966) or Farris (1974) are unambiguous, reflect sister-

group relationships that are objectively stated and interpreted. Although the ICPN does

not prevent the recognition of paraphyletic and polyphyletic taxa (de Queiroz 2006), it

does restrict the application of names to only groups that are hypothesized to be

monophyletic.

In the ICPN named ranks will not necessarily occur. In the ICZN, a family, genus,

or species name is only valid if it is given a rank, and the name may change if its rank is

changed. Ranked names are neither mandated nor prohibited by the ICPN (de Queiroz

2006; Cantino and de Queiroz 2007). Even if an ICPN name has a ranked ending, a

change in name will not occur if hypotheses of relationships are revised later.

The ICPN allows for explicit statements of when a name should be rejected and

no longer used through explicit statement about which taxa can be included and excluded.

Although taxonomies are often stated as hypotheses, there is usually no way to know

179

when the name should be rejected and no longer used as in the ICZN unless there is a

nomenclatural act later on. The ICPN provides a system in which a taxon name can be

explicated stated when it should be used and when it should abandoned in the light of

new data through the use of specifiers and qualifying clauses (Schander and Thollesson

1995; Cantino et al. 1997; Sereno 1999; Bryant and Cantino 2002; Joyce et al. 2004;

Sereno 2005; Bertrand and Härlin 2006).

7.5. Application of the PhyloCode to four ciliate clades

Below are four ciliate clades to which we apply the ICPN. Since they are yet to be

published in the International Code of Phylogenetic Nomenclature’s Companion Volume,

which will be the official start for all valid PhyloCode names and definitions, because

theses names have not been reviewed by outside taxonomists, and because these

definitions are not published in an easily attainable journal, the definitions below are

neither effectively nor validly published here. The entire ciliate clade (currently ranked at

the phylum level in the ICZN) and the Postciliodesmatophora and Intramacronucleata

(the two currently subphyla) and the Colpodea (ranked as a class) are converted here. In

all definitions ranks are ignored. Taxa all italicized following standard formatting of the

ICPN.

180

7.5.1. Ciliophora

F. Doflein 1901 [M. Dunthorn & D. H. Lynn 2009], converted clade name

Definition: The least inclusive clade containing Tetrahymena thermophila Nanney and

McCoy 1976, Blepharisma americanum Suzuki 1954, and Loxodes striatus (Engelmann

1862). This is a node-based definition in which all of the specifiers are extant; it is

intended to apply to a crown clade. Abbreviated definition: < Tetrahymena thermophila

Nanney and McCoy 1976 & Blepharisma americanum Suzuki 1954 & Loxodes striatus

(Engelmann 1862).

Etymology: Derived from the Latin cilium (eyelash) and Greek phoreus (bearer), in

reference to the cilia on the cell bodies of all members of this clade (in at least one stage

of the life cycle).

Reference Phylogeny: The primary reference phylogeny is Hammerschmidt et al. (1996:

Fig. 2). See also Baldauf et al. (2000: Fig. 1) and Yoon et al. (2008: Fig. 2).

Composition: Almost all known ciliates are extant; those few fossils found can be placed

within previously recognized taxa (Lynn 2008). The clade Ciliophora contains

Postciliodesmatophora and Intramacronucleata as defined in this volume, which in turn

include all of the clades listed by Lynn (2008).

181

Diagnostic Apomorphies: Ciliates have three apomorphic characters: 1) dimorphic nuclei,�

with a “germline” micronucleus and a “somatic” macronucleus that are not homologous

with those found in Foraminifera Lee 1990; 2) cilia, in at least one life-cycle stage, that

are derived from a kinetosome (= eukaryotic basal body)� that is associated with three

fibers (a kinetodesmal fiber, a postciliary microtubular �ribbon, and a transverse

microtubular ribbon); and 3) sex in the form of� conjugation, where there is typically

mutual exchange of haploid meiotic products of the �micronucleus (Raikov 1996; Lynn

2008).

Synonyms: Ciliata M. Perty 1852 [approximate], Infusoria Bütschli 1887 [approximate];

see review by Lynn (2008).

Comments: Ciliates have long been recognized as a group because of their distinctive

morphology, and molecular data strongly support this clade (Hammerschmidt et al. 1996;

Lynn 2008). Since the beginning of the 20th century, the name Ciliophora has been the

most widely used (e.g., Corliss 1979; Puytorac 1994; Jankowski 2007; Lynn 2008).

182

7.5.2. Postciliodesmatophora

Z. P. Gerassimova & L. N. Seravin 1976 [M. Dunthorn & D. H. Lynn 2009], converted

clade name

Definition: The least inclusive clade containing Blepharisma americanum

Suzuki 1954 and Loxodes striatus (Engelmann 1862). This is a node-based definition in

which both specifiers are extant; it is intended to apply to a crown clade. Abbreviated

definition: < Blepharisma americanum Suzuki 1954 & Loxodes striatus (Engelmann

1862).

Etymology: Derived from the Latin post (after, behind) and cilium (eyelash) and the

Greek desmos (bond or chain) and phoreus (bearer), in reference to the postciliodesmata

borne by members of this clade (see Diagnostic Apomorphies).

Reference Phylogeny: The primary reference phylogeny is Hammerschmidt et al. (1996:

Fig. 2). See also Hirt et al. (1995: Fig. 2).

Composition: Contains the Karyorelictea and the Heterotrichea; see Lynn (Lynn 2008)

for a list of the contents of these clades.

Diagnostic Apomorphies: The Postciliodesmatophora have postciliodesmata, formed

from postciliary microtubule ribbons overlapping along the right side of a kinety

183

(integrated somatic file of kinetids). Macronuclei either do not divide (Karyorelictea) or

divide using extra-macronuclear microtubules (Heterotrichea).

Synonyms: None.

Comments: Gerassimova and Seravin (1976) recognized this clade based on the shared

postciliodesmata. Subsequent phylogenetic analyses of small subunit ribosomal RNA

genes have supported its monophyly (Hirt et al. 1995; Hammerschmidt et al. 1996; Lynn

2008). The name is defined here in a manner consistent with the intent of the original

conception of the group by Gerassimova and Seravin (1976) as containing only

Karyorelictea and the Heterotrichea. If a lineage is found that has postciliodesmata but

falls below the divergence between Karyorelictea and Heterotrichea, this larger clade

will have to have another name.

184

7.5.3. Intramacronucleata D. H. Lynn 1996 [D. H. Lynn & M. Dunthorn 2009],

converted clade name

Definition: The most inclusive crown clade exhibiting intramacronuclear microtubules

(as described under Diagnostic Apomorphies) synapomorphic with those in Tetrahymena

thermophila Nanney and McCoy 1976. This is an apomorphy-modified node-based

definition in which the specifier is extant; it is intended to apply to a crown clade.

Abbreviated definition: >∇ exhibiting intramacronuclear microtubules (Tetrahymena

thermophila Nanney and McCoy 1976).

Etymology: Derived from the Latin intra (within), Greek makros (large), and Latin

nucleus (kernel), in reference to the presence of intramacronuclear microtubules during

cell division (see Diagnostic Apomorphies).

Reference Phylogeny: The primary reference phylogeny is Hammerschmidt et al.

Hammerschmidt et al. (1996: Fig. 2). See also Hirt et al. (1995: Fig. 2) and Yoon et al.

(2008: Fig. 2).

Composition: The majority of clades within the Ciliophora, as defined in this volume, are

within the Intramacronucleata. Lynn (2008) lists the following taxa as included in the

Intramacronucleata: Armophorea, Colpodea, Litostomatea, Nassophorea,

Oligohymenophorea, Phyllopharyngea, Plagiopylea, Prostomatea, and Spirotrichea.

185

Diagnostic Apomorphies: The most distinctive diagnostic apomorphy is the presence of

microtubules within the macronuclear envelope during nuclear division. In other

Ciliophora (i.e. Postciliodesmotophora), microtubules are either extra-macronuclear, or

macronuclei do not divide.

Synonyms: None

Comments: Lynn (1996) established this taxon, which contains most known ciliates,

based on small subunit ribosomal RNA gene phylogenies as well as the presence of

intranuclear microtubules in dividing macronuclei. Monophyly of this clade is strongly

supported, but internal relationships are unresolved (Lynn 2008).

186

7.5.4. Colpodea E. B. Small & D. H. Lynn 1981 [M. Dunthorn & D. H. Lynn 2009],

converted clade name

Definition: The most inclusive clade exhibiting a LKm fiber (as described below under

Diagnostic Apomorphies) synapomorphic with that in Colpoda cucullus O. F. Müller

(1773) K. C. Gmelin 1790. This is an apomorphy-modified node-based definition in

which the specifier is extant; it is intended to apply to a crown clade. Abbreviated

definition: >∇ exhibiting a LKm fiber (Colpoda cucullus O. F. Müller (1773) K. C.

Gmelin 1790).

Etymology: Derived from the Greek kolpus (womb).

Reference Phylogeny: The primary reference phylogeny is Dunthorn et al. (2008: Fig. 2).

See also Dunthorn et al. (2009), and Lynn et al. (1999).

Composition: All taxa listed in Foissner (1993) and Lynn (2008).

Diagnostic Apomorphies: The most distinctive diagnostic apomorphy of the Colpodea is

the LKm fiber (=transversodesmal fiber) composed of overlapping transverse ribbons of

microtubules extending from the posterior kinetosome (= eukaryotic basal body) of the

somatic dikinetids� (Foissner 1993; Lynn 2008). The Colpodea also have: somatic

stomatogenesis, where parental oral structures are partially to completely reorganized

during cell division; and a reticulate silverline system.

187

Synonyms: None.

Comments: Using the structural conservatism hypothesis (Lynn 1976, 1981), Small and

Lynn (1981) brought once disparate taxa together into the Colpodea. The Colpodea was

expanded by Foissner (1985), who later monographed the group (Foissner 1993). Bardele

(1989) rejects the monophyly of the taxon based on the presence or absence of ciliary

plaques, but there is no support for this claim (Lynn et al. 1999; Dunthorn et al. 2008).

Monophyly of the Colpodea is currently neither supported nor rejected by small subunit

ribosomal DNA sequences (Dunthorn et al. 2008).

188

7.6. References

Adl, S. M., B. S. Leander, A. G. B. Simpson, S. B. Archibald, O. R. Anderson, D. Bass, S. S. Bowser, G. Brugerolle, M. A. Farmer, S. Karpov, M. Kolisko, C. E. Lane, D. J. Lodge, D. G. Mann, R. Meisterfeld, L. Mendoza, Ø. Moestrup, S. E. Mozley-Standridge, A. V. Smirnov, and F. Spiegel. 2007. Diversity, nomenclature, and taxonomy of protists. Syst. Biol. 56:684-689.

APG II. 2003. An update of the Angiosperm Phylogeny Group classification for the oders and families of flowering plants: APG II. Bot. J. Linn. Soc. 141:399-436.

Baldauf, S. L., A. J. Roger, I. Wenk-Siefert, and W. F. Doolittle. 2000. A kingdom-level phylogeny of eukaryotes based on combined protein data. Science 290:972-977.

Bardele, C. F. 1989. From ciliate ontogeny to ciliate phylogeny: a program. Boll. Zool. 56:235-243.

Barkley, T. M., P. DePriest, V. Funk, R. W. Kiger, W. J. Kress, and G. Moore. 2004. Linnaean nomenclature in the 21st Century: A report from a workshop on integrating traditional nomenclature and phylogenetic classification. Taxon 53:153-158.

Benton, M. J. 2000. Stems, nodes, crown clades, and rank-free lists: Is Linnaeus dead? Biol. Rev. 75:633-648.

Berger, H. 1999. Monograph of the Oxytrichidae (Ciliophora, Hypotrichia). Monographiae Biol. 78:i-xii, 1-1080.

Berger, H. 2006. Monograph of the Urostyloidea (Ciliophora, Hypotrichia). Monographiae Biol. 85:i-xv, 1-1303.

Bertrand, Y., and M. Härlin. 2006. Stability and universality in the application of taxon names in phylogenetic nomenclature. Syst. Biol. 55:848-858.

Brothers, D. J. 1983. Nomenclature at the ordinal and higher levels. Syst. Zool. 32:34-42.

Brummitt, R. K. 2003. Further dogged defense of paraphyletic taxa. Taxon 52:803-804.

Bryant, H. N. 1994. Comments on the phylogenetic definitions of taxon names and conventions regarding the naming of crown clades. Syst. Biol. 43:124-130.

Bryant, H. N., and P. D. Cantino. 2002. A review of criticisms of phylogenetic nomenclature: Is taxonomic freedom the fundamental issue? Biol. Rev. 77:39-55.

Bütschli, O. 1887-1889. Protozoa. Abt. III. Infusoria und System der Radiolaria in H. G. Bronn, ed. Klassen und Ordnungen de Thier-Reichs, Vol. I. C. F. Winter, Leipzig.

189

Cantino, P. D., and K. de Queiroz. 2007. International Code of Phylogenetic Nomenclature. Version 4b [http://www.ohiou.edu/phylocode/].

Cantino, P. D., J. A. Doyle, S. W. Graham, W. S. Judd, R. G. Olmstead, D. E. Soltis, P. S. Soltis, and M. J. Donoghue. 2007. Towards a phylogenetic nomenclature of Tracheophyta. Taxon 56:E1-E44.

Cantino, P. D., R. G. Olmstead, and S. J. Wagstaff. 1997. A comparison of phylogenetic nomenclature with the current system: A botanical case study. Syst. Biol. 46:313-331.

Carpenter, J. M. 2003. Critique of pure folly. Bot. Rev. 69:79-92.

Cavalier-Smith, T. 1999. Zooflagellate phylogeny and the systematics of Protozoa. Biol. Bull. 196:393-396.

Cavalier-Smith, T. 2004. Only six kingdoms of life. Proc. R. Soc. Lond. B 271:1251-1262.

Cavalier-Smith, T. 2007. Evolution and relationships of algae: major branches of the tree of life. Pp. 21-55 in J. Brodie, and J. Lewis, eds. Unravelling the algae: the past, present, and future of algal systematics. CRC Press, Boca Raton.

Corliss, J. O. 1972. Common sense and courtesy in nomenclatural taxonomy. Syst. Zool. 21:117-122.

Corliss, J. O. 1979. The ciliated protozoa: characterization, classification and guide to the literature. Pergamon Press, Oxford.

Corliss, J. O. 1983. Consequences of creating new kingdoms of organisms. Bioscience 33:314-318.

de Queiroz, K. 1997. The Linnaean hierarchy and the evolutionatization of taxonomy, with emphasis on the problem of nomenclature. Aliso 15:125-144.

de Queiroz, K. 2006. The Phylocode and the distinction between taxonomy and nomenclature. Syst. Biol. 55:160-162.

de Queiroz, K. 2007. Toward an intergrated system of clade names. Syst. Biol. 956-974.

de Queiroz, K., and J. Gauthier. 1990. Phylogeny as a central principle in taxonomy: phylogenetic definitions of taxon names. Syst. Biol. 39:307-322.

de Queiroz, K., and J. Gauthier. 1992. Phylogenetic taxonomy. Annual Review Of Ecology And Systematics 23:449-480.

de Queiroz, K., and J. Gauthier. 1994. Toward a phylogenetic system of biological nomenclature. Trends Ecol. Evol. 9:27-31.

190

Dominguez, E., and Q. D. Wheeler. 1997. Taxonomic stability is ignorance. Cladistics 13:367-372.

Donoghue, M. J., T. Eriksson, P. A. Reeves, and R. G. Olmstead. 2001. Phylogeny and phylogenetic taxonomy of Dipsacales with special reference to Sinadoxa and Tetradoxa (Adoxaceae). Harv. Pap. Bot. 6:459-479.

Dubois, A. 2005. Proposed Rules for the incorporation of nomina of higher-ranked zoological taxa in the International Code of Zoological Nomenclature. 1. Some general questions, concepts and terms of biological nomenclature. Zoosystema 27:365-426.

Dubois, A. 2006. Proposed Rules for the incorporation of nomina of higher-ranked zoological taxa in the International Code of Zoological Nomenclature. 2. The proposed Rules and their rationale. Zoosystema 28:165-258.

Dundee, H. A. 1989. Higher category name usage for amphibians and reptiles. Syst. Zool. 38:398-406.

Dunthorn, M., M. Eppinger, M. V. J. Schwarz, M. Schweikert, J. Boenigk, L. A. Katz, and T. Stoeck. 2009. Phylogenetic placement of the Cyrtolophosididae Stokes, 1888 (Ciliophora; Colpodea) and neotypification of Aristerostoma marinum Kahl, 1931. Int. J. Syst. Evol. Microbiol. 59:167-180.

Dunthorn, M., W. Foissner, and L. A. Katz. 2008. Molecular phylogenetic analysis of class Colpodea (phylum Ciliophora) using broad taxon sampling. Mol. Phylogenet. Evol. 48:316-327.

Dyke, G. J. 2002. Should paleontologists use "phylogenetic" nomenclature? J. Paleontol. 76:793-796.

Faivovich, J., C. F. B. Haddad, P. C. A. Garcia, D. R. Frost, J. A. Campbell, and W. C. Wheeler. 2005. Systematic review of the frog family Hylidae, with special reference to Hylinae: a phylogenetic analysis and taxonomic revisison. Bull. Amer. Mus. Nat. Hist. 294.

Farris, J. S. 1974. Formal definitions of paraphyly and polyphyly. Syst. Zool. 23:548-554.

Foissner, W. 1985. Klassifikation und phylogenie der Colpodea (Protozoa: Ciliophora). Arch. Protistenk. 129:239-290.

Foissner, W. 1993. Colpodea (Ciliophora). Protozoenfauna 4/1:i-x, 1-798.

Gerassimova, Z. P., and L. N. Seravin. 1976. Ectoplasmic fibrillar system of Infusoria and its role for the understanding of their phylogeny. Zool. Zh. 55:645-656.

Ghiselin, M. T. 1977. On changing the names of higher taxa. Syst. Zool. 26:346-349.

191

Hammerschmidt, B., M. Schlegel, D. Lynn, D. D. Leipe, M. L. Sogin, and I. B. Raikov. 1996. Insights into the evolution of nuclear dualism in the ciliates revealed by phylogenetic analysis of rRNA sequences. J. Eukaryot. Microbiol. 43:225-230.

Hemming, F., ed. 1953. Copenhagen decisions on zoological nomenclature, addirions to, and modeifications of, the Règles internationales de la nomenclature zoologique; approved and adapted by the 14th International Congress of Zoology, Copenhagen, August, 1953. International Trust for Zoological Nomenclature, London.

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

Heywood, V. H., R. K. Brummitt, A. Culham, and O. Seberg. 2007. Flowering Plant Families of the World. Firefly Books, Ontario, Canada.

Hibbett, D. S., and M. J. Donoghue. 1998. Intergrating phylogenetic analysis and classification in fungi. Mycologia 90:347-356.

Hirt, R. P., P. L. Dyal, M. Wilkinson, B. J. Finlay, D. M. Roberts, and T. M. Embley. 1995. Phylogenetic relationships among Karyorelictids and Heterotrichs inferred from small subunit rRNA sequences: resolution at the base of the ciliate tree. Mol. Phylogenet. Evol. 4:77-87.

Hörandl, E. 2007. Neglecting evolution is bad taxonomy. Taxon 56:1-5.

International Commission on Zoological Nomenclature. 1999. International Code of Zoological Nomenclature, 4th ed. The International Trust for Zoological Nomenclature, London.

Jankowski, A. W. 2007. Phylum Ciliophora Doflein, 1901. Pp. 415-993 in A. F. Alimov, ed. Protista. Part 2, Handbook of Zoology. Russian Academy of Sciences, Zoological Institure, St. Petersburg.

Joyce, W. G., J. F. Parham, and J. A. Gauthier. 2004. Developing a protocol for the conversion of rank-based taxon names to phylogenetically defined clade names, as exemplified by turtles. J. Paleontol. 78:989-1013.

Kahl, A. 1926. Neue und wenig bekannte Formen der holotrichen und heterotrichen Ciliaten. Arch. Protistenk. 55:197-438.

Kahl, A. 1930-1935. Urtiere oder Protozoa. I: Wimpertiere oder Ciliata (Infusoria), eine Bearbeitung der freilebenden und ectocommensalen Infusorien der Erde, unter Ausschluss der marinen Tintinnidae. Pp. 1-886 in F. Dahl, ed. Die Teirwelt Deutschlands. G. Fischer, Jena.

Kahl, A. 1931. Urtiere oder Protozoa I: Wimpertiere oder Ciliate (Infusoria) 2. Holotricha außer den im 1. Teil behandelten Prostomata. Tierwelt Dtl. 21:181-398.

192

Keller, R. A., R. N. Boyd, and Q. D. Wheeler. 2003. The illogical basis of phylogenetic nomenclature. Bot. Rev. 69:93-110.

Kojima, J. 2003. Apomorphy-based definition also pinpoints a node, and PhyloCode names prevent effective communication. Bot. Rev. 69:44-58.

Laurin, M., K. de Queiroz, and P. D. Cantino. 2006. Sense and stability of taxon names. Zoologica Scripta 35:113-114.

Lee, M. S. Y. 1996. Stability in meaning and content of taxon names: an evaluation of crown clade definitions. Proc. R. Soc. Lond. B 263:1103-1109.

Lee, M. S. Y. 2001. On recent arguments for phylogenetic nomenclature. Taxon 50:175-180.

Lynn, D. H. 1976. Comparative ultrastructure and systematics of colpodida. Structural conservatism hypothesis and a description of Colpoda steinii Maupas. J. Protozool. 23:302-314.

Lynn, D. H. 1981. The organization and evolution of microtubular organelles in ciliated protozoa. Biol. Rev. 56:243-292.

Lynn, D. H. 1996. My journey in ciliate systematics. J. Eukaryot. Microbiol. 43:253-260.

Lynn, D. H. 2008. The ciliated protozoa: characterization, classification, and guide to the literature, 3rd edition. Springer, Dordrecht.

Lynn, D. H., and E. B. Small. 2002. Phylum Ciliophora Doflein, 1901. Pp. 371-656 in J. J. Lee, G. F. Leedale, and P. C. Bradbury, eds. An illustrated guide to the protozoa, 2nd ed. Society of Protozoologists, Lawrence, Kansas.

Lynn, D. H., A. D. G. Wright, M. Schlegel, and W. Foissner. 1999. Phylogenetic relationships of orders within the class Colpodea (phylum Ciliophora) inferred from small subunit rRNA gene sequences. J. Mol. Evol. 48:605-614.

Matthes, D. 1988. Suctoria und Urceolariidae. Protozoenfauna 7/1:i-ix, 1-309.

Mayr, E. 1998. Two empires or three? Proc. Natl. Acad. Sci., USA 95:9720-9723.

Mayr, E., and W. J. Bock. 2002. Classifications and other ordering systems. J. Zoo. Syst. Evol. Research 40:169-194.

Modesto, S. P., and J. S. Anderson. 2004. The phylogenetic definition of Reptilia. Syst. Biol. 53:815-821.

Moore, G. 1998. A comparison of traditional and phylogenetic nomenclature. Taxon 47:561-579.

Moore, G. 2003. Should taxon names be explicitly defined? Bot. Rev. 69:2-21.

193

Nixon, K. C., and J. M. Carpenter. 2000. On the other "phylogenetic systematics". Cladistics 16:298-318.

Nixon, K. C., J. M. Carpenter, and D. W. Stevenson. 2003. The PhyloCode is fatally flawed, and the "Linnaean" system can easily be fixed. Bot. Rev. 69:111-120.

Nordal, H., and B. Stedje. 2005. Paraphyletic taxa should be accepted. Taxon 54:5-6.

Pleijel, F., and G. W. Rouse. 2003. Ceci n'est pas une pipe: names, clades and phylogenetic nomenclature. J. Zool. Syst. Evol. Research 41:162-174.

Polaszek, A., and E. O. Wilson. 2005. Sense and stability in animal names. Trends Ecol. Evol. 20:421-422.

Puytorac, P. d. 1994. Phylum Ciliophora Doflein, 1901. Pp. 1-15 in P. P. de, ed. Traité de Zoologie, Tome II, Infusoires Ciliés, Fasc. 2, Systématique. Masson, Paris.

Puytorac, P. d., A. Batisse, J. Bohatier, J. O. Corliss, G. Deroux, P. Didier, J. Dragesco, G. Fryd-Versavel, J. Grain, C.-A. Groliére, F. Iftode, M. Laval, M. Roque, A. Savoie, and M. Tuffrau. 1974. Proposition d'une classification du phylum Ciliophora Doflein, 1901. C.R. Acad. Sci. Paris 278:2799-2802.

Raikov, I. B. 1996. Nuclei of ciliates. Pp. 221-242 in K. Hausmann, and P. C. Bradbury, eds. Ciliates: cells as organisms. Gustav Fischer, Stuttgart.

Sangster, G. 2005. A name for the clade formed by owlet-nightjars, swifts and hummingbirds (Aves). Zootaxa:1-6.

Schander, C., and M. Thollesson. 1995. Phylogenetic Taxonomy―Some Comments. Zool. Scr. 24:263-268.

Schuh, R. 2003. The Linnaean system and its 250-year persistence. Bot. Rev. 69:59-78.

Sereno, P. C. 1999. Definitions in phylogenetic taxonomy: Critique and rationale. Syst. Biol. 48:329-351.

Sereno, P. C. 2005. The logical basis of phylogenetic taxonomy. Systematic Biology 54:595-619.

Small, E. B., and D. H. Lynn. 1981. A new macrosystem for the phylum Ciliophora Doflein, 1901. Biosystems 14:387-401.

Sosef, M. S. M. 1997. Hierarchical models, reticulate evolution and the inevitability of paraphyletic supraspecific taxa. Taxon 46:75-85.

Stefanovic, S., D. F. Austin, and R. G. Olmstead. 2003. Classification of Convolvulaceae: A phylogenetic approach. Syst. Bot. 28:791-806.

194

Taylor, M. P., and D. Naish. 2005. The phylogenetic taxonomy of Diplodocoidea (Dinosauria: Sauropoda). PaleoBios 25:1-7.

Thorne, R. F. 2000. The classification and geography of the flowering plants: Dicotyledons of the class angiospermae (Subclasses magnoliidae, ranunculidae, caryophyllidae, dilleniidae, rosidae, asteridae, and lamiidae). Botanical Review 66:441-647.

Wenzel, J. W., K. C. Nixon, and G. Cuccudoro. 2004. Dites non au PhyloCode! Bull. Soc. Fr. Syst. 31:19-23.

Wu, Z. Y., A. M. Lu, Y. C. Tang, Z. D. Chen, and D. Z. Li. 2002. Synopsis of a new "polyphyletic-polychronic-polytopic" system of the angiosperms. Acta Phytotax. Sin. 40:2289-2322.

Yoon, H. S., J. Grant, Y. I. Tekle, M. Wu, B. C. Chaon, J. C. Cole, J. M. Logsdon, D. J. Patterson, D. Bhattacharya, and L. A. Katz. 2008. Broadly sampled multigene trees of eukaryotes. BMC Evol. Biol. 8:14.

195

BIBLIOGRAPHY

Adl, S. M., B. S. Leander, A. G. B. Simpson, S. B. Archibald, O. R. Anderson, D. Bass, S. S. Bowser, G. Brugerolle, M. A. Farmer, S. Karpov, M. Kolisko, C. E. Lane, D. J. Lodge, D. G. Mann, R. Meisterfeld, L. Mendoza, Ø. Moestrup, S. E. Mozley-Standridge, A. V. Smirnov, and F. Spiegel. 2007. Diversity, nomenclature, and taxonomy of protists. Syst. Biol. 56:684-689.

Adl, S. M., A. G. B. Simpson, M. A. Farmer, R. A. Andersen, O. R. Anderson, J. R. Barta, S. S. Browser, G. Brugerolle, R. A. Fensome, S. Fredericq, T. Y. James, S. Karpov, P. Kugrens, J. Krug, C. E. Lane, L. A. Lewis, J. Lodge, D. H. Lynn, D. G. Mann, R. M. McCourt, L. Mendoza, Ø. Moestrup, S. E. Mozley-Standridge, T. A. Nerad, C. A. Shearer, A. V. Smirnov, F. W. Spiegel, and M. Taylor. 2005. The new higher level classification of eukaryotes with emphasis on the taxonomy of protists. J. Eukaryot. Microbiol. 52:399-451.

Aescht, E., W. Foissner, and M. Mulisch. 1991. Ultrastructure of the mycophagous ciliate Grossglockneria acuta (Ciliophora, Colpodea) and phylogenetic affinities of colpodid ciliates. Eur. J. Protistol 26:350-364.

Affa'a, F. M., D. A. Hickey, M. Strüder-Kypke, and D. H. Lynn. 2004. Phylogenetic position of species in the genera Anoplophrya, Plagiotoma, and Nyctotheroides (Phylum Ciliophora), endosymbiotic ciliates of annelids and anurans. J. Eukaryot. Microbiol. 51:301-306.

Agatha, S. 2004a. A cladistic approach for the classification of oligotrichid ciliates (Ciliophora: Spirotricha). Acta Protozool. 43:201-217.

Agatha, S. 2004b. Evolution of ciliary patterns in the oligotrichida (Ciliophora, Spirotricha) and its taxonomic implications. Zoology 107:153-168.

Agatha, S., and W. Foissner. 2009. Conugation in the spirotrich ciliate Halteria grandinella (Müller, 1773) Dujardin, 1841 (Protozoa, Ciliophora) and its phylogenetic implications. Eur. J. Protistol. 45:51-63.

Agatha, S., and M. Strüder-Kypke. 2007. Phylogeny of the order Choreotrichida (Ciliophora, Spriotricha, Oligotrichea) as inferred from morphology, ultrastructure, ontogenesis, and SSrRNA sequences. Europ. J. Protistol. 43:37-63.

Anderson, F. E., and D. L. Swofford. 2004. Should we be worried about long-branch attraction in real data sets? Investigations using metazoan 18S rDNA. Mol. Phylogenet. Evol. 33:440-451.

APG II. 2003. An update of the Angiosperm Phylogeny Group classification for the oders and families of flowering plants: APG II. Bot. J. Linn. Soc. 141:399-436.

196

Ardell, D. H., C. A. Lozupone, and L. F. Landweber. 2003. Polymorphism, recombination and alternative unscrambling in the DNA polymerase alpha gene of the ciliate Stylonychia lemnae (Alveolata, class Spirotrichea). Genetics 165:1761-1777.

Arkhipova, I., and M. Meselson. 2000. Transposable elements in sexual and ancient asexual taxa. Proc. Natl. Acad. Sci. USA 97:14473-14477.

Arkhipova, I., and M. Meselson. 2004. Deleterious transposable elements and the extinction of asexuals. Bioessays 27:76-85.

Ausubel, F. M., R. Brent, R. E. Kingston, D. D. Moore, J. G. Seidman, J. A. Smith, and K. Struhl. 1993. Current Protocols in Molecular Biology. Wiley-Liss, New York.

Baldauf, S. L., A. J. Roger, I. Wenk-Siefert, and W. F. Doolittle. 2000. A kingdom-level phylogeny of eukaryotes based on combined protein data. Science 290:972-977.

Bardele, C. F. 1981. Functional and phylogenetic aspects of the ciliary membrane: a comparative freeze-fracture study. Biosystems 14:403-421.

Bardele, C. F. 1989. From ciliate ontogeny to ciliate phylogeny: a program. Boll. Zool. 56:235-243.

Bardele, C. F., W. Foissner, and R. L. Blanton. 1991. Morphology, morphogenesis and systematic position of the sorocarp forming ciliate Sorogena stoianovitchae Bardbury & Olive 1980. J. Protozool 38:7-17.

Barkley, T. M., P. DePriest, V. Funk, R. W. Kiger, W. J. Kress, and G. Moore. 2004. Linnaean nomenclature in the 21st Century: A report from a workshop on integrating traditional nomenclature and phylogenetic classification. Taxon 53:153-158.

Barth, D., S. Krenek, S. I. Fokin, and T. U. Berendonk. 2006. Intraspecific genetic variation in Paramecium revealed by cytochrome c oxidase I sequences. J. Eukaryot. Microbiol. 53:20-25.

Barth, D., K. Tischer, H. Berger, M. Schlegel, and T. U. Berendonk. 2008. High mitochondrial haplotype diversity of Coleps sp. (Ciliophora: Prostomatida). Environ. Microbiol. 10:626-634.

Bell, G. 1982. The masterpiece of nature: the evolution and genetics of sexuality. University of California Press, Berkeley.

Bell, G. 1988. Sex and death in protozoa: the history of an obsession. Cambridge University Press, Cambridge.

Benton, M. J. 2000. Stems, nodes, crown clades, and rank-free lists: Is Linnaeus dead? Biol. Rev. 75:633-648.

197

Berger, H. 1999. Monograph of the Oxytrichidae (Ciliophora, Hypotrichia). Monographiae Biol. 78:i-xii, 1-1080.

Berger, H. 2006. Monograph of the Urostyloidea (Ciliophora, Hypotrichia). Monographiae Biol. 85:i-xv, 1-1303.

Berger, H., K. A. S. Al-Rasheid, and W. Foissner. 2006. Morphology and cell division of Saudithrix terricola n. gen., n. sp., a large, stichotrich ciliate from Saudi Arabia. J. Eukaryot. Microbiol. 53:260-268.

Berger, H., and W. Foissner. 1997. Cladistic relationships and generic characterization of oxytrichid hypotrichs (Protozoa, Ciliophora). Arch. Protistenk. 148:125-155.

Bergsten, J. 2005. A review of long-branch attraction. Cladistics 21:163-193.

Bertrand, Y., and M. Härlin. 2006. Stability and universality in the application of taxon names in phylogenetic nomenclature. Syst. Biol. 55:848-858.

Bhattacharya, D., T. Friedl, and S. Damberger. 1996. Nuclear-encoded rDNA group I introns: Origin and phylogenetic relationships of insertion site lineages in the green algae. Mol. Biol. Evol. 13:978-989.

Bicknell, R. A., N. K. Borst, and A. M. Koltunow. 2000. Monogenic inheritance of apomixis in two Hieracium species with distinct developmental mechanisms. Heredity 84:228-237.

Birky, C. W. 1996. Heterozygosity, heteromorphy, and phylogenetic trees in asexual eukaryotes. Genetics 144:427-437.

Birky, C. W. 2005. Sex: is Giardia doing in the dark? Curr. Biol. 15:R56-R58.

Bowers, N., T. T. Kroll, and J. R. Pratt. 1998. Diversity and geographic distribution of riboprints from three cosmopolitan species of Colpoda Müller (Ciliophora: Colpodea). Europ. J. Protistol. 34:341-347.

Bradbury, P. C., and L. S. Olive. 1980. Fine structure of the feeding stage of a sorogenic ciliate, Sorogena stoianovitchae gen. n., sp. n. J. Protozool. 27:267-277.

Brothers, D. J. 1983. Nomenclature at the ordinal and higher levels. Syst. Zool. 32:34-42.

Brummitt, R. K. 2003. Further dogged defense of paraphyletic taxa. Taxon 52:803-804.

Bryant, H. N. 1994. Comments on the phylogenetic definitions of taxon names and conventions regarding the naming of crown clades. Syst. Biol. 43:124-130.

Bryant, H. N., and P. D. Cantino. 2002. A review of criticisms of phylogenetic nomenclature: Is taxonomic freedom the fundamental issue? Biol. Rev. 77:39-55.

Bull, J. J., and E. L. Charnov. 1985. On irreversible evolution. 39:1149-1155.

198

Burt, A. 2000. Sex, recombination, and the efficacy of natural selection — was Weisman right? Evolution 54:337-351.

Bütschli, O. 1887-1889. Protozoa. Abt. III. Infusoria und System der Radiolaria in H. G. Bronn, ed. Klassen und Ordnungen de Thier-Reichs, Vol. I. C. F. Winter, Leipzig.

Cannone, J. J., S. Subramanian, M. N. Schnare, J. R. Collett, L. M. D'Souza, Y. S. Du, B. Feng, N. Lin, L. V. Madabusi, K. M. Muller, N. Pande, Z. D. Shang, N. Yu, and R. R. Gutell. 2002. The Comparative RNA Web (CRW) Site: an online database of comparative sequence and structure information for ribosomal, intron, and other RNAs. BMC Bioinformatics 3:2.

Cantino, P. D., and K. de Queiroz. 2007. International Code of Phylogenetic Nomenclature. Version 4b [http://www.ohiou.edu/phylocode/].

Cantino, P. D., J. A. Doyle, S. W. Graham, W. S. Judd, R. G. Olmstead, D. E. Soltis, P. S. Soltis, and M. J. Donoghue. 2007. Towards a phylogenetic nomenclature of Tracheophyta. Taxon 56:E1-E44.

Cantino, P. D., R. G. Olmstead, and S. J. Wagstaff. 1997. A comparison of phylogenetic nomenclature with the current system: A botanical case study. Syst. Biol. 46:313-331.

Carpenter, J. M. 2003. Critique of pure folly. Bot. Rev. 69:79-92.

Catania, F., F. Wurmser, A. A. Potekhin, E. Przyboś, and M. Lynch. 2008. Genetic diversity in the Paramecium aurelia species complex. Mol. Bio. Evol. 26:421-431.

Cavalier-Smith, T. 1999. Zooflagellate phylogeny and the systematics of Protozoa. Biol. Bull. 196:393-396.

Cavalier-Smith, T. 2004. Only six kingdoms of life. Proc. R. Soc. Lond. B 271:1251-1262.

Cavalier-Smith, T. 2007. Evolution and relationships of algae: major branches of the tree of life. Pp. 21-55 in J. Brodie, and J. Lewis, eds. Unravelling the algae: the past, present, and future of algal systematics. CRC Press, Boca Raton.

Ceplitis, A. 2003. Coalescence times and the Meselson effect in asexual eukaryotes. Genet. Res. 82:183-190.

Chantangsi, C., D. H. Lynn, M. T. Brandl, J. C. Cole, N. Hetrick, and P. Ikonomi. 2007. Barcoding ciliates: a comprehensive study of 75 isolates of the genus Tetrahymena. Int. J. Syst. Evol. Microbiol. 57:2412-2425.

199

Chao, A., P. C. Li, S. Agatha, and W. Foissner. 2006. A statistical approach to estimate soil ciliate diversity and distribution based on data from five continents. Oikos 114:479-493.

Chapman, H., G. J. Houliston, B. Robson, and I. Iline. 2003. A case of reversal: the evolution and maintenance of sexuals from parthenogenetic clones in Hieracium pilosella. Int. J. Plant Sci. 164:719-728.

Claff, C. L., V. C. Dewey, and G. W. Kidder. 1941. Feeding mechanisms and nutrition in three species of Bresslaua. Biol. Bull. 81:221-234.

Collin, R., and M. P. Miglietta. 2008. Reversing opinions on Dollo's Law. Trends Ecol. Evol. 23:602-609.

Corliss, J. O. 1961. The ciliated protozoa: Characteriszation, classification, and guide to the literature. Pergamon Press, London.

Corliss, J. O. 1972. Common sense and courtesy in nomenclatural taxonomy. Syst. Zool. 21:117-122.

Corliss, J. O. 1979. The ciliated protozoa: characterization, classification and guide to the literature. Pergamon Press, Oxford.

Corliss, J. O. 1983. Consequences of creating new kingdoms of organisms. Bioscience 33:314-318.

Cummings, M. P., and A. Meyer. 2005. Magic bullets and golden rules: data sampling in molecular phylogenetics. Zoology 108:329-336.

Dacks, J., and A. J. Roger. 1999. The first sexual lineage and the relevance of facultative sex. J. Mol. Evol. 48:779-783.

de Pinna, M. G. G. 1991. Concepts and tests of homology in the cladistic paradigm. Cladistics 7:367-394.

de Queiroz, K. 1997. The Linnaean hierarchy and the evolutionatization of taxonomy, with emphasis on the problem of nomenclature. Aliso 15:125-144.

de Queiroz, K. 2006. The Phylocode and the distinction between taxonomy and nomenclature. Syst. Biol. 55:160-162.

de Queiroz, K. 2007. Toward an intergrated system of clade names. Syst. Biol. 956-974.

de Queiroz, K., and J. Gauthier. 1990. Phylogeny as a central principle in taxonomy: phylogenetic definitions of taxon names. Syst. Biol. 39:307-322.

de Queiroz, K., and J. Gauthier. 1992. Phylogenetic taxonomy. Annual Review Of Ecology And Systematics 23:449-480.

200

de Queiroz, K., and J. Gauthier. 1994. Toward a phylogenetic system of biological nomenclature. Trends Ecol. Evol. 9:27-31.

Detcheva, R. B. 1976. Particularités ultracrturales du cilié Cyrtolophosis mucicola Stokes, 1885. C. r. Séanc. Soc. Biol. Clermon-Ferrand 170:112-114.

Detcheva, R. B. 1982. Recherches, en Bulgarie, sur les cilies de certaines plages de la Mer Niore et des rivieres y aboutissant. Annls. Stn. limnol. Besse 15:231-253.

Detcheva, R. B., and P. d. Puytorac. 1979. Un nouvel de cilie a micronoyau inclus dans l'enveloppe macronucleaire: le genera Aristerostoma Kahl, 1926. Annls. Stn. limnol. Besse 13:247-251.

Díaz, S., A. Martin-González, S. Borniquel, and J. C. Gutiérrez. 2000. Cyrtolophosis elongata (Colpodea, Ciliophora): Some aspects of ciliary pattern, division, cortical and nuclear changes during encystment and resting cyst ultrastructure. Eur. J. Protistol. 36:367-378.

Didier, P., P. Depuytorac, N. Wilbert, and R. Detcheva. 1980. A propos d'observations sur l'ultrastructure de cilié Cyrtolophosis mucicola Stokes, 1885. J. Protozool. 27:72-79.

Dini, F., and D. Nyberg. 1993. Sex in ciliates. Adv. Microb. Ecol. 13:85-153.

Dobell, C. 1932. Antony van Leeuwenhoek and his "little animals". Dover Publications, Inc., New York.

Doherty, M., B. Costas, G. B. McManus, and L. A. Katz. 2007. Culture-independent assessment of planktonic ciliate diversity in coastal northwest Atlantic water. Aquat. Microb. Ecol. 48:141-154.

Domes, K., R. A. Norton, M. Maraun, and S. Scheu. 2007. Reevolution of sexuality breaks Dollo's law. Proc. Natl. Acad. Sci. USA 104:7139-7144.

Dominguez, E., and Q. D. Wheeler. 1997. Taxonomic stability is ignorance. Cladistics 13:367-372.

Donoghue, M. J., T. Eriksson, P. A. Reeves, and R. G. Olmstead. 2001. Phylogeny and phylogenetic taxonomy of Dipsacales with special reference to Sinadoxa and Tetradoxa (Adoxaceae). Harv. Pap. Bot. 6:459-479.

Doyle, J. A., and P. K. Endress. 2000. Morphological phylogenetic analysis of basal angiosperms: comparison and combination with molecular data. Int. J. Plant Sci. 161:S121-S153.

Doyle, J. J. 1992. Gene trees and species trees: molecular systematics as one-character taxonomy. Syst. Bot. 17:144-163.

201

Doyle, J. J. 1997. Trees within trees: genes and species, molecules and morphology. Syst. Biol. 46:537-553.

Dragesco, J., and A. Dragesco-Kernéis. 1979. Cilies muscicoles nouveaux ou peu connus. Acta Protozool. 18:401-416.

Dragesco, J., G. Fryd-versavel, F. Iftode, and P. Didier. 1977. Le cilie Platyophrya spumacola Kahl, 1926: morphologie, stomatogenèse et ultrastructure. Protistologica 13:419-434.

Dubois, A. 2005. Proposed Rules for the incorporation of nomina of higher-ranked zoological taxa in the International Code of Zoological Nomenclature. 1. Some general questions, concepts and terms of biological nomenclature. Zoosystema 27:365-426.

Dubois, A. 2006. Proposed Rules for the incorporation of nomina of higher-ranked zoological taxa in the International Code of Zoological Nomenclature. 2. The proposed Rules and their rationale. Zoosystema 28:165-258.

Dundee, H. A. 1989. Higher category name usage for amphibians and reptiles. Syst. Zool. 38:398-406.

Dunthorn, M., M. Eppinger, M. V. J. Schwarz, M. Schweikert, J. Boenigk, L. A. Katz, and T. Stoeck. 2009. Phylogenetic placement of the Cyrtolophosididae Stokes, 1888 (Ciliophora; Colpodea) and neotypification of Aristerostoma marinum Kahl, 1931. Int. J. Syst. Evol. Microbiol. 59:167-180.

Dunthorn, M., W. Foissner, and L. A. Katz. 2008. Molecular phylogenetic analysis of class Colpodea (phylum Ciliophora) using broad taxon sampling. Mol. Phylogenet. Evol. 48:316-327.

Dunthorn, M., and L. A. Katz. 2008. Richness of morphological hypotheses in ciliate systematics allows for detailed assessment of homology and comparisons with gene trees. Denisia 23:389-394.

Dyke, G. J. 2002. Should paleontologists use "phylogenetic" nomenclature? J. Paleontol. 76:793-796.

Eddy, S. R. 2001. HMMER: Profile hidden markov models for biological sequence analysis. http://hmmer.wustl.edu.

Faivovich, J., C. F. B. Haddad, P. C. A. Garcia, D. R. Frost, J. A. Campbell, and W. C. Wheeler. 2005. Systematic review of the frog family Hylidae, with special reference to Hylinae: a phylogenetic analysis and taxonomic revisison. Bull. Amer. Mus. Nat. Hist. 294.

Farris, J. S. 1974. Formal definitions of paraphyly and polyphyly. Syst. Zool. 23:548-554.

202

Felsenstein, J. 1978. Cases in which parsimony or compatibility methods will be positively misleading. Syst. Zool. 27:401-410.

Finlay, B. J. 2002. Global dispersal of free-living microbial eukaryote species. Science 296:1061-1063.

Finlay, B. J., G. F. Esteban, and T. Fenchel. 1996. Global diversity and body size. Nature 383:132-133.

Fischer, R. A. 1930. The genetical theory of natural selection. Oxford University Press, Oxford.

Foissner, W. 1978. Das Silberliniensystem und die Infraciliatur der Gattungen Platyophrya Kahl, 1926, Cyrtolophosis Stokes, 1885 und Colpoda O.F.M., 1786: Ein Beitrag zur Systematik der Colpidida (Ciliata, Vestibulifera). Acta Protozool. 17:215-231.

Foissner, W. 1985. Klassifikation und phylogenie der Colpodea (Protozoa: Ciliophora). Arch. Protistenk. 129:239-290.

Foissner, W. 1987. Neue und wenig bekannte hypotriche und colpodide Ciliaten (Protozoa: Ciliophora) aus Böden und Moosen. Zool. Beitr. (N. F.) 31:187-282.

Foissner, W. 1991. Basic light and scanning electron microscopic methods for taxonomic studies of ciliated protozoa. Europ. J. Protistol. 27:313-330.

Foissner, W. 1993a. Colpodea (Ciliophora). Protozoenfauna 4/1:i-x, 1-798.

Foissner, W. 1993b. Corticocolpoda kaneshiroae n. g., n. sp., a new colpodid ciliate (Protozoa, Ciliophora) from the bark of ohia trees in Hawaii. J. Eukaryot. Microbiol. 40:764-775.

Foissner, W. 1993c. Idiocolpoda pelobia gen. n., sp. n., a new colpodid ciliate (Protozoa, Ciliophora) from an ephemeral stream in Hawaii. Acta Protozool. 32:175-182.

Foissner, W. 1994. Pentahymena corticicola nov. gen., nov. spec., a new colpodid ciliate (Protozoa, Ciliophora) from bark of acacia trees in Costa Rica. Arch. Protistenk. 144:289-295.

Foissner, W. 1995. Tropical protozoan diversity: 80 ciliate species (Protozoa, Ciliophora) in a soil sample from a tropical dry forest of Costa Rica, with description of four new genera and seven new species. Arch. Protistenk. 145:37-79.

Foissner, W. 1996. Ontogenesis of ciliated protozoa with emphasis on stomatogenesis. Pp. 95-178 in K. Hausmann, and P. C. Bradbury, eds. Ciliates: Cells as Organisms. Gustav Fischer, Stuttgart.

203

Foissner, W. 1997a. Faunistic and taxonomic studies on ciliates (Protozoa, Ciliophora) from clean rivers in Bavaria (Germany), with descriptions of new species and ecological notes. Limnologica 27:179-238.

Foissner, W. 1997b. Global soil ciliate (Protozoa, Ciliophora) diversity: a probability-based approach using large sample collections from Africa, Australia and Antarctica. Biodivers. Conserv. 6:1627-1638.

Foissner, W. 1998. An updated compilation of world soil ciliates (Protozoa, Ciliophora), with ecological notes, new recordsm and descriptions of new species. Europ. J. Protistol. 34:195-235.

Foissner, W. 1999a. Description of two new, mycophagous soil ciliates (Ciliophora, Colpodea): Fungiphrya strobli n. g., n. sp. and Grossglockneria ovata n. sp. J. Eukaryot. Microbiol. 46:34-42.

Foissner, W. 1999b. Protist diversity: estimates of the near-imponderable. Protist 150:363-368.

Foissner, W. 2002. Neotypification of protists, especially ciliates (Protozoa, Ciliophora). Bull. Zool. Nom. 59:165-169.

Foissner, W. 2003. Pseudomaryna australiensis nov gen., nov spec. and Colpoda brasiliensis nov spec., two new colpodids (Ciliophora, Colpodea) with a mineral envelope. Europ. J. Protistol. 39:199-212.

Foissner, W. 2005a. The unusual, lepidosome-coated resting cyst of Meseres corlissi (Ciliophora : Oligotrichea): Transmission electron microscopy and phylogeny. Acta Protozool. 44:217-230.

Foissner, W. 2005b. Two new "flagship" ciliates (Protozoa, Ciliophora) from Venezuela: Sleighophrys pustulata and Luporinophrys micelae. Europ. J. Protistol. 41:99-117.

Foissner, W., S. Agatha, and B. H. 2002. Soil ciliates (Protozoa, Ciliophora) from Namibia (Southwest Africa), with emphasis on two contrasting environments, the Etosha Region and the Namib Desert: Part I. Densia 5:1-1063.

Foissner, W., and H. Berger. 1999. Identification and ontogenesis of the nomen nudum hypotrichs (Protozoa: Ciliophora) Oxytricha nova (= Sterkiella nova sp. n.) and O. trifallax (= S. histriomuscorum). Acta Protozool. 38:215-248.

Foissner, W., H. Berger, and J. Schaumburg. 1999. Indentification and ecology of limnetic plankton ciliates. Inf. Bayer. Landesamt Wasserwirtschaft 3/99:1-793.

Foissner, W., H. Berger, K. Xu, and S. Zechmeister-Boltenstern. 2005a. A huge, undescribed soil ciliate (Protozoa : Ciliophora) diversity in natural forest stands of Central Europe. Biodivers. Conserv. 14:617-701.

204

Foissner, W., H. Blatterer, H. Berger, and F. Kohmann. 1991. Taxonomische und ökologische Revision der Ciliaten des Saprobiensystems—Band I: Cyrtophorida, Oligotrichida, Hypotrichia, Colpodea. Informationsberichte Bayer. Landesmat für Wasserwirtschaft, München.

Foissner, W., A. Chao, and L. A. Katz. 2008. Diversity and geographic distribution of ciliates (Protista: Ciliophora). Biodivers. Conserv. 17:345-363.

Foissner, W., and M. Kreutz. 1998. Systematic position and phylogenetic relationships of the genera Bursaridium, Paracondylostoma, Thylakidium, Bryometopus, and Bursaria (Ciliophora: Colpodea). Acta Protozool. 37:227-240.

Foissner, W., S. Y. Moon-van der Staay, G. W. M. van der Staay, J. H. P. Hackstein, W.-D. Krautgartner, and H. Berger. 2004. Reconciling classical and molecular phylogenies in the stichotrichines (Ciliophora, Spirotrichea), including new sequences from some rare species. Europ. J. Protistol. 40:265-281.

Foissner, W., H. Müller, and S. Agatha. 2007. A comparative fine structural and phylogenetic analysis of resting cysts in oligotrich and hypotrich Spirotrichea (Ciliophora). Europ. J. Protistol. 43:295-314.

Foissner, W., H. Müller, and T. Weisse. 2005b. The unusual, lepidosome-coated resting cyst of Meseres corlissi (Ciliophora : Oligotrichea): Light and scanning electron microscopy, cytochemistry. Acta Protozool. 44:201-215.

Foissner, W., and M. Pichler. 2006. The unusual, lepidosome-coated resting cyst of Meseres corlissi (Ciliophora : Oligotrichea): Genesis of four complex types of wall precursors and assemblage of the cyst wall. Acta Protozool. 45:339-366.

Foissner, W., M. Pichler, K. A. S. Al-Rasheid, and T. Weisse. 2006. The unusual, lepidosome-coated resting cyst of Meseres corlissi (Ciliophora : Oligotrichea): Encystment and genesis and release of the lepidosome. Acta Protozool. 45:323-338.

Foissner, W., and T. Stoeck. 2006. Rigidothrix goiseri nov. gen., nov. spec. (Rigidotrichidae nov. fam.), a new "flagship" ciliate from the Niger floodplain breaks the flexibility-dogma in the classification of stichotrichine spirotrichs (Ciliophora, Spirotrichea). Europ. J. Protistol. 42:249-267.

Foissner, W., M. Struder-Kypke, G. W. M. van der Staay, S. Y. Moon-van der Staay, and J. H. P. Hackstein. 2003. Endemic ciliates (Protozoa, Ciliophora) from tank bromeliads (Bromeliaceae): a combined morphological, molecular, and ecological study. Europ. J. Protistol. 39:365-372.

Foissner, W., and K. Xu. 2007. Monograph of the Spathidiida (Ciliophora: Haptoria), Vol I: Protospathidiidae, Arcuospathidiidae, Apterospathulidae. Monographiae Biol. 81.

205

Foissner, W., K. Xu, and M. Kreutz. 2005c. The Apertospathulidae, a new family of Haptorid ciliates (Protozoa, Ciliophora). J. Eukaryot. Microbiol. 52:360-373.

Gatesy, J., R. DeSalle, and N. Wahlberg. 2007. How many genes should a systematist sample? Conflicting insights from a phylogenomic matrix characterized by replicated incongruence. Syst. Biol. 56:355-363.

Gerassimova, Z. P., and L. N. Seravin. 1976. Ectoplasmic fibrillar system of Infusoria and its role for the understanding of their phylogeny. Zool. Zh. 55:645-656.

Ghiselin, M. T. 1977. On changing the names of higher taxa. Syst. Zool. 26:346-349.

Giribet, G., and W. Wheeler. 2002. On bivalve phylogeny: a high-level analysis of the Bivalvia (Mollusca) based on combined morphology and DNA sequence data. Invert. Biol. 121:271-324.

Givnish, T. J., and K. J. Sytsma. 1997. Consistency, characters, and the likelihood of correct phylogenetic inference. Mol. Phylogenet. Evol. 7:320-330.

Goldberg, E. E., and B. Igic. 2008. On phylogenetic tests of irreversible evolution. Evolution 62:2727-2741.

Golder, T. K. 1976. The macro-micronuclear complex of Woodruffia metabolica. J. Ultrastruct. Res. 54:169-175.

Gong, Y., Y. Yu, P. E. V., F. Zhu, and W. Miao. 2006. Reevaluation of he phylogenetic relationship between mobilid and sessilid peritrichs (Ciliophora, Oligohymenophorea) based on small subunit rRNA gene sequences. J. Eukaryot. Microbiol. 53:397-403.

Gould, S. J. 1970. Dollo on Dollo's Law: irreversibility and the status of evolutionary laws. J. Hist. Biol. 3:189-212.

Grabowski, P. J., A. J. Zaug, and T. R. Cech. 1981. The intervening sequence of the ribosomal RNA precursor is converted to a circular RNA in isolated nuclei of Tetrahymena. Cell 23:467-476.

Graybeal, A. 1998. Is it better to add taxa or characters to a difficult phylogenetic problem? Syst. Biol. 47:9-17.

Greenwood, S. J., M. L. Sogin, and D. H. Lynn. 1991. Phylogenetic relationships within the class oligohymenophorea, phylum Ciliophora, inferred from the complete small subunit rRNA gene sequences of Colpidium campylum, Glaucoma chattoni and Opisthonecta henneguyi. J. Mol. Evol. 33:163-174.

Greslin, A. F., D. M. Prescott, Y. Oka, S. H. Loukin, and J. C. Chappell. 1989. Reordering of 9 exons is necessary to form a functional actin gene in Oxytricha nova. Proc. Natl. Acad. Sci. USA 86:6264-6268.

206

Hamilton, W. D. 2001. Narrow roads of gene land, Vol. 2: the evolution of sex. Oxford University Press, Oxford.

Hammerschmidt, B., M. Schlegel, D. Lynn, D. D. Leipe, M. L. Sogin, and I. B. Raikov. 1996. Insights into the evolution of nuclear dualism in the ciliates revealed by phylogenetic analysis of rRNA sequences. J. Eukaryot. Microbiol. 43:225-230.

Haugen, P., D. H. Coucheron, S. B. Ronning, K. Haugli, and S. Johansen. 2003. The molecular evolution and structural organization of self-splicing group I introns at position 516 in nuclear SSU rDNA of myxomycetes. J. Eukaryot. Microbiol. 50:283-292.

Haugen, P., D. M. Simon, and D. Bhattacharya. 2005. The natural history of group I introns. Trends Genet. 21:111-119.

Hedtke, S. M., T. M. Townsend, and D. M. Hillis. 2006. Resolution of phylogenetic conflict in large data sets by increased taxon sampling. Syst. Biol. 55:522-529.

Hemming, F., ed. 1953. Copenhagen decisions on zoological nomenclature, addirions to, and modeifications of, the Règles internationales de la nomenclature zoologique; approved and adapted by the 14th International Congress of Zoology, Copenhagen, August, 1953. International Trust for Zoological Nomenclature, London.

Hendy, M. D., and D. Penny. 1989. A framework for the quantitative study of evolution. Syst. Zool. 38:297-309.

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

Hewitt, E. A., K. M. Müller, J. Cannone, D. J. Hoga, R. Gutell, and D. M. Prescott. 2003. Phylogenetic relationships among 28 spirotrichous ciliates documented by rDNA. Mol. Phylogenet. Evol. 29:258-267.

Heywood, V. H., R. K. Brummitt, A. Culham, and O. Seberg. 2007. Flowering Plant Families of the World. Firefly Books, Ontario, Canada.

Hibbett, D. S., and M. J. Donoghue. 1998. Intergrating phylogenetic analysis and classification in fungi. Mycologia 90:347-356.

Hillis, D. M. 1998. Taxonomic sampling, phylogenetic accuracy, and investigator bias. Syst. Biol. 47:3-8.

Hillis, D. M., D. D. Pollock, J. A. McGuire, and D. J. Zwickl. 2003. Is sparse taxon sampling a problem for phylogenetic inference? Syst. Biol. 52:124-126.

Hillis, D. M., and J. J. Wiens. 2000. Molecules versus morphology in systematics: conflicts, artifacts, and misconceptions in J. J. Wiens, ed. Phylogenetic analysis of morphological data. Smithsonian Institution Press, Washington.

207

Hirt, R. P., P. L. Dyal, M. Wilkinson, B. J. Finlay, D. M. Roberts, and T. M. Embley. 1995. Phylogenetic relationships among Karyorelictids and Heterotrichs inferred from small subunit rRNA sequences: resolution at the base of the ciliate tree. Mol. Phylogenet. Evol. 4:77-87.

Holder, M. T., J. Sukumaran, and P. O. Lewis. 2008. A justification for reporting the majority-rule consensus tree in Bayesian phylogenetics. Syst. Biol. 57:814-821.

Hörandl, E. 2007. Neglecting evolution is bad taxonomy. Taxon 56:1-5.

Hori, M., I. Tomikawa, E. Przyboś, and M. Fugishima. 2006. Comparison of the evolutionary distances among syngens and sibling species of Paramecium. Molec. Phylogen. Evol. 38:697-704.

Huelsenbeck, J. P., and D. M. Hillis. 1993. Success of phylogenetic methods in the four-taxon case. Syst. Biol. 42:247-264.

Huelsenbeck, J. P., and F. R. Ronquist. 2003. MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics 19:1572-1574.

Hurst, L. D., W. D. Hamilton, and R. J. Ladle. 1992. Covert sex. Trends Ecol. Evol. 7:144-145.

ICZN. 1999. International Code of Zoological Nomenclature, 4th ed. The International Trust for Zoological Nomenclature, London, U.K.

International Commission on Zoological Nomenclature. 1999. International Code of Zoological Nomenclature, 4th ed. The International Trust for Zoological Nomenclature, London.

Israel, R. L., S. L. Kosakovsky Pond, S. V. Muse, and L. A. Katz. 2002. Evolution of duplicated alpha-tubulin genes in ciliates. Evolution 56:1110-1122.

Jackson, S. A., J. J. Cannone, J. C. Lee, R. R. Gutell, and S. A. Woodson. 2002. Distribution of rRNA introns in the three-dimensional structure of the ribosome. J. Mol. Biol. 323:35-52.

Jankowski, A. W. 2007. Phylum Ciliophora Doflein, 1901. Pp. 415-993 in A. F. Alimov, ed. Protista. Part 2, Handbook of Zoology. Russian Academy of Sciences, Zoological Institure, St. Petersburg.

Joyce, W. G., J. F. Parham, and J. A. Gauthier. 2004. Developing a protocol for the conversion of rank-based taxon names to phylogenetically defined clade names, as exemplified by turtles. J. Paleontol. 78:989-1013.

Judo, M. S. B., A. B. Wedel, and C. Wilson. 1998. Stimulation and suppression of PCR-mediated recombination. Nucleic Acids Res. 26:1819-1825.

208

Judson, O. P., and B. B. Normark. 1996. Ancient asexual scandals. Trends Ecol. Evol. 11:A41-A46.

Kahl, A. 1926. Neue und wenig bekannte Formen der holotrichen und heterotrichen Ciliaten. Arch. Protistenk. 55:197-438.

Kahl, A. 1930-1935. Urtiere oder Protozoa. I: Wimpertiere oder Ciliata (Infusoria), eine Bearbeitung der freilebenden und ectocommensalen Infusorien der Erde, unter Ausschluss der marinen Tintinnidae. Pp. 1-886 in F. Dahl, ed. Die Teirwelt Deutschlands. G. Fischer, Jena.

Kahl, A. 1931. Urtiere oder Protozoa I: Wimpertiere oder Ciliate (Infusoria) 2. Holotricha außer den im 1. Teil behandelten Prostomata. Tierwelt Dtl. 21:181-398.

Katz, L. A., J. G. Bornstein, E. Lasek-Nesselquist, and S. V. Muse. 2004a. Dramatic diversity of ciliate histone H4 genes revealed by comparisons of patterns of substitutions and paralog divergences among eukaryotes. Mol. Biol. Evol. 21:555-562.

Katz, L. A., E. Lasek-Nesselquist, J. Bornstein, and S. V. Muse. 2004b. Dramatic diversity of ciliate histone H4 genes revealed by comparisons of patterns of substitutions and paralog divergences among eukaryotes. Mol. Biol. Evol. 21:555-562.

Katz, L. A., G. B. McManus, O. L. O. Snoeyenbos-West, A. Griffin, K. Pirog, B. Costas, and W. Foissner. 2005. Reframing the 'Everything is everywhere' debate: Evidence for high gene flow and diversity in ciliate morphospecies. Aquat. Microb. Ecol. 41:55-65.

Katz, L. A., O. Snoeyenbos-West, and F. P. Doerder. 2006. Patterns of protein evolution in Tetrahymena thermophila: implications for estimates of effective population size. Mol. Biol. Evol. 23:608-614.

Kawakami, H. 1991. An endosymbiotic Chlorella-bearing ciliate: Platyophrya chlorelligera Kawakami 1989. Eur. J. Protistol. 26:245-255.

Keller, R. A., R. N. Boyd, and Q. D. Wheeler. 2003. The illogical basis of phylogenetic nomenclature. Bot. Rev. 69:93-110.

Kim, J. S., H. J. Jeong, D. H. Lynn, J. Y. Park, Y. W. Lim, and W. Shin. 2007. Balanion masanensis n. sp. (Ciliophora: Prostomatea) from the coastal waters of Korea: Morphology and small subunit ribosomal RNA gene sequence. J. Eukaryot. Microbiol. 54:482-494.

Kishino, H., and M. Hasegawa. 1989. Evaluation of the maximum likelihood estimate of the evolutionary tree topologies from DNA sequence data and the branching order in Hominoidea. J. Mol. Evol. 29:170-179.

209

Kojima, J. 2003. Apomorphy-based definition also pinpoints a node, and PhyloCode names prevent effective communication. Bot. Rev. 69:44-58.

Kolaczkowski, B., and J. W. Thornton. 2004. Performance of maximum parsimony and likelihood phylogenetics when evolution is heterogeneous. Nature 431:980-984.

Kondrashov, A. S. 1982. Selection against harmful mutations in large sexual and asexual population. Genet. Res. 40:325-332.

Kondrashov, A. S. 1993. Classification of hypotheses on the advantage of amphimixis. J. Hered. 84:372-387.

Kosakovsky Pond, S. L., S. D. W. Frost, and S. V. Muse. 2005. HyPhy: hypothesis testing using phylogenies. Bioinformatics 21:676-679.

Lasek-Nesselquist, E., and L. A. Katz. 2001. Phylogenetic position of Sorogena stoianovitchae and relationships within the class Colpodea (Ciliophora) based on SSU rDNA sequences. J. Eukaryot. Microbiol. 48:604-607.

Laurin, M., K. de Queiroz, and P. D. Cantino. 2006. Sense and stability of taxon names. Zoologica Scripta 35:113-114.

Lee, M. S. Y. 1996. Stability in meaning and content of taxon names: an evaluation of crown clade definitions. Proc. R. Soc. Lond. B 263:1103-1109.

Lee, M. S. Y. 2000. A worrying systematic decline. Trends Ecol. Evol. 15:346.

Lee, M. S. Y. 2001. On recent arguments for phylogenetic nomenclature. Taxon 50:175-180.

Lefèvre, E., C. Bardot, C. Noël, J. F. Carrias, E. Viscogliosi, C. Amblard, and T. Sime-Ngando. 2007. Unveiling fungal zooflagellates as members of freshwater picoeukaryotes: evidence from a molecular diversity study in a deep meromictic lake. Environ. Microbiol. 9:61-71.

Lin, X., W. Song, and J. Li. 2007. Description of two new marine pleurostomatid ciliates, Loxophyllum choii nov. spec. and L. shini nov spec. (Ciliophora, Pleurostomatida) from China. Europ. J. Protistol. 43:131-139.

Lynch, M., R. Bürger, D. Butcher, and W. Gabriel. 1993. Mutational meltdowns in asexual populations. J. Hered. 84:339-344.

Lynch, M., and J. S. Conery. 2003. The origins of genome complexity. Science 302:1401-1404.

Lynn, D. H. 1976. Comparative ultrastructure and systematics of colpodida. Structural conservatism hypothesis and a description of Colpoda steinii Maupas. J. Protozool. 23:302-314.

210

Lynn, D. H. 1979. Fine structural specializations and evolution of carnivory in Bresslaua (Ciliophora, Colpodida). Trans. Am. Microsc. Soc. 98:353-368.

Lynn, D. H. 1981. The organization and evolution of microtubular organelles in ciliated protozoa. Biol. Rev. 56:243-292.

Lynn, D. H. 1996. My journey in ciliate systematics. J. Eukaryot. Microbiol. 43:253-260.

Lynn, D. H. 2003. Morphology or molecules: how do we identify the major lineages of ciliates (Phylum Ciliophora)? Europ. J. Protistol. 39:356-364.

Lynn, D. H. 2008. The ciliated protozoa: characterization, classification, and guide to the literature, 3rd edition. Springer, Dordrecht.

Lynn, D. H., and E. B. Small. 1997. A revised classification of the phylum Ciliophora Doflein, 1901. Rev. Soc. Mex. Hist. Nat. 47:65-78.

Lynn, D. H., and E. B. Small. 2002. Phylum Ciliophora Doflein, 1901. Pp. 371-656 in J. J. Lee, G. F. Leedale, and P. C. Bradbury, eds. An illustrated guide to the protozoa, 2nd ed. Society of Protozoologists, Lawrence, Kansas.

Lynn, D. H., and M. L. Sogin. 1988. Assessment of phylogenetic relationships among ciliated protists using partial ribosomal RNA sequences derived from reverse transcripts. Biosystems 21:249-254.

Lynn, D. H., and M. Strüder-Kypke. 2002. Phylogenetic position of Licnophora, Lechriopyla, and Schizocaryum, three unusual ciliates (Phylum Ciliophora) endosymbiotic in echinoderms (Phylum echinodermata). J. Eukaryot. Microbiol. 49:460-468.

Lynn, D. H., and M. Strüder-Kypke. 2006. Species of Tetrahymena indentical by small subunut rRNA gene sequences are discriminated by mitochondrial cytochrome c oxidase I gene sequences. J. Eukaryot. Microbiol. 53:385-387.

Lynn, D. H., A.-D. G. Wright, M. Schlegel, and W. Foissner. 1999a. Phylogenetic relationships of orders within the class Colpodea (Phylum Ciliophora) inferred from small subunit rRNA gene sequences. J. Mol. Evol. 48:605-614.

Lynn, D. H., A. D. G. Wright, M. Schlegel, and W. Foissner. 1999b. Phylogenetic relationships of orders within the class Colpodea (phylum Ciliophora) inferred from small subunit rRNA gene sequences. J. Mol. Evol. 48:605-614.

Ma, H. W., R. M. Overstreet, J. H. Sniezek, M. Solangi, and D. W. Coats. 2006. Two new species of symbiotic ciliates from the respiratory tract of cetaceans with establishment of the new genus Planilamina n. gen. (Dysteriida, Kyaroikeidae). J. Eukaryot. Microbiol. 53:407-419.

Maddison, W. P. 1997. Gene trees in species trees. Syst. Biol. 46:523-536.

211

Maddison, W. P., and D. R. Maddison. 2005. MacClade v. 4.0.8. Sinauer Assoc.

Maeda, M. 1986. An illustrated guide to the species of the families Halteriidae and Strobilidiidae (Oligotrichida, Ciliophora), free swimming protozoa common in the aquatic environment.

Mark Welch, D., M. P. Cummings, D. M. Hillis, and M. Meselson. 2004a. Divergent gene copies in the asexual class Bdelloidea (Rotifera) separated before the bdelloid radiation or within bdelloid families. Proc. Natl. Acad. Sci. USA 101:1622-1625.

Mark Welch, D., J. L. Mark Welch, and M. Meselson. 2008. Evidence for degenerate tetraploidy in bdelloid rotifers. Proc. Natl. Acad. Sci. USA 105:5145-5149.

Mark Welch, D., and M. Meselson. 2000. Evidence for the evolution of bdelloid rotifers without sexual reproduction or genetic exchange. Science 288:1211-1215.

Mark Welch, J. L., D. Mark Welch, and M. Meselson. 2004b. Cytogenetic evidence for asexual evolution of bdelloid rotifers. Proc. Natl. Acad. Sci. USA 101:1618-1621.

Marshall, C. R., E. C. Raff, and R. A. Raff. 1994. Dollo's law and the death and resurrection of gene. Proc. Natl. Acad. Sci. USA 91:12283-12287.

Martin, D., and E. Rybicki. 2000. RDP: detection of recombination amongst aligned sequences. Bioinformatics 16:562-563.

Martín-González, A., J. Wierzchos, J. C. Gutiérrez, J. Alonso, and C. Ascaso. 2008. Morphological stasis of protists in lower Cretaceous amber. Protist 159:251-257.

Matthes, D. 1988. Suctoria und Urceolariidae. Protozoenfauna 7/1:i-ix, 1-309.

Maynard Smith, J. 1978. The evolution of sex. Cambridge University Press, Cambridge.

Maynard Smith, J. 1992. Analyzing the mosaic structure of genes. J. Mol. Evol. 34:126-129.

Mayr, E. 1963. Animal species and evolution. The Belknap Press of Harvard University Press, Cambridge.

Mayr, E. 1976. Evolution and the diversity of life. The Belknap Press of Havard University Press, Cambridge.

Mayr, E. 1998. Two empires or three? Proc. Natl. Acad. Sci., USA 95:9720-9723.

Mayr, E., and W. J. Bock. 2002. Classifications and other ordering systems. J. Zoo. Syst. Evol. Research 40:169-194.

212

McGrath, C. L., R. A. Zufall, and L. A. Katz. 2006. Genome evolution in ciliates. Pp. 64-77 in L. A. Katz, and D. Bhattacharya, eds. Genomics and evolution of eukaryotic microbes. Oxford Uniuversity Press, New York.

Mears, J. A., J. J. Cannone, S. M. Stagg, R. R. Gutell, R. K. Agrawal, and S. C. Harvey. 2002. Modeling a minimal ribosome based on comparative sequence analysis. J. Mol. Biol. 321:215-234.

Medlin, L., H. J. Elwood, S. Stickel, and M. L. Sogin. 1988. The characterization of enzymatically amplified eukaryotes 16S-like ribosomal RNA coding regions. Gene 71:491-500.

Miyake, A. 1996. Fertilization and sexuality in ciliates. Pp. 243-290 in K. Hausmann, and P. C. Bradbury, eds. Ciliates: cells as organisms. Gustav Fischer, Stuttgart.

Modesto, S. P., and J. S. Anderson. 2004. The phylogenetic definition of Reptilia. Syst. Biol. 53:815-821.

Moore, G. 1998. A comparison of traditional and phylogenetic nomenclature. Taxon 47:561-579.

Moore, G. 2003. Should taxon names be explicitly defined? Bot. Rev. 69:2-21.

Morin, G. B., and T. R. Cech. 1988. Phylogenetic relationships and altered genome structures among Tetrahymena mitochondrial DNAs. Nucleic Acids Res 16:327-346.

Muller, H. J. 1964. The relation of recombination to mutational advance. Mutation Res. 1:2-9.

Nanney, D. L. 1977. Molecules and morphospecies: the perpetuation of pattern in ciliates protozoa. J. Protozool. 24:27-35.

Nanney, D. L. 1980. Experimental ciliatology: an introduction to genetic and developmental analysis in ciliates. John Wiley & Sons, New York.

Nanney, D. L. 1982. Genes and phenes in Tetrahymena. Bioscience 32:783-788.

Nanney, D. L. 1999. When is a rose?: the kinds of Tetrahymena. Pp. 93-118 in R. A. Wilson, ed. Species: new interdisciplinary essays. The MIT Press, Cambridge.

Nanney, D. L., and J. W. McCoy. 1976. Characterization of the species of the Tetrahymena pyriformis complex. Trans. Am. Microsc. Soc. 95:664-682.

Nanney, D. L., C. Park, R. Preparata, and E. M. Simon. 1998. Comparison of sequence differences in a variable 23S rRNA domain among sets of cryptic species of ciliated protozoa. J. Euk. Microbiol. 45:91-100.

213

Nixon, K. C., and J. M. Carpenter. 2000. On the other "phylogenetic systematics". Cladistics 16:298-318.

Nixon, K. C., J. M. Carpenter, and D. W. Stevenson. 2003. The PhyloCode is fatally flawed, and the "Linnaean" system can easily be fixed. Bot. Rev. 69:111-120.

Njine, T. 1979. Etude ultrastructurale de cilié Kuklikophyra dragescoi gen. n., sp. n. J. Protozool. 26:589-598.

Nordal, H., and B. Stedje. 2005. Paraphyletic taxa should be accepted. Taxon 54:5-6.

Normark, B. B., O. P. Judson, and N. A. Moran. 2003. Genomic signatures of ancient asexual lineages. Biol. J. Linn. Soc. 79:69-84.

Nylander, J. A. 2004. MrModeltest v2. Distrubuted by the author. Evolutionary Biology Center, Uppsala University, Upsalla.

Nylander, J. A., J. C. Wilgenbusch, D. L. Warren, and D. L. Swofford. 2008. AWTY (are we there yet?): a system for graphical exploration of MCMC convergence in Bayesian phyologenetics. Bioinformatics 24:581-583.

Omilian, A. R., M. E. A. Cristescu, J. L. Dudycha, and M. Lynch. 2006. Ameiotic recombination in asexual lineages of Daphnia. Proc. Natl. Acad. Sci. USA 103:18638-18643.

Page, R. D. 1996. TREEVIEW: An application to display phylogenetic tress on personal computers. Comput. Appl. Biosci. 12:357-358.

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, London.

Pernin, P., A. Ataya, and M. L. Cariou. 1992. Genetic structure of natural populations of the free-living ameba, Naegleria lovaniensis. evidence for sexual reproduction. Heredity 68:173-181.

Petz, W., and W. Foissner. 1992. Morphology and morphogenesis of Strobilidium caudatum (Fromentel), Meseres corlissi n. sp., Halteria grandinella (Muller), and Strombidium rehwaldi n. sp., and a proposed phylogenetic system for oligotrich ciliates (Protozoa, Ciliophora). J. Protozool. 39:159-176.

Petz, W., W. Song, and N. Wilbert. 1995. Taxonomy and ecology of the ciliate fauna (Protozoa: Ciliophora) in the endopagial and pelagial of the Weddell Sea, Antarctica. Staphia 4:1-223.

Pfenninger, M., and K. Schwenk. 2007. Cryptic animal species are homogeneously distributed among taxa and biogeographical regions. BMC Evol. Biol. 7:121.

214

Pleijel, F., and G. W. Rouse. 2003. Ceci n'est pas une pipe: names, clades and phylogenetic nomenclature. J. Zool. Syst. Evol. Research 41:162-174.

Poe, S., and D. L. Swofford. 1999. Taxon sampling revisited. Nature 398:299-300.

Polaszek, A., and E. O. Wilson. 2005. Sense and stability in animal names. Trends Ecol. Evol. 20:421-422.

Poljansky, G. 1934. Geschlechtsprozesse bei Bursaria truncatella O.F. Müll. Arch. Protistenk. 81:420-546.

Porter, M. L., and K. A. Crandall. 2003. Lost along the way: significance of evolution in reverse. Trends. Ecol. Evol. 18:541-547.

Posada, D., and K. A. Crandall. 2001. Evaluation of methods for detecting recombination from DNA sequences: computer simulation. Proc. Natl. Acad. Sci. USA 98:13757-13762.

Prescott, D. M. 1992. The unusual organization and processing of genomic DNA in hypotrichous ciliates. Trends Genet. 8:439-445.

Przyboś, E., A. Maciejewska, and B. Skotarczak. 2006. Relationship of species of the Paramecium aurelia complex (Protozoa, ph. Ciliophora, cl. Oligohymenophorea) based on sequences of the histone H4 gene fragment. Folia Biol. (Praha). 54:37-42.

Puytorac, P. d. 1994. Phylum Ciliophora Doflein, 1901. Pp. 1-15 in P. P. de, ed. Traité de Zoologie, Tome II, Infusoires Ciliés, Fasc. 2, Systématique. Masson, Paris.

Puytorac, P. d., A. Batisse, J. Bohatier, J. O. Corliss, G. Deroux, P. Didier, J. Dragesco, G. Fryd-Versavel, J. Grain, C.-A. Groliére, F. Iftode, M. Laval, M. Roque, A. Savoie, and M. Tuffrau. 1974. Proposition d'une classification du phylum Ciliophora Doflein, 1901. C.R. Acad. Sci. Paris 278:2799-2802.

Puytorac, P. d., J. Grain, and P. Legendre. 1994. An attempt at reconstructing a phylogenetic tree of the ciliophora using parsimony methods. Europ. J. Protistol. 30:1-17.

Puytorac, P. d., M. R. Kattar, C. A. Grolière, and I. d. Silva Neto. 1992. Polymorphism and ultrastructure of the colpodean ciliate of the genus Platyophryides Foissner, 1987 J. Protozool. 39:154-159.

Puytorac, P. d., F. Perez-Paniagua, and J. Perez-Silva. 1979. A propos d'observations sur la stomatogenèse et l'ultrastructure de cilié Woodruffia metabolica (Johnson et Larson, 1938). Protistologica 15:231-243.

Raikov, I. B. 1982. The protozoan nucleus: morphology and evolution. Springer-Verlag, Wien.

215

Raikov, I. B. 1996. Nuclei of ciliates. Pp. 221-242 in K. Hausmann, and P. C. Bradbury, eds. Ciliates: cells as organisms. Gustav Fischer, Stuttgart.

Ramesh, M. A., S.-B. Malik, and J. M. Longsdon. 2005. A phylogenomic inventory of meiotic genes: evidence for sex in Giardia and an early eukaryotic origin of meiosis. Curr. Biol. 15:185-191.

Rannala, B., J. P. Huelsenbeck, Z. Yang, and R. Neilsen. 1998. Taxon sampling and the accuracy of large phylogenies. Syst. Biol. 47:702-710.

Reize, I. B., and M. Melkonian. 1989. A new way to investigate living flagellate/ciliated cells in the light microscopr: Immobilization of cells in agarose. Bot. Acta 102:145-151.

Rokas, A., and S. B. Carroll. 2005. More genes or more taxa? The relative contribution of gene number and taxon number to phylogenetic accuracy. Mol. Biol. Evol. 22:1337-1344.

Rokas, A., B. L. Williams, N. King, and S. B. Carroll. 2003. Genome-scale approaches to resolving incongruence in molecular phylogenies. Nature 425:798-804.

Sadler, L. A., and C. F. Brunk. 1992. Phylogenetic relationships and unusual diversity in histone H4 proteins within the Tetrahymena pyriformis complex. Mol. Biol. Evol. 9:70-84.

Sangster, G. 2005. A name for the clade formed by owlet-nightjars, swifts and hummingbirds (Aves). Zootaxa:1-6.

Schander, C., and M. Thollesson. 1995. Phylogenetic Taxonomy―Some Comments. Zool. Scr. 24:263-268.

Schmidt, A. R., H. von Eynatten, and M. Wagreich. 2001. The Mesozoic amber of Schliersee (southern Germany) is Cretaceous in age. Cretaceous Res. 22:423-428.

Schmidt, S. L., D. Bernard, M. Schlegel, and W. Foissner. 2007a. Phylogeny of the Stichotrichea (Ciliophora: Spirotrichea) reconstructed with nulcear small subunit rRNA gene sequences: discrepancies and accordances with morphological data. J. Eukaryot. Microbiol. 54:201-209.

Schmidt, S. L., W. Foissner, M. Schlegel, and D. Bernhard. 2007b. Molecular phylogeny of the heterotrichea (Ciliophora, postciliodesmatophora) based on small subunit rRNA gene sequences. J. Eukaryot. Microbiol. 54:358-363.

Schönborn, W., H. Dörfelt, W. Foissner, L. Krientiz, and U. Schäfer. 1999. A fossilized microcenosis in Triassic amber. J. Eukaryot. Microbiol. 46:571-584.

Schuh, R. 2003. The Linnaean system and its 250-year persistence. Bot. Rev. 69:59-78.

216

Schwander, T., and B. J. Crespi. 2009. Twigs on the tree of life? Neutral and selective models for intergrating macroevolutionary patterns and microevolutionary processes in the analysis of asexuality. Molec. Ecol. 18:28-42.

Scotland, R. W., R. G. Olmstead, and J. R. Bennett. 2003. Phylogeny reconstruction: the role of morphology. Syst. Biol. 52:539-548.

Sereno, P. C. 1999. Definitions in phylogenetic taxonomy: Critique and rationale. Syst. Biol. 48:329-351.

Sereno, P. C. 2005. The logical basis of phylogenetic taxonomy. Systematic Biology 54:595-619.

Shimodaira, H., and M. Hasegawa. 2001. CONSEL: for assessing the confidence of phylogenetic tree selection. Bioinformatics 17:1246-1247.

Shin, M. K., U. W. Hwang, W. Kim, A.-D. G. Wright, C. Krawczyk, and D. H. Lynn. 2000. Phylogenetic position of the ciliates Phacodinium (Order Phacodiniida) and Protocruzia (Subclass Protocruziidia) and systematics of the spirotrich ciliates examined by small subunit ribosomal RNA gene sequences. Europ. J. Protistol. 36:293-302.

Siddall, M. E., and M. F. Whiting. 1999. Long-branch abstractions. Cladistics 15:9-24.

Simon, D., J. Moline, G. Helms, T. Friedl, and D. Bhattacharya. 2005. Divergent histories of rDNA group I introns in the lichen family Physciaceae. J. Mol. Evol. 60:434-446.

Simon, E. M., D. L. Nanney, and F. P. Doerder. 2008. The "Tetrahymena pyformis" complex of cryptic species. Biodivers. Conserv. 17:365-380.

Simpson, G. G. 1953. The major features of evolution. Columbia University Press, New York.

Small, E. B., and D. H. Lynn. 1981. A new macrosystem for the phylum Ciliophora Doflein, 1901. Biosystems 14:387-401.

Small, E. B., and D. H. Lynn. 1985. Phylum Ciliophora, Doflein 1901 in J. J. Lee, S. H. Hutner, and E. C. Bovee eds. An illustrated guide to the protozoa. Society of Protozoologists, Lawrence, KS.

Smith, R. J., T. Kamiya, and D. J. Horne. 2006. Living males of the 'ancient asexual' Darinulidae (Ostracoda: Crustacea). Proc. R. Soc. Lond. B 273:1569-1578.

Snoeyenbos-West, O. L. O., J. Cole, A. Campbell, D. W. Coats, and L. A. Katz. 2004. Molecular phylogeny of phyllopharyngean ciliates and their group I introns. J. Eukaryot. Microbiol. 51:441-450.

217

Snoeyenbos-West, O. L. O., T. Salcedo, G. B. McManus, and L. A. Katz. 2002. Insights into the diversity of choreotrich and oligotrich ciliates (Class : Spirotrichea) based on genealogical analyses of multiple loci. Int. J. Syst. Evol. Microbiol. 52:1901-1913.

Snoke, M. S., T. U. Berendonk, D. Barth, and M. Lynch. 2006. Large global effective population sizes in Paramecium. Mol. Biol. Evol. 23:2474-2479.

Sogin, M. L., and H. J. Elwood. 1986. Primary structure of the Paramecium tetraurelia small-subunit rRNA coding region: phylogenetic relationships within the Ciliophora. J. Mol. Evol. 23:53-60.

Sonneborn, T. M. 1937. Sex, sex inheritance and sex determination in Paramecium aurelia. Proc. Natl. Acad. Sci. 23:378-383.

Sonneborn, T. M. 1957. Breeding systems, reproductive methods, and species problems in protozoa. Pp. 155-324 in E. Mayr, ed. The species problem. American Association for the Advancement of Science, Washington D. C. .

Sonneborn, T. M. 1975. The Paramecium aurelia complex of fourteen sibiling species. Trans. Am. Microsc. Soc. 94:155-178.

Sosef, M. S. M. 1997. Hierarchical models, reticulate evolution and the inevitability of paraphyletic supraspecific taxa. Taxon 46:75-85.

Stamatakis, A. 2006. RAxML-VI-HPC: Maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics 22:2688-2690.

Stamatakis, A., P. Hoover, and J. Rougemont. 2008. A fast boostrap algorithm for the RAxML web-servers. Syst. Biol. 57:758-771.

Stechmann, A., M. Schlegel, and D. H. Lynn. 1998. Phylogenetic relationships between Prostome and Colpodean ciliates tested by small subunit rRNA sequences. Mol. Phyl. Evol. 9:48-54.

Stefanovic, S., D. F. Austin, and R. G. Olmstead. 2003. Classification of Convolvulaceae: A phylogenetic approach. Syst. Bot. 28:791-806.

Stetter, K. O., H. Köning, and E. Stachebrandt. 1983. Pyrodictium gen. nov., a new genus of submarine disc-shaped sulphur reducing archaebacteria growing optimally at 1050C. Appl. Microbiol. 4:535-551.

Stevens, P. F. 1984. Homology and phylogeny: morphology and systematics. Syst. Bot. 9:395-409.

218

Stoeck, T., W. Foissner, and D. H. Lynn. 2007. Small-subunit rDNA phylogenies sugest that Epalxella antiquorum (Penard, 1922) Corliss, 1960 (Ciliophora, Odontostomatida) is a member of the Plagyopylea. J. Eukaryot. Microbiol. 54:436-442.

Stoeck, T., B. Hayward, G. T. Taylor, R. Varela, and S. S. Epstein. 2006. A multiple PCR-primer approach to access the microeukaryotic diversity in environmental samples. Protist 157:31-43.

Stoeck, T., M. V. J. Schwarz, J. Boenigk, M. Schweikert, S. von der Heyden, and A. Behnke. 2005. Cellular identity of an 18S rRNA gene sequence clade within the class Kinetoplastea: the novel genus Actuariola gen. nov. (Neobodonida) with description of the type species Actuariola framvarensis sp. nov. Int. J. Syst. Evol. Microbiol. 55:2623-2635.

Strüder-Kypke, M., O. A. Kornilova, and D. H. Lynn. 2007. Phylogeny of trichostome ciliates (Ciliophora, Litostomatea) endosymbiotic in the Yakut horse (Equus caballus). Europ. J. Protistol. 43:319-328.

Strüder-Kypke, M., A.-D. G. Wright, W. Foissner, and D. H. Lynn. 2006. Molecular phylogeny of Litostome ciliates (Ciliophora, Litostomatea) with emphasis on free-living haptorian genera. Protist 157:261-278.

Strüder-Kypke, M. C., and D. H. Lynn. 2003. Sequence analyses of the small subunit rRNA gene confirm the paraphyly of oligotrich ciliates sensu lato and support the monophyly of the subclasses Oligotrichia and Choreotrichia (Ciliophora, Spirotrichea). J. Zool. 260:87-97.

Swofford, D. 2002. PAUP*. Phylogenetic Analysis Using Parsimony (*and Other Methods). Sinauer Assoc, Sunderland MA.

Szabó, M. 1935. Neuere Beiträge zur Kenntnis der Gattung Halteria (Protozoa, Ciliata). Arch. Protistenk. 86:307-317.

Taylor, M. P., and D. Naish. 2005. The phylogenetic taxonomy of Diplodocoidea (Dinosauria: Sauropoda). PaleoBios 25:1-7.

Teotónio, H., and M. R. Rose. 2001. Perspective: reverse evolution. Evolution 55:653-660.

Thompson, J. D., D. G. Higgins, and T. J. Gibson. 1994. Clustal W: improving the sensitivity of progressive multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix choice. Nucl. Acids Res. 22:4673-4680.

219

Thorne, R. F. 2000. The classification and geography of the flowering plants: Dicotyledons of the class angiospermae (Subclasses magnoliidae, ranunculidae, caryophyllidae, dilleniidae, rosidae, asteridae, and lamiidae). Botanical Review 66:441-647.

Tourancheau, A. B., E. Villalobo, N. Tsao, A. Torres, and R. E. Pearlman. 1998. Protein coding gene trees in ciliates: comparisons with rRNA-based phylogenies. Mol. Phyl. Evol. 10:299-309.

Van Bell, C. T. 1985. The 5s and 5.8s ribosomal RNA sequences of Tetrahymena thermophila and T. pyriformis. J. Protozool. 32:640-644.

van Hannen, E. J., W. M. Mooij, M. P. van Agterveld, H. J. Gons, and H. J. Laanbroek. 1999. Detritus-dependent development of the microbial community in an experimental system: Qualitative analysis by denaturing gradient gel electrophoresis. Appl. Environ. Microbiol. 65:2478-2484.

van Hoek, A. H. A. M., A. S. Akhmanova, M. A. Huynen, and J. H. P. Hackstein. 2000a. A mitochondrial ancestry of the hydrogenosomes of Nyctotherus ovalis. Mol. Biol. Evol. 17:202-206.

van Hoek, A. H. A. M., T. A. van Alen, V. S. I. Sprakel, J. A. M. Leunissen, T. Brigge, G. D. Vogels, and J. H. P. Hackstein. 2000b. Multiple acquisitions of methanogenic Archaeal symbionts by anaerobic ciliates. Mol. Bio. Evol. 17:251-258.

Weisse, T. 2002. The significance of inter- and intraspecific variation in bacterivorous and herbivorous protists. Antonie Van Leeuwenhoek 81:327-341.

Weisse, T., and S. Lettner. 2002. The ecological significance of intraspecific variation among freshwater ciliates. Verh. Internat. Verein. Limnol. 28:1880-1884.

Weisse, T., and S. Rammer. 2006. Pronounced ecophysiological clonal differences of two common freshwater ciliates, Coleps spetai (Prostomatida) and Rimostrombidium lacustris (Oligotrichida), challenge the morphospecies concept. J. Plankton Res. 28:55-63.

Weisse, T., M. C. Strüder-Kypke, H. Berger, and W. Foissner. 2008. Genetic, morphological, and ecological diversity of spatially separated clones of Meseres corlissi Petz & Foissner, 1992 (Ciliophora, Spirotrichea). J. Eukaryot. Microbiol. 55:257-270.

Wenzel, J. W., K. C. Nixon, and G. Cuccudoro. 2004. Dites non au PhyloCode! Bull. Soc. Fr. Syst. 31:19-23.

West, S. A., C. M. Lively, and A. F. Read. 1999. A pluralist approach to sex and recombination. J. Evol. Biol. 12:1003-1012.

220

Wheeler, Q. D. 2008. Undisciplined thinking: morphology and Hennig's unfinished revolution. Syst. Entomol. 33:2-7.

Wikmark, O.-G., P. Haugen, E. W. Lundblad, K. Haugli, and S. D. Johansen. 2007. The molecular evolution and structural organization of group I introns at position 1389 in nuclear small subunit rDNA of myxomycetes. J. Eukaryot. Microbiol. 54:49-56.

Williams, D., and J. C. Clamp. 2007. A molecular phylogenetic investigation of Opisthonecta and related genera (Ciliophora, Peritrichia, Sessilida). J. Eukaryot. Microbiol. 54:317-323.

Williams, G. C. 1975. Sex and evolution. Princeton University Press, Princeton.

Wright, A.-D. G., and D. H. Lynn. 1997. Maximum ages of ciliate lineages estimated using a small subunit rRNA molecular clock: crown eukaryotes date back to the paleoprotoerozoic. Arch. Protistenk. 148:329-341.

Wu, Z. Y., A. M. Lu, Y. C. Tang, Z. D. Chen, and D. Z. Li. 2002. Synopsis of a new "polyphyletic-polychronic-polytopic" system of the angiosperms. Acta Phytotax. Sin. 40:2289-2322.

Ye, A., J., and D. P. Romero. 2002. Phylogenetic relationships amongst tetrahymenine ciliates inferred by a comparison of telomerase RNAs. Int. J. Syst. Evol. Microbiol. 52:2297-3202.

Yi, Z., W. Song, A. Warren, D. M. Roberts, K. A. S. Al-Rasheid, Z. Chen, S. A. Al-Farraj, and X. Hu. 2008. A molecular phylogenetic investigation of Pseudoamphisiellla and Parabirojima (Protozoa, Ciliophora, Spirotrichea), two genera with ambiguous systematic positions. Eur. J. Protistol. 44:45-53.

Yoon, H. S., J. Grant, Y. I. Tekle, M. Wu, B. C. Chaon, J. C. Cole, J. M. Logsdon, D. J. Patterson, D. Bhattacharya, and L. A. Katz. 2008. Broadly sampled multigene trees of eukaryotes. BMC Evol. Biol. 8:14.

Zufall, R. A., C. L. McGrath, S. V. Muse, and L. A. Katz. 2006. Genome architecture drives protein evolution in ciliates. Mol. Biol. Evol. 23:1681-1687.


Recommended