Academia.eduAcademia.edu
Metadata of the book that will be visualized in SpringerLink Publisher Name Springer Berlin Heidelberg Publisher Location Berlin, Heidelberg Series ID SeriesTitle Book ID 272454_1_En Book Title Polyploidy and Genome Evolution Book DOI 10.1007/978-3-642-31442-1 Copyright Holder Name Springer-Verlag Berlin Heidelberg Copyright Year 2012 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 1/8 EC TE D PR OO F Polyploidy and Genome Evolution CO RR 1 UN Editor Proof Layout: T1 Standard SC Chapter No.: FM Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 3/8 Pamela S. Soltis Douglas E. Soltis • Editors 4 Polyploidy and Genome Evolution CO RR EC TE D 5 PR OO 3 6 F 2 UN Editor Proof Layout: T1 Standard SC Chapter No.: FM 123 Editors Pamela S. Soltis Florida Museum of Natural History University of Florida Gainesville, FL USA Douglas E. Soltis Florida Museum of Natural History University of Florida Gainesville, FL USA F 14 15 16 17 18 19 PR OO 20 49 50 51 ISBN 978-3-642-31442-1 (eBook) Springer Heidelberg New York Dordrecht London EC 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 ISBN 978-3-642-31441-4 DOI 10.1007/978-3-642-31442-1 Library of Congress Control Number: 2012945474  Springer-Verlag Berlin Heidelberg 2012 This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now known or hereafter developed. Exempted from this legal reservation are brief excerpts in connection with reviews or scholarly analysis or material supplied specifically for the purpose of being entered and executed on a computer system, for exclusive use by the purchaser of the work. Duplication of this publication or parts thereof is permitted only under the provisions of the Copyright Law of the Publisher’s location, in its current version, and permission for use must always be obtained from Springer. Permissions for use may be obtained through RightsLink at the Copyright Clearance Center. Violations are liable to prosecution under the respective Copyright Law. The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore free for general use. While the advice and information in this book are believed to be true and accurate at the date of publication, neither the authors nor the editors nor the publisher can accept any legal responsibility for any errors or omissions that may be made. The publisher makes no warranty, express or implied, with respect to the material contained herein. CO RR 23 22 24 25 26 27 28 29 TE D 21 UN Editor Proof 7 8 9 10 11 12 13 Printed on acid-free paper Springer is part of Springer Science+Business Media (www.springer.com) Book ISBN: 978-3-642-31441-4 Page: 5/8 Preface 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 D 56 TE 55 Polyploidy (whole-genome duplication; WGD) is common in plants and has long been considered as both an important speciation mechanism and a crucial component of plant genome structure. Analyses of chromosome numbers and hypothesized breaks between diploid and polyploid base numbers have suggested anywhere from 30 to 80 % of all angiosperms are polyploid. While recent polyploids may be easily detected through comparison of chromosome numbers, various processes of diploidization or fractionation may substantially alter chromosome numbers and structure, ultimately masking the evolutionary history of duplication events. In contrast, other footprints of ancient WGD may remain in the genome, even when chromosome numbers no longer carry the signature of past WGDs. Genome sequences and other sources of genomic data tell us that, in fact, all angiosperms, as well as all seed plants, have undergone one or more rounds of polyploidy. Furthermore, ancient WGD characterizes all vertebrates, with subsequent, more recent polyploidization in fishes and amphibians. Ancient WGD is also evident in the genomes of yeast and other fungi. While more common in plants than other major lineages of life, polyploidy is now recognized as a fundamental process in all crown eukaryotes. Polyploidy plays a major role in shaping genome structure and organization and in establishing patterns and mechanisms of gene regulation. In fact, it is now impossible to construct models of genome evolution that do not account for genomic content and genetic interactions contributed by WGD. It has been over 30 years since the publication of a comprehensive treatment of polyploidy [Polyploidy: Biological Relevance, W. H. Lewis (ed.), 1980]. The intervening years have witnessed a technological revolution with a transition from the early days of recombinant DNA to nearly routine genome sequencing of nonmodel organisms and from limited biological computing to high-performance computing networks for the biological sciences. These transformations in methodology and computation permit fresh perspectives on polyploidy and the ability to ask old questions with new tools. Over the past decade, it has been a dream of ours to publish a book that synthesizes the rapid progress in understanding the role of polyploidy in genome EC 54 CO RR 53 PR OO F 52 Book ID: 272454_1_En Date: 16-8-2012 UN Editor Proof Layout: T1 Standard SC Chapter No.: FM v Layout: T1 Standard SC Chapter No.: FM Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 6/8 Preface 106 Gainesville, April 2012 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 PR OO 87 D 86 TE 85 F 105 evolution, and this book is now a reality. In the current volume, we have compiled the expertise of scientists studying polyploid genome evolution from multiple perspectives in phylogenetically diverse organisms. Topics range from the conceptual and theoretical underpinnings of polyploidy (chapters by McGrath and Lynch, Birchler) to processes at work in polyploid genomes (Zielinski and Mittelsten Scheid, Finigan et al., Evans et al.), to patterns of ancient polyploidy and its detection (Burleigh, Paterson et al.), to a series of case studies that both document attributes of genome evolution in focal species and address general properties of polyploid genomes, from ancient polyploids [maize (Schnable and Freeling), legumes (Doyle), vertebrates (Cañestro), fishes (Braasch and Postlethwait), yeast (Hudson and Conant)] to classic model polyploids [cotton (Wendel et al.), tobacco (Kovarik et al.), wheat (Feldman et al.)] to very recent ones [Spartina (Ainouche), Senecio (Hegarty et al.), and Tragopogon (Soltis et al.)]. The emerging paradigm from these studies is that polyploidy—through alterations in genome structure and gene regulation, some of which occur shortly after polyploid formation—generates genetic and phenotypic novelty that manifests itself at the chromosomal, physiological, and organismal levels, with long-term ecological and evolutionary consequences. We thank our many colleagues, students, and postdocs for lively and challenging discussions on polyploidy and its many evolutionary consequences. We further acknowledge the support of the U.S. National Science Foundation (Grants 9624643, 0346437, 0614421, 0919254, and 0922003). 84 CO RR EC 107 UN Editor Proof vi Pamela S. Soltis Douglas E. Soltis Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 7/8 Contents 109 1 Evolutionary Significance of Whole-Genome Duplication . . . . . . . C. L. McGrath and M. Lynch 1 2 Genetic Consequences of Polyploidy in Plants . . . . . . . . . . . . . . . James A. Birchler 21 3 Meiosis in Polyploid Plants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Marie-Luise Zielinski and Ortrun Mittelsten Scheid 33 4 Origins of Novel Phenotypic Variation in Polyploids . . . . . . . . . . Patrick Finigan, Milos Tanurdzic and Robert A. Martienssen 57 5 Identifying the Phylogenetic Context of Whole-Genome Duplications in Plants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . J. Gordon Burleigh 115 116 117 118 119 120 93 7 Genomic Plasticity in Polyploid Wheat . . . . . . . . . . . . . . . . . . . . Moshe Feldman, Avraham Levy, Boulos Chalhoub and Khalil Kashkush 109 8 Maize (Zea mays) as a Model for Studying the Impact of Gene and Regulatory Sequence Loss Following Whole-Genome Duplication . . . . . . . . . . . . . . . . . . . . . . . . . . . . James C. Schnable and Michael Freeling 124 126 127 128 129 130 D Ancient and Recent Polyploidy in Monocots . . . . . . . . . . . . . . . . Andrew H. Paterson, Xiyin Wang, Jingping Li and Haibao Tang 123 125 77 6 121 122 TE 113 114 EC 111 112 CO RR 110 PR OO F 108 UN Editor Proof Layout: T1 Standard SC Chapter No.: FM 9 Polyploidy in Legumes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Jeff J. Doyle 137 147 vii Layout: T1 Standard SC Chapter No.: FM Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 8/8 10 132 133 140 141 142 13 143 144 145 14 146 147 148 149 150 15 151 152 153 16 154 155 PR OO 12 139 Polyploid Evolution in Spartina: Dealing with Highly Redundant Hybrid Genomes. . . . . . . . . . . . . . . . . . . . . . . . . . . . M. Ainouche, H. Chelaifa, J. Ferreira, S. Bellot, A. Ainouche and A. Salmon Allopolyploid Speciation in Action: the Origins and Evolution of Senecio cambrensis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Matthew J. Hegarty, Richard J. Abbott and Simon J. Hiscock D 137 The Early Stages of Polyploidy: Rapid and Repeated Evolution in Tragopogon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Douglas E. Soltis, Richard J. A. Buggs, W. Brad Barbazuk, Srikar Chamala, Michael Chester, Joseph P. Gallagher, Patrick S. Schnable and Pamela S. Soltis 209 225 245 271 TE 136 138 Evolutionary Implications of Genome and Karyotype Restructuring in Nicotiana tabacum L. . . . . . . . . . . . . . . . . . . . . Ales Kovarik, Simon Renny-Byfield, Marie-Angèle Grandbastien and Andrew Leitch 181 F 11 135 Yeast as a Window into Changes in Genome Complexity Due to Polyploidization. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Corey M. Hudson and Gavin C. Conant 293 Two Rounds of Whole-Genome Duplication: Evidence and Impact on the Evolution of Vertebrate Innovations . . . . . . . . Cristian Cañestro 309 EC 134 Jeans, Genes, and Genomes: Cotton as a Model for Studying Polyploidy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Jonathan F. Wendel, Lex E. Flagel and Keith L. Adams CO RR 131 Contents 17 Polyploidy in Fish and the Teleost Genome Duplication . . . . . . . . Ingo Braasch and John H. Postlethwait 18 385 160 Polyploidization and Sex Chromosome Evolution in Amphibians . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Ben J. Evans, R. Alexander Pyron and John J. Wiens 161 Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 411 156 157 158 159 162 UN Editor Proof viii 341 Metadata of the chapter that will be visualized in SpringerLink Book Title Polyploidy and Genome Evolution Series Title Chapter Title Evolutionary Significance of Whole-Genome Duplication Copyright Year 2012 Copyright HolderName Springer-Verlag Berlin Heidelberg Corresponding Author Family Name Lynch Particle Given Name M. Suffix Author Division Department of Biology Organization Indiana University Address Bloomington, IN, USA Email milynch@indiana.edu Family Name McGrath Particle Given Name C. L. Suffix Division Department of Biology Organization Indiana University Address Bloomington, IN, USA Email Abstract Whole-genome duplication (WGD) appears to be a widespread phenomenon, occurring in diverse taxa including many of the model organisms used in molecular, cellular, and developmental biology. It is therefore essential to understand the potential evolutionary consequences for individual duplicated genes, as well as for the lineage as a whole. For example, duplicate genes may undergo pseudogenization or may be maintained due to neofunctionalization, subfunctionalization, or selection for increased dosage or dosage balance. Duplicates created via WGD are maintained at higher rates than single-gene duplicates, perhaps due to dosagebalance constraints. Duplicate-gene maintenance may lead to heterodimerization of an existing homodimer or to the divergence of an entire duplicated network or pathway. Allopolyploids and autopolyploids are likely to undergo different evolutionary pressures due to increased divergence between allopolyploid paralogs and an increased prevalence of multivalent formation at meiosis in autopolyploids. Perhaps most importantly, duplicate-gene loss following a WGD may significantly increase the rate of reproductive isolation between geographically isolated subpopulations and may therefore temporarily increase the speciation rate within polyploid lineages. 1 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 1/20 Chapter 1 4 C. L. McGrath and M. Lynch PR OO F 3 Evolutionary Significance of Whole-Genome Duplication 2 23 1.1 Introduction 11 12 13 14 15 16 17 18 19 20 24 25 26 TE 9 10 EC 8 CO RR 7 D 21 22 Abstract Whole-genome duplication (WGD) appears to be a widespread phenomenon, occurring in diverse taxa including many of the model organisms used in molecular, cellular, and developmental biology. It is therefore essential to understand the potential evolutionary consequences for individual duplicated genes, as well as for the lineage as a whole. For example, duplicate genes may undergo pseudogenization or may be maintained due to neofunctionalization, subfunctionalization, or selection for increased dosage or dosage balance. Duplicates created via WGD are maintained at higher rates than single-gene duplicates, perhaps due to dosage-balance constraints. Duplicate-gene maintenance may lead to heterodimerization of an existing homodimer or to the divergence of an entire duplicated network or pathway. Allopolyploids and autopolyploids are likely to undergo different evolutionary pressures due to increased divergence between allopolyploid paralogs and an increased prevalence of multivalent formation at meiosis in autopolyploids. Perhaps most importantly, duplicate-gene loss following a WGD may significantly increase the rate of reproductive isolation between geographically isolated subpopulations and may therefore temporarily increase the speciation rate within polyploid lineages. 5 6 One of the major findings of the new field of evolutionary genomics is that duplication events involving individual genes or multigene segments arise at rates comparable to the rate of mutation at single-nucleotide sites (Lynch and Conery UN Editor Proof Layout: T1 Standard SC Chapter No.: 1 C. L. McGrath  M. Lynch (&) Department of Biology, Indiana University, Bloomington, IN, USA e-mail: milynch@indiana.edu P. S. Soltis and D. E. Soltis (eds.), Polyploidy and Genome Evolution, DOI: 10.1007/978-3-642-31442-1_1,  Springer-Verlag Berlin Heidelberg 2012 1 Layout: T1 Standard SC Chapter No.: 1 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 2/20 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 F 33 PR OO 32 D 31 TE 30 EC 29 2000, 2003a, b), or possibly at even higher rates (Lipinski et al. 2011). Such observations lend credibility to Ohno’s (1970) early speculation that gene duplication is a major resource for the origin of evolutionary novelties. Moreover, it is now clear that whole-genome duplication (WGD) events have occurred in a wide diversity of phylogenetic lineages, including most of the model systems relied upon in molecular, cellular, and developmental biology. For example, budding yeast is a descendant of an ancient genome duplication (Wolfe and Shields 1997; see Chap. 15, this volume), as is the frog Xenopus laevis (Morin et al. 2006; see Chap. 18, this volume) and the zebrafish (Postlethwait et al. 2000; see Chap. 17, this volume). Many ray-finned fish lineages have experienced additional rounds of WGD (Meyer and Van de Peer 2005; see Chap. 17, this volume), and Ohno’s (1970) suggestion that two WGD events preceded the radiation of the vertebrate lineage has become increasingly credible (Panopoulou and Poustka 2005; Hughes and Liberles 2008; Putnam et al. 2008; see Chap. 16, this volume). Finally, three WGD events are recorded within the genome of Arabidopsis thaliana (Simillion et al. 2002), and nearly all other land-plant genomes appear to harbor a legacy of at least one polyploidization event (Doyle et al. 2008), with a proposed WGD in the ancestor of all seed plants and another in the ancestor of all angiosperms (Jiao et al. 2011). Thus, it is clear that understanding the mechanisms of origin and preservation of duplicate genes promises to reveal not only the ways in which genes acquire new functions and organisms respond to natural selection, but also the roots of organismal diversity across the tree of life. Because genome duplication adds thousands of duplicate genes to the genome, understanding the evolutionary forces that act on individual duplicate genes is critical to our understanding of polyploidization. Processes such as neofunctionalization and subfunctionalization have the potential to influence all gene duplicates, whether created through polyploidization or smaller scale duplication events. It has become increasingly clear, however, that duplicates that arise via polyploidization are subject to unique evolutionary forces, such as increased retention due to dosage-balance constraints. Further, there may be processes that are exclusive to gene duplicates that arise via specific types of polyploidization, such as changes in duplicate-gene expression due to the genomic merger that occurs with allopolyploidization. The relative contributions of these evolutionary forces that give rise to the maintenance and evolution of duplicate genes that arise via WGD, or to the evolution of the genome or species as a whole, are currently unknown. However, discriminating between these forces and their effects is likely to be the subject of much research over the next several years. CO RR 27 28 C. L. McGrath and M. Lynch UN Editor Proof 2 1.2 Fates of Duplicate Genes The fate of the vast majority of duplicate genes arising by segmental duplication is nonfunctionalization of one member of the pair (Lynch and Conery 2000, 2003a, b), and this is expected to occur within a few million years in the absence of any Layout: T1 Standard SC Chapter No.: 1 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 3/20 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 F PR OO 73 74 D 72 TE 70 71 3 intrinsic advantage of a duplicate copy (Watterson 1983; Lynch et al. 2001). Despite this, most genomes that have been studied contain a large number of duplicate genes, some of which are clearly quite ancient (Lynch and Conery 2000). Based on this observation, several mechanisms have been proposed for the permanent preservation of duplicate genes (Hughes 1994; Force et al. 1999; Lynch et al. 2001; Taylor and Raes 2004; Lynch 2007; Innan and Kondrashov 2010): (1) neofunctionalization, whereby one copy acquires a novel, beneficial function at the expense of an essential ancestral function; (2) subfunctionalization, whereby complementary mutations lead to a partitioning of independently mutable subfunctions in the ancestral gene; (3) selection for increased gene product; and (4) the masking of nonfunctional alleles. When a duplicate is maintained by selection for increased gene product, it experiences purifying selection (and may also undergo repeated gene conversion) in order to maintain its ancestral function; this process is likely responsible for the multiple copies of ribosomal RNA genes present in many genomes (e.g., Pinhal et al. 2011). Neofunctionalization, on the other hand, is thought to involve positive selection for the mutation(s) responsible for the new function, generally arising at the expense of an essential original function, thereby preserving both copies. There are many examples of neofunctionalization giving rise to novel gene functions in a variety of organisms, including Arabidopsis (Erdmann et al. 2010), fish (Ngai et al. 1993), vertebrates (Layeghifard et al. 2009), and yeast (Byrne and Wolfe 2007; Tirosh and Barkai 2007). Because one duplicate is undergoing positive selection for a new function while the other is under purifying selection to maintain the ancestral function, asymmetric evolutionary rates between duplicates are often thought to be a hallmark of neofunctionalization (Johnson and Thomas 2007; Han et al. 2009), though purely stochastic mechanisms can also give rise to apparent rate asymmetry (Lynch and Katju 2004). Subfunctionalization may involve positive selection acting on both duplicates if the partitioning of the ancestral functions leads to relaxation of pleiotropic constraints, enabling each ancestral function to be fine-tuned and improved through mutation independently in each copy (Piatigorsky and Wistow 1991; Hughes 1994; Des Marais and Rausher 2008). Alternatively, subfunctionalization may be a completely neutral process if each duplicate copy simply acquires a degenerative mutation that renders it unable to perform one of the ancestral functions (Force et al. 1999). At this point, both copies are needed in order to provide the organism with all of the functionality of the original, single-copy gene, and so both will be maintained in the genome by selection. Although identifying definitive cases of subfunctionalization requires determining that the ancestral gene carried multiple functions that have been partitioned in the daughter duplicates, there are nonetheless several compelling examples (e.g., Force et al. 1999; Altschmied et al. 2002; Yu et al. 2003; Adams and Liu 2007; MacNeil et al. 2008; Semon and Wolfe 2008; Buggs et al. 2010; Deng et al. 2010; Hickman and Rusche 2010; Colon et al. 2011; Froyd and Rusche 2011). In addition to these cases of qualitative subfunctionalization, where duplicates eventually come to be expressed in different tissues or at different times or carry out EC 69 CO RR 68 UN Editor Proof 1 Evolutionary Significance of Whole-Genome Duplication Layout: T1 Standard SC Chapter No.: 1 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 4/20 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 F 118 PR OO 117 D 116 TE 115 different functions relative to each other, quantitative subfunctionalization, in which reduction-of-expression (Force et al. 1999) or activity-reducing mutations (Stoltzfus 1999; Scannell and Wolfe 2008) affect both duplicates, is also possible. In quantitative subfunctionalization, both duplicates acquire partial loss-of-function mutations that affect the same function, again rendering both copies essential for the proper dosage or activity of the gene products. In this case, both copies are preserved while retaining the ancestral gene function. Although few studies demonstrating quantitative subfunctionalization exist, Qian et al. (2010) estimated that this process has been responsible for the maintenance of a large proportion of duplicates in yeast and mammals, whereas Woolfe and Elgar (2007) postulated that sequence evolution in cis-regulatory elements may have caused quantitative subfunctionalization among Fugu duplicates. A related consequence of gene duplication is that it can allow for the differentiation of multimeric subunits, such as the evolution of heterodimers from homodimers. Consider a gene whose protein product forms a homodimer. After duplication of this gene, protein subunits produced by the two duplicates (denoted A and B) may randomly associate to make mixtures of dimers in the ratio 1 AA: 2 AB: 1 BB. If the duplicate genes are identical initially, the AA, AB, and BB dimers will be identical as well. However, subsequent differentiation of the duplicate genes causes the three types of dimer to become distinct. This differentiation could be neutral, or it could be selective. If, for example, there were pleiotropic constraints on the form or function of the pre-duplication homodimer, duplication could allow for escape from these constraints in the AB heterodimer, as each subunit (A and B) can now evolve independently. This can be viewed as a special type of subfunctionalization of duplicates. Winter et al. (2002) showed that a classB floral protein heterodimer had evolved from an ancestral homodimer via this mechanism during the gymnosperm/angiosperm transition. In gymnosperms, GGM2-like genes form homodimers, while the duplicated homologs in eudicots, DEF-like genes and GLO-like genes, form heterodimers. Monocots also have duplicated DEF-like genes and GLO-like genes, but, interestingly, it appears the GLO-like proteins of monocots can both homodimerize and heterodimerize with DEF-like proteins, perhaps representing the transition between the homo- and heterodimerized states (Winter et al. 2002; Kanno et al. 2003; Soltis et al. 2006). A similar process appears to have occurred several times in the evolution of the DUF606 family of transmembrane proteins in bacteria (Lolkema et al. 2008). In bacteria with a single DUF606 gene, the DUF606 proteins are able to insert into the membrane in both orientations, and functional homodimers are formed by two subunits in opposite (antiparallel) orientations. Other species of bacteria, however, have duplicated DUF606 genes located tandemly in an operon. In all of these latter cases, the two protein subunits each have a fixed but opposite orientation in the membrane, and they heterodimerize to form the necessary antiparallel two-domain complex. A phylogenetic analysis of the DUF606 gene family reveals that this process of duplication followed by heterodimerization likely occurred five different times in the history of this gene family lineage (Lolkema et al. 2008). Other proposed examples of this mechanism include SMC proteins (Surcel et al. 2008), EC 114 CO RR 113 C. L. McGrath and M. Lynch UN Editor Proof 4 Layout: T1 Standard SC Chapter No.: 1 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 5/20 5 162 1.3 Fates of Duplicate Genes Arising via WGD 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 184 185 186 187 188 189 190 191 192 193 194 195 196 197 198 PR OO 165 166 In addition to the general preservational processes just mentioned, paralogs resulting from WGD events are subject to unique mechanisms of duplicate-gene maintenance and evolution (Force et al. 1999; Lynch and Conery 2000; Yang et al. 2003; Davis and Petrov 2005; Veitia et al. 2008). Well-studied polyploid species commonly exhibit 25–75 % retention of paralogous gene pairs from the most recent WGD event (reviewed in Lynch 2007; Otto 2007), budding yeast being an exception with only *8 % duplicate-gene preservation (Wolfe and Shields 1997). These are surprisingly high preservational levels, when, as discussed above, the fate of the vast majority of duplicate genes arising by segmental duplication is nonfunctionalization of one duplicate (Lynch and Conery 2000, 2003a, b). Although it is possible that many polyploid species have not yet reached equilibrium and are still in an ongoing phase of duplicate-gene loss, it has become increasingly clear that there are likely to be additional forces acting to preserve duplicate genes arising via WGD. A simple explanation for the large number of preserved duplicates within polyploids is that, unlike single-gene duplicates, WGD duplicates exhibit complete conservation of surrounding regulatory sequences, chromosomal environments, etc. Although this likely contributes somewhat to the pattern of higher duplicate retention in polyploids, it does not explain the observation that different types of genes seem to be preserved following WGD compared to smaller scale duplications. This fact can be better explained by selection for dosage balance among proteins. Due to stoichiometric relationships with other interacting genes (e.g., multi-subunit complexes and numerous pathways involved in metabolism and transcriptional regulation), the functions of a subset of protein-coding loci can be highly influenced by dosage imbalances (Veitia 2002; Papp et al. 2003; Birchler et al. 2005; Veitia et al. 2008). In such cases, duplication of just a single member of a gene interaction may be detrimental and actively selected against. In contrast, following a WGD event, most stoichiometric relationships are initially intact, and therefore subsequent losses of interacting paralogs will be inhibited by selection for proper dosage relationships. Thus, for dosage-dependent genes, the dosage-balance hypothesis predicts an under-representation among duplicates created by singlegene duplications, but an over-representation among those created by WGD (Yang et al. 2003; Davis and Petrov 2005; Veitia et al. 2008). For example, Davis and Petrov (2005) showed that the pool of preserved duplicates from the WGD event in S. cerevisiae is enriched for ribosomal genes (which form a large complex) and regulatory genes encoding transferases, kinases, and transcription factors, while D 164 TE 163 EC 160 CO RR 159 F 161 adenylyl cyclases (Sinha et al. 2005), and mitochondrial peptidases (Brown et al. 2007), all gene families that contain duplicates that form heterodimers in eukaryotes (or eukaryotic mitochondria) with single-copy homologs that form homodimers in prokaryotes. 158 UN Editor Proof 1 Evolutionary Significance of Whole-Genome Duplication Layout: T1 Standard SC Chapter No.: 1 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 6/20 206 207 208 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 236 237 238 239 240 241 242 243 F 205 PR OO 204 D 203 TE 202 EC 201 those involved in ion transport are under-represented. Likewise, the Paramecium tetraurelia genome exhibits elevated retention of duplicate genes involved in known complexes (Aury et al. 2006) and in metabolic pathways (Gout et al. 2009). As in yeast, ribosomal genes, transferases, and kinases are over-represented among surviving paralogs, while ion-transport genes are underrepresented. In Paramecium, there also appears to be an additional effect whereby highly expressed genes are over-retained in duplicate following the most recent polyploidization event (Gout et al. 2010). That certain types of genes are maintained preferentially following a WGD has achieved fairly convincing empirical support from other studies as well (Papp et al. 2003; Yang et al. 2003; Barker et al. 2008; Liang et al. 2008; Qian and Zhang 2008; Edger and Pires 2009), including studies in Arabidopsis (Blanc and Wolfe 2004; Maere et al. 2005; Thomas et al. 2006), vertebrates (Makino and McLysaght 2010), and across divergent species (Paterson et al. 2006). Selection to maintain dosage balance following WGD has also been hypothesized to be the driving force behind the original selective advantage of the WGD in the Saccharomyces cerevisiae lineage (Conant and Wolfe 2007). In this scenario, the maintenance of glycolytic genes and the loss of non-glycolytic genes following WGD might have increased the relative dosage of glycolytic genes, thereby increasing flux through the glycolysis pathway and providing polyploid yeast with a growth advantage over non-polyploids due to increased glucose fermentation ability. Duplicate genes that arise via WGD are further unique in that entire (or partial) duplicated pathways or networks of interacting proteins can diverge in concert. For example, Evlampiev and Isambert (2007) modeled the evolution of protein–protein interaction networks following WGD and concluded that such networks grow under exponential, rather than time-linear, dynamics following WGD. Interestingly, they also found that these exponential dynamics relied on asymmetric divergence between duplicates. Another intriguing possibility is that following WGD, a whole ancestral network may become neofunctionalized or subfunctionalized following polyploidization, with one set of paralogs carrying out one task or reaction and a parallel set of paralogs carrying out a related, but largely independent, task. Obviously, such innovations require the establishment of multiple mutations and the avoidance of pathway crosstalk. Although the essential population genetic theory remains to be worked out, several examples of such paralog coevolution appear to have followed the WGD in yeast: parallel paralogous networks have been identified where the expression of each gene is highly correlated with the other genes within its network but poorly correlated with its paralog (Blanc and Wolfe 2004; Conant and Wolfe 2006). In this way, polyploidy provides a unique mechanism for the evolution of gene networks with new (or subdivided) functions. A final consideration in duplicate-gene evolution is whether the forces that act to preserve duplicates change over evolutionary time. For example, it seems possible that following WGD, a large proportion of genes could be initially maintained due to dosage-balance constraints. Subsequently, however, over longer periods of evolutionary time, some duplicates might accumulate mutations that could lead to neofunctionalization or subfunctionalization. Because these genes CO RR 199 200 C. L. McGrath and M. Lynch UN Editor Proof 6 Layout: T1 Standard SC Chapter No.: 1 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 7/20 7 271 1.4 Autopolyploidy Versus Allopolyploidy 251 252 253 254 255 256 257 258 259 260 261 262 263 264 265 266 267 268 269 272 273 274 275 276 277 278 279 280 281 282 283 284 PR OO 249 250 D 248 TE 247 EC 246 CO RR 245 F 270 are dosage sensitive (hence their initial preservation due to selection for dosage balance), it is likely that such neo- or subfunctionalizing mutations would need to be preceded or rapidly followed by mutations affecting the dosage of one or both copies. For a more detailed example, imagine proteins A and B that must interact in a 1:1 ratio for proper functioning. Both genes become duplicated during a WGD, giving rise to duplicates A1 and A2 and B1 and B2. Initially, all four genes are preserved by selection for dosage balance, as loss of any one gene interrupts the 1:1 interaction ratio. Over evolutionary time, however, slightly deleterious mutations in the A1 promoter that decrease its expression level become fixed due to drift. To compensate, mutations in the A2 promoter that increase its expression level are fixed, which helps to restore the 1:1 A/B ratio. At this point, A1 is contributing fewer products to the overall A protein pool. A subsequent mutation that changes the function of A1, allowing it to take on a new role completely, is now more easily accommodated, as A2 is better able to compensate and take on the full load of the ancestral A activity. Note that, instead of A1 and A2 dosage evolving in concert, as above, A1 and B1 dosage could also evolve in concert to maintain the proper 1:1 A/B ratio, allowing both A1 and B1 to take on new functions. While still just a verbal theory, this scenario has two advantages in terms of allowing for neofunctionalization (or subfunctionalization) of WGD duplicates. First, there is a longer time frame in which neo- or subfunctionalizing mutations can arise, as duplicates are maintained for longer time-scales without becoming nonfunctionalized. This is important because neofunctionalization requires the accumulation of beneficial mutation(s), which are thought to be rare. Second, this process would allow for neo- or subfunctionalization of dosage-sensitive duplicates, both of which might otherwise be constrained to maintain their ancestral function indefinitely following WGD. 244 Whether polyploidization occurs by autopolyploidy or allopolyploidy can have a significant impact on the expression and evolution of duplicate genes. Autopolyploids arise when there is an increase in ploidy within a single species (often within a single individual), while allopolyploids are created by hybridization between two different species, each of which contributes a full complement of chromosomes to the hybrid, thus doubling the genome (reviewed in Coyne and Orr 2004). Many plant and frog polyploids are the result of allopolyploidization (Adams 2007; Evans 2008), while the yeast WGD appears to have been an autopolyploidization event (Scannell et al. 2007), although in practice it is difficult to ascertain the ancestral state once paralog divergence has become high. It has long been assumed that autopolyploids would initially form multivalents at meiosis, with all four homologous chromosomes pairing randomly, while allopolyploids would be more likely to form bivalents, with homologous UN Editor Proof 1 Evolutionary Significance of Whole-Genome Duplication Layout: T1 Standard SC Chapter No.: 1 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 8/20 291 292 293 294 295 296 297 298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 F 290 PR OO 289 D 288 TE 287 chromosomes from each diploid ancestor pairing independently. This would mean that duplicate copies in autopolyploids would not represent true paralogs as the term is usually understood, but would instead represent a doubling of the number of homologs (i.e., four homologs instead of two). The presence of multivalents is significant biologically, as multivalent pairing can lead to intergenomic recombination via segregation, crossing-over, and double reduction. Certain duplicates from one diploid parent could be lost completely via this process, leaving only duplicates from the other diploid parent. This would not represent gene silencing as it is typically understood then, but would rather be a byproduct of multivalent formation and segregation. Evidence from plants indicates that multivalent pairing is indeed more prevalent among autopolyploids, though the difference between the two forms of polyploidy is perhaps less than originally expected: a survey of plant polyploids indicated that the mean percent occurrence of multivalents is 28.8 % in autopolyploids and 8.0 % in allopolyploids (Ramsey and Schemske 2002). Although multivalent formation occurs at a lower rate in allopolyploids, it may be more biologically significant than multivalent formation in autopolyploids, as intergenomic recombination is likely to have a greater effect when genomes are more divergent. Over time, divergence between duplicated chromosomes would lead to increased bivalent formation. Because allopolyploids are the result of a genomic merger between two species, duplicate genes in allopolyploids are already differentiated to some extent immediately after polyploidization, while duplicates in autopolyploids are likely to be more similar in sequence and may even be identical. Allopolyploids often exhibit immediate changes in gene expression due to the genetic differentiation present between homeologs. This can lead to changes in methylation (Salmon et al. 2005; Gaeta et al. 2007), changes in heterochromatin formation and transposable element suppression (Josefsson et al. 2006), biased expression of homeologs (Adams et al. 2003; Bottley et al. 2006; Tate et al. 2006; Udall et al. 2006; Rapp et al. 2009), and non-additive expression effects between homeologs (Hegarty et al. 2006; Wang et al. 2006; Rapp et al. 2009). These initial expression differences between homeologs can, in turn, impact the long-term evolution of duplicates, as selection pressures may be expected to act differently on genes that are differentially expressed. For example, Anderson and Evans (2009) showed that in octoploid and dodecaploid Xenopus species, paralogs of RAG1b were more likely to become pseudogenized than paralogs of RAG1a (the homeolog of RAG1b from an earlier allopolyploidy event), and they inferred that this was due to differences in ancestral expression between RAG1a and RAG1b. Many of these effects seen in allopolyploids are believed to be due to the hybridization between two divergent genomes, rather than genome doubling per se. Flagel et al. (2008) estimated that of the genes with biased expression between homeologs in the allopolyploid Gossypium hirsutum, 24 % exhibit a bias due to the genomic merger (i.e., the bias existed immediately when the allopolyploidization occurred, at time zero), while the bias in the remaining 76 % is due to long-term evolutionary forces such as neofunctionalization and subfunctionalization. The relationship between the magnitude of these alterations in gene expression and the EC 286 CO RR 285 C. L. McGrath and M. Lynch UN Editor Proof 8 Layout: T1 Standard SC Chapter No.: 1 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 9/20 9 336 1.5 Polyploidization and Speciation 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 PR OO 337 Perhaps the most pivotal role that polyploidization plays in evolution is in the creation of new species. The polyploidization event itself can lead to instantaneous reproductive isolation and speciation, as the cross between a new tetraploid (4n) and its diploid progenitor (2n) yields triploid (3n) offspring, which are often sterile due to problems with chromosome pairing/segregation during meiosis and the production of aneuploid gametes (reviewed in Coyne and Orr 2004). It is for this reason that models predict that species capable of self-fertilization are more likely to give rise to a successful polyploid lineage (Rodriguez 1996; Baack 2005; Rausch and Morgan 2005). Perhaps more importantly, however, once a polyploid lineage is established, subsequent silencing of duplicate genes can lead to further reproductive isolation among subpopulations of the polyploids themselves and, therefore, give rise to additional daughter species (Oka 1988; Werth and Windham 1991; Lynch and Conery 2000; Lynch and Force 2000). In this model, we assume a pair of fully functional and redundant duplicate genes, A and B, in an ancestral population, such that each member of the initial population has the genotype AABB (Fig. 1.1). If two subpopulations become geographically isolated and one duplicate becomes nonfunctionalized in each subpopulation, there is a 50 % probability that a different duplicate copy will be lost in each of the two groups. This reciprocal gene loss (or divergent resolution) would result in the genotypes aaBB and AAbb for the two subpopulations, where a and b denote null alleles. Hybridization between the two groups would then lead to offspring with the genotype AaBb. Gametes produced by these F1 individuals would have a 1/4 probability of carrying an ab genotype and would therefore be inviable if a functional copy of the A/B gene were essential for gamete survival or function. Even if this were not the case, 1/16 of the F2 individuals would have the aabb genotype and, if a functional copy were essential for zygote viability or sterility, would be inviable or sterile, whereas another 1/4 would have three null alleles and might experience reduced viability or sterility. Up to 50–65 % of the genes encoding transcription factors, membrane receptors, and members of macromolecular protein complexes are estimated to be haploinsufficient (JimenezSanchez et al. 2001; Veitia 2002), suggesting that only one functional allele of such genes is indeed likely to be deleterious. An appealing aspect of the divergent resolution model is that it is a natural consequence of degenerative mutations, requiring no adaptive evolution at the D 334 TE 333 EC 332 CO RR 331 F 335 genetic divergence between the two parental genomes is not well understood, however, as demonstrated by Brassica allopolyploids (Pires and Gaeta 2011). While the parental species that gave rise to the allopolyploid Brassica napus are more similar to each other than those that gave rise to B. juncea, resynthesized B. napus polyploids exhibit more genomic rearrangements, changes in gene expression, and epigenetic alterations than do resynthesized B. juncea polyploids. 330 UN Editor Proof 1 Evolutionary Significance of Whole-Genome Duplication Layout: T1 Standard SC Chapter No.: 1 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 10/20 C. L. McGrath and M. Lynch Fig. 1.1 Divergent resolution of duplicate genes can lead to hybrid sterility/ inviability and reproductive isolation. Red bar represents a gene that becomes duplicated. See text for details PR OO F Gene Duplication CO RR EC TE D Population Isolation and Divergent Resolution UN Editor Proof 10 Hybridization F1 progeny 1/4 F1 gametes (no functional copy) 1/16 F2 progeny (no functional copy) OR 1/4 F2 progeny (hemizygous) Layout: T1 Standard SC Chapter No.: 1 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 11/20 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 F 376 PR OO 375 D 374 TE 373 11 molecular level for speciation to occur. Moreover, in addition to genes that are reciprocally silenced, duplicate pairs that undergo neofunctionalization or subfunctionalization may also contribute to hybrid sterility/inviability in a similar fashion—for example if a different duplicate becomes neofunctionalized and loses its ancestral function in each subpopulation, or if the two duplicates become subfunctionalized in complementary ways in the two subpopulations (Lynch and Force 2000). The process of reciprocal gene loss has been shown to be responsible for male sterility between hybrids of Drosophila melanogaster and D. simulans (Masly et al. 2006). A gene essential for male fertility, JYAlpha, is located on the fourth chromosome in D. melanogaster and on the third chromosome in D. simulans. This translocation presumably occurred via duplication of the JYAlpha gene and subsequent silencing of one copy. The difference in chromosomal location of the gene in the two species causes a proportion of hybrids to completely lack JYAlpha, leading to their sterility. Two similar cases were recently identified in rice. The first involves reproductive isolation between two subspecies of Oryza sativa. The ancestral O. sativa genome appears to have had a pair of duplicates termed DOPPELGANGER1 (DPL1) and DOPPELGANGER2 (DPL2) (Mizuta et al. 2010). The subspecies japonica and indica have experienced independent losses of one copy each: DPL1 has become a pseudogene in indica, while DPL2 has been nonfunctionalized in japonica. Hybrid pollen lacking a functional copy of either DPL1 or DPL2 is nonfunctional and does not germinate, contributing to the partial reproductive isolation present between the subspecies. This validates an earlier hypothesis by Oka (1988) that F1 sterility between japonica and indica was caused by ‘‘duplicate gametophytic sterility genes’’, japonica being homozygous for one nonfunctional copy and indica being homozygous for another nonfunctional copy. In the second rice example, reciprocal loss of one of the duplicated nuclear genes encoding mitochondrial ribosomal protein L27 in O. sativa and O. glumaepatula again causes a proportion of the pollen produced by F1 hybrids to be sterile (Yamagata et al. 2010). The final example of reproductive isolation through reciprocal gene loss comes from A. thaliana, where the histidinol-phosphate amino-transferase gene appears in different chromosomal locations (as in the Drosophila example, presumably via duplication and subsequent silencing of one copy) in the Columbia and Cape Verde Island accessions (Bikard et al. 2009). F2 offspring homozygous for both null alleles completely lack the gene’s product, HPA, which results in arrested embryo development and seed abortion. In addition, in at least one intermediate heterozygote, a quantitative phenotype termed ‘‘weak root’’ was observed, suggesting that the presence of three null alleles is somewhat deleterious in this cross. As these four examples constitute *1/3 of the dozen or so successful searches for the genes underlying the speciation process (most in Drosophila species; Presgraves 2010), there now seems little question that the passive nonfunctionalization of duplicate genes is a major mechanism of speciation. EC 372 CO RR 371 UN Editor Proof 1 Evolutionary Significance of Whole-Genome Duplication Layout: T1 Standard SC Chapter No.: 1 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 12/20 C. L. McGrath and M. Lynch 439 1.6 Unsolved Problems 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 PR OO 420 421 D 419 TE 418 EC 417 The maintenance of duplicate genes via selection for increased gene product, neofunctionalization, and subfunctionalization has been hypothesized for nearly 40 years (Ohno 1970). Recent genetic and genomic data have now identified compelling examples of these processes and have further contributed to our understanding of the prevalence of whole-genome duplications and the dosagebalance theory of duplicate maintenance. However, a number of unresolved questions related to WGDs and duplicate maintenance merit further scrutiny. The first avenue for future study involves a more comprehensive understanding of the relative importance of the forces behind duplicate-gene maintenance, including maintenance for increased dosage, dosage-balance constraints, neofunctionalization, and subfunctionalization. All of these mechanisms have been demonstrated to be responsible for duplicate maintenance in certain cases, but it remains unclear which, if any, is responsible for maintaining the majority of duplicate genes or how such contributions vary among phylogenetic lineages. Most likely, there will be no single driving force for duplicate maintenance but the relative strength of these forces will differ among taxonomic groups or among functional CO RR 416 F 438 These examples demonstrate that the divergent resolution of even one duplicated gene can lead to detectable reproductive isolation. However, genetic incompatibility between two populations can be magnified substantially when reciprocal gene loss occurs at hundreds or thousands of duplicated loci simultaneously, as is the case in polyploid lineages. The probability that an F2 offspring obtained by outcrossing will be double null for at least one of n pairs of divergently resolved loci is 1-(15/16)n, which takes on values of 0.063, 0.276, 0.476, and 0.998 for n = 1, 5, 10, and 100, respectively. Moreover, in species that undergo autogamy or selfing, such as Paramecium, this probability can be as high as 1-(3/4)n, giving probabilities of 0.250, 0.763, 0.944, and & 1 for n = 1, 5, 10, and 100. Speciation events will continue to occur as long as duplicates are still being resolved between subpopulations, leading to nested rounds of speciation, and, because a large number of duplicates are thought to be silenced quickly following WGD (Scannell et al. 2006), a cluster of speciation events might occur within a brief period of time. The net result is the expected generation of a species radiation following a WGD event. It has been suggested that this nested speciation process might be responsible for the radiations of the polyploid yeast species (Scannell et al. 2006), teleost fishes (Semon and Wolfe 2007; see Chap. 15, this volume), angiosperms (Soltis et al. 2009; though see Mayrose et al. 2011), and the Paramecium aurelia species complex (Aury et al. 2006). This mechanism may also be responsible for reproductive isolation between mutagenized lines of an experimentally derived allotetraploid created by hybridizing two species of Saccharomyces (Maclean and Greig 2010). 415 UN Editor Proof 12 Layout: T1 Standard SC Chapter No.: 1 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 13/20 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 F 462 PR OO 461 D 459 460 TE 458 13 classes of genes. For example, subfunctionalization of duplicates may be more likely within species that have evolved a modular (and therefore independently mutable) regulatory structure. Such modular systems are predicted to arise more easily within species with smaller population sizes (Force et al. 2005), demonstrating how species-level features of an organism may influence the evolutionary forces acting upon duplicate genes. More comprehensive studies of large numbers of duplicates from a variety of organisms are required to address what other features might influence the relative strengths of mechanisms of duplicate maintenance. Such studies must not only detail what genes remain duplicated vs. single-copy, but must also detail whether existing duplicates have the same function as each other (to assess rates of neofunctionalization) or share functions with the pre-duplicated ancestor (to assess rates of subfunctionalization) (Fig. 1.2). There are few data on the rate of duplicate-gene loss over time following a WGD, though data from polyploid yeast species suggest that the rate changes over time (Scannell et al. 2006). Data from additional taxa would aid in determining whether this is a general pattern among all WGDs (Fig. 1.3). A related unresolved question is whether the evolutionary forces controlling duplicate maintenance change over time following a WGD, e.g., whether a dosage-sensitive gene may initially be preserved due to selection for dosage balance, but then evolve a new function concurrent with its release from such dosage constraints. An analysis of this question could be made by comparing the fates of duplicate genes in multiple lineages descended from a single WGD event. Such an analysis might identify duplicates that had been maintained due to dosage constraints in the majority of daughter lineages but that had become neofunctionalized in one lineage, perhaps suggesting a secondary mechanism of retention. The unsolved question that promises to be the hardest to answer is why certain lineages or taxonomic groups appear to contain more WGD events than others. This is not the same as asking why certain groups contain more polyploid species, as this may be a simple reflection of the fact that WGD may promote subsequent reproductive isolation and speciation. The pattern remains, however, that some phylogenetic groups seem to contain more independent WGD events than others in their evolutionary pasts. For example, a recent analysis estimated that among ferns, 31 % of speciation events involve polyploidization, while the value for angiosperms is only 15 % (Wood et al. 2009). Similarly, in the history of the Xenopus lineage, there are many more instances of WGD events than compared to, say, mammals. Several factors could contribute to such patterns, such as the ability to hybridize and form allopolyploids, or the ability to self-fertilize (at least transiently) or undergo asexual reproduction, which helps a polyploid lineage become abundant in a surrounding world of diploids. It is not even understood whether mechanistic reasons (at meiosis, say) or differences in developmental programs would facilitate or hinder creation of a viable polyploid in certain lineages, or whether discrepancies in ecological persistence of polyploid species alone are able to explain the patterns that we see. Perhaps the best way to approach such a question is to study closely related lineages where one exhibits several WGD EC 457 CO RR 456 UN Editor Proof 1 Evolutionary Significance of Whole-Genome Duplication Layout: T1 Standard SC Chapter No.: 1 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 14/20 14 C. L. McGrath and M. Lynch Expression Level PR OO F A Conditions/Tissues B or D Editor Proof Fig. 1.2 Distinguishing between the evolutionary forces that maintain duplicate genes. Panel A shows the expression level of a gene (purple) across different conditions or tissues before duplication. The bottom six panels (B–G) show patterns that might be seen for the two copies (red and blue) once the gene has been duplicated and what evolutionary processes these patterns would indicate. Note that panel D might indicate maintenance for increased dosage in the case of a single-gene duplicate or maintenance for dosage balance in the case of a duplicate arising via WGD TE Nonfunctionalization EC D 500 501 UN CO RR Dosage/dosage balance F Qualitative Subfunctionalization C Nonfunctionalization E Neofunctionalization G Quantitative Subfunctionalization events and the other does not, though teasing apart the mechanistic and ecological differences is certain to remain a challenge for decades to come. Layout: T1 Standard SC Chapter No.: 1 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 15/20 1 Evolutionary Significance of Whole-Genome Duplication (a) D C B Outgroup A D % of duplicates lost (b) 505 506 507 508 509 510 511 512 513 514 515 516 517 518 EC 504 F Species 4 B C D Time Whole-genome duplications are widespread across the tree of life and appear in the evolutionary history of a large number of model organisms. Processes such as neo- and subfunctionalization affect retention of individual gene duplicates, and dosage-balance constraints promote the retention of large sets of genes following polyploidization. Allopolyploidization, through hybridization and subsequent changes or biases in homeolog expression, has the ability to instantaneously create a population of individuals that are ecologically and epigenetically unique from either parent lineage, providing a new lineage upon which natural selection can act. Both allo- and autopolyploidization provide a unique opportunity for the differentiation of new gene networks and pathways through concerted evolution of duplicated, interacting proteins. Most importantly, WGD can lead to reproductive isolation through divergent resolution of duplicated genes, thus creating new species and species groups. Further understanding of the relative importance and the temporal properties of the forces acting on polyploid species and the duplicate genes within their genomes promises to enhance our knowledge of the origins of species as well as genetic, protein network, and organismal complexity. CO RR 503 1.7 Conclusions UN 502 Species 3 PR OO WGD Species 1 Species 2 A TE Editor Proof Fig. 1.3 Determining the rate of duplicate-gene loss over time. a An example of a tree for four species that share a whole-genome duplication (WGD). Duplicate-gene presence/absence information for each of the four species could be used to infer the number of duplicate genes lost on each of the branches labeled A, B, C, and D. b The data on gene retention/loss gathered in A could be used to plot the percentage of duplicate genes lost per unit time (or divergence). Depending on where points A, B, C, or D land on the graph, the data may indicate that the rate of gene loss remains constant (top dashed line) or changes (bottom dotted line) over time 15 Layout: T1 Standard SC Chapter No.: 1 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 16/20 C. L. McGrath and M. Lynch References 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 Adams KL (2007) Evolution of duplicate gene expression in polyploid and hybrid plants. J Hered 98(2):136–141 Adams KL, Cronn R et al (2003) Genes duplicated by polyploidy show unequal contributions to the transcriptome and organ-specific reciprocal silencing. Proc Nat Acad Sci U S A 100:4649– 4654 Adams KL, Liu Z (2007) Expression partitioning between genes duplicated by polyploidy under abiotic stress and during organ development. Curr Biol 17(19):1669–1674 Altschmied J, Delfgaauw J et al (2002) Subfunctionalization of duplicate mitf genes associated with differential degeneration of alternative exons in fish. Genetics 161(1):259–267 Anderson DW, Evans BJ (2009) Regulatory evolution of a duplicated heterodimer across species and tissues of allopolyploid clawed frogs (Xenopus). J Mol Evol 68:236–247 Aury J-M, Jaillon O et al (2006) Global trends of whole-genome duplications revealed by the ciliate Paramecium tetraurelia. Nature 444:171–178 Baack EJ (2005) To succeed globally, disperse locally: effects of local pollen and seed dispersal on tetraploid establishment. Heredity 94:538–546 Barker MS, Kane NC et al (2008) Multiple paleopolyploidizations during the evolution of the compositae reveal parallel patterns of duplicate gene retention after millions of years. Mol Biol Evol 25(11):2445–2455 Bikard D, Patel D et al (2009) Divergent evolution of duplicate genes leads to genetic incompatibilities within A. thaliana. Science 323:623–626 Birchler JA, Riddle NC et al (2005) Dosage balance in gene regulation: biological implications. Trends Genet 21(4):219–226 Blanc G, Wolfe KH (2004) Functional divergence of duplicated genes formed by polyploidy during Arabidopsis evolution. Plant Cell 16:1679–1691 Bottley A, Xia GM et al (2006) Homoeologous gene silencing in hexaploid wheat. Plant J 47:897–906 Brown MT, Goldstone HMH et al (2007) A functionally divergent hydrogenosomal peptidase with protomitochondrial ancestry. Mol Microbiol 64(5):1154–1163 Buggs RJA, Elliott NM et al (2010) Tissue-specific silencing of homoeologs in natural populations of the recent allopolyploid Tragopogon mirus. New Phytol 186:175–183 Byrne KP, Wolfe KH (2007) Consistent patterns of rate asymmetry and gene loss indicate widespread neofunctionalization of yeast genes after whole-genome duplication. Genetics 175(3):1341–1350 Colon M, Hernandez F et al (2011) Saccharomyces cerevisiae Bat1 and Bat2 aminotransferases have functionally diverged from the ancestral-like Kluyveromyces lactis orthologous enzyme. PLoS ONE 6(1):e16099 Conant GC, Wolfe KH (2006) Functional partitioning of yeast co-expression networks after genome duplication. PLoS Biol 4(4):0545–0554 Conant GC, Wolfe KH (2007) Increased glycolytic flux as an outcome of whole-genome duplication in yeast. Mol Syst Biol 3:129 Coyne JA, Orr HA (2004) Speciation. Sinauer Associates, Inc, Sunderland, MA Davis JC, Petrov DA (2005) Do disparate mechanisms of duplication add similar genes to the genome? Trends Genet 21(10):548–551 Deng C, Cheng CHC et al (2010) Evolution of an antifreeze protein by neofunctionalization under escape from adaptive conflict. Proc Nat Acad Sci U S A 107(50):21593–21598 Des Marais DL, Rausher MD (2008) Escape from adaptive conflict after duplication in an anthocyanin pathway gene. Nature 454(7205):762–765 Doyle JJ, Flagel LE et al (2008) Evolutionary genetics of genome merger and doubling in plants. Annu Rev Genet 42:443–461 Edger PP, Pires JC (2009) Gene and genome duplications: the impact of dosage-sensitivity on the fate of nuclear genes. Chromosome Res 17:699–717 CO RR EC TE D PR OO F 519 UN Editor Proof 16 Layout: T1 Standard SC Chapter No.: 1 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 17/20 17 EC TE D PR OO F Erdmann R, Gramzow L et al (2010) GORDITA (AGL63) is a young paralog of the Arabidopsis thaliana B (sister) MADS box gene ABS (TT16) that has undergone neofunctionalization. Plant J 63(6):914–924 Evans BJ (2008) Genome evolution and speciation genetics of clawed frogs (Xenopus and Silurana). Front Biosci 13:4687–4706 Evlampiev K, Isambert H (2007) Modeling protein network evolution under genome duplication and domain shuffling. BMC Syst Biol 1:49 Flagel LE, Udall JA et al (2008) Duplicate gene expression in allopolyploid gossypium reveals two temporally distinct phases of expression evolution. BMC Biol 6:16 Force A, Cresko WA et al (2005) The origin of subfunctions and modular gene regulation. Genetics 170(1):433–446 Force A, Lynch M et al (1999) The preservation of duplicate genes by complementary, degenerative mutations. Genetics 151:1531–1545 Froyd CA, Rusche LN (2011) The duplicated deacetylases sir2 and hst1 subfunctionalized by acquiring complementary inactivating mutations. Mol Cell Biol 31(16):3351–3365 Gaeta RT, Pires JC et al (2007) Genomic changes in resynthesized brassica napus and their effect on gene expression and phenotype. Plant Cell 19(11):3403–3417 Gout J-F, Duret L, et al. (2009) Differential retention of metabolic genes following wholegenome duplication. Mol Biol Evol 26(5):1067–1072 Gout J-F, Kahn D, et al. (2010) The relationship among gene expression, the evolution of gene dosage, and the rate of protein evolution. PLoS Genetics 6(5):e1000944 Han MV, Demuth JP et al (2009) Adaptive evolution of young gene duplicates in mammals. Genome Res 19(5):859–867 Hegarty MJ, Barker GL et al (2006) Transcriptome shock after interspecific hybridization in Senecio is ameliorated by genome duplication. Curr Biol 16:1652–1659 Hickman MA, Rusche LN (2010) Transcriptional silencing functions of the yeast protein Orc1/ Sir3 subfunctionalized after gene duplication. Proc Nat Acad Sci U S A 107(45):19384–19389 Hughes AL (1994) The evolution of functionally novel proteins after gene duplication. Proc R Soc Lond B Biol Sci 256:119–124 Hughes T, Liberles DA (2008) Whole-genome duplications in the ancestral vertebrate are detectable in the distribution of gene family sizes of tetrapod species. J Mol Evol 67(4):343–357 Innan H, Kondrashov F (2010) The evolution of gene duplications: classifying and distinguishing between models. Nat Rev Genet 11:97–108 Jiao YN, Wickett NJ et al (2011) Ancestral polyploidy in seed plants and angiosperms. Nature 473(7345):97–100 Jimenez-Sanchez G, Childs B et al (2001) Human disease genes. Nature 409:853–855 Johnson DA, Thomas MA (2007) The monosaccharide transporter gene family in Arabidopsis and rice: a history of duplications, adaptive evolution, and functional divergence. Mol Biol Evol 24(11):2412–2423 Josefsson C, Dilkes B et al (2006) Parent-dependent loss of gene silencing during interspecies hybridization. Curr Biol 16:1322–1328 Kanno A, Saeki H et al (2003) Heterotopic expression of class B floral homeotic genes supports a modified ABC model for tulip (Tulipa gesneriana). Plant Mol Biol 52:831–841 Layeghifard M, Pirhaji L et al (2009) Adaptive evolution in the Per gene family of vertebrates: neofunctionalization by positive Darwinian selection after two major gene duplications. Biol Rhythm Res 40(6):433–444 Liang H, Plazonic KR et al (2008) Protein under-wrapping causes dosage sensitivity and decreases gene duplicability. PLoS Genet 4(1):0072–0077 Lipinski KJ, Farslow JC et al (2011) High spontaneous rate of gene duplication in Caenorhabditis elegans. Curr Biol 21:306–310 Lolkema JS, Dobrowolski A et al (2008) Evolution of antiparallel two-domain membrane proteins: tracing multiple gene duplication events in the DUF606 family. J Mol Biol 378:596–606 Lynch M (2007) The origins of genome architecture. Sinauer Associates, Sunderland, MA CO RR 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 UN Editor Proof 1 Evolutionary Significance of Whole-Genome Duplication Layout: T1 Standard SC Chapter No.: 1 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 18/20 EC TE D PR OO F Lynch M, Conery JS (2000) The evolutionary fate and consequences of duplicate genes. Science 290(5494):1151–1155 Lynch, M, Conery JS (2003a). The evolutionary demography of duplicate genes. In: Meyer A, Van de Peer Y (eds) Genome evolution. Kluwer Academic Publishers, Dordrecht, 35–44 Lynch M, Conery JS (2003) The origins of genome complexity. Science 302:1401–1404 Lynch M, Force AG (2000) The origin of interspecific genomic incompatibility via gene duplication. Am Naturaliste 156(6):590–605 Lynch M, Katju V (2004) The altered evolutionary trajectories of gene duplicates. Trends Genet 20:544–549 Lynch M, O’Hely M et al (2001) The probability of preservation of a newly arisen gene duplicate. Genetics 159:1789–1804 Maclean CJ, Greig D (2010) Reciprocal gene loss following experimental whole-genome duplication causes reproductive isolation in yeast. Evolution 65(4):932–945 MacNeil AJ, McEachern LA et al (2008) Gene duplication in early vertebrates results in tissuespecific subfunctionalized adaptor proteins: CASP and GRASP. J Mol Evol 67(2):168–178 Maere S, De Bodt S et al (2005) Modeling gene and genome duplications in eukaryotes. Proc Nat Acad Sci U S A 102(15):5454–5459 Makino T, McLysaght A (2010) Ohnologs in the human genome are dosage balanced and frequently associated with disease. Proc Nat Acad Sci U S A 107:9270–9274 Masly JP, Jones CD et al (2006) Gene transposition as a cause of hybrid sterility in Drosophila. Science 313:1448–1450 Mayrose I, Zhan SH et al (2011) Recently formed polyploid plants diversify at lower rates. Science 333(6047):1257 Meyer A, Van de Peer Y (2005) From 2R to 3R: evidence for a fish-specific genome duplication (FSGD). BioEssays 27:937–945 Mizuta Y, Harushima Y et al (2010) Rice pollen hybrid incompatibility caused by reciprocal gene loss of duplicated genes. Proc Nat Acad Sci U S A 107(47):20417–20422 Morin RD, Chang E et al (2006) Sequencing and analysis of 10,967 full-length cDNA clones from Xenopus laevis and Xenopus tropicalis reveals post-tetraploidization transcriptome remodeling. Genome Res 16:796–803 Ngai J, Dowling MM et al (1993) The family of genes encoding odorant receptors in the channel catfish. Cell 72:657–666 Ohno S (1970) Evolution by gene duplication. Springer, Berlin Oka HI (1988). Functions and genetic bases of reproductive barriers. Origin of cultivated rice. Japan Scientific Societies Press/Elsevier, HI Oka, Tokyo, pp 181–209 Otto SP (2007) The evolutionary consequences of polyploidy. Cell 131:452–462 Panopoulou G, Poustka AJ (2005) Timing and mechanism of ancient vertebrate genome duplications—the adventure of a hypothesis. Trends Genet 21:559–567 Papp B, Pal C et al (2003) Dosage sensitivity and the evolution of gene families in yeast. Nature 424:194–197 Paterson AH, Chapman BA et al (2006) Many gene and domain families have convergent fates following independent whole-genome duplication events in Arabidopsis, Oryza, Saccharomyces and Tetraodon. Trends Genet 22(11):597–602 Piatigorsky J, Wistow G (1991) The recruitment of crystallins—new functions precede gene duplication. Science 252(5009):1078–1079 Pinhal D, Yoshimura TS et al (2011) The 5S rDNA family evolves through concerted and birthand-death evolution in fish genomes: an example from freshwater stingrays. BMC Evol Biol 11:151 Pires JC, Gaeta RT (2011) Structural and functional evolution of resynthesized polyploids. In: Schmidt R, Bandcroft I (eds) Genetics and genomics of the brassicaceae, Springer, New York, 9:195–214 Postlethwait JH, Woods IG et al (2000) Zebrafish comparative genomics and the origins of vertebrate chromosomes. Genome Res 10:1890–1902 CO RR 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 C. L. McGrath and M. Lynch UN Editor Proof 18 Layout: T1 Standard SC Chapter No.: 1 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 19/20 19 EC TE D PR OO F Presgraves DC (2010) The molecular evolutionary basis of species formation. Nat Rev Genet 11:175–180 Putnam NH, Butts T et al (2008) The amphioxus genome and the evolution of the chordate karyotype. Nature 453:1064–1071 Qian W, Zhang J (2008) Gene dosage and gene duplicability. Genetics 179:2319–2324 Qian WF, Liao BY et al (2010) Maintenance of duplicate genes and their functional redundancy by reduced expression. Trends Genet 26(10):425–430 Ramsey J, Schemske DW (2002) Neopolyploidy in flowering plants. Annu Rev Ecol Syst 33:589–639 Rapp RA, Udall JA et al (2009) Genomic expression dominance in allopolyploids. BMC Biol 7:18 Rausch JH, Morgan MT (2005) The effect of self-fertilization, inbreeding depression, and population size on autopolyploid establishment. Evolution 59(9):1867–1875 Rodriguez DJ (1996) A model for the establishment of polyploidy in plants. Am Naturalist 147(1):33–46 Salmon A, Ainouche ML et al (2005) Genetic and epigenetic consequences of recent hybridization and polyploidy in Spartina (Poaceae). Mol Ecol 14:1163–1175 Scannell DR, Byrne KP et al (2006) Multiple rounds of speciation associated with reciprocal gene loss in polyploid yeasts. Nature 440:341–345 Scannell DR, Frank AC et al (2007) Independent sorting-out of thousands of duplicated gene pairs in two yeast species descended from a whole-genome duplication. Proc Nat Acad Sci U S A 104(20):8397–8402 Scannell DR, Wolfe KH (2008) A burst of protein sequence evolution and a prolonged period of asymmetric evolution follow gene duplication in yeast. Genome Res 18(1):137–147 Semon M, Wolfe KH (2007) Reciprocal gene loss between Tetraodon and zebrafish after whole genome duplication in their ancestor. Trends Genet 23(3):108–112 Semon M, Wolfe KH (2008) Preferential subfunctionalization of slow-evolving genes after allopolyploidization in Xenopus laevis. Proc Nat Acad Sci U S A 105(24):8333–8338 Simillion C, Vandepoele K et al (2002) The hidden duplication past of Arabidopsis thaliana. Proc Nat Acad Sci U S A 99:13627–13632 Sinha SC, Wetterer M et al (2005) Origin of asymmetry in adenylyl cyclases: structures of Mycobacterium tuberculosis Rv1900c. EMBO J 24:663–673 Soltis DE, Albert VA et al (2009) Polyploidy and angiosperm diversification. Am J Bot 96(1):336–348 Soltis PS, Soltis DE et al (2006) Expression of floral regulators in basal angiosperms and the origin and evolution of ABC-function. Adv Bot Res 44:483–506 Stoltzfus A (1999) On the possibility of constructive neutral evolution. J Mol Evol 49:169–181 Surcel A, Zhou X et al (2008) Long-term maintenance of stable copy number in the eukaryotic SMC family: origin of a vertebrate meiotic SMC1 and fate of recent segmental duplicates. J Syst Evol 46(3):405–423 Tate JA, Ni Z et al (2006) Evolution and expression of homeologous loci in Tragopogon miscellus (Asteraceae), a recent and reciprocally formed allopolyploid. Genetics 173:1599– 1611 Taylor JS, Raes J (2004) Duplication and divergence: The evolution of new genes and old ideas. Annu Rev Genet 38:615–643 Thomas BC, Pedersen B et al (2006) Following tetraploidy in an Arabidopsis ancestor, genes were removed preferentially from one homeolog leaving clusters enriched in dose-sensitive genes. Genome Res 16:934–946 Tirosh I, Barkai N (2007) Comparative analysis indicates regulatory neofunctionalization of yeast duplicates. Genome Biol 8(4):R50 Udall JA, Swanson JM et al (2006) A novel approach for characterizing expression levels of genes duplicated by polyploidy. Genetics 173:1823–1827 Veitia RA (2002) Exploring the etiology of haploinsufficiency. BioEssays 24(2):175–184 CO RR 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 UN Editor Proof 1 Evolutionary Significance of Whole-Genome Duplication Layout: T1 Standard SC Chapter No.: 1 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 20/20 EC TE D PR OO F Veitia RA, Bottani S et al (2008) Cellular reactions to gene dosage imbalance: genomic, transcriptomic, and proteomic effects. Trends Genet 24(8):390–397 Wang J, Tian L et al (2006) Genomewide nonadditive gene regulation in Arabidopsis allotetraploids. Genetics 172:507–517 Watterson GA (1983) On the time for gene silencing at duplicate loci. Genetics 105(3):745–766 Werth CR, Windham MD (1991) A model for divergent, allopatric speciation of polyploid Pteridophytes resulting from silencing of duplicate-gene expression. Am Naturalist 137(4):515–526 Winter KU, Weiser C et al (2002) Evolution of class B floral homeotic proteins: obligate heterodimerization orginated from homodimerization. Mol Biol Evol 19(5):587–596 Wolfe KH, Shields DC (1997) Molecular evidence for an ancient duplication of the entire yeast genome. Nature 387:708–713 Wood TE, Takebayashi N et al (2009) The frequency of polyploid speciation in vascular plants. Proc Nat Acad Sci US A 106(33):13875–13879 Woolfe A, Elgar G (2007) Comparative genomics using Fugu reveals insights into regulatory subfunctionalization. Genome Biol 8(4):R53 Yamagata Y, Yamamoto E et al (2010) Mitochondrial gene in the nuclear genome induces reproductive barrier in rice. Proc Nat Acad Sci U S A 107(4):1494–1499 Yang J, Lusk R et al (2003) Organismal complexity, protein complexity, and gene duplicability. Proc Nat Acad Sci U S A 100(26):15661–15665 Yu W-P, Brenner S et al (2003) Duplication, degeneration and subfunctionalization of the nested synapsin-Timp genes in Fugu. Trends Genet 19:180–183 CO RR 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 C. L. McGrath and M. Lynch UN Editor Proof 20 Metadata of the chapter that will be visualized in SpringerLink Book Title Polyploidy and Genome Evolution Series Title Chapter Title Genetic Consequences of Polyploidy in Plants Copyright Year 2012 Copyright HolderName Springer-Verlag Berlin Heidelberg Corresponding Author Family Name Birchler Particle Given Name James A. Suffix Abstract Division Division of Biological Sciences Organization University of Missouri Address 117 Tucker Hall, 65211, Missouri, Colombia Email birchlerj@missouri.edu Most eukaryotes have a history of whole-genome multiplication events followed by a progressive return to a more diploid state. The initial state of polyploidization, in which more than two copies of the genome are present, is considered here and the various types of genetic consequences that occur depending on the nature of the polyploid formed. The degree of association of chromosomes in meiosis is determined by the relative homology and will affect the segregation of the chromosome which determines the genetic properties. If all the chromosomes are quite similar and form associations of like type, this situation is referred to as autopolyploidy. If the different sets of multiple chromosomes are sufficiently dissimilar to each other, then the homologs will pair in meiosis with themselves and segregate independently of the different but related chromosome pair. This situation is referred to as allopolyploidy. Gene expression in ploidal series typically follows a per cell level correlated more or less with the number of sets of chromosomes present. Variation of individual chromosomes, or aneuploidy, produces a greater number of modulations of gene expression in parallel to classical studies noting that aneuploids have greater impact on the phenotype than changes in the copy number of the whole genome. The genetic properties of odd-number ploidies, such as triploids, are also described as well as higher ploidal levels such as hexaploidy and octoploidy. 1 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 21/32 Chapter 2 4 James A. Birchler 11 12 13 14 15 16 17 18 19 20 21 22 23 24 PR OO D 9 10 TE 8 EC 7 Abstract Most eukaryotes have a history of whole-genome multiplication events followed by a progressive return to a more diploid state. The initial state of polyploidization, in which more than two copies of the genome are present, is considered here and the various types of genetic consequences that occur depending on the nature of the polyploid formed. The degree of association of chromosomes in meiosis is determined by the relative homology and will affect the segregation of the chromosome which determines the genetic properties. If all the chromosomes are quite similar and form associations of like type, this situation is referred to as autopolyploidy. If the different sets of multiple chromosomes are sufficiently dissimilar to each other, then the homologs will pair in meiosis with themselves and segregate independently of the different but related chromosome pair. This situation is referred to as allopolyploidy. Gene expression in ploidal series typically follows a per cell level correlated more or less with the number of sets of chromosomes present. Variation of individual chromosomes, or aneuploidy, produces a greater number of modulations of gene expression in parallel to classical studies noting that aneuploids have greater impact on the phenotype than changes in the copy number of the whole genome. The genetic properties of oddnumber ploidies, such as triploids, are also described as well as higher ploidal levels such as hexaploidy and octoploidy. CO RR 5 6 F 3 Genetic Consequences of Polyploidy in Plants 2 UN Editor Proof Layout: T1 Standard SC Chapter No.: 2 J. A. Birchler (&) Division of Biological Sciences, University of Missouri, 117 Tucker Hall, 65211 Missouri, Colombia e-mail: birchlerj@missouri.edu P. S. Soltis and D. E. Soltis (eds.), Polyploidy and Genome Evolution, DOI: 10.1007/978-3-642-31442-1_2,  Springer-Verlag Berlin Heidelberg 2012 21 Layout: T1 Standard SC Chapter No.: 2 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 22/32 25 J. A. Birchler 2.1 Introduction 56 2.2 Allopolyploids 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 57 58 59 60 61 62 63 PR OO 32 D 31 TE 29 30 EC 28 CO RR 27 F 55 Most eukaryotes have a history of polyploidization followed by fractionation back to a near diploid level (Wolfe and Shields 1997; Simillion et al. 2002; Bowers et al. 2003; Blanc and Wolfe 2004; Chapman et al. 2006; Maere et al. 2005; Blomme et al. 2006; Freeling and Thomas 2006; Barker et al. 2008). Thus, at the least, polyploidy in essence is a matter of degree, and it has played an important role in the composition of the gene repertoire of many species. Typically, it is defined as the presence of more copies of the whole genome than the normal two that constitute a diploid (Stebbins 1947). However, from the standpoint of gene content, the determination of whether a species is a polyploid is somewhat arbitrary and dependent on the time before the present when the copy number of the genome was increased. Nevertheless, for evolutionarily ‘‘recent’’ events, certain principles can apply which will be summarized in this chapter. In the polyploidy literature, the basic chromosome number is designated by x and consists of the complete set of chromosomes, or a genome. The number of chromosomes in the gametophyte generation and hence the gametes is referred to as the gametic chromosome number or n. In diploids, x = n, but at higher levels of polyploidy, this is not the case. Polyploidy is typically divided into at least two categories that are determined by the type of chromosome pairing in meiosis I and the distribution of chromosomes during this process. Indeed, the type of chromosome pairing that occurs in meiosis affects the genetic properties of the species so such classifications have value. If the increase in genome copy number results from the combination of chromosome sets from divergent species, the different types of chromosomes will usually not pair with each other in prophase of meiosis I. In the case of tetraploids, if both divergent genomes are doubled by whatever means, those sets of chromosomes that are similar or identical will preferentially pair with each other to the exclusion of the other genome. This type of pairing is referred to as ‘‘disomic’’ in analogy with the situation in a diploid. A species with this type of scenario is referred to as an allopolyploid because the contributing genomes are different from each other. 26 Genetic ratios in an allotetraploid depend on the constitution of each genome (Clausen and Goodspeed 1925; Clausen 1941; Clausen and Cameron 1944). The different sets of related chromosomes are referred to as homoeologues. If both the homoeologues possess the homoeologous gene copies that are expressed similarly, then both would need to be mutant in order to express a recessive phenotype. Under these circumstances, duplicate gene ratios would typically be observed. In other words, recessive phenotypes would be found in 1/16 (1/4 9 1/4) of the F2 UN Editor Proof 22 Layout: T1 Standard SC Chapter No.: 2 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 23/32 23 70 2.3 Autopolyploids 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 PR OO 72 If on the other hand the increase in genome copy number in a polyploid results from the same species such that the chromosomes are all quite similar, the pairing in prophase of meiosis I forms conglomerates that switch pairing partners along the length of the chromosome (Fig. 2.1). This type of pairing is referred to as ‘‘quadrivalent’’ pairing because all four chromosomes present can be involved with each other. However, 3:1 and 2:2 associations are also observed. The segregation in this case will depend on the position of the locus in question in the chromosome and relative to the respective centromere (Blakeslee et al. 1923; Haldane 1930; Bartlett and Haldane 1934; Mather 1935, 1936; Randolph 1935; Little 1945, 1958; Doyle 1973). Those genes near the centromere will be distributed to the diploid gametes based on the usual case that pairs of centromeres will separate from each other at meiosis I and that the two sets from each chromosome of the complement will do so at random. A homozygous dominant autotetraploid (AAAA) is referred to as a quadruplex and the homozygous recessive (aaaa) as a nulliplex. There are three types of heterozygotes: AAAa (triplex), AAaa (duplex), and Aaaa (simplex). If one designates a hybrid autotetraploid as AA0 aa0 , then there are six types of possible gametes that will be formed: AA0 , Aa, Aa0 , aA, a0 A0 , aa0 . The frequency of diploid homozygous gametes under these circumstances is 1/6 (0.167). A self-pollination will produce 2.77 % of the progeny that are homozygous for the recessive allele (Fig. 2.2). However, as the distance of a gene from the centromere increases, recombination between the locus and the centromere will randomize the distribution of the different alleles into the diploid gametes to the point that the frequency of homozygous diploid gametes will be (4/8 9 3/7 = 0.21) as a maximum. In this case, a self-pollination will produce 4.41 % of the progeny that are homozygous. Recombination between the monitored locus and the centromere can also produce homozygous spores from a triplex heterozygote (AAAa) to produce aa gametes (Catcheside 1956). This process is called double reduction. Again, this result is affected by the position of the locus under consideration from the centromere with greater double reduction increasing with distance. Another factor affecting segregation in autotetraploids is aneuploidy, i.e., altered copy number of individual chromosomes. This circumstance would change the pairing and segregation properties of individual chromosomes. Autotetraploids can also generate spontaneous diploid progeny via parthenogenesis (Randolph and Fischer 1939). D 71 TE 67 68 EC 66 CO RR 65 F 69 from a self of an F1 between parental types that are dominant and recessive. However, if one of the gene copies is missing or expressed in other tissues from one of the homoeologous chromosomes, then genetic ratios typical of a diploid will be found because only one genome will have different alleles in an F1, and they will segregate to produce a 3:1 ratio because the single genome will behave as a diploid. 64 UN Editor Proof 2 Genetic Consequences of Polyploidy in Plants Layout: T1 Standard SC Chapter No.: 2 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 24/32 J. A. Birchler PR OO F Editor Proof 24 Aa Aa Aa TE AA D Fig. 2.1 Comparison of meiotic anaphase I in matched diploid and tetraploid maize plants. a Gray value image of anaphase I of diploid inbred line B73. Note that membrers of each pair of homologs separate from each other. Modified from (Birchler 2011). b Gray value image of anaphase I of a tetraploid derivative of inbred line B73. Note the multivalent associations of chromosomes. Photos by Zhi Gao and Fangpu Han Aa aa AA AA AA Aa AA Aa AA Aa AA Aa AA aa Aa AA Aa Aa Aa Aa Aa Aa Aa Aa Aa Aa aa Aa AA Aa Aa Aa Aa Aa Aa Aa Aa Aa Aa aa Aa AA Aa Aa Aa Aa Aa Aa Aa Aa Aa Aa aa Aa AA Aa Aa Aa Aa Aa Aa Aa Aa Aa Aa aa aa AA aa aa Aa aa Aa aa Aa aa Aa aa aa CO RR EC AA UN Fig. 2.2 Genotypes in a self-pollination of an autotetraploid heterozygote AAaa. In an autotetraploid, the gametes are diploid. With two chromosomes carrying the dominant A allele and two carrying the recessive a allele, the distribution of gametes from the heterozygote is shown across the top and along the side. The combinations of these gametes to produce the tetraploid progeny are shown in the grid. Only one out of 36 are homozygous for A, and one out of 36 are homozygous for the recessive a. Other combinations of A and a are shown. These conditions hold for genes closely linked to centromeres as described in the text Layout: T1 Standard SC Chapter No.: 2 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 25/32 105 25 2.4 Segmental Allopolyploids 118 2.5 Heterosis and Ploidy 113 114 115 116 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 PR OO 112 D 111 Because of the chromosome pairing considerations noted above, allopolyploids will have a diversity of gene products ‘‘fixed’’ in their genomic structure. This circumstance will basically maintain the essence of hybrid vigor even though technically an otherwise high degree of homozygosity might be present. Thus, allopolyploids typically exhibit robust biomass and excellent fertility compared to the diploid progenitor species. Nevertheless, crosses between different isolates of allopolyploids can show even greater heterotic effects when each genome is heterozygous as well (Gustafson 1946). Autopolyploids, which can have up to four different alleles at one locus, also exhibit hybrid vigor and with increasing diversity of alleles present, a phenomenon known as progressive heterosis, there is increasing biomass and fertility (Busbice and Wilsie 1966; Levings et al. 1967; Mok and Peloquin 1975; Groose et al. 1989). However, autotetraploids can be subject to inbreeding depression in which the potential exists for all copies of a chromosome to become homozygous (Busbice and Wilsie 1966; Sockness and Dudley 1989a, b). Polyploids that are entirely homozygous exhibit extreme depression and reduction of stature and fertility (Busbice and Wilsie 1966; Riddle et al. 2006; Abel and Becker 2007; Stupar et al. 2007; Redei 1964; d’Erfurth et al. 2009; Yao et al. 2011). This situation is unlikely under natural circumstances. A discussion of the vigor of polyploids needs to consider the intersection with heterosis or hybrid vigor. In recent years, it has been possible to produce ploidy series for completely or highly homozygous materials (Riddle et al. 2006; Abel and Becker 2007; Stupar et al. 2007; d’Erfurth et al. 2009; Yao et al. 2011). The general rule that emerges from these studies is that with increasing ploidy and the maintenance of homozygosity, there is usually a decline in stature and fertility. The cell and pollen size increases with ploidy and the plants typically take on a TE 109 110 EC 108 CO RR 107 F 117 A third classification based upon empirical chromosome associations is segmental allopolyploid (Stebbins 1947). In this case, some chromosomes exhibit bivalent pairing and others show quadrivalent pairing. The basis of such behavior was not clear until recently. Xiong and colleagues (Xiong et al. 2011) found that in resynthesized Brassica napus derived from the diploid progenitors, B. oleracea and B. rapa, different lineages could form compensating nullisomic-tetrasomic configurations for different chromosomes. In this case, the tetraploid will have some chromosomes that are basically identical and other members of the set that will be divergent. Similar results were reported for naturally occurring tetraploid Tragopogon miscellus (Chester et al. 2012). Such a species will be a composite of allo- and autotetraploid chromosomes for different members of the karyotype and would be expected to exhibit the pairing characterized by a segmental allopolyploid. 106 UN Editor Proof 2 Genetic Consequences of Polyploidy in Plants Layout: T1 Standard SC Chapter No.: 2 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 26/32 J. A. Birchler 153 2.6 Aneuploidy Relative to Ploidy 151 154 155 156 157 158 159 160 161 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 PR OO 150 In contrast to a ploidy series, changes in dosage of individual chromosomes (or substantial parts of chromosomes) have a more dramatic effect on the phenotype (Blakeslee et al. 1920; Blakeslee 1934). Typically, the removal of a chromosome or chromosomal segment has the strongest effects and is lethal in some cases (Kush and Rick 1968; Vizir and Mulligan 1999). All of the monosomics for each of the ten chromosomes in maize have been recovered and studied (Weber 1983), but this is not the case in other species in which this issue has been examined such as tomato (Kush and Rick 1968) and Arabidopsis (Vizir and Mulligan 1999). The addition of a chromosome to produce a trisomic usually also has a detrimental effect on plant vigor but the usual circumstance is that the impact is much less than monosomics (Lee et al. 1996). Indeed, full sets of trisomics have been produced for many plant species (Singh 1993). Tetrasomics for whole chromosome arms, otherwise called secondary trisomics, have been produced in Datura by recovery of extra chromosomes that are duplicated for one or the other arm of the progenitor chromosome (Blakeslee 1934). These secondary trisomics usually have more intensified phenotypic effects and are more intensified when present in haploids (Satina et al. 1937a, b). Extra or missing chromosomes in higher ploidies have less severe phenotypic effects. A comprehensive set of aneuploids was generated in hexaploid wheat (Sears 1944; Sears 1953, 1954). Monosomics and trisomics are regularly produced, and because of the high ploidy state, nullisomics, which are missing both copies of a chromosome, can be produced (Sears 1953, 1954). Nullisomics have a more severe effect than the corresponding monosomic. Tetrasomics can be produced and have a more severe effect than the respective trisomic. Compensating nullisomics for one homoeologue and tetrasomics for another return to a more normal phenotype than exhibited by the nullisomic or tetrasomic alone (Sears 1953, 1954). Newly synthesized B. napus (Xiong et al. 2011) and natural neopolyploids of T. miscellus (Chester et al. 2012) will exhibit aneuploidy that resolves into compensating 4:0 or 3:1 contributions from different progenitor genomes illustrating that the compensating balanced condition is favored in D 149 TE 148 EC 146 147 CO RR 145 F 152 ‘‘stocky’’ appearance but with extreme ploidies, the plants are depauperate (Blakeslee 1941; Randolph 1942; Rhoades and Dempsey 1966; d’Erfurth et al. 2009; Yao et al. 2011). In contrast, hybrids with increasing ploidy tend to exhibit greater biomass and in the species in which it has been examined closely, there is an increase in heterosis with increasing diversity of alleles, i.e., progressive heterosis (Busbice and Wilsie 1966; Mok and Peloquin 1975; Levings et al. 1967; Groose et al. 1989; Bingham et al. 1994; Riddle and Birchler 2008). The common view that polyploids exhibit more robust stature derives from experience with allopolyploids or with heterotic autopolyploids, which are the situations most commonly encountered. 144 UN Editor Proof 26 Layout: T1 Standard SC Chapter No.: 2 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 27/32 27 187 2.7 Gene Expression Studies PR OO 185 F 186 laboratory or natural selection. Together, these results further illustrate that the greater the deviation from the standard set of chromosomes, the more severe the impact on the phenotype. 184 205 2.8 Genomic Balance 192 193 194 195 196 197 198 199 200 201 202 203 206 207 208 209 210 211 212 213 214 215 216 217 218 219 TE 191 EC 190 CO RR 189 D 204 Studies on gene expression in ploidy and aneuploid series parallel the phenotypic results. When individual genes are sampled in a ploidy series, the expression level is more or less proportional to the ploidal level, although there are examples of genes whose expression deviates from this trend both positively and negatively (Birchler and Newton 1981; Guo et al. 1996). Genome-wide studies of gene expression in ploidy series demonstrate a similar pattern (Wang et al. 2004; Albertin et al. 2005; Stupar et al. 2007; Riddle et al. 2010; Yu et al. 2010). In contrast, sampling of individual genes or protein patterns in aneuploids reveals a greater set of changes from the diploid level of expression (Birchler 1979; Birchler and Newton 1981; Guo and Birchler 1994). A dosage series for a particular chromosomal region would alter the amount of expression of a portion of the total gene products encoded across the genome. The effects could be positive or negative correlations with the change in dosage. The more common effect especially with trisomics was a negative correlation between the dosage and the target gene expression (Birchler 1979; Birchler and Newton 1981; Guo and Birchler 1994). Thus, the gene expression patterns show changes in a ploidy series but aneuploid series exhibit greater effects in parallel with the phenotypic relationships. 188 This gene expression relationship led to the suggestion that the stoichiometry of regulatory genes affected the outcome of gene expression (Birchler and Newton 1981) and ultimately the phenotype (Guo and Birchler 1994). Studies to identify single genes that would mimic the aneuploid effects using a partial loss of function mutation in the white eye color gene in Drosophila produced single-gene mutations that would modulate the target’s expression either positively or negatively (Rabinow et al. 1991; Birchler et al. 2001). The molecular identification of many of these genes revealed them to be transcription factors, chromatin modifiers, and components of signal transduction (Birchler et al. 2001). Interestingly, these same classes of genes are typical of those that exhibit preferential retention following a polyploidization event (Blanc and Wolfe, 2004; Freeling and Thomas, 2006) and underrepresentation in segmental duplications (Maere et al. 2005; Freeling et al. 2008). Thus, this evidence suggests that if these classes of genes are out of register with each other, there is a negative fitness UN Editor Proof 2 Genetic Consequences of Polyploidy in Plants Layout: T1 Standard SC Chapter No.: 2 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 28/32 J. A. Birchler 226 2.9 Triploids 222 223 224 PR OO 221 F 225 consequence. Thus, the phenotypic, gene expression and evolutionary studies form a coherent picture that these types of genes form a balance. When individual components exhibit a dosage effect, this will ultimately produce a fitness consequence due to the impact of the altered gene expression on the phenotype (Birchler et al. 2001; Veitia 2002; Veitia 2004; Birchler et al. 2005, 2007; Veitia et al. 2008; Birchler and Veitia 2007, 2010). 220 240 2.10 Higher Ploidal Levels 231 232 233 234 235 236 237 238 241 242 243 244 245 246 247 248 249 250 251 252 253 254 TE 230 EC 229 Ploidal levels above the tetraploid level most commonly involve hexaploids and octoploids although much higher levels have been documented. Chromosome pairing in allohexaploids has been studied in detail, for example in wheat, which is ordinarily disomic in nature (Kihara 1919; Lilienfeld 1951; Dvorak et al. 1988). The Ph system insures pairing of homologs and against pairing of homoeologues but homoeologues can pair in mutant plants (Yousafzai et al. 2010; see Chap. 7, this volume). The wheat genome is composed of three different slightly diverged genomes tracing back through the joining of an allotetraploid composed of two genomes with the third. The Ph system maintains disomic pairing and hence excellent fertility. Octoploids, using sugar cane as an example, have variable chromosome numbers due to the minimal detrimental effects of aneuploidy at this level (Piperidis et al. 2010). In contrast, triticale, which is an octoploid consisting of hexaploid wheat with the addition of a rye genome, exhibits faithful chromosomes numbers. CO RR 228 D 239 Triploids are a polyploid level between diploid and tetraploid. They arise from crosses between diploid and tetraploids of the same or related species or from unreduced gametes from one diploid parent. In meiosis, the chromosomes associate in trivalents, which consists of pairing of any two chromosomes at any one point (McClintock 1929; Punyasingh 1947; Upcott 1935). The distribution of chromosomes is nearly random, resulting in spores that range from 1x to 2x. As a consequence, the gametophytes are mostly highly aneuploid and in some cases abort (Satina and Blakeslee 1937a, b). Fertilization involving gametes of different chromosome numbers in the endosperm will often cause endosperm abortion (Satina et al. 1938; Punyasingh 1947; Brink and Cooper 1947; Cooper 1951). The gametes that are successful tend to be those at or near the 1x or 2x level. Because of the variability of the chromosome numbers in gametes from triploid individuals, this ploidal level is not stable. 227 UN Editor Proof 28 Layout: T1 Standard SC Chapter No.: 2 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 29/32 255 29 2.11 Concluding Remarks 270 Acknowledgments Research supported by National Science Foundation grant DBI 0733857. 271 References 272 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 293 294 295 296 297 Abel S, Becker HC (2007) The effect of autopolyploidy on biomass production in homozygous lines of Brassica rapa and Brassica oleracea. Plant Breed 126:642–643 Albertin W, Brabant P, Catrice O, Eber F, Jenczewski E, Chèvre AM, Thiellement H (2005) Autopolyploidy in cabbage (Brassica oleracea L.) does not alter significantly the proteomes of green tissues. Proteomics 5:2131–2139 Barker MS, Kane NC, Matvienko M, Kozik A, Michelmore RW, Knapp SJ, Rieseberg LH (2008) Multiple paleopolyploidizations during the evolution of the compositae reveal parallel patterns of duplicate gene retention after millions of years. Mol Biol Evol 25:2445–2455 Bartlett MS, Haldane JBS (1934) The theory of inbreeding in Autotetraploids. J Genet 29:175–180 Bingham ET, Groose RW, Woodfield DR, Kidwell KK (1994) Complementary gene interactions in alfalfa are greater in autotetraploids than diploids. Crop Sci 34:823–829 Birchler JA (1979) A study of enzyme activities in a dosage series of the long arm of chromosome one in maize. Genetics 92:1211–1229 Birchler JA (2011) Epigenetic aspects of centromere function in plants. Curr Opin Plant Biol 14:217–222 Birchler JA, Newton KJ (1981) Modulation of protein levels in chromosomal dosage series of maize: the biochemical basis of aneuploid syndromes. Genetics 99:247–266 Birchler JA, Bhadra U, Pal Bhadra M, Auger DL (2001) Dosage dependent gene regulation in multicellular eukaryotes: implications for dosage compensation, aneuploid syndromes and quantitative traits. Dev Biol 234:275–288 Birchler JA, Riddle NC, Auger DL, Veitia RA (2005) Dosage balance in gene regulation: biological implications. Trends Genet 21:219–226 Birchler JA, Veitia RA (2007) The gene balance hypothesis: from classical genetics to modern genomics. Plant Cell 19:395–402 Birchler JA, Veitia RA (2010) The gene balance hypothesis: implications for gene regulation, quantitative traits and evolution. New Phytol 186:54–62 262 263 264 265 266 267 268 PR OO 261 D 260 TE 259 EC 258 CO RR 257 F 269 The genetics of polyploids depends essentially on the pairing properties of the multiple chromosomes in meiosis. If the multiple copies of a genome are sufficiently dissimilar from each other, they tend to pair among themselves and maintain the genetic variation within each genome. If the multiple copies of a genome are similar to each other, then all copies are free to pair and recombine among themselves. In this circumstance, the genetic behavior of a particular gene is dependent on its position on the chromosome and the fidelity of the pairing of homologs. Aneuploidy, i.e., the variation of a single chromosome or chromosomal segment, can have more severe consequences than varying the whole genome. However, as the background ploidy increases, the effect of the same chromosome change of aneuploidy becomes less. The phenotypic effects, gene expression patterns and the evolutionary results of differential gene retention following whole-genome duplications versus segmental duplication suggest the importance of genomic balance. 256 UN Editor Proof 2 Genetic Consequences of Polyploidy in Plants Layout: T1 Standard SC Chapter No.: 2 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 30/32 EC TE D PR OO F Birchler JA, Yao H, Chudalayandi S (2007) Biological consequences of dosage dependent gene regulatory systems. Biochem Biophys Acta-Gene Struct Expr 1769:422–428 Blakeslee AF (1934) New Jimson weeds from old chromosomes. J Hered 24:80–108 Blakeslee AF (1941) Effect of induced polyploidy in plants. Am Nat 75:117–135 Blakeslee AF, Belling J, Farnham ME (1920) Chromosomal duplication and Mendelian phenomena in Datura mutants. Science 52:388–390 Blakeslee AF, Belling J, Farnham ME (1923) Inheritance in tetraploid Daturas. Bot Gaz 76:329–373 Blanc G, Wolfe KH (2004) Functional divergence of duplicated genes formed by polyploidy during Arabidopsis evolution. Plant Cell 16:1679–1691 Blomme T, Vandepoele K, De Bodt S, Simillion C, Maere S, Van de Peer Y (2006) The gain and loss of genes during 600 million years of vertebrate evolution. Genome Biol 7:43 Bowers JE, Chapman BA, Rong J, Paterson AH (2003) Unraveling angiosperm genome evolution by phylogenetic analysis of chromosomal duplication events. Nature 422:433–438 Busbice TH, Wilsie CP (1966) Inbreeding depression and heterosis in autotetraploids with application to Medicago sativa L. Euphytica 15:52–67 Brink RA, Cooper DC (1947) The endosperm in seed development. Bot Rev 13:423–541 Catcheside DG (1956) Double reduction and numerical non-disjunction in tetraploid maize. Heredity 10:205–218 Chapman BA, Bowers JE, Feltus FA, Paterson AH (2006) Buffering of crucial functions by paleologous duplicated genes may contribute cyclicality to angiosperm genome duplication. Proc Natl Acad Sci USA 103:2730–2735 Chester M, Gallagher JP, Symonds VV, Cruz da Silva AV, Mavrodiev EV, Leitch AR, Soltis PS, Soltis DE (2012) Extensive chromosomal variation in a recently formed natural allopolyploid, Tragopogon miscellus (Asterasceae). Proc Natl Acad Sci USA (in press) Clausen RE (1941) Polyploidy in Nicotiana. Am Nat 75:291–306 Clausen RE, Cameron DR (1944) Inheritance in Nicotiana tabacum. XVIII. Monosomic analysis. Genetics 29:447–477 Clausen RE, Goodspeed TH (1925) Interspecific hybridization in Nicotiana II. A tetraploid glutinosa-tabacum hybrid. An experimental verification of Winge’s hypothesis. Genetics 10:278–284 Cooper DC (1951) Caryopsis development following matings between diploid and tetraploid strains of Zea mays. Am J Bot 38:702–708 Doyle GG (1973) Autotetraploid segregation. Theor Appl Genet 43:139–146 d’Erfurth I, Jolivet S, Froger N, Catrice O, Novatchkova M, Mercier R (2009) Turning meiosis into mitosis. PLoS Biol 7:1000124 Dvorak J, McGuire PE, Cassidy B (1988) Apparent sources of the a genomes of wheats inferred from the polymorphism in abundance and restriction fragment length of repeated nucleotide sequences. Genome 30:680–689 Freeling M, Lyons E, Pedersen B, Alam M, Ming R, Lisch D (2008) Many or most genes in Arabidopsis transposed after the origin of the order Brassicales. Genome Res 18:1924–1937 Freeling M, Thomas BC (2006) Gene-balanced duplications, like tetraploidy, provide predictable drive to increase morphological complexity. Genome Res 16:805–814 Guo M, Birchler JA (1994) Trans-acting dosage effects on the expression of model gene systems in maize aneuploids. Science 266:1999–2002 Guo M, Davis D, Birchler JA (1996) Dosage effects on gene expression in a maize ploidy series. Genetics 142:1349–1355 Groose RW, Talbert LE, Kojis WP, Bingham ET (1989) Progressive heterosis in autotetraploid alfalfa: studies using two types of inbreds. Crop Sci 29:1173–1177 Gustafson A (1946) The effect of heterozygosity on viability and vigor. Hereditas 32:263–286 Haldane JBS (1930) The theoretical genetics of autopolyploids. J Genet 22:359–372 Kihara H (1919) Uber zytologische Studien bei einigen Getreidearten. Bot Mag Tokyo 33 Kush GS, Rick CM (1968) Cytogenetic analysis of the tomato genome by means of induced deficiencies. Chromosoma 23:452–484 CO RR 298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 J. A. Birchler UN Editor Proof 30 Layout: T1 Standard SC Chapter No.: 2 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 31/32 31 EC TE D PR OO F Lee EA, Darrah LL, Coe EH (1996) Dosage effects on morphological and quantitative traits in maize aneuploids. Genome 39:898–908 Levings CS, Dudley JW, Alexander DE (1967) Inbreeding and crossing in autotetraploid maize. Crop Sci 7:72–73 Lilienfeld FA (1951) H. Kihara: Genome analysis in Triticum and Aegilops. Concluding review. Cytologia 16:101–123 Little TM (1945) Gene segregation in autotetraploids. Bot Rev 11:60–85 Little TM (1958) Gene segregation in autotetraploids. II. Bot Rev 24:319–339 Maere S, DeBodt S, Raes J, Casneuf T, Van Montagu M, Kuiper M, Van de Peer Y (2005) Modeling gene and genome duplications in eukaryotes. Proc Natl Acad Sci USA 102:5454–5459 Mather K (1935) Reductional and equational separation of the chromosomes in bivalents and tetravalents. J Genet 30:53–78 Mather K (1936) Segregation and linkage in autotetraploids. J Genet 32:287–314 McClintock B (1929) A cytological and genetical study of triploid maize. Genetics 14:180–227 Mok DWS, Peloquin SJ (1975) Breeding value of 2n pollen (diploandroids) in tetraploid x diploid crosses in potato. Theor Appl Genet 46:307–314 Piperidis G, Piperidis N, D’Hont A (2010) Molecular cytogenetic investigation of chromosome composition and transmission in sugarcane. Mol Genet Genomics 284:65–73 Punyasingh K (1947) Chromosome numbers in crosses of diploid, triploid and tetraploid maize. Genetics 32:541–554 Rabinow L, Nguyen-Huynh AT, Birchler JA (1991) A trans-acting regulatory gene that inversely affects the expression of the white, brown and scarlet loci in Drosophila melanogaster. Genetics 129:463–480 Randolph L (1935) Cytogenetics of tetraploid maize. J Agri Res 50:591–605 Randolph LF (1942) The influence of heterozygosis on fertility and vigor in autotetraploid maize. Genetics 27:163 Randolph LF, Fischer HE (1939) The occurrence of parthenogenetic diploids in tetraploid maize. Proc Natl Acad Sci USA 25:161–164 Redei G (1964) Crossing experiences with polyploids. Arabidopsis Inf Serv 1:13 Rhoades MM, Dempsey E (1966) Induction of chromosome doubling at meiosis by the elongate gene in maize. Genetics 54:505–522 Riddle NC, Birchler JA (2008) Comparative analysis of inbred and hybrid maize at the diploid and tetraploid levels. Theor Appl Genet 116:563–576 Riddle NC, Kato A, Birchler JA (2006) Genetic variation for the response to ploidy change in Zea mays L. Theor Appl Genet 114:101–111 Riddle NC, Jiang H, An L, Doerge RW, Birchler JA (2010) Gene expression analysis at the intersection of ploidy and hybridity in maize. Theor Appl Genet 120:341–353 Satina S, Blakeslee AF (1937a) Chromosome behavior in triploid Datura stramonium I. The male gametophyte. Am J Bot 24:518–527 Satina S, Blakeslee AF (1937b) Chromosome behavior in triploid Datura stramonium II. The female gametophyte. Am J Bot 24:621–627 Satina S, Blakeslee AF, Avery AG (1937) Balanced and unbalanced haploids in Datura. J Hered 28:192–202 Satina S, Blakeslee AF, Avery AG (1938) Chromosome behavior in triploid Datura. III. The seed. Am J Bot 25:595–602 Sears ER (1944) Cytogenetic studies with polyploid species of wheat. II. Additional chromosome aberrations in Triticum vulgare. Genetics 29:232–246 Sears ER (1953) Nullisomic analysis in common wheat. Am Nat 87:245–252 Sears ER (1954) The aneuploids of common wheat. Univ Mo Res Bull 572:1–58 Simillion C, Vandepoele K, Montagu MC, Zabeau M, Van de Peer Y (2002) The hidden duplication past of Arabidopsis thaliana. Proc Natl Acad Sci USA 99:13627–13632 Singh RJ (1993) Plant cytogenetics. CRC Press Inc, Boca Raton Sockness BA, Dudley JW (1989a) Performance of single and double cross autotetraploid maize hybrids with different levels of inbreeding. Crop Sci 29:875–879 CO RR 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 UN Editor Proof 2 Genetic Consequences of Polyploidy in Plants Layout: T1 Standard SC Chapter No.: 2 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 32/32 EC TE D PR OO F Sockness BA, Dudley JW (1989b) Morphology and yield of isogenic diploid and tetraploid maize inbreds and hybrids. Crop Sci 29:1029–1032 Stebbins GL Jr (1947) Types of polyploids: their classification and significance. Adv Genet 1:403–429 Stupar RM, Bhaskar PB, Yandell BS, Rensink WA, Hart AL, Ouyang S, Veilleux RE, Busse JS, Erhardt RJ, Cr Buell, Jiang J (2007) Phenotypic and transcriptomic changes associated with potato autopolyploidazation. Genetics 176:2055–2067 Upcott M (1935) The cytology of triploid and tetraploid Lycopersicum esculentum. J Genet 27:105–132 Veitia RA (2002) Exploring the etiology of haploinsufficiency. BioEssays 24:175–184 Veitia RA (2004) Gene dosage balance in cellular pathways: implications for dominance and gene duplicability. Genetics 168:569–574 Veitia RA, Bottani S, Birchler JA (2008) Cellular reactions to gene dosage imbalance: genomic, transcriptomic and proteomic effects. Trends Genet 24:390–397 Vizir IY, Mulligan BJ (1999) Genetics of gamma-irradiation-induced mutations in Arabidopsis thaliana: large chromosomal deletions can be rescued through fertilization of diploid eggs. J Hered 90:412–417 Wang J, Tina L, Madlung A, Lee HS, Chen M, Lee JJ, Watson B, Kagochi T, Comai L, Chen ZJ (2004) Stochastic and epigenetic changes of gene expression in Arabidopsis polyploids. Genetics 167:1961–1973 Weber DF (1983) Monosomic analysis in diploid crop plants. In: Swaminathan MS, Gupta PK, Sinha U (eds) Cytogenetics of crop plants. Macmillan India Limited, New Delhi, pp 351–378 Wolfe KH, Shields DC (1997) Molecular evidence for an ancient duplication of the entire yeast genome. Nature 387:708–713 Xiong Z, Gaeta RT, Pires JC (2011) Homoeologous shuffling and chromosome compensation maintain genome balance in resynthesized allopolyploid Brassica napus. Proc Natl Acad Sci USA 108:7908–7913 Yao H, Kato A, Mooney B, Birchler JA (2011) Phenotypic and gene expression analysis of a ploidy series of maize inbred Oh43. Plant Mol Biol 75:237–251 Yousafzai FK, Al-Kaff N, Moore G (2010) The molecular features of chromosome pairing at meiosis: the polyploidy challenge using wheat as a reference. Funct Integr Genomics 10:147–156 Yu Z, Haberer G, Matthes M, Rattei T, Mayer KF, Gierl A, Torres-Ruiz RA (2010) Impact of natural genetic variation on the transcriptome of autotetraploid Arabidopsis thaliana. Proc Natl Acad Sci USA 107:17809–17814 CO RR 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 J. A. Birchler UN Editor Proof 32 Metadata of the chapter that will be visualized in SpringerLink Book Title Polyploidy and Genome Evolution Series Title Chapter Title Meiosis in Polyploid Plants Copyright Year 2012 Copyright HolderName Springer-Verlag Berlin Heidelberg Corresponding Author Family Name Mittelsten Scheid Particle Given Name Ortrun Suffix Division Author Organization Gregor Mendel Institute of Molecular Plant Biology Address Vienna, Austria Email ortrun.mittelsten_scheid@gmi.oeaw.ac.at Family Name Zielinski Particle Given Name Marie-Luise Suffix Division Organization Gregor Mendel Institute of Molecular Plant Biology Address Vienna, Austria Email Abstract Meiosis is an obligate process during sexual reproduction, which involves the combination of parental genomes and the coordinated segregation of the recombined chromosomes to the gametes. Polyploidy has direct and fundamental consequences on meiosis, which are gradually and individually different between the extreme cases of auto- and allopolyploids. Multiple chromosome complements have a major impact, especially on chromosome pairing during pachytene and on the segregation of genotypes and phenotypes in progeny. At the same time, irregularities during meiosis are a major source of naturally occurring polyploidization events by the formation of unreduced gametes. Although individuals originating from nonhaploid gametes may suffer from reduced vigor and fecundity, their gametogenesis can produce many more chromosomal combinations than regular diploids, and thereby expose more diversity to natural selection. A more relaxed control of pairing and segregation in polyploids, possibly also with increased recombination rates, might be an important contribution to evolution and adaptation potential, especially under drastic or frequent changes in environmental conditions. Book ISBN: 978-3-642-31441-4 Page: 33/54 Chapter 3 2 Meiosis in Polyploid Plants 3 Marie-Luise Zielinski and Ortrun Mittelsten Scheid 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 D 8 TE 7 EC 6 Abstract Meiosis is an obligate process during sexual reproduction, which involves the combination of parental genomes and the coordinated segregation of the recombined chromosomes to the gametes. Polyploidy has direct and fundamental consequences on meiosis, which are gradually and individually different between the extreme cases of auto- and allopolyploids. Multiple chromosome complements have a major impact, especially on chromosome pairing during pachytene and on the segregation of genotypes and phenotypes in progeny. At the same time, irregularities during meiosis are a major source of naturally occurring polyploidization events by the formation of unreduced gametes. Although individuals originating from nonhaploid gametes may suffer from reduced vigor and fecundity, their gametogenesis can produce many more chromosomal combinations than regular diploids, and thereby expose more diversity to natural selection. A more relaxed control of pairing and segregation in polyploids, possibly also with increased recombination rates, might be an important contribution to evolution and adaptation potential, especially under drastic or frequent changes in environmental conditions. CO RR 4 5 PR OO 1 F Book ID: 272454_1_En Date: 16-8-2012 3.1 Introduction Polyploidization can affect the size of the nucleus and the cell, stimulate genetic rearrangements, or modify gene expression patterns, but these changes very likely operate by the same mechanistic principles of nuclear organization, recombination, or gene regulation, as in diploid cells. In contrast, reducing the number of UN Editor Proof Layout: T1 Standard SC Chapter No.: 3 M.-L. Zielinski  O. Mittelsten Scheid (&) Gregor Mendel Institute of Molecular Plant Biology, Vienna, Austria e-mail: ortrun.mittelsten_scheid@gmi.oeaw.ac.at P. S. Soltis and D. E. Soltis (eds.), Polyploidy and Genome Evolution, DOI: 10.1007/978-3-642-31442-1_3,  Springer-Verlag Berlin Heidelberg 2012 33 Layout: T1 Standard SC Chapter No.: 3 Book ID: 272454_1_En Date: 16-8-2012 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 F 30 PR OO 29 D 28 TE 27 chromosomes during the formation of gametes requires a more elementary modification in polyploids. At the same time, irregular gamete formation is a major trigger for polyploidization. Both processes converge in meiosis. This is a sequence of usually two special, subsequent cell divisions in all sexually reproducing eukaryotic organisms in which the number of chromosomes is reduced to half, before two gametes of different parental origin fuse to form a zygote to produce the next generation. Like mitosis, the cell division of somatic cells, meiosis is also preceded by the replication of the nuclear DNA into duplicate sister chromatids which are attached at the centromeres. While in mitosis the replicated chromosomes are arranged individually on the metaphase plate, and chromatids become separated in the subsequent anaphase, each replicated chromosome in meiotic cells is first aligned with its respective homologous partner. This pairing is followed by programed induction of DNA double-strand breaks (DSB), from which at least one per chromosome becomes converted into a recombination event between non-sister chromatids (crossover, CO). Later, these COs become cytologically visible as chiasmata. Recombination within chromosome pairs results in new and unique combinations of paternal and maternal genetic information per chromosome, while it is also prerequisite for subsequent organized chromosome segregation. Paired chromosomes become physically connected by the synaptonemal complex (SC). The process of pairing, synapsis, and recombination takes most of the time during prophase I, which is subdivided into leptotene, zygotene, pachytene, diplotene, and diakinesis according to chromosome condensation and configuration. In the subsequent metaphase I, the recombined meiotic chromosome pairs are arranged on the equatorial plane, sister kinetochores of each pair attach to opposite spindles and are pulled apart in anaphase I. Correct and synchronous pairing is a prerequisite for equal distribution of the chromosomes to the two daughter cells. After telophase I and prophase II, sister chromatids become separated during meta-, ana-, and telo-phase II, resulting in a classical meiosis with four nuclei. Faultless meiosis is an essential factor for fertility and thereby decisive for evolutionary success. Therefore, it is not surprising that it is a tightly controlled process and intensely studied in many organisms. The majority of meiosis research is performed with diploid organisms, and there are excellent reviews available that describe morphological, genetic, and mechanistic aspects of meiosis in many different systems (Dawe 1998; Bhatt et al. 2001; Armstrong et al. 2003; Mezard et al. 2007; Mercier and Grelon 2008; Harrison et al. 2010; Pawlowski 2010). However, characteristic differences in all stages of meiosis in many different polyploids have been observed repeatedly (reviewed in Ramsey and Schemske 2002). Therefore, we will focus on those aspects that are known or expected to be different between meiosis in diploids and polyploids. Although mechanistically intertwined, we will separate three aspects, (1) chromosome pairing, (2) recombination and crossover, and (3) segregation. Whenever relevant, we will distinguish between autopolyploids, with multiples of similar chromosomes, and allopolyploids, with chromosomes of different origin, and therefore higher divergence. EC 26 CO RR 25 M.-L. Zielinski and O. Mittelsten Scheid UN Editor Proof 34 Book ISBN: 978-3-642-31441-4 Page: 34/54 Layout: T1 Standard SC Chapter No.: 3 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 35/54 35 76 3.2 Recognition, Pairing, and Synapsis 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 PR OO 74 Early in meiosis, homologous chromosomes have to find and recognize each other. Interphase chromosomes are thought to occupy nonrandom territories (Schubert and Shaw 2011), and nonrandom spatial organization of homologous chromosomes has been recognized (reviewed in Avivi and Feldman 1980). In many eukaryotes, telomeres and centromeres can cluster at opposite poles of the nucleus (Rabl 1885; Cowan et al. 2001). This Rabl-configuration (Fig. 3.1a) is found in wheat, rye, oats, and barley, but not in maize or Arabidopsis (reviewed in Schubert and Shaw 2011). It is assumed to be related to large genome size and/or chromosome length (Dong and Jiang 1998). Rabl-configuration may support gene expression control and the onset of meiosis (Cowan et al. 2001). After the premeiotic DNA replication, the two sister chromatids are held together by sister chromatid cohesion and gradually become condensed. During leptotene, their chromatin is folded into loops attached to a protein fiber core, the axial element (AE) (Harper et al. 2004). During transition to zygotene, the so-called telomere bouquet is formed, in which the chromosome ends attach to the inner nuclear envelope and cluster (Fig. 3.1b). This association is likely to be actively regulated (Scherthan 2007). Recognition between the homologous chromosomes appears to occur between subtelomeric regions (Corredor et al. 2007). The telomere bouquet, found in almost all studied organisms except Drosophila and C. elegans (Harper et al. 2004), is not absolutely required for the following steps but is proposed to make progression of meiosis more efficient. Both Rabl-configuration and telomere bouquet bundle chromosomes and thereby help reduce the spatial distance between them. After telomere bouquet formation in yeast, the centromeric regions are distributed throughout the nucleus and oscillate to enhance finding homologous sequences (Bass et al. 1997; Scherthan 2007). Chromosome movements are extremely dynamic at this stage (Sheehan and Pawlowski 2009), which increases physical encounters and makes meeting and recognition of the homologs more likely (Pawlowski and Cande 2005). D 73 TE 72 EC 71 CO RR 70 F 75 Since polyploidy in animal germ lines is restricted to relatively few groups, and the research literature about meiotic mechanisms is scarce, most of the evidence summarized here stems from investigations in fungi and plants. There, polyploidy is by far most common among angiosperms, but genome research provides growing evidence for ancient polyploidization events in other groups (Sundstrom et al. 2008; Jiao et al. 2011). Independent of how and when polyploids originate, mastering meiosis is an important checkpoint for survival and evolutionary success. 69 UN Editor Proof 3 Meiosis in Polyploid Plants 3.2.1 Homology Recognition in Polyploids It seems plausible to expect that higher chromosome numbers, and larger nuclei in general, but especially more potential homologous partner chromosomes in Layout: T1 Standard SC Chapter No.: 3 Book ID: 272454_1_En Date: 16-8-2012 Editor Proof 36 Book ISBN: 978-3-642-31441-4 Page: 36/54 M.-L. Zielinski and O. Mittelsten Scheid (b) PR OO F (a) Fig. 3.1 Chromosome arrangement (red circles centromeres; orange triangles telomeres). a Rabl-configuration in interphase, with centromeres clustered at one pole of the nucleus and telomeres oriented toward the other. b Bouquet formation during meiosis at the onset of zygotene, with telomeres clustered at the inner nuclear envelope 125 3.2.2 Pairing in Autopolyploids 113 114 115 116 117 118 119 120 121 122 123 126 127 128 129 130 131 TE 112 EC 111 CO RR 109 110 There is one important difference that distinguishes the pachytene stage in autopolyploids: the possibility of pairing between more than two chromosomes, resulting in the formation of multivalents. This requires a certain degree of homology between pairing partners, and this is reflected in different frequency of multivalent occurrence depending on the type of polyploidy, species, individual chromosomes, and chromosome segments (Sybenga 1996). While allopolyploids UN 108 D 124 autopolyploids, could delay progression through meiosis, since the early stages of recognition and pairing can be more complex. Indeed, meiosis in autotetraploid Saccharomyces cerevisiae strains is delayed in comparison to its diploid counterpart (Trelles-Sticken et al. 2003). While higher C-values (genomic DNA content) in different plant species also increase the duration of meiosis when compared at the same ploidal level, the comparison of plants with different ploidy within the same species revealed a surprising, negative correlation: higher polyploidy seems to shorten the time needed for completing meiosis (Bennett 1977). These discrepancies might be created by defining the duration only by cytologically visible events. Analysis with higher resolution techniques and consideration of molecular events (Carlton et al. 2006) at the onset of meiosis might provide more accurate information on meiosis kinetics in polyploids. In spite of individual differences, it is remarkable that a wide range of genome and nuclear size variation does not modify the length of the recognition process (Moore and Shaw 2009). It appears that homology recognition is not the limiting step, due to supportive elements like Rablconfiguration and telomere bouquet , specialized chromosome pairing sites as found in D. melanogaster and C. elegans (McKee 1996; McKim 2007), and intensive chromosome movements that are effective in diploids and polyploids. 107 Layout: T1 Standard SC Chapter No.: 3 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 37/54 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 159 160 161 162 163 164 165 166 167 168 169 170 171 172 F 138 139 PR OO 136 137 D 135 TE 134 with genomes composed of genetically divergent parents often form only bivalents aligned over the whole length of the chromosome, autopolyploids may have multivalents. These are generated by simultaneous alignment of different partners and at different chromosome ends. The probability depends on the degree of ploidy and homology (random-end pairing model by John and Henderson 1962) and is controlled by genetic factors (see below), but it is independent of chromosome length (Morrison and Rajhathy 1960). Although multivalent formation is more often associated with autopolyploidy, many exceptions delimit the universality of this correlation: newly formed autopolyploids often have a lower rate of multivalents than expected, and allopolyploids may form multivalents to some extent (Ramsey and Schemske 2002). A few of these cases might be due to tri- or tetrasomy compensated by lack of other chromosomes, resulting in apparent euploidy (Mestiri et al. 2010). Progressive pairing starting simultaneously from opposite ends can result in multivalents with pairing partner switches (PPS), so that one chromosome can be aligned with two or more others in different segments (Fig. 3.2a). The distribution of PPS is variable but probably not totally random. More than one switch per chromosome (Fig. 3.2b) indicates the existence of additional, autonomous pairing sites (APS) along the chromosomes, each with a uniform and low probability of generating a PPS (Jones and Vincent 1994). The number of switches indicates the minimal number of pairing sites (Loidl 1995). These are likely not only determined genetically, since a higher number of PPS in autotriploid Crepis capillaris, compared to autotetraploid (Jones 1994), indicates other factors than just APS distribution, but probably some interference between pairing initiation sites. Pairing with one partner at one pairing site preferentially promotes continuation of pairing in a zipper-like manner, due to steric constraints. However, this preference is valid only once synapsis has been initiated. According to the random-end pairing model (John and Henderson 1962), alignment of two chromosome ends enhances the two remaining to pair, although the opposite ends of the chromosome still can pair randomly, resulting in 2/3 multivalents and 1/3 bivalents in tetraploids (only chromosome ends are considered under this model). Exceptions with fewer than expected multivalents (Weiss and Maluszynska 2000; Santos et al. 2003; Carvalho et al. 2010) or bivalent formation due to selective pairing (Simioni and do Valle 2011, and references within) reflect genetic and/or epigenetic heterozygosity between parental chromosome sets, due to ongoing diploidization. This can be different between individual plants or individual chromosomes (Santos et al. 2003) and extremely divergent in complex polyploids, as deduced from genetic maps of sugarcane (Jannoo et al. 2004). Higher than expected ratios of multivalents can also occur, as in newly generated autotetraploids of Arabidopsis thaliana, indicating multiple active pairing initiation sites even in small chromosomes (Santos et al. 2003). EC 133 37 CO RR 132 UN Editor Proof 3 Meiosis in Polyploid Plants Layout: T1 Standard SC Chapter No.: 3 Book ID: 272454_1_En Date: 16-8-2012 Editor Proof 38 Book ISBN: 978-3-642-31441-4 Page: 38/54 M.-L. Zielinski and O. Mittelsten Scheid (a) TE D PR OO F (b) 175 176 177 178 179 180 181 182 183 184 185 186 187 188 CO RR 174 3.2.3 Pairing in Allopolyploids Although newly formed hybrids combining different genomes and different chromosome numbers and shapes can exhibit erroneous pairing or lack of alignment (Ozkan and Feldman 2009), established allopolyploids show usually diploid-like pairing and formation of bivalents. Nevertheless, multivalents or segmental multivalent pairing can be found in allopolyploids, depending on external or genetic factors. In many polyploids, a genetic control system has evolved that regulates the pairing behavior during meiosis (Watanabe 1981; Gupta and Fedak 1985; MartinezPerez et al. 2001; Comai et al. 2003; Martinez-Perez et al. 2003; Jenczewski and Alix 2004). Probably the best studied example was discovered in the hexaploid bread wheat (Triticum aestivum, composed of genomes A, B, and D). In the presence of the Ph1 locus, pairing and recombination occur preferentially between homologous chromosomes (Riley and Chapman 1958); lack of Ph1 leads to a substantial increase of homoeologous pairing. The Ph1 locus from wheat can also affect pairing after introgression into rye (Lukaszewski and Kopecky 2010). The absence of Ph1 in diploids suggests its emergence through polyploidization (Griffiths et al. 2006). The UN 173 EC Fig. 3.2 Pairing partner switches during multivalent formation in polyploids. Alignment and partial synapsis between four chromosomes with one (a) or two (b) pairing partner switches (PPS) on the long arm Layout: T1 Standard SC Chapter No.: 3 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 39/54 39 197 3.2.4 Synapsis 198 199 200 201 202 203 204 205 206 207 208 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 PR OO 195 As soon as meiotic chromosomes align, they become joined by a stable proteinaceous structure, the SC. Its formation starts mainly at telomeres (Stack and Anderson 2002) and is intimately connected with DSBs and recombination (see next paragraph). The SC consists of three components: the two parallel AEs, the connecting central element, and periodically occurring recombination nodules (RNs) (Lohmiller et al. 2008). The RNs are multiprotein complexes thought to be involved in synapsis and recombination and can be distinguished by their structure and appearance as early or late RNs (Stack and Anderson 2002; Anderson and Stack 2005). Synapsis is expected to follow once pairing is established (Loidl 1995). However, unequal chromosome numbers can make a difference in polyploids: in autotriploid, but not in autotetraploid, yeast, trivalent SC formation was observed (Loidl 1995). Such fixed multivalents are prone to nondisjunction at metaphase I. Multivalents with equal numbers seem to resolve more often into regular bivalents, although SC connections between several AEs can occur and become fixed by crossovers. Without crossover at the region of the PPS, the central region is twisted and bivalents are formed (von Wettstein et al. 1984). When, and how, are interesting questions, especially for meiosis in autopolyploids. In summary, recognition, pairing, and synapsis of homologous chromosomes in diploids and polyploids are different in several aspects and determined by various parameters. The high rate of multivalent formation in autopolyploids depends on the type and history of the polyploidization, the number of pairing initiation sites, and odd or even chromosome multiplication. Allopolyploids rather generally form bivalents, but segment-, chromosome-, or genome-specific multivalent formation, and genetic and environmental influence, can also result in meiotic progression different from that in diploids. D 194 TE 193 EC 191 192 CO RR 190 F 196 Ph1 locus is a complex rearrangement between subtelomeric heterochromatin translocated within a cluster of cdk2-like genes localized on one of the chromosomes of the B genome (Griffiths et al. 2006). Ph1 controls meiosis at several levels. While telomere pairing is not affected, Ph1 prevents pairing between centromeres of nonhomologous chromosomes, keeps centromeres at the nuclear periphery, synchronizes chromatin condensation, and controls the expression of synapsis and recombination components (Naranjo and Corredor 2004; Moore and Shaw 2009; Knight et al. 2010; Yousafzai et al. 2010). 189 UN Editor Proof 3 Meiosis in Polyploid Plants 3.3 Recombination and Crossover Meiotic recombination in plants is initiated by multiple DSBs. Some of the breaks are repaired by homologous recombination using the non-sister chromatid as a template, resulting in reciprocal strand exchange (CO events). As the number of Layout: T1 Standard SC Chapter No.: 3 Book ID: 272454_1_En Date: 16-8-2012 234 235 236 237 238 239 240 241 242 243 244 245 246 247 248 249 250 251 252 253 254 255 256 257 258 259 260 261 262 263 264 265 266 267 268 269 270 271 F 232 233 PR OO 231 D 230 TE 229 initial DSBs exceeds the number of COs, the majority of DSBs are processed otherwise, either by separating double Holliday junction into non-crossover (NCO) products or by the synthesis-dependent strand annealing (SDSA) pathway. There are excellent recent reviews that describe details of DSB processing and components of meiotic recombination (Hamant et al. 2006; Mercier and Grelon 2008; Sanchez-Moran et al. 2008; De Muyt et al. 2009; Edlinger and Schlogelhofer 2011; Osman et al. 2011), and there is no reason to assume that the principal mechanisms are different in polyploids. However, there is growing evidence that the frequency with which DSBs give rise to COs can be regionally or generally modified by polyploidy. Therefore, we will describe some factors that influence crossover frequencies. Due to the role of recombination in the regulation of chromosome pairing and segregation, there appears to be a minimum of one CO per chromosome in many species (Youds and Boulton 2011). Beyond this, and in spite of variation between different species, there seems to be no correlation between the length of chromosomes and the number of COs (Brubaker et al. 1999; Mezard 2006). The genomes are rather composed of ‘hot spots’ and ‘cold spots’, with high and low probabilities for meiotic recombination (Drouaud et al. 2006; Mezard 2006; Kim et al. 2007). Even small local sequence divergence, like a transgene insertion, can modify meiotic recombination locally (Sun et al. 2008). Variation in recombination or CO frequency among Arabidopsis accessions (Barth et al. 2001; Sanchez-Moran et al. 2002) also indicates genetic components that regulate the meiotic recombination frequency in trans. Further, temperature and age of flowers within the plant can modify CO events (Francis et al. 2007). Another aspect to consider is CO interference, the inhibition of additional recombination events by COs in their proximity or even along the whole chromosome (Holliday 1977; van Veen and Hawley 2003; Baudat and de Massy 2007; Youds and Boulton 2011). Differences in the number of CO events in polyploids have been described in several systems. The A and D genome components of allotetraploid Gossypium (cotton) are very different in size but marker pairs have, nevertheless, comparable genetic distances if compared between diploid or allopolyploid mapping populations, respectively. Surprisingly, the comparison for the same markers between diploids and allotetraploids indicated a higher recombination rate in the latter (Brubaker et al. 1999; Desai et al. 2006). F1 hybrids between the diploid parents Brassica oleracea and Brassica rapa, compared with F1 plants derived from the same hybrid with doubled chromosome numbers after colchicine treatment, produced more progeny with intergenome recombination (Szadkowski et al. 2010, 2011). A comparison of the CO number among diploid, allotriploid, and allotetraploid hybrids from crosses between Brassica oleracea and Brassica rapa revealed the highest number for allotriploids, intermediate values for allotetraploids, and the lowest numbers in diploids (Nicolas et al. 2008; Leflon et al. 2010). This indicates stimulation of recombination by hybridity and/or polyploidy, however, in a nonlinear correlation with the number of homologous chromosomes. This might be coupled with the occurrence of univalents, in addition to bivalents, in the triploids, whereas the tetraploids exclusively form bivalents EC 228 CO RR 227 M.-L. Zielinski and O. Mittelsten Scheid UN Editor Proof 40 Book ISBN: 978-3-642-31441-4 Page: 40/54 Layout: T1 Standard SC Chapter No.: 3 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 41/54 41 292 3.4 Chromosome and Allele Segregation 293 3.4.1 Chromosome Segregation 280 281 282 283 284 285 286 287 288 289 290 294 295 296 297 298 299 300 301 302 303 304 305 306 307 308 309 310 311 PR OO 279 D 278 TE 277 EC 276 Following pairing and recombination in prophase I, chromosomes are arranged at the equatorial plane during metaphase I. The following anaphase I is distinct from that in mitosis since both centromeres of sister chromatids attach to spindles of the same pole and retain their cohesion when the chromosome pairs get dragged apart, moving to opposite poles. This ‘reductional’ division, therefore, halves the chromosome number. It is followed by metaphase II with the formation of two new equatorial planes, and anaphase II, in which the sister chromatids now become separated and distributed to the resulting four postmeiotic nuclei. A tightly controlled order of maintenance and stepwise release of cohesion specifies all stages of meiosis (Sakuno and Watanabe 2009), but there is no evidence that an increased chromosome number in polyploids affects this control. In contrast, due to the effects of polyploidy on pairing and/or recombination, there is a much higher probability of unequal segregation of chromosomes into the postmeiotic nuclei. Unequal distribution of the chromosomes can occur in many different ways (Pagliarini 2000). It can decrease the regular nuclear DNA content, if unresolved multivalents, unpaired or laggard chromosomes, or parts of chromosomes are not included or translocated to other chromosomes (Madlung et al. 2005; Charles et al. 2010; Gaeta and Pires 2010; Wang et al. 2010). Extreme cases of CO RR 274 275 F 291 (Leflon et al. 2010), indicating a control mechanism to prevent multivalents. A good candidate is the PrBn locus (Jenczewski et al. 2003; Nicolas et al. 2009; Cifuentes et al. 2010). Exploring an assay system for meiotic recombination based on fluorescent proteins expressed in seeds (Melamed-Bessudo et al. 2005), a comparison among diploids, isogenic autotetraploids, and allotetraploids generated by interspecies hybridization revealed an unexpected increase of recombination between the markers for both types of polyploids (Pecinka et al. 2011). The presence of mainly multivalent formation in one, and bivalent formation in the other, argue against a correlation with the pairing behavior. In summary, the frequency of recombination between any two markers is determined by interplay of physical distance, local cis- and trans-acting genetic control elements, the degree of overall genetic divergence, and sex-specific differences (Vizir and Korol 1990; De Vicente and Tanksley 1991; Drouaud et al. 2007; Nelson et al. 2005; Pecinka et al. 2011), regardless of ploidy. The few briefly described examples of increased recombination frequency in polyploid plants suggest that pairing behavior, potential modification of crossover interference by PPS, distinctive segregation patterns (see below), or other still unknown factors add to the complexity. However, it is tempting to speculate that high(er) recombination rates, at least in part, contributed to the prevalence of polyploidy among angiosperms and crop plants bred within the last 12,000 years. 272 273 UN Editor Proof 3 Meiosis in Polyploid Plants Layout: T1 Standard SC Chapter No.: 3 Book ID: 272454_1_En Date: 16-8-2012 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 F PR OO 317 318 D 316 TE 314 315 systematic chromosome elimination were observed in the allopolyploid Paspalum subciliatum (Adamowski et al. 1998) and in pentaploid Brachiaria decumbens (Ricci et al. 2010), due to kinetic asynchrony of the two different genomes after diakinesis. Alternatively, the nuclear DNA may be increased if restitution of nuclei occurs prior to division or around incompletely separated chromosomes (Ramsey 2007; Brownfield and Kohler 2011). This can occur either after the first or the second meiotic division, differing in the degree of heterozygosity transmitted to the progeny (Peloquin et al. 2008). While loss of chromosomal material is often deleterious for the resulting cell and leads to cell death or reduced viability and fertilization, a gain, and optimally a balanced multiplication of all chromosomes, is much less detrimental. Accordingly, unreduced gametes occur in numerous species and are a major source of polyploidization events (Leitch and Leitch 2008; Koehler et al. 2010). The frequency of unreduced gametes varies significantly between hybrids (e.g., Ortiz 1997; Lim et al. 2004), between different ploidal levels (e.g., Burton and Husband 2001; Ramsey 2007), and also depends on genetic factors (reviewed in Brownfield and Kohler 2011), external conditions, and stress factors (reviewed in Ramsey and Schemske 1998). The frequent occurrence of unreduced gametes has been explained by the lack of, or less effective, ‘pachytene checkpoint’ in plants (Li et al. 2009), but may be an inherent element of circumventing meiotic problems in interspecies hybrids by polyploidization. The occurrence of diploid or other unreduced post-meiotic cells also has consequences for the subsequent phase of the life cycle. Haploid gametophytes from diploid parents depend on functionality of each single genomic copy that encodes factors for proper gametophyte viability and fertilization potential. Gametophytes from polyploid species contain more than one copy, providing backups for defective alleles and the potential for heterozygosity even during the two and three postmeiotic cell divisions forming the male and female gametophytes, respectively. Since at least pollen selection is efficient at several levels (Ottaviano et al. 1990), avoiding haploidy might provide reduced selection pressure and increased variability of transmitted genetic information. Not directly related to meiosis but with consequences for gamete formation and fertility is the phenomenon of cytomixis, the fragmentation of chromatin and transfer of fragments among cells through cytoplasmic channels. Observed in many species, and especially in pollen mother cells (e.g., Mursalimov and Deineko 2011), one published comparison between diploid and tetraploid varieties of Withania somnifera indicated that the resulting reduction of fertility is much less pronounced in the tetraploids, along with a reduced extent of intercellular connections (Singhal and Kumar 2008). However, a correlation with polyploidy in other species awaits investigation. EC 313 CO RR 312 M.-L. Zielinski and O. Mittelsten Scheid UN Editor Proof 42 Book ISBN: 978-3-642-31441-4 Page: 42/54 Layout: T1 Standard SC Chapter No.: 3 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 43/54 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 F 357 PR OO 356 D 355 The genetic segregation of traits in polyploids can be quite different from that in diploids. Heterozygosity and multiple alleles per gamete increase the combinatorial possibilities in the progeny population substantially and make the segregation patterns deviate from the Mendelian ratios for diploids. This makes QTL analysis or any mapping process complex, as in the extreme case of highly polyploid and partially aneuploid sugarcane (Andru et al. 2011). Naturally, analysis of inheritance requires consideration of pairing behavior and the resulting chromosome segregation. Variation among hybrids, individuals, chromosomes, and chromosome segments does not allow establishing general rules. Therefore, we will depict the extremes: autotetraploids with quadrivalent formation, unbiased chromosome segregation, and tetrasomic inheritance on one hand, and allopolyploids with strict bivalent formation and disomic inheritance on the other. A Punnett square for segregation of one trait in the progeny of heterozygous autotetraploid parents with two copies of two different alleles has 36 fields (Fig. 3.3b), rather than four in the case of a heterozygous diploid (Fig. 3.3a). Deleterious recessive alleles become apparent as affected phenotypes in only 1:35, rather than 1:3 in diploids. Each individual tetraploid can have one, two, three, or four identical alleles (termed simplex, duplex, triplex, or quadruplex) at each locus, and the expected segregation of the different genotypes upon crossing with a partner of the same genotype is 1:8:18:8:1 (Fig. 3.3b). In the case of allopolyploids with strict bivalent formation between the chromosomes of common origin, each gamete carries one of each chromosome type so that the progeny are uniform (Fig. 3.3c). If the chromosomes carry different alleles, autotetraploids can produce 19 different genotypes at one locus (Fig. 3.4a), while disomic inheritance in allotetraploids is limited to 9 combinations (Fig. 3.4b). While exclusive tetrasomic or disomic inheritance in polyploids does occur, intermediate forms and changes in both directions over time are frequently observed in nature and need to be considered in genetic and population studies (Stift et al. 2008). A genetic peculiarity of polyploid meiosis that does not occur in diploid organisms is the chance of double reduction (Darlington 1929a; Butruille and Boiteux 2000). This term describes the possibility that regions from two sister chromatids become combined in the same gamete. Double reduction results from recombination events between the locus under observation and the centromere. If the two chromosomes that have recombined move to the same pole in anaphase I, there is a high probability that the distal end of the recombined chromatid will be included in the same nucleus as the distal end of the non-recombined chromatid, after anaphase II (locus B in 2 gametes in Fig. 3.5). The frequency of double reduction for a genetic locus depends on its distance from the centromere and the frequency of multivalent formation. Since segregation of traits in polyploids is so complex, but of high importance for population geneticists and plant breeders, there were many attempts to approach the problem via mathematical modeling. Ground-breaking theoretical work (Mather 1936; Fisher 1947) has more recently stimulated several TE 354 3.4.2 Allelic Segregation EC 353 43 CO RR 352 UN Editor Proof 3 Meiosis in Polyploid Plants Layout: T1 Standard SC Chapter No.: 3 Book ID: 272454_1_En Date: 16-8-2012 44 M.-L. Zielinski and O. Mittelsten Scheid PR OO F (a) (b) EC TE D Fig. 3.3 Comparison of segregation for an individual locus with two alleles upon bivalent or quadrivalent formation. Two different alleles (red, blue) in diploid parents are transmitted with equal probability, resulting in three genotypes, forming twice as many heterozygotes as each homozygote (a). Two different alleles with two copies each in tetraploids with random pairing between each of the four chromosomes result in five genotypes, with only one homozygote among 36 individuals (b). The same situation in tetraploids with strict bivalent formation between the more similar chromosome pairs renders only one combination (c) Editor Proof Book ISBN: 978-3-642-31441-4 Page: 44/54 395 396 397 398 UN CO RR (c) computational methods based on maximum-likelihood and other statistical approaches to estimate double reduction frequency and recombination and to support mapping of quantitative trait loci in polyploids with multivalent formation (Doerge and Craig 2000; Ridout et al. 2001; Wu et al. 2001a, b; Luo et al. 2004, Layout: T1 Standard SC Chapter No.: 3 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 45/54 3 Meiosis in Polyploid Plants PR OO F (a) D Editor Proof Fig. 3.4 Comparison of segregation for an individual locus with four alleles upon bivalent or quadrivalent formation. Four different alleles (red, orange, dark blue, light blue) in tetraploids with random pairing between each of the four chromosomes result in nineteen genotypes (a). The same situation in tetraploids with strict bivalent formation between the more similar chromosome pairs renders only nine combinations (b) 45 400 401 402 403 404 405 406 407 408 2006; Li et al. 2010). However, different modeling approaches do not always support the same conclusions (Ma et al. 2002; Cao et al. 2004), and it is likely that we need to extend and refine experimental data collection as well as modeling approaches to provide satisfying tools. Already difficult for strict autopolyploids with mostly polysomic inheritance, the situation is even more complex for polyploids with intermediate types of inheritance, as mentioned above. These can be caused either by segment- or chromosome-specific homology differences, or by stochastic pairing differences. Not even the situation in tetraploid yeast strains, where tetrad analysis provides excellent resolution of segregation analysis, is easy to interpret (Albertin et al. 2009; Stift et al. 2010). Allelic segregation in UN 399 CO RR EC TE (b) Layout: T1 Standard SC Chapter No.: 3 Book ID: 272454_1_En Date: 16-8-2012 M.-L. Zielinski and O. Mittelsten Scheid AA BB PR OO BB AA F Editor Proof 46 Book ISBN: 978-3-642-31441-4 Page: 46/54 A B Meiosis I AA AA BB BB Meiosis II TE D Fig. 3.5 Double reduction. Limited to polyploids, there is a certain probability that distal regions (around locus B) from two sister chromatids are included in the same gamete. This exceptional configuration occurs if two recombined chromosomes move to the same pole in anaphase I, which is excluded in diploids allopolyploid plants, as described, depends very much on the ratio of multivalent formation at the genetic locus under investigation. 411 3.5 External Influence 414 415 416 417 418 419 420 421 422 423 424 425 426 427 As for other developmental and physiological processes, progress and efficiency of meiosis can be modified by environmental influences. Temperature has been most extensively studied, with a range of different responses. Higher temperatures reduced regular meiosis in diploid Allium ursinum (Loidl 1989) and Rosa (Pecrix et al. 2011) and increased the number of unreduced pollen (reviewed in Ramsey and Schemske 1998). This could enhance the formation of polyploids under heat stress and thereby support the described adaptive role of polyploidization under adverse conditions. However, data for polyploids are rare, and partially divergent. In allopolyploid Brassica hybrids, more unreduced pollen was observed at lower temperatures (Mason et al. 2011), and meiosis in wheat was not affected within a physiological range of temperatures (Bayliss and Riley 1972). A more uniform effect across species (yeast, worms, and plants) appears to be a stimulation of meiotic recombination rates at higher temperatures (Rose and Baillie 1979; Borner et al. 2004; Francis et al. 2007). It is likely that adaptation to habitats with more or less drastic environmental challenges has selected a matching range of meiotic responses. Further, the more recently established CO RR 413 UN 412 EC 410 409 Layout: T1 Standard SC Chapter No.: 3 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 47/54 47 polyploid hybrids generated by breeders might respond quite differently than polyploids originating from spontaneous events long ago. 430 3.6 Bypassing Meiosis 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 PR OO 435 436 D 434 TE 433 Ancient polyploidization events and the prevalence of polyploids among living plants leave no doubt that the ecological and metabolic advantages of multiple chromosome complements preponderate over possible disadvantages (Soltis and Soltis 1999; Ramsey and Schemske 2002; Comai 2005; Otto 2007; Parisod et al. 2010; Jiao et al. 2011). As described, successful meiosis and gamete formation are mastered in different ways. Nevertheless, the survival and success of each new hybridization or polyploidization event depends on multiple factors during meiosis. Besides the requirements to adapt chromosome pairing and segregation, asynchrony, incompatibility of protein complex subunits, dosage effects, or differences in gene regulation can present meiotic barriers. One solution in plants to avoid such a decrease in reproduction is to avoid meiosis by apomixis. Apomixis, the asexual production of seeds, leads to offspring that is genetically identical to the mother plant and originate via different modifications of the double fertilization pathway (Gustaffson 1946; Asker and Jerling 1992). Gametophytic apomixis (Nogler 1984) is divided into apospory, where the unreduced embryo sac is formed by a cell of the nucellus, and diplospory, where the unreduced embryo sac is formed by bypassing meiosis I during megasporogenesis. Both result in an embryo sac that contains an unreduced egg cell, which develops into an embryo independent of fertilization. Although apomicts are found in diploids and polyploids, they are more common among polyploids (Asker and Jerling 1992). Apomixis is thought to have evolved several times from sexual ancestors in over 400 plant species (Nogler 1984). It is clearly associated with changes in gene expression, kinetics, or epigenetic regulation (Carman 1997; Sharbel et al. 2010). Whether polyploidization is a driving force toward apomixis, or a consequence of apomictic propagation, is a matter of debate. Apomixis harbors the potential to avoid sexual sterility caused by multivalent formation in polyploids (Horandl et al. 2011). On the other hand, and as described, polyploidy promotes the formation of unreduced gametes (Ramsey and Schemske 1998) and might thereby foster the establishment of higher ploidal levels. Polyploidization of apomicts could also protect them against extinction, by buffering against the irreversible accumulation of deleterious mutations in the absence of recombination as depicted by the metaphor of ‘‘Muller’s ratchet’’ (Darlington 1929b). Nevertheless, it is likely that these conditions are relevant in varying and individual combinations, with additional components such as highly efficient DNA repair system in asexually reproducing species (Schoen and Martens 1998) or occasional or conditional sexual reproduction (D’Souza et al. 2004). EC 432 CO RR 431 F 429 428 UN Editor Proof 3 Meiosis in Polyploid Plants Layout: T1 Standard SC Chapter No.: 3 Book ID: 272454_1_En Date: 16-8-2012 467 M.-L. Zielinski and O. Mittelsten Scheid 3.7 Manipulating Meiosis 500 References 501 502 503 504 505 Adamowski ED, Pagliarini MS, Batista LAR (1998) Chromosome elimination in paspalum subciliatum (Notata group). Sex Plant Reprod 11(5):272–276 Agashe B, Prasad CK, Siddiqi I (2002) Identification and analysis of DYAD: a gene required for meiotic chromosome organisation and female meiotic progression in Arabidopsis. Development 129(16):3935–3943 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 PR OO 473 474 D 472 TE 471 EC 470 CO RR 469 F 499 While natural selection in sexually reproducing species rapidly eradicates failures in meiosis, these are of interest for plant breeders. If not the seed, but the surrounding fruit is the wanted product, for example in Citrus species or melons, customers prefer products with small, reduced, or no seeds. Besides exploiting developmental mutations, one effective and popular way to achieve such fruits is the generation of triploids (Sanchez-Moran et al. 2002) by crossing tetraploids with diploids. Triploid embryos are prone to abort subsequently (Kamiri et al. 2011). Since fruit development is often regulated by phytohormones originating from the developing seeds (Dorcey et al. 2009), this strategy depends on their seed-independent substitution (Pandolfini 2009). However, once generated, triploid plants are often seedless, due to a high number of univalents in meiosis, a resulting low number of balanced gametes, and poor pollination and fertilization rates. This principle is applied in banana (Heslop-Harrison and Schwarzacher 2007) and watermelon (Beaulieu and Lea 2006), or is combined with selection for developmental mutants as for squash (Menezes et al. 2005). Once a beneficial combination of genetic traits by crossing is achieved, breeders want to fix this for further progeny (Wijnker and de Jong 2008; Chan 2010). Programed switches between crossing and apomixis would, therefore, be desirable (Spillane et al. 2004), especially for polyploids with their complex segregation patterns. Recent advances in provoking apomixis have been made by strategies to disturb meiosis. A mutation in the Arabidopsis gene SWI/DYAD (Mercier et al. 2001; Agashe et al. 2002) leads to the formation of unreduced egg cells, albeit at low frequencies. This is due to failure in female meiosis and maintains complete maternal heterozygosity in the triploid progeny (Ravi et al. 2008). Combining three mutations affecting different steps of meiosis (osd1/Atspo11-1/Atrec8) turned meiosis into mitosis and resulted in unreduced male and female gametes (d’Erfurth et al. 2009). Consequently, ploidy in subsequent generations was doubled, accompanied by reduced fertility. Nevertheless, the achievement of fertilization-independent seed development is a step toward introducing apomixis into crop plants, and of special interest for polyploid plants. The work in the OMS lab was supported by grants from the Austrian Science Fund FWF P18986 and I489. 468 UN Editor Proof 48 Book ISBN: 978-3-642-31441-4 Page: 48/54 Layout: T1 Standard SC Chapter No.: 3 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 49/54 49 EC TE D PR OO F Albertin W, Marullo P, Aigle M, Bourgais A, Bely M, Dillmann C, De Vienne D, Sicard D (2009) Evidence for autotetraploidy associated with reproductive isolation in Saccharomyces cerevisiae: towards a new domesticated species. J Evol Biol 22(11):2157–2170 Anderson LK, Stack SM (2005) Recombination nodules in plants. Cytogenet Genome Res 109 (1–3):198–204 Andru S, Pan YB, Thongthawee S, Burner DM, Kimbeng CA (2011) Genetic analysis of the sugarcane (Saccharum spp.) cultivar LCP 85-384’. I. Linkage mapping using AFLP, SSR, and TRAP markers. Theor Appl Genet 123(1):77–93 Armstrong SJ, Franklin FCH, Jones GH (2003) A meiotic time-course for Arabidopsis thaliana. Sex Plant Reprod 16(3):141–149 Asker S, Jerling L (1992) Apomixis in plants. CRC Press, Boca Raton Avivi L, Feldman M (1980) Arrangement of chromosomes in the interphase nucleus of plants. Hum Genet 55(3):281–295 Barth S, Melchinger AE, Devezi-Savula B, Lubberstedt T (2001) Influence of genetic background and heterozygosity on meiotic recombination in Arabidopsis thaliana. Genome 44(6): 971–978 Bass HW, Marshall WF, Sedat JW, Agard DA, Cande WZ (1997) Telomeres cluster de novo before the initiation of synapsis: a three-dimensional spatial analysis of telomere positions before and during meiotic prophase. J Cell Biol 137(1):5–18 Baudat F, de Massy B (2007) Regulating double-stranded DNA break repair towards crossover or non-crossover during mammalian meiosis. Chromosome Res 15(5):565–577 Bayliss MW, Riley R (1972) Analysis of temperature-dependent asynapsis in Triticum aestivum. Genet Res 20(2):193 Beaulieu JC, Lea JM (2006) Characterization and semiquantitative analysis of volatiles in seedless watermelon varieties using solid-phase microextraction. J Agric Food Chem 54(20):7789–7793 Bennett MD (1977) The time and duration of meiosis. Philos Trans R Soc Lond B Biol Sci 277(955):201–226 Bhatt AM, Canales C, Dickinson HG (2001) Plant meiosis: the means to 1 N. Trends Plant Sci 6(3):114–121 Borner GV, Kleckner N, Hunter N (2004) Crossover/noncrossover differentiation, synaptonemal complex formation, and regulatory surveillance at the leptotene/zygotene transition of meiosis. Cell 117(1):29–45 Brownfield L, Kohler C (2011) Unreduced gamete formation in plants: mechanisms and prospects. J Exp Bot 62(5):1659–1668 Brubaker CL, Paterson AH, Wendel JF (1999) Comparative genetic mapping of allotetraploid cotton and its diploid progenitors. Genome 42(2):184–203 Burton TL, Husband BC (2001) Fecundity and offspring ploidy in matings among diploid, triploid and tetraploid Chamerion angustifolium (Onagraceae): consequences for tetraploid establishment. Heredity 87:573–582 Butruille DV, Boiteux LS (2000) Selection-mutation balance in polysomic tetraploids: impact of double reduction and gametophytic selection on the frequency and subchromosomal localization of deleterious mutations. Proc Nat Acad Sci USA 97(12):6608–6613 Cao DC, Osborn TC, Doerge RW (2004) Correct estimation of preferential chromosome pairing in autotetraploids. Genome Res 14(3):459–462 Carlton PM, Farruggio AP, Dernburg AF (2006) A link between meiotic prophase progression and crossover control. PLoS Genet 2:119–128 Carman JG (1997) Asynchronous expression of duplicate genes in angiosperms may cause apomixis, bispory, tetraspory, and polyembryony. Biol J Linn Soc 61(1):51–94 Carvalho A, Delgado M, Baroa A, Frescatada M, Ribeiro E, Pikaard CS, Viegas W, Neves N (2010) Chromosome and DNA methylation dynamics during meiosis in the autotetraploid Arabidopsis arenosa. Sex Plant Reprod 23(1):29–37 Chan SWL (2010) Chromosome engineering: power tools for plant genetics. Trends Biotechnol 28(12):605–610 CO RR 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 UN Editor Proof 3 Meiosis in Polyploid Plants Layout: T1 Standard SC Chapter No.: 3 Book ID: 272454_1_En Date: 16-8-2012 EC TE D PR OO F St Charles J, Hamilton ML, Petes TD (2010) Meiotic chromosome segregation in triploid strains of Saccharomyces cerevisiae. Genetics 186(2):537–550 Cifuentes M, Grandont L, Moore G, Chevre AM, Jenczewski E (2010) Genetic regulation of meiosis in polyploid species: new insights into an old question. New Phytol 186(1):29–36 Comai L (2005) The advantages and disadvantages of being polyploid. Nat Rev Genet 6(11):836– 846 Comai L, Tyagi AP, Lysak MA (2003) FISH analysis of meiosis in Arabidopsis allopolyploids. Chromosome Res 11(3):217–226 Corredor E, Lukaszewski AJ, Pachon P, Allen DC, Naranjo T (2007) Terminal regions of wheat chromosomes select their pairing partners in meiosis. Genetics 177(2):699–706 Cowan CR, Carlton PM, Cande WZ (2001) ‘The polar arrangement of telomeres in interphase and meiosis. Rabl organization and the bouquet. Plant Physiol 125(2):532–538 d’Erfurth I, Jolivet S, Froger N, Catrice O, Novatchkova M, Mercier R (2009) Turning meiosis into mitosis. PLoS Biol 7(6):e1000124 D’Souza TG, Storhas M, Schulenburg H, Beukeboom LW, Michiels NK (2004) Occasional sex in an ‘asexual’ polyploid hermaphrodite. Proc Biol Sci 271(1543):1001–1007 Darlington CD (1929a) Chromosome behaviour and structural hybridity in the tradescantiae. J Genet 21(2):207–286 Darlington CD (1929b) Polyploids and polyploidy. Nature 124:62–64 Dawe RK (1998) Meiotic chromosome organization and segregation in plants. Annu Rev Plant Physiol Plant Mol Biol 49:371–395 De Muyt A, Mercier R, Mezard C, Grelon M (2009) Meiotic recombination and crossovers in plants. Genome Dyn 5:14–25 Desai A, Chee PW, Rong J, May OL, Paterson AH (2006) Chromosome structural changes in diploid and tetraploid A genomes of Gossypium. Genome 49(4):336–345 De Vicente MC, Tanksley SD (1991) Genome-wide reduction in recombination of backcross progeny derived from male versus female gametes in an interspecific cross of tomato. Theor Appl Genet 83(2):173–178 Doerge RW, Craig BA (2000) Model selection for quantitative trait locus analysis in polyploids. Proc Nat Acad Sci USA 97(14):7951–7956 Dong F, Jiang J (1998) Non-Rabl patterns of centromere and telomere distribution in the interphase nuclei of plant cells. Chromosome Res 6(7):551–558 Dorcey E, Urbez C, Blazquez MA, Carbonell J, Perez-Amador MA (2009) Fertilizationdependent auxin response in ovules triggers fruit development through the modulation of gibberellin metabolism in Arabidopsis. Plant J 58(2):318–332 Drouaud J, Camilleri C, Bourguignon PY, Canaguier A, Berard A, Vezon D, Giancola S, Brunel D, Colot V, Prum B et al (2006) Variation in crossing-over rates across chromosome 4 of Arabidopsis thaliana reveals the presence of meiotic recombination ‘‘hot spots’’. Genome Res 16(1):106–114 Drouaud J, Mercier R, Chelyshev L, Bérard A, Falque M, Martin O, Zanni V, Brunel D, Mézard C (2007) Sex-specific crossover distributions and variations in interference level along Arabidopsis thaliana chromosome 4. PLoS Genet 3(6):e106 Edlinger B, Schlogelhofer P (2011) Have a break: determinants of meiotic DNA double strand break (DSB) formation and processing in plants. J Exp Bot 62(5):1545–1563 Fisher RA (1947) The theory of linkage in polysomic inheritance. Philos Trans R Soc Lond B Biol Sci 233(594):55–87 Francis KE, Lam SY, Harrison BD, Bey AL, Berchowitz LE, Copenhaver GP (2007) Pollen tetrad-based visual assay for meiotic recombination in Arabidopsis. Proc Nat Acad Sci USA 104(10):3913–3918 Gaeta RT, Pires JC (2010) Homologous recombination in allopolyploids: the polyploid ratchet. New Phytol 186(1):18–28 Griffiths S, Sharp R, Foote TN, Bertin I, Wanous M, Reader S, Colas I, Moore G (2006) Molecular characterization of Ph1 as a major chromosome pairing locus in polyploid wheat. Nature 439(7077):749–752 CO RR 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 M.-L. Zielinski and O. Mittelsten Scheid UN Editor Proof 50 Book ISBN: 978-3-642-31441-4 Page: 50/54 Layout: T1 Standard SC Chapter No.: 3 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 51/54 51 EC TE D PR OO F Gupta PK, Fedak G (1985) Genetic control of meiotic chromosome pairing in polyploids in the genus Hordeum. Can J Genet Cytol 27(5):515–530 Gustaffson A (1946) Apomixis in higher plants. Lunds Universitets Arsskrift 42(2):1–370 Hamant O, Ma H, Cande WZ (2006) Genetics of meiotic prophase I in plants. Annu Rev Plant Biol 57:267–302 Harper L, Golubovskaya I, Cande WZ (2004) A bouquet of chromosomes. J Cell Sci 117(Pt 18): 4025–4032 Harrison CJ, Alvey E, Henderson IR (2010) Meiosis in flowering plants and other green organisms. J Exp Bot 61(11):2863–2875 Heslop-Harrison JS, Schwarzacher T (2007) Domestication, genomics and the future for banana. Ann Bot 100(5):1073–1084 Holliday R (1977) Recombination and meiosis. Philos Trans R Soc Lond B Biol Sci 277(955): 359–370 Horandl E, Cosendai AC, Rodewald J (2011) Origin and distribution of autopolyploids via apomixis in the alpine species Ranunculus kuepferi (Ranunculaceae). Taxon 60(2):355–364 Jannoo N, Grivet L, David J, D’Hont A, Glaszmann JC (2004) Differential chromosome pairing affinities at meiosis in polyploid sugarcane revealed by molecular markers. Heredity 93(5):460–467 Jenczewski E, Alix K (2004) From diploids to allopolyploids: the emergence of efficient pairing control genes in plants. Crit Rev Plant Sci 23(1):21–45 Jenczewski E, Eber F, Grimaud A, Huet S, Lucas MO, Monod H, Chevre AM (2003) PrBn, a major gene controlling homologous pairing in oilseed rape (Brassica napus) haploids. Genetics 164(2):645–653 Jiao Y, Wickett NJ, Ayyampalayam S, Chanderbali AS, Landherr L, Ralph PE, Tomsho LP, Hu Yi, Liang H, Soltis PS et al (2011) Ancestral polyploidy in seed plants and angiosperms. Nature 473(7345):97–100 John B, Henderson SA (1962) Asynapsis and polyploidy in Schistocerca paranensis. Chromosoma 13(2):111–147 Jones GH (1994) Meiosis in autopolyploid crepis-capillaris 3. Comparison of triploids and tetraploids: evidence for nonindependence of autonomous pairing sites. Heredity 73:215–219 Jones GH, Vincent JE (1994) Meiosis in autopolyploid crepis capillaris 2. Autotetraploids. Genome 37(3):497–505 Kamiri M, Stift M, Srairi I, Costantino G, Moussadik AE, Hmyene A, Bakry F, Ollitrault P, Froelicher Y (2011) Evidence for non-disomic inheritance in a citrus interspecific tetraploid somatic hybrid between C. reticulata and C. limon using SSR markers and cytogenetic analysis. Plant Cell Rep 30(8):1415–1425 Kim S, Plagnol V, Hu TT, Toomajian C, Clark RM, Ossowski S, Ecker JR, Weigel D, Nordborg M (2007) Recombination and linkage disequilibrium in Arabidopsis thaliana. Nat Genet 39(9):1151–1155 Knight E, Greer E, Draeger T, Thole V, Reader S, Shaw P, Moore G (2010) Inducing chromosome pairing through premature condensation: analysis of wheat interspecific hybrids. Funct Integr Genomics 10(4):603–608 Koehler C, Mittelsten OS, Erilova A (2010) The impact of the triploid block on the origin and evolution of polyploid plants. Trends Genet 26(3):142–148 Leflon M, Grandont L, Eber F, Huteau V, Coriton O, Chelysheva L, Jenczewski E, Chevre AM (2010) Crossovers get a boost in brassica allotriploid and allotetraploid hybrids. Plant Cell 22(7):2253–2264 Leitch AR, Leitch IJ (2008) Genomic plasticity and the diversity of polyploid plants. Science 320(5875):481–483 Li J, Das K, Fu G, Tong C, Li Y, Tobias C, Wu R (2010) Em algorithm for mapping quantitative trait Loci in multivalent tetraploids. Int J Plant Genomics 2010:216547 Li XC, Barringer BC, Barbash DA (2009) The pachytene checkpoint and its relationship to evolutionary patterns of polyploidization and hybrid sterility. Heredity 102(1):24–30 CO RR 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 UN Editor Proof 3 Meiosis in Polyploid Plants Layout: T1 Standard SC Chapter No.: 3 Book ID: 272454_1_En Date: 16-8-2012 EC TE D PR OO F Lim KB, Shen TM, Barba-Gonzalez R, Ramanna MS, Van Tuyl JM (2004) Occurrence of SDR 2 N-gametes in Lilium hybrids. Breed Sci 54(1):13–18 Lohmiller LD, De Muyt A, Howard B, Offenberg HH, Heyting C, Grelon M, Anderson LK (2008) Cytological analysis of MRE11 protein during early meiotic prophase I in Arabidopsis and tomato. Chromosoma 117(3):277–288 Loidl J (1989) Effects of elevated temperature on meiotic chromosome synapsis in Allium ursinum. Chromosoma 97(6):449–458 Loidl J (1995) Meiotic chromosome pairing in triploid and tetraploid Saccharomyces cerevisiae. Genetics 139(4):1511–1520 Lukaszewski AJ, Kopecky D (2010) The Ph1 locus from wheat controls meiotic chromosome pairing in autotetraploid rye (Secale cereale L.). Cytogenet Genome Res 129(1–3):117–123 Luo ZW, Zhang RM, Kearsey MJ (2004) Theoretical basis for genetic linkage analysis in autotetraploid species. Proc Nat Acad Sci USA 101(18):7040–7045 Luo ZW, Zhang Z, Zhang RM, Pandey M, Gailing O, Hattemer HH, Finkeldey R (2006) Modeling population genetic data in autotetraploid species. Genetics 172(1):639–646 Ma CX, Casella G, Shen ZJ, Osborn TC, Wu RL (2002) A unified framework for mapping quantitative trait loci in bivalent tetraploids using single-dose restriction fragments: a case study from alfalfa. Genome Res 12(12):1974–1981 Madlung A, Tyagi AP, Watson B, Jiang HM, Kagochi T, Doerge RW, Martienssen R, Comai L (2005) Genomic changes in synthetic Arabidopsis polyploids. Plant J 41(2):221–230 Martinez-Perez E, Shaw P, Aragon-Alcaide L, Moore G (2003) Chromosomes form into seven groups in hexaploid and tetraploid wheat as a prelude to meiosis. Plant J 36(1):21–29 Martinez-Perez E, Shaw P, Moore G (2001) The Ph1 locus is needed to ensure specific somatic and meiotic centromere association. Nature 411(6834):204–207 Mason AS, Nelson MN, Yan G, Cowling WA (2011) Production of viable male unreduced gametes in Brassica interspecific hybrids is genotype specific and stimulated by cold temperatures. BMC Plant Biol 11:103 Mather K (1936) Segregation and linkage in autotetraploids. J Genetics 32(2):287–314 McKee BD (1996) The license to pair: identification of meiotic pairing sites in Drosophila. Chromosoma 105(3):135–141 McKim KS (2007) Meiotic pairing: a place to hook up. Curr Biol 17(5):R165–R168 Melamed-Bessudo C, Yehuda E, Stuitje AR, Levy AA (2005) A new seed-based assay for meiotic recombination in Arabidopsis thaliana. Plant J 43(3):458–466 Menezes CB, Maluf WR, Azevedo SM, Faria MV, Nascimento IR, Nogueira DW, Gomes LA, Bearzoti E (2005) Inheritance of parthenocarpy in summer squash (Cucurbita pepo L.). Genet Mol Res 4(1):39–46 Mercier R, Grelon M (2008) Meiosis in plants: ten years of gene discovery. Cytogenet Genome Res 120(3–4):281–290 Mercier R, Vezon D, Bullier E, Motamayor JC, Sellier A, Lefevre F, Pelletier G, Horlow C (2001) SWITCH1 (SWI1): a novel protein required for the establishment of sister chromatid cohesion and for bivalent formation at meiosis. Genes Dev 15(14):1859–1871 Mestiri I, Chague V, Tanguy AM, Huneau C, Huteau V, Belcram H, Coriton O, Chalhoub B, Jahier J (2010) Newly synthesized wheat allohexaploids display progenitor-dependent meiotic stability and aneuploidy but structural genomic additivity. New Phytol 186(1):86–101 Mezard C (2006) Meiotic recombination hotspots in plants. Biochem Soc Trans 34:531–534 Mezard C, Vignard J, Drouaud J, Mercier R (2007) The road to crossovers: plants have their say. Trends Genet 23(2):91–99 Moore G, Shaw P (2009) Improving the chances of finding the right partner. Curr Opin Genet Dev 19(2):99–104 Morrison JW, Rajhathy T (1960) Frequency of quadrivalents in autotetraploid plants. Nature 187(4736):528–530 Mursalimov SR, Deineko EV (2011) An ultrastructural study of cytomixis in tobacco pollen mother cells. Protoplasma 248(4):717–724 CO RR 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 M.-L. Zielinski and O. Mittelsten Scheid UN Editor Proof 52 Book ISBN: 978-3-642-31441-4 Page: 52/54 Layout: T1 Standard SC Chapter No.: 3 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 53/54 53 EC TE D PR OO F Naranjo T, Corredor E (2004) Clustering of centromeres precedes bivalent chromosome pairing of polyploid wheats. Trends Plant Sci 9(5):214–217 Nelson MN, Nixon J, Lydiate DJ (2005) Genome-wide analysis of the frequency and distribution of crossovers at male and female meiosis in Sinapis alba L. (white mustard). Theor Appl Genet 111(1):31–43 Nicolas SD, Leflon M, Liu Z, Eber F, Chelysheva L, Coriton O, Chevre AM, Jenczewski E (2008) Chromosome ‘speed dating’ during meiosis of polyploid Brassica hybrids and haploids. Cytogenet Genome Res 120(3–4):331–338 Nicolas SD, Leflon M, Monod H, Eber F, Coriton O, Huteau V, Chevre AM, Jenczewski E (2009) Genetic regulation of meiotic cross-overs between related genomes in Brassica napus haploids and hybrids. Plant Cell 21(2):373–385 Nogler GA (1984) Gametophytic apomixis. In: Johri BM (ed) Embryology of angiosperms. Springer, Berlin Ortiz R (1997) Occurrence and inheritance of 2n pollen in Musa. Ann Bot 79(4):449–453 Osman K, Higgins JD, Sanchez-Moran E, Armstrong SJ, Franklin F, Chris H (2011) Pathways to meiotic recombination in Arabidopsis thaliana. New Phytol 190(3):523–544 Ottaviano E, Sari Gorla M, Mulcahy DL (1990) Pollen selection: efficiency and monitoring. In: Ogita ZI, Markert CL (eds) Isozymes: structure, function, and use in biology and medicine. Wiley-Liss, New York, pp 575–588 Otto SP (2007) The evolutionary consequences of polyploidy. Cell 131(3):452–462 Ozkan H, Feldman M (2009) Rapid cytological diploidization in newly formed allopolyploids of the wheat (Aegilops-Triticum) group. Genome 52(11):926–934 Pagliarini MS (2000) Meiotic behavior of economically important plant species: the relationship between fertility and male sterility. Genet Mol Biol 23(4):997–1002 Pandolfini T (2009) Seedless fruit production by hormonal regulation of fruit set. Nutrients 1(2):168–177 Parisod C, Holderegger R, Brochmann C (2010) Evolutionary consequences of autopolyploidy. New Phytol 186(1):5–17 Pawlowski WP, Cande WZ (2005) Coordinating the events of the meiotic prophase. Trends Cell Biol 15(12):674–681 Pawlowski WP (2010) Chromosome organization and dynamics in plants. Curr Opin Plant Biol 13(6):640–645 Pecinka A, Fang W, Rehmsmeier M, Levy Avraham A, Mittelsten Scheid O (2011) Polyploidization increases meiotic recombination frequency in Arabidopsis. BMC Biol 9:24 Pecrix Y, Rallo G, Folzer H, Cigna M, Gudin S, Le Bris M (2011) Polyploidization mechanisms: temperature environment can induce diploid gamete formation in Rosa sp. J Exp Bot 62(10): 3587–3597 Peloquin SJ, Boiteux LS, Simon PW, Jansky SH (2008) A chromosome-specific estimate of transmission of heterozygosity by 2n gametes in potato. J Hered 99(2):177–181 Ramsey J (2007) Unreduced gametes and neopolyploids in natural populations of Achillea borealis (Asteraceae). Heredity 98(3):143–150 Ramsey J, Schemske DW (1998) Pathways, mechanisms, and rates of polyploid formation in flowering plants. Annu Rev Ecol Syst 29:467–501 Ramsey J, Schemske DW (2002) Neopolyploidy in flowering plants. Annu Rev Ecol Syst 33:589–639 Ravi M, Marimuthu MP, Siddiqi I (2008) Gamete formation without meiosis in Arabidopsis. Nature 451(7182):1121–1124 Ricci GCL, Pagliarini MS, Valle CB (2010) Genome elimination during microsporogenesis in two pentaploid accessions of Brachiaria decumbens (Poaceae). Genet Mol Res 9(4): 2364–2371 Ridout MS, Bell JA, Simpson DW (2001) Analysis of segregation data from selfed progeny of allopolyploids. Heredity 87:537–543 Riley R, Chapman V (1958) Genetic control of the cytologically diploid behaviour of hexaploid wheat. Nature 182(4637):713–715 CO RR 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768 769 770 771 772 773 UN Editor Proof 3 Meiosis in Polyploid Plants Layout: T1 Standard SC Chapter No.: 3 Book ID: 272454_1_En Date: 16-8-2012 EC TE D PR OO F Rose AM, Baillie DL (1979) Effect of temperature and parental age on recombination and nondisjunction in Caenorhanditis elegans. Genetics 92(2):409–418 Sakuno T, Watanabe Y (2009) Studies of meiosis disclose distinct roles of cohesion in the core centromere and pericentromeric regions. Chromosome Res 17(2):239–249 Sanchez-Moran E, Armstrong SJ, Santos JL, Franklin FCH, Jones GH (2002) Variation in chiasma frequency among eight accessions of Arabidopsis thaliana. Genetics 162(3): 1415–1422 Sanchez-Moran E, Osman K, Higgins JD, Pradillo M, Cunado N, Jones GH, Franklin FCH (2008) ASY1 coordinates early events in the plant meiotic recombination pathway. Cytogenet Genome Res 120(3–4):302–312 Santos JL, Alfaro D, Sanchez-Moran E, Armstrong SJ, Franklin FC, Jones GH (2003) Partial diploidization of meiosis in autotetraploid Arabidopsis thaliana. Genetics 165(3):1533–1540 Scherthan H (2007) Telomere attachment and clustering during meiosis. Cell Mol Life Sci 64(2): 117–124 Schoen I, Martens K (1998) DNA repair in ancient asexuals: a new solution to an old problem? J Nat Hist 32:943–948 Schubert I, Shaw P (2011) Organization and dynamics of plant interphase chromosomes. Trends Plant Sci 16(5):273–281 Sharbel TF, Voigt ML, Corral JM, Galla G, Kumlehn J, Klukas C, Schreiber F, Vogel H, Rotter B (2010) Apomictic and sexual ovules of Boechera display heterochronic global gene expression patterns. Plant Cell 22(3):655–671 Sheehan MJ, Pawlowski WP (2009) Live imaging of rapid chromosome movements in meiotic prophase I in maize. Proc Natl Acad Sci USA 106(49):20989–20994 Simioni C, do Valle CB (2011) Meiotic analysis in induced tetraploids of Brachiaria decumbens Stapf. Crop Breed Appl Biotech 11(1):43–49 Singhal VK, Kumar P (2008) Cytomixis during microsporogenesis in the diploid and tetraploid cytotypes of Withania somnifera (L.) Dunal, 1852 (Solanaceae). Comp Cytogenet 2(1):85–92 Soltis DE, Soltis PS (1999) Polyploidy: recurrent formation and genome evolution. Trends Ecol Evol 14(9):348–352 Spillane C, Curtis MD, Grossniklaus U (2004) Apomixis technology development-virgin births in farmers fields? Nat Biotechnol 22(6):687–691 Stack SM, Anderson LK (2002) Crossing over as assessed by late recombination nodules is related to the pattern of synapsis and the distribution of early recombination nodules in maize. Chromosome Res 10(4):329–345 Stift M, Reeve R, van Tienderen PH (2010) Inheritance in tetraploid yeast revisited: segregation patterns and statistical power under different inheritance models. J Evol Biol 23(7):1570–1578 Stift Marc, Berenos C, Kuperus P, van Tienderen PH (2008) Segregation models for disomic, tetrasomic and intermediate inheritance in tetraploids: a general procedure applied to Rorippa (Yellow cress) microsatellite data. Genetics 179(4):2113–2123 Sun X, Zhang Y, Yang S, Chen JQ, Hohn B, Tian D (2008) Insertion DNA promotes ectopic recombination during meiosis in Arabidopsis. Mol Biol Evol 25(10):2079–2083 Sundstrom G, Larsson TA, Larhammar D (2008) Phylogenetic and chromosomal analyses of multiple gene families syntenic with vertebrate Hox clusters. BMC Evol Biol 8:254 Sybenga J (1996) Chromosome pairing affinity and quadrivalent formation in polyploids: do segmental allopolyploids exist? Genome 39(6):1176–1184 Szadkowski E, Eber F, Huteau V, Lode M, Coriton O, Jenczewski E, Chevre AM (2011) Polyploid formation pathways have an impact on genetic rearrangements in resynthesized Brassica napus. New Phytol 191(3):884–894 Szadkowski E, Eber F, Huteau V, Lode M, Huneau C, Belcram H, Coriton O, ManzanaresDauleux MJ, Delourme R, King GJ et al (2010) The first meiosis of resynthesized Brassica napus, a genome blender. New Phytol 186(1):102–112 Trelles-Sticken E, Loidl J, Scherthan H (2003) Increased ploidy and KAR3 and SIR3 disruption alter the dynamics of meiotic chromosomes and telomeres. J Cell Sci 116(12):2431–2442 CO RR 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792 793 794 795 796 797 798 799 800 801 802 803 804 805 806 807 808 809 810 811 812 813 814 815 816 817 818 819 820 821 822 823 824 825 826 M.-L. Zielinski and O. Mittelsten Scheid UN Editor Proof 54 Book ISBN: 978-3-642-31441-4 Page: 54/54 Layout: T1 Standard SC Chapter No.: 3 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 55/54 55 EC TE D PR OO F van Veen JE, Hawley RS (2003) Meiosis: when even two is a crowd. Curr Biol 13(21): R831–R833 Vizir IY, Korol AB (1990) Sex difference in recombination frequency in Arabidopsis. Heredity 65(3):379–383 von Wettstein D, Rasmussen SW, Holm PB (1984) The synaptonemal complex in genetic segregation. Annu Rev Genet 18:331–413 Wang Y, Jha AK, Chen R, Doonan JH, Yang M (2010) Polyploidy-associated genomic instability in Arabidopsis thaliana. Genesis 48(4):254–263 Watanabe K (1981) Studies on the control of diploid-like meiosis in polyploid taxa of Chrysanthemim japonese. Cytologia 46(3):459–498 Weiss H, Maluszynska J (2000) Chromosomal rearrangement in autotetraploid plants of Arabidopsis thaliana. Hereditas 133(3):255–261 Wijnker E, de Jong H (2008) Managing meiotic recombination in plant breeding. Trends Plant Sci 13(12):640–646 Wu RL, Gallo-Meagher M, Littell RC, Zeng ZB (2001a) A general polyploid model for analyzing gene segregation in outcrossing tetraploid species. Genetics 159(2):869–882 Wu SS, Wu RL, Ma CX, Zeng ZB, Yang MC, Casella G (2001b) A multivalent pairing model of linkage analysis in autotetraploids. Genetics 159(3):1339–1350 Youds JL, Boulton SJ (2011) The choice in meiosis: defining the factors that influence crossover or non-crossover formation. J Cell Sci 124(Pt 4):501–513 Yousafzai FK, Al-Kaff N, Moore G (2010) The molecular features of chromosome pairing at meiosis: the polyploid challenge using wheat as a reference. Funct Integr Genomics 10(2):147–156 CO RR 827 828 829 830 831 832 833 834 835 836 837 838 839 840 841 842 843 844 845 846 847 848 849 UN Editor Proof 3 Meiosis in Polyploid Plants Metadata of the chapter that will be visualized in SpringerLink Book Title Polyploidy and Genome Evolution Series Title Chapter Title Origins of Novel Phenotypic Variation in Polyploids Copyright Year 2012 Copyright HolderName Springer-Verlag Berlin Heidelberg Corresponding Author Family Name Martienssen Particle Given Name Robert A. Suffix Division Author Organization Cold Spring Harbor Laboratory Address One Bungtown Road, 11724, Cold Spring Harbor, NY, USA Email martiens@cshl.edu Family Name Finigan Particle Given Name Patrick Suffix Division Organization Cold Spring Harbor Laboratory Address One Bungtown Road, 11724, Cold Spring Harbor, NY, USA Email Author Family Name Tanurdzic Particle Given Name Milos Suffix Division Organization Cold Spring Harbor Laboratory Address One Bungtown Road, 11724, Cold Spring Harbor, NY, USA Email Abstract Polyploid species represent a special type of organism in nature, one that can survive and compete with three or more full sets of homologous chromosomes. While less common in the animal and fungal kingdoms, polyploid species are highly prevalent in the plant kingdom. Indeed, most agricultural crops are polyploids, typically because polyploidy confers greater robustness and therefore higher yields. Among many examples of novel phenotypic variation exhibited by polyploids are the production of larger fruits, reduced tillering, delays in the reproductive transition, and even the creation of visually stunning flower pigmentation patterns coveted by gardeners. The source of this novel variation in polyploids is still largely unclear. However, multiple cellular mechanisms have been proposed, with some supporting evidence, to explain novel variation. We review some of these mechanisms here. Book ISBN: 978-3-642-31441-4 Page: 57/75 Chapter 4 4 Patrick Finigan, Milos Tanurdzic and Robert A. Martienssen 9 10 11 12 13 14 15 16 D 8 TE 7 Abstract Polyploid species represent a special type of organism in nature, one that can survive and compete with three or more full sets of homologous chromosomes. While less common in the animal and fungal kingdoms, polyploid species are highly prevalent in the plant kingdom. Indeed, most agricultural crops are polyploids, typically because polyploidy confers greater robustness and therefore higher yields. Among many examples of novel phenotypic variation exhibited by polyploids are the production of larger fruits, reduced tillering, delays in the reproductive transition, and even the creation of visually stunning flower pigmentation patterns coveted by gardeners. The source of this novel variation in polyploids is still largely unclear. However, multiple cellular mechanisms have been proposed, with some supporting evidence, to explain novel variation. We review some of these mechanisms here. EC 5 6 PR OO 3 Origins of Novel Phenotypic Variation in Polyploids 2 19 20 21 22 23 24 25 CO RR 17 18 F 1 Book ID: 272454_1_En Date: 16-8-2012 4.1 Prevalence and Significance of Polyploids Polyploid organisms have three or more complete sets of homologous chromosomes (Winge 1917 #1369; Ramsey 2002 #1334). Polyploidy, or whole-genome duplication (WGD), can arise through multiple ways—most often from unreduced gametes following meiosis (Ahloowalia 1961 #1252; Harlan 1975 #1285; Jørgensen 1928 #1290; Newton 1929 #1318; Skalinska 1946 #1344; Ramsey 1998 #1333; Bretagnolle 1995 #1258). In mammals, WGD events are considered to be rare and typically lethal, presumably due to deleterious effects associated with UN Editor Proof Layout: T1 Standard SC Chapter No.: 4 P. Finigan  M. Tanurdzic  R. A. Martienssen (&) Cold Spring Harbor Laboratory, One Bungtown Road, Cold Spring Harbor, NY 11724, USA e-mail: martiens@cshl.edu P. S. Soltis and D. E. Soltis (eds.), Polyploidy and Genome Evolution, DOI: 10.1007/978-3-642-31442-1_4,  Springer-Verlag Berlin Heidelberg 2012 57 Layout: T1 Standard SC Chapter No.: 4 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 58/75 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 F 32 33 PR OO 31 D 30 TE 29 EC 28 dosage (Bertrand 2010 #1255; Van de Peer 2009 #1361) see also Chap. 18, this volume). However, genomic sequencing has revealed evidence for a few widespread ancient WGD events in the animal kingdom (Maere 2005 #1308; Dehal 2005 #1270; Kellis 2004 #1293; Ohno 1970 #1322; Meyer 2005 #1314, 1999 #1313) (see also Chaps. 16 and 17, this volume). Extant vertebrate genomes in particular are believed to be the result of two or three separate WGD events during evolution (Maere 2005 #1308; Dehal 2005 #1270; Kellis 2004 #1293; Ohno 1970 #1322; Meyer 2005 #1314, 1999 #1313) (Chap. 16, this volume). In plants, polyploidy is generally tolerated, and most plant species are recent or ancient polyploids (Lexer 2003 #1302; Soltis et al. 2009 #1346; Jiao 2011 #1289). Polyploidy is thought to play a major role in speciation (Clausen 1945 #1267; Grant 1981 #1282; Lumaret 1988 #1306), as polyploids are often reproductively isolated from their progenitors—crosses between diploids and tetraploids generally yield triploid progeny with unbalanced meiosis, leading to polyploid infertility (Darlington 1963 #1269; Grant 1981 #1282). On a simplified level, Polyploid organisms can be either autopolyploids or allopolyploids (Kihara 1926 #1296). Following one definition, an autopolyploid is a Polyploid organism in which all of the chromosome sets are derived from the same species, whereas an allopolyploid organism has chromosome sets derived from different species (Kihara 1926 #1296). Allopolyploids are therefore permanent hybrids, but with a complete chromosome set from each parental species (Kihara 1926 #1296). The complete parental chromosome sets in allopolyploids allow for proper pairing of homologous chromosomes during meiosis (disomic segregation), rather than missegregation of trivalents and quadrivalents as may be found in autopolyploids (multisomic segregation) (Stebbins 1971 #1355; Ramsey 2002 #1334). Thus, allopolyploids have often been considered more stable than autopolyploids, and ‘‘fix’’ hybrid genotypes in successive generations, along with any beneficial or detrimental phenotypes that result (Winge 1932 #1370). However, it has been recognized that autopolyploids may have functional (either disomic or multisomic) chromosome pairing and be highly fertile (Soltis et al. 2007 #1349; Soltis and Rieseberg 1986 #1347). Many important agricultural crops are polyploids (Eigsti 1957 #1275). Polyploid crops typically demonstrate increased growth, including flower and fruit size, and novel variation compared to their diploid counterparts that can make them better suited as agricultural products (Stebbins 1971 #1355; Ramsey 2002 #1334; Grant 1981 #1282). The modern-day bread wheat, Triticum aestivum, is an allohexaploid that is the combination of three different diploid species (Dubcovsky 2007 #1272) (see also Chap. 7, this volume). T. aestivum has largely replaced its diploid progenitors and now accounts for about 95 % of the entire wheat crop produced around the world (Dubcovsky 2007 #1272). In addition to wheat, many staple crops (including corn, cotton, coffee, oat, canola, rye, apple, banana, watermelon, potato, sugar cane, and soybean) are also recent or ancient polyploids (Gaut 1997 #1281; Stebbins 1971 #1355; Ohno 1970 #1322; Shoemaker 1996 #1341; Lagercrantz 1996 #1298; Eigsti 1957 #1275). The reduced seed set of certain polyploids (odd ploidal levels) can make them more desirable for CO RR 26 27 P. Finigan et al. UN Editor Proof 58 Layout: T1 Standard SC Chapter No.: 4 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 59/75 59 79 4.2 Origin of Novel Variation in Neopolyploids 77 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 PR OO 76 The prevalence of polyploids in the plant kingdom despite the rarity of their formation (Ramsey 1998 #1333) indicates that polyploids may have a fitness advantage compared to their progenitors (Stebbins 1950 #1354). The explanation for this fitness advantage stems from the observation that many polyploids display novel phenotypic variation compared to their progenitors (Randolph 1941 #1335; Levin 1983 #1300, 2002 #1301; Lumaret 1988 #1306; Ramsey 2002 #1334; Müntzing 1936 #1316). This novel phenotypic variation is believed to enable polyploids to exploit different environmental niches better than their progenitors; this variation bestows a fitness advantage on the polyploids (Clausen 1945 #1267; Stebbins 1950 #1354). One caveat of this hypothesis is that studies documenting the geographic distribution of polyploids and their diploid progenitors used extant diploid relatives instead of the exact diploid progenitors (usually unknown) (Stebbins 1971 #1355; Ramsey 2002 #1334). Thus, it is possible the real progenitors had the same geographic range distribution as the current polyploids. To address this problem, many current studies of polyploids use synthetic polyploids that can be compared to their exact diploid progenitors (Stebbins 1971 #1355; Ramsey 2002 #1334). Future studies should help to clarify if range expansion relative to their progenitors is a general characteristic of polyploids. The emergence of novel phenotypes in polyploids, especially neoallopolyploids (defined as early-generation polyploids), compared to their progenitors has been of great interest because of its application in plant breeding and relevance to biodiversity and evolution (Eigsti 1957 #1275; Stebbins 1971 #1355; Ramsey 2002 #1334; Grant 1981 #1282). Genetic mechanisms encompassing gene dosage/allelic combinations, novel gene interactions, genomic alterations, and epigenomic reorganization have all been proposed to explain the origin of novel phenotypes of polyploids (Fig. 4.1). The relevance of these different mechanisms for causing novel variation is still largely unclear, with important distinctions between autopolyploids and allopolyploids. D 75 TE 74 EC 73 CO RR 72 F 78 consumption, like seedless watermelons and bananas. Moreover, allopolyploids can overcome hybrid incompatibility between different species, allowing production of new varieties and introgression of favorable traits into important crops (Eigsti 1957 #1275). Allohexaploid Nicotiana plants were used as a bridge to transfer the gene responsible for tobacco mosaic virus resistance from wild to commercial varieties of Nicotiana species (Eigsti 1957 #1275). This was only possible through a Polyploid intermediate, as the interspecies Nicotiana hybrids were sterile (Eigsti 1957 #1275) (see also Chap. 11, this volume). 71 UN Editor Proof 4 Origins of Novel Phenotypic Variation in Polyploids Layout: T1 Standard SC Chapter No.: 4 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 60/75 Editor Proof 60 P. Finigan et al. G ag os en D PR OO lic om ic lle Al /A te ic en G F (b) ra tio ns (a) e R e en TE G g in (c) el od eg ul em D at R or ic y et In en te ig ra ct Ep io ns Novel Phenotypic Variation (d) 109 110 111 112 113 114 4.2.1 Gene Dosage and Allelic Combinations May Result in Novel Variation in Neopolyploids Genome-wide increases in DNA content have been proposed to contribute to novel phenotypes in neopolyploids (Randolph 1941 #1335). The ‘‘gigas effect’’ is a welldocumented phenotype associated with polyploid formation, where certain organs are larger in the polyploids than their progenitors—but not necessarily the entire plant (Randolph 1941 #1335; Stebbins 1971 #1355). The gigas effect is believed to UN 108 CO RR EC Fig. 4.1 Molecular mechanisms that may contribute to novel phenotypic variation. Four distinct molecular mechanisms have been suggested to contribute to changes in gene expression and function that may result in novel phenotypic variation in neoallopolyploids compared to their progenitors—adapted from (Osborn 2003 #1324). The figure diagrams each mechanism by showing the two progenitor states in comparison with their resultant polyploid offspring. a Changes in genic or allelic dosage are diagramed by the inheritance of novel chromosome sets. b Genomic alterations are shown by the inheritance of recombined chromosomes. c Novel gene regulatory interactions are diagrammed by unique protein–protein interactions. d Epigenomic reorganization is shown as the inheritance of large-scale changes in chromatin compaction. The progenitors and their polyploid offspring are diagrammed by ellipses, with the progenitors represented by the ellipses furthest from the central ‘‘novel phenotypic variation’’ ellipse. The polyploids are represented by the ellipses closest to the central ellipse. Layout: T1 Standard SC Chapter No.: 4 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 61/75 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 F 120 PR OO 119 D 118 TE 117 61 be a consequence of increasing the genome size, leading to the increase in cell size and reduction of cell divisions that are commonly associated with neopolyploids compared to their progenitors (Noggle 1946 #1321; Stebbins 1971 #1355). Additionally, some evidence suggests that neopolyploids have a protracted reproductive growth phase, with a delay in the onset of flowering and longer reproductive growth (Stebbins 1971 #1355). While these changes in morphology and growth may be the result of simply increasing the amount of DNA in neopolyploids, they may also stem from increasing the relative level of gene and protein expression. Gene and protein expression may be sensitive to gene dosage effects, whereby increases in ploidy result in changes in expression that are not proportional to the increase in gene dosage. Gene and protein expression studies in Zea, Solanum, Helianthus, and Arabidopsis autoploid series have suggested that the vast majority of changes in expression are proportional to gene dosage (Yu 2010 #1376; Pignatta 2010 #1327; Wang 2006a, b #1366; Stupar 2007 #1357; Guo 1996 #1283; Yao 2011 #1375; Riddle 2010 #1337; Church 2009 #1266). Pignatta et al. (2010) created independent neo-autotetraploid lines of Arabidopsis thaliana (Columbia) from the same homogeneous diploid progenitor. The researchers performed genome-wide gene expression studies to identify reproducible changes in gene expression in the autotetraploids compared to the diploid progenitor. Only a few genes, out of 26,107, were found that potentially displayed changes in gene expression that were not proportional to the ploidal levels. Of note, resolution or sampling limitations may have missed some dosage-sensitive genes (Pignatta 2010 #1327). A recent study with Columbia and Landsberg accessions of Arabidopsis ploidal series found similar results to those above, with slightly higher numbers of genes whose expression was disproportionately affected by changes in ploidy (Yu 2010 #1376). A greater role for gene dosage was identified in potato and maize studies, as increasing the ploidal levels resulted in roughly 10 % of genes with disproportionate effects on expression (Stupar 2007 #1357; Guo 1996 #1283; Yao 2011 #1375; Riddle 2010 #1337; Church 2009 #1266). Taken together, these studies suggest that most gene expression is proportional to ploidal level. None of these studies identified dosage-sensitive genes whose increase in expression correlated with any ploidy-dependent phenotype. One unanswered question is: what effect does increasing the overall level of gene expression, without changing the relative abundance of gene products, have on biological pathways? In theory, increased gene dosage and mixtures in polyploids could result in unique allelic combinations that would expand phenotypic diversity (Bingham 1979 #1256; Grant 1981 #1282; Stebbins 1971 #1355). Instead of possessing only two alleles for a given gene as in a diploid, a tetraploid could possess four alleles that may allow additional variability for a phenotypic trait. This mechanism of allelic interaction is a natural extension of phenotypic variation in diploids attributed to combinations of different alleles that affect quantitative traits (Guo 1994 #1377; Birchler 2001 #1257; Osborn 2003 #1324). However, it remains unclear what allelic combinations contribute to novel variation in polyploids EC 116 CO RR 115 UN Editor Proof 4 Origins of Novel Phenotypic Variation in Polyploids Layout: T1 Standard SC Chapter No.: 4 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 62/75 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 184 185 186 187 188 189 190 191 192 193 194 F PR OO 164 Hybrid gene interactions may contribute to phenotypic variation in allopolyploids (Osborn 2003 #1324; Ramsey 2002 #1334). Similar to interspecies hybrids, allopolyploids have divergent genetic contributions from their progenitors. This heterogeneity could result in perturbations in regulatory networks that could dramatically alter their outcomes (Ramsey 2002 #1334; Stebbins 1947 #1353; Stebbins 1950 #1354; Osborn 2003 #1324). In synthetic Arabidopsis allopolyploids, altered regulatory interactions between the two progenitor-derived chromosome sets are thought to be responsible for delaying flowering time (Wang 2006a, b #1365). The A. arenosa FRIGIDA (FRI) gene is a transcription factor that trans-activates the A. thaliana (Landsberg) Flowering Locus C (FLC) gene (Wang 2006a, b #1365). FLC is a major negative regulator of the reproductive transition in Arabidopsis species, and its over-expression is predicted to delay flowering in the Arabidopsis neoallopolyploids (Shindo 2005 #1340; Simpson 2002 #1342). While FRI and FLC genes are present in both the progenitors, A. thaliana (Landsberg) FRI and A. arenosa FLC alleles have reduced expression or function (Wang 2006a, b #1365). Therefore, it is possible that the flowering delay might be explained by the novel genic interaction between A. arenosa FRI and A. thaliana FLC in the neoallopolyploids (Wang 2006a, b #1365). D 163 TE 162 4.2.2 Hybrid Regulatory Interactions Can Cause Novel Phenotypic Variation EC 161 (Osborn 2003 #1324), and indeed allelic variation in neopolyploids may be limited due to genetic ‘‘bottlenecking’’ in their formation. 4.2.3 Chromosome Missegregation Can Lead to Novel Variation in Neopolyploids CO RR 159 160 P. Finigan et al. Perturbations in chromosome segregation have been proposed to be a major source of novel phenotypic variation in polyploids (Stebbins 1971 #1355; Müntzing 1937 #1317; Ramsey 2002 #1334; Song 1993 #1352, 1995 #1351; Soltis and Soltis 1999 #1348). During meiotic prophase, the increased chromosome sets in polyploids will often lead to multivalent chromosome pairing that can result in slower cell division, unequal chromosome partitioning to daughter cells, and even homeologous recombination (Storchova 2006 #1356; Stebbins 1971 #1355; Ramsey 2002 #1334). For example, autotetraploids have four homologous chromosomes instead of two, like their diploid counterparts. These four homologous chromosomes may form univalents, trivalents, or quadrivalents during prophase that will cause lagging or unequal chromosome partitioning during anaphase and result in aneuploid gametes (Ramsey 2002 #1334). The prevalence of multivalent formation in UN Editor Proof 62 Layout: T1 Standard SC Chapter No.: 4 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 63/75 201 202 203 204 205 206 207 208 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 236 237 238 239 F 200 PR OO 199 D 198 TE 197 63 polyploids is widespread; a survey of the available literature by Ramsey et al. (2002) found that the mean multivalent frequency for auto- and allopolyploids was estimated to be 28.8 and 8.0 %, respectively. The lower percentage of allopolyploids with multivalent pairing versus autopolyploids, is most likely the result of sequence divergence between homeologous chromosomes in allopolyploids that favors disomic pairing. Multivalent pairing is expected to result in a much higher degree of aneuploidy than disomic pairing and ultimately reduce the fertility of polyploids (Ramsey 2002 #1334; Stebbins 1971 #1355). Ramsey et al. (2002) compared the occurrence of aneuploidy in gametic and sporophytic cells in different Polyploid species. They found that the mean frequency of aneuploid pollen was approximately 40 %, while the mean frequency of aneuploid progeny was an estimated 29 %. Importantly, there were no studies on the frequency of aneuploid ovules, but if one were to assume a similar ratio of aneuploidy as found in pollen, it would imply that 64 % of all zygotes would be aneuploids. These differences in the expected and observed percentages of aneuploid progeny suggest that some aneuploid gametes or progeny do not survive and that chromosome segregation defects correlate with the reduced fertility of polyploid. In contrast to the very different multivalent frequencies observed at meiosis, the frequency of aneuploid gametes and progeny was similar in auto- and allopolyploids. If aneuploidy requires multivalent pairing, it is unclear how allopolyploids could have a similar level of aneuploidy as autopolyploids. Either there is a much higher level of multivalent pairing in allopolyploids than reported, or other factors contribute to the unexpected level of aneuploidy in allopolyploids (Ramsey 2002 #1334; Chester 2012 #1265). In addition to aneuploidy, multivalent pairing of homeologous chromosomes can also result in genetic recombination that can severely disrupt genome organization and contribute to phenotypic variation (Stebbins 1971 #1355; Ramsey 2002 #1334; Song 1993 #1352, 1995 #1351). In allopolyploids, unequal recombination between chromosomes can result from synteny between homeologous chromosomes. Changes in gene dosage, gene expression, the epigenetic landscape, and even gene conversion are all possible outcomes of homeologous recombination in allopolyploids. Homeologous recombination has been observed in many different resynthesized allopolyploid species (Wendel 2000 #1368; Osborn 2003 #1324; Doyle 2008 #1271; Hegarty 2008 #1287; Leitch 2008 #1299; Soltis and Soltis 2009 #1346; Stebbins 1971 #1355; Grant 1981 #1282; Ramsey 2002 #1334; Chester 2012 #1265; Pires 2004 #1328; Buggs 2009 #1259), suggesting it may be a major mechanism of genome alteration and evolution. In synthetic Arabidopsis allopolyploids derived from A. thaliana and A. arenosa progenitors, the nucleolus organizer region (NOR) and 5S rDNA genetic regions have been shown to be lost or recombined between the chromosomes derived from A. thaliana and A. arenosa by the F3 generation (Pontes 2004 #1329). Rearrangements of the NOR and 5S rDNA regions were also demonstrated in the naturally occurring Arabidopsis allopolyploid, A. suecica. Genetic rearrangements in resynthesized Brassica napus allopolyploids have been linked to phenotypic variation (Gaeta 2007 #1280; Pires 2004 #1328; Xiong 2011 #1372; Song 1995 #1351). In one striking example, EC 196 CO RR 195 UN Editor Proof 4 Origins of Novel Phenotypic Variation in Polyploids Layout: T1 Standard SC Chapter No.: 4 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 64/75 247 248 249 250 251 252 253 254 255 256 257 258 259 260 261 262 263 264 265 266 267 268 269 270 271 272 273 274 275 276 277 278 F PR OO 246 4.2.4 Epigenetic Remodeling Can Result in Novel Variation in Neopolyploids Epigenetic remodeling of polyploid genomes has also been suggested to play a role in the origin of novel phenotypes (Osborn 2003 #1324). Epigenetic changes involving DNA methylation, histone post-translational modifications (PTMs), histone replacement, and sRNA-mediated silencing have all been demonstrated to affect gene expression levels in diploids (Wolffe 1999 #1371; Calarco 2011 #1378). Moreover, the dynamic nature of these epigenetic modifications in combination with the instability of neopolyploid genomes would suggest that there is the potential for rampant epigenetic remodeling in polyploids that could affect gene expression (Osborn 2003 #1324; Wolffe 1999 #1371; Ramsey 2002 #1334; Matzke 1999 #1310). Concurrently, there have been many studies documenting novel changes in DNA methylation (Li 2010 #1303; Chen 2008 #1264; Lukens 2006 #1305; Xu 2009 #1373; Wang 2009 #1364, 2004 #1367; Liu 2003 #1304; Madlung 2002 #1307; Kenan-Eichler 2011 #1294; Parisod 2009 #1325; Yaakov 2011 #1374) as well as changes in sRNA profiles (Kenan-Eichler 2011 #1294; Ha 2009 #1284; Preuss 2008 #1331) within allopolyploid genomes compared to their progenitors, but there has been limited data demonstrating links of these epigenetic changes to phenotypic consequences. One classic example of epigenetic reprogramming in allopolyploids is the occurrence of nucleolar dominance among interspecies hybrids, including allopolyploids (McStay 2006 #1311; Preuss 2007 #1330). Nucleolar dominance occurs when rRNA genes from one parent, or progenitor species, are preferentially silenced in a hybrid; this silencing is neither the result of random inactivation, nor correlated with imprinting and sexual parent of origin (McStay 2006 #1311; Preuss 2007 #1330). In Arabidopsis allopolyploids, small interfering RNAs (sRNA), DNA methylation, and histone deacetylation result in the silencing of A. thaliana-derived rRNA genes in A. suecica (Preuss 2007 #1330, 2008 #1331; Earley 2006 #1274). While no clear phenotype is associated with nucleolar dominance, it is believed to be important for genome stability and prevention of premature aging, which has been linked to perturbations in rDNA genic regions (Finigan 2008 #1277). D 245 TE 243 244 EC 242 variation for flowering time between two different B. napus allopolyploid lines was found to correlate with a genomic rearrangement disrupting the expression of FLC (Pires 2004 #1328). These allopolyploid lines also displayed phenotypic variation in flowering time that exceeded the range of their progenitors (Gaeta 2007 #1280; Pires 2004 #1328). Together, these results provide powerful evidence for the role of genomic rearrangements in neopolyploids driving the creation of novel phenotypic variation that is not present in their diploid progenitors. CO RR 240 241 P. Finigan et al. UN Editor Proof 64 Layout: T1 Standard SC Chapter No.: 4 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 65/75 286 287 288 289 290 291 292 293 294 295 296 297 298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 F PR OO 284 285 D 283 TE 282 EC 281 65 For histone PTMs, a few studies in Arabidopsis neoallopolyploids demonstrate changes in histone modification patterns that correlate with changes in gene expression (Wang 2006a, b #1365; Ni et al. 2009 #1320). Interestingly, these changes in gene expression patterns were linked to phenotypic changes in flowering time (Wang 2006a, b #1365) and chlorophyll and sugar content (Ni et al. 2009 #1320). However, these perturbations are most likely the downstream consequences of novel regulatory interactions (Ni et al. 2009 #1320; Wang 2006a, b #1365). Additionally, a tentative link between seed death and perturbations in histone H3 lysine 27 methylation (H3K27me) has been identified in Arabidopsis neoallopolyploids (Josefsson 2006 #1291; Walia 2009 #1363). Josefsson et al. (2006) reported that the expression of PHERES1, MEDEA, and other imprinted genes were deregulated in developing Arabidopsis neoallopolyploid seeds. In addition to the imprinted genes, Athila retrotransposons from the pollen parent were also deregulated in Arabidopsis allotriploid seed, which are inviable. This deregulation of imprinted genes and Athila elements was suggested to result from a lack of silencing of their pollen parent copies by the POLYCOMB REPRESSIVE COMPLEX (PRC), which is responsible for H3K27me in Arabidopsis plants (Josefsson 2006 #1291). Subsequent work suggested that the PRC complex was not functioning in these neoallopolyploids because of a 2–10-fold down-regulation in the expression of FERTILIZATION-INDEPENDENT SEED 2 (FIS2), one of the subunits of the PRC (Walia 2009 #1363). Why the PRC complex, and FIS2, are misregulated in the developing neoallopolyploids is still an unanswered question, but major disruptions of at least H3K27 methylation patterns in these neoallopolyploids would be predicted by this model (Josefsson 2006 #1291; Walia 2009 #1363). Future studies to analyze H3K27 methylation patterns in the neoallopolyploids will be important to validate this hypothesis. This misregulation of the PRC complex is only present in the neoallopolyploids but absent in its autopolyploid progenitors and depends on both the ploidy level and genetic diversity of the progenitors (Josefsson 2006 #1291; Walia 2009 #1363). So far, the limited evidence for epigenetic remodeling has been primarily identified in allo-, and not autopolyploids. While allopolyploids inherit epigenetic and genomic divergence from their progenitors, one might still expect to see epigenomic remodeling in autopolyploids due to their dynamic nature. One role for epigenetic mechanisms in autopolyploids is in the origin of unreduced gametes from which they arise. In the ovules of most sexual flowering plants, female gametogenesis is initiated from a single surviving gametic cell, the functional megaspore, formed after meiosis of the somatically derived megaspore mother cell (MMC). The Arabidopsis small RNA binding protein ARGONAUTE 9 (AGO9) controls female gamete formation by restricting the specification of gametophyte precursors to the MMC (Olmedo-Monfil 2010 #1323; Durán-Figueroa 2010 #1273). Mutations in AGO9 lead to the differentiation of diploid (unreduced) gametic cells from the surrounding ovule that are able to initiate gametogenesis. Mutations in the maize AGO9 homolog also lead to unreduced gametes, through a related but distinct mechanism (Singh 2011 #1343). AGO9 preferentially interacts with 24-nucleotide sRNAs derived from transposable elements (TEs), and its CO RR 279 280 UN Editor Proof 4 Origins of Novel Phenotypic Variation in Polyploids Layout: T1 Standard SC Chapter No.: 4 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 66/75 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 F 329 PR OO 328 4.2.5 Nonadditive Gene Expression and Novel Phenotypic Variation Mechanisms encompassing gene dosage/allelic combination, novel gene interactions, genomic alterations, and epigenomic reorganization must all converge on gene expression and function to induce novel phenotypic variation. While little is known about perturbations in gene function in neopolyploids, changes in gene expression (Wang 2006a, b #1366, #1365; Akhunova 2010 #1253; Pumphrey 2009 #1332; Chaudhary 2009 #1263; Chagué 2010 #1261; Flagel 2008 #1278, 2010 #1279; Hegarty 2006 #1286) and protein levels (Albertin 2006 #1254; Ng 2011 #1319) have been demonstrated in allopolyploids compared to their progenitors. Transcript profiling experiments in synthetic Arabidopsis allopolyploids were the first demonstrations of genome-wide nonadditive gene expression in allopolyploids (Wang 2006a, b #1366). Nonadditive gene expression occurs when the expression level of genes in hybrids differs from the average expression level of the orthologous genes in the parental species (i.e. midparent level). In Arabidopsis neoallopolyploids, an estimated 5–35 % of all genes were nonadditively expressed—depending on the statistical method employed (Wang 2006a, b #1366). Subsequent research in other neoallopolyploid species has suggested that nonadditive gene expression is a general trend found associated with recent allopolyploidization; however, the amount of nonadditive gene expression in specific allopolyploids is highly variable (Hegarty 2006 #1286; Chaudhary 2009 #1263; Flagel 2008 #1278, 2010 #1279; Akhunova 2010 #1253; Pumphrey 2009 #1332; Chagué 2010 #1261). For example, in bread wheat allohexaploids nonadditive gene expression has been reported as anywhere from 7 to 40 % (Pumphrey 2009 #1332; Chagué 2010 #1261; Kashkush 2002 #1292; Akhunova 2010 #1253). The discovery of nonadditive gene expression in neoallopolyploids implies that the progenitor-derived chromosome sets must be interacting to result in novel gene expression patterns. How these chromosome sets interact is just beginning to be unraveled; but, by definition, these novel interactions must stem directly from changes in cis- and trans-regulatory divergence between the progenitor-derived genomes. Cis- and/or trans-regulatory differences could result in changes in the expression of homeologous genes (homeoalleles) from each progenitor-derived chromosome set. Cis-regulatory divergence can act directly on single genes or on D 327 TE 326 activity is necessary to silence TEs in female gametes and their accessory cells (Olmedo-Monfil 2010 #1323; Durán-Figueroa 2010 #1273). That AGO9-dependent sRNA silencing is also crucial to specify ploidy in the gametes indicates that epigenetic reprogramming may link transposon silencing to germ cell fate (Slotkin 2009 #1345). This may reflect the coevolution of transposon activity and sexual reproduction, as well as the increased tolerance of polyploids to mutations caused by TE insertions (Martienssen 2010 #1309). EC 325 CO RR 324 P. Finigan et al. UN Editor Proof 66 Layout: T1 Standard SC Chapter No.: 4 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 67/75 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 F 370 PR OO 368 369 D 367 TE 366 localized chromatin domains, such as promoters or enhancers, resulting in asymmetric accumulation of homeologous transcripts in allopolyploids. Transregulatory divergence between the parental genomes will affect homeologous genes equally and result in equal accumulation of homeologous transcripts. Thus, both cis- and trans-regulatory differences could result in nonadditive homeologous gene expression, but only cis-regulatory differences could lead to biased homeoallele-specific gene expression. The relative contribution of cis- and trans-regulatory divergence to nonadditive gene expression is still unclear, but some evidence has arisen recently implicating a major role of cis-regulatory differences for this phenomenon. Transcriptome profiling studies in Gossypium (Flagel 2010 #1279; Doyle 2008 #1271), Arabidopsis (Wang 2006a, b #1366), and maize (Schnable 2011 #1339), and small-scale analyses in Tragopogon (Buggs 2011 #1260; Tate 2006 #1359) polyploids have suggested that there is a common theme of ‘‘genome dominance’’ in homeologous gene expression in allopolyploids (Rapp 2009 #1336). In the genome dominance model, one of the progenitor transcription profiles outcompetes the other to shift the overall expression in the allopolyploids toward a progenitor-specific expression profile. In Gossypium allopolyploids, five different allopolyploids have been suggested to have a 54–60 % transcriptional bias in favor of the D genome versus the A genome (Flagel 2010 #1279). A similar phenomenon has been reported in Arabidopsis allopolyploids, where a 55 % bias toward A. arenosa over A. thaliana homeoallele expression has been reported for natural allopolyploids (Chang 2010 #1262) and suggested for synthetic allopolyploids (Wang 2006a, b #1366). In contrast, the evidence for genome dominance has been inconclusive in Triticum (Akhunova 2010 #1253; Pumphrey 2009 #1332; Chagué 2010 #1261). However, this phenomenon is still un-resolved, because the majority of the evidence for genome dominance are based on comparing the expression profiles between the progenitors and their allopolyploid offspring, without distinguishing between homeoalleles in the allopolyploids (Flagel 2008 #1278, 2010 #1279; Wang 2006a, b #1366). Further, in cases where genome-wide profiles of homeologous gene expression have suggested biased expression of one genome over the other (Akhunova 2010 #1253; Chang 2010 #1262; Schnable 2011 #1339), the exact progenitors of these established allopolyploids were extinct and true comparisons could not be made. Future studies should address these discrepancies and help to unravel the phenotypic consequences of nonadditive gene expression. EC 365 67 CO RR 364 4.3 Technical Considerations When Comparing Gene Expression Studies in Polyploids UN Editor Proof 4 Origins of Novel Phenotypic Variation in Polyploids Many of the controversies surrounding genome dominance and the variability of nonadditive gene expression in neoallopolyploids may stem from the differences in the methodologies employed and the organisms studied. One important distinction Layout: T1 Standard SC Chapter No.: 4 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 68/75 Editor Proof 68 P. Finigan et al. (a) Species α‘ (4x = 2n) Polyploidization Species α‘ β‘ (4x = 2n) Hybridization Polyploidization (b) Species α (2x = 2n) ‘ Species α‘ β (2x = 1n) PR OO Species β‘ (4x = 2n) Species β (2x = 2n) F Species α (2x = 2n) Species α‘ β‘ (4x = 2n) Polyploidization Hybridization TE D Species β (2x = 2n) 405 406 407 408 409 410 411 412 413 414 415 416 417 418 is whether polyploidization preceded hybridization or vice versa (Hegarty 2008 #1287). If polyploidization precedes hybridization then the resulting neoallopolyploid will have duplicate chromosomes that can accurately pair at meiosis (Fig. 4.2). Synthetic Arabidopsis allopolyploids are an example, and natural Arabidopsis allopolyploids are believed to have arisen this way (Hegarty 2008 #1287; Josefsson 2006 #1291). On the other hand, if polyploidization follows hybridization, then the resulting hybrid will first transit through an unbalanced chromosome state, most likely severely infertile, before restoration with a polyploidization event to create a neoallopolyploid (Fig. 4.2) (Hegarty 2006 #1286, 2008 #1287). Gossypium, Triticum, Tragopogon, Spartina, and Senecio allopolyploids are examples of the latter type (Hegarty 2006 #1286). Transcriptome profiling studies in Senecio (Hegarty 2006 #1286) and Triticum (Feldman 2005 #1276) hybrids before and after polyploidization suggested that the homologous gene expression levels were more perturbed (vs. the midparent) in the sterile hybrids before chromosome doubling. These results may suggest that gene UN 404 CO RR EC Fig. 4.2 The origin of allopolyploids. Two types of allopolyploids are distinguished by the relative order of polyploidization and hybridization. a Type I allopolyploids arise when polyploidization precedes hybridization, typically via unreduced gametes. Type I neoallopolyploids will have duplicate chromosomes that properly pair in meiosis. b Type II allopolyploids are created when hybridization precedes polyploidization. Type II allopolyploids transit through a hybrid state, which is usually sterile, as chromosomes do not pair during meiosis. A subsequent polyploidization event restores 2n chromosome number in allopolyploids Layout: T1 Standard SC Chapter No.: 4 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 69/75 69 455 4.4 Conclusions 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 456 457 458 459 PR OO 425 D 423 424 TE 422 EC 421 CO RR 420 F 454 expression profiles in the resulting allopolyploid may be influenced by the creation of a hybrid intermediate. Another important distinction is the nature of the progenitor species used to compare with allopolyploid expression profiles. Many allopolyploids do not have extant progenitors, and relatives must be used instead (Chang 2010 #1262; Rapp 2009 #1336; Flagel 2010 #1279). These relatives may not be exact genetic matches (Chang 2010 #1262; Rapp 2009 #1336; Flagel 2010 #1279) and may contain epigenetic differences as well, as is the case for different accessions of A. thaliana (Vaughn 2007 #1362; Kliebenstein 2006 #1297). A recent study illustrates this point. An estimated 6,790 homologous genes were differentially expressed between A. suecica and synthetic Arabidopsis allopolyploids (Chang 2010 #1262), and comparative genome sequencing revealed that 938 homeologous genes were missing from A. suecica compared to the synthetic allopolyploid (Chang 2010 #1262). These genomic differences are comparable to those between different Arabidopsis thaliana accessions (Vaughn 2007 #1362; Kliebenstein 2006 #1297) and may represent differences between the A. suecica progenitors and their extant relatives. Additionally, some studies employ progenitor species that have been severely genetically manipulated (Akhunova 2010 #1253; Chang 2010 #1262; Kerber 1964 #1295; Wang 2006a, b #1366). Previous analysis of fifth-generation Arabidopsis allopolyploids have relied on a tetraploidized A. thaliana (Landsberg accession) line produced spontaneously from root explant regenerants that underwent a callus phase and cell culture treatment (Chang 2010 #1262; Valvekens 1988 #1360; Wang 2006a, b #1366; Comai 2000 #1268). Plant lines regenerated from callus/cell cultures typically display heritable changes in genetic and epigenetic regulation—namely, somaclonal variation (Phillips 1994 #1326; Meins 2003 #1312; Mohan Jain 2001 #1315; Tanurdzic 2008 #1379). These changes could have a profound effect on gene expression and epigenetic regulation in neoallopolyploids derived from such progenitors. Classical approaches based on colchicine treatment have also been applied to the generation of tetraploid A. thaliana lines (Henry 2005 #1288; Santos 2003 #1338). Studies in Gossypium (Rapp 2009 #1336; Chaudhary 2009 #1263), Triticum (Akhunova 2010 #1253; Pumphrey 2009 #1332; Chagué 2010 #1261), and Tragopogon (Tate et al. 2009) have also relied on colchicine treatments to polyploidize sterile triploid hybrids after fertilization. Importantly, there was no difference in genetic or DNA methylation changes between spontaneous or colchicine produced Brassica polyploids (Gaeta 2007 #1280). 419 UN Editor Proof 4 Origins of Novel Phenotypic Variation in Polyploids The prevalence of polyploidy in nature suggests it has been positively selected for during evolution. Indeed, the higher yields of polyploids among agricultural crops and their usefulness to bridge species barriers have led to their widespread use and creation for the agricultural industry (Eigsti 1957 #1275). Understanding the Layout: T1 Standard SC Chapter No.: 4 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 70/75 P. Finigan et al. 472 References 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 Ahloowalia B, Garber F (1961) The genus Collinsia. XIII. Cytogenetic studies of interspecific hybrids involving species with pediceled flowers. Bot Gaz 122:219 Akhunova A, Matniyazov R, Liang H, Akhunov E (2010) Homoeolog-specific transcriptional bias in allopolyploid wheat. BMC Genomics 11(1):505 Albertin W, Balliau T, Brabant P, Chevre A-M, Eber F, Malosse C, Thiellement H (2006) Numerous and rapid Nonstochastic modifications of Gene products in newly synthesized 10.1534/ Brassica napus Allotetraploids. Genetics 173(2):1101–1113. doi: genetics.106.057554 Bertrand D, Gagnon Y, Blanchette M, El-Mabrouk N (2010) Reconstruction of ancestral genome subject to whole genome duplication, speciation, rearrangement and loss. Paper presented at the Proceedings of the 10th international conference on algorithms in bioinformatics, Liverpool, UK Bingham E (1979) Maximizing heterozygosity in autopolyploids. Basic Life Sci 13:471–489 Birchler JA, Bhadra U, Bhadra MP, Auger DL (2001) Dosage-dependent Gene regulation in multicellular Eukaryotes: implications for dosage compensation, Aneuploid syndromes, and quantitative traits. Dev Biology 234(2):275–288. doi: 10.1006/dbio.2001.0262 Bretagnolle F, Thompson J (1995) Tansley review no. 78. Gametes with the somatic chromosome number: mechanisms of their formation and role in the evolution of Autopolyploid plants. New Phytol 129:1 Buggs RJA, Doust AN, Tate JA, Koh J, Soltis K, Feltus FA, Paterson AH, Soltis PS, Soltis DE (2009) Gene loss and silencing in Tragopogon miscellus (Asteraceae): comparison of natural and synthetic Allotetraploids. Heredity 103 (1):73–81. doi: http://www.nature.com/hdy/ journal/v103/n1/suppinfo/hdy200924s1.html Buggs Richard JA, Zhang L, Miles N, Tate Jennifer A, Gao L, Wei W, Schnable Patrick S, Barbazuk WB, Soltis Pamela S, Soltis Douglas E (2011) Transcriptomic shock generates evolutionary novelty in a newly formed. Natural allopolyploid plant. Curr Biol 21(7):551–556. doi: 10.1016/j.cub.2011.02.016 Calarco JP, Martienssen RA (2011) Genome reprogramming and small interfering RNA in the Arabidopsis germline. Curr Opin Genet Dev 21(2):134–139. doi: 10.1016/j.gde.2011.01.014 Chagué V, Just J, Mestiri I, Balzergue S, Tanguy A-M, Huneau C, Huteau V, Belcram H, Coriton O, Jahier J, Chalhoub B (2010) Genome-wide gene expression changes in genetically stable synthetic and natural wheat allohexaploids. New Phytol 187(4):1181–1194. doi: 10.1111/ j.1469-8137.2010.03339.x 466 467 468 469 470 PR OO 465 D 464 TE 463 EC 462 CO RR 461 F 471 emergent properties of neopolyploids, such as novel phenotypic variation and reproductive barriers, is vitally important to further advance the usefulness of polyploids for crop breeding. While we are beginning to understand the molecular mechanisms that contribute to novel variation in polyploids, there is still a lack of specific links between genes, proteins, and phenotypes. We also do not understand the interplay between these different mechanisms and the important drivers of phenotypic diversity. What role do genomic alterations or allelic diversity play in the origin of this novel phenotypic variation? Many homeologous genes are nonadditively expressed in neoallopolyploids, but what is the source of this differential expression? If nonadditive gene expression is a result of genome dominance, than how is such a mechanism established? The answers to these questions are likely to emerge in the coming years. 460 UN Editor Proof 70 Layout: T1 Standard SC Chapter No.: 4 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 71/75 71 EC TE D PR OO F Chang P, Dilkes B, McMahon M, Comai L, Nuzhdin S (2010) Homoeolog-specific retention and use in allotetraploid Arabidopsis suecica depends on parent of origin and network partners. Genome Biol 11(12):R125 Chaudhary B, Flagel L, Stupar R, Udall J, Verma N, Springer N, Wendel J (2009) Reciprocal silencing, transcriptional bias and functional divergence of homeologs in polyploid cotton (Gossypium). Genetics 182:503–517 Chen M, Ha M, Lackey E, Wang J, Chen Z (2008) RNAi of met1 reduces DNA methylation and induces genome-specific changes in gene expression and centromeric small RNA accumulation in Arabidopsis Allopolyploids. Genetics 178:1845–1858 Chester M, Gallagher JP, Symonds VV, Cruz da Silva AV, Mavrodiev EV, Leitch AR, Soltis PS, Soltis DE (2012) Extensive chromosomal variation in a recently formed natural allopolyploid species, Tragopogon miscellus (Asteraceae). Proc Nat Acad Sci 109(4):1176–1181. doi: 10.1073/pnas.1112041109 Church SA, Spaulding EJ (2009) Gene expression in a wild Autopolyploid sunflower series. J Hered 100(4):491–495. doi: 10.1093/jhered/esp008 Clausen J, Keck D, Hiesey W (1945) Experimental studies on the nature of species. II. Plant evolution through Amphiploidy and autoploidy, with examples from the Madiinae. Carnegie Inst Wash Publ 564 Comai L, Tyagi AP, Winter K, Holmes-Davis R, Reynolds SH, Stevens Y, Byers B (2000) Phenotypic instability and rapid Gene silencing in newly formed Arabidopsis Allotetraploids. Plant Cell 12(9):1551–1568. doi: 10.1105/tpc.12.9.1551 Darlington CD (1963) Chromosome botany, and the origins of cultivated plants. Allen and Unwin, London Dehal P, Boore JL (2005) Two rounds of whole genome duplication in the Ancestral Vertebrate. PLoS Biol 3(10):e314 Doyle JJ, Flagel LE, Paterson AH, Rapp RA, Soltis DE, Soltis PS, Wendel JF (2008) Evolutionary genetics of genome merger and doubling in plants. Annu Rev Genet 42(1):443–461. doi: 10.1146/annurev.genet.42.110807.091524 Dubcovsky J, Dvorak J (2007) Genome plasticity a key factor in the success of polyploid wheat under domestication. Science 316(5833):1862–1866. doi: 10.1126/science.1143986 Durán-Figueroa N, Vielle-Calzada J-P (2010) ARGONAUTE9-dependent silencing of transposable elements in pericentromeric regions of Arabidopsis. Plant Signal Behav 5(11):1476–1479 Earley K, Lawrence RJ, Pontes O, Reuther R, Enciso AJ, Silva M, Neves N, Gross M, Viegas W, Pikaard CS (2006) Erasure of histone acetylation by Arabidopsis HDA6 mediates large-scale gene silencing in nucleolar dominance. Genes Dev 20(10):1283–1293. doi: 10.1101/ gad.1417706 Eigsti OJ (1957) Induced Polyploidy. Am J Bot 44(3):272–279 Feldman M, Levy AA (2005) Allopolyploidy—a shaping force in the evolution of wheat genomes. Cytogenet Genome Res 109(1–3):250–258 Finigan P, Martienssen RA (2008) Nucleolar dominance and DNA methylation directed by small interfering RNA. Mol Cell 32(6):753–754 Flagel L, Udall J, Nettleton D, Wendel J (2008) Duplicate gene expression in allopolyploid Gossypium reveals two temporally distinct phases of expression evolution. BMC Biol 6:16 Flagel LE, Wendel JF (2010) Evolutionary rate variation, genomic dominance and duplicate gene expression evolution during allotetraploid cotton speciation. New Phytol 186(1):184–193. doi: 10.1111/j.1469-8137.2009.03107.x Gaeta RT, Pires JC, Iniguez-Luy F, Leon E, Osborn TC (2007) Genomic changes in Resynthesized Brassica napus and their effect on Gene expression and Phenotype. Plant Cell Online 19(11):3403–3417. doi: 10.1105/tpc.107.054346 Gaut BS, Doebley JF (1997) DNA sequence evidence for the segmental allotetraploid origin of maize. Proc Nat Acad Sci U S A 94(13):6809–6814 Grant V (1981) Plant speciation, 2nd edn. Columbia University Press, New York Guo M, Birchler JA (1994) Trans-acting dosage effects on the expression of model Gene systems in Maize Aneuploids. Science 266(5193):1999–2002. doi: 10.1126/science.266.5193.1999 CO RR 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 UN Editor Proof 4 Origins of Novel Phenotypic Variation in Polyploids Layout: T1 Standard SC Chapter No.: 4 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 72/75 EC TE D PR OO F Guo M, Davis D, Birchler JA (1996) Dosage effects on Gene expression in a Maize Ploidy series. Genetics 142(4):1349–1355 Ha M, Lu J, Tian L, Ramachandran V, Kasschau KD, Chapman EJ, Carrington JC, Chen X, Wang X-J, Chen ZJ (2009) Small RNAs serve as a genetic buffer against genomic shock in Arabidopsis interspecific hybrids and allopolyploids. Proc Nat Acad Sci 106(42):17835–17840. doi: 10.1073/pnas.0907003106 Harlan J, deWet J (1975) On Ö. Winge and a prayer: the origins of polyploidy. Bot Rev 41 (4):361–390. doi: 10.1007/bf02860830 Hegarty MJ, Barker GL, Wilson ID, Abbott RJ, Edwards KJ, Hiscock SJ (2006) Transcriptome shock after interspecific hybridization in senecio is ameliorated by genome duplication. Curr Biol 16(16):1652–1659. doi: 10.1016/j.cub.2006.06.071 Hegarty MJ, Hiscock SJ (2008) Genomic clues to the evolutionary success of polyploid plants. Curr Biol: CB 18(10):R435–R444 Henry IM, Dilkes BP, Young K, Watson B, Wu H, Comai L (2005) Aneuploidy and genetic variation in the Arabidopsis thaliana Triploid response. Genetics 170(4):1979–1988. doi: 10.1534/genetics.104.037788 Jiao Y, Wickett NJ, Ayyampalayam S, Chanderbali AS, Landherr L, Ralph PE, Tomsho LP, Hu Y, Liang H, Soltis PS, Soltis DE, Clifton SW, Schlarbaum SE, Schuster SC, Ma H, LeebensMack J, dePamphilis CW (2011) Ancestral polyploidy in seed plants and angiosperms. Nature 473 (7345):97–100. doi: http://www.nature.com/nature/journal/v473/n7345/abs/10.1038nature09916-unlocked.html#supplementary-information Jørgensen C (1928) The experimental formation of heteroploid plants in the genus Solanum. J Genet 11:133 Josefsson C, Dilkes B, Comai L (2006) Parent-dependent loss of Gene silencing during interspecies hybridization. Curr Biol 16(13):1322–1328. doi: 10.1016/j.cub.2006.05.045 Kashkush K, Feldman M, Levy A (2002) Gene loss, silencing and activation in a newly synthesized wheat Allotetraploid. Genetics 160:1651–1659 Kellis M, Birren BW, Lander ES (2004) Proof and evolutionary analysis of ancient genome duplication in the yeast Saccharomyces Cerevisiae. Nature 428 (6983):617–624. doi: http:// www.nature.com/nature/journal/v428/n6983/suppinfo/nature02424_S1.html Kenan-Eichler M, Leshkowitz D, Tal L, Noor E, Melamed-Bessudo C, Feldman M, Levy AA (2011) Wheat hybridization and Polyploidization results in deregulation of small RNAs. Genetics 188(2):263–272. doi: 10.1534/genetics.111.128348 Kerber E (1964) Wheat: reconstitution of the tetraploid component (AABB) of hexaploids. Science 143:53–255 Kihara H, Ono T (1926) Chromosomenzahlen und systematische Gruppierung der Rumex-Arten. Z Zellforsch Mikr Anat 4:475 Kliebenstein DJ, West MAL, van Leeuwen H, Kim K, Doerge RW, Michelmore RW, St. Clair DA (2006) Genomic survey of Gene expression diversity in Arabidopsis thaliana. Genetics 172(2):1179–1189. doi: 101534/genetics.105.049353 Lagercrantz U, Lydiate DJ (1996) Comparative genome mapping in Brassica. Genetics 144(4):1903–1910 Leitch AR, Leitch IJ (2008) Genomic plasticity and the diversity of polyploid plants. Science 320(5875):481–483. doi: 10.1126/science.1153585 Levin DA (1983) Polyploidy and novelty in flowering plants. Am Nat 122(1):1–25 Levin DA (2002) The role of chromosomal change in plant evolution. Oxford University Press, Oxford Lexer C, Welch ME, Raymond O, Rieseberg LH (2003) The origin of ecological divergence in Helianthus paradoxus (Asteraceae): selection on transgressive characters in a novel hybrid habitat. Evolution 57(9):1989–2000. doi: 10.1111/j.0014-3820.2003.tb00379.x Li X, Guo W, Wang B, Li X, Chen H, Wei L, Wang Y, Wu J, Long H (2010) Instability of chromosome number and DNA methylation variation induced by hybridization and amphidiploid formation between Raphanus sativus L. and Brassica Alboglabra Bailey. BMC Plant Biol 10(1):207 CO RR 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 P. Finigan et al. UN Editor Proof 72 Layout: T1 Standard SC Chapter No.: 4 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 73/75 73 EC TE D PR OO F Liu B, Wendel JF (2003) Epigenetic phenomena and the evolution of plant allopolyploids. Mol Phylogenet Evol 29(3):365–379. doi: 10.1016/s1055-7903(03)00213-6 Lukens LN, Pires JC, Leon E, Vogelzang R, Oslach L, Osborn T (2006) Patterns of sequence loss and cytosine methylation within a population of newly Resynthesized Brassica napus allopolyploids. Plant Physiol 140(1):336–348. doi: 10.1104/pp.105.066308 Lumaret R (1988) Adaptive strategies and ploidy levels. Acta Oecol Oecol Plant 9:83 Madlung A, Masuelli RW, Watson B, Reynolds SH, Davison J, Comai L (2002) Remodeling of DNA Methylation and Phenotypic and transcriptional changes in synthetic Arabidopsis Allotetraploids. Plant Physiol 129(2):733–746. doi: 10.1104/pp.003095 Maere S, De Bodt S, Raes J, Casneuf T, Van Montagu M, Kuiper M, Van de Peer Y (2005) Modeling gene and genome duplications in Eukaryotes. Proc Nat Acad Sci U S A 102(15):5454–5459. doi: 10.1073/pnas.0501102102 Martienssen RA (2010) Heterochromatin, small RNA and post-fertilization dysgenesis in allopolyploid and interploid hybrids of Arabidopsis. New Phytol 186(1):46–53. doi: 10.1111/ j.1469-8137.2010.03193.x Matzke MA, Scheid OM, Matzke AJM (1999) Rapid structural and epigenetic changes in polyploid and aneuploid genomes. BioEssays 21(9):761–767. doi: 10.1002/(sici)15211878(199909)21:9\761:aid-bies7[3.0.co;2-c McStay B (2006) Nucleolar dominance: a model for rRNA gene silencing. Genes Dev 20(10):1207–1214. doi: 10.1101/gad.1436906 Meins F, Thomas M (2003) Meiotic transmission of epigenetic changes in the cell-division factor requirement of plant cells. Development 130(25):6201–6208. doi: 10.1242/dev.00856 Meyer A, Schartl M (1999) Gene and genome duplications in vertebrates: the one-to-four (-toeight in fish) rule and the evolution of novel gene functions. Curr Opin Cell Biol 11(6):699–704. doi: 10.1016/s0955-0674(99)00039-3 Meyer A, Van de Peer Y (2005) From 2R to 3R: evidence for a fish-specific genome duplication (FSGD). BioEssays 27(9):937–945. doi: 10.1002/bies.20293 Mohan Jain S (2001) Tissue culture-derived variation in crop improvement. Euphytica 118(2):153–166. doi: 10.1023/a:1004124519479 Müntzing A (1936) The evolutionary significance of Autopolyploidy. Hereditas 21(2–3):363–378. doi: 10.1111/j.1601-5223.1936.tb03204.x Müntzing A (1937) The effects of chromosomal variation in Dactylis. Hereditas 23:113 Newton W, Pellew C (1929) Primula kewensis and its derivatives. J Genet 20:405 Ng DWK, Zhang C, Miller M, Shen Z, Briggs SP, Chen ZJ (2011) Proteomic divergence in Arabidopsis autopolyploids and allopolyploids and their progenitors. Heredity. doi: http:// www.nature.com/hdy/journal/vaop/ncurrent/suppinfo/hdy201192s1.html Ni Z, Kim E-D, Ha M, Lackey E, Liu J, Zhang Y, Sun Q, Chen ZJ (2009) Altered circadian rhythms regulate growth vigour in hybrids and Allopolyploids. Nature 457 (7227):327–331. doi: http://www.nature.com/nature/journal/v457/n7227/suppinfo/nature07523_S1.html Noggle G (1946) The physiology of polyploidy in plants. Lloydia 9:153 Ohno S (1970) Evolutoin by gene duplication. Springer, Berlin Olmedo-Monfil V, Duran-Figueroa N, Arteaga-Vazquez M, Demesa-Arevalo E, Autran D, Grimanelli D, Slotkin RK, Martienssen RA, Vielle-Calzada J-P (2010) Control of female gamete formation by a small RNA pathway in Arabidopsis. Nature 464 (7288):628–632. doi: http://www.nature.com/nature/journal/v464/n7288/suppinfo/nature08828_S1.html Osborn T, Pires J, Birchler J, Auger D, Chen Z, Lee H, Comai L, Madlung A, Doerge R, Colot V, Martienssen R (2003) Understanding mechanisms of novel gene expression in polyploids. Trends Genet 19:141–147 Parisod C, Salmon A, Zerjal T, Tenaillon M, Grandbastien M-A, Ainouche M (2009) Rapid structural and epigenetic reorganization near transposable elements in hybrid and allopolyploid genomes in Spartina. New Phytol 184(4):1003–1015. doi: 10.1111/j.1469-8137.2009.03029.x Phillips RL, Kaeppler SM, Olhoft P (1994) Genetic instability of plant tissue cultures: breakdown of normal controls. Proc Nat Acad Sci U S A 91(12):5222–5226 CO RR 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 UN Editor Proof 4 Origins of Novel Phenotypic Variation in Polyploids Layout: T1 Standard SC Chapter No.: 4 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 74/75 EC TE D PR OO F Pignatta D, Dilkes BP, Yoo S-Y, Henry IM, Madlung A, Doerge RW, Jeffrey Chen Z, Comai L (2010) Differential sensitivity of the Arabidopsis thaliana transcriptome and enhancers to the effects of genome doubling. New Phytol 186(1):194–206. doi: 10.1111/j.1469-8137.2010. 03198.x Pires JC, Zhao J, Schranz ME, Leon EJ, Quijada PA, Lukens LN, Osborn TC (2004) Flowering time divergence and genomic rearrangements in resynthesized Brassica polyploids (Brassicaceae). Biol J Linn Soc 82(4):675–688. doi: 10.1111/j.1095-8312.2004.00350.x Pontes O, Neves N, Silva M, Lewis MS, Madlung A, Comai L, Viegas W, Pikaard CS (2004) Chromosomal locus rearrangements are a rapid response to formation of the allotetraploid Arabidopsis suecica genome. Proc Nat Acad Sci U S A 101(52):18240–18245. doi: 10.1073/ pnas.0407258102 Preuss S, Pikaard CS (2007) rRNA gene silencing and nucleolar dominance: insights into a chromosome-scale epigenetic on/off switch. Biochimica et Biophysica Acta (BBA)—Gene structure and expression 1769 (5–6):383–392. doi: 10.1016/j.bbaexp.2007.02.005 Preuss SB, Costa-Nunes P, Tucker S, Pontes O, Lawrence RJ, Mosher R, Kasschau KD, Carrington JC, Baulcombe DC, Viegas W, Pikaard CS (2008) Multimegabase silencing in nucleolar dominance involves siRNA-directed DNA methylation and specific methylcytosinebinding proteins. Mol Cell 32(5):673–684. doi: 10.1016/j.molcel.2008.11.009 Pumphrey M, Bai J, Laudencia-Chingcuanco D, Anderson O, Gill B (2009) Nonadditive expression of homoeologous genes is established upon polyploidization in hexaploid wheat. Genetics 181:1147–1157 Ramsey J, Schemske DW (1998) Pathways, mechanisms, and rates of polyploid formation in flowering plants. Annu Rev Ecol Syst 29(1):467–501. doi: 10.1146/annurev.ecolsys.29.1.467 Ramsey J, Schemske DW (2002) Neopolyploidy in flowering plants. Annu Rev Ecol Syst 33(1):589–639. doi: 10.1146/annurev.ecolsys.33.010802.150437 Randolph L (1941) An evaluation of induced polyploidy as a method of breeding crop plants. Am Nat 75:347 Rapp R, Udall J, Wendel J (2009) Genomic expression dominance in allopolyploids. BMC Biol 7:18 Riddle N, Jiang H, An L, Doerge R, Birchler J (2010) Gene expression analysis at the intersection of ploidy and hybridity in maize. Theor Appl Genet 120(2):341–353. doi: 10.1007/s00122009-1113-3 Santos JL, Alfaro D, Sanchez-Moran E, Armstrong SJ, Franklin FCH, Jones GH (2003) Partial diploidization of meiosis in Autotetraploid Arabidopsis thaliana. Genetics 165(3):1533–1540 Schnable JC, Springer NM, Freeling M (2011) Differentiation of the maize subgenomes by genome dominance and both ancient and ongoing gene loss. Proc Nat Acad Sci 108(10):4069–4074. doi: 10.1073/pnas.1101368108 Shindo C, Aranzana MJ, Lister C, Baxter C, Nicholls C, Nordborg M, Dean C (2005) Role of FRIGIDA and flowering locus c in determining variation in flowering time of Arabidopsis. Plant Physiol 138(2):1163–1173. doi: 10.1104/pp.105.061309 Shoemaker RC, Polzin K, Labate J, Specht J, Brummer EC, Olson T, Young N, Concibido V, Wilcox J, Tamulonis JP, Kochert G, Boerma HR (1996) Genome duplication in soybean (Glycine subgenus soja). Genetics 144(1):329–338 Simpson GG, Dean C (2002) Arabidopsis, the rosetta stone of flowering time? Science 296(5566):285–289. doi: 10.1126/science.296.5566.285 Singh M, Goel S, Meeley RB, Dantec C, Parrinello H, Michaud C, Leblanc O, Grimanelli D (2011) Production of viable gametes without Meiosis in Maize deficient for an ARGONAUTE protein. Plant Cell Online 23(2):443–458. doi: 10.1105/tpc.110.079020 Skalinska M (1946) Polyploidy in valeriana officinalis Linn. In relation to its ecology and distribution. J Linn Soc London, Bot 53:159 Slotkin RK, Vaughn M, Borges F, Tanurdzic M, Becker JD, Feijo JA, Martienssen RA (2009) Epigenetic reprogramming and small RNA silencing of transposable elements in pollen. Cell 136 (3):461–472. doi: S0092-8674(08)01644-9 [pii] 10.1016/j.cell.2008.12.038 CO RR 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 P. Finigan et al. UN Editor Proof 74 Layout: T1 Standard SC Chapter No.: 4 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 75/75 75 EC TE D PR OO F Soltis DE, Albert VA, Leebens-Mack J, Bell CD, Paterson AH, Zheng C, Sankoff D, dePamphilis CW, Wall PK, Soltis PS (2009) Polyploidy and angiosperm diversification. Am J Bot 96(1):336–348. doi: 10.3732/ajb.0800079 Soltis DE, Rieseberg LH (1986) Autopolyploidy in Tolmiea menziesii (Saxifragaceae): genetic insights from enzyme electrophoresis. Am J Bot 73(2):310–318 Soltis DE, Soltis PS (1999) Polyploidy: recurrent formation and genome evolution. Trends Ecol Evol 14(9):348–352. doi: 10.1016/s0169-5347(99)01638-9 Soltis DE, Soltis PS, Schemske DW, Hancock JF, Thompson JN, Husband BC, Judd WS (2007) Autopolyploidy in angiosperms: have we grossly underestimated the number of species? Taxon 56(1):13–30 Soltis PS, Soltis DE (2009) The role of hybridization in plant speciation. Annu Rev Plant Biol 60(1):561–588. doi: 10.1146/annurev.arplant.043008.092039 Song K, Lu P, Tang K, Osborn TC (1995) Rapid genome change in synthetic polyploids of Brassica and its implications for polyploid evolution. Proc Nat Acad Sci 92(17):7719–7723 Song K, Tang K, Osborn T (1993) Development of synthetic Brassica amphidiploids by reciprocal hybridization and comparison to natural amphidiploids. Theor Appl Genet 86:811 Stebbins G (1947) Types of polyploids: their classification and significance. Adv Genet 1:403 Stebbins GL (1950) Variation and evolution in plants. Columbia University Press, New York Stebbins GL (1971) Chromosomal evolution in higher plants [by] G. Ledyard Stebbins. Contemporary biology, vol Accessed from http://nla.gov.au/nla.cat-vn1859678. Edward Arnold, London Storchova Z, Breneman A, Cande J, Dunn J, Burbank K, O’Toole E, Pellman D (2006) Genomewide genetic analysis of polyploidy in yeast. Nature 443 (7111):541–547. doi: http:// www.nature.com/nature/journal/v443/n7111/suppinfo/nature05178_S1.html Stupar RM, Bhaskar PB, Yandell BS, Rensink WA, Hart AL, Ouyang S, Veilleux RE, Busse JS, Erhardt RJ, Buell CR, Jiang J (2007) Phenotypic and Transcriptomic changes associated with Potato Autopolyploidization. Genetics 176(4):2055–2067. doi: 10.1534/genetics.107.074286 Tanurdzic M, Vaughn MW, Jiang H, Lee T-J, Slotkin RK, Sosinski B, Thompson WF, Doerge RW, Martienssen RA (2008) Epigenomic consequences of immortalized plant cell suspension culture. PLoS Biol 6(12):e302. doi: 10.1371/journal.pbio.0060302 Tate JA, Symonds VV, Doust AN, Buggs RJA, Mavrodiev E, Koh J, Soltis PS, Soltis DE (2009) Synthetic polyploids of Tragopogon miscellus and T. mirus (Asteraceae): 50 ? years after Ownbey’s discovery. Am J Bot 96:979–988 Tate JA, Ni Z, Scheen A-C, Koh J, Gilbert CA, Lefkowitz D, Chen ZJ, Soltis PS, Soltis DE (2006) Evolution and expression of Homeologous Loci in Tragopogon miscellus (Asteraceae), a recent and reciprocally formed allopolyploid. Genetics 173(3):1599–1611. doi: 10.1534/genetics.106.057646 Valvekens D, Montagu MV, Lijsebettens MV (1988) Agrobacterium tumefaciens-mediated transformation of Arabidopsis thaliana root explants by using kanamycin selection. Proc Nat Acad Sci U S A 85(15):5536–5540 Van de Peer Y, Maere S, Meyer A (2009) The evolutionary significance of ancient genome duplications. Nat Rev Genet 10(10):725–732 Vaughn MW, Tanurdzic M, Lippman Z, Jiang H, Carrasquillo R, Rabinowicz PD, Dedhia N, McCombie WR, Agier N, Bulski A, Colot V, Doerge RW, Martienssen RA (2007) Epigenetic natural variation in Arabidopsis thaliana. PLoS Biol 5 (7):e174. doi: 06-PLBI-RA-2115 [pii] 10.1371/journal.pbio.0050174 Walia H, Josefsson C, Dilkes B, Kirkbride R, Harada J, Comai L (2009) Dosage-dependent deregulation of an AGAMOUS-LIKE gene cluster contributes to interspecific incompatibility. Curr Biol: CB 19(13):1128–1132 Wang H, Chai Y, Chu X, Zhao Y, Wu Y, Zhao J, Ngezahayo F, Xu C, Liu B (2009) Molecular characterization of a rice mutator-phenotype derived from an incompatible cross-pollination reveals transgenerational mobilization of multiple transposable elements and extensive epigenetic instability. BMC Plant Biol 9(1):63 CO RR 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768 769 770 771 772 UN Editor Proof 4 Origins of Novel Phenotypic Variation in Polyploids Layout: T1 Standard SC Chapter No.: 4 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 76/75 EC TE D PR OO F Wang J, Tian L, Lee H-S, Chen ZJ (2006a) Nonadditive regulation of FRI and FLC Loci mediates flowering-time variation in Arabidopsis Allopolyploids. Genetics 173(2):965–974. doi: 10.1534/genetics.106.056580 Wang J, Tian L, Lee H-S, Wei NE, Jiang H, Watson B, Madlung A, Osborn TC, Doerge RW, Comai L, Chen ZJ (2006b) Genomewide nonadditive Gene regulation in Arabidopsis Allotetraploids. Genetics 172(1):507–517. doi: 10.1534/genetics.105.047894 Wang J, Tian L, Madlung A, Lee H-S, Chen M, Lee JJ, Watson B, Kagochi T, Comai L, Chen ZJ (2004) Stochastic and Epigenetic changes of Gene expression in Arabidopsis Polyploids. Genetics 167(4):1961–1973. doi: 10.1534/genetics.104.027896 Wendel JF (2000) Genome evolution in polyploids. Plant Mol Biol 42(1):225–249. doi: 10.1023/ a:1006392424384 Winge O (1917) The chromosomes: their number and general importance. C.R. Trav Lab, Carlsberg Winge Ö (1932) On the origin of constant species-hybrids. Sven Bot Tidskr 26:107 Wolffe AP, Matzke MA (1999) Epigenetics: regulation through repression. Science 286(5439):481–486. doi: 10.1126/science.286.5439.481 Xiong Z, Gaeta RT, Pires JC (2011) Homoeologous shuffling and chromosome compensation maintain genome balance in resynthesized allopolyploid Brassica napus. Proc Nat Acad Sci 108(19):7908–7913. doi: 10.1073/pnas.1014138108 Xu Y, Zhong L, Wu X, Fang X, Wang J (2009) Rapid alterations of gene expression and cytosine methylation in newly synthesized Brassica napus allopolyploids. Planta 229(3):471–483. doi: 10.1007/s00425-008-0844-8 Yaakov B, Kashkush K (2011) Massive alterations of the methylation patterns around DNA transposons in the first four generations of a newly formed wheat allohexaploid. Genome 54(1):42–49. doi: 10.1139/g10-091 Yao H, Kato A, Mooney B, Birchler J (2011) Phenotypic and gene expression analyses of a ploidy series of maize inbred Oh43. Plant Mol Biol 75(3):237–251. doi: 10.1007/s11103-0109722-4 Yu Z, Haberer G, Matthes M, Rattei T, Mayer KFX, Gierl A, Torres-Ruiz RA (2010) Impact of natural genetic variation on the transcriptome of autotetraploid Arabidopsis thaliana. Proc Nat Acad Sci 107(41):17809–17814. doi: 10.1073/pnas.1000852107 CO RR 773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792 793 794 795 796 797 798 799 800 801 802 803 P. Finigan et al. UN Editor Proof 76 Metadata of the chapter that will be visualized in SpringerLink Book Title Polyploidy and Genome Evolution Series Title Chapter Title Identifying the Phylogenetic Context of Whole-Genome Duplications in Plants Copyright Year 2012 Copyright HolderName Springer-Verlag Berlin Heidelberg Corresponding Author Family Name Burleigh Particle Given Name J. Gordon Suffix Abstract Division Department of Biology Organization University of Florida Address 32611, Gainesville, FL, USA Email gburleigh@ufl.edu Although evolutionary biologists have long recognized the transformative evolutionary potential of wholegenome duplications (WGDs) in plants, identifying the precise phylogenetic location of WGDs presents many challenges. This chapter reviews some new approaches to map WGDs on a phylogeny, the first step for understanding the large-scale evolutionary and ecological consequences of WGDs in plants. Specifically, it examines approaches for using chromosome and gene copy number data, gene trees, and other genomic insights to identify the evolutionary location of WGDs. The abundance of genomic sequence data and advances in phylogenetic methods present unprecedented opportunities to place WGDs within the plant tree of life. Still, there exist few direct tests to identify and place WGDs, and analyses of complex data are often susceptible to error. 1 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 77/91 Chapter 5 4 J. Gordon Burleigh 9 10 11 12 13 14 15 16 18 19 20 21 22 23 24 25 26 27 28 29 A central challenge in evolutionary biology is to determine the genetic mechanisms that generate species diversity as well as new traits, functions, and adaptations. Plant evolutionary biologists have long recognized the transformative evolutionary potential of polyploidy or whole-genome duplication (WGD) (e.g., Stebbins 1950; Grant 1963; Levin 2002; Soltis and Soltis 2000; Soltis et al. 2009). However, to study the evolutionary consequences of whole-genome duplications (WGDs) and link WGDs to phenotypic changes or diversification, the WGDs must first be placed in a phylogenetic context. Unfortunately, the rapid gene loss and genome rearrangements that frequently follow WGDs erase evidence of WGD (e.g., Wolfe 2001; Doyle et al. 2008). This process of diploidization brings about a paradox of the study of polyploidy; despite its apparent pervasiveness throughout the evolutionary history of land plants and its evolutionary importance, clear, unambiguous evidence of ancient WGDs can be remarkably difficult to locate within a phylogeny. CO RR 17 PR OO D 8 TE 7 Abstract Although evolutionary biologists have long recognized the transformative evolutionary potential of whole-genome duplications (WGDs) in plants, identifying the precise phylogenetic location of WGDs presents many challenges. This chapter reviews some new approaches to map WGDs on a phylogeny, the first step for understanding the large-scale evolutionary and ecological consequences of WGDs in plants. Specifically, it examines approaches for using chromosome and gene copy number data, gene trees, and other genomic insights to identify the evolutionary location of WGDs. The abundance of genomic sequence data and advances in phylogenetic methods present unprecedented opportunities to place WGDs within the plant tree of life. Still, there exist few direct tests to identify and place WGDs, and analyses of complex data are often susceptible to error. EC 5 6 F 3 Identifying the Phylogenetic Context of Whole-Genome Duplications in Plants 2 UN Editor Proof Layout: T1 Standard SC Chapter No.: 5 J. G. Burleigh (&) Department of Biology, University of Florida, Gainesville, FL 32611, USA e-mail: gburleigh@ufl.edu P. S. Soltis and D. E. Soltis (eds.), Polyploidy and Genome Evolution, DOI: 10.1007/978-3-642-31442-1_5,  Springer-Verlag Berlin Heidelberg 2012 77 Layout: T1 Standard SC Chapter No.: 5 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 78/91 J. Gordon Burleigh 48 5.1 Chromosome Evolution 32 33 34 35 36 37 38 39 40 41 42 43 44 45 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 Many early surveys of polyploidy in plants used chromosome counts to estimate the percentage of polyploid species (e.g., Grant 1963, 1982; Stebbins 1938, 1950, 1971; Goldblatt 1980). This work laid the foundation for understanding the role of WGDs in plant evolution. However, lacking a phylogenetic context, it is difficult to estimate the frequency of WGD events, let alone the evolutionary placement of WGDs simply from the percentage of polyploids (although see Levin and Wilson 1976; Otto and Whitton 2002). For example, if a plant family has 100 species, 50 of which are recent polyploids, this could be the result of as few as one WGD or as many as 50 WGDs (assuming at most one WGD per lineage). Understanding the relationships among species is necessary to infer the history of WGD. Furthermore, chromosome number alone is not necessarily indicative of polyploidy. Zea mays, with a haploid chromosome number of 10, has had multiple WGDs in the last *20 million years (Gaut and Doebley 1997; Gaut 2001), and Arabidopsis thaliana, with a haploid chromosome number of five, has experienced 3–5 WGDs since the origin of seed plants (Vision et al. 2000; Blanc et al. 2003; Bowers et al. 2003; Jiao et al. 2011). Yet few chromosome number surveys would not have considered either polyploid. Mapping chromosome numbers on a phylogenetic hypothesis can help reveal the frequency and evolutionary placement of WGD events. Informal phylogenetic observations were first used to deduce ancestral chromosome numbers. For example, numerous studies surmised a low base chromosome number for angiosperms based on the chromosome counts of the ‘‘basal’’ angiosperm lineages (e.g., EC 50 CO RR 49 TE D 46 PR OO 31 F 47 WGD in plants has been and studied for more than 100 years (e.g., Digby 1912; Winge 1917). Although sometimes cryptic, evidence of WGDs may be gleaned from such disparate sources as chromosome counts (e.g., Stebbins 1971), guard cell size (Masterson 1994), age distributions of gene copies (e.g., Blanc and Wolfe 2004), or large segmental duplications within genomes (e.g., Vision et al. 2000). Still, it is difficult to infer the phylogenetic location of the WGDs with any precision from these observations alone. To place WGDs in an evolutionary context requires a robust, and ideally well–sampled, phylogenetic hypothesis, large-scale comparative data indicating WGDs, and models and methods to map these data onto a phylogeny. Only recently have advances in phylogenetics, comparative methods, and genome sequencing made this possible on a large scale. These advances have enabled the first studies addressing some of the basic evolutionary questions about polyploidy, such as the frequency of polyploid speciation (Wood et al. 2009) or the effect of WGDs on diversification (Vamosi and Dickinson 2006; Soltis et al. 2009; Mayrose et al. 2011), in a rigorous phylogenetic framework. This chapter reviews some new approaches to map WGDs on a phylogeny, the first step for understanding the large-scale evolutionary and ecological consequences of WGDs in plants. 30 UN Editor Proof 78 Layout: T1 Standard SC Chapter No.: 5 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 79/91 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 F 76 PR OO 75 D 74 TE 73 79 Ehrendorfer et al. 1968; Stebbins 1971; Walker 1972; Raven 1975). These observations implied a history of WGD near the root of angiosperms, although they could not map this precisely; indeed, the relationships among these lineages were unknown. With the growth of phylogenetic methods and data, more formal maximum parsimony approaches were used to reconstruct ancestral chromosome number on the inner nodes of phylogenetic trees (e.g., Stace et al. 1997; Schultheis 2001; Mishima et al. 2002; Guggisberg et al. 2006; Hipp et al. 2007). To do this, chromosome number can be treated as a discrete variable, and the ancestral states can be reconstructed using linear or squared change parsimony. If the ancestral states are far higher than the base chromosome number, then that ancestral node may represent a polyploid. It is possible to construct elaborate chromosome number transition matrices for parsimony analyses, for example, allowing chromosome doubling as well as single chromosome changes, or to weight different changes, like down-weighting chromosome losses, but these analyses are rarely performed. In any case, parsimony reconstructions often have difficulty accounting for multiple transitions on a single branch or quantifying uncertainty in ancestral state reconstructions. More recently, probabilistic models of chromosome number evolution have been developed (Meyers and Levin 2006; Mayrose et al. 2010). In a simple formulation, the chromosome models allow transitions that add a chromosome, remove a chromosome, or double the number of chromosomes (Mayrose et al. 2010). Thus, although the transition matrix among chromosome states (chromosome numbers) may be extremely large, the evolutionary process can be modeled with only a few parameters. The performance of these models has not been characterized in detail; however, they appear to infer more ancient WGDs than parsimony methods (Wood et al. 2009) and may also provide quite different ancestral state reconstructions of chromosome number (Cusimano et al. 2012). In the future, these models may link chromosome evolution to diversification rates or phenotypes related to the frequency of WGDs to obtain even more accurate estimates of WGDs in a phylogeny. Studies of plant chromosome numbers have provided a wealth of insight into polyploidy in plants and have contributed substantially to canonical views of plant speciation and evolution. Yet chromosome number is a sort of summary statistic for WGD, a simple observation that is meant to represent a complex, large-scale genomic change, and chromosome number alone may not be sufficient to detect WGDs. A small chromosome number, as in Arabidopsis thaliana, does not necessarily imply the absence of historical WGDs, and high chromosome numbers are not necessarily evidence of WGDs. Without additional cytological or genetic data, it is impossible to distinguish between a WGD and increasing dysploidy, a change in the chromosome number that is not associated with a change in the amount of genetic material, based solely on chromosome counts. Also, as with any phenotype, there are limitations and biases associated with ancestral state reconstruction (e.g., Schluter et al. 1997; Ané 2008). Often reconstruction is most difficult for characters with high rates of evolution or high degrees of homoplasy. Chromosome numbers may be unstable following a WGD EC 72 CO RR 71 UN Editor Proof 5 Identifying the Phylogenetic Context of Whole-Genome Layout: T1 Standard SC Chapter No.: 5 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 80/91 J. Gordon Burleigh 133 5.2 Gene Copy Models 122 123 124 125 126 127 128 129 130 131 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 PR OO 121 D 120 TE 119 Form availability of large-scale genomic data from an increasing number of plant species, estimates of copy numbers for gene families are increasingly available for many plant species. Since WGD should change not only the chromosome numbers but also the copy numbers for all gene families, gene family copy number should provide more data to estimate a WGD than simply a single chromosome number. Like chromosome number evolution, ancestral gene copy numbers can be reconstructed using parsimony methods (Snel et al. 2002; Kunin and Ouzounis 2003; Mirkin et al. 2003; Csürös 2010; Ames et al. 2012; Librado et al. 2012). The parsimony models can be implemented in numerous ways, including weighting gains and losses differently. It is not necessarily easy to find evidence of WGDs based on the number of gene gains or losses on a branch in the species tree, but we might expect WGDs will result in far more gains, and subsequently losses, per unit time than are found on other branches. Hahn et al. (2005) developed a stochastic birth and death model that assumes a homogeneous process of duplication and loss throughout the species tree. The ML implementation of this model in CAFÉ (De Bie et al. 2006), as well as a similar Bayesian approach (Liu et al. 2011), estimates gene family gain and loss rates across the tree and can identify anomalous gene families and branches on the tree. These branches may reflect the effects of WGDs. More complex models that account for heterogeneity in the rates of duplications and losses across lineages, and in some cases also allowing gains of genes or gene families by lateral transfer, also have been proposed (Iwasaki and Takagi 2007; Csurös 2010; Ames et al. 2012; Librado et al. 2012). EC 118 CO RR 117 F 132 (see Lim et al. 2008; Chester et al. 2012) and can decrease quickly following a WGD. Thus, modeling approaches likely will have difficulty for identifying ancient WGDs. In fact, it appears that the frequency of chromosome loss and diploidization was not always appreciated in studies that only examined chromosome numbers, and even with large-scale genomic data, the mechanisms for rapid chromosome loss are not clear (Doyle et al. 2008). This lack of appreciation for the lability of chromosome numbers may have contributed to the failure to detect, or even surmise, the extent of ancient WGDs, and also may have encouraged the idea that WGDs were evolutionary dead ends (e.g., Stebbins 1950). Despite the limitations of chromosome numbers alone, data are available for many thousands of plant species, for example on the online Index to Plant Chromosome Numbers (IPCN) database (http://www.tropicos.org/Project/IPCN). Thus, until large-scale genomic sequence data sets become available for thousands of phylogenetically diverse taxa, chromosome number may provide the best opportunity to identify putative WGD events, especially recent events, with phylogenetic precision and to examine the macroevolutionary consequences of WGD throughout the history of all plants. 116 UN Editor Proof 80 Layout: T1 Standard SC Chapter No.: 5 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 81/91 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 184 185 186 187 188 189 190 191 192 193 194 195 196 F 161 PR OO 160 D 159 TE 158 In spite of much recent work on developing models of gene family copy number, all of the gene copy models assume that gene duplications or gene gains are independent. Thus, a WGD in a plant might be viewed as 20,000 gene duplications rather than a single duplication event. Consequently, while an increased duplication rate on a branch or increased loss rates on subsequent branches may suggest a WGD, there is no definitive test of WGD, and it may be difficult to distinguish WGD from simply an elevated duplication rate or a large-scale duplication. One approach may be to create a model that could estimate a rate of doubling for all gene family numbers. This transition matrix could be applied to different branches to test for either a rate of WGD [0 or different rates between clades or branches. Although gene copy number provides more detailed assessment of genomic content than chromosome numbers, inferring the histories of gene copy number and chromosome number have similar limitations. For example, gene copies appear to be rapidly silenced and lost immediately following a WGD (e.g., Tate et al. 2006; Buggs et al. 2009, 2012; see Chap. 14, this volume), which may quickly obscure the evidence for WGDs. However, simply obtaining accurate estimates of gene copy number for extant taxa may be a challenge. Without complete genome sequencing, it can be difficult to distinguish a gene loss from simply a failure to sample a gene. In fact, the lack of complete sequencing across a broad range of plant species may explain the lack of studies of gene copy number evolution in plants. Even with complete genome sequences, estimates of gene copy number depend on the vagaries of the extremely complex genome annotation and gene family circumscription problems. Furthermore, there are high levels of intraspecific variation in gene copy number in some plants (e.g., Springer et al. 2009; Debolt 2010; Zheng et al. 2011). It is possible to account for uncertainty in the gene copy numbers or incomplete sampling in a likelihood model, although such approaches have not been implemented. Still, gene copy number does not always provide direct evidence for the location of gene or WGDs. For example, take the case in Fig. 5.1, in which an outgroup has a single gene copy, and two sister taxa each have two gene copies. The parsimonious explanation for these data would be that a gene duplication preceded the most recent common ancestor of the sister taxa, although it is possible that there were independent duplications in each sister lineage. In this case, the gene topologies can provide much more insight into the history of duplication than simply looking at copy number and can easily distinguish between the two duplication scenarios in Fig. 5.1. The additional information from evolutionary history of the genes can further help identify the placement of historical duplications and WGDs. EC 157 81 CO RR 156 UN Editor Proof 5 Identifying the Phylogenetic Context of Whole-Genome 5.3 Gene Tree Reconciliation The general problem of gene tree reconciliation is based on the observation that population-level processes, such as coalescence (lineage sorting), as well as evolutionary events, such as gene duplications and loss, recombination, Layout: T1 Standard SC Chapter No.: 5 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 82/91 J. Gordon Burleigh D PR OO F Editor Proof 82 TE Fig. 5.1 Gene copy number data and corresponding gene trees. The gene copy number data mapped on a species tree implies at least one gene duplication, but do not specify the location of the duplication(s). Gene tree 1 implies a single duplication preceding the most recent common ancestor of A and B. Gene tree 2 implies independent duplications in the lineages leading to A and B 209 5.3.1 Parsimony Approaches 200 201 202 203 204 205 206 207 210 211 212 CO RR 199 UN 198 EC 208 hybridization, or lateral gene transfer can result in gene topologies that differ from the phylogeny of the species in which the genes evolve (e.g., Maddison 1997). The general challenge of gene tree reconciliation is to find the evolutionary scenario that best explains the gene tree topologies. For the case of WGD, we might ask if a collection of gene trees is consistent with a WGD event, or at least a large number of duplications, at a particular point in the species phylogeny. While this approach seeks to directly map the location of gene duplications, and thus provide more direct evidence of WGD than simply examining changes in chromosome or gene copy numbers, in practice it often is complicated by the number of different scenarios that can cause gene tree incongruence and the inherent difficulty of accurately inferring a gene tree from only a gene sequence alignment (for more detailed, general reviews see Eulenstein et al. 2010; Doyon et al. 2011b). 197 Much of the initial work in gene tree reconciliation was based on optimizing a parsimony criterion; that is, finding a mapping of a gene tree topology onto the species tree that implies the fewest evolutionary events. The gene duplication model was first Layout: T1 Standard SC Chapter No.: 5 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 83/91 5 Identifying the Phylogenetic Context of Whole-Genome 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 F PR OO EC 215 introduced by Goodman et al. (1979; also see Page 1994 and Guigó et al. 1996) to find the minimum number of gene duplications needed to explain the incongruence between a gene tree and species tree. To do this, a gene tree can be embedded into a species tree through least common ancestor mapping (lca-mapping), which maps every node in the gene tree (tips and internal nodes) to the most recent node in the species tree that could have contained the gene node (Fig. 5.2). A duplication occurred if a parent and child node in the gene tree share the same lca-mapping in the species tree. The lca-mapping identifies the most recent possible location of the gene duplication in the species tree, but this often is not the only possible location of the gene duplication. In many cases, the duplication could have preceded the lca-mapping, although the earlier placement of the duplication may imply additional gene losses. However, as in the gene copy number analyses, gene losses are often difficult to distinguish from incomplete sampling. This simple gene reconciliation can provide a direct estimate of the phylogenetic location of gene duplication events with a precision that is impossible with chromosome or gene copy number data. Also, it does not even require the presence of duplicated, paralogs genes; incongruence between a single-copy gene tree topology and the species phylogeny may be evidence of a hidden history of gene duplication and loss. On the other hand, gene duplications and losses are not the only explanation of gene tree incongruence. For example, incomplete lineage sorting or reticulation also can confound gene tree topologies. In this case, the gene duplication model may mistakenly imply a large number of duplications CO RR 214 UN 213 TE D Editor Proof Fig. 5.2 An example of lca-mapping of gene duplications. The parent and child nodes in the gene tree (h and t) map to the same node (X) in the species tree. This implies a single duplication occurring prior to or at node X in the species tree 83 Layout: T1 Standard SC Chapter No.: 5 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 84/91 241 242 243 244 245 246 247 248 249 250 251 252 253 254 255 256 257 258 259 260 261 262 263 264 265 266 267 268 269 270 271 272 273 274 275 F 240 PR OO 239 D 238 TE 237 preceding rapid cladogenesis in the species tree, where we might expect high levels of incomplete lineage sorting. Perhaps the most difficult problem underlying the gene reconciliation approaches is that, in many cases, the incongruence between a gene tree and the species phylogeny may simply be due to error. The gene tree model will interpret any topological error as evidence of duplications. Consequently, this approach often implies far more duplications rather than biologically plausible (e.g., Rasmussen and Kellis 2011). The errors in the gene tree tend to place erroneously large numbers of duplications near the root of the species tree, which may falsely suggest large-scale duplication events at the origins of major clades (Hahn 2007; Burleigh et al. 2010). In the parsimony context, several strategies have been proposed to ameliorate the problems of gene tree error. First, poorly supported clades in the gene tree may be collapsed. Several algorithmic approaches have extended the gene duplication model to deal with reconciling nonbinary trees (Berglund-Sonnhammer et al. 2006; Chang and Eulenstein 2006; Durand et al. 2006). Also, several approaches have been developed to allow minor modifications of the gene tree topology if they reduce the number of implied duplications (e.g., Chen et al. 2000; Chaudhury et al. 2011, 2012; Gorecki and Eulenstein 2012). For example, Chaudhury et al. (2012) introduced an algorithm that, given a gene tree and a species tree, finds a gene topology in a subtree pruning and regrafting (SPR) neighborhood of the original gene tree that minimizes the number of implied duplications. These local rearrangements can massively reduce the number of estimated gene duplication events. In spite of the many issues related to gene tree reconciliation, simple and informal gene tree reconciliations have been effective at helping to identify the phylogenetic location of WGDs in plants. These approaches are usually limited to small gene trees with paralogs, that is, gene trees in which at least one duplication must have occurred. In a simple three-taxon approach, a gene tree is constructed with a pair of paralogs genes from a test taxon, and the best homologs from a sister taxon and from an outgroup taxon (e.g., Bowers et al. 2003). If paralogs from the test taxon form a clade, they diverged after the common ancestor with the sister taxon; if they do not, they diverged before the most recent common ancestor. This three-taxon phylogenetic approach provides only a limited phylogenetic context for the duplications, but it has been used to determine the timing of WGDs in Arabidopsis relative to its divergence from pines, rice, and other eudicots (Bowers et al. 2003) and rice relative to its divergence from pines, Arabidopsis, and other monocots (Vandepoele et al. 2003; Chapman et al. 2004). More recently, Jiao et al. (2011) counted the gene trees that were consistent with different scenarios of WGD to infer WGDs at the root of angiosperms and seed plants. EC 236 CO RR 235 J. Gordon Burleigh UN Editor Proof 84 5.3.2 Parsimony Methods to Identify WGDs The gene duplication problem described above treats each duplication independently. Although it may identify places in the species tree with high numbers of Layout: T1 Standard SC Chapter No.: 5 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 85/91 85 306 5.3.3 Likelihood-Based Approaches 282 283 284 285 286 287 288 289 290 291 292 293 294 295 296 297 298 299 300 301 302 303 304 307 308 309 310 311 312 313 314 315 316 PR OO 281 D 280 TE 279 EC 278 CO RR 277 F 305 gene duplications, it does not attempt to find large-scale duplication events. Several proposed approaches attempt to identify the minimum number of gene duplication events, where an event may include duplications of many or all genes, rather than simply the number of duplications. One indirect approach is to examine all the possible locations of each gene duplication and find a mapping that minimizes the number of locations (nodes) on the species tree where gene duplications occur (Guigó et al. 1996; Page and Cotton 2002; Burleigh et al. 2009; Luo et al. 2009). This approach does not directly infer WGDs; but ideally, it can identify places in the species tree that are possible locations of clusters of many duplications. With a limited number of gene trees, this approach appears to be effective at identifying some locations of ancient WGDs in plants (Burleigh et al. 2009). Unfortunately, with a large number of gene trees, it is likely that all possible duplication mappings will require duplication events at every node in the species tree. In this case, every possible mapping of gene duplications will be equally optimal, and this approach will be uninformative. Another approach seeks to find a gene duplication mapping that implies the fewest gene duplication episodes (Guigó et al. 1996; Page and Cotton 2002; Bansal and Eulenstein 2008; Luo et al. 2009). Given a single gene tree and species tree, any set of gene duplications, from the same or different gene trees, that occur on the same node in a species tree can be explained by a single gene duplication episode (or event) as long as none of the gene duplications in the set have an ancestor–descendant relationship with each other (Fig. 5.3). This approach appears to help identify WGDs, which should be very large episodes, but the largest episode is simply the largest episode found on any single gene tree (Page and Cotton 2002; Burleigh et al. 2010). In practice, the mapping that minimizes the number of episodes is determined by only the largest gene trees. Furthermore, randomizing the leaf labels (taxon names) on the gene trees can result in gene tree mappings that imply fewer episodes (Burleigh et al. 2010). Thus, although the notion of finding gene tree mappings that are consistent with large-scale duplications is desirable, it is not clear that this problem has been properly formulated. 276 If the gene trees are accurate, the parsimony criterion for mapping gene duplications appears to perform well when the rates of duplication and loss are low (Åkerborg et al. 2009; Doyon et al. 2009). However, these approaches do not consider branch lengths in the gene or species trees, and they have a limited ability to allow multiple duplications and losses on a single branch. Perhaps more important, it is difficult to incorporate the parsimony criterion into a rigorous statistical framework to examine evolutionary hypotheses associated with gene duplication. Numerous likelihood-based models of gene duplication and loss for reconciling gene trees and species trees have been proposed (e.g., Arvestad et al. 2003, 2004, 2009; Åkerborg et al. 2009; Doyon et al. 2009, 2011a; Rasmussen and UN Editor Proof 5 Identifying the Phylogenetic Context of Whole-Genome Layout: T1 Standard SC Chapter No.: 5 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 86/91 J. Gordon Burleigh PR OO F Editor Proof 86 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 336 337 338 EC 319 320 Kellis 2011; Gorecki et al. 2011). Most modeling approaches are based on using birth–death processes to model duplications and losses of genes as they evolve within a species tree (e.g., Arvestad et al. 2003, 2004, 2009; Åkerborg et al. 2009; Rasmussen and Kellis 2011). Also, Doyon et al. (2009, 2011a) introduced a reconciliation algorithm that calculates the likelihood of possible reconciliations based on a constant duplication and loss rate model, and this appears to produce similar reconciliations as a parsimony approach. Gorecki et al. (2011) used a Poisson model to identify the most likely distribution of reconciliations across branches in the tree based on the species branch lengths and the gene tree topologies. Most of the likelihood model-based approaches require as input a species tree with branch lengths. The method of Gorecki et al. (2011) also requires gene tree topologies. Thus, it may be susceptible to the same problems with gene tree error as parsimony approaches. Several approaches, however, take a gene sequence alignment as input and use a Markov chain Monte Carlo (MCMC) approach to simultaneously obtain the posterior distributions of gene tree topologies and gene duplication and loss mappings (e.g., Arvestad et al. 2004, 2009; Åkerborg et al. 2009; Rasmussen and Kellis 2011). This approach provides an elegant, although computationally difficult, way to incorporate gene tree uncertainty into the gene tree reconciliation. Rasmussen and Kellis (2011) demonstrated that this approach can produce more accurate gene trees and consequently greatly reduce the number of implied duplication events compared to a parsimony approach. CO RR 318 UN 317 TE D Fig. 5.3 Examples of gene duplication episodes. Duplications in the gene trees are noted with a * followed by their location in the species tree. Both gene trees have three duplications, one at the root node and two at species node C. In gene tree 1, the two duplications at species node C can be explained by a single duplication episode (since there is not a parent–child relationship between the duplication nodes). However, in gene tree 2, the two duplication episodes must have occurred at species node C since one duplication preceded the other Layout: T1 Standard SC Chapter No.: 5 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 87/91 87 352 5.4 Other Genomic Data 347 348 349 350 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 PR OO 345 346 D 344 From availability of large-scale genomic data, evidence of cryptic ancient WGDs often comes from either identifying large, syntenic (duplicated) blocks within a single genome, or by looking at the age distribution of duplicated genes within a chromosome (see Van de Peer 2004). Since these approaches use only data from a single species and are focused more on identifying ancient WGDs than placing the WGDs in an evolutionary context, I will not cover them in detail. The presence of duplicated chromosomal segments may provide direct, unambiguous evidence of WGDs that may be difficult to obtain from simply gene copy numbers or gene trees (Vision et al. 2000). However, in practice, rapid gene losses and rearrangements after polyploidy can make it extremely difficult to detect such duplications, and different methods of detecting duplicated blocks and using different criteria for defining a syntenic block can greatly affect interpretations of the history of large-scale duplications (see Durand and Hoberman 2006). Although simply examining the genome of a single species cannot reveal the phylogenetic context of a WGD, the dates of the ancient divergences can be estimated based on the molecular divergence of paralogs. It may be possible to map the evolution of large duplicated segments on a tree, but in plants, this may require extending the taxonomic sampling of species with adequate genomic mapping data. Perhaps the greater contribution of these duplicated regions is that they can define sets of paralogs that originated from WGDs, and this information can be used to validate mappings of duplications from gene copy number or gene reconciliation analyses. The rapidly increasing abundance of large-scale transcriptome data sets for plants provides an opportunity to define WGDs based on the age distribution of duplicated genes (see Cui et al. 2006). The methods first can define pairs of most recent gene duplicates using methods such as an all-by-all BLAST. If gene duplication and loss occur at a constant rate, the frequency of duplicated genes in a genome will decrease exponentially with time. In contrast, a large-scale duplication TE 342 343 EC 341 CO RR 340 F 351 The computational complexity of these likelihood-based approaches raises concerns that they may have difficulty exploiting the magnitude of new genomic data. Yet, in many ways, they still greatly simplify the complexity of genome evolution. Rasmussen and Kellis (2012) have described the first models of both duplication and coalescence, and other modeling approaches estimate the effects of hybridization and coalescence, but not duplication and loss (e.g., Meng and Kubatko 2009; Gerard et al. 2011). These represent important steps in simultaneously accounting for the many processes that affect gene tree topologies. Still, all of the modeling approaches for duplications and losses assume that the genes and all gene duplications are independent. Thus, they may detect branches in the species tree with high rates of duplication or loss, but they do not directly assess the likelihood of WGDs. The development of such models that allow for simultaneous duplications across many genes can allow for rigorous statistical tests of the placement of WGD events. 339 UN Editor Proof 5 Identifying the Phylogenetic Context of Whole-Genome Layout: T1 Standard SC Chapter No.: 5 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 88/91 J. Gordon Burleigh 391 5.5 Conclusions 382 383 384 385 386 387 388 389 PR OO 381 F 390 event, like a WGD, should result in an overrepresentation of duplicated gene pairs at the time corresponding to the large-scale duplication event. In practice, in a plot of the age distribution, usually represented by synonymous substitution distance, of duplicated genes, peaks in the age distribution curves or evidence of multiple distributions, may indicate WGDs. Again, it is difficult to precisely place a WGD just from the pairwise divergence of duplicated sequences, but with data available from many taxa, comparisons of these age plots from related species can be informative. In some cases, analyses of the age distribution plots have failed to detect known WGDs (e.g., Blanc and Wolfe 2004; Paterson et al. 2004). However, unlike gene tree reconciliation methods, they will not be misled by incomplete lineage sorting or gene tree error. 380 407 408 409 410 411 Acknowledgments This chapter was developed and written in parts with support from the Gene Tree Reconciliation Working Group at NIMBioS through NSF award EF-0832858, with additional support from the University of Tennessee. Discussions with members of the working group including Cecile Ané, Oliver Eulenstein, Pawel Gorecki, and Brian O’Meara were helpful for this manuscript. 412 References 396 397 398 399 400 401 402 403 404 405 413 414 415 416 TE 395 EC 394 CO RR 393 D 406 New genomic sequence data and advances in phylogenetic methods presents unprecedented opportunities to place WGDs within the plant tree of life. Still, there is much work to do. Although numerous data sources and methods may provide evidence of WGDs, there exist few statistical tests of WGD. A rigorous statistical framework still must be developed to examine hypotheses about the locations of WGDs. Also, examinations of the phylogenetic placement of WGDs often are based on available data; data sets are rarely generated solely for the purpose of placing the location of WGDs. Thus, there has been little discussion about the optimal methods or optimal data sets for mapping WGDs. Indeed, this is a complex issue. For example, gene trees may allow direct observations of the patterns of gene duplication and loss, but they also are susceptible to many errors and biases that may not be problems with simpler data, such as gene copy number. Ideally, learning more about the evolutionary context and implications of WGDs in plants (e.g., their effect on diversification rates and their relationship to phenotypes such as life history or mating system) will also help to identify and place WGDs in plants. 392 UN Editor Proof 88 Åkerborg Ö, Sennlad B, Arvestad L, Lagergren J (2009) Simultaneous bayesian gene tree reconstruction and reconciliation analysis. Proc Natl Acad Sci USA 106:5714–5719 Ames RM, Money D, Ghatge VP, Whelan S, Lovell SC (2012) Determining the evolutionary history of gene families. Bioinformatics (In press) Layout: T1 Standard SC Chapter No.: 5 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 89/91 89 EC TE D PR OO F Ané C (2008) Analysis of comparative data with hierarchical autocorrelation. Ann Appl Stat 2:107–1102 Arvestad L, Berglund A-C, Lagergren J, Sennblad B (2003) Bayesian gene/species tree reconciliation and orthology analysis using MCMC. Bioinformatics 19:i7–i15 Arvestad L, Berglund A-C, Lagergren J, Sennblad B (2004) Gene tree reconstruction and orthology analysis based on an integrated model for duplications and sequence evolution. RECOMB 2004:326–335 Arvestad L, Lagergren J, Sennblad B (2009) The gene evolution model and computing its associated probabilities. J ACM 56:7 Bansal MS, Eulenstein O (2008) The multiple gene duplication problem revisited. Bioinformatics 24:i132–i138 Berglund-Sonnhammer A-C, Steffansson P, Betts MJ, Liberles DA (2006) Optimal gene-trees from sequences and species trees using a soft interpretation of parsimony. J Mol Evol 63:240–250 Blanc G, Wolfe KH (2004) Widespread paleopolyploidy in model plant species inferred from age distributions of duplicate genes. Plant Cell 16:1093–1101 Blanc G, Hokamp K, Wolfe KH (2003) A recent polyploidy superimposed on older large-scale duplications in the Arabidopsis genome. Genome Res 13:137–144 Bowers JE, Chapman BA, Rong J, Paterson AH (2003) Unravelling angiosperm genome eolution by phylogenetic analysis of chromosomal duplication events. Nature 422:433–438 Buggs RJA, Doust AN, Tate JA, Koh J, Soltis K, Feltus FA, Paterson AH, Soltis PS, Soltis DE (2009) Gene loss and silencing in Tragopogon miscellus (Asteraceae): comparison of natural and synthetic allotetraploids. Heredity 103:73–81 Buggs RJA, Chamala S, Wu W, Tate JA, Schnable PS, Soltis DE, Soltis PS, Barbazuk WB (2012) Rapid, repeated, and clustered loss of duplicated genes in allopolyploid plant populations of independent origin. Curr Biol 22:1–5 Burleigh JG, Bansal MS, Wehe A, Eulenstein O (2009) Locating large-scale gene duplication events through reconciled trees: implications for identifying ancient polyploidy in plants. J Comput Biol 16:1071–1083 Burleigh JG, Bansal M, Eulenstein O, Vision TJ (2010) Inferring species trees from gene duplication episodes. Proc BCB 2010:198–203 Chang W-C, Eulenstein O (2006) Reconciling gene trees with apparent polytomies. COCOON 2006. LNCS 4112:235–244 Chapman BA, Bowers JE, Schulze SR, Paterson AH (2004) A comparative phylogenetic approach for dating whole genome duplication events. Bioinformatics 20:180–185 Chaudhary R, Burleigh JG, Eulenstein O (2011) Algorithms for rapid error correction for the gene duplication problem (ISBRA) 2011. LNCS 6674:184-196 Chaudhary R, Burleigh JG, Eulenstein O (2012) Efficient error correction algorithms for gene tree reconciliation based on duplication, duplication and loss, and deep coalescence. BMC Bioinformatics 13:s11 Chen K, Durand D, Farach-Colton M (2000) Notung: a program for dating gene duplications and optimizing gene family trees. J Comput Biol 7:429–447 Chester M, Gallagher JP, Symonds VV, Cruz da Silva AV, Mavrodiev EV, Leitch AR, Soltis PS, Soltis DE (2012) Extensive chromosomal variation in a recently formed natural allopolyploid species, Tragapogon miscellus (Asteraceae). Proc Nat Acad Sci USA 109:1176–1181 Csurös M (2010) Count: evolutionary analysis of phylogenetic profiles with parsimony and likelihood. Bioinformatics 26:1910–1912 Cui L, Wall PK, Leebens-Mack JH, Lindsay BG, Soltis DE, Doyle JJ, Soltis PS, Carlson JE, Arumuganathan K, Barakat A, Albert VA, Ma H, de Pamphilis CW (2006) Widespread genome duplications throughout the history of flowering plants. Genome Res 16:738–749 Cusimano N, Sousa A, Renner SS (2012) Maximum likelihood inference implies a high, not a low, ancestral haploid chromosome number in araceae, with a critique of the bias introduced by ‘x’. Ann Bot 109:681-692 CO RR 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 UN Editor Proof 5 Identifying the Phylogenetic Context of Whole-Genome Layout: T1 Standard SC Chapter No.: 5 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 90/91 EC TE D PR OO F De Bie T, Cristianini N, Demuth JD, Hahn MW (2006) CAFÉ: a computational tool for the study of gene family evolution. Bioinformatics 22:1269–1271 DeBolt S (2010) Copy number variation shapes genome diversity in Arabidopsis over immediate family generational scales. Genome Biol Evol 2:441–453 Digby L (1912) The cytology of Primula kewensis and of other related Primula hybrids. Ann Bot 26:357–388 Doyle JJ, Flagel LE, Paterson AH, Rapp RA, Soltis DE, Soltis PS, Wendel JF (2008) Evolutionary genetics of genome merger and doubling in plants. Annu Rev Genet 42:443–461 Doyon J-P, Chauve C, Hamel S (2009) Space of gene/species tree reconciliations and parsimonious models. J Comput Biol 16:1399–1418 Doyon J-P, Hamel S, Chauve C (2011a) An efficient method for exploring the space of gene tree/ species tree reconciliations in a probabilistic framework. IEEE/ACM Trans. Comput Biol Bioinform 99: (In press) Doyon J-P, Ranwez V, Daubin V, Berry V (2011b) Models, algorithms and programs for phylogeny reconciliation. Briefings Bioinform 12:392–400 Durand D, Halldórsson B, Vernot B (2006) A hybrid micro-macroevolutionary approach to gene tree reconstruction. J Comput Biol 13:320–335 Durand D, Hoberman R (2006) Diagnosing duplications—can it be done? Trends Genet 22:156–164 Ehrendorfer F, Krendl F, Habeler E, Sauer W (1968) Chromosome numbers and evolution in primitive angiosperms. Taxon 17:337–468 Eulenstein O, Huzurbazar S, Liberles DA (2010) Reconciling phylogenetic trees. In: Dittmar K, Liberles D (eds) Evolution after gene duplication. Wiley, Hoboken, pp 185–206 Gaut BS (2001) Patterns of chromosomal duplication in maize and their implications for comparative maps of the grasses. Genome Res 11:55–66 Gaut BS, Doebley JF (1997) DNA sequence evidence for the segmental allotetraploid origin of maize. Proc Natl Acad Sci USA 94:6809–6814 Gerard D, Gibbs HL, Kubatko L (2011) Estimating hybridization in the presence of coalescence using phylogenetic intraspecific sampling. BMC Evol Biol 11:291 Goldblatt P (1980) Polyploidy in angiosperms: monocotyledons. In: Lewis WH (ed) Polyploidy: biological relevance. Plenum Press, New York, pp 219–239 Goodman M, Czelusniak J, Moore GW, Romero-Herrera AE, Matsuda G (1979) Fitting the gene lineage into its species lineage, a parsimony strategy illustrated by cladograms constructed by globin sequences. Syst Zool 28:132–163 Gorecki P, Eulenstein O (2012) Simultaneous error correction and rooting for gene tree reconciliation and the gene duplication problem. BMC Bioinformatics (In press) Gorecki P, Eulenstein O, Burleigh JG (2011) Maximum likelihood models and algorithms for gene tree evolution with duplications and losses. BMC Bioinform 12:S15 Grant V (1963) The origin of adaptations. Columbia University Press, New York Grant V (1982) Periodicities in the chromosome numbers of the angiosperms. Bot Gaz 143:379–389 Guggisberg A, Mansion G, Kelso S, Conti E (2006) Evolution of biogeographic patterns, ploidy levels, and breeding systems in a diploid-polyploid species complex in primula. New Phytol 171:617–632 Guigó R, Muchnik I, Smith TF (1996) Reconstruction of ancient molecular phylogeny. Mol Phylogenet Evol 6:189–213 Hahn MW (2007) Bias in phylogenetic tree reconciliation methods: implications for vertebrate genome evolution. Genome Biol 8:R141 Hahn MW, De Bie T, Stajich JE, Nguyen C, Cristianni N (2005) Estimating the tempo and mode of gene family evolution from comparative data. Genome Res 15:1153–1160 Hipp AL, Rothrock PE, Reznicek AA, Berry PE (2007) Chromosome number changes associated with speciation in sedges: a phylogenetic study in Carex section Ovales (Cyperaceae) using AFLP data. Aliso 23:193–203 CO RR 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 J. Gordon Burleigh UN Editor Proof 90 Layout: T1 Standard SC Chapter No.: 5 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 91/91 91 EC TE D PR OO F Iwasaki W, Takagi T (2007) Reconstruction of highly heterogeneous gene-content evolution across the three domains of life. Bioinformatics 23:i230–i239 Jiao Y, Wickett NJ, Ayampalayam S, Chanderbali AS, Landherr L, Ralph PE, Tomsho LP, Hu Y, Liang H, Soltis PS, Soltis DE, Clifton SW, Schlarbaum SE, Schuster SC, Ma H, LeebensMack J, de Pamphilis CW (2011) Ancestral polyploidy in seed plants and angiosperms. Nature 473:97–102 Kunin V, Ouzounis CA (2003) GeneTRACE-reconstruction of gene content of ancestral species. Bioinformatics 19:1412–1416 Levin DA (2002) The role of chromosomal change in plant evolution. Oxford University Press, New York Levin DA, Wilson AC (1976) Rates of evolution in seed plants: net increase in diversity of chromosome numbers and species numbers through time. Proc Natl Acad Sci 73:2086–2090 Librado P, Vieira FG, Rozas J (2012) BadiRate: estimating family turnover rates by likelihoodbased methods. Bioinformatics 28:279–281 Lim KY, Soltis DE, Soltis PS, Tate J, Matyasek R, Srubarova H, Kovarik A, Pires JC, Xiong Z, Leitch AR (2008) Rapid chromosome evolution in recently formed polyploids in Tragapogon (Asteraceae). PLoS ONE 3:e3353 Liu L, Yu L, Kalavacharla V, Liu Z (2011) A bayesian model for gene family evolution. BMC Bioinform 12:426 Luo CW, Chen MC, Chen YC, Yang RWL, Liu HF, Chao KM (2009) Linear-time algorithms for the multiple gene duplication problems. IEEE/ACM Trans Comput Biol Bioinform 99:5555 Maddison WP (1997) Gene trees in species trees. Syst Biol 46:523–536 Masterson J (1994) Stomatal size in fossil plants: evidence for polyploidy in majority of angiosperms. Science 264:421–424 Mayrose I, Barker MS, Otto SP (2010) Probabilistic models of chromosome evolution and the inference of polyploidy. Syst Biol 59:132–144 Mayrose I, Zhan SH, Rothfels CJ, Magnus-Ford K, Barker MS, Rieseberg LH, Otto SP (2011) Recently formed polyploidy plants diversify at lower rates. Science 333:1257 Meng C, Kubatko LS (2009) Detecting hybrid speciation in the presence of incomplete lineage sorting using gene tree incongruence: a model. Theor Popul Biol 75:35–45 Meyers LA, Levin DA (2006) On the abundance of polyploids in flowering plants. Evolution 60:1198–1206 Mirkin BG, Fenner TI, Galperin MY, Koonin EV (2003) Algorithms for computing parsimonious evolutionary scenarios for genome evolution, the last universal common ancestor and dominance of horizontal gene transfer in the evolution of prokaryotes. BMC Evol Biol 3:2 Mishima M, Ohmido N, Fukui K, Yahara T (2002) Trends in site-number change of rDNA loci during polyploidy evolution in Sanguisorba (Rosaceae). Chromosoma 110:550–558 Page RDM (1994) Maps between trees and cladistic analysis of historical associations among genes, organisms, and areas. Syst Biol 43:58–77 Page RDM, Cotton JA (2002) Vertebrate phylogenomics: reconciled trees and gene duplication. Pac Symp Biocomput, 536–547 Paterson AH, Bowers JE, Chapman BA (2004) Ancient polyploidization predating divergence of the cereals, and its consequences for comparative genomics. Proc Natl Acad Sci 101:9903–9908 Rasmussen MD, Kellis M (2011) A Bayesian approach for fast and accurate gene tree reconstruction. Mol Biol Evol 28:273–290 Rasmussen MD, Kellis M (2012) Unified modeling of gene duplication, loss and coalescence using a locus tree. Genome Res 22:755-765 Raven PH (1975) The bases of angiosperm phylogeny: cytology. Ann Mo Bot Gard 62:724–764 Schluter D, Price T, Mooers AØ, Ludwig D (1997) Likelihood of ancestor states in adaptive radiation. Evolution 41:1239–1251 Schultheis LM (2001) Systematics of Downingia (Campanulaceae) based on molecular sequence data: implications for floral and chromosome evolution. Syst Bot 26:603–621 CO RR 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 UN Editor Proof 5 Identifying the Phylogenetic Context of Whole-Genome Layout: T1 Standard SC Chapter No.: 5 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 92/91 EC TE D PR OO F Soltis DE, Albert VA, Leebens-Mack J, Bell CD, Patterson AH, Zheng C, Sankoff D, de Pamphilis CW, Wall PK, Soltis PS (2009) Polyploidy and angiosperm diversification. Am J Bot 96:336–348 Soltis PS, Soltis DE (2000) The role of genetic and genomic attributes in the success of polyploids. Proc Natl Acad Sci 97:7051–7057 Snel B, Bork P, Huynen MA (2002) Genomes in flux: the evolution of archael and proteobacterial gene content. Genome Res 12:17–25 Springer NM, Ying K, Fu Y, Ji T, Yeh C-T, Jia Y, Wu W, Richmond T, Kitzman J, Rosenbaum H, Iniguez AL, Barbazuk WB, Jeddeloh JA, Nettleton D, Schnable PS (2009) Maize inbreds exhibit high levels of copy number variation (CNV) and presence/absence (PAV) in genome content. PLoS Genet 5:e1000734 Stace HM, Chapman AR, Lemson KL, Powell JM (1997) Cytoevolution, phylogeny, and taxonomy in Epacridaceae. Ann Bot 79:283–290 Stebbins GL (1938) Cytological characteristics associated with the different growth habits in the dicotyledons. Am J Bot 25:189–198 Stebbins GL (1950) Variation and evolution in plants. Columbia University Press, New York Stebbins GL (1971) Chromosomal evolution in higher plants. Addison-Wesley, London Tate JA, Ni Z, Scheen A-C, Koh J, Gilbert CA, Lefkowitz D, Chen ZJ, Soltis PS, Soltis DE (2006) Evolution and expression of homeologous loci in Tragopogon miscellus (Asteraceae), a recent and reciprocally formed allopolyploid. Genetics 173:1599–1611 Vamosi JC, Dickinson TA (2006) Polyploidy and diversification: a phylogenetic investigation in Rosaceae. Int J Plant Sci 167:349–358 Vandepoele K, Simillion C, Vande Peer Y (2003) Evidence that rice and other cereals are ancient aneuploids. Plant Cell 15:2192–2202 Van de Peer Y (2004) Computational approaches to unveiling ancient genome duplications. Nat Rev Genet 5:752–763 Vision TJ, Brown DG, Tanksley SD (2000) The origins of genomic duplication in arabidopsis. Science 290:2114–2117 Walker JW (1972) Chromosome numbers, phylogeny, phytogeography of the Annonaceae and their bearing on the (original) basic chromosome number of angiosperms. Taxon 21:57–65 Winge Ö (1917) The chromosomes. Their numbers and general importance. Comptes Rendus des Travaux Laboratoire Carlsberg 13:131–275 Wolfe KH (2001) Yesterday’s polyploids and the mystery of diploidization. Nat Rev Genet 2:333–341 Wood TE, Takebayashi N, Barker MS, Mayrose I, Greenspoon PB, Rieseberg LH (2009) The frequency of polyploidy speciation in vascular plants. Proc Natl Acad Sci USA 106:13875–13879 Zheng L-Y, Guo X-S, He B, Sun L-J, Peng Y, Dong S-S, Liu T-F, Jiang S, Ramachandran S, Liu C-M, Jing H-C (2011) Genome-wide patterns of genetic variation in sweet and grain sorghum (Sorghum bicolor). Genome Biol 12:R114 CO RR 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 J. Gordon Burleigh UN Editor Proof 92 Metadata of the chapter that will be visualized in SpringerLink Book Title Polyploidy and Genome Evolution Series Title Chapter Title Ancient and Recent Polyploidy in Monocots Copyright Year 2012 Copyright HolderName Springer-Verlag Berlin Heidelberg Corresponding Author Family Name Paterson Particle Given Name Andrew H. Suffix Author Division Plant Genome Mapping Laboratory Organization University of Georgia Address Athens, GA, USA Email paterson@plantbio.uga.edu Family Name Wang Particle Given Name Xiyin Suffix Author Division Plant Genome Mapping Laboratory Organization University of Georgia Address Athens, GA, USA Division Center for Genomics and Computational Biology Organization Hebei United University Address Tangshan, People’s Republic of China Email wang.xiyin@gmail.com Family Name Li Particle Given Name Jingping Suffix Author Division Plant Genome Mapping Laboratory Organization University of Georgia Address Athens, GA, USA Email jingpingli@gmail.com Family Name Tang Particle Given Name Haibao Suffix Division Plant Genome Mapping Laboratory Organization University of Georgia Address Athens, GA, USA Division Abstract Organization J. Craig Venter Institute Address Rockville, MD, USA Email tanghaibao@gmail.com At least two whole-genome duplications (WGD) have profoundly influenced the evolution of most, if not all, grass (Poaceae) genomes, with the most recent of these predating the divergence of these lineages by 20 million or more years. Taxa within each major lineage of Poaceae (e.g., Panicoideae, Ehrhartoideae, Pooideae) have independently experienced additional polyploidizations that have been of central importance to the evolution and productivity of some of our most important crop plants [for example, sugarcane (Saccharum spp.), and durum and bread wheat (Triticum spp.)]. Following polyploidy, adaptation to the duplicated state is evident at the levels of transmission genetics, chromosome structure, and gene repertoire. While most duplicated chromosomal regions re-establish largely independent evolution within a few million years, 70million-year-old duplicated chromosome segments in one unusual region of the rice genome and its orthologs in other grasses have continued to exhibit concerted evolution more recently than the divergence of rice subspecies japonica and indica an estimated 400,000 years ago. 1 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 93/107 Chapter 6 4 Andrew H. Paterson, Xiyin Wang, Jingping Li and Haibao Tang 11 12 13 14 15 16 17 19 18 PR OO D 9 10 TE 8 EC 7 Abstract At least two whole-genome duplications (WGD) have profoundly influenced the evolution of most, if not all, grass (Poaceae) genomes, with the most recent of these predating the divergence of these lineages by 20 million or more years. Taxa within each major lineage of Poaceae (e.g., Panicoideae, Ehrhartoideae, Pooideae) have independently experienced additional polyploidizations that have been of central importance to the evolution and productivity of some of our most important crop plants [for example, sugarcane (Saccharum spp.), and durum and bread wheat (Triticum spp.)]. Following polyploidy, adaptation to the duplicated state is evident at the levels of transmission genetics, chromosome structure, and gene repertoire. While most duplicated chromosomal regions re-establish largely independent evolution within a few million years, 70-million-year-old duplicated chromosome segments in one unusual region of the rice genome and its orthologs in other grasses have continued to exhibit concerted evolution more recently than the divergence of rice subspecies japonica and indica an estimated 400,000 years ago. CO RR 5 6 F 3 Ancient and Recent Polyploidy in Monocots 2 A. H. Paterson (&)  X. Wang  J. Li  H. Tang Plant Genome Mapping Laboratory, University of Georgia, Athens, GA, USA e-mail: paterson@plantbio.uga.edu X. Wang e-mail: wang.xiyin@gmail.com J. Li e-mail: jingpingli@gmail.com H. Tang e-mail: tanghaibao@gmail.com X. Wang Center for Genomics and Computational Biology, Hebei United University, Tangshan, People’s Republic of China UN Editor Proof Layout: T1 Standard SC Chapter No.: 6 H. Tang J. Craig Venter Institute, Rockville, MD, USA P. S. Soltis and D. E. Soltis (eds.), Polyploidy and Genome Evolution, DOI: 10.1007/978-3-642-31442-1_6,  Springer-Verlag Berlin Heidelberg 2012 93 Layout: T1 Standard SC Chapter No.: 6 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 94/107 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 F 25 26 PR OO 24 D 23 Monocotyledons, also known as monocots, are one of the major clades of angiosperms. Most recent phylogenetic analyses (based largely on plastid sequence data) reveal a successive series of basal angiosperm lineages, with monocots often sister to Ceratophyllum plus eudicots (e.g., Soltis et al. 2011). According to the IUCN there are 59,300 species of monocots (http:// cmsdata.iucn.org/). The most species-rich family in this clade (and indeed, one of the largest of all angiosperms) are the orchids (family Orchidaceae), with more than 20,000 species (Raven et al. 2005). Among the best-studied monocots are the grasses, family Poaceae (Gramineae), which provide much of the world’s food and plant biomass and include economically important grains, such as rice (Oryza), wheat (Triticum), maize (Zea), barley (Hordeum), and sorghum (Sorghum), turf and forage/pasture grasses, sugar cane, and the bamboos. Other economically important monocot families are the palms (Arecaceae), bananas (Musaceae), gingers (Zingiberaceae), and the onion family Alliaceae. As of this writing, the vast majority of monocot genetic and genomic information, including all whole-genome-scale DNA sequences of sufficient contiguity for synteny analysis, are for members of Poaceae, necessarily constraining the focus of this chapter. However, this constraint is expected to be relieved in the very near future, with genome sequences in progress for several non-grass monocots (e.g. Phoenix dactylifera, Musa acuminata, Ananas comosus, Asparagus officinalis, Phalaenopsis equestris, Zostera marina), adding important new dimensions to understanding of monocot gene and genome evolution. A draft of the date palm (P. dactylifera) genome is publicly available, but is presently of too low contiguity for robust synteny analysis. High-quality sequences for representatives of all three major grass clades have been published, including rice (International Rice Genome Sequencing Project 2005; Yu et al. 2005) (Ehrhartoideae), sorghum (Paterson et al. 2009b) and maize (Schnable et al. 2009) (Panicoideae), and Brachypodium (The International Brachypodium Initiative 2010) (Pooideae) (Fig. 6.1). Draft genome sequence of barley (Hordeum) and wheat (Triticum) group 1 chromosomes were recently made available (Mayer et al. 2011; Wicker et al. 2011). Initial analyses of the available genome sequences available for Poaceae have shown at least two whole-genome duplications (WGD) influencing most, if not all, grass genomes, with the most recent of these predating the divergence of these lineages by an estimated 20 million or more years. Taxa within each lineage have independently experienced additional polyploidizations and readaptation to the duplicated state (for example, sugarcane and durum and bread wheat), with the model genomes constituting a good starting point for accelerating progress in the study and improvement of many additional taxa. Although chloridoid and arundinoid grasses are explored only at the EST level to date, these data show that at least the Chloridoideae experienced an evolutionary history similar to those of the other major grass clades, including both of the two WDGs shared by other TE 22 EC 21 6.1 Monocot Comparative Genomics CO RR 20 A. H. Paterson et al. UN Editor Proof 94 Layout: T1 Standard SC Chapter No.: 6 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 95/107 95 CO RR EC TE D PR OO F Editor Proof 6 Ancient and Recent Polyploidy in Monocots 63 64 65 66 UN Fig. 6.1 Phylogenetic relationships among major lineages of monocots and estimated positions of early paleopolyploidy events (modified from Paterson et al. 2009a). Notice that some lack sufficient contiguity (Phoenix) or data (Ananas, Zostera) to infer genome duplication, and accordingly the position of sigma (r) (Tang et al. 2010) cannot be precisely determined at this time; this uncertainty is indicated by arrows Poaceae. Chloridoids also have an additional lineage-specific genome duplication (Kim et al. 2009). By developing and using comparative genomics tools, research starting from the sequenced genomes may transitively shed light on monocots that are still Layout: T1 Standard SC Chapter No.: 6 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 96/107 A. H. Paterson et al. 73 6.2 Rho (q), the Most Recent Pan-Poaceae Polyploidy 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 PR OO 74 The notion that even relatively small and structurally simple grass genomes may be paleopolyploid is of long-standing interest. Indeed, secondary associations of grass chromosomes have been known for 80 years (Lawrence 1931), and genetic mapping nearly 2 decades ago suggested duplication of scattered chromosome segments in both rice (Kishimoto et al. 1994; Nagamura et al. 1995) and sorghum (Chittenden et al. 1994). Paleopolyploidy in the grasses was demonstrated beyond a reasonable doubt with the first availability of the whole-genome sequence of rice, Oryza sativa ssp. japonica (Goff et al. 2002). A brief controversy about whether the scope of duplication was confined to specific chromosomes (Vandepoele et al. 2003) or the whole genome (Paterson et al. 2003, 2004) was soon reconciled, by analyzing the genome sequence of another rice cultivar, O. sativa ssp. indica (Yu et al. 2005), with the controversy attributed to differences in the approaches used to infer the duplicated blocks (Wang et al. 2005). This polyploidy event (q) was dated to *70 million years ago (mya) based on putatively neutral DNA substitution rates between duplicated genes, and suggested to be shared by all main lineages of grasses. The genome sequences of sorghum (Sorghum bicolor) and Brachypodium (The International Brachypodium Initiative 2010) confirmed that these taxa share the paleoduplication first discerned in rice, and that neither has been affected by additional polyploidization after the evolutionary split of their respective lineages. Moreover, Brachypodium and rice have preserved a very high level of gene collinearity (Paterson et al. 2009b; The International Brachypodium Initiative 2010), making it possible to take them as a single genetic system to perform transitive genetics research across different grasses (Freeling 2001). Only a small fraction of genes shows differential gene losses after the split of rice (1.8 %) and sorghum (3.1 %). These findings suggest that, after the 70-mya polyploidization event, the genome of the last common ancestor of grasses had already experienced most gene loss and reached a relatively stable state prior to the divergence of the extant major grass lineages about 50 mya (Paterson 2008; The International Brachypodium Initiative 2010). Some prior reports of deviations from collinearity in the grasses may be accounted for by these low rates of differential gene loss on homoeologous chromosomes. D 71 TE 70 EC 69 CO RR 68 F 72 lacking de novo genomic data (Van de Peer 2004; Wang et al. 2006; Lohithaswa et al. 2007; Tang et al. 2008b). The sequences of additional grasses and non-grass monocots such as Elaeis (Arecaceae), Musa (Musaceae), and Zostera (Zosteraceae) will clarify the functional innovation of their gene sets, further elucidating the structural and functional evolution of this ecologically and economically important plant family. 67 UN Editor Proof 96 Layout: T1 Standard SC Chapter No.: 6 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 97/107 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 F 111 PR OO 110 Rho accounted for much, but not all, of the paleo-duplication that could be discerned in modern cereal genomes (Paterson et al. 2004). Shortly after the discovery of rho, an early report (Zhang et al. 2005), later supported by two independent studies (Jaillon et al. 2007; Salse et al. 2008) hinted at the presence of additional, earlier monocot duplications. Detailed elucidation of sigma (r), an additional duplication event preceding rho in the grass lineage, utilized a ‘bottom–up’ approach, first collapsing 15,640 rice genes and 15,636 sorghum genes into 13,308 rho-nodes that computationally reverse post-rho gene loss, increasing the sensitivity of subsequent analysis. These nodes (genes) were compared among themselves, revealing collinear patterns of correspondence that involve all nine major synteny blocks resulting from the rho duplication. Some collinear patterns between pairs of rho-blocks are one-to-one, while others are higher order, suggesting that multiple events may have been identified. The eight largest synteny blocks that retained collinearity following sigma contain a total of 4,168 sigma-nodes, covering 5,747 rice genes and 5,738 sorghum genes (*20 % of the rice and sorghum transcriptomes). Further study of these sigma-duplicated regions highlights an important constraint in monocot genomics, albeit one that soon will be relieved. There is little remaining intragenomic correspondence between sigma-derived rice segments, although relationships between some of these duplicated segments can still be identified through transitive comparisons of cereal genomes to outgroups. At present, however, the only available outgroups are eudicots such as grape (Vitis). Similarities between monocot and eudicot genomes resulting from common ancestry have been obscured by many paleo-polyploidy events and numerous genome rearrangements (Liu et al. 2001; Jaillon et al. 2007). The availability of non-cereal monocot outgroup genomes would significantly increase power to study these segments, particularly outgroups with genomes that have not experienced lineage-specific duplication and associated fractionation of ancestral gene orders. The lack of an ideal (non-cereal non-reduplicated monocot) outgroup notwithstanding, better understanding of monocot paleopolyploidy has improved our ability to compare monocot and eudicot genomes, a long-standing goal (Paterson et al. 1996) that has been complicated by paleopolyploidy events. Toward this goal, we applied a hierarchical clustering approach, first dividing the chromosomes into small bins and then comparing all pairs of rice and grape bins. Duplicated segments retained in grape following the eudicot ‘gamma’ hexaploidy event (3), and homologous segments retained in rice following at least two rounds of duplication (rho and sigma), contain 38 ‘‘putative ancestral regions’’ (PAR) that collectively explain 19.1 % of all observed homolog pairs and 31.0 % of reciprocal best hits between grape and rice genes, *10-fold more than would be explicable by chance. The PARs interleave multiple grape and rice genomic regions collectively D 109 TE 108 EC 107 97 6.3 Sigma (r), and Clarifying the Genome Composition of the Last Universal Common Ancestor of Monocots CO RR 106 UN Editor Proof 6 Ancient and Recent Polyploidy in Monocots Layout: T1 Standard SC Chapter No.: 6 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 98/107 156 157 158 159 160 161 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 184 185 186 187 188 189 F 155 PR OO 153 154 D 152 TE 151 covering *70 % of each genome. By consolidating much of the redundancy in each genome, the PARs create syntenic blocks with much less ambiguity and in most cases show association between one gamma block and one sigma block. We did not find any PAR that simultaneously mapped to two different gamma or sigma blocks (Tang et al. 2010). Some ‘‘ghost duplications’’ (Van de Peer 2004) in rice that we failed to identify through intragenomic comparisons (due to reciprocal gene losses in largely complementary fashion) are much clearer in cross-species comparisons (PARs). Compared to the WGD events in grape where 94.5 % of the genome appears duplicated (Jaillon et al. 2007), the rice WGDs are more complicated and degenerate. The 38 grape–rice PARs are a qualitative advance toward a global view of monocot-eudicot synteny. Collinearity appears to be disrupted around the peri-centromeric regions in 10 of the 12 rice chromosomes, suggesting dynamic reorganization of heterochromatic portions of the rice genome that may result from massive transpositions or gene losses (Bowers et al. 2005). Our unique approach to synteny analysis provides new insight into the number(s) of WGD events experienced by modern cereal genomes. In many lineages, the existence and the numbers of WGD events have been contentious. In 22 of the 38 PARs, grape–rice collinearity is clear, with 12 PARs being threefold redundant in grape, consistent with hexaploidy (Jaillon et al. 2007). The level of redundancy in rice is less clear, ranging from as little as twofold (1 PAR) to sevenfold (3 PARs) and eightfold (5 PARs). In line with the intragenomic evidence from our bottom–up analysis, these high redundancies suggest that the rice lineage experienced more than two, perhaps three, rounds of ancient WGD that have collectively been lumped into the single event that we are presently referring to as ‘sigma’ (r). Once again, the availability of non-cereal and non-reduplicated monocot outgroup genomes would significantly improve our understanding of monocot evolutionary history. EC 150 6.4 Lineage-Specific Monocot Polyploidies and Their Consequences CO RR 149 A. H. Paterson et al. Polyploidy in monocots is, of course, not limited to ancient events, and indeed more recent events have been integral to the evolution and productivity of many major crops. A classical textbook example of a more recent polyploidization is the evolution of wheat (see also Chap. 7, this volume), with formation of a tetraploid and subsequently a hexaploid leading to independent crops, durum wheat used for pasta (Triticum turgidum) and bread wheat (T. aestivum), respectively (Feldman and Levy 2005). Indeed, the synthesis by humans of octoploid triticale (9 Triticosecale a hybrid between Triticum and Secale) in the past century illustrates that there may be further gains to be made by mimicking the natural tendency of some plant lineages to form polyploids. Polyploidy has been particularly closely associated with productivity in autopolyploids such as many forage and biomass grasses. One of the most extreme UN Editor Proof 98 Layout: T1 Standard SC Chapter No.: 6 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 99/107 197 198 199 200 201 202 203 204 205 206 207 208 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 F PR OO 196 D 194 195 TE 193 EC 192 99 cases of a successful autoploid is sugar cane (Saccharum spp.), in which cultivated forms are typically octoploid—dodecaploid (i.e., with 8–12 chromosome sets, and sometimes variations in between!). These are interspecific hybrids (and therefore not strict autopolyploids, but a combination of auto- and allopolyploidy) composed of about 85–90 % chromatin from Saccharum officinarum and 10–15 % from S. spontaneum, a wild relative. Studies using DNA markers to quantify ‘dosages’ of homologous genomic regions in sugar cane show clearly nonlinear consequences of allele dosage, with one or two copies of favorable alleles at homologous loci usually having favorable effects, but additional copies yielding diminishing or even reduced returns (Ming et al. 2001, 2002a, b). Comparative genomic studies, in particular with sorghum, show that the Saccharum lineage experienced at least one WGD since its divergence from the Sorghum lineage something less than 10 mya (Ming et al. 1998; Jannoo et al. 2007). Ploidy in Saccharum is particularly interesting, in that it has been further complicated by a high frequency of chromosome non-reduction, yielding 2n ? n progeny in S. officinarum (female) 9 S. spontaneum (male) crosses (Bremer 1923). In efforts to broaden genetic variability in ‘noble canes’ that were the foundation of sugar cane production until the early part of the twentieth century, F1 progeny of interspecific crosses between S. officinarum and S. spontaneum were found to be distinctively more robust than either parent. When S. officinarum clones were used as the female parent, progeny tended to be larger stalked, higher in sucrose levels, and generally more vigorous than when S. spontaneum clones were used as the female parent. The pollination of noble cane with S. spontaneum followed by repeated backcrosses to the noble canes has come to be called ‘nobilization’, with selected hybrid progenies referred to as ‘nobilized’ canes (Bremer 1961). A key event in the evolution of modern sugarcane cultivars was the production of the nobilized cultivar, ‘POJ2878’, of Proefstation Oost, Java, in 1921 (Jeswiet 1929). Further study of the clade that includes Saccharum may prove intriguing. This clade also includes Miscanthus, which has biomass yields that are similarly high as Saccharum, but is better adapted to temperate climates and therefore has stimulated much interest for bioenergy production in the USA, Europe, and China (Heaton et al. 2008). Miscanthus species have a basic set of n = x = 19 chromosomes, versus the x = 10 that is characteristic of many Saccharinae including Saccharum. One attractive hypothesis to explain the transition from 10 to 19 chromosomes is that Miscanthus, like Saccharum (Ming et al. 1998), may have experienced a polyploidization event in the 8–9 my since its divergence from sorghum—but unlike Saccharum, which is largely autopolyploid, Miscanthus homologs have diverged sufficiently that they no longer normally pair with one another—that is, there is now preferential pairing of chromosomes. Genome evolution may have included a chromosomal fusion to get from n = 20 to 19. If Miscanthus and Saccharum shared a genome doubling event, a possibility that we are continuing to investigate, it would be an intriguing and perhaps unprecedented case in which one lineage (Miscanthus) adapted to the duplicated state by re-establishing disomy, while a sister lineage continued to have the option of polysomy. CO RR 190 191 UN Editor Proof 6 Ancient and Recent Polyploidy in Monocots Layout: T1 Standard SC Chapter No.: 6 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 100/107 A. H. Paterson et al. 234 6.5 Adapting to the Polyploid State 235 6.5.1 Centromeric Divergence and Restoration of Disomy 255 6.5.2 Karyotype Evolution 242 243 244 245 246 247 248 249 250 251 252 253 256 257 258 259 260 261 262 263 264 265 266 267 268 269 270 271 PR OO 241 D 240 TE 239 EC 238 A polyploidy event may result in genomic instability, consequently incurring a process of diploidization characterized by widespread DNA rearrangements often accompanied by large-scale gene losses (Paterson et al. 2004; Van de Peer 2004; Wang et al. 2006; Xiong et al. 2011). These DNA rearrangements may result in chromosome number variations. Grasses range from n = 2 to 18 in their basic chromosome sets (Soderstrom et al. 1987; Hilu 2004). In the sequenced genomes, rice, sorghum, and Brachypodium have n = 12, 10, and 5 chromosomes, respectively. Although maize (Zea mays) experienced a WGD since divergence from sorghum, modern-day maize retains the same chromosome number (n = 10) as sorghum. Some lineages of Sorghum have experienced chromosome condensations even in the absence of polyploidization (Spangler et al. 1999; Spangler 2003). Comparison of grass genomes has shed light on chromosome number evolution and ancestral grass karyotypes (Salse et al. 2009; Murat et al. 2010). An ancestral karyotype of n = 5 chromosomes was inferred (Salse et al. 2009; Murat et al. 2010) before the pan-grass polyploidization, with n = 2x = 10 chromosomes after the duplication, then two fissions to result in n = 2x = 12 chromosomes in the common CO RR 237 F 254 One important challenge facing a newly formed autopolyploid may be that the presence of sets of four homologous chromosomes may tend to hinder purging of deleterious alleles. Transition to diploid inheritance would, in principle, allow more rapid allele frequency changes and reduced genetic load. Centromeric divergence may have been a mechanism by which paleopolyploid grasses restored diploid inheritance (Bowers et al. 2005). A high concentration of rice genes duplicated by ancient polyploidy falls near Ks 0.85, while rice gene pairs with Ks 0.2–0.6 tend to be located in peri/centromeric regions. This suggests that shortly after polyploidization, a substantial restructuring of centromeric regions began that lasted until about 16 mya [based on the synonymous substitution rate used (Lynch and Conery 2000)]. About 18 % of the rice genome shows highly significant concentrations of matches (p \ 1 9 10-5) in the Ks range of 0.2–0.6. The restructuring of non-syntenic regions largely involves migration of DNA between pericentromeric regions of different chromosomes. The concentrations of relatively recent single–locus duplications near the centromeres may, but does not necessarily, reflect a higher duplication rate in these regions. Alternatively, recent duplications may be preserved more frequently in pericentromeric regions, i.e., there may be more rapid loss of single-gene duplications in euchromatic regions (Bowers et al. 2005). 236 UN Editor Proof 100 Layout: T1 Standard SC Chapter No.: 6 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 101/107 101 286 6.5.3 Gene Retention and Loss 278 279 280 281 282 283 284 287 288 289 290 291 292 293 294 295 296 297 298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 PR OO 277 D 276 Among the genes duplicated in the polyploidy events shared by the grass lineage, the vast majority of pairs have lost at least one duplicated copy. The finding that only a small fraction of genes show differential gene losses after the split of rice (1.8 %) and sorghum (3.1 %) demonstrates empirically that, after the 70-mya polyploidization, the genome of the last universal common ancestor of grasses had already experienced most gene loss and reached a relatively stable state prior to the divergence of the major grass lineages about 50 mya (Paterson 2008; The International Brachypodium Initiative 2010). Gene losses have often occurred in a complementary and segmental manner, that is, with non-random patterns of retention/loss on corresponding duplicated DNA segments, in a process known as fractionation (Thomas et al. 2006). Genes may be removed by a short-DNA deletion mechanism (Woodhouse et al. 2010), and in a pair of duplicates, gene loss may be universally biased to preserve the gene that is responsible for the majority of expression (Schnable et al. 2011). For more than 90 % of the preserved duplicated genes, the two copies have the same transcriptional orientations (Wang et al. 2005), and the exceptions may be a result of local DNA inversions or differential gains/losses of new tandemly duplicated genes in the paleo-duplicated regions. In grasses, and indeed across angiosperms, we find three ‘fates’ of individual gene pairs following duplication (see also Chap. 1, this volume). Most gene functional groups show post-duplication gene preservation/loss rates that are indistinguishable from the genome-wide average. Such ‘neutral’ loss of duplicated genes presumably involves inactivating mutations opposed by very weak selection (Haldane 1933), as the fate of the vast majority of duplicated genes. Population genetic models suggest that loss of duplicated genes may happen over a few million years (Lynch and Conery 2000). Genes in some specific functional categories duplicate and reduplicate TE 275 EC 274 CO RR 273 F 285 ancestor of major cereals. However, the authors noted that an ancestral karyotype of n = 6–7 was also possible. They inferred that chromosome number variation/ reduction from the common ancestor may be attributable to non-random centric double-strand break repair events. It was suggested that centromeric/telomeric illegitimate recombination between non-homologous chromosomes led to nested chromosome fusions and synteny break points, and concluded that these break points were meiotic recombination hotspots that corresponded to high sequence turnover loci through repeat invasion. These rules seem to explain most chromosome number changes in the grass genomes sequenced so far, especially the previously observed nested chromosome fusions in Brachypodium (The International Brachypodium Initiative 2010). However, many details related to dynamics of centromeres and telomeres during the rearrangements remain unclear, and the wide range in possible ancestral karyotypes (from 12 to 24) suggests that further revision of thinking on this subject is likely. 272 UN Editor Proof 6 Ancient and Recent Polyploidy in Monocots Layout: T1 Standard SC Chapter No.: 6 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 102/107 A. H. Paterson et al. 325 6.6 The Unique Case of Rice Chromosomes 11 and 12 322 323 326 327 328 329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 353 PR OO 320 321 Rice chromosomes 11 and 12 (R11 and R12) are a striking exception among chromosomes affected by the 70-mya polyploidization (Wang et al. 2011). R11 and R12 share a *3-Mb duplicated DNA segment at the termini of their short arms, the formation of which had been dated based on synonymous substitutions to *5–7 mya (The Rice Chromosomes 11 and 12 Sequencing Consortia 2005; Wang et al. 2005; Yu et al. 2005). Remarkably, in the second grass genome sequence, sorghum, the corresponding region(s) of its orthologous chromosomes (S5 and S8, respectively) also contained such an apparently recent duplication despite having diverged from rice about 50 mya (Paterson et al. 2009b). Physical and genetic maps also suggest shared terminal segments of the corresponding chromosomes in wheat (4, 5), foxtail millet (VII, VIII), and pearl millet (linkage groups 1, 4) (Devos et al. 2000; Singh et al. 2007). It would be exceedingly unlikely for segmental duplications to happen independently at such closely corresponding locations in reproductively isolated lineages. A much more parsimonious hypothesis is that the R11/12 and S5/8 regions each resulted from the pan-grass duplication 70 mya but have an unusual evolutionary history (Paterson et al. 2009b). Detailed analysis of R11 and R12 suggested that illegitimate recombination has continued for millions of years after the 70-mya divergence of these homoeologs, indeed remaining ongoing in the past 400,000 years since divergence of subspecies japonica and indica (Wang et al. 2007). Gradual and step-by-step restrictions on recombination, starting around the time of the 70-mya polyploidization, have resulted in ‘strata’ along the chromosome pair that differ in the degree of DNA sequence similarity between homoeologous genes (Wang et al. 2011). Sequence similarity between homoeologs in the strata reflects the time(s) of recombination suppression rather than the times of their origination. Indeed, the most terminal stratum in rice (RSA) appears \0.5 my old, while two more internal strata (RSB, RSC) appear to have been restricted in their ability to recombine 9.4 and 39.1 mya, D 319 TE 317 318 EC 316 CO RR 315 F 324 (Blanc and Wolfe 2004; Seoighe and Gehring 2004; Maere et al. 2005; Chapman et al. 2006; Paterson et al. 2006; Tang et al. 2008a), and in many instances can be related to the ‘‘gene balance’’ hypothesis, that stoichiometry among members of pathways and networks is important to biological function (Birchler et al. 2005; Birchler and Veitia 2007; Veitia et al. 2008). Other specific genes and gene functional groups show more extensive loss of duplicate copies than the genome-wide average, and this loss has often been convergent following independent duplications separated by hundreds of millions of years during the evolution of grasses, Arabidopsis, yeast, and Tetraodon (pufferfish) (Paterson et al. 2006). Much greater knowledge of gene functions, particularly regarding those less-explored genes that are recurrently restored to the singleton state, may provide new insights into the ‘adaptation’ of a newly formed polyploid to the duplicated state. 313 314 UN Editor Proof 102 Layout: T1 Standard SC Chapter No.: 6 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 103/107 103 PR OO F Editor Proof 6 Ancient and Recent Polyploidy in Monocots Fig. 6.2 Gene repertoire and organization in a terminal segment of homoeologous rice and sorghum chromosomes experiencing concerted evolution. This region includes 4 rice-sorghum quartets (Q), 3 triplets [T: losses on S5 (2), S8], 13 pairs of taxon-specific genes [P: 5 sorghum, 8 rice]; 8 taxon-specific singletons [S: 6 sorghum, 2 rice], and one lineage-specific duplication (X: two terminal genes on S5 and S8 share common ancestor with single rice genes). Classification letters for a gene family appear in only one member (members connected by lines). Reprinted from Wang et al. (2011) with permission. Copyright is owned by the American Society of Plant Biology 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 D TE 357 358 EC 356 respectively. In sorghum, even the most terminal (of only two) strata dates to 13.4 mya, reflecting the parallel but independent evolution of this unusual chromosome segment in divergent lineages (Wang et al. 2011). The corresponding regions in maize and Brachypodium also show prominent homoeologous recombination. However, widespread chromosomal rearrangement, especially in maize after its lineage-specific polyploidization, makes the stratification patterns more difficult to compare than in rice and sorghum. Both intriguing and perplexing is that the distal chromosomal region with the greatest DNA similarity between surviving duplicated genes also has the highest concentration of lineage-specific gene pairs found anywhere in these genomes, and a significantly elevated gene evolutionary rate (Wang et al. 2011). Of 33 and 23 riceand sorghum-specific gene pairs on these chromosomes, a respective 100 and 90 % of them are in the young strata. Both members of a remarkable 50 % of the 16 duplicated RSA gene pairs are absent from sorghum, and 15 (38 %) of 39 SSA pairs are absent from rice (Fig. 6.2). Gene losses on either one of a pair of homoeologs experiencing concerted evolution may be commuted to the other, perhaps explaining the more than tenfold higher rate of gene loss in the RSA and SSA regions than the genome-wide averages of 1.8 % in rice and 3.1 % in sorghum since their divergence about 50 mya. The singular evolutionary history of this pair of grass chromosomes needs further exploration. Elevated gene loss rates and elevated evolutionary rates of the preserved genes in young strata may facilitate speciation in that the loss of alternative copies of duplicated genes leads to reproductive isolation (Werth and Windham 1991; Lynch and Force 2000; Scannell et al. 2006; see also Chap. 1, this volume). The recently proposed inter-relationship between reproductive isolation and autoimmune responses (Bomblies et al. 2007; Yin et al. 2008) draws attention to the finding that orthologs R11 and S5 each contain *25 % of the NBS-LRR resistance genes (Zhou et al. 2004; Paterson et al. 2009b). Second, a high level of CO RR 355 UN 354 Layout: T1 Standard SC Chapter No.: 6 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 104/107 A. H. Paterson et al. 392 References 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 Birchler JA, Veitia RA (2007) The gene balance hypothesis: from classical genetics to modern genomics. Plant Cell 19:395–402 Birchler JA, Riddle NC, Auger DL, Veitia RA (2005) Dosage balance in gene regulation: biological implications. Trends Genet 21:219–226 Blanc G, Wolfe KH (2004) Functional divergence of duplicated genes formed by polyploidy during Arabidopsis evolution. Plant Cell 16:1679–1691 Bomblies K, Lempe J, Epple P, Warthmann N, Lanz C, Dangl JL, Weigel D (2007) Autoimmune response as a mechanism for a dobzhansky-muller-typeincompatibility syndrome in Plants. PLOS Biology 5:1962-1972 Bowers JE, Arias MA, Asher R, Avise JA, Ball RT, Brewer GA, Buss RW, Chen AH, Edwards TM, Estill JC, Exum HE, Goff VH, Herrick KL, Steele CLJ, Karunakaran S, Lafayette GK, Lemke C, Marler BS, Masters SL, McMillan JM, Nelson LK, Newsome GA, Nwakanma CC, Odeh RN, Phelps CA, Rarick EA, Rogers CJ, Ryan SP, Slaughter KA, Soderlund CA, Tang HB, Wing RA, Paterson AH (2005) Comparative physical mapping links conservation of microsynteny to chromosome structure and recombination in grasses. Proc Nat Acad Sci USA 102:13206–13211 Bremer G (1923) A cytological investigation of some species and species-hybrids of the genus Saccharum. Genetica 5:273–326 Bremer G (1961) Problems in breeding and cytology of sugar cane. 4. Origin of increase of chromosome number in species hybrids of Saccharum. Euphytica 10:325–342 Chapman BA, Bowers JE, Feltus FA, Paterson AH (2006) Buffering crucial functions by paleologous duplicated genes may impart cyclicality to angiosperm genome duplication. Proc Nat Acad Sci USA 103:2730–2735 Charlesworth B (2002) The evolution of chromosomal sex determination. Novartis Found Symp 244:207–219 (discussion 220–204, 253–207) Chittenden LM, Schertz KF, Lin YR, Wing RA, Paterson AH (1994) A detailed Rflp map of Sorghum-bicolor X S-propinquum, suitable for high-density mapping, suggests ancestral duplication of sorghum chromosomes or chromosomal segments. Theor Appl Genet 87:925–933 Devos KM, Pittaway TS, Reynolds A, Gale MD (2000) Comparative mapping reveals a complex relationship between the pearl millet genome and those of foxtail millet and rice. Theor Appl Genetics 100:190–198 Feldman M, Levy AA (2005) Allopolyploidy—a shaping force in the evolution of wheat genomes. Cytogenet Genome Res 109:250–258 Freeling M (2001) Grasses as a single genetic system: reassessment 2001. Plant Physiol 125:1191–1197 389 390 PR OO 388 D 387 TE 385 386 EC 384 CO RR 383 F 391 concerted evolution, associated stratification of chromosomal segments, and extensive homoeologous gene loss are each characteristics of sex chromosomes in organisms from divergent branches of the tree of life, including humans (Lahn and Page 1999), chickens (Lawson Handley et al. 2006), fungi (Charlesworth 2002), and plants (Ming and Moore 2007). Moreover, unexpectedly close proximity between, and co-expansion of, NBS-LRR and several sex-determining gene analogs is found, particularly on S5 (Wang et al. 2011). A hypothesis for further study is whether genes on the various orthologs and paralogs of these chromosomes (or regions therein) could have some ‘functional coherence’ resembling that of the human Y chromosome (Lahn and Page 1999). 382 UN Editor Proof 104 Layout: T1 Standard SC Chapter No.: 6 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 105/107 105 EC TE D PR OO F Goff SA, Ricke D, Lan TH, Presting G, Wang RL, Dunn M, Glazebrook J, Sessions A, Oeller P, Varma H, Hadley D, Hutchinson D, Martin C, Katagiri F, Lange BM, Moughamer T, Xia Y, Budworth P, Zhong JP, Miguel T, Paszkowski U, Zhang SP, Colbert M, Sun WL, Chen LL, Cooper B, Park S, Wood TC, Mao L, Quail P, Wing R, Dean R, Yu YS, Zharkikh A, Shen R, Sahasrabudhe S, Thomas A, Cannings R, Gutin A, Pruss D, Reid J, Tavtigian S, Mitchell J, Eldredge G, Scholl T, Miller RM, Bhatnagar S, Adey N, Rubano T, Tusneem N, Robinson R, Feldhaus J, Macalma T, Oliphant A, Briggs S (2002) A draft sequence of the rice genome (Oryza sativa L. ssp japonica). Science 296:92–100 Haldane JBS (1933) The part played by recurrent mutation in evolution. Am Nat 67:5–19 Heaton EA, Dohleman FG, Long SP (2008) Meeting US biofuel goals with less land: the potential of Miscanthus. Glob Change Biol 14:2000–2014 Hilu KW (2004) Phylogenetics and chromosomal evolution in the Poaceae (grasses). Aust J Bot 52:10 International Rice Genome Sequencing P (2005) The map-based sequence of the rice genome. Nature 436:793–800 Jaillon O, Aury JM et al (2007) The grapevine genome sequence suggests ancestral hexaploidization in major angiosperm phyla. Nature 449:463–467 Jannoo N, Grivet L, Chantret N, Garsmeur O, Glaszmann JC, Arruda P, D’Hont A (2007) Orthologous comparison in a gene-rich region among grasses reveals stability in the sugarcane polyploid genome. Plant J 50:574–585 Jeswiet J (1929) The development of selection and breeding of the sugarcane in Java. Int Soc Sugar Cane Technol 3:44–57 Kim C, Tang H, Paterson AH (2009) Duplication and divergence of grass genomes: integrating the chloridoids. Trop Plant Biol 2:51–62 Kishimoto N, Higo H, Abe K, Arai S, Saito A, Higo K (1994) Identification of the duplicated segments in rice chromosomes 1 and 5 by linkage analysis of cDNA markers of known functions. Theor Appl Genet 88:722–726 Lahn BT, Page DC (1999) Four evolutionary strata on the human X chromosome. Science (New York) 286:964–967 Lawrence WJC (1931) The secondary association of chromosomes. Cytologia 2:352–384 Lawson Handley LJ, Hammond RL, Emaresi G, Reber A, Perrin N (2006) Low Y chromosome variation in Saudi-Arabian hamadryas baboons (Papio hamadryas hamadryas). Heredity 96:298–303 Liu H, Sachidanandam R, Stein L (2001) Comparative genomics between rice and Arabidopsis shows scant collinearity in gene order. Genome Res 11:2020–2026 Lohithaswa HC, Feltus FA, Singh HP, Bacon CD, Bailey CD, Paterson AH (2007) Leveraging the rice genome sequence for comparative genomics in monocots. Theor Appl Genetics 115:237–243 Lynch M, Conery JS (2000) The evolutionary fate and consequences of duplicate genes. Science 290:1151–1155 Lynch M, Force AG (2000) The origin of interspecific genomic incompatibility via gene duplication. Am Nat 156:590–605 Maere S, De Bodt S, Raes J, Casneuf T, Van Montagu M, Kuiper M, Van de Peer Y (2005) Modeling gene and genome duplications in eukaryotes. Proc Nat Acad Sci USA 102:5454–5459 Mayer KFX, Martis M, Hedley PE, Simkova H, Liu H, Morris JA, Steuernagel B, Taudien S, Roessner S, Gundlach H, Kubalakova M, Suchankova P, Murat F, Felder M, Nussbaumer T, Graner A, Salse J, Endo T, Sakai H, Tanaka T, Itoh T, Sato K, Platzer M, Matsumoto T, Scholz U, Dolezel J, Waugh R, Stein N (2011) Unlocking the barley genome by chromosomal and comparative genomics. The Plant Cell 23:1249–1263 Ming R, Moore PH (2007) Genomics of sex chromosomes. Curr Opin Plant Biol 10:123–130 Ming R, Liu SC, Lin YR, da Silva J, Wilson W, Braga D, van Deynze A, Wenslaff TF, Wu KK, Moore PH, Burnquist W, Sorrells ME, Irvine JE, Paterson AH (1998) Detailed alignment of CO RR 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 UN Editor Proof 6 Ancient and Recent Polyploidy in Monocots Layout: T1 Standard SC Chapter No.: 6 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 106/107 EC TE D PR OO F Saccharum and sorghum chromosomes: comparative organization of closely related diploid and polyploid genomes. Genetics 150:1663–1682 Ming R, Liu SC, Moore PH, Irvine JE, Paterson AH (2001) QTL analysis in a complex autopolyploid: genetic control of sugar content in sugarcane. Genome Res 11:2075–2084 Ming R, Del Monte TA, Hernandez E, Moore PH, Irvine JE, Paterson AH (2002a) Comparative analysis of QTLs affecting plant height and flowering among closely-related diploid and polyploid genomes. Genome 45:794–803 Ming R, Wang YW, Draye X, Moore PH, Irvine JE, Paterson AH (2002b) Molecular dissection of complex traits in autopolyploids: mapping QTLs affecting sugar yield and related traits in sugarcane. Theor Appl Genet 105:332–345 Murat F, Xu JH, Tannier E, Abrouk M, Guilhot N, Pont C, Messing J, Salse J (2010) Ancestral grass karyotype reconstruction unravels new mechanisms of genome shuffling as a source of plant evolution. Genome Res 20:1545–1557 Nagamura Y, Inoue T, Antonio B, Shimano T, Kajiya H, Shomura A, Lin S, Kuboki Y, Harushima Y, Kurata N, Minobe Y, Yano M, Sasaki T (1995) Conservation of duplicated segments between rice chromosomes 11 and 12. Breed Sci 45:373–376 Paterson AH (2008) Paleopolyploidy and its impact on the structure and function of modern plant genomes. Genome Dyn 4:1–12 Paterson AH, Lan TH, Reischmann KP, Chang C, Lin YR, Liu SC, Burow MD, Kowalski SP, Katsar CS, DelMonte TA, Feldmann KA, Schertz KF, Wendel JF (1996) Toward a unified genetic map of higher plants, transcending the monocot-dicot divergence. Nat Genet 14:380–382 Paterson A, Bowers J, Peterson D, Estill J, Chapman B (2003) Structure and evolution of cereal genomes. Curr Opin Genet Dev 13:644–650 Paterson AH, Bowers JE, Chapman BA (2004) Ancient polyploidization predating divergence of the cereals, and its consequences for comparative genomics. Proc Nat Acad Sci USA 101:9903–9908 Paterson AH, Chapman BA, Kissinger J, Bowers JE, Feltus FA, Estill J, Marler BS (2006) Convergent retention or loss of gene/domain families following independent whole-genome duplication events in Arabidopsis, Oryza, Saccharomyces, and Tetraodon. Trends Genet 22:597–602 Paterson AH, Bowers JE, Feltus FA, Tang H, Lin L, Wang X (2009a) Comparative genomics of grasses promises a bountiful harvest. Plant Physiol 149:125–131 Paterson AH, Bowers JE, Bruggmann R, Dubchak I, Grimwood J, Gundlach H, Haberer G, Hellsten U, Mitros T, Poliakov A, Schmutz J, Spannagl M, Tang H, Wang X, Wicker T, Bharti AK, Chapman J, Feltus FA, Gowik U, Grigoriev IV, Lyons E, Maher CA, Martis M, Narechania A, Otillar RP, Penning BW, Salamov AA, Wang Y, Zhang L, Carpita NC, Freeling M, Gingle AR, Hash CT, Keller B, Klein P, Kresovich S, McCann MC, Ming R, Peterson DG, Mehboob ur R, Ware D, Westhoff P, Mayer KF, Messing J, Rokhsar DS (2009b) The Sorghum bicolor genome and the diversification of grasses. Nature 457:551–556 Raven PH, Evert RF, Eichhorn SE (2005) Biology of plants, 7th edn. W. H. Freeman, New York Salse J, Abrouk M, Bolot S, Guilhot N, Courcelle E, Faraut T, Waugh R, Close TJ, Messing J, Feuillet C (2009) Reconstruction of monocotelydoneous protochromosomes reveals faster evolution in plants than in animals. Proceedings of the National academy of sciences of the United States of America 106:14908–14913 Salse J, Bolot S, Throude M, Jouffe V, Piegu B, Quraishi UM, Calcagno T, Cooke R, Delseny M, Feuillet C (2008) Identification and characterization of shared duplications between rice and wheat provide new insight into grass genome evolution. Plant Cell 20:11–24 Scannell DR, Byrne KP, Gordon JL, Wong S, Wolfe KH (2006) Multiple rounds of speciation associated with reciprocal gene loss in polyploid yeasts. Nature 440:341–345 Schnable PS, Ware D, Fulton RS, Stein JC, Wei FS, Pasternak S, Liang CZ, Zhang JW, Fulton L, Graves TA, Minx P, Reily AD, Courtney L, Kruchowski SS, Tomlinson C, Strong C, Delehaunty K, Fronick C, Courtney B, Rock SM, Belter E, Du FY, Kim K, Abbott RM, Cotton M, Levy A, Marchetto P, Ochoa K, Jackson SM, Gillam B, Chen WZ, Yan L, CO RR 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 A. H. Paterson et al. UN Editor Proof 106 Layout: T1 Standard SC Chapter No.: 6 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 107/107 107 EC TE D PR OO F Higginbotham J, Cardenas M, Waligorski J, Applebaum E, Phelps L, Falcone J, Kanchi K, Thane T, Scimone A, Thane N, Henke J, Wang T, Ruppert J, Shah N, Rotter K, Hodges J, Ingenthron E, Cordes M, Kohlberg S, Sgro J, Delgado B, Mead K, Chinwalla A, Leonard S, Crouse K, Collura K, Kudrna D, Currie J, He RF, Angelova A, Rajasekar S, Mueller T, Lomeli R, Scara G, Ko A, Delaney K, Wissotski M, Lopez G, Campos D, Braidotti M, Ashley E, Golser W, Kim H, Lee S, Lin JK, Dujmic Z, Kim W, Talag J, Zuccolo A, Fan C, Sebastian A, Kramer M, Spiegel L, Nascimento L, Zutavern T, Miller B, Ambroise C, Muller S, Spooner W, Narechania A, Ren LY, Wei S, Kumari S, Faga B, Levy MJ, McMahan L, Van Buren P, Vaughn MW, Ying K, Yeh CT, Emrich SJ, Jia Y, Kalyanaraman A, Hsia AP, Barbazuk WB, Baucom RS, Brutnell TP, Carpita NC, Chaparro C, Chia JM, Deragon JM, Estill JC, Fu Y, Jeddeloh JA, Han YJ, Lee H, Li PH, Lisch DR, Liu SZ, Liu ZJ, Nagel DH, McCann MC, SanMiguel P, Myers AM, Nettleton D, Nguyen J, Penning BW, Ponnala L, Schneider KL, Schwartz DC, Sharma A, Soderlund C, Springer NM, Sun Q, Wang H, Waterman M, Westerman R, Wolfgruber TK, Yang LX, Yu Y, Zhang LF, Zhou SG, Zhu Q, Bennetzen JL, Dawe RK, Jiang JM, Jiang N, Presting GG, Wessler SR, Aluru S, Martienssen RA, Clifton SW, McCombie WR, Wing RA, Wilson RK (2009) The B73 maize genome: complexity, diversity, and dynamics. Science 326:1112–1115 Schnable JC, Springer NM, Freeling M (2011) Differentiation of the maize subgenomes by genome dominance and both ancient and ongoing gene loss. Proc Nat Acad Sci USA 108:4069–4074 Seoighe C, Gehring C (2004) Genome duplication led to highly selective expansion of the Arabidopsis thaliana proteome. Trends Genet 20:461–464 Singh NK, Dalal V, Batra K, Singh BK, Chitra G, Singh A, Ghazi IA, Yadav M, Pandit A, Dixit R, Singh PK, Singh H, Koundal KR, Gaikwad K, Mohapatra T, Sharma TR (2007) Singlecopy genes define a conserved order between rice and wheat for understanding differences caused by duplication, deletion, and transposition of genes. Funct Integr Genomics 7:17–35 Soderstrom TR, Hilu KW, Campbell CS, Barkworth MA (1987) Grass systematics and evolution. Smithsonian Institution Press, Washington Soltis DE, Smith S, Cellinese N, Refulio-Rodriquez NF, Olmstead R, Crawley S, Black C, Diouf D, Hilu KW, Latvis M, Wurdack K, Xi Z, Davis C, Donoghue M, Soltis PS (2011) Inferring angiosperm phylogeny: a 17-gene analysis. Am J Bot 98:704–730 Spangler R (2003) Taxonomy of Sarga, Sorghum, and Vacoparis (Poaceae: Andropogoneae). Aust Syst Bot 16:279–299 Spangler R, Zaitchik B, Russo E, Kellogg E (1999) Andropogoneae evolution and generic limits in Sorghum (Poaceae) using ndhF sequences. Syst Bot 24:267–281 Tang H, Wang X, Bowers JE, Ming R, Alam M, Paterson AH (2008) Unraveling ancient hexaploidy through multiply-aligned angiosperm gene maps. Genome Res 18:1944–1954 Tang H, Bowers JE, Wang X, Ming R, Alam M, Paterson AH (2008) Synteny and colinearity in plant genomes. Science 320:486–488 Tang HB, Bowers JE, Wang XY, Paterson AH (2010) Angiosperm genome comparisons reveal early polyploidy in the monocot lineage. Proc Nat Acad Sci USA 107:472–477 The International Brachypodium Initiative (2010) Genome sequencing and analysis of the model grass Brachypodium distachyon. Nature 463:763–768 Thomas BC, Pedersen B, Freeling M (2006) Following tetraploidy in an Arabidopsis ancestor, genes were removed preferentially from one homeolog leaving clusters enriched in dosesensitive genes. Genome Res 16:934–946 Van de Peer Y (2004) Computational approaches to unveiling ancient genome duplications. Nat Rev Genet 5:752–763 Vandepoele K, Simillion C, Van de Peer Y (2003) Evidence that rice and other cereals are ancient aneuploids. Plant Cell 15:2192–2202 Veitia RA, Bottani S, Birchler JA (2008) Cellular reactions to gene dosage imbalance: genomic, transcriptomic and proteomic effects. Trends Genet 24:390–397 Wang X, Shi X, Hao B, Ge S, Luo J (2005) Duplication and DNA segmental loss in the rice genome: implications for diploidization. New Phytol 165:937–946 CO RR 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 UN Editor Proof 6 Ancient and Recent Polyploidy in Monocots Layout: T1 Standard SC Chapter No.: 6 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 108/107 EC TE D PR OO F Wang XY, Shi XL, Li Z, Zhu QH, Kong L, Tang W, Ge S, Luo JC (2006) Statistical inference of chromosomal homology based on gene colinearity and applications to arabidopsis and rice. BMC Bioinformatics 7:447 Wang X, Tang H, Bowers JE, Feltus FA, Paterson AH (2007) Extensive concerted evolution of rice paralogs and the road to regaining independence. Genetics 177:1753–1763 Wang X, Tang H, Paterson AH (2011) Seventy million years of concerted evolution of a homoeologous chromosome pair, in parallel, in major Poaceae lineages. Plant Cell 23:27–37 Werth CR, Windham MD (1991) A model for divergent, allopatric speciation of polyploid pteridophytes resulting from silencing of duplicate-gene expression. Am Nat 137:515–526 Wicker T, Mayer KFX, Gundlach H, Martis M, Steuernagel B, Scholz U, Simkova H, Kubalakova M, Choulet F, Taudien S, Platzer M, Feuillet C, Fahima T, Budak H, Dolezel J, Keller B, Stein N (2011) Frequent gene movement and pseudogene evolution is common to the large and complex genomes of wheat, barley, and their relatives. The Plant Cell 23:1706–1718 Woodhouse MR, Schnable JC, Pedersen BS, Lyons E, Lisch D, Subramaniam S, Freeling M (2010) Following tetraploidy in maize, a short deletion mechanism removed genes preferentially from one of the two homologs. PLoS Biol 8:e1000409 Xiong Z, Gaeta RT, Pires JC (2011) Homoeologous shuffling and chromosome compensation maintain genome balance in resynthesized allopolyploid Brassica napus. Proc Nat Acad Sci USA 108:7908–7913 Yin T, Difazio SP, Gunter LE, Zhang X, Sewell MM, Woolbright SA, Allan GJ, Kelleher CT, Douglas CJ, Wang M, Tuskan GA (2008) Genome structure and emerging evidence of an incipient sex chromosome in Populus. Genome Res 18:422–430 Yu J, Wang J, Lin W, Li SG, Li H, Zhou J, Ni PX, Dong W, Hu SN, Zeng CQ, Zhang JG, Zhang Y, Li RQ, Xu ZY, Li ST, Li XR, Zheng HK, Cong LJ, Lin L, Yin JN, Geng JN, Li GY, Shi JP, Liu J, Lv H, Li J, Deng YJ, Ran LH, Shi XL, Wang XY, Wu QF, Li CF, Ren XY, Wang JQ, Wang XL, Li DW, Liu DY, Zhang XW, Ji ZD, Zhao WM, Sun YQ, Zhang ZP, Bao JY, Han YJ, Dong LL, Ji J, Chen P, Wu SM, Liu JS, Xiao Y, Bu DB, Tan JL, Yang L, Ye C, Zhang JF, Xu JY, Zhou Y, Yu YP, Zhang B, Zhuang SL, Wei HB, Liu B, Lei M, Yu H, Li YZ, Xu H, Wei SL, He XM, Fang LJ, Zhang ZJ, Zhang YZ, Huang XG, Su ZX, Tong W, Li JH, Tong ZZ, Li SL, Ye J, Wang LS, Fang L, Lei TT, Chen C, Chen H, Xu Z, Li HH, Huang HY, Zhang F, Xu HY, Li N, Zhao CF, Dong LJ, Huang YQ, Li L, Xi Y, Qi QH, Li WJ, Hu W, Zhang YL, Tian XJ, Jiao YZ, Liang XH, Jin JA, Gao L, Zheng WM, Hao BL, Liu SQ, Wang W, Yuan LP, Cao ML, McDermott J, Samudrala R, Wong GKS, Yang HM (2005) The genomes of Oryza sativa: a history of duplications. Plos Biology 3:266–281 Zhou JH, Wang JL, Xu JC, Lei CL, Ling ZZ (2004) Identification and mapping of a rice blast resistance gene Pi-g(t) in the cultivar Guangchangzhan. Plant Pathol. 53:191–196 Zhang Y, Xu GH, Guo XY, Fan LJ (2005) Two ancient rounds of polyploidy in rice genome. J Zhejiang Univ Sci B 6:87–90 CO RR 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 A. H. Paterson et al. UN Editor Proof 108 Metadata of the chapter that will be visualized in SpringerLink Book Title Polyploidy and Genome Evolution Series Title Chapter Title Genomic Plasticity in Polyploid Wheat Copyright Year 2012 Copyright HolderName Springer-Verlag Berlin Heidelberg Corresponding Author Family Name Kashkush Particle Given Name Khalil Suffix Author Division Department of Life Sciences Organization Ben-Gurion University Address 84105, Beer-Sheva, Israel Email kashkush@bgu.ac.il Family Name Feldman Particle Given Name Moshe Suffix Division Plant Sciences Department Organization The Weizmann Institute of Science Address 76100, Rehovot, Israel Email Author Family Name Levy Particle Given Name Avraham Suffix Division Plant Sciences Department Organization The Weizmann Institute of Science Address 76100, Rehovot, Israel Email Author Family Name Chalhoub Particle Given Name Boulos Suffix Division UMR INRA 1165—CNRS 8114—UEVE, Organization and Evolution of Plant Genomes (OEPG) Organization Unité de Recherche En Génomique Végétale (URGV) Address 2 Rue Gaston Crémieux, 91057, Evry Cedex, France Email Abstract The importance of hybridization and polyploidization in wheat speciation has been recognized for close to a century (Sakamura 1918; Kihara 1919, 1924, 1954; Percival 1921; Sax 1927). Following these pioneering works, it quickly became apparent that polyploid wheats are not the sum of their constituent genomes. This is not unexpected because the nascent hybrids/polyploids are equipped with a complex set of regulatory elements and of copy number variation that originate from two or more divergent genomes and that generate novel types of interactions and dosage effects. Moreover, they have to adjust at the cytological level, at the level of gene expression, and at the protein level. They also have to maintain genome stability through the regulation of meiotic pairing and recombination, the orchestration of cell division, and the silencing of transposons. The recent studies described here provide an impressive account with regard to the extent and the rapid time course at which a new genetic variant was established upon hybridization and polyploidization. We describe here the current knowledge on the changes that occurred in the wheat genome upon allopolyploidization, starting from the early evolutionary and cytological studies to the recent genomic analyses. Book ISBN: 978-3-642-31441-4 Page: 109/134 Chapter 7 2 Genomic Plasticity in Polyploid Wheat 3 5 4 Moshe Feldman, Avraham Levy, Boulos Chalhoub and Khalil Kashkush 6 Abstract PR OO 1 7 13 14 15 16 17 18 19 20 21 22 23 D 12 TE 10 11 The importance of hybridization and polyploidization in wheat speciation has been recognized for close to a century (Sakamura 1918; Kihara 1919, 1924, 1954; Percival 1921; Sax 1927). Following these pioneering works, it quickly became apparent that polyploid wheats are not the sum of their constituent genomes. This is not unexpected because the nascent hybrids/polyploids are equipped with a complex set of regulatory elements and of copy number variation that originate from two or more divergent genomes and that generate novel types of interactions and dosage effects. Moreover, they have to adjust at the cytological level, at the level of gene expression, and at the protein level. They also have to maintain genome stability through the regulation of meiotic pairing and recombination, the orchestration of cell division, and the silencing of transposons. The recent studies described here provide an impressive account with regard to the extent and the rapid time course at which a new genetic variant was established upon hybridization and polyploidization. We describe here the current knowledge on the changes that occurred in the wheat genome upon allopolyploidization, starting from the early evolutionary and cytological studies to the recent genomic analyses. EC 9 CO RR 8 F Book ID: 272454_1_En Date: 16-8-2012 M. Feldman  A. Levy Plant Sciences Department, The Weizmann Institute of Science, 76100, Rehovot, Israel B. Chalhoub UMR INRA 1165—CNRS 8114—UEVE, Organization and Evolution of Plant Genomes (OEPG), Unité de Recherche En Génomique Végétale (URGV), 2 Rue Gaston Crémieux, 91057, Evry Cedex, France UN Editor Proof Layout: T1 Standard SC Chapter No.: 7 K. Kashkush (&) Department of Life Sciences, Ben-Gurion University, 84105, Beer-Sheva, Israel e-mail: kashkush@bgu.ac.il P. S. Soltis and D. E. Soltis (eds.), Polyploidy and Genome Evolution, DOI: 10.1007/978-3-642-31442-1_7,  Springer-Verlag Berlin Heidelberg 2012 109 Layout: T1 Standard SC Chapter No.: 7 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 110/134 110 M. Feldman et al. Diploids (2n = 2x = 14) Tetraploids (2n = 4x = 28) Hexaploids (2n = 6x = 42) 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 D 28 TE 27 Spelt (hulled) wheats Naked (free threshing) wheats The discovery of wild emmer wheat, the progenitor of most domesticated wheats (Aaronsohn and Schweinfurth 1906; Aaronsohn 1910), made it possible for Schulz (1913) to assemble the first natural classification of the wheats. He divided the genus Triticum into three major groups: einkorn, emmer, and dinkel. This classification was supported by the pioneering cytological study of Sakamura (1918), who was the first to determine the correct chromosome number of the wheats. Sakamura discovered that Schulz’s three groups of wheats differ in their chromosome number: the einkorns are diploids (2n = 14), the emmers are tetraploids (2n = 28), and the dinkels are hexaploids (2n = 42) (Table 7.1). It then became obvious that the species of Triticum represent a polyploid series with diploid, tetraploid, and hexaploid species. Since then, the species of the wheat group (Triticum and its closely related genus Aegilops) have been subjected to extensive taxonomic, cytogenetic, genetic, biochemical, molecular, and evolutionary study by numerous scientists (see reviews of Kihara 1954; Mac Key 1966; Morris and Sears 1967; Kimber and Sears 1987; Feldman et al. 1995; Feldman 2001; Gupta et al. 2005; Dvorak 2009). Due to these extensive studies the allopolyploid species of the wheat group became a classic example of evolution through allopolyploidy. The cytogenetic studies of Kihara (1919, 1924, 1954), Percival (1921), Sax (1927), and others on chromosome pairing in hybrids among species of different ploidy levels showed that all the polyploid species of the group form an allopolyploid series based on x = 7. Each allopolyploid species was identified as a product of hybridization followed by chromosome doubling (Fig. 7.1). Von Tschermak and Bleier (1926) were the first to identify a spontaneous chromosome doubling in the cross of wild emmer (T. turgidum ssp. dicoccoides) with Aegilops geniculata, thus demonstrating the possibility of species formation via allopolyploidy in the wheat group. Subsequent studies showed that the frequency of unreduced gametes in intergeneric hybrids of wheat could be in some hybrids as high as 50 % (Kihara and Lilienfeld 1949). Therefore, one might assume that there is a high potential for the frequent and EC 26 Domesticated wheats Einkorn (one-grained T. aegilopoides T. monococcum None wheat) —[ Emmer (two-grained T. dicoccoides T. dicoccum — T. durum wheat) —[ [ T. turgidum T. polonicum Dinkel None T. spelta — [ T. compactum T. vulgare CO RR 25 Wild progenitors 7.1 The Wheat Group: Natural and Synthetic Polyploids UN 24 Series F Ploidy level PR OO Editor Proof Table 7.1 Superimposition of Sakamura’s (1918) finding of the right chromosome number on Schulz’s (1913) natural classification of the wheats Layout: T1 Standard SC Chapter No.: 7 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 111/134 111 PR OO F Editor Proof 7 Genomic Plasticity in Polyploid Wheat 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 TE 58 EC 56 57 recurrent formation of interspecific or intergeneric hybrids and allopolyploids in the wheat group. The discovery by Blakeslee (1937) that colchicine can induce chromosome doubling opened new possibilities for the study of wheat evolution through allopolyploidization. It also provided an easy method to synthesize different wheat allopolyploids, some of which have similar genomes to natural allopolyploids and others having new genomic combinations (see examples in Ozkan et al. 2001). Synthetic allopolyploids, either induced or occurring spontaneously, offer excellent tools to mimic the evolutionary speciation events that occurred in nature and to test in a controlled manner; the new features of the hybrid/polyploid genome compared to those of its parents. Since the discovery that the polyploid species of wheat comprise an allopolyploid series, attempts have been made to identify the diploid donors of the different genomes to the allopolyploids of the wheat group. Most of these attempts used the cytogenetic approach of genome analysis, developed by Kihara (1919, 1924) and based on the concept of genome stability, assuming that the genomes of the allopolyploid species remain similar to those of their diploid parents. However, the accumulating cytogenetic and molecular evidence has indicated that this is not the case; while one genome remains relatively unchanged, the second genome(s) of the allopolyploids have changed considerably from those of their parental diploids. These genomes were termed modified genomes by Kihara (1954) and other wheat cytogeneticists. Every allopolyploid species of Aegilops and Triticum contains an unchanged genome side-by-side with a modified one whose diploid origin has been intricate and difficult to trace (Zohary and Feldman 1962). CO RR 55 UN 54 D Fig. 7.1 Evolutionary history of allotetraploid and allohexaploid wheat. Diploid wheats (2n = 2x = 14) from the Triticum-Aegilops group diverged *4 million years ago (MYA) from a common progenitor (Huang et al. 2002). Interspecific hybridization between the diploid T. urartu (genome AA) as male and the donor of the B genome (an unknown species similar to Ae. speltoides) as female, followed by chromosome doubling, gave rise, *0.5 MYA, to wild allotetraploid wheat, Triticum turgidum ssp. dicoccoides (2n = 4x = 28, genome BBAA), an allotetraploid considered as the direct progenitor of durum wheat. Domestication of allotetraploid wheat took place *10,500 YA and was rapidly followed (*9,500 YA) by a second round of intergeneric hybridization and chromosome doubling between domesticated allotetraploid wheat and the donor of the D genome, Ae. tauschii (2n = 2x = 14, genome DD), giving rise to bread wheat, an allohexaploid with 2n = 6x = 42 chromosomes (genome BBAADD). Adapted from Levy and Feldman (2004) Layout: T1 Standard SC Chapter No.: 7 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 112/134 M. Feldman et al. 95 7.2 Cytological Diploidization 87 88 89 90 91 92 93 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 PR OO 85 86 D 83 84 TE 81 82 Because of the close relationship of the progenitors and the similarity of the two different genomes of allotetraploid wheat and of the three different genomes of allohexaploid wheat (Morris and Sears 1967), the successful establishment of the polyploid species in nature required the acquisition during the allopolyploidization process of molecular and genetic systems that would prevent intergenomic pairing and recombination. By restricting pairing to fully homologous chromosomes (intragenomic pairing), the cytological-diploidizing systems ensure exclusive bivalent pairing at meiosis and, consequently, regular segregation of genetic material, complete fertility, and genetic stability. Cytological diploidization has been brought about in allopolyploid wheat by two independent systems that complement each other: One system is based on genetic control of pairing, and the second is based on physical divergence of chromosomes. Historically, the first mechanism of cytological diploidization that received much attention was a genetic system involved in sustaining the exclusive bivalent pairing in allopolyploids. It consisted of the activity of the genetic loci Ph1 on chromosome arm 5BL and Ph2 on chromosome arm 3DS (Sears 1976). These loci suppress pairing of homoeologous chromosomes while allowing homologs to pair regularly. The mechanism controlling the Ph1 mode of action is still unclear. Mapping data suggests the involvement of a complex locus that affects cell cycle progression through the regulation of Cyclin-dependent kinases (Griffiths et al. 2006); however, direct evidence through complementation studies is still missing. It was suggested by several authors that Ph-like genes exist in diploid species of the wheat group (Okamoto and Inomata 1974; Avivi 1976; Waines 1976; Maan EC 80 CO RR 79 F 94 Nevertheless, genome analysis studies revealed that allotetraploid wheat (genome BBAA) originated from hybridization events involving two diploid progenitors classified in the genera Aegilops and Triticum. Genome B, which is a modified genome, was derived from Ae. speltoides and underwent changes on the polyploid level or, more likely, from a closely related species to Ae. speltoides which is extinct or extant (Feldman et al. 1995). Genome A, which has been modified relatively little, was derived from T. urartu (Chapman et al. 1976; Dvorak 1976). Allohexaploid wheat (genome BBAADD) originated from hybridization between allotetraploid wheat and Ae. tauschii, the donor of the D genome (Kihara 1944; McFadden and Sears 1944, 1946). The evolution of the wheats is presented in Fig. 7.1 (for details, see Feldman et al. 1995; Feldman 2001). Modern classification for the Triticum group (Van Slageren 1994; Table 7.2) recognizes two diploid species, T. monococcum L. and T. urartu Tum. ex Gand., two tetraploid species, T. turgidum L. and T. timopheevii (Zhuk.) Zhuk., and two hexaploid species, T. aestivum L. and T. zhukovskyi Men. & Er. The economically important wheats are T. aestivum (bread wheat, comprising 95 % of the global wheat production) and T. turgidum (macaroni wheat). 78 UN Editor Proof 112 Layout: T1 Standard SC Chapter No.: 7 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 113/134 7 Genomic Plasticity in Polyploid Wheat 113 AmAm AA Tetraploids (2n = 4x = 28) GGAA Triticum urartu Triticum timopheevii ssp. armeniacum ssp. timopheevii Triticum turgidum ssp. dicoccoides ssp. dicoccon ssp paleocochicum ssp. parvicoccumb ssp. durum ssp. turgidum ssp. polonicum ssp. turanicum ssp. carthlicum Triticum zhukovskyi D BBAA Triticum monococcum ssp. aegilopoides ssp. monococcum EC TE Hexaploids (2n = 6x = 42) GGAA AmAm BBAADD Triticum aestivum ssp. spelta ssp. macha ssp. aestivum ssp. compactum ssp. sphaerococcum a Wild einkorn Domesticated einkorn or small spelt None (wild form) F Diploids (2n = 2x = 14) PR OO Editor Proof Table 7.2 The nomenclature of the commercially cultivated wheats and their immediate wild relatives (after van Slageren, 1994)a Ploidy level Genome Species and subspecies Common name Wild timopheevii Domesticated timopheevii Wild emmer Domesticated emmer Georgian wheat None Macaroni or hard wheat Rivet, cone or pollard wheat Polish wheat Khorassan wheat Persian wheat None Dinkel or large spelt None Common or bread wheat Club wheat Indian dwarf or short wheat 120 121 122 123 124 125 126 127 128 129 130 131 1977) and they became more effective at the polyploid level as a result of duplication. This dosage-dependent effect might have been selected to improve the fertility of the allopolyploid. This genetic system superimposes itself on, takes advantage of, and thereby reinforces the system of physical homoeologous differentiation already in existence and described below. The genetic system is very effective in suppressing homoeologous pairing in interspecific and intergeneric F1 hybrids. However, its suppressive effect on homoeologous pairing in allopolyploid wheat might not be essential since in plants deficient for Ph1 there is relatively very little such pairing (Sears 1976). Interestingly, and in accord with the above, gene(s) like Ph were not found in all the allopolyploid species of the closely related genus Aegilops. In spite of this, these species, relying solely on the structural homoeologous differentiation, exhibit exclusive bivalent pairing of fully homologous chromosomes, i.e., exclusive intragenomic pairing. It might be that UN 119 CO RR Taxa derived from a single spontaneous or induced mutation and are not commercially cultivated, such as diploid T. sinskajae Filat and Kurk., tetraploid T. militinae Zhuk. and Migush., and hexaploid T. vavilovii (Tum.) Jakubz., are not included b Extinct, described by Kislev 1980 Layout: T1 Standard SC Chapter No.: 7 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 114/134 114 M. Feldman et al. Occur during or immediately after allopolyploidization F Genetic and epigenetic changes Species-specific Lead to cytological diploidization Occur during the life of the allopolyploid species Mostly genetic changes Population- or biotype-specific Promote genetic diversity, flexibility, and adaptability Improve harmonic functioning of the divergent genomes Stabilize the nascent allopolyploid and facilitate its establishment as a new species in nature PR OO Editor Proof Table 7.3 Types and characteristics of genomic changes in allopolyploid wheat Revolutionary changes (triggered by allopolyploidization) Evolutionary changes (Facilitated by allopolyploidy) 143 7.3 Genomic Structural Changes 137 138 139 140 141 144 145 146 147 148 149 150 151 152 153 154 155 156 TE 136 EC 135 Studies with synthetic polyploids as well as genome sequencing data indicate that a broad range of DNA rearrangements occurred during, or soon after, hybridization and polyploidization. What triggers these changes is a fascinating and still open question, but what is clear is that these changes are extensive, including DNA loss, transposon activation, gene duplication, and pseudogenization, and are relatively rapid. Feldman and Levy (2005) have distinguished between the revolutionary changes that occur rapidly and evolutionary changes that take place throughout the evolution of the allopolyploid (Tables 7.3 and 7.4). Note that evolutionary changes might also occur in an accelerated manner, thanks to the buffering of mutations in the polyploid background (Mac Key 1954, 1958; Sears 1972; Thompson et al. 2006), leading to rapid neo- or subfunctionalization of genes and to a process of diploidization and of divergence from the diploid progenitor genomes. CO RR 134 UN 133 D 142 stringent selection for fertility under domestication has favored the development of two systems to ensure suppression of multivalent formation and to promote bivalent pairing. Moreover, the cytological diploidization of the allopolyploid wheats that leads to disomic inheritance prevents independent segregation of genes from the different genomes. This mode of inheritance leads to permanent maintenance of favorable intergenomic genetic interactions. It enables fixation of heterotic interaction between genomes and sustained division of tasks (genome asymmetry) between genomes. A series of DNA rearrangements in the allopolyploid further contributes to the physical divergence between the homoeologous chromosomes and to the strengthening of the disomic genetic system. These changes are described below. 132 Layout: T1 Standard SC Chapter No.: 7 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 115/134 7 Genomic Plasticity in Polyploid Wheat 115 Structural • Elimination of low-copy DNA sequences Chromatin remodeling Chromatin modifications Heterochromatinization DNA methylation Small RNAs activation or repression Methylation (leading to silencing) Demethylation (leading to gene activation) Release transposons from silencing Silencing transposons PR OO • Elimination, reduction, or amplification of highcopy sequences • Inter-genomic invasion of DNA sequences • Elimination of rRNA and 5S RNA genes Functional • Gene loss • • • • • F Editor Proof Table 7.4 Revolutionary changes induced by allopolyploidization in wheat Level Genetic Epigenetic • 161 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 TE 160 Allopolyploidization causes immediate nonrandom elimination of specific noncoding, lowcopy, and high-copy DNA sequences. These sequences are present in all the diploid species of Aegilops and Triticum but occur in only one pair of chromosomes (chromosome-specific sequences) or in several chromosome pairs of one genome (genome-specific sequences) at the polyploid level (Fig. 7.2) (Feldman et al. 1997; Liu et al. 1998a; Liu et al. 1998b; Ozkan et al. 2001; Shaked et al. 2001; Han et al. 2003; Salina et al. 2004; Han et al. 2005). The extent of DNA elimination was estimated by the determination of the nuclear DNA amount in natural allopolyploids and in their diploid progenitors as well as in newly synthesized allopolyploids and in their parental plants (Ozkan et al. 2003; Eilam et al. 2008; Eilam et al. 2010). Natural wheat allopolyploids contain 2–10 % less DNA than the sum of their diploid parents, and synthetic allopolyploids exhibit a similar loss, indicating that DNA elimination occurs soon after allopolyploidization (Eilam et al. 2008; Eilam et al. 2010). Also, from the very little variation in DNA amount that exists at the intraspecific level, it was concluded that the reduction of DNA content occurred immediately after the formation of the polyploids, and that after this there was almost no change in DNA amount during the life of the allopolyploid species. In triticale (an allopolyploid between wheat and rye, Secale cereale), Boyko et al. (1984, 1988) found that there was a great reduction in DNA content in the course of triticale formation with about 9 % for octoploid and 28–30 % for hexaploid triticale. The different genomes were not affected equally in triticale; wheat genomic sequences were relatively conserved, whereas rye genomic sequences were predominantly involved in a very high level of variation and elimination (Ma et al. 2004; Ma and Gustafson 2005, 2006). EC 159 CO RR 158 7.3.1 Revolutionary changes UN 157 D • Rewiring of gene expression through novel inter- • genomic interactions • New dosage response (positive, negative, dosage • compensation) • Gene suppression or activation • • Transcriptional activation of transposons (that may affect nearby genes in cis) • New transpositions of transposons Layout: T1 Standard SC Chapter No.: 7 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 116/134 M. Feldman et al. D PR OO F Editor Proof 116 184 185 186 187 188 189 190 191 192 193 194 195 196 Also in hexaploid wheat the genomes were not affected equally; genome D underwent considerable reduction in DNA amount whereas the wheat A and B genomes did not shrink in size (Eilam et al. 2008; Eilam et al. 2010). DNA elimination seems to be nonrandom at the intrachromosomal level as well: Liu et al. (1997) found in allohexaploid wheat that the chromosome-specific sequences of chromosome arm 5BL are not distributed at random along this chromosome arm but cluster in terminal (subtelomeric), subterminal, and interstitial regions of this arm, making these regions extremely chromosome-(homologous-)specific. Hence, it was tempting to suggest that these homologous-specific regions play a critical role in homology search and initiation of meiotic pairing (the classical pairing initiation sites) (Feldman et al. 1997). In some studies, DNA elimination was not observed in the first generations of synthetic allopolyploids (Mestiri et al. 2010; Zhao et al., 2011). It is possible that the timing, as well as the extent of elimination depends on the type of genomic combinations involved or on the type of sequences analyzed. CO RR 183 UN 182 EC TE Fig. 7.2 Schematic representation of the wheat karyotype. The wheat karyotype is arranged into genomes A, B, and D and into seven homoeologous groups. Examples of the different types of sequences are drawn on top of the chromosomes, namely: triplicated sequences (group-specific sequences), chromosome-specific sequences (CSSs) that are present in only one chromosome pair, genome-specific sequences (GSSs) that can be on more than one chromosome pair but only in one of the genomes, and dispersed repeats that are present on both homoeologous and nonhomoeologous chromosomes. Adapted from Levy and Feldman (2004) Layout: T1 Standard SC Chapter No.: 7 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 117/134 7 Genomic Plasticity in Polyploid Wheat 117 Editor Proof Table 7.5 Evolutionary changes facilitated by allopolyploidy in wheat Level Type • Chromosomal repatterning (intra- and inter-genomic translocations) • Introgression of chromosomal segments from alien genomes and production of recombinant genomes Functional • Nonfunctionalization (deletion or pseudogenization) • Subfunctionalization • Neofunctionalization • Copy number changes • New allelic variation 202 203 204 205 206 207 208 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 D 201 TE 200 Several structural changes occur in the allopolyploid wheat genomes, which generate new variants that could not take place in the diploid parental genomes and that occur almost exclusively in an allopolyploid background (Table 7.5). This includes events, such as intergenomic horizontal transfer of chromosomal segments, repetitive sequences, transposons, or genes among the constituent genomes. These events may occur sporadically throughout the history of the allopolyploid species. Intergenomic translocations that are population- or biotype-specific are widespread in allohexaploid wheat (Maestra and Naranjo 1999). Invasion of the A genome by sequences from the B genome—most probably transposons—was detected in wild allotetraploid wheat using GISH (Belyayev et al. 2000). The possibility of intergenomic transfer adds to the allopolyploid genomes’ plasticity and enables the creation of new genetic combinations that are beyond the addition of two genomes. Moreover, in contrast to diploids, which are genetically isolated from each other and have undergone divergent evolution, allopolyploids in the wheat group exhibit convergent evolution, because they contain genetic material from two or more different diploid genomes and can exchange genes with each other via hybridization and introgression, resulting in the production of new genomic combinations. Examples of such introgression between allotetraploid Aegilops species that share one genome and differ in the other genome(s) were provided by Zohary and Feldman (1962) and Feldman (1965a, b, c). Such hybridizations are eased by the shared genome, which acts as a buffer and ensures some fertility in the resulting hybrids. In such hybrids, chromosomes of two dissimilar genomes, brought together from different parents, may pair and exchange genetic material and recombine (Feldman 1965a). Additional evidence for the existence of introgressed genomes in allopolyploid Aegilops was obtained from C-banding analysis (Badaeva et al. 2004). Introgression of a DNA sequence from allopolyploid wheat to the allotetraploid Aegilops species, Ae. peregrina, was described by Weissmann et al. (2005). In addition to evolutionary changes that are almost unique to an allopolyploid background, it might be that other types of mutations that can cause structural or functional changes on an evolutionary scale (e.g., point mutations, satellite EC 199 CO RR 198 7.3.2 Evolutionary Changes UN 197 PR OO F Structural Layout: T1 Standard SC Chapter No.: 7 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 118/134 M. Feldman et al. 240 7.4 Functional Changes 232 233 234 235 236 237 238 PR OO 231 F 239 instability, transposition, etc.) can occur in an accelerated manner in the polyploid background. The presence of duplication or triplication of the genetic material in wheat allopolyploids might have relaxed constraints on gene structure and function. Thus, the accumulation of genetic variation through mutations or hybridization might be tolerated more readily in allopolyploid than in diploid species. While there is no direct evidence for this assertion, there is indirect support from experimental data showing a higher resistance of allohexaploid wheat to irradiation compared to the diploid progenitors (Mac Key 1954, 1958; Sears 1972). Such increase in resistance to mutation with increased ploidy level was shown to be correlated with increased evolvability and fitness in yeast (Thompson et al. 2006). 230 254 7.4.1 Functional Diploidization 246 247 248 249 250 251 252 255 256 257 258 259 260 261 262 263 264 265 266 TE 244 245 EC 243 CO RR 242 D 253 Allopolyploidy affects gene or protein function through a variety of mechanisms (Table 7.4). It has been widely suggested that following polyploidization, individual genes follow one of many possible evolutionary fates including nonfunctionalization (deletion or pseudogenization), neofunctionalization (evolution of novel functions among alleles or homoeoalleles), generating new phenotypes, or subfunctionalization (evolution of partitioned ancestral functions among alleles or homoeoalleles) (Lynch and Force 2000; Prince and Pickett 2002; Chaudhary et al. 2009). Functional changes are often regulated by genetic and epigenetic interactions among homoeoalleles and might provide the plasticity that is required to improve the fitness and adaptation of the newly formed allopolyploid and to increase its competitive efficiency with its parental species as well as other plant species, leading to its successful establishment in nature. Moreover, as discussed above, polyploidization seems to facilitate gene evolvability. 241 Increased gene dosage may lead to redundancy or in some cases may have a deleterious effect, for example, due to the formation of an unbalanced system (Veitia et al. 2008). Regulating gene action in a duplicated genome might be achieved through dosage compensation (nonlinear gene dosage response). A regulatory process that can bring redundant or unbalanced gene systems in polyploids toward a diploid-like mode of expression is functional diploidization (Ohno 1970). Functional diploidization is the process whereby existing genes in multiple doses can be eliminated, become inactive via mutations (base substitution, insertion, or deletion leading to pseudogenization) or diverted to new functions. Examples of functional diploidization in polyploid wheat involve mainly genes that code for structural or storage proteins, e.g., histones, subunits of tubulins, UN Editor Proof 118 Layout: T1 Standard SC Chapter No.: 7 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 119/134 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 293 294 295 296 297 298 299 300 301 302 303 304 305 306 307 308 309 310 311 F 272 PR OO 271 D 270 TE 269 119 subunits of glutenins and gliadins, and ribosomal RNA (and possibly also tRNA). In such genes, expression of all homoeoalleles might be redundant and even deleterious, due to overproduction and inefficiency. Also, activity of all homoeoalleles may produce intermediate phenotypes in several traits that decrease the viability of the plants. In this case, traits controlled by genes from only one genome may have a higher adaptive value. It is, therefore, expected that such gene loci would have been targets for genetic diploidization. In hexaploid wheat, intergenomic suppression, as seen by the disappearance of a storage protein subunit, was observed immediately upon formation of a wheat allohexaploid (genome BBAADD; see Galili and Feldman 1984; Galili et al. 1986). This is a common way to instantaneously reduce the negative effect of overproduction and inefficiency of genes that exist in super-optimal dose. Fascinatingly, suppression was reversible: the storage protein reappeared upon extraction of the tetraploid BBAA genomes and disappeared when the D genome was added. Similarly, attempts to transfer a leaf-rust resistance gene from tetraploid to hexaploid wheat failed because of a suppressor gene that was mapped to the D genome (Kerber and Green 1980). Intergenomic suppression of disease resistance genes is a common phenomenon as was noticed in natural and in several newly formed allopolyploids (Anikster, Y., Manisterski, J., and Feldman, M., unpublished data). Comparable results were obtained by Dhaliwal and co-workers (Aghaee-Sarbarzeh et al. 2001) who found that in Triticum durum-Aegilops amphiploids dominant leaf rust and stripe rust resistance genes from the Aegilops parents were suppressed by genes on the AB genomes of the wheat parent. Another well-studied example of intergenomic suppression is the silencing of rye ribosomal RNA genes in the presence of the wheat genome. Cytosine methylation is involved in this silencing, as suggested by reactivation of the rye ribosomal RNA genes upon treatment with 5-aza-cytidine and analyzed by the use of methylation sensitive/insensitive isoschizomers (Houchins et al. 1997). Recent analysis of the sequences of wheat group 1 chromosomes shows significant deviations from synteny with many of the nonsyntenic genes representing pseudogenes (Wicker et al. 2011). Part of this pseudogenization might have occurred after polyploidization as suggested from the analysis of the Q-factor. A recent study shows that a combination of mutations in Q genes contributed to the domestic spike phenotype, namely nonfragile, soft glumes, and free threshing (Zhang et al. 2011). The mutation with the most significant phenotypic effect is an amino acid substitution in the protein coded by the 5A locus, while other mutations, such as pseudogenization of the locus on 5B or subfunctionalization of the locus on 5D, also contributed to the domestication phenotype, but to a lesser extent (Zhang et al. 2011). Remarkably, these mutations occurred after polyploidization. The Hardness (Ha) gene constitutes another example of genetic diploidization, through gene deletion, in polyploid wheat (Chantret et al. 2005). Genetic diploidization might also be achieved through epigenetic control. Epigenetic silencing can be brought about via cytosine methylation of DNA sequences (Kashkush et al. 2002) or chromatin modifications or remodeling as well as the activity of small RNA molecules (Kenan-Eichler et al. 2011). Gene EC 268 CO RR 267 UN Editor Proof 7 Genomic Plasticity in Polyploid Wheat Layout: T1 Standard SC Chapter No.: 7 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 120/134 314 315 316 317 318 319 320 silencing can also be achieved via novel regulatory interactions such as intergenomic suppression (Galili and Feldman 1984). Shaked et al. (2001) and Kashkush et al. (2002) reported the occurrence of alterations in cytosine methylation in 13 % of the loci affecting both repetitive DNA sequences and low-copy DNA in approximately equal proportions. Changes in microRNAs, such as miR168 which targets the Argonaute1 gene, were shown to occur in newly synthesized wheat (Kenan-Eichler et al. 2011). F 313 PR OO 312 M. Feldman et al. 7.4.2 Subfunctionalization Through Partitioning and Compensation of Duplicated Gene Expression 338 7.4.3 New Interactions Between Maintained Genes/Proteins 326 327 328 329 330 331 332 333 334 335 336 339 340 341 342 343 344 345 346 347 TE 324 325 EC 323 CO RR 322 D 337 Functional diversification of duplicated genes (subfunctionalization), i.e., differential or partitioning of expression of homoeoalleles in different tissues and/or in different developmental stages, is also a form of genetic diploidization (Adams et al. 2003). Subfunctionalization, an important aspect of allopolyploidy, has been studied relatively little in allopolyploid wheat. Koebner and co-workers (Bottley et al. 2006) found that differential expression of homoeoalleles in different plant tissues is common in hexaploid wheat. The activity of silenced genes could be restored in aneuploid lines, suggesting that no mutation was involved but rather new cis–trans interactions or reversible epigenetic alterations were responsible. The data of Bottley et al. (2006) suggest that for leaf transcripts, there is a modest bias toward silencing of the D genome copies, but this pattern does not extend to root transcripts. Mochida et al. (2006) also presented evidence for differential expression of homoeoalleles in wheat and suggested that inactivation of homoeoalleles is a nonrandom effect. The molecular basis for these cases of instant subfunctionalization remains to be determined. Similarly, subfunctionalization of all three homoeologs of the Q-locus in hexaploid wheat was recently described (Zhang et al. 2011). 321 In allohexaploid wheat many gene loci exist in triplicate dose, and the extra gene dosage per se may produce improved or even novel traits. Homoeoalleles may differ from one another by allelic variation, and in this case, activity of all the duplicated genes may produce desirable intergenomic interactions and heterotic effects. These are mainly genes that code for functional proteins (enzymes). This occurrence of enzyme diversity (isozymes) increases the biochemical potential of the allopolyploids. In fact, in wheat allopolyploids the expression of most homoeoalleles coding for functional enzymes is retained (Mitra and Bhatia 1971; Hart 1983a, b, 1987). Also, intergenomic gene interactions may be, in some cases, UN Editor Proof 120 Layout: T1 Standard SC Chapter No.: 7 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 121/134 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 F 354 PR OO 353 D 351 352 TE 350 121 expressed in novel traits that do not exist in their parental diploids. Some of these traits may have great adaptive value. Intergenomic gene interactions have direct relevance also to wheat cultivation. For example, the baking quality of allohexaploid wheat (bread wheat) is due to the unique properties of its gluten—a product derived from the combined contribution of the three genomes of hexaploid wheat, and thus exists only at the hexaploid level. In addition, the combination of a large number of spikelets per spike derived from T. urartu (the donor of the A genome), with several fertile florets per spikelet originating from the donors of the B genome, facilitated the high fertility of durum and bread wheat. Likewise, merging of Ae. tauschii (the donor of the D genome) displaying the cold hardiness phenotype with the prolific nature of allotetraploid wheat (the donor of the A and B genomes) enabled the expansion of wheat cultivation into colder regions. Enzyme multiplicity, derived from the activity of all homoeoalleles, increases the ability of the allopolyploid to adapt to a wider range of environments. This might account for the very wide distribution of wheat under cultivation—much wider than that of any other cultivated plant. Genome-wide reprograming in gene expression may occur in a polyploid as a result of new interactions among regulatory factors of the parents. For example, novel interactions between the trans factor from one species and the cis or trans factors of the other parental species, as was shown in yeast in a cross between wide hybrids (Tirosh et al. 2009), may account for the observed cases of gene repression or for cases of activation via overexpression. The consequences of allopolyploidy on gene expression have been widely studied at the genome-wide level in several natural and synthetic allopolyploids of Triticum. The majority of studies relied on the comparison of the expression level in the allopolyploid to those of its parents and/or to the average of its parents, expressed as the mid-parental value (MPV). In hexaploid wheat, Pumphrey et al. (2009) found that approximately 16 % of the 825 analyzed genes displayed nonadditive expression in the first generation of synthetic hexaploid wheat. Chagué et al. (2010) analyzed 55,052 transcripts in two lines of synthetic allohexaploid wheat and found that 7 % of the genes had nonadditive expression, while Akhunova et al. (2010) found in synthetic allohexaploid wheat that about 19 % of the studied genes showed nonadditive expression. Similar studies (He et al. 2003) showed that the expression of a significant fraction of genes (7.7 %) was altered in the synthetic allohexaploid T. turgidum-Ae. tauschii, and that Ae. tauschii genes were affected much more frequently than those of T. turgidum. Strikingly, these different studies show that deviation of gene expression in allopolyploids from their parents or average of parents appears to be a common feature, although evaluated in various allopolyploids and using different techniques and approaches. Interestingly, silencing of the same genes was also found in natural hexaploid wheat, i.e., in the variety Chinese Spring (He et al. 2003). Chagué et al. (2010) suggested, based on similar gene expression patterns observed between natural and synthetic wheat allohexaploids, that regulation of gene expression is established immediately after allohexaploidization and maintained over generations. It is of interest to note that several genes that are silent in EC 349 CO RR 348 UN Editor Proof 7 Genomic Plasticity in Polyploid Wheat Layout: T1 Standard SC Chapter No.: 7 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 122/134 M. Feldman et al. 407 7.5 Genome asymmetry 400 401 402 403 404 405 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 PR OO 399 D 398 The rapid processes of cytological and genetic diploidization allow for the development and occurrence of two contrasting and highly important genetic phenomena in allopolyploid wheat that contribute to the evolutionary success of these polyploids: (1) build up and maintenance of enduring intergenomic favorable genetic combinations, and (2) genome asymmetry in the control of a variety of morphological, physiological, and molecular traits, i.e., complete or principal control of certain traits by only one of the constituent genomes. However, while the first phenomenon was taken for granted by plant geneticists, genomic asymmetry in interspecific hybrids and allopolyploids was mainly known in ribosomal RNA genes (reviewed in Pikaard 2000) and only recently has also been documented in other traits (Peng et al. 2003a, b; Fahima et al. 2006; Feldman and Levy 2009; Flagel et al. 2009; Rapp et al. 2009; Flagel and Wendel 2010). The phenomena of diploidization and of dominance in gene expression lead to genome asymmetry, which is manifested in a clear-cut division of tasks among the constituent genomes of allopolyploid wheat (Levy and Feldman 2004; Feldman and Levy 2009; Feldman et al., in press). Genome A controls morphological traits while genome B in allotetraploid wheat and genomes B and D in allohexaploid wheat control the reaction to biotic and abiotic factors (Tables 7.6 and 7.7). Intergenomic pairing would lead to both disruption of the linkage of the homoeoalleles that contribute to positive intergenomic interactions and segregation of genes that participate in the control of certain traits by a single genome. Intergenomic recombination may, therefore, result in many intermediate phenotypes that may affect, in a negative manner, the functionality, adaptability, and stability of the allopolyploids. TE 396 397 EC 395 CO RR 394 F 406 the parental species became active in the newly formed allohexaploid (He et al. 2003). Similarly, cDNA-AFLP gels also revealed several cDNAs that were expressed only in the allopolyploids and not in the diploid progenitors (Shaked et al. 2001; Kashkush et al. 2002). The proportion of genes for which expression in the polyploids is different from the average of the parents is underestimated in these studies, as most of the technologies used do not allow separate analysis of the expression level of each of the gene copies (homoeoalleles) and their respective contribution to overall gene expression (Pumphrey et al. 2009; Chagué et al. 2010; Akhunova et al. 2010). It is not possible to detect, for example, a situation where repression of one homoeoallele is compensated by the activation of the other using microarray technologies. The progress towards sequencing of wheat genomes and the promise of next-generation technologies should allow in the future the resolution of expression at the level of individual homoeologues. 393 UN Editor Proof 122 Layout: T1 Standard SC Chapter No.: 7 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 123/134 7 Genomic Plasticity in Polyploid Wheat 123 Glumes with keels The shape of the edge of the glumes (beaked glumes) Hairs at the base of every spikelet Plant habitus Growth habit Autogamous behavior Many domestication genes 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 D In the preceding sections, we have discussed structural changes and changes in gene expression resulting from genetic or epigenetic phenomena. Transposable elements (TEs) are discussed here separately due to their specific mode of action, their abundance (up to 90 % of the wheat genomes; Sabot et al. 2005), and their impact on both structure and expression of the genome. TE 435 EC 434 7.6.1 Transcriptional Activation of Retrotransposons in Synthetic Wheat Polyploids CO RR 433 Higher polymorphism of HMW glutenin genes Larger amount of repetitive sequences Activity on nucleolar organizers Larger number of rRNA genes 7.6 Response of Transposable Elements to Allopolyploidization It is now clear that some eukaryote retrotransposon promoters retain activity under normal conditions and initiate either read-in transcripts of the transposon itself or read-out transcripts into flanking host sequences (Vicient et al. 2001; Kashkush et al. 2002; Nigumann et al. 2002; Kashkush et al. 2003; Kashkush and Khasdan 2007). Following allopolyploidization events in wheat, the steady-state level of expression of LTR retrotransposons was massively elevated (Kashkush et al. 2002, 2003, unpublished data), similarly to what was observed in synthetic Arabidopsis allopolyploid hybrids (Madlung et al. 2005). In addition, the transcriptional activity of a LTR element termed Wis2-1A (Lucas et al. 1992) leads to the production of read-out transcripts toward flanking host DNA sequences, a process that occurred in a genome-wide manner (Kashkush et al. 2003). In many cases, these read-out transcripts were associated with the expression of adjacent genes, depending on their orientation: knocking down or knocking out the gene product if the read-out transcript was in the antisense orientation relative to the orientation of UN 432 Regulation of ecological adaptation Double the number of disease resistance genes Contains more stress–related genes? Higher polymorphism of molecular markers F Inflorescence morphology Free caryopsis PR OO Editor Proof Table 7.6 Genome asymmetry in the control of various traits in the wild allotetraploid wheat, T. turgidum subsp. dicoccoides (genome BBAA) Traits under control of Genome A Traits under control of Genome B Layout: T1 Standard SC Chapter No.: 7 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 124/134 M. Feldman et al. Table 7.7 Genome asymmetry in the control of agronomic traits in domesticated durum (genome BBAA) and bread wheat (genome BBAADD)a Traits Traits under control of br B1 on 3BS tg2 on 2BS PR OO Eg P2 on 7BL (?) Rht B1 on 4BS; Rh4 on 2Bl; Ga1, Ga3 on 4BS W1 on 2BS W1I on 2BS; W3I on 1BL ms1 on 4BS Ph1 on 5BL Ne1 on 5BL; Ne2 on 2BS EC CO RR Hybrid chlorosis Ch1 on 2A Aluminum tolerance Boron tolerance Low cadmium uptake Iron deficiency Herbicide response Difenzoquat insensitivity Chlortoluron insensitivity Imi3 on 6AL Imidazolinone resistance Response to photoperiod Response to vernalization Vrn-A1 on 5AL Response to salinity Frost resistance Number of resistance genes to diseases and pests a Fr1 on 5AL 45 tg1 on 2DS Rht D1 on 4DS; Rht8 on 2DL; Rht5 on 3BS; Rht9 on 7BS; Rht13 on 7BS Gpc B1 on 6BS Pro1 on 5DL; Pro2 on 5Ds Ha on 5DS Pin D1 on 5DS TE Grain protein content Grain hardness Puroindolines and grain softness protein Gibberellic acid response Waxiness Epistatic inhibitors of waxiness Male sterility Ms3 on 5AS; ms5 on 3A Pairing homoeologous Hybrid necrosis Genome D F Genome B D Genome A Inflorescence morphology Elongated glumes Eg P1 on 7AL Branched spikes Bh on 2AS Nonbrittle rachis br A1 on 3AS br A2 on 2A Nontenacious glume (lax glume) Reduce plant height Rht7 on 2A; Rht12 on 5AL; UN Editor Proof 124 Bo1 on 7BL Cdu1 on 5BL Fe2 on 7BS Dfg 1 on 2BL Su1 on 6BS Imi2 on 6BL Ppd-B1 on 2BS Vrn-B1 on 5BL; Vrn-B3 on 7BS 88 Ga2 on 4DS W2 on 2DS (?) W2I on 2DS Ms2 on 4DS; Ms4 on 4DS Ph2 on 3DS Ch2 on 3DL Alt2 on 4DL Fe1 on 7DL Imi1 on 6DL Ppd-D1 on 2DS Vrn-D1 on 5DL; Vern-D4 on 5DL;Vern-D5 on 5DL Kna1 on 4DL Fr2 on 5DL 51 Data from the 2008 Wheat Gene Catalogue (http://wheat.pw.usda.gov/GG2/index.shtml) Layout: T1 Standard SC Chapter No.: 7 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 125/134 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 F 460 PR OO 459 7.6.2 Massive Methylation of TE-Adjacent DNA Sequences Following Allopolyploidization D 458 Alterations in the genomic methylation patterns following allopolyploidization have been examined in several polyploid systems, including Arabidopsis (Madlung et al. 2002; Belzile et al. 2009), Spartina (Salmon et al. 2005; Parisod et al. 2009), Brassica (Lukens et al. 2006; Wang et al. 2009), and wheat (Shaked et al. 2001). The methylation alterations are either hyper- or hypomethylation, depending on the sequence analyzed, and are reproducible. Recent studies in wheat have investigated in detail the methylation of CCGG sites flanking several TE families (Kraitshtein et al. 2010; Yaakov and Kashkush 2011a, b; Zhao et al. 2011). In one study (Kraitshtein et al. 2010), transposon methylation display (TMD) analysis was applied (see Kashkush and Khasdan 2007) to analyze a terminal-repeat retrotransposon in miniature (TRIM), termed Veju, in Triticum turgidum ssp. durum (genome AABB) and Aegilops tauschii (genome DD), and the first four generations of the derived allohexaploid. It was estimated that over 50 % of the CCGG sites flanking Veju elements showed altered TMD patterns in the first four generations of the newly formed allohexaploid. Hypomethylation of Veju-flanking CCGG sites was predominant in the first generation of the newly formed allohexaploid, while hypermethylation was predominant in subsequent generations. This might indicate reduced Veju transcriptional activity after the third generation of the synthetic allohexaploid. A similar pattern of hypomethylation of Veju elements was also observed in the first three generations of a synthetic allotetraploid that was derived from a cross between Ae. sharonensis (genome SlSl) and T. monococcum ssp. aegilopoides (genome AmAm) (Yaakov and Kashkush 2011b). However, unlike in the synthetic allohexaploid, Veju elements remained hypomethylated up to the fourth generation of the synthetic allotetraploid. In support of these studies, Zhao et al. (2011) observed massive methylation changes around Veju in three different combinations of newly formed wheat allohexaploids. TE 457 125 the gene transcript (such as the iojap-like gene) or overexpressing the gene if the read-out transcript was in the sense orientation (such as the puroindoline-b gene). The mechanisms by which transcriptional activation of TEs influences the expression of neighboring genes are poorly understood. In some cases, the correlation between the reduction of the sense expression of the gene and the production of the antisense strand that initiated from the adjacent transposon promoter (Kashkush et al. 2003; Puig et al. 2004) might indicate that post-transcriptional gene silencing is a major mechanism for inactivating adjacent genes. Recent studies on tracking methylation changes around a LTR retrotransposon in the first four generations of a newly formed wheat allopolyploid (Kraitshtein et al. 2010) may indicate that this read-out activity is restricted to the first generations of the nascent polyploid species. EC 456 CO RR 455 UN Editor Proof 7 Genomic Plasticity in Polyploid Wheat Layout: T1 Standard SC Chapter No.: 7 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 126/134 M. Feldman et al. 507 7.6.3 Changes in TE Composition Following Allopolyploidization 501 502 503 504 505 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 PR OO 500 The prevalence of TEs and their inherent sequence similarity make them a prime target for illegitimate and nonhomologous recombination. TEs have been shown to undergo rearrangements following allopolyploidization in Spartina (Parisod et al. 2009), tobacco (Petit et al. 2010), and Triticale (Bento et al. 2008). Recent data for synthetic allohexaploid wheat indicate that rearrangements of retrotransposoncontaining sequences occur rapidly and reproducibly (repeated in independently newly formed allopolyploid lines) in the first generations following polyploidization. In addition, a change in the methylation status (usually hypomethylation) in the first generation was followed by deletion of retrotransposon-containing sequences in subsequent generations (Kraitshtein et al. 2010). These data suggest a correlation between methylation and post-allopolyploidization rearrangements that occur via a mechanism that has yet to be identified. One possible explanation is that hypomethylation confers an open chromatin structure to the TE sequences, which exposes these demethylated elements to be targeted for deletion by the host. There is evidence that small RNAs corresponding to Veju elements might play a pivotal role in Veju methylation in the newly formed wheat allohexaploid (KenanEichler et al. 2011). Despite the altered methylation status and transcriptional activation of TEs following allopolyploidization, there are very few reports on the transpositional activity of transposons. Madlung et al. (2005) showed both methylation alterations and limited transpositional activation of a Sunfish transposon in polyploid Arabidopsis. Petit et al. (2010) reported an increase in the copy number of a Tnt1 retrotransposon in allotetraploid tobacco. No transposition bursts were reported in Spartina (Parisod et al. 2009) or in wheat (Kashkush et al. 2003; Kraitshtein et al. 2010; Yaakov and Kashkush 2011a). These reports suggest that the transpositional activity of TEs following allopolyploidization might be restricted to specific TE families (Parisod et al. 2010). Recently, it was shown that the immense loss of Veju sequences in the first generation of the synthetic allohexaploid is probably D 499 TE 498 EC 497 CO RR 496 F 506 The methylation patterns of three TEs [Balduin (belonging to the CACTA superfamily), Apollo (belonging to the MuDR/Foldback superfamily), and Thalos (a stowaway-like MITE belonging to the Tc1/mariner superfamily)] have also been analyzed in allopolyploid wheat. The CCGG sites flanking the three elements underwent massive hypermethylation in the first four generations of the synthetic allohexaploid (Yaakov and Kashkush 2011a, b), while they underwent massive hypomethylation in the first four generations of the synthetic allotetraploid (Yaakov and Kashkush 2011b). The massive hypermethylation of these elements in the synthetic allohexaploid might be connected to the lack of transpositional activity (Yaakov and Kashkush 2011a). It is important to mention that transcriptional activation of LTR retrotransposons does not correlate with transpositional activity (Kashkush et al. 2003). 495 UN Editor Proof 126 Layout: T1 Standard SC Chapter No.: 7 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 127/134 127 552 7.7 Concluding Remarks 543 544 545 546 547 548 549 550 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 PR OO 542 D 541 TE 539 540 The studies reviewed above indicate that, in the wheat group, hybridization and chromosome doubling induce a burst of genomic alterations, some of which could not occur at the diploid level. Some of these changes might improve the ability of the newly formed allopolyploids to survive in nature and to compete with their parental species, corresponding thus to phenotypic and adaptive novelty. Other changes are probably deleterious. TEs seem to play an important role in the various responses to hybridization and polyploidization due to their abundance and also due to their tendency to be dysregulated as a result of genomic shocks. The balance between the beneficial and deleterious changes associated with allopolyploidization is probably what determines the fate of the nascent species. The formation of an allopolyploid species is accomplished rapidly via the combined processes of hybridization and genome doubling, but its establishment in nature as a successful species probably depends on a high degree of plasticity that enables it to overcome potential incompatibilities and to gain new traits. The studies reported here suggest that wheat can achieve genomic plasticity through the induction of a series of cardinal nonadditive genomic changes. Some changes, genetic and epigenetic, are rapid and nonMendelian, occurring during or immediately after the formation of the F1 hybrid or the allopolyploid (revolutionary changes). Other changes occur sporadically over a long time period during the life of the allopolyploid species (evolutionary changes). From a population point of view, the chance of a new individual, such as a nascent hybrid/allopolyploid, to establish itself as a new species is almost nil, unless it has some increased fitness over its parents. This fitness advantage must be manifested within a few generations of formation or the new species will rapidly be extinct. The revolutionary EC 538 CO RR 537 F 551 followed by retrotransposition in subsequent generations, a process that causes new insertions to accumulate in allohexaploids (Kraitshtein et al. 2010). The same study also suggests that these new insertions are targeted for methylation. Methylation of the new Veju elements protects the genome from deleterious transposon insertions. Investigating the scale of eliminated DNA sequences, including TE sequences, by identifying the deletion breakpoints will allow a better understanding of the mechanism(s) involved and of the nature of the connection between methylation and rearrangements. In summary, different classes of wheat TEs appear to respond differently to the allopolyploidization event. Table 7.8 summarizes the type of response of 12 different class I and class II elements. It can be seen clearly that epigenetic response, mainly methylation changes, is a common factor for all studied TEs, while genetic response that includes rearrangements and/or transpositions is restricted to specific TE families. Transcriptional data for most studied TEs are still lacking (Table 7.8). However, there is a good basis for suggesting a connection among methylation changes in TEs with alteration in their expression patterns. 536 UN Editor Proof 7 Genomic Plasticity in Polyploid Wheat Layout: T1 Standard SC Chapter No.: 7 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 128/134 128 M. Feldman et al. Editor Proof Table 7.8 Summary of the genetic and epigenetic responses of several TE families to allopolyploidization in wheat Reference Family TE Genetic and epigenetic alterations in synthetic allopolyploids (compared to parental lines)a Methylation Transcription Rearrangementb Transpositionc 4 4 4 Copia Angela BARE1 Wis2-1 A 4 4 NA NA NA 9 NA 4 9 NA NA 9 9 MuDR/ Foldback Apollo 4 NA 4 NA a b c 578 579 580 581 582 583 584 585 4 9 9 9 9 9 9 4 4 4 4 ? 4 ? ? EC NA NA NA NA Kraitshtein et al. (2010) KenanEichler et al. (2011) Zhao et al. (2011) Unpublished Zhao et al. (2011) Kashkush et al. (2003) Unpublished Yaakov and Kashkush (2011a) Yaakov and Kashkush (2011a) Yaakov and Kashkush (2011a) Unpublished Unpublished Unpublished Unpublished 4 altered, 9 no change, ? not validated, NA data not available Include deletion and/or insertion Typical TE transposition changes described here may contribute to the establishment of the new species. Instantaneous elimination of sequences from one genome in the newly formed allopolyploids increases the divergence of the homoeologous chromosomes, and thus leads to exclusive intragenomic pairing that improves fertility. Mechanisms such as loss of deleterious genes (e.g. genetic incompatibilities) or positive dosage effects or new intergenomic heterotic interactions may all rapidly increase the fitness of the nascent species. The evolutionary changes, on the other hand, contribute to the build-up of genetic variability and thus increase adaptability, fitness, competitiveness, and colonizing ability. It is clear that most hybridization events in UN 577 4 4 CO RR Stowaway- Thalos like MITE Eos Minos Oleus Fortuna 9 D Sabrina 4 Balduin 4 9 9 TE Gypsy CACTA ? F Veju PR OO TRIM Layout: T1 Standard SC Chapter No.: 7 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 129/134 129 610 References 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 Aaronsohn A (1910) Agricultural and botanical explorations in Palestine. Bull Plant Ind 180:1–63 Aaronsohn A, Schweinfurth G (1906) Die auffindung des wilden emmers (Triticum dicoccum) in Nordpalästina. Altneuland Monatsschrift für die irtschaft. Erschliessung Palästinas 7(8):213–220 Adams KL, Cronn R, Percifield R, Wendel JF (2003) Genes duplicated by polyploidy show unequal contributions to the transcriptome and organ-specific reciprocal silencing. Proc Natl Acad Sci U S A 100(8):4649–4654 Aghaee-Sarbarzeh M, Dhaliwal HS, Harjit-Singh (2001) Suppression of rust resistance genes from distantly related species in Triticum durum-Aegilops amphiploids. In: Johnson R, Yahyaoui A, Wellings C, Saidi A, Ketata H (eds) Meeting the challenge of yellow rust in cereal crops. Proceedings of the First Regional Conference on Yellow Rust in the Central and West Asia and North Africa Region, Karaj, Iran. pp 8–14 Akhunova AR, Matniyazov RT, Liang H, Akhunov ED (2010) Homoeolog-specific transcriptional bias in allopolyploid wheat. BMC genomics 11:505 Avivi L (1976) The effect of genes controlling different degrees of homoeologous pairing on quadrivalent frequency in induced autotetraploid lines of Triticum longissimum. Can J Genet Cytol 18:357–364 Badaeva ED, Amosova AV, Samatadze TE, Zoshchuk SA, Shostak NG, Chikida NN, Zelenin AV, Raupp WJ, Friebe BR, Gill BS (2004) Genome differentiation in Aegilops. 4. Evolution of the U-genome cluster. Plant Syst Evol 246:45–76 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 PR OO 591 D 590 TE 589 EC 588 CO RR 587 F 609 nature do not lead to the formation of a new species, but remarkably, the wheat group is equipped with a battery of molecular mechanisms that provide the potential for phenotypic novelty and for successful speciation to occur. Future work should enable better understanding of the role of specific genes and DNA sequences in speciation, the mechanisms that confer robustness of the genome to the shock of allopolyploidy and to the activation of TEs, and the mechanisms that enable the orchestration of chromosome division and the control of bivalent pairing during meiosis. Altogether, the reported revolutionary and evolutionary genomic changes emphasize the dynamic plasticity of the wheat allopolyploid genome with regard to both structure and function. Presumably, these changes have improved the adaptability of the newly formed allopolyploids and facilitated their rapid colonization of new ecological niches. No wonder, therefore, that cultivated allopolyploid wheats exhibit a wider range of genetic flexibility than diploid wheats and could adapt themselves to a great variety of environments. In contrast to Stephens (1951), who had the insight that allopolyploidy might lead to new evolutionary opportunities, Stebbins (1971, 1980) stated that while polyploidy has been of great importance for the origin of species it has contributed little to progressive evolution. He assumed that polyploids evolve more slowly than their diploid relatives. Stebbins (1971, 1980) did not take into consideration that allopolyploidization triggers a burst of genomic alterations that are not feasible at the diploid level and that lead to new evolutionary opportunities. Allopolyploidy has proved to be a powerful evolutionary factor that has played a decisive role in the evolution of the wheat group. 586 UN Editor Proof 7 Genomic Plasticity in Polyploid Wheat Layout: T1 Standard SC Chapter No.: 7 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 130/134 EC TE D PR OO F Belyayev A, Raskina O, Korol A, Nevo E (2000) Coevolution of A and B genomes in allotetraploid Triticum dicoccoides. Genome 43(6):1021–1026 Belzile F, Beaulieu J, Jean M (2009) The allotetraploid Arabidopsis thaliana-Arabidopsis lyrata subsp petraea as an alternative model system for the study of polyploidy in plants. Mol Genet Genomics 281(4):421–435 Bento M, Pereira HS, Rocheta M, Gustafson P, Viegas W, Silva M (2008) Polyploidization as a retraction force in plant genome evolution: sequence rearrangements in Triticale. PLoS ONE 3:1402–1413 Blakeslee AF (1937) Redoublement du nombre de chromosomes chez les plantes par traitement chimique. Compt Rend Acad Sci Paris 205:476–479 Bottley A, Xia GM, Koebner RMD (2006) Homoeologous gene silencing in hexaploid wheat. Plant J 47(6):897–906 Boyko EV, Badaev NS, Maximov NG, Zelenin AV (1984) Does DNA content change in the course of triticale breeding. Cereal Res Commun 12(1–2):99–100 Boyko EV, Badaev NS, Maximov NG, Zelenin AV (1988) Regularities of genome formation and organization in cereals. I. DNA quantitative changes in the process of allopolyploidization. Genetika 24:89–97 Chague V, Just J, Mestiri I, Balzergue S, Tanguy AM, Huneau C, Huteau V, Belcram H, Coriton O, Jahier J, Chalhoub B (2010) Genome-wide gene expression changes in genetically stable synthetic and natural wheat allohexaploids. New phytol 187(4):1181–1194 Chaudhary B et al (2009) Reciprocal silencing, transcriptional bias and functional divergence of homeologs in polyploid cotton (gossypium). Genetics 182:503–517 Chantret N, Salse J, Sabot F, Rahman S, Bellec A, Laubin B, Dubois I, Dossat C, Sourdille P, Joudrier P, Gautier MF, Cattolico L, Beckert M, Aubourg S, Weissenbach J, Caboche M, Bernard M, Leroy P, Chalhoub B (2005) Molecular basis of evolutionary events that shaped the hardness locus in diploid and polyploid wheat species (Triticum and Aegilops). Plant Cell 17(4):1033–1045 Chapman V, Miller TE, Riley R (1976) Equivalence of the A genome of bread wheat and that of Triticum urattu. Genet Res 27:69–76 Dvorak J (1976) The relationship between the genome of Triticum urattu and the A and B genomes of Triticum aestivum. Can J Genet Cytol 18:371–377 Dvorak J (2009) Triticeae genome structure and evolution. In: Feuiller C, Muehlbauer GJ (eds) Genetics and genomics of the Triticeae, plant genetics and genomics: crops and models 7. Springer, Berlin. pp 685–711 Eilam T, Anikster Y, Millet E, Manisterski J, Feldman M (2008) Nuclear DNA amount and genome downsizing in natural and synthetic allopolyploids of the genera Aegilops and Triticum. Genome 51(8):616–627 Eilam T, Anikster Y, Millet E, Manisterski J, Sagi-Assif O, Feldman M (2010) Genome size in diploids, allopolyploids, and autopolyploids of mediterranean triticeae. doi:10.1155/2010/ 341380 Fahima T, Cheng JP, Peng JH, Nevo E, Korol A (2006) Asymmetry distribution of disease resistance genes and domestication synrome QTLs in tetraploid wheat genome. 8th International Congress of Plant Molecular Biology, Adelaide, Australia Feldman M (1965a) Chromosome pairing between differential genomes in hybrids of tetraploid Aegilops species. Evolution 19:563–568 Feldman M (1965b) Fertility of interspecific F1 hybrids and hybrid derivatives involving tetraploid species of Aegilops Section Pleionathera. Evolution 19:556–562 Feldman M (1965c) Further evidence for natural hybridization between tetraploid tetraploid species of Aegilops Section Pleionathera. Evolution 19:162–174 Feldman M (2001) The origin of cultivated wheat. In: Bonjean A, Angus W (eds) The wheat book. Lavoisier Tech and Doc, Paris, pp 1–56 Feldman M, Levy AA (2005) Allopolyploidy—a shaping force in the evolution of wheat genomes. Cytogenet Genome Res 109(1–3):250–258 CO RR 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 M. Feldman et al. UN Editor Proof 130 Layout: T1 Standard SC Chapter No.: 7 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 131/134 131 EC TE D PR OO F Feldman M, Levy AA (2009) Genome evolution in allopolyploid wheat—a revolutionary reprogramming followed by gradual changes. J Genet Genomics 36(9):511–518 Feldman M, Levy AA, Fahima T, Korol A (2012) Genomic asymmetry in allopolyploid plants wheat as a model. J. Exp. Bot. (in press) Feldman M, Liu B, Segal G, Abbo S, Levy AA, Vega JM (1997) Rapid elimination of low-copy DNA sequences in polyploid wheat: a possible mechanism for differentiation of homoeologous chromosomes. Genetics 147(3):1381–1387 Feldman M, Lupton FGH, Miller TE (1995) Wheats. In: Smartt J, Simmonds NW (eds) Evolution of crop plants, 2nd edn. Longman Scientific, London, pp 184–192 Flagel LE, Chen LP, Chaudhary B, Wendel JF (2009) Coordinated and fine-scale control of homoeologous gene expression in allotetraploid cotton. J Hered 100(4):487–490 Flagel LE, Wendel JF (2010) Evolutionary rate variation, genomic dominance and duplicate gene expression evolution during allotetraploid cotton speciation. New Phytol 186(1):184–193 Galili G, Feldman M (1984) Inter-genomic suppression of endosperm- protein genes in common wheat. Can J Genet Cytol 26:651–656 Galili G, Levy AA, Feldman M (1986) Gene-dosage compensation of endosperm proteins in hexaploid wheat Triticum aestivum. Proc Natl Acad Sci U S A 83:6524–6528 Griffiths S, Sharp R, Foote TN, Bertin I, Wanous M, Reader S, Colas I, Moore G (2006) Molecular characterization of Ph1 as a major chromosome pairing locus in polyploid wheat. Nature 439(7077):749–752 Gupta PK, Kulwal PL, Rustgi S (2005) Wheat cytogenetics in the genomics era and its relevance to breeding. Cytogenet Genome Res 109(1–3):315–327 Han FP, Fedak G, Guo WL, Liu B (2005) Rapid and repeatable elimination of a parental genomespecific DNA repeat (pGcIR-1a) in newly synthesized wheat allopolyploids. Genetics 170(3):1239–1245 Han FP, Fedak G, Ouellet T, Liu B (2003) Rapid genomic changes in interspecific and intergeneric hybrids and allopolyploids of Triticeae. Genome 46(4):716–723 Hart GH (1983a) Genetic and evolution of mulilocus isozymes in hexaploid wheat. In: Ratazzi MC, Scandalios JG, Whitt GS (eds) Isozymes: current topics in biological and medical research, vol 10., Genetics and Evolution Alan R. Liss., Inc., New York, pp 365–380 Hart GH (1983b) Hexaploid wheat (Triticum aestivum L. em Thell.). In: Tanksley SD, Orton TJ (eds) Isozymes in plant genetics and breeding, Part. B, Elsvier Science Publishers B.V., Amsterdam, pp 35–56 Hart GH (1987) Genetic and biochemical studies of enzymes. In: Heyne EG (ed) Wheat and wheat improvement, Second Ed., Amer. Soc. Agronomy, Madison, Wisconsin, USA He P, Friebe BR, Gill BS, Zhou JM (2003) Allopolyploidy alters gene expression in the highly stable hexaploid wheat. Plant Mol Biol 52(2):401–414 Houchins K, ODell M, Flavell RB, Gustafson JP (1997) Cytosine methylation and nucleolar dominance in cereal hybrids. Mol Gen Genet 255(3):294–301 Huang S, Sirikhachornkit A, Su X, Faris J, Gill B, Haselkorn R, Gornicki P (2002) Genes encoding plastid acetyl-CoA carboxylase and 3-phosphoglycerate kinase of the Triticum/ Aegilops complex and the evolutionary history of polyploid wheat. Proc Natl Acad Sci U S A 99(12):8133–8138 Kashkush K, Feldman M, Levy AA (2002) Gene loss, silencing and activation in a newly synthesized wheat allotetraploid. Genetics 160(4):1651–1659 Kashkush K, Feldman M, Levy AA (2003) Transcriptional activation of retrotransposons alters the expression of adjacent genes in wheat. Nature Genet 33(1):102–106 Kashkush K, Khasdan V (2007) Large-scale survey of cytosine methylation of retrotransposons, and the impact of readout transcription from LTRs on expression of adjacent rice genes. Genetics 177:1975–1985 Kenan-Eichler M, Leshkowitz D, Tal L, Noor E, Melamed-Bessudo C, Feldman M, Levy AA (2011) Wheat hand polyploidization results in deregulation of small RNAs. Genetics 188:263–272 CO RR 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 UN Editor Proof 7 Genomic Plasticity in Polyploid Wheat Layout: T1 Standard SC Chapter No.: 7 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 132/134 EC TE D PR OO F Kerber ER, Green GJ (1980) Suppression of stem rust resistance in hexaploid wheat cv Canthach by chromosome 7DL. Can J Bot 58:1347–1350 Kihara H (1919) Über cytologische studien bei einigen getreidearten. I. Species-bastarde des weizens und weizenroggen-bastarde. Bot Mag 33:17–38 Kihara H (1924) Cytologische und genetische studien bei wichtigen getreidearten mit besonderer rücksicht ouf das verhalten der chromosomen und die sterilitat in den bastarden. Mem Cell Sci, Kyoto Imp University, B1: 1–200 Kihara H (1944) Discovery of the DD-analyser, one of the ancestors of Triticum vulgare. Agric Hortic 19:13–14 Kihara H (1954) Considerations on the evolution and distribution of Aegilops species based on the analyzer-method. Cytologia 19:336–357 Kihara H, Lilienfeld F (1949) A new synthesized 6x-wheat. In: Larsson GBaR (ed) Proceedings of Eighth International Congress of Genetics, Stockholm, Sweden, 1949. Hereditas (Suppl), pp 307–319 Kimber G, Sears ER (1987) Evolution in the genus Triticum and the origin of cultivated wheat. In: Heyne EG (ed) Wheat and wheat improvement. American Society of Agronomy, Madison, pp 154–164 Kislev ME (1980) Triticum parvicoccum sp. nov., the oldest naked wheat. Isr J Bot 28:95–107 Kraitshtein Z, Yaakov B, Khasdan V, Kashkush K (2010) Genetic and epigenetic dynamics of a retrotransposon after allopolyploidization of wheat. Genetics 186(3):U801–U889 Levy AA, Feldman M (2004) Genetic and epigenetic reprogramming of the wheat genome upon allopolyploidization. Biol J Linn Soc 82(4):607–613 Liu B, Segal G, Vega JM, Feldman M, Abbo S (1997) Isolation and characterization of chromosome-specific DNA sequences from a chromosome arm genomic library of common wheat. Plant J 11(5):959–965 Liu B, Vega JM, Feldman M (1998a) Rapid genomic changes in newly synthesized amphiploids of Triticum and Aegilops. II. Changes in low-copy coding DNA sequences. Genome 41(4):535–542 Liu B, Vega JM, Segal G, Abbo S, Rodova H, Feldman M (1998b) Rapid genomic changes in newly synthesized amphiploids of Triticum and Aegilops. I. Changes in low-copy noncoding DNA sequences. Genome 41(2):272–277 Lucas H, Moore G, Murphy G, Flavell RB (1992) Inverted repeats in the long-terminal repeats of the wheat retrotransposon wis 2–1A. Mol Bio Evol 9(4):716–728 Lukens LN, Pires JC, Leon E, Vogelzang R, Oslach L, Osborn T (2006) Patterns of sequence loss and cytosine methylation within a population of newly resynthesized Brassica napus allopolyploids. Plant Physiol 140(1):336–348 Lynch M, Force A (2000) The probability of duplicate gene preservation by subfunctionalization. Genetics 154:459–473 Ma XF, Fang P, Gustafson JP (2004) Polyploidization-induced genome variation in triticale. Genome 47(5):839–848 Ma XF, Gustafson JP (2005) Genome evolution of allopolyploids: a process of cytological and genetic diploidization. Cytogenet Genome Res 109(1–3):236–249 Ma XF, Gustafson JP (2006) Timing and rate of genome variation in triticale following allopolyploidization. Genome 49(8):950–958 Maan SS (1977) Fertility of amphiploids in Triticinae. J Heredity 68:87–94 Mac Key J (1954) Mutation breeding in polyploid cereals. Acta Agriculturae Scandinavica 4:549–557 Mac Key J (1958) Mutagenic response in Triticum at different lrvels of ploidy. In: Jenkins CB (ed) Proceedings 1st Mac Key J (1966) Species relationship in Triticum. Proceedings 2nd International Wheat Genetics Symposium, Lund 1963, Hereditas Suppl. 2, pp 237–276 Madlung A, Masuelli RW, Watson B, Reynolds SH, Davison J, Comai L (2002) Remodeling of DNA methylation and phenotypic and transcriptional changes in synthetic Arabidopsis allotetraploids. Plant Physiol 129(2):733–746 CO RR 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768 769 770 771 772 773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 M. Feldman et al. UN Editor Proof 132 Layout: T1 Standard SC Chapter No.: 7 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 133/134 133 EC TE D PR OO F Madlung A, Tyagi AP, Watson B, Jiang HM, Kagochi T, Doerge RW, Martienssen R, Comai L (2005) Genomic changes in synthetic Arabidopsis polyploids. Plant J 41(2):221–230 Maestra B, Naranjo T (1999) Structural chromosome differentiation between Triticum timopheevii and T-turgidum and T-aestivum. Theor Appl Genet 98(5):744–750 McFadden ES, Sears ER (1944) The artificial synthesis of Triticum spelta. Records Genet Soc Amer 13:26–27 McFadden ES, Sears ER (1946) The origin of Triticum spelta and its free-threshing hexaploid relatives. J Heredity 37(81–89):107–116 Mestiri I, Chague V, Tanguy AM, Huneau C, Huteau V, Belcram H, Coriton O, Chalhoub B, Jahier J (2010) Newly synthesized wheat allohexaploids display progenitor-dependent meiotic stability and aneuploidy but structural genomic additivity. New phytol 186(1):86–101 Mitra R, Bhatia C (1971) Isoenzymes and polyploidy. 1. Qualitative and quantitative isoenzyme studies in the Triticinae. Genet Res Camb 18:57–69 Mochida K, Kawaura K, Shimosaka E, Kawakami N, Shin-I T, Kohara Y, Yamazaki Y, Ogihara Y (2006) Tissue expression map of a large number of expressed sequence tags and its application to in silico screening of stress response genes in common wheat. Mol Genet Genomics 276(3):304–312 Morris R, Sears ER (1967) The cytogenetics of wheat and its relatives. In: Quisenberry KS, Reitz LP (eds) Wheat and wheat improvement. Madison, U.S.A., pp 19–87 Nigumann P, Redik K, Matlik K, Speek M (2002) Many human genes are transcribed from the antisense promoter of L1 retrotransposon. Genomics 79(5):628–634 Ohno S (1970) Evolution by gene duplication. Springer, Berlin Okamoto M, Inomata N (1974) Possibility of 5B-like effect in diploid species. Wheat Inform Serv 38:15–16 Ozkan H, Levy AA, Feldman M (2001) Allopolyploidy-Induced rapid genome evolution in the wheat (Aegilops-Triticum) group. Plant Cell 13:1735–1747 Ozkan H, Tuna M, Arumuganathan K (2003) Nonadditive changes in genome size during allopolyploidization in the wheat (Aegilops-Triticum) group. J Hered 94(3):260–264 Parisod C, Alix K, Just J, Petit M, Sarilar V, Mhiri C, Ainouche M, Chalhoub B, Grandbastien MA (2010) Impact of transposable elements on the organization and function of allopolyploid genomes. New Phytol 186(1):37–45 Parisod C, Salmon A, Zerjal T, Tenaillon M, Grandbastien MA, Ainouche M (2009) Rapid structural and epigenetic reorganization near transposable elements in hybrid and allopolyploid genomes in Spartina. New Phytol 184(4):1003–1015 Peng I, Ronin Y, Fahima T, Röder MS, Li Y, Nevo E, Korol A (2003a) Genomic distribution of domestication QTLs in wild emmer wheat, Triticum dicoccoides. In Proceedings 10th International Wheat Genetics Symposium, Paestum, Italy, pp 34–37 Peng JH, Ronin Y, Fahima T, Roder MS, Li YC, Nevo E, Korol A (2003b) Domestication quantitative trait loci in Triticum dicoccoides, the progenitor of wheat. Proc Natl Acad Sci U S A 100(5):2489–2494 Percival J (1921) The wheat plant. E.P. Dutton and Company, New York, pp 1–463 Petit M, Guidat C, Daniel J, Denis E, Montoriol E, Bui QT, Lim KY, Kovarik A, Leitch AR, Grandbastien MA, Mhiri C (2010) Mobilization of retrotransposons in synthetic allotetraploid tobacco. New Phytol 186(1):135–147 Pikaard CS (2000) The epigenetics of nucleolar dominance. Trends Genet 16(11):495–500 Prince VE, Pickett FB (2002) Splitting pairs: the diverging fates of duplicated genes. Nat Rev Genet 3:827–837 Puig M, Caceres M, Ruiz A (2004) Silencing of a gene adjacent to the breakpoint of a widespread Drosophila inversion by a transposon-induced antisense RNA. P Natl Acad Sci U S A 101(24):9013–9018 Pumphrey M, Bai J, Laudencia-Chingcuanco D, Anderson O, Gill BS (2009) Nonadditive expression of homoeologous genes is established upon polyploidization in hexaploid wheat. Genetics 181(3):1147–1157 CO RR 791 792 793 794 795 796 797 798 799 800 801 802 803 804 805 806 807 808 809 810 811 812 813 814 815 816 817 818 819 820 821 822 823 824 825 826 827 828 829 830 831 832 833 834 835 836 837 838 839 840 841 842 843 UN Editor Proof 7 Genomic Plasticity in Polyploid Wheat Layout: T1 Standard SC Chapter No.: 7 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 134/134 EC TE D PR OO F Rapp RA, Udall JA, Wendel JF (2009) Genomic expression dominance in allopolyploids. Bmc Biol 7 Sabot F, Guyot R, Wicker T, Chantret N, Laubin B, Chalhoub B, Leroy P, Sourdille P, Bernard M (2005) Updating of transposable element annotations from large wheat genomic sequences reveals diverse activities and gene associations. Mol Genet Gen 274(2):119–130 Sakamura T (1918) Kurze mitteilung über die chromosomenzahalen und die verwandtschaftsverhältnisse der Triticum Arten. Bot Mag 32(1918):151–154 Salina EA, Numerova OM, Ozkan H, Feldman M (2004) Alterations in subtelomeric tandem repeats during early stages of allopolyploidy in wheat. Genome 47(5):860–867 Salmon A, Ainouche ML, Wendel JF (2005) Genetic and epigenetic consequences of recent hybridization and polyploidy in Spartina (Poaceae). Mol Ecol 14(4):1163–1175 Sax K (1927) Chromosome behavior in Triticum hybrids, Verhandlungen des V Int. Kongresses für Vererbungswissenchaft, Berlin, 2:1267–1284 Schulz A (1913) Die geschichte der kultivierten getreide. Nebert, Halle Sears ER (1972) The nature of mutation in hexaploid wheat. Symp Biol Hung 12:73–82 Sears ER (1976) Genetic control of chromosome pairing in wheat. Annu Rev Genet 10:31–51 Shaked H, Kashkush K, Ozkan H, Feldman M, Levy AA (2001) Sequence elimination and cytosine methylation are rapid and reproducible responses of the genome to wide hybridization and allopolyploidy in wheat. Plant Cell 13:1749–1759 Stebbins GLJ (1980) Polyploidy in plants: unsolved problems and prospect, in polyploidy— biological relevance. In: Lewis WH (ed) Plenum Press, New York Stebbins GLJ (1971) Chromosomal evolution in higher plants. Addison-Wesley, New York Stephens SG (1951) Possible significance of duplication in evolution. Adv Genet 4:247–265 Thompson DA, Desai MM, Murray AW (2006) Ploidy controls the success of mutators and nature of mutations during budding yeast evolution. Curr Biol: CB 16(16):1581–1590 Tirosh I, Reikhav S, Levy AA, Barkai N (2009) A yeast hybrid provides insight into the evolution of gene expression regulation. Science 324(5927):659–662 Van Slageren MW (1994) Wild wheats: a monograph of Aegilops L. and Amblyopyrum (Jaub. and Spach) Eig (Poaceae). Agricultural University, Wageningen, The Netherlands Veitia RA, Bottani S, Birchler JA (2008) Cellular reactions to gene dosage imbalance: genomic, transcriptomic and proteomic effects. Trends Genet 24(8):390–397 Vicient CM, Jaaskelainen MJ, Kalendar R, Schulman AH (2001) Active retrotransposons are a common feature of grass genomes. Plant Physiol 125(3):1283–1292 Von Tschermak E, Bleier H (1926) Über fruchtbare Aegilops-weizenbastarde, der deutsch. Bot Ges 44:110–132 Waines JG (1976) A model for the origin of diploidizing mechanisms in polyploid species. Amer Natur 110:415–430 Wang JB, Xu YH, Zhong L, Wu XM, Fang XP (2009) Rapid alterations of gene expression and cytosine methylation in newly synthesized Brassica napus allopolyploids. Planta 229(3):471–483 Weissmann S, Feldman M, Gressel J (2005) Sequence evidence for sporadic intergeneric DNA introgression from wheat into a wild Aegilops species. Mol Biol Evol 22:2055–2062 Wicker T, Mayer KFX, Gundlach H, Martis M, Steuernagel B, Scholz U, Simkova H, Kubalakova M, Choulet F, Taudien S, Platzer M, Feuillet C, Fahima T, Budak H, Dolezel J, Keller B, Stein N (2011) Frequent gene movement and pseudogene evolution is common to the large and complex genomes of wheat, barley, and their relatives. Plant Cell 23(5):1706–1718 Yaakov B, Kashkush K (2011a) Massive alterations of the methylation patterns around DNA transposons in the first four generations of a newly formed wheat allohexaploid. Genome 54(1):42–49 Yaakov B, Kashkush K (2011b) Methylation, transcription, and rearrangements of transposable elements in synthetic allopolyploids. Int J Plant Genomics. doi:10.1155/2011/569826 Zhang ZC, Belcram H, Gornicki P, Charles M, Just J, Huneau C, Magdelenat G, Couloux A, Samain S, Gill BS, Rasmussen JB, Barbe V, Faris JD, Chalhoub B (2011) Duplication and CO RR 844 845 846 847 848 849 850 851 852 853 854 855 856 857 858 859 860 861 862 863 864 865 866 867 868 869 870 871 872 873 874 875 876 877 878 879 880 881 882 883 884 885 886 887 888 889 890 891 892 893 894 895 896 897 M. Feldman et al. UN Editor Proof 134 Layout: T1 Standard SC Chapter No.: 7 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 135/134 135 EC TE D PR OO F partitioning in evolution and function of homoeologous Q loci governing domestication characters in polyploid wheat. P Natl Acad Sci U S A 108(46):18737–18742 Zhao N, Zhu B, Li M, Wang L, Xu L, Zhang H, Zheng S, Qi B, Han F, Liu B (2011) Extensive and heritable epigenetic remodeling and genetic stability accompany allohexaploidization of wheat. Genetics. doi:10.1534/genetics.111.127688 Zohary D, Feldman M (1962) Hybridization between amphiploids and the evolution of polyploids in the wheat (Aegilops-Triticum) group. Evolution 16:44–61 CO RR 898 899 900 901 902 903 904 UN Editor Proof 7 Genomic Plasticity in Polyploid Wheat Metadata of the chapter that will be visualized in SpringerLink Book Title Polyploidy and Genome Evolution Series Title Chapter Title Maize (Zea Mays) as a Model for Studying the Impact of Gene and Regulatory Sequence Loss Following Whole-Genome Duplication Copyright Year 2012 Copyright HolderName Springer-Verlag Berlin Heidelberg Corresponding Author Family Name Freeling Particle Given Name Michael Suffix Author Division Department of Plant and Microbial Biology Organization University of California Address Berkeley, USA Email freeling@uclink.berkeley.edu Family Name Schnable Particle Given Name James C. Suffix Division Department of Plant and Microbial Biology Organization University of California Address Berkeley, USA Email Abstract Modern maize (2n = 20) is functionally diploid, and its chromosomes pair normally, forming 10 bivalents during meiosis. Sufficient genomic rearrangement has occurred that no two maize chromosomes are homologous across their entire lengths. Yet comparisons of genetic maps, duplicate gene sequences, and later genome assemblies revealed maize is descended from a polyploid ancestor which lived 5–12 million years ago. In the time since that polyploid ancestor lived 8,000–9,000 genes conserved at syntenic positions in other grass species have been reduced to single copy in maize while 4,000–5,000 genes are still retained as homologous gene pairs. The consequences of this polyploidy are continuing to resolve in modern maize accessions. With a wide range of data sets generated by an active research community, maize is an unparalleled model for the in silico study of the changes in genome structure, gene content, and gene regulation that a successful polyploidy brings about in a plant lineage. 1 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 137/144 Chapter 8 6 James C. Schnable and Michael Freeling 11 12 13 14 15 16 17 18 19 22 23 24 25 26 27 CO RR 20 21 PR OO 9 10 Abstract Modern maize (2n = 20) is functionally diploid, and its chromosomes pair normally, forming 10 bivalents during meiosis. Sufficient genomic rearrangement has occurred that no two maize chromosomes are homologous across their entire lengths. Yet comparisons of genetic maps, duplicate gene sequences, and later genome assemblies revealed maize is descended from a polyploid ancestor which lived 5–12 million years ago. In the time since that polyploid ancestor lived 8,000–9,000 genes conserved at syntenic positions in other grass species have been reduced to single copy in maize while 4,000–5,000 genes are still retained as homologous gene pairs. The consequences of this polyploidy are continuing to resolve in modern maize accessions. With a wide range of data sets generated by an active research community, maize is an unparalleled model for the in silico study of the changes in genome structure, gene content, and gene regulation that a successful polyploidy brings about in a plant lineage. D 7 8 TE 4 EC 3 F 5 Maize (Zea Mays) as a Model for Studying the Impact of Gene and Regulatory Sequence Loss Following Whole-Genome Duplication 2 8.1 Background on the Maize Polyploidy Suspicion of a polyploid origin for maize first came from the prevalence of duplicate mutants identified at unlinked locations throughout the genome; these have been observed since the early days of maize genetics (Rhoades 1951). Over time, the evidence that the maize lineage was descended from an ancient polyploid grew to include the arrangement of duplicate genes and markers in similar orders on multiple maize chromosomes reported in a number of genetic maps developed UN Editor Proof Layout: T1 Standard SC Chapter No.: 8 J. C. Schnable  M. Freeling (&) Department of Plant and Microbial Biology, University of California, Berkeley, USA e-mail: freeling@uclink.berkeley.edu P. S. Soltis and D. E. Soltis (eds.), Polyploidy and Genome Evolution, DOI: 10.1007/978-3-642-31442-1_8,  Springer-Verlag Berlin Heidelberg 2012 137 Layout: T1 Standard SC Chapter No.: 8 Book ID: 272454_1_En Date: 16-8-2012 J. C. Schnable and M. Freeling 42 8.2 Timing of the Maize Polyploidy 37 38 39 40 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 PR OO 35 36 D 34 Unlike other important crops species, such as bread wheat and potato, the maize polyploidy was not directly associated with domestication, but occurred millions of years earlier. Analysis of divergence between duplicate genes on opposite maize subgenomes—known as homeologs—has placed the split of the two progenitor genomes found within modern maize at *12 million years before present (Swigoňová et al. 2004). The date of divergence between the two subgenomes of maize does not necessarily reflect how long ago the actual event of polyploidization occurred, as fertile allopolyploids can form between related species that have been evolving independently for millions of years. However, based on a single case of gene conversion it can be concluded that the two subgenomes of maize have shared a single nucleus for at least 5 million years (Swigoňová et al. 2004). Recent evidence has allowed the phylogenetic placement of the maize tetraploidy to be narrowed to a discrete interval within the diversification of the Andropogoneae—the tribe of grasses within which maize is placed (Mathews et al. 2002). Phylogenetic trees of the homologous genes zfl1 and zfl2 show that these whole genome duplicates had already diverged in the common ancestor of the genus Zea (maize and teosinte) and the sister genus Tripsacum (Bomblies and Doebley 2005). Comparison of genetic maps in maize and sorghum prior to the publication of complete genome sequences for these species demonstrated that sorghum—and by extension other relatives within the ‘‘core’’ Andropogoneae— did not share the maize polyploidy (Wei et al. 2007). Following the publication of complete genome assembles for both maize (Schnable et al. 2009) and sorghum (Paterson et al. 2009), a study of thousands of maize homologous gene pairs found that both maize subgenomes appear equally diverged from sorghum (Woodhouse et al. 2010). Given these constraints, the phylogenetic placement of the maize tetraploidy can be inferred (Fig. 8.1). TE 33 EC 31 32 CO RR 30 F 41 during the 1980s (Goodman et al. 1980; Wendel et al. 1986; Helentjaris et al. 1988). With the availability of abundant sequence data for cloned duplicate genes, it became possible to definitively classify maize as an ancient polyploid (Gaut and Doebley 1997). In the absence of diploid species showing higher genetic similarity to one maize subgenome or the other, it will be impossible to conclusively prove maize is not an ancient autopolyploid. The most closely related genus to Zea, Tripsacum, is descended from the same polyploid ancestor (Bomblies and Doebley 2005). The most closely related species with a sequenced genome is Sorghum bicolor, which shows an equal divergence from both subgenomes of maize (Swigoňová et al. 2004; Woodhouse et al. 2010). Almost no molecular data are available for the most closely related, and apparently diploid, genera to maize, Elionurus (13 sequences in GenBank) and Coelorachis (14 sequences in GenBank). The question of autoversus allo-polyploidy may never be conclusively answered for maize. 28 29 UN Editor Proof 138 Book ISBN: 978-3-642-31441-4 Page: 138/144 Layout: T1 Standard SC Chapter No.: 8 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 139/144 139 PR OO F Editor Proof 8 Maize (Zea Mays) as a Model for Studying the Impact of Gene Fig. 8.1 A highly pruned tree of grass species. Branches are scaled by modal synonymous substitutions per site among all syntenic orthologs as measured by SynMap and QuotaAlign (Lyons et al. 2008b; Tang et al. 2011). The position of the branch between Zea and Tripsacum is only approximated, as the genome of Tripsacum has not yet been sequenced 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 D TE 72 While all plant genomes sequenced to date have experienced at least one ancient polyploidization (Paterson et al. 2010), in the vast majority of cases, duplicate regions have been too heavily rearranged to reconstruct which regions can be grouped together as originating from the same parental genome. Such was not the case with maize. The polyploid ancestor of maize possessed 20 chromosomes, 2 equivalent to each of the 10 chromosomes of sorghum. Researchers could assign segments of the ten chromosomes of modern maize back to each of these ancestral chromosomes even before the completion of either the sorghum or maize genomes (Wei et al. 2007). With the publication of the completed maize and sorghum genomes, the same reconstruction can now be carried out using web-based tools by anyone with a fondness for puzzles (Fig. 8.2). A cursory examination of Fig. 8.2 will reveal for any position in the sorghum genome there are two syntenic orthologous regions in maize, while for any position in the maize genome there is only one syntenic orthologous region in sorghum. In no case are the two regions orthologous to the same region in sorghum are present on the same maize chromosome. In 15 of the 20 inferred ancestral chromosomes, all segments are contained within a single chromosome of modern maize. In the remaining five cases, orthologous regions belonging to a single inferred ancestral maize chromosome are split between two—or in one case three—modern maize chromosomes. Fusion of chromosomes has often resulted from the insertion of one chromosome in between the arms of another—often linked with inversions of whole chromosome arms—as previously observed in the reduction of the Brachypodium distachyon genome to five chromosomes from the EC 71 CO RR 70 8.3 Changes in Genome Arrangement Following Polyploidy in Maize UN 69 Layout: T1 Standard SC Chapter No.: 8 Book ID: 272454_1_En Date: 16-8-2012 J. C. Schnable and M. Freeling TE D PR OO F Editor Proof 140 Book ISBN: 978-3-642-31441-4 Page: 140/144 CO RR EC Fig. 8.2 A dotplot generated using SynMap (Lyons et al. 2008b) to compare the arrangement of orthologous regions of the sorghum (x-axis) and maize (y-axis) genome. Each dot represents a pair of homologous genes in maize and sorghum within a syntenic region, and each dot is color coded by synonymous substitutions per site between those two genes. In general, purple lines represent syntenic orthologous regions between the maize and sorghum genomes whilegreen diagonals represent homologous syntenic regions from a more ancient whole-genome duplication shared by all grass species. An interactive version of this graphic can be regenerated using the following link: http://genomevolution.org/r/3vpl 99 8.4 Ancient Gene Loss in Maize 95 96 97 100 101 102 UN 98 inferred ancestral number of 12 (The International Brachypodium Initiative 2010). Researchers can search for differences across whole pairs of reconstructed maize ancestral chromosomes, and, by using differential gene loss as a mark to distinguish chromosome copies, even compare entire subgenomes (Schnable et al. 2011). 94 Being able to compare pairs of complete ancestral chromosomes enabled a thorough investigation of an odd observation reported for Arabidopsis paleopolyploidy: the loss of genes was biased between homologous regions of the genome Layout: T1 Standard SC Chapter No.: 8 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 141/144 141 135 8.5 Ongoing Gene Loss in Maize 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 136 137 138 139 140 141 142 143 PR OO 108 D 107 TE 106 EC 105 CO RR 104 F 134 (‘‘biased fractionation’’), with one copy of each region retaining more total genes than the other (Thomas et al. 2006). Depending on the definition of a functional gene, between 4,000 and 5,000 homologous gene pairs are retained within the maize genome (Schnable et al. 2009; Woodhouse et al. 2010). Another 8,000–9,000 genes conserved at syntenic location in both rice and sorghum have fractionated back to a single copy in maize (Schnable et al. 2011). The same fractionation bias was observed across each of the ten pairs of ancestral maize chromosomes, with one chromosome copy retaining 70–90 % of the ancestral grass gene content and the other chromosome copy retaining only 40–60 %. By grouping high gene loss ancestral chromosomes together into one parental subgenome and low gene loss ancestral chromosomes into the other parental subgenome and using the wealth of maize RNA-seq data being produced by other research projects, it was shown that genes on the subgenome with low gene loss, referred to as maize1, tend to be expressed at higher levels than their duplicates on the subgenome with high gene loss, or maize2 (Schnable et al. 2011). This inequality of expression between parental subgenomes has previously been reported in allopolyploid cotton—among other species—where it was termed ‘‘genomic dominance’’ (Flagel and Wendel 2010). The correlation between bias in gene expression levels and bias in gene loss rates suggests a simple explanation for the bias in gene loss rates observed in all ancient polyploid species studied to date (Sankoff et al. 2010). The deletion of the gene copy that contributed the majority of total gene pair expression should be more likely to have a functional impact than the deletion of the gene copy which contributes the minority of total gene pair expression. Therefore, even if the base rate of gene deletion is equal on both subgenomes, null alleles of more highly expressed gene copies would be more likely to be purged from the population by purifying selection. Over time, deletion of redundant gene copies would tend to cluster on the less expressed subgenome. In support of this hypothesis, it was noted that known maize morphological mutants are significantly more likely to result from the disruption of genes on the high-expression subgenome than expected given the overall distribution of expressed genes between the subgenomes (Schnable and Freeling 2011). 103 In addition to ancient fixed patterns of gene loss, fractionation continues within maize today, with genes present in some maize inbreds lost in other accessions. Concurrent with the publication of the maize genome, the first report emerged describing a high incidence of presence–absence variation (PAV) between different maize inbreds. A comparison of two of the most-studied maize lines, B73 and Mo17 (the former being the line used to generate the reference genome), determined that thousands of sequences were present in the former but entirely absent from the latter, including at least 180 single-copy genes (Springer et al. 2009). UN Editor Proof 8 Maize (Zea Mays) as a Model for Studying the Impact of Gene Layout: T1 Standard SC Chapter No.: 8 Book ID: 272454_1_En Date: 16-8-2012 J. C. Schnable and M. Freeling 160 8.6 Sequence Deletion in Maize 151 152 153 154 155 156 157 158 161 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 PR OO 150 D 149 TE 147 148 The loss of genes is the result of the deletion of sequence from the genome. Given the extensive observation of chromosomal mispairing, rearrangements, and partial or complete aneuploidy in synthetic and recent natural polyploids, it might be expected that much duplicate gene loss is the result of large deletions which remove whole chromosomal segments from one subgenome. The fact that only a portion of the genomes of ancient polyploid species such as rice—65.7 % of the genes covered by duplicated blocks (Yu et al. 2005)—and Arabidopsis—89 % of genes covered (Bowers et al. 2003)—would seem consistent with this expectation. However, the pattern of gene loss in maize is not consistent with large deletions following whole-genome duplication. A search for regions of the sorghum genome represented by only one syntenic orthologous region within maize revealed only one putative deletion of [30 genes (Schnable et al. 2012). A comparison of the patterns of gene loss observed in duplicate regions of the maize genome to simulations assuming different lengths of sequence deletion found that the pattern observed is consistent with [85 % of deletions removing only a single gene, or a portion there of, with the remainder of deletions removing two, or occasionally three, adjacent genes (Woodhouse et al. 2010). The present genome of maize is littered with the partially deleted fragments of homologous genes (Fig. 8.3a). The prevalence of short direct repeats flanking deletions within these fragmentary genes indicates the loss of genes in polyploids is a result of intrastrand non-homologous recombination: short direct repeats found by chance throughout the genome pair with each other, splicing out any intervening sequence, and leaving a single copy of the repeat sequence (Woodhouse et al. 2010). EC 146 CO RR 145 F 159 A follow-up study which examined 33 inbreds and accessions of wild teosinte identified 3,410 high-confidence genes in the genome of B73 which had been lost in one or more of these lines—a total that represents more than 10 % of all highconfidence genes present in the maize genome (Swanson-Wagner et al. 2010). While most PAV in maize was initially observed using comparative genomic hybridization to microarrays, the same variation in gene content was observed in a recent study that resequenced six inbreds using Illumina short read technology (Lai et al. 2010). While many genes which exhibit PAVs between B73 and other maize inbreds possess no syntenic orthologs in other grass species and may have recently inserted into their current locations, 4–6 % of maize genes with syntenic orthologs in both sorghum and rice—indicating these genes have been functionally conserved for at least 50 million years—are also involved in PAV between diverse maize lines (Schnable et al. 2011). Among genes where both homologous are still retained in B73, genes on the ‘high-gene-loss, low-expression’ subgenome (maize2) are more likely to be involved in PAV (Swanson-Wagner et al. 2010; Schnable et al. 2011). 144 UN Editor Proof 142 Book ISBN: 978-3-642-31441-4 Page: 142/144 Layout: T1 Standard SC Chapter No.: 8 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 143/144 143 PR OO F Editor Proof 8 Maize (Zea Mays) as a Model for Studying the Impact of Gene TE D Fig. 8.3 a GEvo panel (Lyons et al. 2008a) illustrating an example of a gene conserved in rice, sorghum, and one of two maize subgenomes (maize1), while on the second maize subgenome (maize 2) only the 5’ and 3’ ends of the gene remain conserved. The colored boxes mark regions of sequence similarity, as determined by blastn, between the orthologous genomic regions in this figure. An interactive version of the same panel can be regenerated at the following link: http:// genomevolution.org/r/3vbu. b A second GEvo panel comparing a single sorghum gene to its two coorthologs in maize. Purple rectangles represent functionally constrained noncoding sequences identified by the comparison of this gene to its syntenic ortholog in rice. The red box highlights a cluster of conserved noncoding sequences conserved upstream of one maize gene but lost from the promoter of the other. c A comparison of the relative expression of these two maize genes in a wide range of tissues using RNA-seq data produced by a number of maize research groups (Wang et al. 2009; Jia et al. 2009; Li et al. 2010; Davidson et al. 2011; Waters et al. 2011) 191 8.7 Future Prospects 187 188 189 192 193 194 195 196 197 198 199 CO RR 186 The mechanism responsible for the unequal expression of duplicate genes from different parents in polyploid species remains unknown. With an extensive collection of well-characterized mutants, including knockouts of a variety of epigenetic mechanisms, maize is an excellent system for the investigation of this inexplicable behavior. If the whole-genome duplication in maize was indeed the result of allopolyploidy, the identification of a wild species more closely related to one maize subgenome than the other remains an exciting possibility. The wealth of RNA-seq data being generated in maize for unrelated purposes creates an opportunity to UN 185 EC 190 Fractionation and sequence deletion are not confined to only the protein-coding regions of genes. Sequence deletion, presumably by the same mechanism of intrastrand non-homologous recombination, can also remove conserved noncoding regulatory sequences from one of two duplicate copies of a maize gene (Fig. 8.3b). These genes often show different patterns of tissue-specific expression in existing maize RNA-seq data sets, allowing researchers to develop testable hypotheses about the function of specific regulatory sequences (Fig. 8.3c) (Freeling et al. 2012). 184 Layout: T1 Standard SC Chapter No.: 8 Book ID: 272454_1_En Date: 16-8-2012 J. C. Schnable and M. Freeling 210 211 212 213 Acknowledgments We thank Vincent Li, a high school intern in the Freeling lab from Project SEED, for identifying the gene fragment shown in Fig. 8.3 and Addie M. Thompson for critical reading of an early version of this text. Funding provided by NSF Plant Genome Research Program grant 0701871 to MF and a Chang-Lin Tien Graduate Fellowship to JCS. 214 References 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 236 237 238 239 240 241 242 243 244 245 Bomblies K, Doebley JF (2005) Molecular evolution of FLORICAULA/LEAFY orthologs in the Andropogoneae (Poaceae). Mol Biol Evol 22:1082–1094 Bowers JE et al (2003) Unravelling angiosperm genome evolution by phylogenetic analysis of chromosomal duplication events. Nature 422:433–438 Davidson RM et al (2011) Utility of RNA sequencing for analysis of maize reproductive transcriptomes. Plant Genome 4:191–203 Flagel LE, Wendel JF (2010) Evolutionary rate variation, genomic dominance and duplicate gene expression evolution during allotetraploid cotton speciation. New Phytol 186:184–193 Freeling M et al (2012) Fractionation mutagenesis and similar consequences of mechanisms removing dispensable or less-expressed DNA in plants. Curr Opin Plant Biol, Advance Online Publication Gaut BS, Doebley JF (1997) DNA sequence evidence for the segmental allotetraploid origin of maize. Proc Natl Acad Sci USA 94:6809–6814 Goodman MM et al (1980) Linkage relationships of 19 enzyme Loci in maize. Genetics 96:697–710 Helentjaris T et al (1988) Identification of the genomic locations of duplicate nucleotide sequences in maize by analysis of restriction fragment length polymorphisms. Genetics 118:353–363 Jia Y et al (2009) Loss of RNA–dependent RNA polymerase 2 (RDR2) function causes widespread and unexpected changes in the expression of transposons, genes, and 24-nt small RNAs. PLoS Genet 5:e1000737 Lai J et al (2010) Genome-wide patterns of genetic variation among elite maize inbred lines. Nat Genet 42:1027–1030 Li P et al (2010) The developmental dynamics of the maize leaf transcriptome. Nat Genet 42:1060–1067 Lyons E et al (2008a) Finding and comparing syntenic regions among Arabidopsis and the outgroups papaya, poplar, and grape: CoGe with Rosids. Plant Physiol 148:1772–1781 Lyons E et al (2008b) The value of non-model genomes and an example using SynMap within CoGe to dissect the hexaploidy that predates the Rosids. Trop Plant Biol 1:181–190 Mathews S et al (2002) Phylogeny of Andropogoneae inferred from phytochrome B, GBSSI, and ndhF. Int J Plant Sci 163:441–450 Paterson AH et al (2010) Insights from the comparison of plant genome sequences. Annu Rev Plant Biol 61:349–372 206 207 208 PR OO 205 D 204 TE 203 EC 202 CO RR 201 F 209 investigate how the expression patterns of homologous gene pairs in maize have diverged in silico. Soon, it may also be possible to study how the functions of duplicate genes have diverged in the 5–12 million years since the maize polyploidy in silico, as high-resolution genome-wide associate studies begin to identify the loci responsible for variation in a wide range of maize phenotypes. In addition to a higher likelihood of being lost entirely, has the reduced importance of less expressed genes on the non-dominant subgenome also given these genes greater freedom for innovation, even if their new role comes at the express of their ancestral function? We predict that higher levels of neofunctionalization will be observed on the nondominant subgenome, but this hypothesis remains untested. 200 UN Editor Proof 144 Book ISBN: 978-3-642-31441-4 Page: 144/144 Layout: T1 Standard SC Chapter No.: 8 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 145/144 145 EC TE D PR OO F Paterson AH et al (2009) The sorghum bicolor genome and the diversification of grasses. Nature 457:551–556 Rhoades MM (1951) Duplicate genes in maize. Am Nat 85:105–110 Sankoff D et al (2010) The collapse of gene complement following whole genome duplication. BMC Genomics 11:313 Schnable JC et al (2011) Differentiation of the maize subgenomes by genome dominance and both ancient and ongoing gene loss. Proc Natl Acad Sci USA 108:4069–4074 Schnable JC et al (2012) Genome-wide analysis of syntenic gene deletion in the grasses. Genome Biol Evol 4:265–277 Schnable JC, Freeling M (2011) Genes identified by visible mutant phenotypes show increased bias toward one of two subgenomes of maize. PLoS ONE 6:e17855 Schnable PS et al (2009) The B73 maize genome: complexity, diversity, and dynamics. Science 326:1112–1115 Springer NM et al (2009) Maize inbreds exhibit high levels of copy number variation (CNV) and presence/absence variation (PAV) in genome content. PLoS Genet 5:e1000734 Swanson-Wagner RA et al (2010) Pervasive gene content variation and copy number variation in maize and its undomesticated progenitor. Genome Res 20:1689–1699 Swigoňová Z et al (2004) Close split of sorghum and maize genome progenitors. Genome Res 14:1916–1923 Tang H et al (2011) Screening synteny blocks in pairwise genome comparisons through integer programming. BMC Bioinf 12:102 The International Brachypodium Initiative (2010) Genome sequencing and analysis of the model grass Brachypodium distachyon. Nature 463:763–768 Thomas BC et al (2006) Following tetraploidy in an Arabidopsis ancestor, genes were removed preferentially from one homologue leaving clusters enriched in dose-sensitive genes. Genome Res 16:934–946 Wang X et al (2009) Genome-wide and organ-specific landscapes of epigenetic modifications and their relationships to mRNA and small RNA transcriptomes in maize. Plant Cell 21:1053–1069 Waters AJ et al (2011) Parent-of-origin effects on gene expression and DNA methylation in the maize endosperm. Plant Cell 23:4221–4233 Wei F et al (2007) Physical and genetic structure of the maize genome reflects its complex evolutionary history. PLoS Genet 3:e123 Wendel JF et al (1986) Duplicated chromosome segments in maize (Zea mays L.): further evidence from hexokinase isozymes. Theoret Appl Genet 72:178–185 Woodhouse MR et al (2010) Following tetraploidy in maize, a short deletion mechanism removed genes preferentially from one of the two homeologs. PLoS Biol 8:e1000409 Yu J et al (2005) The Genomes of Oryza sativa: a history of duplications. PLoS Biol 3:e38 CO RR 246 247 248 249 250 251 252 253 254 255 256 257 258 259 260 261 262 263 264 265 266 267 268 269 270 271 272 273 274 275 276 277 278 279 280 281 282 UN Editor Proof 8 Maize (Zea Mays) as a Model for Studying the Impact of Gene Metadata of the chapter that will be visualized in SpringerLink Book Title Polyploidy and Genome Evolution Series Title Chapter Title Polyploidy in Legumes Copyright Year 2012 Copyright HolderName Springer-Verlag Berlin Heidelberg Corresponding Author Family Name Doyle Particle Given Name Jeff J. Suffix Abstract Division Department of Plant Biology Organization Cornell University Address 14850, Ithaca, NY, USA Email jjd5@cornell.edu Legumes are the third largest family of flowering plants, with over 700 genera and more than 19,000 species. Genomic evidence has shown that a whole-genome duplication (WGD) occurred shortly after the origin of the family, in an ancestor that gave rise to the papilionoids, the clade that comprises 65 % of the genera and 71 % of the species, including nearly all of the economically important crop legumes. This polyploidy event may have been associated with the origin of nitrogen-fixing symbiosis (nodulation) in the papilionoids. Nodulation most likely evolved independently in other legumes outside the papilionoids, hence there appears to be no requirement for polyploidy in the evolution of this important symbiosis. More recent polyploidy, as inferred from chromosome counts, occurs in approximately a quarter of all legume genera for which data are available. In most cases, polyploidy is confined to individual genera, species within genera, or cytotypes within species. An exception is the core clade of the genistoid legumes, a major papilionoid group that includes lupines (Lupinus). This group is probably fundamentally polyploid and also has a propensity for further polyploidy and aneuploidy in many of its genera. The frequency of polyploidy varies considerably among clades of the family, being most common (outside the genistoids) in the largely temperate, herbaceous Hologalegina (including pea and clover), and low in woody tropical groups such as the caesalpinioids. Book ID: 272454_1_En Date: 16-8-2012 Chapter 9 2 Polyploidy in Legumes 3 Jeff J. Doyle 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 PR OO D 8 TE 7 EC 6 Abstract Legumes are the third largest family of flowering plants, with over 700 genera and more than 19,000 species. Genomic evidence has shown that a wholegenome duplication (WGD) occurred shortly after the origin of the family, in an ancestor that gave rise to the papilionoids, the clade that comprises 65 % of the genera and 71 % of the species, including nearly all of the economically important crop legumes. This polyploidy event may have been associated with the origin of nitrogen-fixing symbiosis (nodulation) in the papilionoids. Nodulation most likely evolved independently in other legumes outside the papilionoids, hence there appears to be no requirement for polyploidy in the evolution of this important symbiosis. More recent polyploidy, as inferred from chromosome counts, occurs in approximately a quarter of all legume genera for which data are available. In most cases, polyploidy is confined to individual genera, species within genera, or cytotypes within species. An exception is the core clade of the genistoid legumes, a major papilionoid group that includes lupines (Lupinus). This group is probably fundamentally polyploid and also has a propensity for further polyploidy and aneuploidy in many of its genera. The frequency of polyploidy varies considerably among clades of the family, being most common (outside the genistoids) in the largely temperate, herbaceous Hologalegina (including pea and clover), and low in woody tropical groups such as the caesalpinioids. CO RR 4 5 F 1 Book ISBN: 978-3-642-31441-4 Page: 147/179 UN Editor Proof Layout: T1 Standard SC Chapter No.: 9 J. J. Doyle (&) Department of Plant Biology, Cornell University, Ithaca, NY 14850, USA e-mail: jjd5@cornell.edu P. S. Soltis and D. E. Soltis (eds.), Polyploidy and Genome Evolution, DOI: 10.1007/978-3-642-31442-1_9,  Springer-Verlag Berlin Heidelberg 2012 147 Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 148/179 24 J. J. Doyle 9.1 Introduction 38 9.2 A Brief Overview of Legume Phylogeny 31 32 33 34 35 36 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 PR OO 30 D 29 TE 28 Along with Polygalaceae, Surianaceae, and Quillajaceae, Leguminosae form the order Fabales, one of eight-orders in the Fabidae clade of rosid eudicots (Wang et al. 2009). Bello et al. (2009) suggested that the Fabales are the product of a rapid radiation, with legumes probably sister to Surianaceae plus Quillajaceae. The Leguminosae has been the focus of considerable phylogenetic study, culminating in solid, chloroplast based, working hypotheses of generic relationships (Fig. 9.1), notably those of Wojciechowski et al. (2004) and Bruneau et al. (2008). The older classification of the family into three subfamilies, Caesalpinioideae, Mimosoideae, and Papilionoideae (sometimes treated as separate families), is not supported by molecular phylogenetic studies, in that although the Mimosoideae (mimosas, acacias) and Papilionoideae (pea, bean, soybean, etc.) are monophyletic, the former is embedded in one clade of a paraphyletic caesalpinioid grade. Relationships at the base of the family are uncertain and differ among the studies of Wojciechowski et al. (2004), which focused most heavily on Papilionoideae, and Bruneau et al. (2008), which emphasized caesalpinioids and included few papilionoids. However, both studies identified caesalpinioids as the earliestdiverging lineages, including such taxa as the tribe Cercideae, which includes Cercis (the redbud or Judas tree) and the large genus, Bauhinia (orchid tree). Relationships at the bases of the two monophyletic subfamilies are also uncertain. Fossil evidence places the origin of the family in the Paleocene, around 60 million years ago (MYA; see Lavin et al. 2005 for discussion; see also Bell et al. 2010). Divergence times of all major groups within the family have been estimated from fossil-calibrated molecular data (Lavin et al. 2005; Bruneau et al. 2008) and suggest rapid diversification of many clades, such that within EC 27 CO RR 26 F 37 The legumes (Leguminosae or, less preferably, Fabaceae, according to Lewis et al. 2005) are the third largest family of flowering plants, and are tremendously diverse ecologically, morphologically, chemically, and cytologically (Doyle and Luckow 2003; Lewis et al. 2005). Not surprisingly, the family is also cytologically diverse. As in other families, polyploidy is implicated as a major force at all levels of legume evolution, from the early stages of radiation in the family to the origin and recent diversification of modern genera, such as Glycine (soybean and allies) and species within genera, such as the Medicago sativa complex (alfalfa and allies). After summarizing progress in understanding the phylogeny of the family, this review will discuss the role of paleopolyploidy during the early stages of the radiation of the entire family and the possible connections with nodulation. The occurrence of polyploidy in each of the major clades of the family will then be reviewed. 25 UN Editor Proof 148 Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 149/179 Editor Proof 9 Polyploidy in Legumes 149 x = 14 } Early-diverging papilionoid lineages PR OO polyploidy (to x= 28) Genistoids x = 9 Dalbergioids x = 10 Baphioids x = 11 Hypocalyptus Mirbelioids x = 9 H-Loteae x = 7, 8 H-Robinieae x = 10, 11 H-IRLC x = 8 Indigofereae x = 7, 8 Millettioids x = 10, 11 F Cercideae x = 7 or 14 Detarieae x = 12 Duparquetia Dialiinae x = 14 MCC clade (incl. Mimosoideae) x = 14 (incl. Phaseoleae) 60 55 50 million years TE D Fig. 9.1 Phylogeny of legumes. Caesalpinioids are shown in green, papilionoids in black. Relationships among caesalpinoid lineages are shown as unresolved due to conflict among published studies. Papilionoid taxa marked ‘‘H’’ are members of the Hologalegina. Divergence dates for the origins of major clades are from Lavin et al. (2005). Base chromosome numbers are given for groups with published counts. Minimum (solid arrow) and maximum (dashed arrow) dates for the papilionoid polyploidy event are shown 66 9.3 Polyploidy and the Early Diversification of Legumes 68 69 70 71 72 73 74 75 CO RR 67 To what degree has polyploidy shaped the radiation of legumes? Given the growing understanding that polyploidy can drive phenotypic diversification (e.g., Freeling et al. 2006) and has played a role in the preservation of lineages during periods of extinction (Fawcett et al. 2009), it might be expected that polyploidy would be an important feature of evolution in a family that is ‘‘successful’’ as judged by its sheer size and ecological dominance in some tropical biomes (e.g., rain forests, woody savannas, and dry forests), and in which a significant adaptive novelty—the symbiotic association with nitrogen-fixing soil bacteria, termed nodulation—has arisen (e.g., Doyle 2011). UN 64 EC 65 10 million years of divergence from a common ancestor, all of the major lineages in the family had evolved, including the two monophyletic subfamilies and all of the major clades within the Papilionoideae. 63 Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 150/179 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 F 81 PR OO 80 D 79 Until the advent of genomic data, chromosome number was the prime source of information available for inferring the existence of polyploidy. Goldblatt’s (1981) review of the distribution of chromosome numbers in Leguminosae, published in Advances in Legume Systematics, Part 2, remains the most comprehensive treatment of chromosomal variation in the family, and includes hypotheses concerning the base numbers and ploidy levels of its constituent subfamilies and tribes. The information from which his summary was drawn was included in the descriptions of genera in Advances in Legume Systematics, Part 1 (Polhill and Raven 1981), known to researchers in the family as the legume ‘‘bible.’’ The taxonomic treatments provided by this key resource were recently updated in Legumes of the World (Lewis et al. 2005), taking into account the rapid progress in legume phylogenetics. The phylogenetic studies that have revolutionized our understanding of relationships within the family also provide a new phylogenetic context for understanding chromosome number evolution that was not available previously, but unfortunately the otherwise excellent Legumes of the World does not include any cytological information. The key contribution of objective phylogenetic data to our understanding of cytological evolution in the family is the confirmation of caesalpinioid legumes as a grade rather than as a natural subfamily. Chromosome numbers for the major clades that comprise the caesalpinioid grade are relatively constant, principally 2n = 24–28 (Fig. 9.1). Standing out from these higher chromosome numbers are Chamaecrista and Cercis. The large, mostly, pantropical genus, Chamaecrista, is cytologically complex, with 2n = 14, 16, and 28. Goldblatt (1981) considered its lower numbers to be the products of aneuploid reduction, and this hypothesis has been recently supported (Torres et al. 2011, see below). Phylogenetic studies (Wojciechowski et al. 2004; Bruneau et al. 2008) now nest Chamaecrista and other Cassieae s.s. within the Mimosoideae-Caesalpinioideae-Cassieae (MCC) clade (2n = 28), supporting this hypothesis. Cercis is a small genus (10 species) with disjunct worldwide distribution and 2n = 14. It is a member of the Cercideae, all other members of which are 2n = 28, including the large pantropical genus, Bauhinia s.l. (ca. 250 species). Phylogenetic studies show that Cercis is sister to the remaining genera (Bruneau et al. 2008), which may be consistent with Goldblatt’s (1981) conclusion that it is diploid and the remainder of the tribe is fundamentally polyploid. This is of some importance given the relatively early divergence of Cercideae in some phylogenies. In the rbcL phylogeny of Kajita et al. (2001) the tribe was sister to the remainder of the family, though with relatively weak support, and this topology also appears in the phylogenetic summary of Lewis et al. (2005). The chloroplast matK tree of Wojciechowski et al. (2004), which emphasized Papilionoideae, placed Cercideae, and Detarieae (mainly 2n = 24) together as the first-diverging legume lineage. In these topologies, it is possible that, as Goldblatt (1981) suggested, the legumes are fundamentally x = n = 7, with subsequent independent TE 78 EC 77 9.3.1 Chromosome Number Evidence for Polypoidy in Legumes CO RR 76 J. J. Doyle UN Editor Proof 150 Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 151/179 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 F 125 126 PR OO 123 124 D 122 TE 121 chromosomal increase both within Cercideae and in the ancestor of all remaining legumes. In contrast, the concatenated chloroplast matK/trnK ? trnL-F tree of Bruneau et al. (2008) placed Detarieae as the first branch in the legume phylogeny, sister to a trichotomy composed of Cercideae, Duparquetia, and the remainder of the family. In this phylogeny, then, the base number for the family would be x = n = 12, with 2n = 14 in Cercis representing a reduction. Interestingly, the genome size of Cercis canadensis is comparable to measurements from the several species of Bauhinia in the Kew C-value database (http://data.kew.org/cvalues/; Leitch and Doyle, unpublished data), supporting this reduction hypothesis. Even a high base number for early diverging lineages, as suggested by the Bruneau et al. (2008) topology, would not definitively suggest polyploidy at the base of the family, given what is known of chromosome numbers from other Fabales. No information is available for Quillaja in the Index of Plant Chromosome Numbers (IPCN; http://www.tropicos.org/Project/IPCN), but Surianaceae is represented by a single species of Stylobasium, with a number of 2n = 30, suggesting that the common ancestor of legumes and Surianaceae could have had a high chromosome number. Patterns of chromosomal evolution among major groups of legumes are complex even outside of the earliest branching. The bulk of the family belongs to two sister clades: the MCC clade and the Papilionoideae. The two tribes that comprise the MCC clade along with Mimosoideae (Caesalpinieae and Cassieae s.s.) are both diploid based on x = 14. Given the presence of taxa with 2n = 28 in the grade at the base of Papilionoideae, it is likely that the common ancestor of that group and the MCC was diploid based on x = 14 as well. The majority of papilionoids, however, have lower base chromosome numbers, ranging from x = 7–11, depending on the tribe. These presumably represent reductions in chromosome number, as discussed below; they certainly give no evidence for polyploidy. EC 120 151 CO RR 119 9.3.2 Genetic and Genomic Evidence for Polyploidy in the Early Evolution of Legumes Genomic studies, starting with linkage maps and continuing through studies of expressed sequence tags (ESTs) and genome sequencing, have revolutionized understanding of polyploidy in seed plants. Although it was long known that diploidization can erase chromosomal evidence of polyploidy over time, it is now clear that plant genomes comprise nested sets of WGD. The common ancestor of all seed plants underwent a polyploid duplication, with a later WGD in the ancestor of all angiosperms (Jiao et al. 2011) and numerous lineage-specific duplications in various groups of flowering plants (Soltis et al. 2009). It has been known for some time that cryptic polyploidy occurs in legume genomes. For example, Shoemaker et al. (1996) used linkage map information to UN Editor Proof 9 Polyploidy in Legumes Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 152/179 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 184 185 186 187 188 189 190 191 192 193 194 195 196 197 198 199 200 201 202 203 F 164 PR OO 163 D 162 TE 161 hypothesize that the soybean genome shows evidence of a more ancient duplication than the one that is responsible for its high chromosome number relative to allied phaseoloid genera (millettioid clade, Fig. 9.1). In 2004, two different groups mined the extensive EST collections of soybean and the diploid model legume, Medicago truncatula (2n = 14; a member of the Hologalegina IRLC clade), to search for the genomic signature of ancient polyploidy events (Blanc and Wolfe 2004; Schleueter et al. 2004). This signature is produced when all genes in the genome are duplicated by autopolyploidy or when homoeologous loci are brought together by allopolyploidy. It is observed by plotting the frequency distribution of pairwise Ks (synonymous substitutions per synonymous site—a stand-in for time) values for hundreds to thousands of paralogous gene pairs. Simple gene duplication is an ongoing phenomenon in all eukaryotes, but most duplicates are purged from the genome rapidly, producing a characteristic distribution with many recent duplicates with low Ks values and relatively few older pairs with high Ks (Lynch and Conery 2003). Polyploid duplications appear as additional components (‘‘peaks’’) against this background; the mode of such a Ks peak is taken as an estimate for the age of the polyploid, though it is generally an overestimate of that age (Doyle and Egan 2010). Both Blanc and Wolfe (2004) and Schleueter et al. (2004) identified two Ks peaks in soybean, as expected (Fig. 9.2); both reported similar Ks modes for these peaks but because they used different substitution rates for plant nuclear genes, this led to different estimates of the age of polyploidy (or homoeologue divergence). The Schleueter et al. (2004) estimates are more in keeping with divergence dates for papilionoid legume taxa (Lavin et al. 2005) and are preferred for that reason (Shoemaker et al. 2006); in addition, the rate used by Schleueter et al. (2004) is much closer to the rate recently estimated for Arabidopsis (Ossowski et al. 2010). Of great interest was the finding, by both groups, of two Ks peaks in the M. truncatula EST collection (Fig. 9.2). The younger of the two peaks is recent enough that if due to polyploidy, it would most likely have left chromosomal evidence, and has yet to be explained (Young et al. 2011). The older Medicago peak, on the other hand, was estimated by Schleueter et al. (2004) to be around 54.6 MYA, very close to the 54 MY age estimated by Lavin et al. (2005) for the divergence of the soybean (millettioid) and Medicago (Hologalegina) lineages, and also similar to the age estimated for the older soybean peak (41.6 MYA). This raised the possibility that the two species shared an ancient WGD. This hypothesis was tested by Pfeil et al. (2005) using a phylogenomic approach with 39 gene pairs chosen from among those used by Schleueter et al. (2004) to identify the Ks peak in soybean. Topologies of gene trees overwhelmingly favored the hypothesis that their common ancestor was polyploid. Comparisons of linkage relationships between Medicago and the other legume model species, Lotus japonicus (in the Hologalegina Loteae clade, Fig. 9.1), provided further support for this hypothesis, and also showed that the duplication was not found in poplar. Using the Lavin et al. (2005) date for the divergence of the millettioid (Glycine) and Hologalegina (Lotus, Medicago) clades, the WGD event had taken place by around 54 MYA. Thus, the common ancestor of the two major EC 160 CO RR 159 J. J. Doyle UN Editor Proof 152 Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 153/179 153 CO RR EC TE D PR OO F Editor Proof 9 Polyploidy in Legumes UN Fig. 9.2 Evidence for polyploidy in the genomes of: a Glycine max and b Medicago truncatula. The graphs plot the number of pairs of paralogous sequences (‘‘density’’ or ‘‘percent of pairs’’) versus binned Ks (‘‘synonymous distances’’) classes. Pairs with very low divergence (produced by ongoing recent duplications) were not plotted. Curves were fit to the binned divergence data and are interpreted as groups of genes duplicated simultaneously in large-scale genomic events such as polyploidy; modes of peaks provide a maximum age for allopolyploid events (Doyle and Egan 2010). Divergence time was estimated from synonymous distances using standard clock methods; note the different estimated ages (modes of curves) for the older event in the two species. Data are from Schleueter et al. (2004), who used expressed sequence tags (ESTs). Figure courtesy of Jessica Schlueter (UNC-Charlotte) Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 154/179 J. J. Doyle 237 9.3.3 Polyploidy and Nodulation in Legumes 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 238 239 240 241 242 243 PR OO 210 211 D 209 TE 207 208 EC 206 CO RR 205 F 236 sister clades that comprise nearly 9,000 species––around 45 % of all legumes and 63 % of papilionoids––was polyploid. Subsequently, Bertioli et al. (2009) studied Arachis (peanut) and showed that its genome also shows evidence of the [54 MYA event, indicating that the large dalbergioid clade is also fundamentally polyploid. Unpublished information reported by McClean (personal communication) at the 2009 International Conference of Legume Genetics and Genomics showed that Lupinus, and hence the genistoid clade, also shares this WGD. Thus, all of the lineages of the main radiation of the papilionoids share a polyploid ancestor. The possibility that all legumes share this polyploidy event remained tenable until transcriptomic data from the caesalpinioid genus, Chamaecrista, became available (Singer et al. 2009). Cannon et al. (2010) produced and analyzed over 1,200 gene phylogenies from these data, which overwhelmingly supported the conclusion that the Chamaecrista genome shows no evidence of any polyploidy event subsequent to the prerosid triplication, and notably lacks the WGD found in core papilionoids (all but the early diverging lineages in Fig. 9.1). Chamaecrista belongs to the MCC clade, which is sister to the papilionoid clade. Therefore, the absence of the WGD in Chamaecrista indicates that the common ancestor of the MCC clade and papilionoids was not polyploid, placing the WGD within papilionoids (Fig. 9.1). Whether the WGD took place in the papilionoid common ancestor is still unknown, because the lineages that comprise the paraphyletic grade lacking the putative molecular synapomorphy for the major papilionoid radiation (chloroplast genome 50 kb inversion) remain to be sampled. Thus, it is possible that this core papilionoid WGD facilitated the radiation of the most species-rich lineage of legumes, comprising 69 % of the species (13,390/ 19,327) and 59 % (438/741) of the genera of the third largest family of flowering plants. This is the group that is more uniformly characterized by the eponymous legume fruit, by the bilaterally symmetric papilionoid flower, and by the ability to nodulate. The early diverging grade of papilionoids does contain some genera with papilionoid flowers and legume fruits, but many lineages in this part of the tree are characterized by unusual, nonpapilionaceous corollas and drupaceous or samaroid fruits (Pennington et al. 2000); this grade also contains nearly all of the papilionoid genera that do not nodulate (Doyle 2011). 204 The correspondence between nodulation and polyploidy in the family is interesting. Core papilionoids nearly all nodulate, but this is also true of Mimosoideae, and Chamaecrista is among a handful of caesalpinioids known to be able to form a nodulation symbiosis (Sprent 2009). It remains unclear whether there was a single origin of nodulation in the common ancestor of the papilionoid and MCC clades, followed by many losses of nodulation, or whether there were multiple origins of UN Editor Proof 154 Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 155/179 155 264 9.3.4 Harmonizing Chromosomal and Genomic Evolution 252 253 254 255 256 257 258 259 260 261 262 265 266 267 268 269 270 271 272 273 274 275 276 277 278 279 280 281 282 283 284 PR OO 250 251 D 249 TE 248 EC 246 247 Whatever the original basic chromosome number of the family, the earliest radiation from the common ancestor does not seem to have involved polyploidy, despite Goldblatt’s (1981, p. 457) conclusion that the ‘‘… initial phase of polyploidy is probably very ancient and may have taken place in the late Cretaceous, when major groups of Leguminosae began differentiating and were probably evolving rapidly into new habitats.’’ Evidence against Goldblatt’s view of polyploidy in the ancestor of the entire family is the absence of any trace of polyploidy in gene families of Chamaecrista other than the prerosid whole-genome triplication (WGT; Jaillon et al. 2007). This indicates that the ancestor of the MCC and older ancestors back to the prerosid WGT did not experience polyploid duplications. The uniformity of base chromosome numbers in the major radiations of the family— Cercideae, detarioids, MCC, and probably papilionoids (see above) suggest that relatively high numbers (2n = 24–28) are plesiomorphic in the family. The most parsimonious hypothesis for papilionoids is that the earliest papilionoid ancestor was also 2n = 28. Shortly after the divergence of this ancestor from the MCC ancestor, the papilionoids radiated rapidly, and polyploidy occurred nearly simultaneously, no later than the divergence of the first major lineage to diverge in the core clade (genistoids, e.g., Lupinus). This WGD did not leave evidence in higher chromosome numbers; to the contrary, polyploidy is associated with chromosome number reduction in core papilionoids (Fig. 9.1). Goldblatt CO RR 245 F 263 nodulation in the MCC clade and an independent origin in the ancestor of the core papilionoids (Doyle 2011). The demonstration that Chamaecrista not only lacks the core papilionoid WGD, but also does not show any genomic evidence of other polyploidy events, indicates that polyploidy is not a prerequisite for nodulation in legumes as a whole (Cannon et al. 2010). Thus, in a model of a single origin of nodulation in the family, polyploidy would have played no role (Fig. 9.3). At the other extreme, polyploidy could not have been involved in an origin of nodulation unique to Chamaecrista, nor would it have been essential for the origin of nodulation in a model where the symbiosis evolved in a common ancestor of Chamaecrista and other members of the MCC clade (e.g., Mimosoideae). This is not to say, of course, that polyploidy was not important in the origin or evolution of nodulation in core papilionoids or in other nodulating taxa whose genomes have yet to be explored, such as mimosoids. In the case of papilionoids, it is possible that nodulation and the WGD will be found to coincide, either in the ancestor of the main radiation of core papilionoids or at the first papilionoid ancestor, but better phylogenetic resolution is required before this can be tested (Pennington et al. 2001). In either case, refinement of the nodulation symbiosis in taxa such as Medicago may well have been facilitated by the availability of homoeologues produced in the core papilionoid WGD (Young et al. 2011). 244 UN Editor Proof 9 Polyploidy in Legumes Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 156/179 PR OO F Cercideae Detarieae Amherstieae Dialiinae Umtiza clade Batesia Chamaecrista Melanoxylon Cassia Senna Caesalpinia clade Moldenhawera Erythrophleum Calpocalyx Pentaclethra Parkia core Mimosoideae Campsiandra Dimorphandra Burkea Dinizia Arapatiella Jacqueshuberia Sclerolobium Tachigali Peltophorum clade Swartzia clade Non-nod. papil. Mimosoideae-Caesalpinieae-Cassieae (MCC) clade J. J. Doyle Editor Proof 156 TE D N 60 50 40 EC P 20 30 million years 10 “core papilionoids” 0 285 286 287 UN CO RR Fig. 9.3 Relationship between polyploidy and nodulation in Leguminosae. Genera known to nodulate are shown in bold face; Pentaclethra includes both nodulating species and species that apparently cannot nodulate. Possible origins of nodulation are indicated with symbols and colors. A single origin of nodulation for the entire family could have occurred in the common ancestor of the papilionoids and the MCC clade (indicated by ‘‘N’’); this would have required many independent losses in the course of legume evolution. Independent origins within the MCC clade and Papilionoideae would require fewer losses. In the MCC clade, a single origin could have occurred in the common ancestor of all genera known to nodulate (triangle), or once in each major lineage of nodulating taxa (circles), or additional times within some clades (squares). Similarly, a single origin could be hypothesized for papilionoids (green circle), or twice (squares). The placement of the papilionoid polyploidy event (red ‘‘P’’) is indicated as in Fig. 9.1. Polyploidy is associated with nodulation only in the papilionoids and might not be directly associated there, given the uncertainty about the placement of both the polyploidy event and the origin(s) of nodulation. Figure adapted from Doyle (2011) (1981, p. 457) notes that ‘‘the cytological history of legumes seems to involve some descending aneuploidy in every major evolutionary line but is most pronounced in Papilionoideae, in which most predominantly herbaceous tribes or Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 157/179 157 299 9.4 Polyploidy in Tribes and Genera of Legumes 295 296 297 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 PR OO 294 Genomic and phylogenomic information currently is limited to the few genera discussed above. More information can be expected with the advent of new sequencing technologies, notably (at this writing) Illumina and 454. Twenty-five legume genera, representing all three subfamilies, are included in the 1,000 Transcriptomes project (1 kp, see http://www.onekp.com/angiosperms.html), but results from these species have not yet been analyzed (Steven Cannon, personal communication). Chromosome numbers provide a better guide at this level, but clearly should be interpreted with the caveat that high numbers are likely to indicate polyploidy, but low numbers cannot be assumed to be fundamentally diploid, given the potential for cryptic polyploidy, a phenomenon discussed further below. Three sources of data were used for the following survey of polyploidy in legume clades: Goldblatt (1981); chromosome numbers provided for each genus listed in the tribal treatments in Advances in Legume Systematics, Part 1 (Polhill and Raven 1981); and the on-line Index of Plant Chromosome Numbers (IPCN; http://www.tropicos.org/projectwebportal.aspx?pagename=Home&projectid=9). IPCN was searched for all genera in Lewis et al. (2005) for which no data were available in Goldblatt (1981) or Polhill and Raven (1981), and for genera with polyploid counts, to identify infrageneric patterns of polyploidy, particularly when published phylogenies were available. For genera with evidence of polyploidy, BIOSIS was searched using the genus name in conjunction with either \polyploid*[and/or\phylogeny[. For some larger genera, IPCN was also consulted to search for polyploid counts published since 1981. There remain many gaps in our knowledge of legume chromosome numbers. Overall, only around 54 % of legume genera have counts reported (Table 9.1), and the percentage is far lower for tropical woody groups such as the caesalpinioid tribe Detarieae (30 %). This is perhaps not too surprising, given the size of the family, the large number of small genera, and the tropical distributions of many groups. It is also no doubt a commentary on how counting chromosomes has fallen from favor in this age of highthroughput science. D 292 293 TE 291 EC 290 CO RR 289 F 298 genera have achieved relatively low base numbers.’’ The details of this process await elucidation of phylogenetic relationships of the early diverging papilionoid lineages, and sampling of these taxa for the presence of polyploidy. What is clear is that, at some point early in the history of Papilionoideae, chromosome numbers fell from as high as n = 28 to n = 7–11. This process could have begun in, or prior to, the common ancestor of the core papilionoids (Fig. 9.1). In any case, the initial reduction was rapid, taking place within 10 MY after the polyploid event, and perhaps within only 5 MY of the WGD. Detailed studies of synteny in the various core papilionoid lineages should help elucidate whether initial reorganization occurred in a common ancestor or was completely independent in the longdiverged major clades of the core papilionoids. 288 UN Editor Proof 9 Polyploidy in Legumes Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 158/179 158 J. J. Doyle 171 (2251) 82 (3271) 40 (415) 83 (2354) 53 (1514) 7 (58) 32 (763) 34 (414) 62 39 15 59 32 6 22 23 4 10 2 24 11 1 5 9 6 26 13 41 34 17 23 39 54 (4351) 7 (768) 40 2 15 1 38 50 56 (1104) 22 1 5 112 (2064) 731 71 393 12 95 17 24 a F Percent genera with polyploidy (%) PR OO Caesalpinioidsb Mimosoideae Swartzieae ? Sophoreae Genistoids Dalbergioids Baphioids Mirbelioids Hologalegina: Robinioids ? Loteae Hologalegina: IRLC Millettioids: indigofereae Millettioids: core Millettioids Millettioids: phaseoloids Total based on chromosome numbers Genera with Number of chromosome genera with counts polyploidy D Editor Proof Table 9.1 Polyploidy in legume clades, Clade Number of genera (species)a Source: Lewis et al. (2005) Summation of the following monophyletic groups: Cercideae, 12 genera total, 2 of 6 genera with chromosome counts have reports of polyploidy, 33 % (12, 2/6, 33 %); Dialiinae (17, 0/7, 0 %); Detarieae (82, 2/25, 8 %); MCC clade minus mimosoids (59, 0/24, 0 %) TE b 342 9.4.1 Caesalpinioids and Mimosoideae 332 333 334 335 336 337 338 339 340 343 344 345 CO RR 331 UN 330 EC 341 Goldblatt (1981) made a distinction among genera that are exclusively polyploid relative to genera in the same tribe, and genera that include species with both ‘‘diploid’’ and ‘‘polyploid’’ chromosome numbers. In light of our current understanding of polyploidy, this is an artificial distinction, but exclusively polyploid genera are perhaps worth noting, because they represent lineages where polyploids may have replaced their diploid progenitors entirely. Such lineages would constitute evidence of polyploid ‘‘success’’ (e.g., Mayrose et al. 2011) in measures of diversity a million years from now. The results of this survey show that polyploidy, as inferred solely from chromosome numbers, occurs in nearly a quarter of all legume genera, but varies widely in frequency among different lineages (Table 9.1). As Goldblatt (1981) noted, polyploidy is rare in woody, tropical groups such as caesalpinioids and early diverging papilionoid lineages. 329 9.4.1.1 Clades of the Caesalpinioid Grade, Excluding the MCC Clade Goldblatt (1981) stated that polyploidy (assuming a basic diploid number of 2n = 28) is uncommon in caesalpinioid groups. As noted above, unpublished Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 159/179 159 360 9.4.1.2 The MCC Clade 355 356 357 358 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 PR OO 353 354 D 351 352 Chromosome numbers of the caesalpinioid members of the MCC clade range from 2n = 20–28, with the exception of Chamaecrista, in which numbers of 2n = 14, 16, and 28 occur. Goldblatt (1981) raised the possibility that low chromosome numbers in this genus are not ancestral, but instead represent aneuploid reduction from ancestral x = 14 species. This makes good sense given the prevalence of 2n = 28 counts throughout much of the MCC clade. A recent molecular phylogenetic analysis of Chamaecrista (Torres et al. 2011) supported the hypothesis that the 2n = 14 species of sect. Xerocalyx form a monophyletic group nested within species having higher chromosome numbers. Chamaecrista is a genus of considerable interest as a potential model for nonpapilionoid legumes (Singer et al. 2009), making it an attractive system for exploring chromosome evolution in the caesalpinioids. Mimosoideae comprise a monophyletic group embedded in the MCC clade, and the subfamily is dominated by taxa with n = 13. Goldblatt (1981) listed only three genera with base numbers higher than n = 14: Schleinitzia (2n = 52, 54), Leucaena (2n = 52, 56, 104, 112), and Dichrostachys (2n = 50, 56). These genera are all members of tribe Mimoseae and are relatively closely related within that tribe (Lewis et al. 2005), with the former two being members of the same clade and Dichrostachys being part of a sister clade (Luckow et al. 2003). However, Schleinitzia and Leucaena are not sisters within their clade (Hughes et al. 2003; Luckow et al. 2005), suggesting that polyploidy has originated independently in each case. Leucaena itself has been fertile ground for systematic investigation. Boff and Schifino-Wittmann (2003) concluded that its species are segmental paleopolyploids. A series of studies has built a strong foundation for understanding the complex history of hybridization and polyploidy in the genus, and the impact of these phenomena on characters such as nrDNA ITS pseudogene evolution, and TE 350 EC 349 CO RR 348 F 359 genome size data do not support the hypothesis that Bauhinia (commonly 2n = 28) is tetraploid relative to Cercis (2n = 14). However, this is negative evidence, and given the prevalence of genome downsizing (Leitch and Bennett 2004) in polyploids, it remains possible that Bauhinia (along with Adenolobus and Griffonia) is polyploid. More recent polyploidy occurs in the group, with Tylosema having 2n = 52; Tylosema is nested within Bauhinia s. l. (Sinou et al. 2009). The large Detarieae s.l. clade is overwhelmingly 2n = 24. The two detarioid genera having higher chromosome numbers (Hardwickia, 2n = 34; Colophospermum, 2n = 36) are strongly supported as sisters within the Prioria clade of Bruneau et al. (2008); they could represent independent aneuploid reduction from a polyploid ancestor. Goldblatt (1981) mentioned polyploidy in Anthonotha (2n = 24, 28, 72), but only 2n = 24 and 28 are listed by Cowan and Polhill (1981), and no counts are listed in IPCN. All known counts from the Dialiinae clade are 2n = 24 or 28. 346 347 UN Editor Proof 9 Polyploidy in Legumes Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 160/179 J. J. Doyle 418 9.4.2 Papilionoideae 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 419 420 421 422 423 424 425 426 PR OO 394 D 393 TE 392 EC 391 CO RR 390 F 417 their role in domestication (Hughes et al. 2002, 2007; Govindarajulu et al. 2011a, b). Govindarajulu et al. (2011b) concluded that ‘‘… a comprehensive picture of the complex evolutionary dynamics of polyploidy in Leucaena is emerging. This includes paleotetraploidization, diploidization of the last common ancestor to Leucaena, allopatric divergence among diploids, and recent allopolyploid origins for tetraploid species likely associated with human translocation of seed.’’ Acacia (sens. lat.) is reported to have 2n = 26, 52, 76, and 104, though the majority of its species are diploid (Gallagher et al. 2011). Polyploidy has not been a focus of recent phylogenetic analyses (e.g., Brown et al. 2010; Murphy et al. 2010). Similarly, phylogenetic studies of Prosopis (2n = 28, 52, 56) or Prosopidastrum (2n = 28, 56) do not discuss polyploidy (e.g., Bessega et al. 2006; Catalano et al. 2008). No phylogenetic studies appear to have addressed polyploidy in Neptunia (2n = 28, 36, 54, 56, 72, 78), though Pandit et al. (2006) note that N. plena, an invasive species in Singapore, is a polyploid (2n = 72). No studies appear to exist on polyploidy and phylogeny of Inga (2n = 26, 52; the latter reported by Hanson 1995), Albizia (2n = 26, c. 78), or Calliandra (2n = 16, 22, 32, 44). Mimosa (2n = 24, 26, 28, 40, 52) is a genus of around 500 species; a second genus whose species vary in ploidy, Schrankia (2n = 16, 22, 24, 26, 52), is deeply nested within Mimosa (Simon et al. 2011). Dahmer et al. (2011) concluded that the phylogenetic pattern ‘‘… suggests that duplication of chromosome numbers evolved several times in the genus and that polyploidy is not restricted to any particular clade within Mimosa. On the contrary, it seems that polyploids arose independently from ancestors with lower ploidy levels and are present in divergent lineages in the genus.’’ Seijo and Fernandez (2001) reported chromosome numbers from the southern extreme of the range and discovered polyploidy within M. balansae. Chromosomal and morphological studies by Morales et al. (2010) clarified relationships in the M. debilis-M. nuda complex, demonstrating that hybridization and polyploidy are responsible for taxonomic complexity in the group. 388 389 Relatively few papilionoid genera appear to be exclusively polyploid based on chromosome number. As noted above, the early diverging papilionoid lineages have relatively high numbers, like the caesalpinioid and mimosoid groups. Ateleia (Swartzieae) is 2n = 40, presumably representing a second polyploidy event followed by aneuploid reduction. Goldblatt listed Dipteryx (Dipterygeae) as being a polyploid genus but with a questionable count of 2n = 32; this number was not reported in the treatment of the tribe by Polhill (1981a), nor is a count for the genus listed in IPCN. UN Editor Proof 160 Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 161/179 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 F 432 433 PR OO 431 D 430 The genistoid clade, which is weakly supported as sister to the remaining core papilionoids (Fig. 9.1), is one of the most complex groups with respect to polyploidy, and detailed discussion is beyond the scope of this review. The frequency of polyploidy in the genistoid clade is the highest for any well-sampled group of legumes (Table 9.1). Cusma-Velari and Feoli-Chiapella (2009) discussed cytology in ‘‘so-called ‘primitive’ genera of Genisteae’’ in light of molecular phylogenetic data. Genistoids clearly have a base number of x = 9 (Goldblatt 1981), with 2n = 18 being common in most of its tribes. Sophora s.l. has been divided into several segregates that vary in chromosome number. Sophora s.s. is part of the genistoid clade and is x = 9. Boatwright and van Wyk (2011) reported on the relationships of several of these based on nrDNA ITS sequences. They focused on the placement of the South African species, S. ihambanensis, which is polyploid (2n = 36); in their tree it is sister to S. tomentosa, a diploid, but they do not discuss origins of the polyploid. A count of 2n = 18 is common in Sophora s.l., and several additional species are polyploids with 2n = 36 (S. alopecuroides, S. pachycarpa, and S. songarica); S. leachiana is listed by ICPN as having 2n = 36 and 2n = 54 cytotypes. The small segregate, Calia, has 2n = 18 and may be sister to the entire genistoid clade. The remainder of Sophora s.l. comprises 2n = 28 species transferred to Styphnolobium in the early diverging papilionoid grade. In Thermopsideae, Thermopsis has both diploid and polyploid species (2n = 18, 36). In Podalyrieae, two genera are exclusively polyploid: Virgilia (2n = 54) and Cyclopia (2n = 36); they are not supported as sisters in Boatwright et al. (2008). Crotalarieae are sister to Genisteae and include the exclusively polyploid Buchenroedera (2n = 28; Van Wyk and Schutte 1988), as well as polyploids within Crotalaria (2n = 14, 16, 32) and Lotononis (2n = 18, 28, 36). Genisteae are by far the most cytologically complex group in the entire Leguminosae. Even genera with low numbers may be polyploid, such as Anarthrophyllum (2n = 24; Goldblatt 1981) and Dichilus (2n = 28). These genera, along with the polyploid Polhillia (2n = 32) and complex Melolobium (2n = 18, 32), were once placed in Crotalarieae. It is in the core Genisteae that polyploidy and aneuploidy have run rampant. The group includes Argyrolobium (2n = 24, 26, 30, 32, 48), Adenocarpus (2n = 26, 46, 48, 52, 54), Laburnum (2n = 48, 50), Cytisophyllum (2n = 50, 52), Petteria (2n = 52), Argyrocytisus (2n = 50), Chamaecytisus (2n = 48, 96), Cytisus (2n = 22, 24, 46, 48, 92, 96), Calicotome (2n = 24, 48, 50), Erinacea (2n = 52), Spartium (2n = 48, 52, 54, 56), Retama (2n = 48), Genista (2n = 18, 22, 24, 26, 28, 30, 32, 36, 40, 42, 44, 46, 48, 50, 52, 56, 72, 80, 96), Echinospartium (2n = 44, 52), Stauracanthus (2n = 28, 48, ca. 128), Ulex (2n = 32, 64, 80, 96), and Lupinus (2n = 24, 30, 32, 34, 36, 38, 40, 42, 48, 50, 52, 96). Bisby (1981) considered the plethora of chromosome numbers attributed to individual genera to be partly a real phenomenon, but also due to difficulties in obtaining reliable counts given the small size and high numbers of chromosomes, TE 429 9.4.2.1 Genistoids EC 428 161 CO RR 427 UN Editor Proof 9 Polyploidy in Legumes Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 162/179 J. J. Doyle 497 9.4.2.2 Dalbergioids 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 PR OO 478 D 477 TE 475 476 EC 473 474 CO RR 472 F 496 combined with the taxonomic complexity of the groups. Some of the taxonomic complexity is being resolved by molecular phylogenetic studies focused on Cytisus and Genista (Cubas et al. 2002; Pardo et al. 2004), but these studies do not discuss polyploidy per se. For Lupinus, three studies have documented the rapid radiation of the genus in its New World center of diversity (Hughes and Eastwood 2006; Drummond 2008; Drummond et al. 2012), and another phylogenetic study focused on the Old World species (Ainouche et al. 2004). Drummond (2008) noted that, ‘‘While a complex history of aneuploidy (2n = 32, 34, 36, 38, 40, 42, 50, 52) in the Old World and eastern New World … implies that allopolyploidy may have provided an additional mechanism for reproductive isolation and evolutionary divergence, chromosomal numbers in the western New World species (2n = 48 with occasional autopolyploids of 2n = 96) are relatively stable.’’ It is this western group that has radiated explosively, presumably driven by ecology and not due to polyploidy per se (Drummond 2008). Conterato and Schifino-Wittmann (2006) described chromosome numbers and meiotic behavior in diploid and polyploid American lupines, and noted consistencies with phylogenetic relationships in the genus. The placement of polyploid former Crotalarieae in the same clade with core Genisteae may lend support to the idea that polyploidy arose early in the entire clade, as suggested by Goldblatt (1981), Lavin et al. (2005) dated the common ancestor of Crotalarieae and Genisteae at around 41 MYA. On the other hand, Goldblatt also noted (1981, p. 452) that ‘‘basic numbers for these genera (Genista, Ulex, Cytisus) are however in the diploid range and a basic number of x = 12 for the group as a whole and for several genera has been suggested …’’. Genomic data for members of Genisteae should eventually allow the determination of the number and relative timing of polyploid events in the group. 471 The dalbergioid s.l. clade is split into two major subclades, Amorpheae and a second clade comprising Adesmieae, Aeschynomeneae, and many members of the polyphyletic Dalbergieae (Lewis et al. 2005). The entire dalbergioid clade is dominated by 2n = 20 species. Within Amorpheae, polyploidy occurs in genera from each of the major subclades described by McMahon (2005). In the daleoid clade, Dalea has 2n = 14, 16, 28, and 42. Spellenberg (1981) hypothesized that tetraploids and hexaploids of D. formosa (2n = 28, 42) were autopolyploids derived from the diploid (2n = 14) cytotype. In the amorphoid clade, Amorpha includes both diploids and polyploids (2n = 20, 40), all native to the New World. The widespread A. fruticosa is exclusively polyploid and morphologically complex (Wilbur 1975). It has become an invasive weed in Europe (e.g., Hulina 2010), illustrating a common feature of polyploidy (e.g., Pandit et al. 2011; te Beest et al. 2011). Its relationships to other members of the genus appear to be complex, sharing chloroplast haplotypes with different sympatric diploids across its range (Straub and Doyle, unpublished data). Studies of the A. georgiana complex UN Editor Proof 162 Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 163/179 163 547 9.4.2.3 Baphioids 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 548 549 550 551 552 PR OO 520 D 518 519 TE 517 EC 515 516 CO RR 514 F 546 identified mixed populations of diploids and polyploids in what had been assumed to be an exclusively diploid species; these included an apparent allopolyploid between A. georgiana and A. herbacea (Straub and Doyle 2009). Straub (unpublished data) has identified additional polyploid species in the genus and hypothesized their origins. The core dalbergioid clade is split into the small Adesmia clade (six genera) and the much larger clade comprising the Dalbergia and Pterocarpus sister clades (Lewis et al. 2005). Polyploidy occurs in all three clades. In the Adesmia clade, Adesmia includes both diploids and polyploids (2n = 20, 40), and Amicia is exclusively polyploid (2n = 38). In the Dalbergia clade, Smithia (2n = 38) is exclusively polyploid and is considered closely related to Kotschya (Lewis et al. 2005), a genus that includes species with chromosome numbers indicative of polyploidy and aneuploidy (2n = 28, 30, 36, 40). These two genera are grouped with Aeschynomene species (Lavin et al. 2001), a genus that also includes diploids and polyploids (2n = 18, 20, 40). Another dalbergioid-clade genus, Ormocarpum (2n = 24, 26), was not listed in Goldblatt’s discussion of polyploid genera, but could potentially be a cryptic polyploid with aneuploid reduction. Information on chromosome numbers of other members of the Ormocarpum group (Thulin and Lavin 2001) would be useful in addressing this issue. The Pterocarpus clade also includes several genera with both diploids (including presumed aneuploids) and polyploids scattered among its subclades: Platymiscium (2n = 16, 18, 20, 32), Pterocarpus (2n = 22, 24, 44), Geoffroea (2n = 20, 60), and Arachis (2n = 20, 40). Arachis includes the tetraploid peanut or groundnut (A. hypogaea), as well as three other tetraploid species, one of which (A. glabrata) is a tropical forage crop. Peanut is hypothesized to be an allopolyploid derived from A- and B-genome species (e.g., Burow et al. 2009). Seijo et al. (2007) provide a useful summary of hypotheses concerning the origin (or origins) of peanut; controversy exists concerning such issues as the exact progenitor species of both homoeologous genomes, mode and number of origins, and whether there was subsequent introgression from wild species into the cultigen. They identified likely diploid progenitors of A. hypogaea using GISH and studied meiotic behavior of two other tetraploids (Ortiz et al. 2011) and a spontaneous autotriploid of A. pintoi (Lavia et al. 2011). 513 This small group appears to be interesting from the standpoint of polyploidy. Chromosome numbers are known from six of its seven genera. Of these, four are listed as 2n = 22, one (Dalhousiea) is 2n = 44, and Baphia was reported in Polhill and Raven (1981) to have both numbers, though no reports for any of the 47 species of the genus exists in IPCN. UN Editor Proof 9 Polyploidy in Legumes Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 164/179 553 J. J. Doyle 9.4.2.4 Mirbelioids 576 9.4.2.5 Hologalegina 562 563 564 565 566 567 568 569 570 571 572 573 574 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 PR OO 561 D 559 560 TE 557 558 EC 556 The Hologalegina clade includes robinioids and the Inverted Repeat Loss Clade (IRLC; named for the absence of a prominent feature of the chloroplast genome). Goldblatt (1981) concluded that ‘‘Species polyploidy is overwhelmingly concentrated in temperate to cool Eurasia’’ so it is not surprising that this largest clade of legumes, which includes many temperate genera, has a higher frequency of polyploidy than the family average, nearly 40 % in both of its major subclades (Table 9.1). The robinioids comprise two clades: one with Sesbania plus Loteae (including Coronilleae), the other being Robinieae (s.s.). Diploid chromosome numbers vary considerably within Robinieae s.s., and Goldblatt (1981) suggested several possible base numbers, the most likely being x = 10 or 11. There is one apparently exclusively polyploid genus, Poissonia, which Goldblatt (1989) counted as (2n = ca. 32). Although only numbers of 2n = 10 and 11 were given for Robinia in Polhill and Sousa (1981), more recent counts of 2n = 30 for R. hispida suggest polyploidy within this small genus. Even with the limited sampling of Loteae in Wojciechowski et al. (2004), it is clear that numerous problems exist with the genera as circumscribed, notably that CO RR 555 F 575 Polyploidy in the mirbeliod clade (23 %) is close to the average for the whole family (Table 9.1). Isotropis (2n = 16, 18, 32) is sister to the large ‘‘Pultenaea s.l. group’’ in Orthia et al. (2005), which includes Oxylobium (2n = 16). Chandler et al. (2001) sank both Brachysema (2n = 16, 32) and the monotypic Jansonia (2n = 32) in Gastrolobium, previously a genus with only 2n = 16 species. Chandler et al. (2001) placed the single sampled species of Jansonia sister to Brachysema celsianum (not listed in IPCN), in a clade that also included B. praemorsum; that species is a diploid at 2n = 16, as is at least one species in the sister clade, Nemcia coriacea. Thus, it is likely that polyploidy has arisen more than once just within Gastrolobium s.l., and another time in Isotropis. Among several ‘‘strongly paraphyletic’’ Pultenaea s.l. genera, nearly all of which are 2n = 16, are two genera with polyploidy: Pultenaea s.s. (2n = 8, 12, 14, 16, 18, 27, 32) and Chorizema (2n = 16, 32). Sorting out how the various chromosome numbers in Pultenaea s.s. are related will be of considerable interest but will require more complete phylogenies than appear to be available at present. Of the 10 Chorizema species (out of 27 in the genus) listed in IPCN, polyploidy is only reported from C. aciculare, which has both diploid and tetraploid cytotypes. Smaller genera in Pultenaea s.l. with known polyploidy are Eutaxia (2n = 16, 32) and Dillwynia (2n = 14, 21, 28), which are in the same weakly supported clade in Orthia et al. (2005). One of the two Dillwynia species (D. phylicoides) included in the Orthia et al. (2005) tree has both diploid and tetraploid cytotypes listed in IPCN. 554 UN Editor Proof 164 Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 165/179 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 F PR OO 601 602 D 600 TE 598 599 Anthyllis (2n = 10, 12, 14, 16, 28) and Ornithopus (2n = 14) are nested within Lotus (2n = 10, 12, 14, 24, 28). Other genera with known polyploidy are Coronilla (2n = 12, 20, 24); Hippocrepis (2n = 14, 28); Dorycnium (2n = 14, 28; both cytotypes in D. axilliflorum), and Scorpiurus (S. muricatus has 2n = 14, 16, 28; other species are 2n = 14 or 28). Degtjareva et al. (2006) provided phylogenetic hypotheses for Lotus, with sampling of other genera, but did not discuss polyploidy. Rosello and Castro (2008) discussed polyploidy in the flora of the Balearic Isles, among which are species of Anthyllis and Coronilla. The genus Lotus includes the genomic model legume, L. japonicus (Sato et al. 2008), which is part of the L. corniculatus (birdsfoot trefoil) polyploid complex. Grant and Small (1996) summarized many studies of this complex and concluded that it was a fertile topic for further study, particularly to identify the diploid progenitors of L. corniculatus itself, which they considered to be an allopolyploid. Gauthier et al. (1998a, b) discussed evolutionary patterns in the L. corniculatus/L. alpinus polyploid complex in the Alps of Europe; they described morphological and genetic consequences of autopolyploidy in L. alpinus and suggested introgression at the tetraploid level between it and L. corniculatus. The majority of genera and species in the IRLC clade belong to two sister clades in Wojciechowski et al. (2004): one includes the Astragalean clade (Astragalus and allies) and Hedysareae, and the second includes the Vicioid clade. The remainder of the IRLC phylogeny, moving successively further from these clades, consists of a clade with Wisteria (2n = 16) and one species of Callerya, followed by a clade with Glycyrrhiza (2n = 16) and a second species of Callerya. The Astragalean clade has extensive polyploidy. Perhaps most striking is the clade that includes the New Zealand endemic tribe Carmichaelieae plus the Australian Swainsona and an additional New Zealand genus, Montigera, all of which are polyploid (Wagstaff et al. 1999). The only exclusively polyploid genera that Goldblatt (1981) listed for the IRLC clade belong to this clade: Swainsona (2n = 32); Clianthus (2n = 32); Carmichaelia (2n = 32, ca. 96), Chordospartium (2n = 32); and Corallospartium (2n = 32), with the latter two subsumed in Carmichaelia in Lewis et al. (2005). Wagstaff et al. (1999) concluded that the New Zealand radiation was recent, involved an already polyploid colonizer, and may have been associated with orogeny and glaciation. Elsewhere in the Astragalean clade are two large genera with extensive polyploidy and aneuploidy, Oxytropis (300–400 species: 2n = 16, 32, 48, 64, 96) and Astragalus (ca. 2500 species: 2n = 16, 22, 24, 26, 28, 32, 44, 48, 64). In Astragalus, polyploidy appears to be more common among Old World than among New World species. Wojciechowski (2005) summarized phylogenetic results for this huge genus, showing that aneuploid species form a clade. According to Gohil and Ashraf (2008), polyploidy occurs in around only 17 % of Astragalus species. However, Astragalus is one of the largest genera of plants with as many as 2500 species (Lewis et al. 2005), so if this percentage is correct, then there are over 400 polyploid species in the genus. There does not seem to be a comprehensive phylogeny that discusses origins of polyploidy in Oxytropis. However, a series of papers describe autopolyploidy, including multiple autopolyploid origins, in EC 596 597 165 CO RR 594 595 UN Editor Proof 9 Polyploidy in Legumes Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 166/179 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 F PR OO 644 645 D 642 643 TE 641 Oxytropis chankaensis (e.g., Artyukova et al. 2011). Jorgensen et al. (2003) suggested ‘‘a scenario of multiple formations of polyploids, possibly including hybridization among diverged Alaskan Oxytropis populations.’’ Within the Hedysareae, phylogenetic studies of Caragana (2n = 16, 24, 32, 48) suggest that polyploidy is confined to a single group of species, and that triploids, tetraploids, and hexaploids may all be autopolyploid in origin (Zhang et al. 2009). Neither Hedysarum (2n = 14, 16, 48) nor Onobrychis (2n = 14, 16 28, 32) appears to be monophyletic on the basis of nrDNA ITS phylogenies (Ahangarian et al. 2007). Hejazi et al. (2010) discussed karyotypic evolution in diploid and polyploid species but did not provide a phylogenetic context or identify origins of polyploids. Based on IPCN listings, most polyploidy reported for Hedysarum in IPCN appears to involve multiple cytotypes within a single species (e.g., H. arcticum and H. hedysaroides, both 2n = 14, 28; H. dasycarpum and H. mackenziei, both 2n = 16, 32; H. gmelinii, 2n = 16, 28, 56), but some species are exclusively polyploid (e.g., H. inundatum, 2n = 28). Similarly, in Onobrychis there is variation within species (e.g., O. aequidentata, 2n = 14, 16, 28; O. arenaria, O. bobrovii, 2n = 14, 28; O. crista-galli, 2n = 16, 32), with other species being exclusively polyploid (e.g., O. biebersteinii, O. cyri, O. dielsii, all 2n = 28). There appear to be no phylogenies or evolutionary studies of polyploidy in Alhagi (2n = 16, 28; the latter number is not listed in IPCN). The majority of the vicioid clade forms two sister clades, one with Fabeae (Vicieae) plus Trifolium, and a second comprising the remaining Trifolieae genera; polyploidy occurs in both clades. Successive sisters to this clade (Cicer, Galega, and Parochetus) are all 2n = 16. Vicia includes both diploids and tetraploids (2n = 10, 12, 14, 24, 28), but polyploidy was considered rare in the genus by Kupicha (1981). Indeed, Vicia is best known for its extensive non-polyploid variation in genome size (Chooi 1971; Neumann et al. 2006), which shows only weak correlation with ploidy: diploid (2n = 14) V. peregrina has a genome size of 9.48 pg/1C, double that of tetraploid (2n = 24) V. tenuifolia (4.73 pg/1C). Endo et al. (2008) did not discuss either issue in their phylogenetic study of New Wold Vicia. Travnicek et al. (2010) studied the history of polyploidy in V. cracca, determining the ploidy of over 6,500 individuals at more than 250 localities in Europe and mapping the distributions of diploids, triploids, and tetraploids; they noted the rarity of triploids, suggesting strong reproductive barriers between diploids and tetraploids. Polyploidy is also noted to be rare in Lathyrus (2n = 14, 28, 42; Kupicha 1981). Gutierrez et al. (1994) hypothesized autopolyploid origins of L. pratensis and L. palustris from conspecific diploids, but an allopolyploid origin of L. venosus from two diploid species (L. ochroleucus and L. palustris). Only 2n = 14 counts are listed for the closely related Pisum in IPCN, for which Kupicha (1981) listed polyploidy as ‘‘rare.’’ Turini et al. (2010) reconstructed nrDNA ITS and chloroplast phylogenies for 69 of the 86 species of Ononis (2n = 16, 20, 30, 32, 60, 64) and identified several well-supported clades. They concluded that, ‘‘Unfortunately, only limited information is available … on chromosome numbers to test support for these groups’’. EC 640 CO RR 639 J. J. Doyle UN Editor Proof 166 Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 167/179 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 F 690 PR OO 689 D 687 688 TE 686 However, chromosome counts are available for nearly half of the species in their phylogeny in IPCN, and some conclusions can be drawn. For example, the clade that is strongly supported as sister to the remainder of the genus in their nrDNA ITS tree includes only polyploids (O. tridentata and O. fruticosa are both 2n = 30; O. rotundifolia is 2n = 32), suggesting that the genus as a whole could be polyploid. Only three species have low, potentially non-polyploid numbers in IPCN. These occur in different clades, and in two cases species with low numbers have higher numbers as well (O. variegata, 2n = 16, 30; O. ornithopodioides, 2n = 20, 32), raising the possibility that they are independent reductions from typical polyploid numbers. The exception, O. adenotricha, is only reported as 2n = 16; its position varies between the nrDNA ITS and trnL-F trees of Turini et al. (2010), being sister to the tridentata clade in the trnL-F tree; however, this entire group is not resolved as sister to the remainder of the genus in that tree. Elsewhere in the genus, O. spinosa has multiple cytotypes (2n = 30, 32, 60), whereas O. pendula is only known at 2n = 64. Kloda et al. (2008) studied patterns of genetic diversity in several diploid and polyploid species in England and concluded that gene flow was occurring within ploidy levels, but not between diploids and tetraploids. Medicago includes the genomic model legume, M. truncatula (Young et al. 2011). Steele et al. (2010) provided a phylogeny for Medicago (2n = 14, 16, 32, 48), including multispecies sampling of its sister clade, which comprises the interdigitated species of the two paraphyletic genera Melilotus (2n = 16, 24, 32; though tetraploids are not reported in IPCN) and Trigonella (2n = 16, 28, 32, 44). Aneuploid change from 2n = 16 to 2n = 14 has occurred several times in Medicago (Steele et al. 2010). Polyploidy is concentrated in a clade that comprises most species of sect. Medicago, along with M. arborea (sect. Dendrotelis); an additional polyploid species, M. scutellata, occurs in the clade sister to this sect. Medicago clade. Rosato et al. (2008) used fluorescence in situ hybridization (FISH) to study relationships between polyploids and diploids in sect. Dendrotelis. Some Medicago species are exclusively polyploid whereas others possess multiple cytotypes. The M. sativa complex, which includes cultivated autotetraploid alfalfa (M. sativa ssp. sativa) as well as other diploid (2n = 16) and autopolyploid (2n = 32) species and their hybrids, has been the focus of several recent studies (Sakiroglu et al. 2010; Havananda et al. 2010, 2011, and unpublished data). Two major autopolyploid pairs in the complex are: (1) M. s. caerulea and M. s. sativa, both with blue flowers and coiled pods, distinguishable by the larger size of the tetraploid (M. s. sativa) for several characters; and (2) M. s. falcata, a yellowflowered taxon with falcate pods whose diploid and polyploid cytotypes are indistinguishable morphologically. Interestingly, although M. s. sativa and M. s. caerulea are undifferentiated for chloroplast haplotypes, the two cytotypes of M. s. falcata possess nearly mutually exclusive sets of haplotypes, with haplotypes in the tetraploid most likely derived by introgression from M. prostrata, a species from outside the complex (Havananda et al. 2011). Jenczewski et al. (1999) reported gene flow between wild and cultivated M. sativa populations; however, based on chloroplast data, there does not appear to be significant gene flow between blue- and yellow-flowered taxa in the complex either at the diploid or EC 685 167 CO RR 684 UN Editor Proof 9 Polyploidy in Legumes Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 168/179 J. J. Doyle 743 9.4.2.6 Indigofereae 736 737 738 739 740 741 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768 769 770 PR OO 735 D 733 734 Schrire et al. (2009) provided a detailed phylogeny for this tribe, a monophyletic group that is sister to the millettioid clade. Schrire et al. (2009) did not comment on chromosomal variation or polyploidy, but numerous records are readily available in IPCN, and mapping these onto the phylogeny provides some insights into cytological evolution of the group. The tribe is dominated by the very large genus Indigofera (ca. 700 species), which Goldblatt (1981) and Polhill (1981b) listed as having 2n = 14, 16, 32, 48. The higher numbers thus would be interpreted as representing tetraploids and hexaploids. However, Frahm-Leliveld (1966), summarizing the cytotaxonomy of the tribe, cited two x = 6 species, I. macrocalyx (2n = 12) and I. emarginella (2n = 24), and concluded that ‘‘… the 48-chromosome Himalayan and EastAsiatic shrubby Indigoferas may not be hexaploids with base number x = 8, but octoploids in an x = 6 range.’’ None of the species is listed in IPCN, but Reddy and Revathi (1993) reported 2n = 12 for I. anil, confirming the presence of x = 6 in the genus. The Schrire et al. (2009) phylogeny does not support the Frahm–Leliveld (1966) hypothesis. One of the two 2n = 12 species, I. macrocalyx, is placed in the large Palaeotropical clade of Schrire et al. (2009) and is sister to a group of species that includes I. pulchra (2n = 16). All three sampled species with 2n = 48 are in the Palaeotropical clade, but are placed nowhere near I. macrocalyx. In the Pantropical clade, I. rhynchocarpa (2n = 16) is sister to the clade that includes I. emarginella, which is on a long branch sister to several other species; the only other species counted from this subclade is also 2n = 16. Thus, there is no evidence that 2n = 48 species are derived from x = 12 species. Tetraploids based on x = 8 are scattered throughout the phylogeny (Schrire et al. 2009), supporting the observation of Frahm-Leliveld (1966) that 2n = 32 is common in the genus. In the Palaeotropical clade, I. atriceps (2n = 32) is in a TE 732 EC 731 CO RR 730 F 742 tetraploid levels, despite the existence of morphologically intermediate hybrid subspecies (Havananda et al. 2011, and unpublished data). Much is known about the genetics of polyploidy in alfalfa, where unreduced gametes have received considerable study as a breeding tool (e.g., Bingham 1972; Veronesi et al. 1986; Tondini et al. 1993; Calderini and Mariani 1997). Ellison et al. (2006) constructed a phylogeny of Trifolium (2n = 10, 14, 16, 28, 32) that included 218 of its ca. 255 species, as well as species from 11 genera of the vicioid clade. They showed that the genus is monophyletic; incongruence within the genus between nuclear and chloroplast markers suggests considerable hybridization. They also hypothesized a minimum of 19 shifts to aneuploidy and 22 instances of polyploidy from a base number of 2n = 16. They identified the progenitors of two important species, both shown to be allopolyploids: the widespread weed, T. dubium, and the most commonly cultivated clover species, T. repens (Ellison et al. 2006). 729 UN Editor Proof 168 Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 169/179 169 794 9.4.2.7 Millettioids 780 781 782 783 784 785 786 787 788 789 790 791 792 795 796 797 798 799 800 801 802 803 804 805 806 807 808 809 810 811 812 PR OO 778 779 D 776 777 TE 775 EC 773 774 With the recognition that Wisteria and Callerya are part of the IRLC, and that Cyclolobium and Poecilanthe belong in the Brongniartieae, chromosome numbers in the Millettieae (Tephrosieae in Polhill and Raven (1981)) are mostly 2n = 20 or 22, with 2n = 24 in Xeroderris, though many genera have no reported counts in IPCN. Interestingly, Xeroderris is placed as sister to the remainder of the entire millettioid clade in Wojciechowski et al. (2004), suggesting that 2n = 20 or 22 may be synapomorphic for the remainder of the millettioid clade (including phaseoloids; see below). Millettieae comprises the bulk of one of the two major millettioid clades (core millettioids), along with Abrus (Abreae), and much of the subtribe Diocleinae of tribe Phaseoleae. Both diploid and tetraploid cytotypes (2n = 22, 44) have been reported in three species of the ca. 40 IPCN records for the large (ca. 350 spp.) genus Tephrosia (e.g., Srivastav and Raina 1986). The low frequency of polyploidy in the core millettioid clade (5 %) is nearly identical to the frequency in caesalpinioids, both of which are largely woody, tropical groups. The other large clade (phaseoloids) contains most of the tribe Phaseoleae as well as the tribes Desmodieae and Psoraleeae and is dominated by 2n = 20 or 22 counts. Polyploidy is more frequent in the phaseoloid clade (17 %), but still less than half as common as in Hologalegina (Table 9.1). Within the phaseoloid clade, CO RR 772 F 793 subclade that also includes diploids. In another subclade, I. mysorensis includes both 2n = 16 and 32 cytotypes; other two members of its subclade are diploid. Indigofera microcalyx, in yet another subclade, is 2n = 32; no other members of its subclade has been counted, but the only counts from its sister clade are diploid. In the pantropical clade, a small subclade in Schrire et al. (2009) includes I. koreana (2n = 32), I. grandiflora (2n = 32, 48), I. decora (2n = 48 in Choi and Kim (1997) but not listed in IPCN), as well as I. venulosa (no count available) and I. kirilowii (2n = 16). Topologies differ between Schrire et al. (2009) and Choi and Kim (1997), who focused on this group of mostly Korean endemics. Choi and Kim (1997) listed I. grandiflora as 2n = 16, and given this count their topology could suggest independent derivation of polyploidy in I. koreana (from I. grandiflora) and I. decora (from I. venulosa if it is diploid). An alternative explanation is a single derivation of polyploidy within this clade. Elsewhere in the pantropical clade, I. heterantha (2n = 48) is sister to I. hebepetala (2n = 16); the clade sister to these two species includes I. amblyantha (2n = 48) and I. cassioides (2n = 16). In a different subclade, I. suffruticosa is reported to have both 2n = 16 and 32 cytotypes; the closest reported species to it is diploid. In the Tethyan clade, I. sessiliflora (2n = 32) is the only member of its subclade with a count in IPCN. Indigofera hochstetteri has both 2n = 16 and 32 counts; its sister species, I. arabica, is diploid. Indigofera angulosa is 2n = 32; no other species in its clade has counts in IPCN. Thus, there appear to be no large clades composed exclusively of polyploids in Indigofera. Instead, as in other large legume genera, polyploidy is sporadic. 771 UN Editor Proof 9 Polyploidy in Legumes Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 170/179 820 821 822 823 824 825 826 827 828 829 830 831 832 833 834 835 836 837 838 839 840 841 842 843 844 845 846 847 848 849 850 851 852 853 854 855 856 F PR OO 818 819 D 816 817 TE 815 chromosome numbers suggest that several genera are polyploid. The best known of these is Glycine, with around 30 species whose lowest chromosome numbers are 2n = 38 and 40, in contrast to most of its phylogenetic neighbors (e.g., Doyle et al. 2003; Stefanovic et al. 2009) which are typical millettioids with 2n = 20 or 22. As noted above, genomic data confirm the presence of two cycles of polyploidy in G. max (soybean) since the origin of the legumes. The more recent of these has resulted in homoeologous gene pairs that diverged around 10 MYA (Shoemaker et al. 2006; Egan and Doyle 2010), setting a maximum date for the polyploidy event, with the minimum date set by the earliest divergence of the various Glycine species around 5 MYA (Innes et al. 2008; Doyle and Egan 2010). Phylogenetic evidence is consistent either with autopolyploidy or with allopolyploidy from extinct species more closely related to one another than to any extant genera outside of Glycine (Straub et al. 2006). The presence of two classes of centromeric heterochromatin repeats suggests that Glycine could be an allopolyploid, with the two repeat types each derived from one of the diploid progenitor species (Gill et al. 2009). Such a hypothesis is difficult to test due to the extensive rearrangement of homoeologous segments in the soybean genome (Schmutz et al. 2010) and also requires complex patterns of concerted evolution among repeats on different chromosomes. At least three other phaseoloid genera are likely to be exclusively polyploid based on chromosome number alone (Lackey 1981). In Erythrina (coral bean), all sampled species are 2n = 42. The single count in IPCN for the small genus Cologania is 2n = 44, and the monotypic Teyleria is also 2n = 44 (Kumar and Hymowitz 1989). Goldblatt (1981) considered Calopogonium, with counts of 2n = 36 and ca. 37 in C. mucunoides, to be a polyploid, presumably with aneuploid reduction from a base of x = 10; however, Gill and Husaini (1986) reported a count of 2n = 24, which could suggest a more recent derivation of polyploidy within the genus. Similarly, counts of 2n = 28—considered polyploid in T. mollis by Kumari and Bir (1990)—predominate in Teramnus species, though T. labialis is variously listed as 2n = 20, 22, and 28. Strongylodon is also 2n = 28. Polyploidy appears to be rare within Phaseoleae genera. Even relatively large and well-studied genera such as Rhynchosia (ca. 230 species), Phaseolus (60–65 species), and Vigna (ca. 100 species) were reported in Polhill and Raven (1981) as being exclusively diploid, though Sen and Bhattacharya (1988) later reported a count of 2n = 44 in V. glabrescens. Polyploidy has also been reported within species of Amphicarpaea and Neonotonia by Kumar and Hymowitz (1989; both 2n = 22, 44). Apios americana includes both diploid and triploid cytotypes (2n = 22, 33); Joly and Bruneau (2004) reported multiple origins of autotriploidy and high heterozygosity in this species. Glycine not only is a relatively recent polyploid at the generic level (see above) but also includes several recently formed allopolyploid species whose genomic relationships to extant diploids have been worked out using molecular phylogenies (reviewed by Doyle et al. 2004), and which are the focus of physiological and transcriptomic studies (Coate and Doyle 2010; Ilut et al. (in press)). EC 814 CO RR 813 J. J. Doyle UN Editor Proof 170 Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 171/179 865 866 867 868 869 870 871 872 873 874 875 876 877 878 879 880 881 882 883 884 885 886 887 888 889 890 891 892 893 894 895 896 897 F PR OO 864 9.5 Searching for Cryptic Polyploidy in the Phaseoloid Legumes Clearly, all papilionoids are fundamentally polyploid, even those with low chromosome numbers. The tempo and mechanism(s) of chromosomal diploidization are unknown (Doyle et al. 2008; Soltis et al. 2010), and without that information it is difficult to estimate the prevalence of cryptic polyploidy. As noted above, consideration of the divergence times for major lineages suggests that the rate of chromosomal diploidization is rapid—likely 10 MY or less. On the other hand, ‘‘polyploid’’ chromosome numbers have persisted for at least 5–10 MY in Glycine (Fig. 9.2). A cryptic polyploid papilionoid legume is a taxon that has experienced an additional polyploidy event subsequent to the ca. 50 MYA papilionoid WGD but has a low chromosome number typical of its clade. Thus, in the phaseoloid clade, cryptic polyploids would have chromosome numbers of 2n = 20 or 22. Polyploids on the way to diploidization would have chromosome numbers between these numbers and 2n = 40-44. As noted above, Calopogonium and Teramnus are candidates for this class; one perennial Glycine species with 2n = 38 is likely at the first stages of this process. We know from the paralog Ks profile of Glycine (Schleueter et al. 2004) that no additional polyploidy events took place between the two WGD episodes detectable in its genome. Therefore, we can infer that all of the ancestors of Glycine experienced only the papilionoid WGD; this includes the ancestors that form the backbone of the phaseoloid clade (Fig. 9.4), as well as the common ancestor of phaseoloids and Indigofereae, and also its common ancestor with the IRLC clade. Given these conditions, candidates for cryptic polyploidy include lineages that are connected to the phaseoloid backbone by branches longer than 5–10 MY. This includes several major groups, such as subtribes Phaseolinae and Cajaninae (Fig. 9.4). Initial sampling of one species of each lineage would provide information on another set of ancestors, and subsequent searches can then be focused on lineages connected to these ancestors by suitably long branches. This has now been done for Cajanus cajan (pigeonpea), which Varshney et al. (2011) have shown has no history of recent polyploidy. We are sampling other phaseoloids using 454 D 862 863 TE 861 Polyploidy occurs within at least one genus of Desmodieae, Lespedeza, which was listed in Polhill and Raven (1981) as 2n = 18, 20, 22, 36. However, IPCN gives higher numbers, for example L. bicolor with both 2n = 22 and 42 cytotypes, as well as L. daurica and L. potaninii, both exclusively 2n = 42. Triploidy occurs in Campylotropis polyantha var. leiocarpa (2n = 22, 33) and possibly in the genus Pseudarthria, listed as 2n = 22, 26, 34. Only diploid counts (2n = 20, 22) were reported by Ohashi et al. (1981) from the large (ca. 275 spp.) genus, Desmodium. However, additional counts are found in IPCN, both at the diploid (2n = 24, 26) and tetraploid levels, the latter in D. styracifolium (2n = 42) and D. incanum (2n = 22, 44). EC 859 860 171 CO RR 857 858 UN Editor Proof 9 Polyploidy in Legumes Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 172/179 Bituminaria Cullen Glycine 2n = 38, 40 Teramnus 2n = 28 Amphicarpaea 2n = 22, 44 Pueraria lobata Neonotonia 2n = 22, 44 Pachyrhizus Pueraria phaseoloides Pseudovigna Dumasia Cologania 2n = 44 Macroptilium Vigna Macrotyloma Psophocarpus Erythrina 2n = 42 Cajanus Bolusafra Hardenbergia Kennedia Lespedeza Campylotropis Desmodium pauciflorum Desmodium barbatum Mucuna Shuteria Apios 2n = 22, 33 20 10 million years TE D PR OO } polyploidy F J. J. Doyle Editor Proof 172 CO RR EC Fig. 9.4 Phylogeny and polyploidy in the phaseoloid clade. Topology and dates of the chronogram are taken from Stefanovic et al. (2009). Chromosome numbers of the species used in that study are color coded as follows: Green, 2n = 20; blue, 2n = 22; black, 2n = 18; red, known or possible polyploid numbers, with the numbers shown following the taxon name. Species known to have multiple cytotypes are indicated by multiple colors corresponding to chromosome numbers listed above. The two numbers shown for Glycine are from different species, only one of which (G. max, 2n = 40) was used in the Stefanovic et al. (2009) study. The range of dates for the polyploidy event in Glycine is indicated. Yellow dots indicate ancestral nodes that lacked any polyploidy event subsequent to the papilionoid WGD 903 9.6 Conclusions 899 900 901 904 905 906 UN 902 transcriptome sequencing, including taxa with both low and high chromosome numbers; for the latter we wish to estimate maximum ages of polyploidy. Thus far we have not found examples of cryptic polyploidy, but have determined that polyploidy in Erythrina probably took place on roughly the same time scale as in Glycine (\10 MYA; Egan and Doyle, unpublished data). 898 Among the most persistent questions concerning polyploidy in plants are how successful the phenomenon is as an evolutionary mechanism. Is polyploidy a ticket to innovation, adaptation, invasiveness, survival in the face of global catastrophes, Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 173/179 173 942 943 944 945 946 947 948 Acknowledgments I thank many colleagues and lab members for discussions of polyploidy, and Jane Doyle for her support and encouragement. I also thank Jane Doyle, Sue Sherman-Broyles, Iben Sorensen, and Toby Pennington for critical reading of the manuscript, and Melissa Luckow for help with mimosoid systematics. I am grateful for many years of funding from the National Science Foundation for work on polyploidy, most recently grants DEB-0948800, IOS-0939423, IOS-0822258, and IOS-0744306. I thank Doug Soltis for helpful suggestions in review of the manuscript. 913 914 915 916 917 918 919 920 921 922 923 924 925 926 927 928 929 930 931 932 933 934 935 936 937 938 939 940 PR OO 912 D 911 TE 910 EC 909 CO RR 908 F 941 or is it an evolutionary dead end … or both? Based on rates of polyploid formation and extinction in the phylogenetic record, Mayrose et al. (2011) conclude that ‘‘polyploidy is most often an evolutionary dead end, but the possibility remains that the expanded genomic potential of those polyploids that do persist drives longterm evolutionary success.’’ Legumes may illustrate both of these points. The most diverse and species-rich clade of this third largest family of flowering plants, the core papilionoids, is ancestrally polyploid. Clearly, the ancestor of this group of around 450 genera and 13,000 species, like the ancestor of seed plants and the ancestor of flowering plants (Jiao et al. 2011), was most emphatically not a ‘‘dead end.’’ It remains to be determined whether there is a perfect correlation between the papilionoid polyploidy event and the origin of nodulation in core papilionoids, and it will take much more work to demonstrate that the two are causally related (Doyle 2011; Young et al. 2011). It is also clear that nodulation is not sufficient to explain the explosive radiation of papilionoid legumes, because other nodulating groups both in legumes and elsewhere in the rosids have not proliferated to the same extent as papilionoids (Doyle 2011). Despite the obvious success of the core papilionoid lineage, the pattern of evolution within the core papilionoids suggests that polyploidy has not been a major feature in establishing new lineages, similar to the conclusion of Mayrose et al. (2011) for angiosperms generally. It is not that polyploidy is rare within the family—indeed, around a quarter of all legumes for which chromosome data are available have one or more species that are polyploid (Table 9.1). However, much of the polyploidy in the family occurs as single polyploid genera embedded within diploids, as scattered species within genera, or as multiple cytotypes within species. Two significant exceptions are the Genisteae, which may be entirely polyploid and within which nearly all genera show a propensity for polyploidy and aneuploidy, and the lineage that includes the IRLC tribe Carmichaelieae. The largest papilionoid genera, including Astragalus, are not fundamentally polyploid. ‘‘Success’’ is a very ambiguous term and can be measured in many ways. Species with short evolutionary histories that have not been involved in subsequent speciation, yet have invaded extensive new territories and had major impact on the environment, certainly could be considered ‘‘successful.’’ Many plant polyploids, including genera and species of legumes, fit this description. So does Homo sapiens. 907 UN Editor Proof 9 Polyploidy in Legumes Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 174/179 TE D PR OO F Ahangarian S, Osaloo SK, Maassoumi AA (2007) Molecular phylogeny of the tribe Hedysareae with special reference to Onobrychis (Fabaceae) as inferred from nrDNA ITS sequences. Iran J Bot 13:64–74 Ainouche A, Bayer RJ, Misset M- (2004) Molecular phylogeny, diversification and character evolution in Lupinus (Fabaceae) with special attention to Mediterranean and African lupines. Plant Syst Evol 246:211–222 Artyukova EV, Kozyrenko MM, Kholina AB, Zhuravlev YN (2011) High chloroplast haplotype diversity in the endemic legume Oxytropis chankaensis may result from independent polyploidization events. Genetica 139:221–232 Bell CD, Soltis DE, Soltis PS (2010) The age and diversification of the angiosperms re-revisited. Am J Bot 97:296–313 Bello MA, Bruneau A, Forest F, Hawkins JA (2009) Elusive relationships within order Fabales: phylogenetic analyses using matK and rbcL sequence data. Syst Bot 34:102–114 Bertioli D, Moretzsohn M, Madsen L, Sandal N, Leal-Bertioli S, Guimaraes P, Hougaard B, Fredslund J, Schauser L, Nielsen A, Sato S, Tabata S, Cannon S, Stougaard J (2009) An analysis of synteny of Arachis with Lotus and Medicago sheds new light on the structure, stability and evolution of legume genomes. BMC Genomics 10:45 Bessega C, Vilardi JC, Saidman BO (2006) Genetic relationships among American species of the genus Prosopis (Mimosoideae, Leguminosae) inferred from ITS sequences: evidence for longdistance dispersal. J Biogeogr 33:1905–1915 Bingham ET (1972) Sexual poly ploidy in Medicago-Sativa-D. Genetics 71:S5 Bisby FA (1981) Genisteae. In: Polhill RM, Raven PH (eds) Advances in legume systematics, Part 1. Royal Botanic Gardens, Kew, pp 409–425 Blanc G, Wolfe KH (2004) Widespread paleopolyploidy in model plant species inferred from age distributions of duplicate genes. Plant Cell 16:1667–1678 Boatwright JS, Van Wyk B- (2011) The systematic position of Sophora inhambanensis (Fabaceae: Sophoreae). S Afr J Bot 77:249–250 Boatwright JS, Savolainen V, Van Wyk B, Schutte-Vlok AL, Forest F, van der Bank M (2008) Systematic position of the anomalous genus Cadia and the phylogeny of the tribe Podalyrieae (Fabaceae). Syst Bot 33:133–147 Boff T, Schifino-Wittmann MT (2003) Segmental allopolyploidy and paleopolyploidy in species of Leucaena Benth: evidence from meiotic behaviour analysis. Hereditas (Lund) 138:27–35 Brown GK, Clowes C, Murphy DJ, Ladiges PY (2010) Phylogenetic analysis based on nuclear DNA and morphology defines a clade of eastern Australian species of Acacia s.s. (section Juliflorae): the ‘Acacia longifolia group’. Aust Syst Bot 23:162–172 Bruneau A, Mercure M, Lewis GP, Herendeen PS (2008) Phylogenetic patterns and diversification in the caesalpinioid legumes. Botany-Botanique 86:697–718 Burow MD, Simpson CE, Faries MW, Starr JL, Paterson AH (2009) Molecular biogeographic study of recently described B- and A-genome Arachis species, also providing new insights into the origins of cultivated peanut. Genome 52:107–119 Calderini O, Mariani A (1997) Increasing 2n gamete production in diploid alfalfa by cycles of phenotypic recurrent selection. Euphytica 93:113–118 Cannon SB, Ilut D, Farmer AD, Maki SL, May GD, Singer SR, Doyle JJ (2010) Polyploidy did not predate the evolution of nodulation in all legumes. PLoS ONE 5:e11630 Catalano SA, Vilardi JC, Tosto D, Saidman BO (2008) Molecular phylogeny and diversification history of Prosopis (Fabaceae: Mimosoideae). Biol J Linn Soc 93:621–640 Chandler GT, Bayer RJ, Crisp MD (2001) A molecular phylogeny of the endemic Australian genus Gastrolobium (Fabaceae: Mirbelieae) and allied genera using chloroplast and nuclear markers. Am J Bot 88:1675–1687 Choi B, Kim J (1997) ITS sequences and speciation on far eastern Indigofera (Leguminosae). J Plant Res 110:339–346 EC 950 951 952 953 954 955 956 957 958 959 960 961 962 963 964 965 966 967 968 969 970 971 972 973 974 975 976 977 978 979 980 981 982 983 984 985 986 987 988 989 990 991 992 993 994 995 996 997 998 999 1000 References CO RR 949 J. J. Doyle UN Editor Proof 174 Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 175/179 175 EC TE D PR OO F Chooi WY (1971) Variation in nuclear DNA content in the genus Vicia-D. Genetics 68:195–211 Coate JE, Doyle JJ (2010) Quantifying whole transcriptome size, a prerequisite for understanding transcriptome evolution across species: an example from a plant allopolyploid. Genome Biol Evol 2:534–546 Conterato IF, Schifino-Wittmann MT (2006) New chromosome numbers, meiotic behaviour and pollen fertility in American taxa of Lupinus (Leguminosae): contributions to taxonomic and evolutionary studies. Bot J Linn Soc 150:229–240 Cowan RS, Polhill RM (1981) Amherstieae. In: Polhill RM, Raven PH (eds) Advances in legume systematics, Part 1. Royal Botanic Gardens, Kew, pp 135–142 Cubas P, Pardo C, Tahiri H (2002) Molecular approach to the phylogeny and systematics of cytisus (Leguminosae) and related genera based on nucleotide sequences of nrDNA (ITS region) and cpDNA (trnL-trnF intergenic spacer). Plant Syst Evol 233:223–242 Cusma-Velari T, Feoli-Chiapella L (2009) The so-called primitive genera of Genisteae (Fabaceae): systematic and phyletic considerations based on karyological data. Bot J Linn Soc 160:232–248 Dahmer N, Simon MF, Schifino-Wittmann MT, Hughes CE, Sfoggia Miotto ST, Giuliani JC (2011) Chromosome numbers in the genus Mimosa L.: cytotaxonomic and evolutionary implications. Plant Syst Evol 291:211–220 Degtjareva GV, Kramina TE, Sokoloff DD, Samigullin TH, Valiejo-Roman CM, Antonov AS (2006) Phylogeny of the genus Lotus (Leguminosae, Loteae): evidence from nrITS sequences and morphology. Can J Bot 84:813–830 Doyle JJ (2011) Phylogenetic perspectives on the origins of nodulation. Molec Plant Microbe Interact 24:1289–1295 Doyle JL, Rauscher JT, Brown AHD (2004) Diploid and polyploid reticulate evolution throughout the history of the perennial soybeans (Glycine …. New Phytologist) Doyle JJ, Egan AN (2010) Dating the origins of polyploidy events. New Phytol 186:73–85 Doyle JJ, Flagel LE, Paterson AH, Rapp RA, Soltis DE, Soltis PS, Wendel JF (2008) Evolutionary genetics of genome merger and doubling in plants. Annu Rev Genet 42:443–461 Doyle JJ, Doyle JL, Harbison C (2003) Chloroplast-expressed glutamine synthetase in Glycine and related Leguminosae: phylogeny, gene duplication, and ancient polyploidy. Syst Bot 28:567–577 Doyle JJ, Luckow MA (2003) The rest of the iceberg. Legume diversity and evolution in a phylogenetic context. Plant Physiol (Rockville) 131:900–910 Drummond CS (2008) Diversification of Lupinus (Leguminosae) in the western new world: derived evolution of perennial life history and colonization of montane habitats. Mol Phylogenet Evol 48:408–421 Drummond CS, Eastwood RJ, Miotto STS, Hughes CE (2012) Multiple continental radiations and correlates of diversification in lupinus (Leguminosae): testing for key innovation with incomplete taxon sampling. Syst Biol 61:443–460 Egan AN, Doyle J (2010) A comparison of global, gene-specific, and relaxed clock methods in a comparative genomics framework: dating the polyploid history of soybean (Glycine max). Syst Biol 59:534–547 Ellison NW, Liston A, Steiner JJ, Williams WM, Taylor NL (2006) Molecular phylogenetics of the clover genus (Trifolium––Leguminosae). Mol Phylogenet Evol 39:688–705 Endo Y, Choi B, Ohashi H, Delgado-Salinas A (2008) Phylogenetic relationships of New World Vicia (Leguminosae) inferred from nrDNA internal transcribed spacer sequences and floral characters. Syst Bot 33:356–363 Fawcett JA, Maere S, Van de Peer Y (2009) Plants with double genomes might have had a better chance to survive the cretaceous-tertiary extinction event. Proc Natl Acad Sci U S A 106:5737–5742 Frahm-Leliveld JA (1966) Cytotaxonomic notes on the genera Indigofera L. and Cyamopsis DC. [Leguminosae]. Genetica 37:403–426 Freeling M, Thomas BC (2006) Gene-balanced duplications, like tetraploidy, provide predictable drive to increase morphological complexity. Genome Res 16:805–814 CO RR 1001 1002 1003 1004 1005 1006 1007 1008 1009 1010 1011 1012 1013 1014 1015 1016 1017 1018 1019 1020 1021 1022 1023 1024 1025 1026 1027 1028 1029 1030 1031 1032 1033 1034 1035 1036 1037 1038 1039 1040 1041 1042 1043 1044 1045 1046 1047 1048 1049 1050 1051 1052 1053 1054 UN Editor Proof 9 Polyploidy in Legumes Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 176/179 EC TE D PR OO F Gallagher RV, Leishman MR, Miller JT, Hui C, Richardson DM, Suda J, Travnicek P (2011) Invasiveness in introduced Australian Acacias: the role of species traits and geneome size. Divers Distrib 17:884–897 Gauthier P, Lumaret R, Bedecarrats A (1998a) Genetic variation and gene flow in Alpine diploid and tetraploid populations of Lotus (L. alpinus (D.C.) Schleicher/L. corniculatus L.). I. Insights from morphological and allozyme markers. Heredity 80:683–693 Gauthier P, Lumaret R, Bedecarrats A (1998b) Genetic variation and gene flow in Alpine diploid and tetraploid populations of Lotus (L. alpinus (D.C.) Schleicher/L. corniculatus L.). II. Insights from RFLP of chloroplast DNA. Heredity 80:694–701 Gill LS, Husaini (1986) Cytological observations in Leguminosae from southern Nigeria. Willdenowia 15:521–527 Gill N, Findley S, Walling JG, Hans C, Ma J, Doyle J, Stacey G, Jackson SA (2009) Molecular and chromosomal evidence for allopolyploidy in soybean. Plant Physiol 151:1167–1174 Gohil RN, Ashraf M (2008) Cytological parameters viz a viz probable modes of evolution in Astragalus L. Proc Nat Acad Sci India Sect B-Biol Sci 78:281–287 Goldblatt P (1989) Miscellaneous chromosome counts in Asteraceae Bignoniaceae Proteaceae and Fabaceae. Ann Mo Bot Gard 76:1186–1188 Goldblatt P (1981) Cytology and the phylogeny of leguminosae. In: Polhill RM, Raven PH (eds) Advances in legume systematics, Part 2. Royal Botanic Gardens, Kew, pp 427–464 Govindarajulu R, Hughes CE, Bailey D (2011a) Phylogenetic and population genetic analyses of diploid Leucaena (Leguminosae-Mimosoideae) reveal cryptic species diversity and patterns of divergent allopatric speciation. Am J Bot 98:2049–2063 Govindarajulu R, Hughes CE, Alexander P, Bailey D (2011b) The complex dynamics of ancient and recent polyploidy in Leucaena (Leguminosae). Am J Bot 98:2064–2076 Grant WF, Small E (1996) The origin of the Lotus corniculatus (Fabaceae) complex: a synthesis of diverse evidence. Can J Bot 74:975–989 Gutierrez JF, Vaquero F, Vences FJ (1994) Allopolyploid vs. autopolyploid origins in the genus Lathyrus (Leguminosae). Heredity 73:29–40 Hanson L (1995) Some new chromosome counts in the genus Inga (Leguminosae: Mimosoideae). Kew Bull 50:801–804 Havananda T, Brummer EC, Maureira-Butler IJ, Doyle JJ (2010) Relationships among diploid members of the Medicago sativa (Fabaceae) species complex based on chloroplast and mitochondrial DNA sequences. Syst Bot 35:140–150 Havananda T, Brummer EC, Doyle JJ (2011) Complex patterns of autopolyploid evolution in alfalfa and allies (Medicago sativa: Leguminosae). Am J Bot 98:1633-1646 Hejazi H, Mohsen S, Nasab MZ (2010) Cytotaxonomy of some Onobrychis (Fabaceae) species and populations in Iran. Caryologia 63:18–31 Hughes CE, Eastwood R (2006) Island radiation on a continental scale: exceptional rates of plant diversification after uplift of the Andes. Proc Natl Acad Sci U S A 103:10334–10339 Hughes CE, Govindarajulu R, Robertson A, Filer DL, Harris SA, Bailey CD (2007) Serendipitous backyard hybridization and the origin of crops. Proc Natl Acad Sci U S A 104:14389–14394 Hughes CE, Bailey CD, Harris SA (2002) Divergent and reticulate species relationships in Leucaena (Fabaceae) inferred from multiple data sources: insights into polyploid origins and nrDNA polymorphism. Am J Bot 89:1057–1073 Hughes CE, Bailey CD, Krosnick S, Luckow MA (2003) Relationships among genera of the informal Dichrostachys and Leucaena groups (Mimosoideae) inferred from nuclear ribosomal ITS sequences. In: Klitgaard B, Bruneau A (eds) Advances in legume systematics, Part 1.0. Royal Botanic Gardens, Kew, pp 221–238 Hulina N (2010) ‘‘Planta hortifuga’’ in flora of the continental part of Croatia. Agriculturae Conspectus Scientificus 75:57–65 Ilut DC, Coate JE, Luciano AK, Owens TG, May GD, Farmer A, Doyle JJ (2012) A comparative transcriptomic study of an allotetraploid and its diploid progenitors illustrates the unique advantages and challenges of RNA-Seq in plant species. Am J Bot 99:383–396 CO RR 1055 1056 1057 1058 1059 1060 1061 1062 1063 1064 1065 1066 1067 1068 1069 1070 1071 1072 1073 1074 1075 1076 1077 1078 1079 1080 1081 1082 1083 1084 1085 1086 1087 1088 1089 1090 1091 1092 1093 1094 1095 1096 1097 1098 1099 1100 1101 1102 1103 1104 1105 1106 1107 J. J. Doyle UN Editor Proof 176 Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 177/179 177 EC TE D PR OO F Innes RW, Ameline-Torregrosa C, Ashfield T, Cannon E, Cannon SB, Chacko B, Chen NWG, Couloux A, Dalwani A, Denny R, Deshpande S, Egan AN, Glover N, Hans CS, Howell S, Ilut D, Jackson S, Lai H, Mammadov J, del Campo SM, Metcalf M, Nguyen A, O’Bleness M, Pfeil BE, Podicheti R, Ratnaparkhe MB, Samain S, Sanders I, Segurens B, Sevignac M, Sherman-Broyles S, Thareau V, Tucker DM, Walling J, Wawrzynski A, Yi J, Doyle JJ, Geffroy V, Roe BA, Maroof MAS, Young ND (2008) Differential accumulation of retroelements and diversification of NB-LRR disease resistance genes in duplicated regions following polyploidy in the ancestor of soybean. Plant Physiol (Rockville) 148:1740–1759 Jaillon O, Aury J, Noel B, Policriti A, Clepet C, Casagrande A, Choisne N, Aubourg S, Vitulo N, Jubin C, Vezzi A, Legeai F, Hugueney P, Dasilva C, Horner D, Mica E, Jublot D, Poulain J, Bruyere C, Billault A, Segurens B, Gouyvenoux M, Ugarte E, Cattonaro F, Anthouard V, Vico V, Del Fabbro C, Alaux M, Di Gaspero G, Dumas V, Felice N, Paillard S, Juman I, Moroldo M, Scalabrin S, Canaguier A, Le Clainche I, Malacrida G, Durand E, Pesole G, Laucou V, Chatelet P, Merdinoglu D, Delledonne M, Pezzotti M, Lecharny A, Scarpelli C, Artiguenave F, Pe ME, Valle G, Morgante M, Caboche M, Adam-Blondon A, Weissenbach J, Quetier F, Wincker P, French-Italian Public (2007) The grapevine genome sequence suggests ancestral hexaploidization in major angiosperm phyla. Nature (London) 449:463 Jenczewski E, Prosperi J, Ronfort J (1999) Evidence for gene flow between wild and cultivated Medicago sativa (Leguminosae) based on allozyme markers and quantitative traits. Am J Bot 86:677–687 Jiao Y, Wickett NJ, Ayyampalayam S, Chanderbali AS, Landherr L, Ralph PE, Tomsho LP, Hu Y, Liang H, Soltis PS, Soltis DE, Clifton SW, Schlarbaum SE, Schuster SC, Ma H, LeebensMack J, dePamphilis CW (2011) Ancestral polyploidy in seed plants and angiosperms. Nature 473:97–100 Joly S, Bruneau A (2004) Evolution of triploidy in Apios americana (Leguminosae) revealed by genealogical analysis of the histone H3-D gene. Evolution 58:284–295 Jorgensen JL, Stehlik I, Brochmann C, Conti E (2003) Implications of ITS sequences and RAPD markers for the taxonomy and biogeography of the Oxytropis campestris and O. arctica (Fabaceae) complexes in Alaska. Am J Bot 90:1470–1480 Kajita T, Ohashi H, Tateishi Y, Bailey CD, Doyle JJ (2001) rbcL and legume phylogeny, with particular reference to phaseoleae, millettieae, and allies. Syst Bot 26:515–536 Kloda JM, Dean PDG, Maddren C, MacDonald DW, Mayes S (2008) Using principle component analysis to compare genetic diversity across polyploidy levels within plant complexes: an example from British Restharrows (Ononis spinosa and Ononis repens). Heredity 100:253–260 Kumar PS, Hymowitz T (1989) Where are the diploid 2n equals 2x equals 20 genome donors of glycine willd. Leguminosae Papilionoideae. Euphytica 40:221–226 Kumari S, Bir SS (1990) Karyomorphological evolution in Papilionaceae. J Cytol Genet 25:173–219 Kupicha FK (1981) Vicieae. In: Polhill RM, Raven PH (eds) Advances in legume systematics, Part 1. Royal Botanic Gardens, Kew, pp 377–381 Lackey JA (1981) Phaseoleae. In: Polhill RM, Raven PH (eds) Advances in legume systematics, Part 1. Royal Botanic Gardens, Kew, pp 301–328 Lavia GI, Ortiz MA, Robledo G, Fernandez A, Seijo G (2011) Origin of triploid Arachis pintoi (Leguminosae) by autopolyploidy evidenced by FISH and meiotic behaviour. Ann Bot (London) 108:103–111 Lavin M, Herendeen PS, Wojciechowski MF (2005) Evolutionary rates analysis of Leguminosae implicates a rapid diversification of lineages during the tertiary. Syst Biol 54:575–594 Lavin M, Pennington RT, Klitgaard BB, Sprent JI, de Lima HC, Gasson PE (2001) The dalbergioid legumes (Fabaceae): delimitation of a pantropical monophyletic clade. Am J Bot 88:503–533 Leitch IJ, Bennett MD (2004) Genome downsizing in polyploid plants. Biol J Linn Soc 82:651–663 CO RR 1108 1109 1110 1111 1112 1113 1114 1115 1116 1117 1118 1119 1120 1121 1122 1123 1124 1125 1126 1127 1128 1129 1130 1131 1132 1133 1134 1135 1136 1137 1138 1139 1140 1141 1142 1143 1144 1145 1146 1147 1148 1149 1150 1151 1152 1153 1154 1155 1156 1157 1158 1159 1160 UN Editor Proof 9 Polyploidy in Legumes Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 178/179 EC TE D PR OO F Lewis GP, Schrire B, MacKinder B, Lock M (2005) Legumes of the world. Royal Botanic Gardens, Kew Luckow M, Fortunato RH, Sede S, Livshultz T (2005) The phylogenetic affinities of two mysterious monotypic mimosoids from southern South America. Syst Bot 30:585–602 Luckow M, Miller JT, Murphy DJ, Livshultz T (2003) A phylogenetic analysis of the Mimosoideae (Leguminosae) based on chloroplast DNA sequence data. In: Klitgaard B, Bruneau A (eds) Advances in legume systematics, Part 1.0. Royal Botanic Gardens, Kew, pp 197–220 Lynch M, Conery JS (2003) The origins of genome complexity. Science 302:1401–1404 Mayrose I, Zhan SH, Rothfels CJ, Magnuson-Ford K, Barker MS, Rieseberg LH, Otto SP (2011) Recently formed polyploid plants diversify at lower rates. Science (Washington DC) 333:1257 McMahon MM (2005) Phylogenetic relationships and floral evolution in the papilionoid legume clade Amorpheae. Brittonia 57:397–411 Morales M, Wulff AF, Fortunato RH, Poggio L (2010) Chromosome and morphological studies in the Mimosa debilis complex (Mimosoideae, Leguminosae) from southern South America. Aust J Bot 58:12–22 Murphy DJ, Brown GK, Miller JT, Ladiges PY (2010) Molecular phylogeny of Acacia Mill. (Mimosoideae: Leguminosae): evidence for major clades and informal classification. Taxon 59:7–19 Neumann P, Koblizkova A, Navratilova A, Macas J (2006) Significant expansion of Vicia pannonica genome size mediated by amplification of a single type of giant retroelement. Genetics 173:1047–1056 Ohashi H, Polhill RM, Schubert BG (1981) Desmodieae. In: Polhill RM, Raven PH (eds) Advances in legume systematics, Part 1. Royal Botanic Gardens, Kew, pp 292–300 Orthia LA, Cook LG, Crisp MD (2005) Generic delimitation and phylogenetic uncertainty: an example from a group that has undergone an explosive radiation. Aust Syst Bot 18:41–47 Ortiz MA, Guillermo Seijo J, Fernandez A, Lavia GI (2011) Meiotic behavior and pollen viability of tetraploid Arachis glabrata and A. nitida species (section Rhizomatosae, Leguminosae): implications concerning their polyploid nature and seed set production. Plant Syst Evol 292:73–83 Ossowski S, Schneeberger K, Lucas-Lledo JI, Warthmann N, Clark RM, Shaw RG, Weigel D, Lynch M (2010) The rate and molecular spectrum of spontaneous mutations in Arabidopsis thaliana RID E-2139-2011 RID C-1418-2008. Science 327:92–94 Pandit MK, Tan HTW, Bisht MS (2006) Polyploidy in invasive plant species of Singapore. Bot J Linn Soc 151:395–403 Pandit MK, Pocock MJO, Kunin WE (2011) Ploidy influences rarity and invasiveness in plants. J Ecol 99: 1108-1115 Pardo C, Cubas P, Tahiri H (2004) Molecular phylogeny and systematics of Genista (Leguminosae) and related genera based on nucleotide sequences of nrDNA (ITS region) and cpDNA (trnL-trnF intergenic spacer). Plant Syst Evol 244:93–119 Pennington, RT, Klitgaard BB, Ireland H, Lavin M (2000) New insights into floral evolution of basal Papilionoideae from molecular phylogenies. In: Herendeen PS Bruneau A (eds.) Advances in legume systematics Part 9. Royal Botanic Gardens, Kew p 233–248 Pennington, RT, Lavin M, Ireland H, Klitgaard BB, Preston J, Hu J-M (2001) Phylogenetic relationships of basal papilionoid legumes based upon sequences of the chloroplast trnL intron. Syst Bot 26:537–556 Pfeil BE, Schlueter JA, Shoemaker RC, Doyle JJ (2005) Placing paleopolyploidy in relation to taxon divergence: a phylogenetic analysis in legumes using 39 gene families. Syst Biol 54:441–454 Polhill RM (1981a) Dipteryxeae. In: Polhill RM, Raven PH (eds) Advances in legume systematics, Part 1. Royal Botanic Gardens, Kew, pp 231–232 Polhill RM (1981b) Indigofereae. In: Polhill RM, Raven PH (eds) Advances in legume systematics, Part 1. Royal Botanic Gardens, Kew, pp 289–291 CO RR 1161 1162 1163 1164 1165 1166 1167 1168 1169 1170 1171 1172 1173 1174 1175 1176 1177 1178 1179 1180 1181 1182 1183 1184 1185 1186 1187 1188 1189 1190 1191 1192 1193 1194 1195 1196 1197 1198 1199 1200 1201 1202 1203 1204 1205 1206 1207 1208 1209 1210 1211 1212 1213 1214 J. J. Doyle UN Editor Proof 178 Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 179/179 179 EC TE D PR OO F Polhill RM, Raven PH (1981) Advances in legume systematics, Part 1. Royal Botanic Gardens, Kew Polhill RM, Sousa M (1981) Robinieae. In: Polhill RM, Raven PH (eds) Advances in legume systematics, Part 1. Royal Botanic Gardens, Kew, pp 283–288 Reddy VRK, Revathi R (1993) Chemotaxonomic studies in the genus Indigofera Linn. J Econ Taxon Bot 17:115–120 Rosato M, Castro M, Rossello JA (2008) Relationships of the woody Medicago species (section Dendrotelis) assessed by molecular cytogenetic analyses. Ann Bot (London) 102:15–22 Rossello JA, Castro M (2008) Karyological evolution of the angiosperm endemic flora of the Balearic Islands. Taxon 57:259–273 Sakiroglu M, Doyle JJ, Brummer EC (2010) Inferring population structure and genetic diversity of broad range of wild diploid alfalfa (Medicago sativa L.) accessions using SSR markers. Theor Appl Genet 121:403–415 Sato S, Nakamura Y, Kaneko T, Asamizu E, Kato T, Nakao M, Sasamoto S, Watanabe A, Ono A, Kawashima K, Fujishiro T, Katoh M, Kohara M, Kishida Y, Minami C, Nakayama S, Nakazaki N, Shimizu Y, Shinpo S, Takahashi C, Wada T, Yamada M, Ohmido N, Hayashi M, Fukui K, Baba T, Nakamichi T, Mori H, Tabata S (2008) Genome structure of the legume, Lotus japonicus. DNA Res 15:227–239 Schleueter J, Dixon P, Granger C, Grant D, Clark L, Doyle JJ, Shoemaker RC (2004) Mining EST databases to resolve evolutionary events in major crop species. Genome 47:868–876 Schmutz J, Cannon SB, Schlueter J, Ma J, Mitros T, Nelson W, Hyten DL, Song Q, Thelen JJ, Cheng J, Xu D, Hellsten U, May GD, Yu Y, Sakurai T, Umezawa T, Bhattacharyya MK, Sandhu D, Valliyodan B, Lindquist E, Peto M, Grant D, Shu S, Goodstein D, Barry K, FutrellGriggs M, Du J, Tian Z, Zhu L, Gill N, Joshi T, Libault M, Sethuraman A, Zhang XC, Shinozaki K, Nguyen HT, Wing RA, Cregan P, Specht J, Grimwood J, Rokhsar D, Stacey G, Shoemaker RC, Jackson SA (2010) Genome sequence of the paleopolyploid soybean. Nature 463:178–183 Schrire BD, Lavin M, Barker NP, Forest F (2009) Phylogeny of the Tribe Indigofereae (Leguminosae-Papilionoideae): geographically structured more in succulent-rich and temperate settings than in grass-rich environments. Am J Bot 96:816–843, 844–852 Seijo G, Lavia GI, Fernandez A, Krapovickas A, Ducasse DA, Bertioli DJ, Moscone EA (2007) Genomic relationships between the cultivated peanut (Arachis hypogaea, Leguminosae) and its close relatives revealed by double GISH. Am J Bot 94:1963–1971 Seijo G, Fernandez A (2001) Chromosome numbers of some southernmost species of Mimosa L. (Leguminosae). Cytologia 66:19–23 Sen O, Bhattacharya S (1988) Cytomixis in vigna-glabrescens Ttk-1 wild. Cytologia 53:437–440 Shoemaker RC, Polzin K, Labate J, Specht J, Brummer EC, Olson T, Young N, Concibido V, Wilcox J, Tamulonis JP, Kochert G, Boerma HR (1996) Genome duplication in soybean (Glycine subgenus soja). Genetics 144:329–338 Shoemaker RC, Schlueter J, Doyle JJ (2006) Paleopolyploidy and gene duplication in soybean and other legumes. Curr Opin Plant Biol 9:104–109 Simon MF, Grether R, de Queiroz LP, Saerkinen TE, Dutra VF, Hughes CE (2011) The evolutionary history of Mimosa (Leguminosae): toward a phylogeny of the sensitive plants. Am J Bot 98:1201–1221 Singer SR, Maki SL, Farmer AD, Ilut D, May GD, Cannon SB, Doyle JJ (2009) Venturing beyond beans and peas: what can we learn from chamaecrista? Plant Physiol 151:1041–1047 Sinou C, Forest F, Lewis GP, Bruneau A (2009) The genus Bauhinia s.l. (Leguminosae): a phylogeny based on the plastid trnL-trnF region. Botany-Botanique 87:947–960 Soltis DE, Buggs RJA, Doyle JJ, Soltis PS (2010) What we still don’t know about polyploidy. Taxon 59:1387–1403 Soltis DE, Albert VA, Leebens-Mack J, Bell CD, Paterson AH, Zheng C, Sankoff D, dePamphilis CW, Wall PK, Soltis PS (2009) Polyploidy and angiosperm diversification. Am J Bot 96:336–348 CO RR 1215 1216 1217 1218 1219 1220 1221 1222 1223 1224 1225 1226 1227 1228 1229 1230 1231 1232 1233 1234 1235 1236 1237 1238 1239 1240 1241 1242 1243 1244 1245 1246 1247 1248 1249 1250 1251 1252 1253 1254 1255 1256 1257 1258 1259 1260 1261 1262 1263 1264 1265 1266 1267 UN Editor Proof 9 Polyploidy in Legumes Layout: T1 Standard SC Chapter No.: 9 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 180/179 EC TE D PR OO F Spellenberg R (1981) Poly ploidy in Dalea-Formosa Fabaceae on the Chihuahuan desert. Brittonia 33:309–324 Sprent JI (2009) Legume nodulation: a global perspective. Wiley-Blackwell, Ames Srivastav PK, Raina SN (1986) Cytogenetics of Tephrosia Vi. Meiotic systems in some taxa. Cytologia 51:359–374 Steele KP, Ickert-Bond SM, Zarre S, Wojciechowski MF (2010) Phylogeny and character evolution in Medicago (Leguminosae): evidence from analyses of plastid Trnk/matk and nuclear Ga3ox1 sequences. Am J Bot 97:1142–1155 Stefanovic S, Pfeil BE, Palmer JD, Doyle JJ (2009) Relationships among phaseoloid legumes based on sequences from eight chloroplast regions. Syst Bot 34:115–128 Straub SCK, Doyle JJ (2009) Conservation genetics of Amorpha georgiana (Fabaceae), an endangered legume of the Southeastern United States. Mol Ecol 18:4349–4365 Straub SCK, Pfeil BE, Doyle JJ (2006) Testing the polyploid past of soybean using a low-copy nuclear gene––is Glycine (Fabaceae: Papilionoideae) an auto- or allopolyploid? Mol Phylogenet Evol 39:580–584 te Beest M, Le Roux JJ, Richardson DM, Brysting AK, Suda J, Kubešová M, Pyšek P (2011) The more the better? The role of polyploidy in facilitating plant invasions. Ann Bot 109:19–45 Thulin M, Lavin M (2001) Phylogeny and biogeography of the Ormocarpum group (Fabaceae): a new genus Zygocarpum from the horn of Africa region. Syst Bot 26:299–317 Tondini F, Tavoletti S, Mariani A, Veronesi F (1993) A statistical approach to estimate the frequency of n, 2n and 4n pollen grains in diploid alfalfa. Euphytica 69:109–114 Torres DC, Matos Santos Lima JP, Fernandes AG, Nunes EP, Grangeiro TB (2011) Phylogenetic relationships within chamaecrista sect. Xerocalyx (Leguminosae, Caesalpinioideae) inferred from the cpDNA trnE- trnT intergenic spacer and nrDNA ITS sequences. Genet Mol Biol 34:244–251 Travnicek P, Eliasova A, Suda J (2010) The distribution of cytotypes of Vicia cracca in Central Europe: the changes that have occurred over the last four decades. Preslia (Prague) 82:149–163 Turini FG, Braeuchler C, Heubl G (2010) Phylogenetic relationships and evolution of morphological characters in Ononis L. (Fabaceae). Taxon 59:1077–1090 van Wyk BE, Schutte AL (1988) Chromosome numbers in Lotononis and Buchenroedera (Fabaceae-Crotalarieae). Ann Missouri Bot Gard 75:1603–1607 Varshney RK, Chen W, Li Y, Bharti AK, Saxena RK, Schlueter JA et al (2011) Draft genome sequence of Pigeonpea (Cajanus cajan), an orphan legume crop of resource-poor farmers. Nat Biotech. doi:10.1038/nbt.2022 Veronesi F, Mariani A, Bingham ET (1986) Unreduced gametes in diploid medicago and their importance in alfalfa breeding. Theor Appl Genet 72:37–41 Wagstaff S, Heenan P, Sanderson M (1999) Classification, origins, and patterns of diversification in New Zealand Carmichaelinae (Fabaceae). Am J Bot 86:1346–1356 Wang H, Moore MJ, Soltis PS, Bell CD, Brockington SF, Alexandre R, Davis CC, Latvis M, Manchester SR, Soltis DE (2009) Rosid radiation and the rapid rise of angiosperm-dominated forests. Proc Nat Acad Sci 106:3853–3858 Wilbur RL (1975) A revision of the North American genus Amorpha Leguminosae Psoraleae. Rhodora 77:337–409 Wojciechowski MF, Lavin M, Sanderson MJ (2004) A phylogeny of legumes (Legumenosae) based on analyses of the plastid matK gene resolves many well-supported subclades within the family. Am J Bot 91:1846–1862 Wojciechowski M (2005) Astragalus (Fabaceae): a molecular phylogenetic perspective. Brittonia 57:382–396 Young N, Debellé F, Oldroyd G, Geurts R, Cannon SB et al (2011) The medicago genome provides insight into the evolution of rhizobial symbioses. Nature. doi:10.1038/nature10625 Zhang M, Fritsch PW, Cruz BC (2009) Phylogeny of Caragana (Fabaceae) based on DNA sequence data from rbcL, trnS-trnG, and ITS. Mol Phylogenet Evol 50:547–559 CO RR 1268 1269 1270 1271 1272 1273 1274 1275 1276 1277 1278 1279 1280 1281 1282 1283 1284 1285 1286 1287 1288 1289 1290 1291 1292 1293 1294 1295 1296 1297 1298 1299 1300 1301 1302 1303 1304 1305 1306 1307 1308 1309 1310 1311 1312 1313 1314 1315 1316 1317 1318 1319 1320 J. J. Doyle UN Editor Proof 180 Metadata of the chapter that will be visualized in SpringerLink Book Title Polyploidy and Genome Evolution Series Title Chapter Title Jeans, Genes, and Genomes: Cotton as a Model for Studying Polyploidy Copyright Year 2012 Copyright HolderName Springer-Verlag Berlin Heidelberg Corresponding Author Family Name Wendel Particle Given Name Jonathan F. Suffix Author Division Department of Ecology, Evolution, and Organismal Biology Organization Iowa State University Address 50011, Ames, IA, USA Email jfw@iastate.edu Family Name Flagel Particle Given Name Lex E. Suffix Author Division Department of Biology Organization Duke University Address 90338, 27708, Durham, NC, USA Email lex.flagel@duke.edu Family Name Adams Particle Given Name Keith L. Suffix Abstract Division Department of Botany, and UBC Botanical Garden and Centre for Plant Research Organization University of British Columbia Address 6268 University Blvd, V6T 1Z4, Vancouver, BC, Canada Email keitha@mail.ubc.ca We present an overview of the cotton genus (Gossypium) as a model for the study of polyploidy. A synopsis of the origin and evolution of polyploid cotton is provided, offering an organismal framework and phylogenetic perspective that is critical for understanding modes and mechanisms of gene and genome evolution. Sequence data from thousands of genes implicate a mid-Pleistocene (1–2 mya) origin of polyploid cotton, following trans-oceanic dispersal of an Old World, A-genome diploid to the New World and subsequent hybridization with an indigenous D-genome diploid. This chance biological reunion, occurring after 5–10 million years of diploid evolution in isolation, has led to an array of molecular genetic interactions in the newly formed allopolyploid lineage, including nonreciprocal homoeologous recombination and perhaps other forms of interlocus concerted evolution, differential rates of genomic evolution, intergenomic spread of transposable elements, and myriad forms of alterations in duplicate expression relative to that experienced in the ancestral diploids. The latter include developmental, organ-, tissue-, and cell-specific biases in homoeologous gene expression, which can be sensitive to various forms of environmental perturbation and stress. The allopolyploid Gossypium transcriptome is exceptionally dynamic, with homoeolog expression ratios being subject to change even during development of the single-celled cotton fiber. Expression evolution is temporally partitioned into changes accompanying genome merger (hybridization) at the diploid level, polyploidization, and longer term evolution at the allopolyploid level. Evidence indicates that allopolyploidy facilitated colonization of a new ecological niche for the genus and led to an enhanced capacity for developing agronomically superior cotton varieties. The myriad mechanisms that underlie genomic and regulatory evolution are suggested to have contributed to both ecological success and agronomic potential. Book ISBN: 978-3-642-31441-4 Page: 181/206 Chapter 10 4 Jonathan F. Wendel, Lex E. Flagel and Keith L. Adams 9 10 11 12 13 14 15 16 17 18 19 20 21 22 D 8 TE 7 Abstract We present an overview of the cotton genus (Gossypium) as a model for the study of polyploidy. A synopsis of the origin and evolution of polyploid cotton is provided, offering an organismal framework and phylogenetic perspective that is critical for understanding modes and mechanisms of gene and genome evolution. Sequence data from thousands of genes implicate a mid-Pleistocene (1–2 mya) origin of polyploid cotton, following trans-oceanic dispersal of an Old World, A-genome diploid to the New World and subsequent hybridization with an indigenous D-genome diploid. This chance biological reunion, occurring after 5–10 million years of diploid evolution in isolation, has led to an array of molecular genetic interactions in the newly formed allopolyploid lineage, including nonreciprocal homoeologous recombination and perhaps other forms of interlocus concerted evolution, differential rates of genomic evolution, intergenomic spread of transposable elements, and myriad forms of alterations in duplicate expression relative to that experienced in the ancestral diploids. The latter include developmental, organ-, tissue-, and cell-specific biases in homoeologous gene expression, which can be sensitive to various forms of environmental perturbation and stress. The allopolyploid Gossypium transcriptome is exceptionally dynamic, with homoeolog expression ratios being subject to change even during development of the EC 6 CO RR 5 PR OO 3 Jeans, Genes, and Genomes: Cotton as a Model for Studying Polyploidy 2 F 1 Book ID: 272454_1_En Date: 16-8-2012 J. F. Wendel (&) Department of Ecology, Evolution, and Organismal Biology, Iowa State University, Ames, IA 50011, USA e-mail: jfw@iastate.edu L. E. Flagel Department of Biology, Duke University, 90338 Durham, NC 27708, USA e-mail: lex.flagel@duke.edu UN Editor Proof Layout: T1 Standard SC Chapter No.: 10 K. L. Adams Department of Botany, and UBC Botanical Garden and Centre for Plant Research, University of British Columbia, 6268 University Blvd, Vancouver, BC V6T 1Z4, Canada e-mail: keitha@mail.ubc.ca P. S. Soltis and D. E. Soltis (eds.), Polyploidy and Genome Evolution, DOI: 10.1007/978-3-642-31442-1_10,  Springer-Verlag Berlin Heidelberg 2012 181 Layout: T1 Standard SC Chapter No.: 10 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 182/206 25 26 27 28 29 30 single-celled cotton fiber. Expression evolution is temporally partitioned into changes accompanying genome merger (hybridization) at the diploid level, polyploidization, and longer term evolution at the allopolyploid level. Evidence indicates that allopolyploidy facilitated colonization of a new ecological niche for the genus and led to an enhanced capacity for developing agronomically superior cotton varieties. The myriad mechanisms that underlie genomic and regulatory evolution are suggested to have contributed to both ecological success and agronomic potential. F 24 PR OO 23 J. F. Wendel et al. 31 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 D 36 TE 35 Because of its economic importance, the cotton genus (Gossypium L.) has long attracted the attention of agricultural scientists, taxonomists, and biologists in multiple disciplines. Accordingly, a great deal is understood about the origin and diversification of the genus, its basic plant biology, and its properties as a crop plant (Paterson 2009; Stewart et al. 2010; Wendel et al. 2009). One of the most salient features of the genus is that its history is so encompassing in scope, involving a global (mostly austral) phylogenetic diversification and a repeated history of trans-oceanic dispersal. Superimposed on this natural diversification has been a many-thousand-year history of human manipulation, tracing to ancient human cultures on several continents who independently domesticated four species, two from the Americas, G. hirsutum and G. barbadense, and two from Africa-Asia, G. arboreum and G. herbaceum. In each of these four cases, aboriginal peoples discovered that the unique properties of cotton ‘‘fibers’’, which are unicellular epidermal seed trichomes, made them useful for ropes, textiles, and other applications. Each of these crop species has its own history of domestication, diversification, and current utilization (Brubaker et al. 1999a; Hutchinson 1951, 1954; Hutchinson et al. 1947; Percy and Wendel 1990; Wendel et al. 2009). This rich history of scientific study has made Gossypium one of the best systems for studies of polyploidy in plants. As testified by the many papers in this volume, plant evolution has been characterized by repeated rounds of whole-genome doubling. Evolutionary footprints of paleopolyploidy in ‘‘diploid’’ (n = 13) cotton have long been evidenced using both classic and more modern techniques, including chromosome banding, comparative genome mapping, and analysis of synonymous substitution rate curves (Muravenko et al. 1998; Brubaker et al. 1999b; Reinisch et al. 1994; Paterson 2009; Lin and Paterson 2009; Lin et al. 2011). These more ancient events, which must have profoundly impacted morphological, ecological, and physiological diversification, predate a more recent allopolyploidy event that traces to the mid-Pleistocene (Wendel 1989; Wendel et al. 2009; Wendel and Cronn 2003). The most experimentally tractable polyploid event is this most recent one, resulting from the merger that led to the evolution of the five extant species of New World allopolyploid cotton (2n = 4x = 52), including the two EC 33 34 10.1 Introduction CO RR 32 UN Editor Proof 182 Layout: T1 Standard SC Chapter No.: 10 Book ISBN: 978-3-642-31441-4 Page: 183/206 Jeans, Genes, and Genomes: Cotton as a Model for Studying Polyploidy 183 88 10.1.1 Origin and Diversification of the Diploid Cottons 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 PR OO 71 D 70 TE 68 69 EC 67 The cotton genus belongs to a small tribe, the Gossypieae, that includes only eight genera and *120 species (Fryxell 1968, 1979). Four of these genera are either monotypic or contain only several species with restricted geographic distributions, Lebronnecia (Marquesas Islands), Cephalohibiscus (New Guinea, Solomon Islands), Gossypioides (east Africa, Madagascar), and Kokia (Hawaii). The tribe also includes four moderately sized genera with broader ranges: Hampea, with 21 neotropical species; Cienfuegosia, with 25 species from the neotropics and parts of Africa; Thespesia, with 17 tropical species; and last but not least, Gossypium, the largest and most widely distributed genus in the tribe with more than 50 species (Fryxell 1992). Gossypium species collectively have achieved a nearly worldwide distribution, with several primary centers of diversity in the arid or seasonally arid tropics and subtropics. Species-rich regions include Australia, especially the Kimberley region in NW Australia, the Horn of Africa and southern Arabian Peninsula, and the western part of central and southern Mexico. Recognition of these groups of species reflects decades of accumulated understanding that emerged from basic plant exploration and taxonomic analysis (Fryxell 1979, 1992; Hutchinson et al. CO RR 66 F 87 cultivated species G. hirsutum (Upland cotton) and G. barbadense (Pima cotton, Sea Island cotton). Gossypium hirsutum presently is responsible for over 90 % of the cotton crop internationally, having spread from its original home in Mesoamerica to over 50 countries in both eastern and western hemispheres. Accumulating evidence indicates that polyploidy per se has played a critical role in enabling the development of modern, agronomically elite varieties. Thus, it is of interest to explore what cotton has taught us about the evolutionary consequences of genome doubling, and in turn, what the process of genome doubling might reveal about phenotypic diversification and the goals of agronomic improvement. We begin by reviewing our current understanding of the diversity of the genus and the origin of the Gossypium clade, which includes one of the classic examples of polyploidy. This phylogenetic and temporal perspective provides the organismal framework that serves as the foundation for all analyses of the consequences of polyploid evolution, including those focused on genomic, epigenomic, and phenotypic levels. We then provide a synopsis of the myriad genomic consequences that were set in motion by the evolutionary processes of genome merger, chromosome doubling, and subsequent evolutionary diversification which collectively gave rise to modern allopolyploid cottons, drawing attention to the relationships among evolutionary processes and temporal scale of divergence. Thus, we distinguish phenomena and processes that might characterize the earliest stages of polyploid formation from those that are responsible for longer term genomic and phenotypic changes. Finally, we summarize evidence that polyploidy enables the evolution of transgressive or novel phenotypes in cotton, as exemplified by the differences between modern cultivated diploid versus allopolyploid cotton. 64 65 UN Editor Proof 10 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 10 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 184/206 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149 F 111 112 PR OO 109 110 D 108 TE 107 1947; Saunders 1961; Watt 1907; Wendel et al. 2009). The genus is extraordinarily diverse; species morphologies range from fire-adapted, herbaceous perennials in NW Australia to trees in SW Mexico that escape the dry season by dropping their leaves. Corolla colors span a rainbow of blue to purple (G. triphyllum), mauves and pinks (‘‘Sturt’s Desert Rose’’, G. sturtianum, is the official floral emblem of the Northern Territory, Australia), whites and pale yellows (NW Australia, Mexico, Africa-Arabia), and even a deep sulfur-yellow (G. tomentosum from Hawaii). Seed coverings range from nearly glabrous (e.g., G. klotzschianum and G. davidsonii), to short stiff, dense, brown hairs that aid in wind-dispersal (G. australe, G. nelsonii), to the long, fine white fibers that characterize highly improved forms of the four cultivated species (Fig. 10.1). There are even seeds that produce fat bodies to facilitate ant-dispersal (Seelanan et al. 1999). At the other end of the ant coevolution spectrum is G. tomentosum from the Hawaiian Islands, which lost the foliar and extra-floral nectaries that are common in other Gossypium species, presumably in response to the absence of native ants. The evolution of this morphological and geographic diversity was accompanied by a parallel diversification at the chromosomal level (Lin and Paterson 2009). Although all diploid species share the same chromosome number (n = 13), there is more than 3-fold variation in DNA content per genome (Hendrix and Stewart 2005). Chromosome morphology is similar among closely related species, as reflected in the ability of related species to form hybrids that display normal meiotic pairing and sometimes high F1 fertility. In contrast, crosses among more distant relatives may be difficult to achieve, and those that are successful are characterized by meiotic abnormalities. The collective observations of pairing behavior, chromosome size, and relative fertility in interspecific hybrids led to the designation of single-letter genome symbols (Beasley 1941) for related clusters of species. Presently, eight diploid genome groups (A through G, plus K) are recognized (Endrizzi et al. 1985; Stewart 1995). A genealogical framework for the genus is provided from analyses of multiple molecular phylogenetic investigations (reviewed in Wendel and Cronn 2003). A key phylogenetic conclusion has been the demonstration that the group of species recognized as belonging to Gossypium do, in fact, constitute a single natural lineage, despite their exceptionally broad geographic distribution and extraordinary morphological and cytogenetic diversity. A second important result is that the closest relative of Gossypium is the sister clade that includes the AfricanMadagascan genus Gossypioides and the Hawaiian endemic genus Kokia; these latter genera may thus be used as phylogenetic outgroups for studying evolutionary patterns and processes within Gossypium. A third phylogenetic conclusion is that each of the classically recognized genome groups comprises a monophyletic group. This information is summarized in a depiction of our present understanding of relationships (Fig. 10.2), which shows four major lineages of diploid species corresponding to three continents: Australia (C-, G-, K-genomes), the Americas (D-genome), and Africa/Arabia (two lineages: one comprising the A-, B-, and F-genomes, and a second containing the E-genome species). Embedded in this result is the observation that the earliest divergence in the genus separated the New EC 106 CO RR 105 J. F. Wendel et al. UN Editor Proof 184 Layout: T1 Standard SC Chapter No.: 10 Book ISBN: 978-3-642-31441-4 Page: 185/206 Jeans, Genes, and Genomes: Cotton as a Model for Studying Polyploidy 185 TE D PR OO F Editor Proof 10 Book ID: 272454_1_En Date: 16-8-2012 151 152 153 154 155 156 157 158 159 World D-genome lineage from the ancestor of all Old World taxa, and thus, that New World and Old World diploids are phylogenetic sister groups. Following this split in the genus, cottons comprising the Old World lineage divided into three groups, namely, the Australian cottons (C-, G-, and K-genome species), the African-Arabian E-genome species, and the African A-, B-, and F-genome cottons. The African F-genome clade, which consists of the sole species G. longicalyx, is diagnosed as sister to the A-genome species, an important realization in that this relationship identifies the wild forms most closely related to the clade (the A-genome) that ‘‘invented’’ long, or spinable, fiber since these two clades (A- and F-) shared a common ancestor. UN 150 CO RR EC Fig. 10.1 Representative seed and trichome diversity in Gossypium. Seed and trichome size and morphology are exceedingly variable in the genus. Most wild species have relatively small seeds (\5 mm in any dimension) with equally short fibers. Long (spinnable) fiber evolved only once, in the ancestor of modern A-genome cottons, which subsequently donated this capacity to modern tetraploid species, including the commercially important G. hirsutum and G. barbadense, at the time of allopolyploid formation. Key to species: Cult. AD1 = G. hirsutum TM1; Wild AD1 = G. hirsutum Tx2094 from the Yucatan Peninsula; AD3 = G. tomentosum WT936 from Hawaii; C1 = G. sturtianum C1-4 from Australia; Cult. A2 = G. arboreum AKA8401; Wild A1 = G. herbaceum subsp. africanum from Botswana; D5 = G. raimondii from Peru; D3 = G. davidsonii D3d-32 from Baja California; F1 = G. longicalyx F1-3 from Tanzania; B1 = G. anomalum B1-1 from Africa Layout: T1 Standard SC Chapter No.: 10 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 186/206 J. F. Wendel et al. TE D PR OO F Editor Proof 186 161 162 163 164 165 166 167 168 169 170 A temporal framework for the origin of Gossypium and its diversification is provided by sequence divergence data, which may serve as a proxy for time (Senchina et al. 2003). These analyses indicate that Gossypium diverged from its closest relatives during the Miocene, perhaps 10–15 mya, subsequently spreading around the world via trans-oceanic dispersal to acquire its modern geographic range. These early estimates of divergence, and hence times, have recently been supported by an enormous amount of new sequence data derived from a global assembly of ESTs (Flagel et al. 2012). Consideration of the phylogeny of Fig. 10.2 in a temporal context and in light of plate tectonic history leads to the inference that the history of Gossypium has entailed multiple episodes of trans-oceanic dispersal. These include at least one dispersal between Australia and Africa, UN 160 CO RR EC Fig. 10.2 Evolutionary history of Gossypium, as inferred from multiple molecular phylogenetic data sets. The closest relative of Gossypium is a lineage containing the African-Madagascan genus Gossypioides and the Hawaiian endemic genus Kokia. Following its likely origin 5–10 mya, Gossypium split into three major diploid lineages: the New World clade (D-genome); the African-Asian clade (A-, B-, E-, and F-genomes); and the Australian clade (C-, G-, and K-genomes). This global radiation involved several trans-oceanic dispersal events and was accompanied by morphological, ecological, and chromosomal differentiation (2C genome sizes shown in white ellipses). Allopolyploid cottons formed following trans-oceanic dispersal of an A-genome diploid to the Americas, where the immigrant underwent hybridization, as female, with a native D-genome diploid similar to modern G. raimondii. Polyploid cotton probably originated during the Pleistocene (1–2 mya), with the five modern species comprising the descendants of an early and rapid colonization of the New World tropics and subtropics Layout: T1 Standard SC Chapter No.: 10 Book ISBN: 978-3-642-31441-4 Page: 187/206 Jeans, Genes, and Genomes: Cotton as a Model for Studying Polyploidy 187 174 10.1.2 Origin and Diversification of the Polyploid Cottons 179 180 181 182 183 184 185 186 187 188 189 190 191 192 193 194 195 196 197 198 199 200 201 202 203 204 205 206 207 208 209 210 211 PR OO 178 D 177 A rich body of cytogenetic and experimental evidence has demonstrated that the tetraploid species, which are entirely New World in their distribution, are allopolyploids containing two genomes, an A-genome from Africa or Asia, and a D-genome similar to those found in the American diploids (Endrizzi et al. 1985; Wendel et al. 2009; Wendel and Cronn 2003). The hemisphere-scale allopatry of these two diploid genome groups led to more than 50 years of mystery surrounding the timing of formation and parentage of the New World allopolyploids. With respect to the question of ‘‘when’’, gene sequence data convincingly demonstrate that allopolyploid Gossypium originated prior to the evolution of modern humans but relatively recently in geological terms, perhaps 1–2 mya, or in the midPleistocene (Senchina et al. 2003; Wendel 1989; Flagel et al. 2012). With respect to the second part of the question, that of polyploid parentage, it is now clear that both extant A-genome species (G. arboreum, G. herbaceum) are equally divergent from the A-genome of allopolyploid cottons and that the closest living relative of the progenitor D-genome donor is G. raimondii (Endrizzi et al. 1985; Wendel et al. 2009; Wendel and Cronn 2003). Studies using nuclear (biparentally inherited) genes led to the same conclusion. Additionally, all allopolyploids contain an A-genome cytoplasm, as evidenced from analysis of both mitochondrial and plastid genomes (Galau and Wilkins 1989; Small and Wendel 1999; Wendel 1989). Finally, the studies mentioned above, and additional, extensive DNA sequence data from ongoing studies (Grover et al. unpublished), support a single origin for allopolyploid cotton. Given a Pleistocene origin for allopolyploid cotton species, one may infer that their morphological diversification and spread must have been relatively rapid following polyploidization. At present, five allopolyploid species are widely recognized, although a sixth species (G. ekmanianum) was recently proposed (Krapovickas and Seijo 2008). Gossypium darwinii is native to the Galapagos Islands, where it may form large populations in some areas (Percy and Wendel 1990). Gossypium tomentosum, from the Hawaiian Islands, has a more diffuse population structure, occurring mostly as scattered individuals and small populations on several islands (DeJoode and Wendel 1992). A third allopolyploid, G. mustelinum, is restricted to a small region of northeast Brazil (Wendel et al. 1994). In addition to these three truly wild species, there are two cultivated species (G. barbadense and G. hirsutum), each of which has a large indigenous range, collectively encompassing a wealth of morphological forms that span the wildto-domesticated continuum (Brubaker and Wendel 1993, 1994, 2001; Fryxell 1979; Hutchinson 1951; Percy and Wendel 1990). Gossypium hirsutum is widely TE 176 EC 175 CO RR 172 F 173 another to the Americas leading to the evolution of the D-genome diploids, and a second, much later colonization of the New World by the A-genome ancestor of the AD-genome allopolyploids (see below). 171 UN Editor Proof 10 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 10 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 188/206 J. F. Wendel et al. 232 10.2 Evolution Following Genome Duplication 233 10.2.1 Chromosomal Stabilization and Structural Stasis 220 221 222 223 224 225 226 227 228 229 230 234 235 236 237 238 239 240 241 242 243 244 245 246 247 248 249 250 251 PR OO 219 D 218 TE 216 217 EC 214 215 Classical cytogenetic evidence indicates that chromosomes of the A- and D-genomes of allopolyploid Gossypium are less able to pair with one another than are chromosomes of the living descendants of their diploid progenitors (reviewed in Endrizzi et al. 1985). For example, allopolyploid-derived haploids form an average of less than one bivalent per cell at meiotic metaphase, whereas chromosomes in hybrids between extant A- and D- genome diploids average 5.8 and 7.8 bivalents (reviewed in Endrizzi et al. 1985). These and similar observations indicate that natural selection has favored the evolution of mechanisms that promote exclusive bivalent formation in the allopolyploid. Neither the pace at which such mechanisms operate nor their nature are understood, but it seems rational to postulate that selection would be most intense in the first generations following allopolyploid formation, where the fitness cost of unbalanced gametes would be the greatest. One hypothesis for this apparent lack of bivalent formation is that genome stabilization following polyploidization involved genomic reorganization of the two resident genomes such that they no longer are capable of homoeologous pairing. To evaluate the extent of structural change, genetic maps were generated and compared among interspecific F2 progenies for diploid (both A- and D-) and allopolyploid cottons (Brubaker et al. 1999b; Rong et al. 2004). Comparisons of CO RR 213 F 231 distributed in Central and northern South America, the Caribbean, and even reaches distant islands in the Pacific (eg., Solomon Islands, Marquesas). Gossypium hirsutum is thought to have a more northerly distribution than G. barbadense, with wild populations occurring as far north as Tampa Bay, Florida (27̊38’N) (Stewart, personal observation). Gossypium barbadense has a more southerly indigenous range, centered in the northern third of South America but with a large region of range overlap with G. hirsutum in the Caribbean. Consideration of the distribution of the allopolyploid species suggests that polyploidy led to the invasion of a new ecological niche. Fryxell (1965, 1979) noted that in contrast to the majority of diploid species, allopolyploid species typically occur in coastal habitats, at least those forms that arguably are truly wild. Two species, both island endemics (G. darwinii and G. tomentosum), are restricted to near coastlines, and for two others (G. barbadense and G. hirsutum), wild forms occur in littoral habitats ringing the Gulf of Mexico, northwest South America, and distant Pacific Islands. Fryxell speculated that following initial formation, adaptation of the newly evolved allopolyploid to littoral habitats enabled it to exploit the fluctuating sea levels that characterized the Pleistocene. This ecological innovation is envisioned to have facilitated initial establishment of the new polyploid lineage and also may have provided a means for the rapid dispersal of the salt-water-tolerant seeds. 212 UN Editor Proof 188 Layout: T1 Standard SC Chapter No.: 10 258 259 260 261 262 263 264 265 266 267 268 269 270 271 272 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 F 257 PR OO 256 D 255 10.2.2 Genome Sizes, Transposable Element Mobilization, and Genomic Downsizing TE 254 189 gene order and synteny among the A- and D-genome maps, as well as those for both genomes of the allopolyploid (A vs. AT and D vs. DT), demonstrate that relatively few structural rearrangements have arisen in the 1–2 my since allopolyploid formation; conservation of collinear linkage groups is the rule rather than the exception. Thus, allopolyploidy in Gossypium has not been accompanied by extensive chromosomal rearrangement. This implies that structural rearrangement has not been a significant aspect of the process of polyploid genome stabilization in cotton. Additional support for this idea emerges from experiments involving synthetic allopolyploids; for example, Liu et al. (2001) used AFLP analysis to demonstrate almost exclusive fragment additivity for 22,000 genomic loci in nine sets of newly synthesized allotetraploid and allohexaploid Gossypium. Thus, and in contrast to some other plant models described in this volume (e.g., wheat, Tragopogon), the polyploid Gossypium genome appears to be relatively quiescent, at least with respect to the phenomenon of rapid genome change. A corollary is that the evolutionary enforcement of homologous pairing discussed above originated through means other than structural rearrangements. As shown in Fig. 10.2, genome sizes vary widely among diploid cotton species, from *900 Mb in the D-genome diploids to *2,600 Mb in the Australian diploids (Hendrix and Stewart 2005), reflecting primarily the differential and punctuated proliferation of various families of copia and gypsy transposable elements (TEs), as well as lineage-specific differences in the rate of deletions (Hawkins et al. 2008, 2006, 2009). The two progenitor genomes of allopolyploid cotton differ 2-fold in size, and moreover, they differ in their complement of resident TEs. Thus, allopolyploidization entailed the merger of two different complements of TEs, creating the potential for activation of TEs due to the generalized disruption of epigenetic suppression of TE activity following the merger of two diverged regulatory systems, a process commonly referred to as ‘‘genomic shock’’. To evaluate the possibility that polyploidization in Gossypium was accompanied by a transpositional burst, as in some other species (Kashkush et al. 2002; Shan et al. 2005; Ungerer et al. 2006), Hu et al. (2010) used phylogenetic and quantitative methods to identify changes in TE populations. These data showed that the major LTR retrotransposon classes in the AD genome phylogenetically clustered with either their A- or Dgenome antecedent elements in a genome-specific fashion, with no evidence of an impressive, recent, TE burst. Thus, hybridization and polyploidy do not appear to have stimulated a massive TE proliferation in Gossypium. Notwithstanding the relative TE quiescence indicated by these studies, evidence using FISH (Hanson et al. 1999, 1998) implicates at least a modest level of TE activity in allopolyploid cotton. These data show that a family of copia-like retrotransposable elements ‘‘horizontally’’ transferred across genomes following EC 253 CO RR 252 Book ISBN: 978-3-642-31441-4 Page: 189/206 Jeans, Genes, and Genomes: Cotton as a Model for Studying Polyploidy UN Editor Proof 10 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 10 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 190/206 J. F. Wendel et al. 324 10.2.3 Genic Evolution in Diploid Versus Allopolyploid Cotton 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 325 326 327 328 329 330 331 332 333 PR OO 298 D 297 TE 296 EC 295 CO RR 294 F 323 allopolyploid formation. This result highlights the phenomenon of TE spread across previously separated genomes following polyploid formation, raising the possibility that this process has played a role in diversification and adaptation via novel TE insertions. An attractive feature of the Gossypium model system is that species representing both progenitor genomes remain extant, and that they vary so dramatically (nearly 2-fold) in genome size yet retain collinearity with their orthologs in the allopolyploid genome. Although the Gossypium genome has yet to be sequenced, it is likely that a high-quality D-genome sequence (from G. raimondii) will be published by the time this volume is published, with perhaps the A-genome sequence not far behind. It is an exciting prospect to contemplate the availability of both diploid sequences as well as one or more from allopolyploid Gossypium. These data will offer a veritable gold mine for generating insight into the pace, patterns, and dynamics of genome evolution that accompany diploid divergence, allopolyploid formation, and subsequent evolution at the polyploid level. A glimpse of this promise is provided by current work focused on comparative sequencing of bacterial artificial chromosomes (BACs) from Gossypium in polyploid and diploid species. Two comparative BAC sequencing studies have been published, a region surrounding the CesA1 gene (Grover et al. 2004) from both homoeologous genomes of G. hirsutum, and a region surrounding the AdhA gene (Grover et al. 2007) from G. hirsutum and diploids representing models of the two progenitor genomes, i.e., G. arboreum and G. raimondii. Data generated to date indicate that small deletions are more prevalent in the polyploid genomes (AT and DT) than in either diploid genome, illustrative of the general phenomenon of genomic downsizing in polyploid genomes (Bennett and Leitch 2005), while providing a glimpse into an underlying mechanism (illegitimate recombination, which is biased in the allopolyploid toward deletion). Extensions of this work into 20 regions and in more genomes, including a phylogenetic outgroup are underway (Grover, Wendel, and Paterson, unpubl.), offering an opportunity to learn more about features of genome evolution that distinguish polyploid cotton from its diploid progenitors. 293 The most immediate and important genomic consequence of allopolyploid formation in Gossypium was simultaneous duplication of all nuclear genes. From a phylogenetic perspective, the various fates of gene duplication may partially be modeled as shown in Fig. 10.3. The null hypothesis for sequence evolution in allopolyploids derives from the organismal history; if both duplicated genes evolve independently following allopolyploid formation, then each homoeolog should be phylogenetically sister to its ortholog from its progenitor diploid, rather than to the other homoeolog. Similarly, if rates of sequence evolution are maintained between the diploid and allopolyploid level, branch lengths for the two A-genome sequences UN Editor Proof 190 Layout: T1 Standard SC Chapter No.: 10 Book ISBN: 978-3-642-31441-4 Page: 191/206 Recombination TE D Null hypothesis PR OO Gene loss Unequal rates 191 F Jeans, Genes, and Genomes: Cotton as a Model for Studying Polyploidy Editor Proof 10 Book ID: 272454_1_En Date: 16-8-2012 Novel expression Intergenomic transfer 334 335 336 337 UN CO RR EC Fig. 10.3 A model of various possibilities for duplicate gene evolution after allopolyploidy in Gossypium. The null expectation (center) derives from the organismal history (left): if homoeologs evolve independently following allopolyploid formation, then each should be phylogenetically sister to its ortholog from the donor diploid, rather than to the other homoeolog. Similarly, if rates of sequence evolution are similar at the diploid and allopolyploid level, branch lengths for the two A-genome sequences (one from the diploid, ‘‘A’’, and the other from the allopolyploid, ‘‘AT’’) should be similar, as they should for the two D-genome sequences (‘‘D’’ and ‘‘DT’’). The utility of this null hypothesis lies in its falsification; if homoeologous sequences interact via concerted evolutionary forces or nonreciprocal homoeologous recombination, for example, a different tree may be recovered (‘‘Recombination’’, top center), or if there is strong directional selection or pseudogenization, rate inequalities may become evident (‘‘Unequal rates’’, bottom center). Additional possibilities include loss of one of the homeologs (‘‘Gene loss’’, top right), replicative transfer of sequences from one genome to the other (‘‘Intergenomic transfer’’, bottom right), and evolutionary divergence in duplicate gene expression domains or amounts (‘‘Novel expression’’, right middle). This latter category, novel expression, encompasses multiple phenomena, including developmentally or environmentally regulated biases in homoeolog expression ratios, organ- or tissue- or cell-specific homoeolog silencing, novel expression domains, and transgressive (higher or lower than either progenitor diploid) expression amounts (one from the diploid, the other from the allopolyploid) should be similar, as they should for the two D-genome sequences. The utility of the null hypothesis lies in its falsification; if homoeologous sequences interact, for example, a different tree may be recovered, or if there is strong directional selection or pseudogenization, rate Layout: T1 Standard SC Chapter No.: 10 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 192/206 J. F. Wendel et al. 342 10.2.3.1 Molecular Evolution of Homoeologs and Orthologs 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 PR OO 346 D 345 Early work (Cronn et al. 1999; Senchina et al. 2003) demonstrated that duplicated genes typically are both retained even in natural (1–2 mya -ld) allopolyploids, and that they evolve essentially as modeled in the central panel of Fig. 10.3, i.e., equally, additively, and at equivalent rates. A more recent analysis (Flagel et al. 2012), based on a global assembly of 5 million Sanger and 454 ESTs supplemented by *150 million 82-bp Illumina reads, has provided a vastly expanded and detailed view of genic evolution in diploid and allopolyploid cotton, perhaps the most extensive yet for any plant genus. These data, representing analysis of *10,000 genes in each comparison, show that rates of synonymous substitution (Ks) between A- and D-genome orthologs (0.036) are nearly identical to that experienced by their homoeologous descendants in the allopolyploid genome (0.037), with non-synonymous substitution rates being identical in the two contrasts (0.009). These results demonstrate unequivocally that allopolyploidy has not, in general, been accompanied by an enhanced rate of nucleotide substitution in coding regions, as might be expected from an assumption of rapid decay of ‘‘redundant’’ duplicated copies. An additional perspective provided by Flagel et al. (2012) is that the data also provide an extraordinarily accurate depiction of the ancestry of polyploid cotton. By using thousands of genic alignments, the relative distances of the A- and D-genome diploids to their counterparts in the allopolyploid (AT, DT) can be calculated and effectively translated into relative divergence times because of the demonstration of rate homogeneity discussed above. The results show that the A genome diploid has a mean Ks of 0.009 from the AT-genome, whereas the comparable figure for the D genome comparison is 0.015. From this we infer that modern A-genome diploids are a better model (by about 50 %) of the actual genome donor of allopolyploid cotton than are extant D-genome diploids, consistent with previous suggestions based on diverse data sources (Wendel et al. 2009; Wendel and Cronn 2003; Senchina et al. 2003). These new data provide deep insight into the cotton model system, furthering its value for comparative analyses. A general expectation of molecular evolution in allopolyploids is that mutations occur randomly among homoeologs, and hence that evolutionary rates will be equivalent for homoeologs (Fig. 10.3). A corollary expectation is that duplicated gene copies will accumulate diversity, within and among populations, at equivalent rates. Small et al. (1999) tested this hypothesis of rate equivalence among homoeologs using the gene AdhA, for which both homoeologs were sequenced for TE 343 344 EC 340 CO RR 339 F 341 inequalities may become evident. Additional possibilities include silencing or loss of one of the duplicated copies, ‘‘horizontal transfer’’ of sequences from one genome to the other, and novel expression, the latter encompassing a variety of phenomena discussed under Duplicate gene expression (Sect. 10.4). 338 UN Editor Proof 192 Layout: T1 Standard SC Chapter No.: 10 Book ISBN: 978-3-642-31441-4 Page: 193/206 Jeans, Genes, and Genomes: Cotton as a Model for Studying Polyploidy 193 407 10.2.3.2 Interaction Among Homoeologs at the DNA Level 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 408 409 410 411 412 413 414 415 416 417 418 419 PR OO 386 D 385 TE 384 EC 382 383 CO RR 381 F 406 22 accessions (44 alleles per genome) of G. hirsutum and for five accessions (10 alleles per genome) of G. barbadense. In both allopolyploids, estimates of nucleotide diversity were higher for AdhA from the D-genome than from the Agenome. In a follow-up study (Small and Wendel 2002) using a second ADH gene (AdhC), this conclusion was even more strongly supported; here 24 different alleles were detected for the D-genome homoeolog versus only 7 for the A-genome homoeolog, with a similar increase in allelic diversity for the D-homoeolog in G. barbadense. These observations indicate that at least for some genes, homoeologs may accumulate synonymous substitutions at vastly different rates. At present, the responsible forces and underlying molecular mechanisms are obscure, but a logical suggestion is that they are causally connected to the nearly 2fold difference in genome size between the co-resident genomes in the allopolyploid nucleus (the AT genome is approximately twice the size of the DT genome). At present, little information exists for Gossypium that enables a thorough analysis of the relative rates of pseudogenization between diploids and allopolyploids, or between the two homoeologous genomes of allopolyploid cotton. The ongoing extensions of the comparative BAC sequencing approach taken by Grover and colleagues (Grover et al. 2004, 2007) are promising in this respect, as are the ongoing genome sequencing projects. Similarly, relatively little information exists on rates of gene loss in cotton, though evidence to date based on comparative BAC sequencing (Grover et al. 2004, 2007), comparative mapping analyses (Rong et al. 2004; Brubaker et al. 1999b), EST collections (Flagel et al. 2012), and AFLP studies (Liu et al. 2001) suggests that rates of gene loss are neither high nor particularly biased with respect to genomic origin. A recent study using Southern hybridization, however, detected three losses (of 27 genes studied) of D-homoeologs from allopolyploid cotton and no losses of the A-homoeolog, suggesting a possible bias (Rong et al. 2010). Again, key data likely will be generated soon that will permit these speculations to be evaluated. 379 380 One of the first indications that duplicated genes could behave in an evolutionarily dependent fashion was the study of Wendel et al. (1995), who demonstrated interaction among the 18S–26S ribosomal genes that exist at multiple loci in the A- and D-genomes. Instead of evolving independently, as expected if homoeologous repeats did not interact, repeats at the different arrays in allopolyploid cotton have been ‘‘homogenized’’ to the same sequence (either ‘‘A-like’’ or ‘‘D-like’’) by one or more processes of concerted evolution (reviewed by Elder and Turner 1995). In four of the five allopolyploid species, interlocus homogenization has created exclusively D-genome like rDNAs, whereas in G. mustelinum nearly all rDNA repeats have been homogenized to an A-like form. This example showed that since polyploid formation 1–2 mya, some 3,800 repeats, each approximately 10 kb in length, were ‘‘overwritten’’ with the alternative form originating from the UN Editor Proof 10 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 10 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 194/206 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 F 426 PR OO 425 D 423 424 TE 422 other parental genome, probably through unequal crossing over or gene conversion, and that this phenomenon operated bidirectionally, in different directions in different allopolyploid lineages. Interlocus concerted evolution of rDNA repeats has since been documented in many other plant polyploids, including Nicotiana and Tragopogon (see this Chaps. 10 and 14, this volume). From a mechanistic perspective, it seems likely that the homogenization has resulted from unequal crossing over within and among arrays and nonreciprocal homoeologous recombination events, with homoeolog exchanges being facilitated by the sub-telomeric location of rDNAs in many plant lineages, including Gossypium (Wendel 2000). This demonstration that some repeated sequences could interact across genomes in the allopolyploid nucleus led to additional investigations of the scope of the phenomenon. In an analogous study, Cronn et al. (1996) showed that in contrast to 18S–26S arrays, 5S rDNA genes are not homogenized by concerted evolutionary forces in the allopolyploid. Similarly, and as noted above, early studies demonstrated apparent independent evolution of homoeologs (Cronn et al. 1999; Senchina et al. 2003), consistent with the null hypothesis of Fig. 10.3. Thus, until recently it was thought that most nuclear genes duplicated by allopolyploidy largely evolve independently of one another in the polyploid nucleus, a proposition that seemed entirely reasonable given the absence of evident cytogenetic interactions among homoeologs. This presumption of homoeolog independence turns out to be incorrect; however, and for a surprisingly large percentage of duplicated genes. Using ESTs generated from both A- and D-genome diploids and AD-genome allopolyploids, Salmon et al. (2010) sought evidence of small ‘‘gene conversion’’ or nonreciprocal exchanges among homoeologs, based on the presence in allopolyploid cotton of genes that have genome-diagnostic (A or D) SNPs (single nucleotide polymorphisms) that occurred in patterns suggestive of genic interactions (e.g., a single expressed sequence containing diagnostic SNPs in the following order, AADDAA, implicating conversion of the middle section of an AT homoeolog by the DT homoeolog). These bioinformatic inferences were validated by de novo PCR and sequencing. Results were convincing and compelling, demonstrating that about 2 % of contigs in G. hirsutum have experienced nonreciprocal homoeologous exchanges since the origin of polyploid cotton 1–2 mya. Moreover, when a sampling of these homoeologous interactions was studied throughout the polyploid clade using a phylogenetic approach, nonreciprocal homoeologous exchanges were shown to have occurred throughout polyploid divergence and speciation, as opposed to being concentrated at the root of the polyploid tree. Among six homoeologous interactions, five occurred in only one species, with the sixth event being shared. This result refutes a logical a priori prediction, namely that interactions among homoeologs should be most frequent early in allopolyploid evolution (and hence shared among species), prior to reinforcement of apparently strict bivalent pairing. Finally, some genomic regions showed multiple patterns of homoeologous recombination among species, suggesting that some regions or genes may be ‘‘hot-spots’’ for nonreciprocal homoeologous exchanges. EC 421 CO RR 420 J. F. Wendel et al. UN Editor Proof 194 Layout: T1 Standard SC Chapter No.: 10 Book ISBN: 978-3-642-31441-4 Page: 195/206 Jeans, Genes, and Genomes: Cotton as a Model for Studying Polyploidy 195 486 10.3 Duplicate Gene Expression 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 PR OO 469 D 468 TE 467 EC 466 In addition to evolutionary changes in gene and genome structure, a key component of polyploid evolution concerns the consequences of genome doubling on gene expression. The biological reunion in the allopolyploid nucleus of two regulatory systems that had evolved in isolation for 5–10 my (A- and D-genome), in only one of the two parental cytoplasms (A-genome), might be expected to lead to violations of the equal and additive expression modeled in Fig. 10.3. In principle, both genomes in an allotetraploid could contribute equally to the transcriptome, for any pair of homoeologs or overall, or alternatively, there may be preferential transcription of one genome due to intergenomic interactions that could bias the transcription machinery. Many aspects of this problem are of interest, including the scope and scale of preferential homoeolog expression, its tissue and organ specificity, the level of genomic bias, the mechanistic underpinnings that result in regulatory responses, the temporal scale at which alterations in homoeolog expression evolve, and ultimately, its physiological and evolutionary relevance. Steps toward answering these questions for Gossypium have been taken using many different experimental approaches over the last decade. We summarize these results below, focusing on lessons regarding (1) developmental and environmental CO RR 465 F 485 This demonstration of non-independent evolution of homoeologs has recently been confirmed and extended to tens of thousands of genes using an expanded EST set (Flagel et al. 2012). This analysis resulted in an even higher percentage (5 %) of contigs in the assembly that show evidence of homoeolog contact in at least one of the two polyploid species studied (G. hirsutum and G. barbadense), as well as confirm the suggestion of hot-spots for homoeologous recombination. This high frequency of homoeologous contact is an astonishing result given the absence of prior cytogenetic or other observations that would have suggested this possibility, and the apparently complete bivalent formation at meiosis. Flagel et al. (2012) also detected 50 % more homoeologous exchanges in G. hirsutum than in G. barbadense, suggesting that following polyploid formation, rates of nonreciprocal homoeologous recombination may diverge, even among closely related species that share nearly all evident ecological and life-history characteristics. These results are fascinating and lead to a number of questions. For example, what is the genomic distribution of genes subject to nonreciprocal recombination events, and does this information suggest a causal mechanism? Are nonreciprocal recombination events random with respect to outcome (A vs D), or alternatively, is the nuclear genome of allopolyploid cotton slowly becoming more ‘‘A-like’’ or D-like’’? Do any of the detected ‘‘gene conversion’’ events have physiological consequences and hence possible adaptive significance? Insight into these and related questions in cotton and in other allopolyploids is likely to emerge in the next several years. 464 UN Editor Proof 10 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 10 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 196/206 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 F PR OO 508 509 The first indication that polyploidy in Gossypium is accompanied by extensive organ-specific changes in duplicate gene expression emerged nearly a decade ago in a study of 40 homoeologous gene pairs in different organs of G. hirsutum using SSCP-cDNA (Adams et al. 2003). Almost one-third of the genes revealed bias toward one homoeolog or the other, or only expression of one homoeolog, in at least one organ. Transcript levels for the two members of each gene pair varied by gene and, unexpectedly, by organ type. Especially noteworthy were genes that showed organ-specific, reciprocal silencing of alternative homoeologs; that is, one member of a duplicated gene pair displayed minimal to no transcription in some organs, whereas a reciprocal pattern was exhibited by the alternative homoeolog in other organ(s). In particular, floral organs showed dramatic expression patterns in this regard, with major differences among petals, stamens, and stigmas/styles (Adams et al. 2003, 2004). Organ-specific expression of homoeologs was assayed more extensively by Chaudhary et al. (2009), who employed a novel high-resolution mass-spectrometry technology (Sequenom) to investigate relative expression levels of each homoeolog for 63 gene pairs in 24 tissues. Results from over 2,000 assays demonstrated that 40 % of homoeologs are transcriptionally biased in at least one stage of cotton development, that genome merger per se has a large effect on relative expression of homeologs (see section on temporal partitioning, below), and that the majority of alterations are caused by cis-regulatory divergence between diploid progenitors. The study also revealed 15 cases of probable regulatory neofunctionalization among 8 tissues, perhaps the first such demonstration in allopolyploid plants. These surprising indications of partitioning of duplicate gene expression during development have since been found in many other allopolyploid plants (see also Chap. 14, this volume, and, e.g., Buggs et al. 2010; Bottley and Koebner 2008; Bottley et al. 2006; Wang et al. 2006), opening the experimental floodgates on the study of duplicate gene expression. Because differences in duplicate gene expression were found between different organ types, the question arose as to patterns of duplicate gene expression during development of any single organ. Liu and Adams (2007) examined the expression ratio of homoeologous AdhA genes in eight developmental stages of hypocotyls, cotyledons, and roots, and 11 developmental stages of ovary walls and ovules. They showed that expression ratios of the two homoeologs changed considerably when comparing some stages of organ development, indicating that differential expression of homoeologous genes is developmentally regulated and that determining the extent of homoeologous gene D 507 10.3.1 Developmental and Environmental Effects on Duplicate Gene Expression TE 506 effects on gene expression; (2) global analyses of genomic interactions; and (3) the temporal scale at which expression evolution arises. EC 505 CO RR 504 J. F. Wendel et al. UN Editor Proof 196 Layout: T1 Standard SC Chapter No.: 10 Book ISBN: 978-3-642-31441-4 Page: 197/206 Jeans, Genes, and Genomes: Cotton as a Model for Studying Polyploidy 197 577 10.3.2 Global Biases in Duplicate Gene Expression 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 578 579 580 581 582 583 584 PR OO 549 D 548 TE 547 EC 546 CO RR 545 F 576 expression changes requires analysis of multiple developmental stages and organs. An important extension of this line of analysis was to ask whether biased expression of homoeologs can be cell type-specific. In this respect, cotton offers an excellent opportunity because the single-celled fibers are easily accessible. To address this question, Hovav et al. (2008a, b) used homoeolog-specific microarrays to assay homoeologous gene expression during fiber development. In these remarkable studies, discussed further below, a full range of duplicate gene expression was observed among the approximately 1,400 genes, from balanced expression of homoeologs, to varying homoeolog ratios that shifted during development. Notably, four genes showed complete reciprocal silencing of alternative homoeologs during fiber development. Collectively, the foregoing studies show that variation in homoeologous gene expression during development is the rule rather than the exception, and that it occurs even during development of a single cell type. An additional twist on duplicate gene expression has been the demonstration that homoeolog expression patterns are not simply developmentally fixed, but instead are responsive to environmental conditions. Liu and Adams (2007) examined expression of AdhA homoeologs in response to abiotic stress conditions, with a focus on cold and water submersion (flooding simulation), at different developmental stages. They found that some stress treatments significantly altered AT to DT expression ratios, including a case of reciprocal silencing of homoeologs in response to abiotic stress, an unprecedented finding. In a more recent study, Dong and Adams (2011) extended their analysis to 30 gene pairs using three organs (leaves, roots, and cotyledons) and five abiotic stress treatments (heat, cold, drought, high salt, and water submersion). Over 70 % of the genes showed stressinduced changes in the relative expression levels of the duplicates under one or more stress treatments, and 12 sets of homoeologs showed opposite changes in expression levels in response to different abiotic stress treatments. These results indicate that abiotic stress conditions can have considerable effects on duplicate gene expression in a polyploid, with the effects varying by gene, stress, and organ type. Together with the profound developmental partitioning noted above, differential expression in response to environmental stresses or cues may all be factors that contribute to the preservation of duplicated genes in polyploids. 544 To provide a more global perspective on homologous gene expression in allopolyploids, Udall et al. (2006) developed homoeolog-specific microarrays that utilized genome-diagnostic SNPs from ESTs generated from the two genomes of allopolyploid cotton and the diploids G. arboreum (A-genome) and G. raimondii (D-genome). Using leaf RNA, they found that 199 of 461 gene pairs (43 %) deviated from equal expression, thereby providing an initial quantitative perspective on the scale of biased homoeolog expression. This microarray UN Editor Proof 10 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 10 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 198/206 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 F 591 PR OO 590 D 589 TE 588 EC 587 approach was refined and extended to a larger set of gene pairs (n = 1383) by Flagel et al. (2008), who reported that 70 % of the genes in petals have biased homoeolog expression ratios, and that more of these genes are D-genome (39.5 %) than A-genome (30.5 %) biased. In addition, they found that the D-genome copies of 69 genes and the A-genome copies of 46 genes were silenced, collectively representing about 8 % of all genes. Similarly, Hovav et al. (2008a, b) used the same SNP-specific microarray technology on RNAs extracted from trichomes harvested from three developmental time points in wild and modern accessions of two independently domesticated cotton species, G. hirsutum and G. barbadense. Among these species 25–37 % of genes were significantly biased toward one of the two genomes at each developmental stage, but these biases were not random with respect to genome-of-origin; instead, duplicate gene expression was biased toward the D-genome at all three time points studied, accounting for 63–76 % of the biased genes. Finally, and to place homoeolog expression evolution in a phylogenetic context, Flagel and Wendel (2010) extended their analysis of duplicate gene expression in petals to all five extant allopolyploid species. Several aspects of this study are notable, including the demonstration that all five species display an overall preference for D-genome expression (D-genome bias accounting for 54–60 % of genes with biased homoeolog expression), that the percentage of duplicate genes that are biased varies widely among species (from 48 to 88 %), and that the overall magnitude of bias (as opposed to simply whether or not a gene exhibits bias) similarly varies widely among species. The foregoing synopsis shows that gene expression is massively altered in polyploid cotton relative to its diploid progenitors. Unequal expression of one of the two homoeologs likely is the rule rather than the exception, when integrated across organs and tissues, with gene silencing representing only the endpoint in a continuum of biased homoeolog expression. A global bias in gene expression from one of two co-resident genomes (the D-genome; see, however, Yang et al. 2006) is evident in allopolyploid cotton. At present, little evidence connects these observations to physiology or metabolism, and so connections to function and ecological relevance remain obscure (see, however, section on fiber below). Recent data (Hu et al. 2011) have extended the concept of bias to the proteomic level, where allopolyploid cotton seeds have been shown to preferentially accumulate proteins from the D-genome diploid parent, suggesting the possibility of a functional connection. A promising direction for future research is targeted, functional studies of specific gene families, protein complexes, and metabolic pathways and networks, aiming to place homoeolog expression into a context that permits insight into the biological significance of expression bias. An emerging phenomenon in the study of polyploid gene expression on a large scale is the concept of genomic dominance in expression. This phenomenon, originally discovered and elaborated in cotton (Flagel and Wendel 2010; Rapp et al. 2009), and just recently detected in Coffea allopolyploids (Bardil et al. 2011), is defined as the state where total expression of a homoeolog pair mimics the expression level of one of two diploid parents of an allopolyploid. That is, if parent A is upregulated relative to parent B, the allopolyploid would exhibit the CO RR 585 586 J. F. Wendel et al. UN Editor Proof 198 Layout: T1 Standard SC Chapter No.: 10 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 F 635 PR OO 634 10.3.3 Temporal Dynamics of Duplicate Gene Expression Evolution One of the early (Adams et al. 2003) and subsequently further characterized findings about duplicate gene expression in cotton is that expression evolution has a temporal dimension; phenomena that typify genome merger and doubling need not be the same, qualitatively or quantitatively, as those that characterize natural allopolyploids that are 1–2 million years distant from polyploidization (Fig. 10.4). In some cases, such as organ-specific silencing of particular homoeologs, it has been shown that this can arise immediately upon allopolyploid formation (Adams et al. 2003, 2004; Chaudhary et al. 2009), leading to the notion (Adams et al. 2003; Rapp and Wendel 2005) that homoeolog silencing may be significant in duplicate gene retention and that it may actually represent a form of instantaneous subfunctionalization. These studies of individual genes have demonstrated that some of the changes detected in synthetic allopolyploids are mirrored in natural allopolyploids, whereas others are distinct between the synthetic and natural allopolyploids (see examples in Adams et al. 2004; Chaudhary et al. 2009). One possibility raised by these observations is that in some cases there is an immediate epigenetic response to genome merger and doubling that becomes either epigenetically stable or genetically fixed, for 1–2 million years. With respect to the phenomena of genomic bias, the evolution of novel quantitative (transgressive) expression states, and genomic dominance, the studies of Flagel et al. (2008, 2010) have been particularly illuminating with respect to this temporal dimension to expression evolution. For example, whereas only about one-third of genes exhibit biased homoeolog expression in petals of AxD-genome, diploid hybrids, half or more do in petals of natural allopolyploids. Moreover, many of the genes that display bias in the initial F1 retain or even magnify the degree of bias after 1–2 million years. Also, whereas allopolyploid formation appears to result in relatively few genes that exhibit quantitatively transgressive expression levels, relative to their diploid parents, this number rises sharply in natural allopolyploids, suggesting novel opportunities for exploring gene expression space. Finally, a comparison of genomic dominance in all five natural allopolyploids (Flagel and Wendel 2010) to that observed in synthetic allopolyploids (Rapp et al. 2009) demonstrated that the magnitude of genomic dominance remains high after D 633 TE 632 199 expression state of parent A; in the case where parent A is downregulated relative to parent B, so is the allopolyploid. In this example, the A parent is genomically dominant with respect to gene expression. Using two different genomic combinations in synthetic allopolyploid Gossypium, Rapp et al. (2009) demonstrated that there exists a quantitative bias in genomic dominance toward one of the two parents. In the case of the synthetic AD allopolyploid, more than 10,000 genes display this dominant, quantitative expression phenotype. At present, the functional significance of genomic dominance is unknown. EC 631 CO RR 630 Book ISBN: 978-3-642-31441-4 Page: 199/206 Jeans, Genes, and Genomes: Cotton as a Model for Studying Polyploidy UN Editor Proof 10 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 10 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 200/206 J. F. Wendel et al. Editor Proof 200 G. hirsutum AD AADD PR OO G. barbadense G. hirsutum F G. tomentosum G. darwinii G. barbadense G. mustelinum “Gene conversion” Gene loss Genomic dominance Concerted evolution Genetic sub- and neofunctionalization Intergenomic transfer Biased homoeolog expression Transgressive expression Homoeologous expression change Biased genomic dominance Epigenetic subfunctionalization 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 1–2 my of evolution, but that the bias in its direction dissipates (shifting from predominantly D-genome dominance in the synthetics to similar levels of A- and D-genome dominance in the natural allopolyploids). Thus, it appears that natural allopolyploids have adjusted to more equally utilize the transcriptomes of the two co-resident genomes, notwithstanding the residual D-genome homoeolog bias described above. An intriguing possibility is that this D-genome homoeolog bias is connected causally to the biased D-genome expression dominance that arises following genome merger; to the extent that it is, it leads to the suggestion that homoeolog bias in other allopolyploids may be predicted by the initial conditions established by genomic merger in the distant past, and which may be experimentally mirrored in many systems through the use of synthetic hybrids and allopolyploids. The partitioning of gene expression evolution into its temporal components leads to the suggestion that these different components may entail different or at least complementary mechanisms. The first, involving rapid or instantaneous gene expression alteration as a consequence of genome merger and doubling, reflects the myriad novel interactions accompanying a biological reunion of two differentiated genomes into a common nucleus. The precise nature of these interactions CO RR 672 UN 671 EC TE D Fig. 10.4 A temporal depiction of phenomena that characterize polyploid evolution in Gossypium. Shown are the formation and diversification of the allopolyploid clade, several accompanying phenomena (dispersal, chromosomal stabilization), and the parallel domestication of G. hirsutum and G. barbadense. Images to the right represent wild and domesticated states, with close-ups of single seeds and the transformations in morphology accompanying domestication and crop improvement. Genetic and epigenetic phenomena associated with polyploidy in cotton are shown below, with a timeline (bottom), illustrating the temporal context for each phenomenon and whether there are increases or decreases (where known) in magnitude over time (blue to purple gradation indicates increase, the reverse indicates decrease) Layout: T1 Standard SC Chapter No.: 10 Book ISBN: 978-3-642-31441-4 Page: 201/206 Jeans, Genes, and Genomes: Cotton as a Model for Studying Polyploidy 201 699 10.4 Polyploidy and Ecological Novelty 695 696 697 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 PR OO 694 The foregoing discussion of the genomic and genetic attributes of allopolyploid cotton demonstrates that polyploid formation has led to a diverse array of genetic and genomic responses, including non-Mendelian transmission and various forms of non-additive gene expression. The question naturally arises as to whether allopolyploidy has stimulated novel adaptation or physiological capability. A voluminous literature in plants documents the frequency of polyploids in various habitats, their morphological and physiological attributes, and their ecological success relative to diploids (Grant 1981; Stebbins 1947, 1950; Soltis and Soltis 2000). One generalization that has emerged is that polyploidy often is associated with broader ecological amplitude and novel evolutionary opportunity, often suggested to be mediated by the increased ‘‘buffering’’ capacity afforded by duplicated genes and the enhanced vigor resulting from the ‘‘fixed heterozygosity’’ of their duplicated genomes. We might now rephrase or expand these suggested mechanistic explanations to encompass a network perspective, and the vastly increased combinatorial possibilities for regulation and evolution enabled by a suddenly duplicated complement of genes and merged regulatory systems. With respect to Gossypium, allopolyploidy led to the apparent invasion of a new ecological niche. In considering the Pleistocene origin of allopolyploid cotton, Fryxell (1965, 1979) noted that in contrast to the majority of diploid species, which occur mostly inland in various arid to seasonally arid environments, allopolyploid species typically occur in coastal habitats, at least those forms that arguably are truly wild (see also Brubaker and Wendel 1994). Thus, among the five allopolyploid species, two are completely restricted to near coastlines, in that they are island endemics (G. darwinii and G. tomentosum), and for two others (G. barbadense and G. hirsutum), wild forms occur disparately in littoral habitats ringing the Gulf of Mexico, northwest South America, and distant Pacific Islands. The capacity for oceanic dispersal in Gossypium (Fryxell 1965, 1979; Stephens 1958, 1966) was associated at the allopolyploid level with specialization for establishment in coastal communities. Fryxell (1965, 1979) forwarded the tantalizing suggestion that following initial formation, adaptation of the newly evolved D 693 TE 692 EC 691 CO RR 690 F 698 is not known, but probably includes disruptions in gene dosage balance, stoichiometric changes resulting from differences in competition for transcription factors, differences in microRNA expression, and a host of novel cis and trans interactions (Birchler et al. 2005; Birchler and Veitia 2010; Chen 2007; Osborn et al. 2003; Veitia 2005) (see also Chaps. 1, 2, and 4, this volume). Superimposed on these rapid evolutionary responses to polyploidy are those that arise more slowly during the stabilization of the new polyploid genome and during evolution and speciation over much longer periods of time. Thus, the presence of duplicated genomes would seem to provide evolutionary opportunity and consequences immediately on polyploid formation and for millions of years thereafter. 689 UN Editor Proof 10 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 10 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 202/206 J. F. Wendel et al. 737 10.5 Polyploidy and Fiber 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768 769 770 PR OO 735 Finally, we would be remiss if we failed to consider the consequences of polyploidy on the development of agronomically advanced cultivars of allopolyploid cotton. Although four separate species of Gossypium were independently domesticated for their seed hairs, the characteristic that attracted the attention of the earliest domesticators, the seed ‘‘lint’’ itself, evolved only once in the progenitor of the A-genome diploids (Hovav et al. 2008c; Wendel et al. 2009). Applequist et al. (2001) generated growth curves for trichomes from cultivated and wild diploid and allopolyploid species and demonstrated that the evolution of an extended primary wall elongation occurred in the ancestor of wild A-genome cotton prior to domestication and in Africa. Follow-up comparative expression profiling experiments (Hovav et al. 2008a, b; Rapp et al. 2010) further identify some of the metabolic pathways that were modified to enable this evolution in fiber properties. These results led to the fascinating implication that domestication of the New World allopolyploid cottons (which contain an A-genome, in addition to a D-genome) that presently dominate cotton agriculture worldwide was first precipitated by developmental and physiological transformations that occurred hundreds of thousands of years ago in a different hemisphere. Because fibers from all D-genome diploids are short and non-spinnable, it is particularly interesting that fiber from the cultivated (New World) allopolyploids is agronomically superior to that of the cultivated A-genome diploids; in this sense, the cultivated allotetraploid fiber morphology is ‘‘non-additive’’, or perhaps ‘‘heterotic’’. A number of studies have noted this point (Jiang et al. 1998; Paterson 2005; Wright et al. 1998), suggesting that allopolyploidization provided novel opportunities for agronomic improvement. A recent meta-analysis (Rong et al. 2007) of a large number of QTL studies in allopolyploid cotton leads to a general picture consistent with this interpretation; more loci affecting fiber yield and quality traits are found in the DT (n = 221) than the AT (n = 184) genome, possibly explaining the superiority of the lint of the allopolyploids relative to the A-genome diploids. Support for this speculation that ‘‘recruitment’’ of D-genome genes has been important in enabling the development of advanced allotetraploid cultivars is also emerging from comparative expression profiling studies (Hovav et al. 2008a, b; Rapp et al. 2010), which reveal in exquisite detail the thousands of gene expression differences that distinguish wild from domesticated cotton fiber D 734 TE 733 EC 732 CO RR 731 F 736 allopolyploid to littoral habitats enabled it to exploit the fluctuating sea levels that characterized the Pleistocene. This ecological innovation not only is envisioned to have permitted the initial establishment of the nascent polyploid lineage, but is also suggested to have provided a means for the rapid dispersal of the salt-water-tolerant seeds. By this means, perhaps, the mobile shorelines of the Pleistocene facilitated exploitation of a new ecological niche, and hence colonization of the New World tropics. 730 UN Editor Proof 202 Layout: T1 Standard SC Chapter No.: 10 Book ISBN: 978-3-642-31441-4 Page: 203/206 Jeans, Genes, and Genomes: Cotton as a Model for Studying Polyploidy 203 788 789 790 791 792 Acknowledgments Research in the Wendel lab has largely been funded by the NSF Plant Genome Program, with additional support from other NSF programs, the USDA NRI, and Cotton Incorporated. Research in the Adams lab has been supported by the Natural Science and Engineering Research Council of Canada and by the USDA NRI program. We gratefully acknowledge all of these sources of support. 793 References 794 795 796 797 798 799 800 801 802 803 804 805 806 807 808 809 810 811 812 813 Adams KL, Cronn R, Percifield R, Wendel JF (2003) Genes duplicated by polyploidy show unequal contributions to the transcriptome and organ-specific reciprocal silencing. Proc Nat Acad Sci USA 100:4649–4654 Adams KL, Percifield R, Wendel JF (2004) Organ-specific silencing of duplicated genes in a newly synthesized cotton allotetraploid. Genetics 168:2217–2226 Applequist WL, Cronn RC, Wendel JF (2001) Comparative development of fiber in wild and cultivated cotton. Evol Dev 3:3–17 Bardil A, Dantas de Almeida J, Combes MC, Lashermes P, Bertrand B (2011) Genomic expression dominance in the natural allopolyploid Coffea arabica is massively affected by growth temperature. New Phytol (in press) Beasley JO (1941) Hybridization, cytology, and polyploidy of Gossypium. Chronica Bot 6:394–395 Bennett MD, Leitch IJ (2005) Genome size evolution in plants. In: Gregory TR (ed) The evolution of the genome. Elsevier, San Diego, pp 89–162 Birchler JA, Riddle NC, Auger DL, Veitia RA (2005) Dosage balance in gene regulation: biological implications. Trends Genet 21:219–226 Birchler JA, Veitia RA (2010) The gene balance hypothesis: implications for gene regulation, quantitative traits and evolution. New Phytol 186:54–62 Bottley A, Koebner RM (2008) Variation for homoeologous gene silencing in hexaploid wheat. Plant J 56:297–302 778 779 780 781 782 783 784 785 786 PR OO 777 D 775 776 TE 774 EC 773 CO RR 772 F 787 development, as well as a bias toward preferential expression of D-genome homoeologs. Finally, similar implications emanate from genetic mapping experiments, where it has been shown that for 535 genes implicated in cotton fiber development, more transcription factors were from DT than AT genome, whereas the reverse was true for fiber development genes (Xu et al. 2010). These data are interpreted to suggest that the D-genome ancestor provided key transcription factors that regulate the expression of fiber genes donated by the ancestral A-genome parent. Taken together, these studies may provide actual genetic evidence for a speculation forwarded 75 years ago by Harland (1936), who stated If as a consequence of polyploidy a large number of genes become duplicated, and the characters governed by such genes are of importance to the species, one of the members may mutate, leaving the character unimpaired, with the further possibility that the mutation may be of benefit to the species. An exciting prospect is that in the near future we will develop a deeper understanding of the nature of these genes, the molecular genetic meaning of Harland’s invocation of the word ‘‘mutation’’, and their effects on the developmental networks that underlie altered morphology and agronomic improvement. 771 UN Editor Proof 10 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 10 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 204/206 EC TE D PR OO F Bottley A, Xia G, Koebner R (2006) Homoeologous gene silencing in hexaploid wheat. Plant J 47:897–906 Brubaker CL, Bourland FM, Wendel JF (1999a) The origin and domestication of cotton. In: Smith CW, Cothren JT (eds) Cotton; origin, history, technology and production. Wiley, New York, pp 3–31 Brubaker CL, Paterson AH, Wendel JF (1999b) Comparative genetic mapping of allotetraploid cotton and its diploid progenitors. Genome 42:184–203 Brubaker CL, Wendel JF (1993) On the specific status of Gossypium lanceolatum Todaro. Genet Res Crop Evol 40:165–170 Brubaker CL, Wendel JF (1994) Reevaluating the origin of domesticated cotton (Gossypium hirsutum; Malvaceae) using nuclear restriction fragment length polymorphisms (RFLPs). Am J Bot 81:1309–1326 Brubaker CL, Wendel JF (2001) RFLP diversity in cotton. In: Jenkins JN, Saha S (eds) Genetic improvement of cotton: emerging technologies. Science Publishers, Inc., Enfield, pp 81–102 Buggs RJA, Elliott NM, Zhang L, Koh J, Viccini LF, Soltis DE, Soltis PS (2010) Tissue-specific silencing of homoeologs in natural populations of the recent allopolyploid Tragopogon mirus. New Phytol 186:175–183 Chaudhary B, Flagel L, Stupar RM, Udall JA, Verma N, Springer NM, Wendel JF (2009) Reciprocal silencing, transcriptional bias and functional divergence of homeologs in polyploid cotton (Gossypium). Genetics 182:503–517 Chen ZJ (2007) Genetic and epigenetic mechanisms for gene expression and phenotypic variation in plant polyploids. Annu Rev Plant Biol 58:377–406 Cronn R, Small RL, Wendel JF (1999) Duplicated genes evolve independently following polyploid formation in cotton. Proc Nat Acad Sci USA 96:14406–14411 Cronn RC, Zhao X, Paterson AH, Wendel JF (1996) Polymorphism and concerted evolution in a tandemly repeated gene family: 5S ribosomal DNA in diploid and allopolyploid cottons. J Mol Evol 42(6):685–705 DeJoode DR, Wendel JF (1992) Genetic diversity and origin of the Hawaiian Islands cotton, Gossypium tomentosum. Amer J Bot 79:1311–1319 Dong S, Adams KL (2011) Differential contributions to the transcriptome of duplicated genes in response to abiotic stresses in natural and synthetic polyploids. New Phytol 190(4):1045–1057 Elder JF, Turner BJ (1995) Concerted evolution of repetitive DNA sequences in eukaryotes. Quart Rev Biol 70:297–320 Endrizzi JE, Turcotte EL, Kohel RJ (1985) Genetics, cytology, and evolution of Gossypium. Adv Genet 23:271–375 Flagel L, Udall J, Nettleton D, Wendel J (2008) Duplicate gene expression in allopolyploid Gossypium reveals two temporally distinct phases of expression evolution. BMC Biol 6:11 Flagel LE, Wendel JF (2010) Evolutionary rate variation, genomic dominance and duplicate gene expression evolution during allotetraploid cotton speciation. New Phytol 186:184–193 Flagel LE, Wendel JF, Udall JA (2012) Duplicate gene evolution, homoeologous recombination, and transcriptome characterization in allopolyploid cotton. BMC Genomics (in press) Fryxell PA (1965) Stages in the evolution of Gossypium. Adv Frontiers Plant Sci 10:31–56 Fryxell PA (1968) A redefinition of the tribe Gossypieae. Bot Gaz 129:296–308 Fryxell PA (1979) The natural history of the cotton tribe Texas. A&M University Press College Station, College Station Fryxell PA (1992) A revised taxonomic interpretation of Gossypium L. (Malvaceae). Rheedea 2:108–165 Galau GA, Wilkins TA (1989) Alloplasmic male sterility in AD allotetraploid Gossypium hirsutum upon replacement of its resident a cytoplasm with that of D species G. harknessii. Theor Appl Genet 78:23–30 Grant V (1981) Plant speciation. Columbia University Press, New York Grover CE, Kim H, Wing RA, Paterson AH, Wendel JF (2004) Incongruent patterns of local and global genome size evolution in cotton. Genome Res 14:1474–1482 CO RR 814 815 816 817 818 819 820 821 822 823 824 825 826 827 828 829 830 831 832 833 834 835 836 837 838 839 840 841 842 843 844 845 846 847 848 849 850 851 852 853 854 855 856 857 858 859 860 861 862 863 864 865 866 J. F. Wendel et al. UN Editor Proof 204 Layout: T1 Standard SC Chapter No.: 10 205 EC TE D PR OO F Grover CE, Kim H, Wing RA, Paterson AH, Wendel JF (2007) Microcolinearity and genome evolution in the AdhA region of diploid and polyploid cotton (Gossypium). Plant J 50: 995–1006 Hanson RE, Islam-Faridi MN, Crane CF, Zwick MS, Czeschin DG, Wendel JF, Mcknight TD, Price HJ, Stelly DM (1999) Ty1-copia-retrotransposon behavior in a polyploid cotton. Chromosome Res 8:73–76 Hanson RE, Zhao X-P, Islam-Faridi MN, Paterson AH, Zwick MS, Crane CF, McKnight TD, Stelly DM, Price HJ (1998) Evolution of interspersed repetitive elements in Gossypium (Malvaceae). Am J Bot 85:1364–1368 Harland SC (1936) The genetical conception of the species. Cambridge Philos Soc Biol Rev 11:83–112 Hawkins JS, Hu G, Rapp RA, Grafenberg JL, Wendel JF (2008) Phylogenetic determination of the pace of transposable element proliferation in plants: copia and LINE-like elements in Gossypium. Genome 51:11–18 Hawkins JS, Kim H, Nason JD, Wing RA, Wendel JF (2006) Differential lineage-specific amplification of transposable elements is responsible for genome size variation in Gossypium. Genome Res 16(10):1252–1261 Hawkins JS, Proulx SR, Rapp RA, Wendel JF (2009) Rapid DNA loss as a counterbalance to genome expansion through retrotransposon proliferation in plants. Proc Nat Acad Sci USA 106(42):17811–17816 Hendrix B, Stewart JM (2005) Estimation of the nuclear DNA content of Gossypium species. Ann Bot 95:789–797 Hovav R, Chaudhary B, Udall JA, Flagel L, Wendel JF (2008a) Parallel domestication, convergent evolution and duplicated gene recruitment in allopolyploid cotton. Genetics 179(3):1725–1733 Hovav R, Udall J, Chaudhary B, Flagel L, Rapp R, Wendel J (2008b) Partitioned expression of duplicated genes during development and evolution of a single cell in a polyploid plant. Proc Nat Acad Sci USA 105:6191 Hovav R, Udall JA, Chaudhary B, Hovav E, Flagel L, Hu G, Wendel JF (2008c) The evolution of spinnable cotton fiber entailed prolonged development and a novel metabolism. PLoS Genet 4:e25 Hu G, Hawkins JS, Grover CE, Wendel JF (2010) The history and disposition of transposable elements in polyploid Gossypium. Genome 53:599–607 Hu G, Houston NL, Pathak D, Schmidt L, Thelan JJ, Wendel JF (2011) Genomically biased accumulation of seed storage proteins in allopolyploid cotton. Genetics 189:1103–1115 Hutchinson JB (1951) Intra-specific differentiation in Gossypium hirsutum. Heredity 5:161–n193 Hutchinson JB (1954) New evidence on the origin of the old world cottons. Heredity 8: 225–241 Hutchinson JB, Silow RA, Stephens SG (1947) The evolution of Gossypium and the differentiation of the cultivated cottons. Oxford University Press, London Jiang C, Wright R, El-Zik K, Paterson A (1998) Polyploid formation created unique avenues for response to selection in Gossypium (cotton). Proc Nat Acad Sci USA 95:4419–4424 Kashkush K, Feldman M, Levy AA (2002) Gene loss, silencing, and activation in a newly synthesized wheat allotetraploid. Genetics 160:1651–1659 Krapovickas A, Seijo G (2008) Gossypium ekmanianum (Malvaceae), algodon silvestre de la Republica Dominicana. Bonplandia 17:55–63 Lin L, Paterson AH (2009) Physical composition and organization of the Gossypium genomes. In: Paterson AH (ed) Genomics of cotton, plant genetics and genomics, crops and models 3. Springer, New York, pp 141–156 Lin LF, Tang HB, Compton RO, Lemke C, Rainville LK, Wang XY, Rong JK, Rana MK, Paterson AH (2011) Comparative analysis of Gossypium and Vitis genomes indicates genome duplication specific to the Gossypium lineage. Genomics 97:313–320 Liu B, Brubaker CL, Mergeai G, Cronn RC, Wendel JF (2001) Polyploid formation in cotton is not accompanied by rapid genomic changes. Genome 44:321–330 CO RR 867 868 869 870 871 872 873 874 875 876 877 878 879 880 881 882 883 884 885 886 887 888 889 890 891 892 893 894 895 896 897 898 899 900 901 902 903 904 905 906 907 908 909 910 911 912 913 914 915 916 917 918 919 920 Book ISBN: 978-3-642-31441-4 Page: 205/206 Jeans, Genes, and Genomes: Cotton as a Model for Studying Polyploidy UN Editor Proof 10 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 10 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 206/206 EC TE D PR OO F Liu Z, Adams KL (2007) Expression partitioning between genes duplicated by polyploidy under abiotic stress and during organ development. Curr Biol 17:1669–1674 Muravenko O, Fedotov AR, Punina EO, Federova LI, Grif VG, Zelenin AV (1998) Comparison of chromosome BrdU-hoechst-giemsa banding patterns of the A1 and (AD)2 genomes of cotton. Genome 41:616–625 Osborn TC, Chris Pires J, Birchler JA, Auger DL, Jeffery Chen Z, Lee H-S, Comai L, Madlung A, Doerge RW, Colot V, Martienssen RA (2003) Understanding mechanisms of novel gene expression in polyploids. Trends Genet 19(3):141–147 Paterson AH (2005) Polyploidy, evolutionary opportunity, and crop adaptation. Genetica 123:191 Paterson AH (2009) Genomics of cotton, plant genetics and genomics, crops and models 3. Springer, New York Percy RG, Wendel JF (1990) Allozyme evidence for the origin and diversification of Gossypium barbadense L. Theor Appl Genet 79:529–542 Rapp R, Haigler C, Flagel L, Hovav R, Udall J, Wendel J (2010) Gene expression in developing fibres of Upland cotton (Gossypium hirsutum L.) was massively altered by domestication. BMC Biol 8:139 Rapp R, Wendel J (2005) Epigenetics and plant evolution. New Phytol 168:81 Rapp RA, Udall JA, Wendel JF (2009) Genomic expression dominance in allopolyploids. BMC Biol 7:18 Reinisch AJ, Dong J, Brubaker CL, Stelly DM, Wendel JF, Paterson AH (1994) A detailed RFLP map of cotton, Gossypium hirsutum x G. barbadense: chromosome organization and evolution in a disomic polyploid genome. Genetics 138:829–847 Rong J, Abbey C, Bowers JE, Brubaker CL, Chang C, Chee PW, Delmonte TA, Ding X, Garza JJ, Marler BS, Park C, Pierce GJ, Rainey KM, Rastogi VK, Schulze SR, Trolinder NL, Wendel JF, Wilkins TA, Williams-Coplin TD, Wing RA, Wright RJ, Zhao X, Zhu L, Paterson AH (2004) A 3347-locus genetic recombination map of sequence-tagged sites reveals features of genome organization, transmission and evolution of cotton (Gossypium). Genetics 166:389–417 Rong J, Feltus EA, Waghmare VN, Pierce GJ, Chee PW, Draye X, Saranga Y, Wright RJ, Wilkins TA, May OL, Smith CW, Gannaway JR, Wendel JR, Paterson AH (2007) Metaanalysis of polyploid cotton QTL shows unequal contributions of subgenomes to a complex network of genes and gene clusters implicated in lint fiber development. Genetics 176(4):2577–2588 Rong J, Feltus FA, Liu L, Lin L, Paterson AH (2010) Gene copy number evolution during tetraploid cotton radiation. Heredity 105(5):463–472 Salmon A, Flagel L, Ying B, Udall JA, Wendel JF (2010) Homoeologous nonreciprocal recombination in polyploid cotton. New Phytol 186:123–134 Saunders JH (1961) The wild species of Gossypium and their evolutionary history. Oxford University Press, London Seelanan T, Brubaker CL, Stewart JM, Craven LA, Wendel JF (1999) Molecular systematics of Australian Gossypium section grandicalyx (Malvaceae). Syst Bot 24:183–208 Senchina DS, Alvarez I, Cronn RC, Liu B, Rong JK, Noyes RD, Paterson AH, Wing RA, Wilkins TA, Wendel JF (2003) Rate variation among nuclear genes and the age of polyploidy in Gossypium. Mol Biol Evol 20:633–643 Shan X, Liu Z, Dong Z, Wang Y, Chen Y, Lin X, Long L, Han F, Dong Y, Liu B (2005) Mobilization of the active MITE transposons mPing and Pong in rice by introgression from wild rice (Zizania latifolia Griseb.). Mol Biol Evol 22:976–990 Small RL, Ryburn JA, Wendel JF (1999) Low levels of nucleotide diversity at homoeologous Adh loci in allotetraploid cotton (Gossypium L.). Mol Biol Evol 16:491–501 Small RL, Wendel JF (1999) The mitochondrial genome of allotetraploid cotton (Gossypium L.). J Hered 90:251–253 Small RL, Wendel JF (2002) Differential evolutionary dynamics of duplicated paralogous Adh loci in allotetraploid cotton (Gossypium). Mol Biol Evol 19:597–607 Soltis PS, Soltis DE (2000) The role of genetic and genomic attributes in the success of polyploids. Proc Nat Acad Sci USA 97:7051–7057 CO RR 921 922 923 924 925 926 927 928 929 930 931 932 933 934 935 936 937 938 939 940 941 942 943 944 945 946 947 948 949 950 951 952 953 954 955 956 957 958 959 960 961 962 963 964 965 966 967 968 969 970 971 972 973 974 J. F. Wendel et al. UN Editor Proof 206 Layout: T1 Standard SC Chapter No.: 10 207 EC TE D PR OO F Stebbins GL (1947) Types of polyploids: their classification and significance. Adv Genet 1: 403–429 Stebbins GL (1950) Variation and evolution in plants. Columbia University Press, New York Stephens SG (1958) Salt water tolerance of seeds of Gossypium species as a possible factor in seed dispersal. Amer Nat 92:83–92 Stephens SG (1966) The potential for long range oceanic dispersal of cotton seeds. Amer Nat 100:199–210 Stewart JM (1995) Potential for crop improvement with exotic germplasm and genetic engineering. In: Constable GA, Forrester NW (eds) Challenging the future: proceedings of the world cotton research, CSIRO, Melbourne, pp 313–327 Stewart JM, Oosterhuis D, Heithholt JJ, Mauney JR (2010) Physiology of cotton. Springer, The Netherlands Udall JA, Swanson JM, Nettleton D, Percifield RJ, Wendel JF (2006) A novel approach for characterizing expression levels of genes duplicated by polyploidy. Genetics 173(3): 1823–1827 Ungerer MC, Strakosh SC, Zhen Y (2006) Genome expansion in three hybrid sunflower species is associated with retrotransposon proliferation. Curr Biol 16R:872–873 Veitia RA (2005) Gene dosage balance: deletions, duplications and dominance. Trends Genet 21:33 Wang JL, Tian L, Lee HS, Wei NE, Jiang HM, Watson B, Madlung A, Osborn TC, Doerge RW, Comai L, Chen ZJ (2006) Genomewide nonadditive gene regulation in Arabidopsis allotetraploids. Genetics 172:507–517 Watt G (1907) The wild and cultivated cotton plants of the world. Longmans, Green and Co, London Wendel JF (1989) New world tetraploid cottons contain old world cytoplasm. Proc Nat Acad Sci USA 86:4132–4136 Wendel JF (2000) Genome evolution in polyploids. Plant Mol Biol 42:225–249 Wendel JF, Brubaker CL, Alvarez JP, Cronn RC, Stewart JM (2009) Evolution and natural history of the cotton genus. In: Paterson AH (ed) Genomics of cotton, plant genetics and genomics, crops and models 3. Springer, New York, pp 3–22 Wendel JF, Cronn RC (2003) Polyploidy and the evolutionary history of cotton. Adv Agron 78:139–186 Wendel JF, Rowley R, Stewart JM (1994) Genetic diversity in and phylogenetic relationships of the Brazilian endemic cotton, Gossypium mustelinum (Malvaceae). Pl Syst Evol 192:49–59 Wendel JF, Schnabel A, Seelanan T (1995) Bidirectional interlocus concerted evolution following allopolyploid speciation in cotton (Gossypium). Proc Nat Acad Sci USA 92:280–284 Wright RJ, Thaxton PM, El-Zik KM, Paterson AH (1998) D-subgenome bias of Xcm resistance genes in tetraploid Gossypium (cotton) suggests that polyploid formation has created novel avenues for evolution. Genetics 149:1987–1996 Xu Z, Yu JZ, Cho J, Yu J, Kohel RJ, Percy RG (2010) Polyploidization altered gene functions in cotton (Gossypium spp.). PLoS ONE 5:e14351 Yang SS, Cheung F, Lee JJ, Ha M, Wei NE, Sze SH, Stelly DM, Thaxton P, Triplett B, Town CD, Chen ZJ (2006) Accumulation of genome-specific transcripts, transcription factors and phytohormonal regulators during early stages of fiber cell development in allotetraploid cotton. Plant J 47:761–775 CO RR 975 976 977 978 979 980 981 982 983 984 985 986 987 988 989 990 991 992 993 994 995 996 997 998 999 1000 1001 1002 1003 1004 1005 1006 1007 1008 1009 1010 1011 1012 1013 1014 1015 1016 1017 1018 1019 1020 Book ISBN: 978-3-642-31441-4 Page: 207/206 Jeans, Genes, and Genomes: Cotton as a Model for Studying Polyploidy UN Editor Proof 10 Book ID: 272454_1_En Date: 16-8-2012 Metadata of the chapter that will be visualized in SpringerLink Book Title Polyploidy and Genome Evolution Series Title Chapter Title Evolutionary Implications of Genome and Karyotype Restructuring in Nicotiana tabacum L Copyright Year 2012 Copyright HolderName Springer-Verlag Berlin Heidelberg Corresponding Author Family Name Leitch Particle Given Name Andrew Suffix Author Division School of Biological and Chemical Sciences Organization Queen Mary University of London Address London, UK Email a.r.leitch@qmul.ac.uk Family Name Kovarik Particle Given Name Ales Suffix Division Organization Institute of Biophysics Address Brno, Czech Republic Email Author Family Name Renny-Byfield Particle Given Name Simon Suffix Division School of Biological and Chemical Sciences Organization Queen Mary University of London Address London, UK Email Author Family Name Grandbastien Particle Given Name Marie-Angèle Suffix Division Organization Institute Jean-Pierre Bourgin Address Versailles, France Email Abstract Nicotiana tabacum is an allopolyploid that formed within the last 200,000 years from relatives of the extant diploids N. sylvestris and N. tomentosiformis, the donors of the S- and T-genomes, respectively. Here we review progress in our understanding of the divergence of N. tabacum subsequent to its formation, by comparing the N. tabacum genome with those of its diploid progenitors. We also review the data from synthetic N. tabacum, where there is evidence for much genetic change in early generations, including various chromosomal translocations, allopolyploid-induced retroelement mobility and loss, and reductions in the copy numbers of some tandem repeats. These observations are similar to patterns found in natural N. tabacum, suggesting that rapid genetic divergence is induced by allopolyploidy. The T-genome of N. tabacum shows the greatest number of genetic changes and appears to be less stable than the S-genome. We describe possible mechanisms that may have stimulated these genetic changes and propose that these can lead to enhanced fertility, more regular chromosome pairing, and the evolution of disomic inheritance. 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 F PR OO 5 Ales Kovarik, Simon Renny-Byfield, Marie-Angèle Grandbastien and Andrew Leitch Abstract Nicotiana tabacum is an allopolyploid that formed within the last 200,000 years from relatives of the extant diploids N. sylvestris and N. tomentosiformis, the donors of the S- and T-genomes, respectively. Here we review progress in our understanding of the divergence of N. tabacum subsequent to its formation, by comparing the N. tabacum genome with those of its diploid progenitors. We also review the data from synthetic N. tabacum, where there is evidence for much genetic change in early generations, including various chromosomal translocations, allopolyploid-induced retroelement mobility and loss, and reductions in the copy numbers of some tandem repeats. These observations are similar to patterns found in natural N. tabacum, suggesting that rapid genetic divergence is induced by allopolyploidy. The T-genome of N. tabacum shows the greatest number of genetic changes and appears to be less stable than the S-genome. We describe possible mechanisms that may have stimulated these genetic changes and propose that these can lead to enhanced fertility, more regular chromosome pairing, and the evolution of disomic inheritance. D 4 Evolutionary Implications of Genome and Karyotype Restructuring in Nicotiana tabacum L. TE 3 Chapter 11 EC 2 Book ISBN: 978-3-642-31441-4 Page: 209/224 CO RR 1 Book ID: 272454_1_En Date: 16-8-2012 A. Kovarik Institute of Biophysics, Brno, Czech Republic S. Renny-Byfield  A. Leitch (&) School of Biological and Chemical Sciences, Queen Mary University of London, London, UK e-mail: a.r.leitch@qmul.ac.uk UN Editor Proof Layout: T1 Standard SC Chapter No.: 11 M.-A. Grandbastien Institute Jean-Pierre Bourgin, Versailles, France P. S. Soltis and D. E. Soltis (eds.), Polyploidy and Genome Evolution, DOI: 10.1007/978-3-642-31442-1_11,  Springer-Verlag Berlin Heidelberg 2012 209 Layout: T1 Standard SC Chapter No.: 11 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 210/224 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 F PR OO 28 D 26 27 The genus Nicotiana is the sixth largest genus in the family Solanaceae and includes 76 species found naturally in the Americas, Australia and surrounding islands, and with a single species in Namibia (Goodspeed 1954; Knapp et al. 2004). However, several species, including N. tabacum (tobacco), now occur more widely. The genus has received much recent attention because of evidence for recurrent polyploidy, resulting in nearly half the species in the genus being chromosomally polyploid. It is likely there have been six independent polyploidy events within the genus: at about 10 million years ago (mya) (section Suaveolentes, c. 24 species), 5 mya (section Repandae, 4 species), 1–2 mya (section Polydicleae, 2 species), and within the last 200,000 years (N. rustica, N. arentsii, and N. tabacum) (Lim et al. 2004a). There is also reticulation at the diploid level, with at least three species likely to be of homoploid hybrid origin (N. glauca, N. linearis, and N. spegazzinii) (Goodspeed 1954; Kelly et al. 2010). N. tabacum is an autotetraploid (2n = 4x = 48) derived from diploid species that most closely resemble N. sylvestris (2n = 2x = 24), the maternal S-genome donor, and N. tomentosiformis (2n = 2x = 24, section Tomentosae), the paternal T-genome donor. Nicotiana sylvestris is the only species in section Sylvestres, and detailed genetic analysis of multiple accessions indicates that there is little genetic diversity in this species (4.2 % polymorphisms in AFLPs—Amplified Fragment Length Polymorphisms, Petit et al. 2007). However, there is considerably more genetic diversity among accessions of N. tomentosiformis (31.7 % AFLP polymorphism), which form two distinct groups, one of which most closely resembles the T-genome of N. tabacum (Murad et al. 2002; Petit et al. 2007). These data indicate that N. tabacum formed subsequent to the divergence of the two groups of N. tomentosiformis. Molecular clock estimates of plastid and internally transcribed spacer sequences (ITS) of nuclear ribosomal DNA (rDNA) suggest that N. tabacum formed less than 200,000 years ago. These estimates were calibrated using likely maximum ages of some endemic species occurring on islands of known geological ages (Clarkson et al. 2005; Leitch et al. 2008). However, given that only feral populations of N. tabacum have been found and no truly ‘‘wild’’ population exists (Knapp personal communication), it is quite possible that N. tabacum was born subsequent to the origin of human agriculture, within the last 10,000 years. The purpose of this paper is to review current understanding of the genetic consequences of polyploidy in N. tabacum. We show that genetic change can occur rapidly within a few generations, perhaps as a consequence of the ‘‘genomic shock’’ of allopolyploidy (McClintock 1984), a process that generates variants from which selection favors those with enhanced fertility. This unstable phase is presumably transient; however, there remains evidence for genome dynamism over time scales of thousands of years, including the replacement of multiple copies of rDNA sequences, the turnover of retroelements, and the loss of sequences targeted at the T-genome. We discuss how such changes may promote fertility and lead to the fixation of rearranged karyotypes during polyploid species establishment. TE 25 EC 24 11.1 Introduction to Nicotiana CO RR 23 A. Kovarik et al. UN Editor Proof 210 Layout: T1 Standard SC Chapter No.: 11 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 F PR OO 71 D 69 70 Kenton et al. (1993) were the first to use genomic in situ hybridization (GISH) to study the chromosomes of N. tabacum. The fluorescent probes derived from total genomic DNA of N. sylvestris and N. tomentosiformis hybridized to a separate subset of 24 chromosomes, corresponding to the chromosomes of the S- and T-genomes, respectively. However, the efficacy of the T-genome labeling was inferior (Lim et al. 2000b), likely because of a loss of T-genome sequences (Renny-Byfield et al. 2011). GISH also revealed in all N. tabacum accessions 4–9 intergenomic translocations (Kenton et al. 1993; Lim et al. 2004a; Moscone et al. 1996). Similar translocations (up to 3) were also observed in some synthetic N. tabacum lines that were only a few generations old (Skalicka et al. 2005). Perhaps these translocations have arisen as a consequence of multivalent formation (see Chap. 7), or they may represent hotspots of recombination. While some translocations are fixed (i.e. occur in all varieties), others are specific for particular accessions. There are more translocations of S-genome origin chromatin to T-genome chromosomes (T/s chromosome) than the reverse (S/t chromosome). Kenton et al. (1993) hypothesized that this could be caused by selection against S/t chromosomes. Alternatively, some of the ‘‘translocations’’ may actually be S-genome subtelomeric satellite sequences that now occur on T-genome chromosomes, perhaps arising in their new location via recombination-based homogenization processes (Koukalova et al. 2010). Next-generation sequencing (NGS) may enable us to distinguish among these alternative hypotheses. As with most other plant species (Heslop-Harrison and Schwarzacher 2011), a large fraction of Nicotiana genomes is composed of various types of retroelements. Genome sampling using NGS revealed that the major component of the genome of N. tabacum and its diploid relatives comprises long terminal repeat (LTR)-retrotransposons, with at least 17.1–22.5 % of the genome being Ty3/gypsy elements and 2.2–3.4 % being Ty1/copia elements. DNA transposons comprise around 1.7 % of the genome of these species (Renny-Byfield et al. 2011). In addition, there are several unrelated satellite repeat families that have been characterized (the distributions of some are shown Fig. 11.1, Lim et al. 2000b). These repeats include: (1) The HRS60 family, which is the best-characterized tandem repeat family in Nicotiana and includes an interstitial repeat in N. tomentosiformis and N. tabacum (called GRS, Gazdova et al. 1995) and predominantly subtelomeric repeats in N. sylvestris and N. tabacum (called HRS60 and NSYL2, Koukalova et al. 2010; Koukalova et al. 1989); (2) NTS9 in N. sylvestris and N. tabacum (Jakowitsch et al. 1998); (3) NTRS in N. tomentosiformis and N. tabacum (Fig. 11.1, Matyasek et al. 1997), and (4) NicCL3, a long 2.2 kb tandem repeat comprising *2 % of the N. tomentosiformis genome and in lower abundance in N. tabacum (Fig. 11.2, Renny-Byfield et al. 2011 and Renny-Byfield et al. 2012). There are also a number of satellite repeat families of known origin; these are: (5) A1/A2 satellite repeats derived from the intergenic spacer (IGS) of rDNA, which have transposed and amplified to multiple locations across the genome (Lim et al. 2004b); (6) tandem repeats of TE 68 EC 67 211 11.2 Nicotiana tabacum Genome Structure CO RR 66 Book ISBN: 978-3-642-31441-4 Page: 211/224 Evolutionary Implications of Genome UN Editor Proof 11 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 11 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 212/224 A. Kovarik et al. TE D PR OO F Editor Proof 212 111 112 113 114 115 116 117 118 119 120 geminivirus-related DNA (GRD), occurring in two distinct families in some accessions of N. tomentosiformis and inherited to N. tabacum (Murad et al. 2004; Murad et al. 2002), the first endogenous viruses discovered in plants (Bejarano et al. 1996); (7) sequences of pararetroviral origin (Matzke et al. 2004), including distinct variants found in N. tomentosiformis (NtoEPRV) and N. sylvestris (NsEPRV), both of which have been inherited in N. tabacum. CO RR 110 11.3 Retroelement Response to Allopolyploid ‘‘Genomic Shock’’ UN 109 EC Fig. 11.1 Ideograms showing the distribution of tandem repeat sequences in N. tabacum and its diploid progenitors, N. sylvestris and N. tomentosiformis. Taken with permission from Murad et al. (2002). See text for a description of the repeats shown Nicotiana hybrids were among the first in which chromosomal changes following hybridization were demonstrated. Gerstel and Burns (1967) reported that N. otophora 9N. tabacum hybrids have genetic instabilities of two kinds. Firstly, the heterochromatin from N. otophora undergoes breakage causing chromatin loss; this Layout: T1 Standard SC Chapter No.: 11 Book ISBN: 978-3-642-31441-4 Page: 213/224 Evolutionary Implications of Genome 213 TE D PR OO F Editor Proof 11 Book ID: 272454_1_En Date: 16-8-2012 121 122 123 124 125 UN CO RR EC Fig. 11.2 Reduction in copy number of T-genome repeats in N. tabacum from expectation given their abundance in N. tomentosiformis. a Fluorescence in situ hybridization (FISH) showing the distribution of dispersed A1/A2 repeats (green fluorescence). The 18-5.8-26S rDNA locus on chromosome 3 (N. tomentosiformis) and T3 (N. tabacum) is labeled with yellow fluorescence due to the overlap of red and green rDNA fluorescence. Note that there is much reduced A1/A2 signal in N. tabacum (for further information see Lim et al. 2004b). b FISH showing the distribution of NicCL3 in both species (green fluorescence). Note the reduced abundance of this repeat in N. tabacum. The 18-5.8-26S rDNA locus on chromosome 3 (N. tomentosiformis) and T3 (N. tabacum) is labeled with red fluorescence. For further information see Renny-Byfield et al. (2012). c, d. NGS analysis of repeat clusters in N. tabacum and its diploid progenitors. The graphs show the genome proportions (arcsine transformed) of repeats uniquely inherited in N. tabacum from c N. tomentosiformis and d N. sylvestris. If N. tabacum had faithfully inherited the repeats as found in the diploids, then all repeats would fall on the blue lines. However, in N. tomentosiformis c most repeats at higher genome proportions are underrepresented in N. tabacum. This is not the case for N. sylvestris d, where repeats are more or less in their expected abundance. This trend is reflected in the abundance of repeats that are biparentally inherited, i.e., the more abundant they are in N. tomentosiformis, the more likely they will be underrepresented in N. tabacum. For further information see Renny-Byfield et al. (2011) can be observed cytologically and also phenotypically by the leaf variegation it causes. Secondly, in some cells heterochromatic blocks from N. otophora proliferate enormously to many times their normal length, forming ‘‘megachromosomes’’. This pioneering work revealed that interspecific hybridization can induce chromosomal rearrangements and rapid sequence losses and gains. Layout: T1 Standard SC Chapter No.: 11 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 214/224 A. Kovarik et al. 157 11.4 Loss of DNA: Targeting the T-Genome 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 158 159 160 161 162 163 164 165 PR OO 132 D 131 TE 130 EC 128 129 CO RR 127 F 156 Melayah et al. (2001) were the first to show, using N. tabacum, that environmental stresses can induce retrotransposition. There is also evidence that allopolyploidy can stimulate transposition in N. tabacum and synthetic mimics of N. tabacum made from the diploid progenitors (Petit et al. 2007, 2010). These authors analyzed four populations of Ty1/copia LTR-retrotransposons (Tnt1-ol16, Tnt1-ol13, Tnt2, and Tto1) using sequence-specific amplification polymorphism (SSAP). They showed that the parental diploid species share essentially similar classes of retroelements. In N. tabacum there is evidence of retrotransposon diversification subsequent to allopolyploidy, with sequence losses concomitant with gains. Losses of retroelements were more frequent to the T-genome, while novel insertions of some populations, such as Tnt2, were shown to preferentially locate to the S-genome. However, each retrotransposon population seems to behave differently, with some populations undergoing rapid turnover, while others display relative stasis. There was little or no diversification of retrotransposons among individual accessions of natural N. tabacum, indicating that retroelement diversification occurred before the accessions analyzed had diverged, perhaps early in the species’evolution and arising as a consequence of allopolyploidy and induced ‘‘genomic shock’’ (McClintock 1984). To test the hypothesis that allopolyploidy can induce retroelement diversification, Petit et al. (2010) compared insertion patterns of the Tnt1 family in the synthetic N. tabacum Th37 (Burk 1973) with its diploid progenitors. In some Th37 individuals in the fourth synthetic generation (S4), there was evidence of new Tnt1 insertions. Newly, transposed copies were amplified from elements located on the N. sylvestrisderived genome and were highly similar to the Tnt1A tobacco copies amplified in response to microbial factors (Grandbastien et al. 2005). Furthermore, a high proportion of parental SSAP bands was lost in Th37, particularly those from the N. tomentosiformis-derived genome, again as observed in natural N. tabacum. Together, these data indicate that retrotransposon amplification and molecular restructuring in, or around, insertion sites occur rapidly in response to allopolyploidy, and that similar sequences have responded in the same way to allopolyploidy in natural and synthetic N. tabacum. 126 As with many other allopolyploid species (Leitch and Bennett 2004), N. tabacum has a lower genome size than is expected from the sum of parental genomes (Leitch et al. 2008, N. tabacum genome size, 1C = 5110 Mpb; N. sylvestris, 1C = 2636 Mbp; N. tomentosiformis, 1C = 2683 Mbp (http://data.kew.org/cvalues/). Recently, (Renny-Byfield et al. 2011) used NGS analysis of N. tabacum and its two diploid progenitors and showed that repeats derived from N. tomentosiformis was underrepresented in N. tabacum, a trend that was not observed for repeats from N. sylvestris (Fig. 11.2). These observations lead to the conclusion that the T-genome of UN Editor Proof 214 Layout: T1 Standard SC Chapter No.: 11 Book ISBN: 978-3-642-31441-4 Page: 215/224 Evolutionary Implications of Genome 215 189 11.5 Ribosomal DNA Homogenization is Rapid and Ongoing 174 175 176 177 178 179 180 181 182 183 184 185 186 187 190 191 192 193 194 195 196 197 198 199 200 201 202 203 204 205 206 PR OO 173 D 171 172 TE 169 170 EC 168 In most eukaryotes, 5S and 18–5.8–26S rDNA units occur in tandem arrays at one or several loci. Each large rDNA unit contains the 18S, 5.8S, and 26S ribosomal RNA (rRNA) genes, ITS, and IGS sequences. Whilst these sequences are vital for cell functioning, they can also be highly recombinogenic and labile sequences influencing genome stability (Kobayashi 2011). The genes themselves are highly conserved; however, their spacers diverge at suitable rates for resolving species relationships within most genera, including Nicotiana (Chase et al. 2003). Early studies revealed that the restriction fragment length polymorphism (RFLP) patterns of the N. tabacum IGS is not additive of the diploid progenitors (Kovarik et al. 1996), but that N. tabacum has evolved its own distinct rRNA gene family(ies). Sequence analysis revealed that the tobacco-specific units arose by reorganization of N. tomentosiformis-inherited units followed by their subsequent amplification (Volkov et al. 1999). The sequence changes mainly involved amplification and reduction of subrepeats upstream and downstream of the transcription start site within the IGS. However, the newly evolved units still occur at the four rDNA loci that N. tabacum inherited from its parents (see Fig. 11.1). Thus, it is likely that the parental units of S-genome origin were overwritten by the newly amplified CO RR 167 F 188 N. tabacum has undergone extensive sequence losses, and that genetic changes have occurred more rapidly to the paternally derived T-genome than the S-genome. N. tabacum harbors two families of endogenous pararetrovirus DNA, one from each progenitor (Jakowitsch et al. 1999). These are NsEPRV from N. sylvestris, which in N. tabacum is at the expected abundance, and NtoEPRV from N. tomentosiformis, which occurs in N. tabacum with reduced copy number. Likewise, the A1/A2 repeats (Lim et al. 2004b) and NicCL3 (Renny-Byfield et al. 2012), which are found in multiple locations around the genome in N. tomentosiformis (Fig. 11.2), are also underrepresented in N. tabacum. Similarly, NGS data show that all families of Ty3/gypsy elements have undergone sequence loss in N. tabacum (Renny-Byfield et al. 2011), and SSAP analysis of Tnt retroelements (Ty1/copia) reveals more deletion events influencing loci from the T-genome than the S-genome of N. tabacum (Petit et al. 2007). There is evidence of rapid and targetted loss of repeats from the T-genome even from the earliest generations, since in synthetic N. tabacum lines, many individuals already have reduced copy numbers of NtoEPRV, NTRS repeats (Skalicka et al. 2005), and NicCL3 (Renny-Byfield et al. 2012), whilst most losses of Tnt1 retroelements were from the T-genome (Petit et al. 2010). All these data suggest rapid, targetted changes early in allopolyploid evolution. The targeting of repeat losses from the T-genome may support the nuclear-cytoplasmic interaction hypothesis, whereby the paternally derived genome functions in a maternal cytoplasm (Gill 1991; Leitch et al. 2006), which in allopolyploids may lead to incompatibilities that preferentially destabilize the paternally derived genome. 166 UN Editor Proof 11 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 11 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 216/224 A. Kovarik et al. 211 N.tabacum-specific units (Lim et al. 2000a), although a few units of S genomeorigin remain intact, perhaps because they are methylated and inactive (Kovarik et al. 2008). In contrast, the 5S rDNA follows a different evolutionary pattern in N. tabacum, with no evidence for intergenomic homogenization, although shifts in copy numbers of parental gene families have been observed (Fulnecek et al. 2002). 212 11.6 The Fate of Duplicated Genes 207 208 209 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 236 237 238 239 240 241 242 243 244 245 246 247 D 216 217 TE 215 Although there has been substantial restructuring of the repetitive fraction of the N. tabacum genome since its formation, most genic sequences analyzed have remained in duplicate copies, e.g., genes for drug resistance (Schenke et al. 2003), putrescine N-methytransferases (Riechers and Timko 1999), a family of small GTP-binding proteins (Takumi et al. 2002), lignin forming peroxidase (Matassi et al. 1991), nitrite reductase (Kronenberger et al. 1993), nitrate reductase (Vaucheret et al. 1989), glutamine synthetase (Clarkson et al. 2010), phytochrome A (Intrieri et al. 2008), the developmental gene LEAFY (McCarthy 2010), and families of DNA methyltransferases (Fulnecek et al. 2009). The only current exception is a family of N. tabacum glucan endo-1,3-beta-glucosidase genes that appeared to be recombinants of both ancestral sequences (Sperisen et al. 1991). Locus additivity does not always seem to be a rule even in recently formed synthetic allopolyploids. For example, frequent deletions of homoeologs were found among 70 protein-coding loci in Tragopogon miscellus that formed within the last 80 years (Buggs et al. 2009, 2012; see also Chap. 14, this volume). These changes may not be random, as Buggs et al. (2012) showed that clusters of genes are repeatedly lost or retained and the likelihood of retention reflects gene ontology categories or their predicted levels of dosage sensitivity. Similarly, synthetic lines of Brassica napus seem to eliminate much of the parental DNA, although it is currently unknown whether coding or noncoding sequences are preferentially targeted (Song et al. 1995; Szadkowski et al. 2011). The occurrence of both homoeologs in N. tabacum may reflect limited sampling, or high levels of retention. Additivity is observed in synthetic lines of Gossypium (Liu et al. 2002; see also Chap. 10, this volume) and in synthetic wheat allopolyploids, although in the latter case there remains some controversy (Mestiri et al. 2010; Feldman and Levy 2009; see also Chap. 7, this volume). Additivity has also been demonstrated at the expression level. Alleles of phytochrome A, lignin forming peroxidases, nitrite reductase, and DNA methyltransferases are expressed from both parental homeologs. There are also reports of epigenetic silencing of one subset of the parental alleles, including rRNA genes that have escaped homogenization (Kovarik et al. 2008) and the CYP82E4 locus (Chakrabarti et al. 2007) involved in the alkaloid biosynthesis pathway. Duplicate copies arising through polyploidy can be retained through a number of mechanisms (Doyle et al. 2008): (1) one copy can evolve a new function (neofunctionalization) that can become fixed through a selective advantage EC 214 CO RR 213 PR OO F 210 UN Editor Proof 216 Layout: T1 Standard SC Chapter No.: 11 Book ISBN: 978-3-642-31441-4 Page: 217/224 Evolutionary Implications of Genome 217 259 11.7 Genome Revolution and Allopolyploid Establishment 255 256 257 260 261 262 263 264 265 266 267 268 269 270 271 272 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 PR OO 254 Until the early 1990s, it was commonly thought that chromosomal changes establish isolation barriers among populations because of reduced fertility in heterozygotes. Rearranged karyotypes may, then, become fixed through inbreeding or meiotic drive, particularly in small populations (King 1993). However, it is difficult to distinguish between karyotype changes that drive speciation events and those arising by genetic drift after speciation. Certainly, many rearrangements are deleterious and can cause embryo lethality or hybrid breakdown prior to reproduction (Rieseberg 1997). Only neutral or advantageous recombinants will pass the bottleneck of selection during the early phase of speciation. Furthermore, evolutionary geneticists pointed out that genetic mutations at the DNA level are more common than karyotype change and are therefore likely to be more important in speciation processes (Butlin 1993). However, recent evidence from a number of sources, e.g., Anopheles gambiae (Turner and Hahn 2010) and Helianthus homoploid hybrids (Strasburg et al. 2009), indicates that chromosomal inversions can, indeed, have a role in establishing barriers between populations and drive speciation. Strasburg et al. (2009) suggested that inversions may become foci for the accumulation of adaptive genes, particularly at the junctions between collinear and inverted regions, where gene flow is likely to be most impeded. Likewise, computer modeling has shown that when recombination is eliminated at an inverted region, and in the presence of strong selection and minimal or no gene flow, then species isolation can be driven by such a rearrangement (Feder and Nosil 2009). In the context of early polyploid evolution, karyotype rearrangement may drive enhanced fertility and potentially be fundamental to allopolyploid species establishment. Certainly, high levels of chromosomal change have been observed in young polyploids, e.g., karyotype variability (dysploidy and intergenomic translocations) within and among populations of Tragopogon polyploids that formed within the last 80 years (Chester et al. 2010, 2012; Lim et al. 2008), and largescale chromosomal deletions occurring in individual plants of early generation D 253 TE 252 EC 251 CO RR 250 F 258 (Lynch et al. 2001); (2) duplicate copies can diverge through complementary loss of function at a particular point in the development or in particular tissues (subfunctionalization), a process that can occur through the action of drift alone; and (3) selection can occur against gene loss because that loss would compromise appropriate levels of gene product in relation to another gene(s) (gene balance hypothesis, cf. Birchler and Veitia 2010). In N. tabacum, a mechanism termed ‘‘nonfunctionalization’’ has been proposed which considers a combination of degenerative mutation and epigenetic silencing (Chakrabarti et al. 2007). Here, ‘‘non-functionalization’’ has altered alkaloid metabolism through reduced conversion of nicotine to nornicotine. Perhaps this mutation has been important to the ‘‘success’’ of N. tabacum as a recreational drug. 248 249 UN Editor Proof 11 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 11 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 218/224 296 297 298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 F 295 • Reduced chiasma frequency. Selection may favor the formation of fewer and/or focussed chiasma at recombination hotspots. This would result in quadrivalents formed in prophase I falling into two bivalents in metaphase I and then segregating normally (Fig. 11.3a). Computer modeling has shown that reduced chiasma frequency results in the evolution of disomic inheritance (Le Comber et al. 2010), although we are unaware of empirical evidence to support this assertion. • Structural and epigenetic divergence of homeologues. Allopolyploidy can trigger extensive karyotype and molecular restructuring in early generations (Gaeta et al. 2007; Renny-Byfield et al. 2011). This may result in genetic divergence of the parental genomes, e.g., through the loss or rearrangement of sequences shared by homeologues and recognized by the homolog recognition machinery in meiosis (Fig. 11.3b). Such sequences are unlikely to be those that are highly repeated across the genome since they are not chromosome specific. Nevertheless, genetic changes may introduce regions among homeologues that do not recombine, forming ‘‘islands of divergence’’, as in the establishment of homoploid hybrid species (Strasburg et al. 2009). Without recombination, the local regions around these islands will continue to diverge, leading to regular homolog pairing, and bivalent formation. Similarly, allopolyploidy can trigger epigenetic changes across the genome (Parisod et al. 2009). Epigenetic changes, which are thought to alter patterns of ectopic recombination (Colot et al. 1995), may also influence patterns of meiotic recombination and initiate islands of divergence. In addition, newly formed epi-alleles may establish tissue or temporal patterns of gene expression (subfunctionalization, Adams and Wendel 2005). Subfunctionalization is likely to favor the evolution of regular disomic inheritance and to select against multivalent formation. This is because multivalents result in the inappropriate segregation of homeoalleles and the generation of individuals with reduced fitness, because they have aberrant patterns of gene expression (Le Comber et al. 2010). • Intergenomic translocations. In the context of a diploid, translocations among heterologues can promote quadrivalent formation, leading to chromosome loss, and reduced fertility. Similarly, in the context of polyploids, translocations and aneuploidy are associated with low fertility, as observed in synthetic B. napus lines (Xiong et al. 2011). However, it is possible that a more complex dynamic can occur (Fig. 11.3c). If homeologues already form multivalents, then it is conceivable that the gain of genetic material from a heterologous chromosome may promote bivalent formation, disomic inheritance, and increased fertility. Alternatively, complex rearrangements can result through cascades of induced PR OO 294 D 293 TE 292 EC 291 synthetic allopolyploids of B. napus (Gaeta et al. 2007). Typically, a major hurdle that a newly formed polyploid must overcome is reduced fertility, often arising through multivalent formation, where chromosomes can segregate aberrantly leading to aneuploidy and reduction in fertility. We can envisage several allopolyploid-induced processes that may act to reduce multivalent formation and be favored by selection, these are: CO RR 289 290 A. Kovarik et al. UN Editor Proof 218 Layout: T1 Standard SC Chapter No.: 11 Editor Proof 11 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 219/224 Evolutionary Implications of Genome 219 (a) Reduced chiasma frequency X Focused / fewer chiasma may result in elevated bivalent formation -enhancing fertility X F X X PR OO (b) Structural divergence of homologues X X X X X X X X (c) Intergenomic translocations X X X X X X X X X X Rapid homeologue divergence may drive bivalent formation, enhancing fertility Promotes homologue pairing, disomic inheritance and homologue divergence D X TE X X X X X X EC X Homeologous translocations promote multivalent formation, generating further translocations. Meiotic abnormalities stimulate unequal reciprocal translocations, inversions, end-to-end chromosome fusions, resulting in dysploidy . 334 335 336 337 338 340 339 341 342 343 344 pairing problems resulting from recurrent multivalent formation (unequal reciprocal translocations, inversions, end-to-end chromosome fusions), potentially generating novel chromosomes with segments from multiple ancestral chromosomes. Such a process may be associated with a reduction in chromosome number to a diploid-like level, as reported in the divergence of Arabidopsis karyotypes (Mandakova et al. 2010). UN 333 CO RR Fig. 11. 3 Fast molecular and chromosome evolution may enhance bivalent formation, driving disomic inheritance, and fertility in young polyploids. Homologs are shown in the same color (pink or pale green). Homeologues diverge over time, this is reflected in their colors becoming more distinct (red and green). A translocation from a heterologue is shown in blue. Crossovers among chromosomes are shown with crosses From Nicotiana allopolyploids, we have evidence for several of the proposed modes of chromosome evolution. First, in synthetic N. tabacum, we have observed plants that were homozygous for an intergenomic translocation that is similar to translocations observed in natural N. tabacum (Lim et al. 2004a). Perhaps this translocation is selected because it enhances fertility, and that in the synthetic Layout: T1 Standard SC Chapter No.: 11 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 220/224 347 348 349 350 351 352 353 354 355 356 357 358 N. tabacum evolution is repeating itself. Second, the allotetraploids of sections Polydicliae and Repandae, which formed c. 1 mya and 5 mya, respectively, have chromosome numbers that are additive of the progenitor diploids (2n = 4x = 48). However, the sequence organization along their chromosomes has diverged considerably, particularly in the latter. Indeed, Lim et al. (2007) suggested that in some species of section Repandae there has been near complete genome turnover. Perhaps these changes contributed to the establishment of disomic inheritance. Finally, in section Suaveolentes, several species show evidence of chromosome number reduction from expectation (chromosome numbers range from 2n = 4x = 32 to 46, depending on species, Goodspeed 1954; Knapp et al. 2004). We anticipate that this reduction is associated with karyotype restructuring, as in the polyploids of Arabidopsis (Mandakova et al. 2010). F 346 PR OO 345 A. Kovarik et al. 11.8 Advantages of the Nicotiana System and Future Perspectives 369 Acknowledgments We thank NERC and the Czech Science Foundation for support. 370 References 371 372 373 374 375 376 377 378 379 380 Adams KL, Wendel JF (2005) Allele-specific, bidirectional silencing of an alcohol dehydrogenase gene in different organs of interspecific diploid cotton hybrids. Genetics 171:2139–2142 Bejarano ER, Khashoggi A, Witty M, Lichtenstein C (1996) Integration of multiple repeats of geminiviral DNA into the nuclear genome of tobacco during evolution. Proc Nat Acad Sci USA 93:759–764 Birchler JA, Veitia RA (2010) The gene balance hypothesis: implications for gene regulation, quantitative traits and evolution. New Phytol 186:54–62 Buggs RJ, Doust AN, Tate JA, Koh J, Soltis K, Feltus FA, Paterson AH, Soltis PS, Soltis DE (2009) Gene loss and silencing in Tragopogon miscellus (Asteraceae): comparison of natural and synthetic allotetraploids. Heredity 103:73–81 364 365 366 367 TE 363 EC 361 362 CO RR 360 D 368 Nicotiana provides many opportunities for studying allopolyploid genome divergence and addressing key questions facing polyploidy researchers [outlined in Soltis et al. (2010)]. It is important, even vital for the future of humankind, to know how allopolyploid genomes interact together. This is because many of our most important crop species are recognizably polyploid based on their chromosome numbers, and recent evidence is emerging that all seed-bearing plants have undergone at least one round of polyploidy in their ancestry (Jiao et al. 2011). The allopolyploid species of Nicotiana and the synthetic allopolyploids provide a unique opportunity to study snap shots of polyploid divergence over 10 million years of evolution. 359 UN Editor Proof 220 Layout: T1 Standard SC Chapter No.: 11 221 EC TE D PR OO F Buggs RJA, Chamala S, Wu W, Tate JA, Schnable PS, Soltis DE, Soltis PS, Barbazuk WB (2012) Rapid, repeated, and clustered loss of duplicate genes in allopolyploid plant populations of independent origin. Curr Biol 22:248–252 Burk LG (1973) Partial self-fertility in a theoretical amphiploid progenitor of N. tabacum. J Hered 64:348–350 Butlin RK (1993) Species evolution—the role of chromosome change—King, M. Nature 366:27 Chakrabarti M, Meekins KM, Gavilano LB, Siminszky B (2007) Inactivation of the cytochrome P450 gene CYP82E2 by degenerative mutations was a key event in the evolution of the alkaloid profile of modern tobacco. New Phytol 175:565–574 Chase MW, Knapp S, Cox AV, Clarkson JJ, Butsko Y, Joseph J, Savolainen V, Parokonny AS (2003) Molecular systematics, GISH and the origin of hybrid taxa in Nicotiana (Solanaceae). Ann Bot 92:107–127 Chester M, Gallagher JP, Symonds VV, Cruz da Silva AV, Mavrodiev EV, Leitch AR, Soltis PS and Soltis DE (2012) Extensive chromosomal variation in a recently formed natural allopolyploid species, Tragopogon miscellus (Asteraceae). Proceedings of the National Academy of Sciences of the United States of America, 109:1176-1181 Chester M, Sykorova E, Fajkus J, Leitch AR (2010) Single integration and spread of a Copia-like sequence nested in rDNA intergenic spacers of Allium cernuum (Alliaceae). Cytogenet Genome Res 129:35–46 Clarkson JJ, Kelly LJ, Leitch AR, Knapp S, Chase MW (2010) Nuclear glutamine synthetase evolution in Nicotiana: phylogenetics and the origins of allotetraploid and homoploid (diploid) hybrids. Mol Phylogenet Evol 55:99–112 Clarkson JJ, Lim KY, Kovarik A, Chase MW, Knapp S, Leitch AR (2005) Long-term genome diploidization in allopolyploid Nicotiana section Repandae (Solanaceae). New Phytol 168:241–252 Colot V, Goyon C, Faugeron G, Rossignol JL (1995) Methylation of repeated DNA sequences and genome stability in Ascobolus immersus. Can J Bot-Rev Canadienne De Botanique 73:S221–S225 Doyle JJ, Flagel LE, Paterson AH, Rapp RA, Soltis DE, Soltis PS, Wendel JF (2008) Evolutionary genetics of genome merger and doubling in plants. Ann Rev Genet 443–461 Feder JL, Nosil P (2009) Chromosomal inversions and species differences: when are geens affecting adaptive divergence and reproductive isolation expected to reside within inversions? Evolution 63:3061–3075 Feldman M, Levy AA (2009) Genome evolution in allopolyploid wheat-a revolutionary reprogramming followed by gradual changes. J Genet Genomics 36:511–518 Fulnecek J, Lim KY, Leitch AR, Kovarik A, Matyasek R (2002) Evolution and structure of 5S rDNA loci in allotetraploid Nicotiana tabacum and its putative parental species. Heredity 88:19–25 Fulnecek J, Matyasek R, Kovarik A (2009) Faithful inheritance of cytosine methylation patterns in repeated sequences of the allotetraploid tobacco correlates with the expression of DNA methyltransferase gene families from both parental genomes. Mol Genet Genomics 281:407–420 Gaeta RT, Pires JC, Iniguez-Luy F, Leon E, Osborn TC (2007) Genomic changes in resynthesized Brassica napus and their effect on gene expression and phenotype. Plant Cell 19:3403–3417 Gazdova B, Siroky J, Fajkus J, Brzobohaty B, Kenton A, Parokonny A, Heslop-Harrison JS, Palme K, Bezdek M (1995) Characterization of a new family of tobacco highly repetitive DNA, GRS, specific for the Nicotiana tomentosiformis genomic component. Chromosome Res 3:245–254 Gerstel DU, Burns JA (1967) Phenotypic and chromosomal abnormalities associated with the introduction of heterochromatin from Nicotiana otophora into N. tabacum. Genetics 56:483–502 Gill BS (1991) Nucleocytoplasmic interaction (NCI) hypothesis of genome evolution and speciation in polyploid plants. In: Sasakuma T, KinoshitaT (ed.) Proceedings of the Kihara CO RR 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 Book ISBN: 978-3-642-31441-4 Page: 221/224 Evolutionary Implications of Genome UN Editor Proof 11 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 11 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 222/224 EC TE D PR OO F memorial international symposium on cytoplasmic engineering in wheat, Kihara Memorial Foundation, Yokohama, Japan, pp 48–53 Goodspeed TH (1954) The genus Nicotiana Massachusetts. Chronica Botanica Company, USA Grandbastien M, Audeon C, Bonnivard E, Casacuberta JM, Chalhoub B, Costa APP, Le QH, Melayah D, Petit M, Poncet C, Tam SM, Van Sluys MA, Mhiri C (2005) Stress activation and genomic impact of Tnt1 retrotransposons in Solanaceae. Cytogenetic Genome Res 110:229–241 Heslop-Harrison JS, Schwarzacher T (2011) Organisation of the plant genome in chromosomes. Plant J 66:18–33 Intrieri MC, Muleo R, Buiatti M (2008) Phytochrome A as a functional marker of phyletic relationships in Nicotiana genus. Biol Plant 52:36–41 Jakowitsch J, Mette MF, van der Winden J, Matzke MA, Matzke AJM (1999) Integrated pararetroviral sequences define a unique class of dispersed repetitive DNA in plants. Proc Nat Acad Sci USA 96:13241–13246 Jakowitsch J, Papp I, Matzke MA, Matzke AJM (1998) Identification of a new family of highly repetitive DNA, NTS9, that is located predominantly on the S9 chromosome of tobacco. Chromosome Res 6:649–659 Jiao YN, Wickett NJ, Ayyampalayam S, Chanderbali AS, Landherr L, Ralph PE, Tomsho LP, Hu Y, Liang HY, Soltis PS, Soltis DE, Clifton SW, Schlarbaum SE, Schuster SC, Ma H, LeebensMack J, dePamphilis CW (2011) Ancestral polyploidy in seed plants and angiosperms. Nature 473:97–113 Kelly LJ, Leitch AR, Clarkson JJ, Hunter RB, Knapp S, Chase MW (2010) Intragenic recombination events and evidence for hybrid speciation in Nicotiana (Solanaceae). Mol Biol Evol 27:781–799 Kenton A, Parokonny AS, Gleba YY, Bennett MD (1993) Characterization of the Nicotiana tabacum L. genome by molecular cytogenetics. Mol Gen Genet 240:159–169 King M (1993) Species evolution the role of chromosome change. Cambridge University Press, Cambridge Knapp S, Chase MW, Clarkson JJ (2004) Nomenclatural changes and a new sectional classification in Nicotiana (Solanaceae). Taxon 53:73–82 Kobayashi T (2011) Regulation of ribosomal RNA gene copy number and its role in modulating genome integrity and evolutionary adaptability in yeast. Cell Mol Life Sci 68:1395–1403 Koukalova B, Moraes AP, Renny-Byfield S, Matyasek R, Leitch AR, Kovarik A (2010) Fall and rise of satellite repeats in allopolyploids of Nicotiana over c. 5 million years. New Phytol 186:148–160 Koukalova B, Reich J, Matyasek R, Kuhrova V, Bezdek M (1989) A BamHI family of highly repeated DNA sequecnes of Nicotiana tabacum. Theor Appl Genet 78:77–80 Kovarik A, Dadejova M, Lim K, Chase M, JJ C, Knapp S, Leitch A (2008) Evolution of rDNA in Nicotiana allopolyploids: a potential link between rDNA homogenization and epigentics. Ann Bot 101:815–823 Kovarik A, Fajkus J, Koukalova B, Bezdek M (1996) Species-specific evolution of telomeric and rDNA repeats in the tobacco composite genome. Theor Appl Genet 92:1108–1111 Kronenberger J, Lepingle A, Caboche M, Vaucheret H (1993) Cloning and expession of distinct nitrite reductases in tobacco-leaves and roots. Mol Gen Genet 236:203–208 Le Comber SC, Ainouche ML, Kovarik A, Leitch AR (2010) Making a functional diploid: from polysomic to disomic inheritance. New Phytol 186:113–122 Leitch AR, Lim KY, Skalicka K, Kovarik A (2006) Nuclear cytoplasmic interaction hypothesis and the role of translocations in Nicotiana allopolyploids. In: Cigna AAD, M. Yerevan (eds.) Radiation risk estimates in normal and emergency situations book series: NATO security through science series B: Physics and Biophysics (ARMENIA) pp 319–326 Leitch IJ, Bennett MD (2004) Genome downsizing in polyploid plants. Biol J Linn Soc 82:651–663 CO RR 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 A. Kovarik et al. UN Editor Proof 222 Layout: T1 Standard SC Chapter No.: 11 223 EC TE D PR OO F Leitch IJ, Hanson L, Lim KY, Kovarik A, Chase MW, Clarkson JJ, Leitch AR (2008) The ups and downs of genome size evolution in polyploid species of Nicotiana (Solanaceae). Ann Bot 101:805–814 Lim KY, Kovarik A, Matyasek R, Bezdek M, Lichtenstein CP, Leitch AR (2000a) Gene conversion of ribosomal DNA in Nicotiana tabacum is associated with undermethylated, decondensed and probably active gene units. Chromosoma 109:161–172 Lim KY, Kovarik A, Matyasek R, Chase MW, Clarkson JJ, Grandbastien MA, Leitch AR (2007) Sequence of events leading to near-complete genome turnover in allopolyploid Nicotiana within five million years. New Phytol 175:756–763 Lim KY, Matyasek R, Kovarik A, Leitch AR (2004a) Genome evolution in allotetraploid Nicotiana. Biol J Linn Soc 82:599–606 Lim KY, Matyasek R, Lichtenstein CP, Leitch AR (2000b) Molecular cytogenetic analyses and phylogenetic studies in the Nicotiana section Tomentosae. Chromosoma 109:245–258 Lim KY, Skalicka K, Koukalova B, Volkov RA, Matyasek R, Hemleben V, Leitch AR, Kovarik A (2004b) Dynamic changes in the distribution of a satellite homologous to intergenic 26-18S rDNA spacer in the evolution of Nicotiana. Genetics 166:1935–1946 Lim KY, Soltis DE, Soltis PS, Tate J, Matyasek R, Srubarova H, Kovarik A, Pires JC, Xiong ZY, Leitch AR (2008) Rapid chromosome evolution in recently formed polyploids in Tragopogon (Asteraceae). Plos One, 3 Liu B, Brubaker CL, Mergeai G, Cronn RC, Wendel JF (2002) Polyploid formation in cotton is not accompanied by rapid genomic changes. Genome 44:321–330 Lynch M, O’Hely M, Walsh B, Force A (2001) The probability of preservation of a newly arisen gene duplicate. Genetics 159:1789–1804 Mandakova T, Joly S, Krzywinski M, Mummenhoff K, Lysak MA (2010) Fast diploidization in close mesopolyploid relatives of Arabidopsis. Plant Cell 22:2277–2290 Matassi G, Melis R, Macaya G, Bernardi G (1991) Compositional bimodality of the nuclear genome of tobacco. Nucleic Acids Res 19:5561–5567 Matyasek R, Gazdova B, Fajkus J, Bezdek M (1997) NTRS, a new family of highly repetitive DNAs specific for the T1 chromosome of tobacco. Chromosoma 106:369–379 Matzke M, Gregor W, Mette MF, Aufsatz W, Kanno T, Jakowitsch J, Matzke AJM (2004) Endogenous paragretroviruses of polyploid Nicotiana tabacum and its diploid progenitors, N. sylvestris and N. tomentosiformis. Biol J Linn Soc 82:627–638 McCarthy EW (2010) The effect of molecular and spectral reflectance evolution on Nicotiana polyploids of different ages. In School of Biological and Chemical Sciences. London: Ph.D. thesis from Queen Mary University of London, p 259 McClintock B (1984) The significance of responses of the genome to challenge. Science 226:792–801 Melayah D, Bonnivard E, Chalhoub B, Audeon C, Grandbastien MA (2001) The mobility of the tobacco Tnt1 retrotransposon correlates with its transcriptional activation by fungal factors. Plant Journal, 28:159-168 Mestiri I, Chague V, Tanguy AM, Huneau C, Huteau V, Belcram H, Coriton O, Chalhoub B, Jahier J (2010) Newly synthesized wheat allohexaploids display progenitor-dependent meiotic stability and aneuploidy but structural genomic additivity. New Phytol 186:101–186 Moscone EA, Matzke MA, Matzke AJM (1996) The use of combined FISH/GISH in conjunction with DAPI counterstaining to identify chromosomes containing transgene inserts in amphidiploid tobacco. Chromosoma 105:231–236 Murad L, Bielawski JP, Matyasek R, Kovarik A, Nichols RA, Leitch AR, Lichtenstein CP (2004) The origin and evolution of Geminivirus-related DNA sequences in Nicotiana. Heredity 92:352–358 Murad L, Lim KY, Christopodulou V, Matyasek R, Lichtenstein CP, Kovarik A, Leitch AR (2002) The origin of tobacco’s T genome is traced to a particular lineage within Nicotiana tomentosiformis (Solanaceae). Am J Bot 89:921–928 CO RR 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 Book ISBN: 978-3-642-31441-4 Page: 223/224 Evolutionary Implications of Genome UN Editor Proof 11 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 11 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 224/224 EC TE D PR OO F Parisod C, Salmon A, Zerjal T, Tenaillon M, Grandbastien MA, Ainouche M (2009) Rapid structural and epigenetic reorganization near transposable elements in hybrid and allopolyploid genomes in Spartina. New Phytol 184:1003–1015 Petit M, Guidat C, Daniel J, Denis E, Montoriol E, Bui QT, Lim KY, Kovarik A, Leitch AR, Grandbastien MA, Mhiri C (2010) Mobilization of retrotransposons in synthetic allotetraploid tobacco. New Phytol 186:135–147 Petit M, Lim KY, Julio E, Poncet C, de Borne FD, Kovarik A, Leitch AR, Grandbastien MA, Mhiri C (2007) Differential impact of retrotransposon populations on the genome of allotetraploid tobacco (Nicotiana tabacum). Mol Genet Genomics 278:1–15 Renny-Byfield S, Chester M, Kovarik A, Le Comber S, Grandbastien M, Deloger M, Nichols R, Macas J, Novak P, Chase M, Leitch A (2011) Next generation sequencing reveals genome downsizing in allotetraploid Nicotiana tabacum, predominantly through the elimination of paternally derived repetitive DNAs Molecular Biology and Evolution 28:2843–2854 Renny-Byfield S, Kovarik A, Chester M, Nichols RA, Macas J, Novak P and Leitch AR (2012) Independent, rapid and targeted loss of highly repetitive DNA in natural and synthetic allopolyploids of Nicotiana tabacum. Plos One, 7:e36963 Riechers DE, Timko MP (1999) Structure and expression of the gene family encoding putrescine N-methyltransferase in Nicotiana tabacum: new clues to the evolutionary origin of cultivated tobacco. Plant Mol Biol 41:387–401 Rieseberg LH (1997) Hybrid origins of plant species. Annu Rev Ecol Syst 28:359–389 Schenke D, Sasabe M, Toyoda K, Inagaki Y, Shiraishi T, Ichinose Y (2003) Genomic structure of the NtPDR1 gene, harboring the two miniature inverted-repeat transposable elements, NtToya1 and NtStowaway101. Genes Genet Syst 78:409–418 Skalicka K, Lim KY, Matyasek R, Matzke M, Leitch AR, Kovarik A (2005) Preferential elimination of repeated DNA sequences from the paternal, Nicotiana tomentosiformis genome donor of a synthetic, allotetraploid tobacco. New Phytol 166:291–303 Soltis DE, Buggs RJA, Doyle JJ, Soltis PS (2010) What we still don’t know about polyploidy. Taxon 59:1387–1403 Song KM, Lu P, Tang KL and Osborn TC (1995) Rapid genome change in synthetic polyploids of Brassica and its implications for polyploid evolution. Proceedings of the National Academy of Sciences of the United States of America, 92:7719-7723 Sperisen C, Ryals J, Meins F (1991) Comparison of cloned genes provides evidence for intergenomic exchange of DNA in the evolution of a tobacco glucan endo-1,3 betaglucosidase gene family. Proc Nat Acad Sci USA 88:1820–1824 Strasburg JL, Scotti-Saintagne C, Scotti I, Lai Z, Rieseberg LH (2009) Genomic patterns of adaptive divergence between chromosomally differentiated sunflower species. Mol Biol Evol 26:1341–1355 Szadkowski E, Eber F, Huteau V, Lode M, Coriton O, Jenczewski E and Chevre AM (2011) Polyploid formation pathways have an impact on genetic rearrangements in resynthesized Brassica napus. New Phytologist, 191:884-894 Takumi S, Ida M, Haisa Y, Ando S, Nakamura C (2002) Genomic structure and homoeologous relationship of the two alpha-subunit genes of a heterotrimeric GTP-binding protein in tobacco. Genome 45:626–633 Turner TL, Hahn MW (2010) Genomic islands of speciation or genomic islands and speciation? Mol Ecol 19:848–850 Vaucheret H, Vincentz M, Kronenberger J, Caboche M, Rouze P (1989) Molecular cloning and characterization of the 2 homologous genes coding for nitrate reductase in tobacco. Mol Gen Genet 216:10–15 Volkov RA, Borisjuk NV, Panchuk II, Schweizer D, Hemleben V (1999) Elimination and rearrangement of parental rDNA in the allotetraploid Nicotiana tabacum. Mol Biol Evol 16:311–320 Xiong ZY, Gaeta RT, Pires JC (2011) Homoeologous shuffling and chromosome compensation maintain genome balance in resynthesized allopolyploid Brassica napus. Proc Nat Acad Sci USA 108:7908–7913 CO RR 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 A. Kovarik et al. UN Editor Proof 224 Metadata of the chapter that will be visualized in SpringerLink Book Title Polyploidy and Genome Evolution Series Title Chapter Title Polyploid Evolution in Spartina: Dealing with Highly Redundant Hybrid Genomes Copyright Year 2012 Copyright HolderName Springer-Verlag Berlin Heidelberg Corresponding Author Family Name Ainouche Particle Given Name M. Suffix Author Division University of Rennes 1 Organization UMR CNRS 6553 Ecobio Address Bât. 14A, Campus Scientifique de Beaulieu, 35042, Rennes Cedex, France Email malika.ainouche@univ-rennes1.fr Family Name Chelaifa Particle Given Name H. Suffix Division University of Rennes 1 Organization UMR CNRS 6553 Ecobio Address Bât. 14A, Campus Scientifique de Beaulieu, 35042, Rennes Cedex, France Email Author Family Name Ferreira Particle Given Name J. Suffix Division University of Rennes 1 Organization UMR CNRS 6553 Ecobio Address Bât. 14A, Campus Scientifique de Beaulieu, 35042, Rennes Cedex, France Email Author Family Name Bellot Particle Given Name S. Suffix Division University of Rennes 1 Organization UMR CNRS 6553 Ecobio Address Bât. 14A, Campus Scientifique de Beaulieu, 35042, Rennes Cedex, France Email Author Family Name Ainouche Particle Given Name A. Suffix Division University of Rennes 1 Organization UMR CNRS 6553 Ecobio Address Bât. 14A, Campus Scientifique de Beaulieu, 35042, Rennes Cedex, France Email Author Family Name Salmon Particle Given Name A. Suffix Division University of Rennes 1 Organization UMR CNRS 6553 Ecobio Address Bât. 14A, Campus Scientifique de Beaulieu, 35042, Rennes Cedex, France Email Abstract Polyploidy and recurrent interspecific hybridization represent major features of Spartina evolution, resulting in several superimposed divergent genomes that coexist in the currently living species. This chapter summarizes what we presently know about Spartina history, emphasizing the recent hybridization and polyploidization events that have important ecological and evolutionary consequences. Particular attention is devoted to the recent formation of the allododecaploid invasive Spartina anglica, a salt-marsh “ecosystem engineer” that resulted from hybridization between the hexaploid S. alterniflora (introduced from North America) and tetraploid S. maritima (a European native) and subsequent genome duplication of the F1 hybrid S. x townsendii during the nineteenth century in Western Europe. Allopolyploidy was not accompanied by substantial restructuring of the parental genomes, as observed in some other allopolyploid systems. The major evolutionary events affect the regulatory systems controlling gene expression (including epigenetic regulation), which appear to have been profoundly altered by the merger of different genomes. Methodological challenges in exploring non-model, highly redundant genomes resulting from superimposed events of polyploidization (such as those encountered in Spartina) and the contribution of the new massive parallel sequencing technologies are discussed. 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 F M. Ainouche, H. Chelaifa, J. Ferreira, S. Bellot, A. Ainouche and A. Salmon PR OO 5 Abstract Polyploidy and recurrent interspecific hybridization represent major features of Spartina evolution, resulting in several superimposed divergent genomes that coexist in the currently living species. This chapter summarizes what we presently know about Spartina history, emphasizing the recent hybridization and polyploidization events that have important ecological and evolutionary consequences. Particular attention is devoted to the recent formation of the allododecaploid invasive Spartina anglica, a salt-marsh ‘‘ecosystem engineer’’ that resulted from hybridization between the hexaploid S. alterniflora (introduced from North America) and tetraploid S. maritima (a European native) and subsequent genome duplication of the F1 hybrid S. x townsendii during the nineteenth century in Western Europe. Allopolyploidy was not accompanied by substantial restructuring of the parental genomes, as observed in some other allopolyploid systems. The major evolutionary events affect the regulatory systems controlling gene expression (including epigenetic regulation), which appear to have been profoundly altered by the merger of different genomes. Methodological challenges in exploring non-model, highly redundant genomes resulting from superimposed events of polyploidization (such as those encountered in Spartina) and the contribution of the new massive parallel sequencing technologies are discussed. D 4 Polyploid Evolution in Spartina: Dealing with Highly Redundant Hybrid Genomes TE 3 Chapter 12 EC 2 Book ISBN: 978-3-642-31441-4 Page: 225/242 CO RR 1 Book ID: 272454_1_En Date: 16-8-2012 UN Editor Proof Layout: T1 Standard SC Chapter No.: 12 M. Ainouche (&)  H. Chelaifa  J. Ferreira  S. Bellot  A. Ainouche  A. Salmon University of Rennes 1, UMR CNRS 6553 Ecobio, Bât. 14A, Campus Scientifique de Beaulieu, 35042, Rennes Cedex, France e-mail: malika.ainouche@univ-rennes1.fr P. S. Soltis and D. E. Soltis (eds.), Polyploidy and Genome Evolution, DOI: 10.1007/978-3-642-31442-1_12,  Springer-Verlag Berlin Heidelberg 2012 225 Layout: T1 Standard SC Chapter No.: 12 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 226/242 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 F 30 PR OO 29 D 28 Polyploidy and recurrent interspecific hybridization represent major features of Spartina evolution, resulting in several superimposed divergent genomes that coexist in the currently living species. Particularly fascinating is the rapid range expansion of the recently formed allododecaploid species S. anglica Hubbard that formed in Western Europe during the end of the nineteenth century. The ecological impact and genetic determinants of the spectacular propagation of this invasive species that is now (deliberately or accidentally) introduced on several continents generated an abundant literature (reviewed in Gray et al. 1990, Ainouche et al. 2004a, Triplet and Gallicé 2008, Ainouche et al. 2009) and much interest, as illustrated by the successive International Invasive Spartina Conferences or Forums held in 1990 (Seattle WA, USA), 1997 (Olympia, WA, USA), 2004 (San Francisco, CA, USA), and 2011 (Oakland, CA, USA). In the field of evolutionary biology, Spartina has long represented a textbook example of recent allopolyploid speciation in which the historical context of species formation is particularly well documented (Huskin 1930; Stebbins 1950). In this chapter, we summarize the current state of knowledge about Spartina evolutionary history, including recent insights from evolutionary genetics and genomic approaches. We also examine how genome evolution following natural interspecific hybridization and polyploidization has contributed to diversification and adaptation. Methodological challenges in exploring highly redundant genomes resulting from superimposed events of polyploidization will be discussed. TE 27 EC 26 12.1 Introduction 12.2 Recurrent Reticulate Evolution and Polyploidy in Spartina CO RR 25 M. Ainouche et al. The grass genus Spartina (‘‘cordgrasses’’) belongs to the Chloridoideae subfamily (Fig. 12.1), one of the most poorly understood lineages of the Poaceae. Divergence between Spartina and various grass model species is currently estimated as 35–40 MYA with Sorghum—maize–sugar cane (subfamily Panicoideae) and 50 MYA with rice (subfamily Erhartoideae) (Christin et al. 2008, but see Prasad et al. 2011). Phylogenetic relationships among genera of Chloridoideae are not fully resolved and still are under debate (Hilu and Alice 2001, GPWG 2001). In recent molecular phylogenies, Spartina appears closely related to Sporobolus, Calamovilfa and Zoysia (Columbus et al. 2007, Fortune et al. 2007). Species from these genera share a C4-type photosynthetic system that evolved in the Chloridoideae 25–32 MYA (Christin et al. 2008). C4 photosynthesis is generally considered an adaptation conferring higher productivity under warm temperatures. However, species of Spartina exhibit geographic distributions that cover a range of climatic conditions, from temperate to tropical–subtropical regions; the species exhibit salt and/or drought tolerance on coastal or inland marshes or sand dunes. UN Editor Proof 226 Layout: T1 Standard SC Chapter No.: 12 (a) Book ISBN: 978-3-642-31441-4 Page: 227/242 Polyploid Evolution in Spartina 227 (c) Panicoideae Cynodon dactylon S. argentinensis Hybrids (9x) 6x S. foliosa CHLORIDOIDEAE Hybrids (6x) 4x Cynodonteae (b) Zoy sieae Pooideae S. pectinata 6x , 8x S. cynusoroides Hybrids S. densiflora (7x) PR OO Aristidoideae Bambusoideae S. anglica (12x) S. alterniflora Danthonioideae Ehrhartoideae S. x townsendii (6x) S. maritima Arundinoideae + Micrairoideae S. x neyrautii (6x) F Editor Proof 12 Book ID: 272454_1_En Date: 16-8-2012 S. x caespitosa SPARTINA S. bakeri Calamovilfa Sporobolus Zoysia S patens Eragrostideae S. arundinacea Hybrids Fig. 12.1 Phylogeny of Spartina (see references in the text) a position within the grass family b position within the Chloridoideae subfamily c phylogenetic relationships among Spartina species, recurrent hybridization and polyploidy 66 67 68 69 70 71 72 73 74 Marchant (1968a) Marchant (1968a) Marchant (1968a) Marchant (1968a) Marchant. (1968a) Marchant (1968a) Kim et al. (2010) Ayres et al. (2008), Fortune et al. . (2008) Marchant. (1968b) Ayres et al. (2008) Marchant (1968b) Marchant (1968b) Marchant (1977) Marchant. (1968b) TE densiflora Brongn. maritima (Curtis) Fern. foliosa Trin. alterniflora Loisel. x townsendii H & J Groves x neyrautii Foucaud anglica C.E. Hubbard EC S. S. S. S. S. S. S. 40 40 40 40 40 40 40, 60, 80 70 60 62 62 62 62 120, 122, 124 CO RR 65 patens (Aiton) Muhl. cynusoroides (L.) Roth backeri Merr. gracilis Trin. arundinacea (Thouars) Carmich. pectinata Link Spartina is composed of about 15 perennial species that have mostly diversified in the New World (Mobberley 1956). Accidental or deliberate introduction of species outside their native range over the past 150 years has accelerated diversification by facilitating hybridization with native species, introgression or speciation. The basic (haploid) chromosome number in this lineage is considered to be x = 10 (Marchant 1968a), and all Spartina species recorded to date are polyploid, ranging from tetraploids to dodecaploids (Table 12.1). Molecular phylogenies based on nuclear and chloroplast DNA sequences indicate that Spartina comprises two main lineages that include tetraploid and hexaploid species, respectively (Baumel et al. 2002a). The tetraploid lineage is composed of species native to the New World; these species colonize coastal or UN 64 S. S. S. S. S. S. D Table 12.1 Chromosome numbers reported for Spartina. Additional chromosome numbers resulting from interspecific hybridization and/or backcrosses are presented in the text Taxa 2n References Layout: T1 Standard SC Chapter No.: 12 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 228/242 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 F PR OO 81 D 80 TE 78 79 EC 77 inland salt marshes from either North America (S. patens, S. bakeri, S. gracilis, S. cynusoroides, S. pectinata) or South America (S. ciliata, S. arundinacea). S. argentinensis (syn. S. spartinae), which has a disjunct distribution in Central America and South America, is sister to the hexaploid lineage (Fig. 12.1c). This hexaploid clade is composed of S. maritima, S. alterniflora, and S. foliosa, all colonizing low-marsh zones. S. maritima, native to the Atlantic coasts of Western Europe and Africa was, until the nineteenth century, the only Old World native species, if we exclude recent taxa of hybrid origin. S. foliosa, a species limited to the Pacific coast of North America (California and Mexico), is a weakly supported sister species to S. alterniflora (Baumel et al. 2002a; Ainouche et al. 2004b) and is distributed along the east coast of North and South America from Canada to Argentina. This species has now one of the largest geographic distributions in the genus, as it has been introduced to several continents (Europe, Asia). The nature of polyploidy (auto- vs. allopolyploidy) and the origin of the hexaploid clade are not fully understood. Up to three different duplicated (homoeologous) genes were distinguished in hexaploids for the low-copy nuclear gene Waxy, with substitution rates ranging from 2.18 to 4.79 % among homoeologs (Fortune et al. 2007). The presence of three different homoeologous copies of Waxy could support a hybrid (allopolyploid) origin of this lineage (Fortune et al. 2007). These hexaploid species exhibit regular bivalent pairing (Marchant 1968b). Although allopolyploidy is the currently favored hypothesis for the origin of the hexaploid clade, and is supported by the propensity for interspecific hybridization in this genus, an autopolyploid origin cannot be ruled out. It has been shown that genetic diploidization via disomic inheritance may occur rapidly following autopolyploid formation (Le Comber et al. 2010), leading to a progressive divergence of genes duplicated by polyploidy, and to a gene topology that becomes similar to what would be expected for allopolyploids (Staub et al. 2003). Dating the origin of a polyploid species from (biparentally inherited) nuclear gene divergence data is difficult in that (selective and neutral) population-level evolutionary processes and differential evolution of genes duplicated by polyploidy may give different estimates of the date of origin at different loci. Moreover, in allopolyploids, the divergence of homoeologs actually reflects the divergence between the parental species, thus leading to an overestimated age of the polyploid (Doyle and Negan 2009). We recently attempted to estimate the divergence times between different Spartina lineages from maternally inherited chloroplast sequences. Using reconstruction of the Spartina chloroplast genome from 454 Roche GS-FLX (Titanium) pyrosequencing (Bellot S. and Ainouche M., unpublished), we were able to analyze three coding (matK, rbcL, ndhF) and eight non-coding (intergenic spacers or introns) chloroplast sequences. Using a Bayesian phylogenetic analysis calibrated with known divergence times in the grass family (Wolfe et al. 1989; Gaut 2002; Prasad et al. 2005; Chalupska et al. 2008; Christin et al. 2008), we estimated the divergence between the tetraploid and hexaploid Spartina maternal genomes at 5 MYA, and the divergence time between the two hexaploid S. alterniflora and S. maritima chloroplast genomes at approximately 3 MYA. CO RR 75 76 M. Ainouche et al. UN Editor Proof 228 Layout: T1 Standard SC Chapter No.: 12 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 159 160 161 162 F PR OO 125 D 123 124 TE 122 EC 121 229 Several additional recent hybridization events between Spartina species have led to the formation of homoploid hybrids or new allopolyploids. These events not only have involved parental species of the same ploidal level within the tetraploid or the hexaploid clades, but also involved crosses between tetraploid and hexaploid species (Fig. 12.1c). In the tetraploid lineage, early morphological analyses suggested that S. x caespitosa was a hybrid between S. patens (Eastern North America) and S. pectinata, which has a wide distribution across much of North America (Mobberley 1956). Moreover, a recent survey of S. pectinata populations across North America has revealed hexaploid and octoploid cytotypes that are morphologically indistinguishable from the tetraploids (Kim et al. 2010). The origin of these octoploids is unknown, and the current hypothesis that the hexaploids derive from backcrosses between octoploids and tetraploids has to be verified. Genetic and genomic resources are currently being developed for this species, which is considered a good candidate for bioenergy due to its high biomass production (Gedye et al. 2010). The South American species S. densiflora illustrates well the genetic exchanges occurring between the tetraploid and hexaploid lineages. The history of this species was only recently elucidated: this vigorous, high marsh species originated from southeastern coast of South America, but has been introduced to Chile, California and Spain (Bortolus 2006). Genetic analyses and molecular phylogenies (Baumel et al. 2002a; Ayres et al. 2008; Fortune et al. 2008) have revealed that this species is heptaploid (2n = 70), with a hybrid origin from a tetraploid maternal species closely related to S. arundinacea (with which it shares high chloroplast sequence similarity), which occurs in Southern Hemisphere islands (Amsterdam, St Paul and Tristan de Cunha islands), and a hexaploid paternal species related to S. alterniflora that also occurs in Argentina. Interestingly, S. densiflora appears to have hybridized with each species from the hexaploid clade in both its native area and in more recently colonized regions where it came in contact with native species: (1) At the mouth of the Rio de la Plata in Argentina and Uruguay, morphologically intermediate plants named Spartina longispica were early recognized as hybrids between S. densiflora and S. alterniflora (Saint-Yves 1932; Mobberley 1956). (2) S. densiflora was accidentally introduced to California, probably from Chile, as suggested by geographic, historical, and molecular data (Bortolus 2006; Fortune et al. 2008). In the San Francisco Bay, Ayres et al. (2008) have reported hybrids between S. densiflora and the native S. foliosa (2n = 62). Different chromosome numbers (2n = 66, 94–96) suggest that the F1 hybrids deriving from a heptaploid and a hexaploid parent might also have backcrossed with the parental species. (3) On the Atlantic coast of the Iberian Peninsula, Castillo et al. (2010) have recently detected hybrids between introduced highmarsh S. densiflora and the native low-marsh hexaploid S. maritima (2n = 60). As in California, different chromosome numbers (2n = 64–66, 2n = ca. 94) were recorded for the hybrids, with either S. densiflora or S. maritima as maternal genome donors. No genome doubling (allopolyploid speciation) is recorded to date in these perennial (apparently sterile) hybrids, as occurred for S. anglica. CO RR 119 120 Book ISBN: 978-3-642-31441-4 Page: 229/242 Polyploid Evolution in Spartina UN Editor Proof 12 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 12 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 230/242 170 171 172 173 174 175 176 177 178 179 180 181 182 183 184 185 186 187 188 189 190 191 192 193 194 195 196 197 198 199 200 201 202 203 204 205 206 207 F 168 169 PR OO 167 D 166 TE 165 Within the hexaploid clade, the outcomes of interspecific hybridization are consistent with the phylogenetic relationships and genetic divergence among species (Baumel et al. 2002a; Ainouche et al. 2004b): production of fertile invasive hybrids between closely related taxa on the one hand, and formation of sterile hybrids (followed by genome doubling) between more divergent species on the other (Fig. 12.1c). These hybridizations resulted from recent introductions (nineteenth and twentieth centuries) of S. alterniflora from the eastern coast of the Americas to the Pacific coast of North America and to the Atlantic coast of Western Europe. In California, plants of S. alterniflora spread rapidly (Daehler and Strong 1997; Civille et al. 2005). Hybridization with its sister species S. foliosa (native to California) has been shown to occur in both directions (Ayres et al. 1999). The greater male fitness of S. alterniflora and recurrent backcrosses has resulted in hybrid swarms that progressively replace original S. foliosa plants (Antilla et al. 2000; Ayres et al. 2007). Adding an additional layer of complexity, some of these plants might also be involved in the formation of new hybrid genotypes with the introduced S. densiflora (Ayres et al. 2008). In Europe, S. alterniflora was accidentally introduced by ship ballast in southern England and western France, where it hybridized with S. maritima. In England, hybridization recorded in 1870 gave rise to S. x townsendii, a perennial, sterile hybrid (Groves and Groves 1880). Another sterile hybrid between S. alterniflora and S. maritima was discovered in 1892 in southwestern France in the Bidassoa Estuary (Foucaud 1897) and named Spartina x neyrautii (Jovet 1941). Because of their different morphology, some authors suggested that S. x neyrautii and S. x townsendii might result from reciprocal crosses; however, molecular data revealed that both hybrids share the same chloroplast genome of S. alterniflora, identifying it as the maternal parent of both hybrids (Baumel et al. 2003). After 1890 in England, fertile plants were recorded that appeared to have resulted from chromosome doubling in S. x townsendii (Marchant 1963), thus leading to the formation of a new allododecaploid species named S. anglica (Hubbard 1968), with chromosome numbers 2n = 120, 122, or 124 (Marchant et al. 1968b) suggesting aneuploidy. The vigorous plants (Fig. 12.2) have rapidly colonized Western European salt marshes (Raybould et al. 1991a; Thompson 1991; Genegou and Levasseur 1993). Robust shoots, rhizomes and root systems enable this new species to accumulate large volumes of tidal sediments. For this reason, S. anglica was deliberately introduced in several parts of the world (northern Europe, Australia, New Zealand, China, North America) for land reclamation and marsh restoration purposes. S. anglica has rapidly expanded in its introduced range and now has a worldwide distribution (Ainouche et al. 2009). The rapid spread of the introduced populations has led to various attempts to control or eradicate the species; in fact, it is now listed among the 100 ‘‘world’s worst’’ invaders (IUCN 2000). A recent survey of the ploidal levels in the original population (Southampton area, UK) where S. anglica and S. x townsendii still coexist, has revealed that the perennial, sterile F1 hybrids represent more than 90 % of the population (Renny-Byfield et al. 2010). Using genomic in situ EC 164 CO RR 163 M. Ainouche et al. UN Editor Proof 230 Layout: T1 Standard SC Chapter No.: 12 Book ISBN: 978-3-642-31441-4 Page: 231/242 Polyploid Evolution in Spartina 231 TE D PR OO F Editor Proof 12 Book ID: 272454_1_En Date: 16-8-2012 Fig. 12.2 Spartina anglica in the Baie des Veys (Cotentin), first colonized site (1906) in France (Corbière 1926) 211 212 213 214 215 216 217 218 219 220 221 222 223 224 EC 210 hybridization (GISH), this study also confirmed the existence of nonaploid plants (2n = ca. 90), most likely resulting from backcrosses between S. anglica and its maternal parent S. alterniflora. The recurrent and continuing hybridization and genome duplication in Spartina (Fig. 12.1) make it a particularly useful model with which to explore the consequences of these processes at various evolutionary time scales. Recent events allow ecological, phenotypic, and genetic comparisons between the newly formed hybrids or polyploids and their parents that are still extant in natural populations, a situation that is only met in a few biological systems (see also Chaps. 13 and 14, this volume). CO RR 209 12.3 Ecological and Adaptive Consequences of Hybridization and Polyploidy in Spartina UN 208 Spartina species play an important ecological role in the sedimentary dynamics of salt marshes, where the plants are considered to be ‘‘ecosystem engineers’’(Crooks 2002). The ecological range of S. anglica along the shore is larger than either of its parents. Spartina anglica tolerates several hours of immersion at high tides and thus is able to occupy a vacant niche as a pioneer species in the low-tide zone. This Layout: T1 Standard SC Chapter No.: 12 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 232/242 232 233 234 235 236 237 238 239 240 241 242 243 244 245 246 247 248 249 250 251 252 253 254 255 256 257 258 259 260 261 262 263 264 265 266 267 268 269 F 231 PR OO 230 D 228 229 TE 227 species may accrete large volumes of tidal sediments, making the habitat more terrestrial, and allowing colonization by other salt marsh plant species, which modifies the physical structure of intertidal coastal zones. Recently formed Spartina hybrids and allopolyploids display hybrid vigor and rapid expansion in their native range and invasive abilities when introduced, with important implications for ecosystem management (Lambrinos 2008). S. alterniflora x foliosa hybrids have rapidly invaded the San Francisco Bay (Ayres et al. 2007). Several generations of introgressive hybridization make it difficult to differentiate invasive ‘‘cryptic hybrids’’ from ‘‘pure’’ native S. foliosa plants, and this complicates eradication plans (The Invasive Spartina Project, www.spartina.org). The heptaploid South American S. densiflora has also rapidly colonized Californian marshes where it increased fivefold in distribution during the last 25 years following its introduction (Ayres et al. 2004). In Spain, the hybrid S. densiflora x maritima exhibits greater ecological amplitude than either parental species: hybrids are able to survive both in lower elevations in the intertidal zone where S. maritima naturally grows and also in high marshes where S. densiflora invades (Castillo et al. 2010). Hybridization and polyploidy in Spartina are accompanied by various biological changes that have influenced important adaptive traits such as breeding system, physiology, and morphology. The Californian S. densiflora x foliosa hybrids, derived from self-incompatible, outcrossing parents, have evolved selffertility that has contributed substantially to their rapid spread (Sloop et al. 2009). This breakdown in self-incompatibility, also observed following allopolyploidy in Brassica and Arabidopsis, is most likely triggered by epigenetic mechanisms (Nasrallah et al. 2007). Several transgressive traits in height and biomass, vegetative growth rates, intertidal amplitude, and salinity tolerance are also reported in the Spartina hybrids from California (Ayres et al. 2007). Hybridization between S. alterniflora and S. maritima had very different morphological consequences in the two independent events that occurred in England (S. x townsendii) and France (S. x neyrautii), even though these hybridization events involved crosses in the same direction (S. alterniflora being the maternal genome donor) and similar parental genotypes (Baumel et al. 2003; Yannic et al. 2004; Salmon et al. 2005). S. x neyrautii has shorter spikelets and is distinctly more slender than S. x townsendii, which has longer fleshy leaves, resembling more closely the maternal parent S. alterniflora, whereas S. x townsendii has intermediate morphological features between S. maritima and S. alterniflora (Mobberley 1956). The phenotypic differences between these two F1 hybrids of similar genetic origin are puzzling and most likely represent differential effects of the ‘‘genomic shock’’ resulting from the merger of divergent genomes. S. x townsendii is almost indistinguishable from its allopolyploid derivative S. anglica; moreover, the latter species exhibits larger phenotypic plasticity (Thompson et al. 1991). Physiological and anatomic adaptations are important components of Spartina ecology and distribution (Maricle et al. 2006, 2009). As observed in many polyploids (Otto 2007), stomatal cell size increases with ploidy level in Spartina (Marchant 1967; Kim et al. 2010), which may affect photosynthetic rates (Warner EC 226 CO RR 225 M. Ainouche et al. UN Editor Proof 232 Layout: T1 Standard SC Chapter No.: 12 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 293 294 295 296 297 298 299 300 301 302 303 304 305 306 307 308 309 310 F 276 PR OO 275 D 274 TE 273 and Edwards 1993). The larger ecological amplitude of the allopolyploid S. anglica compared to its parents has to be related to increased tolerance to highly reducing and sulfidic sediment conditions. This increased tolerance may explain the ability of S. anglica to colonize low-marsh zones (Maricle et al. 2006). Survival of S. anglica in anoxic sediments likely is facilitated by its particular ability to develop aerenchyma systems that supply the submerged plants with atmospheric oxygen and efficiently transport oxygen to the roots (Maricle and Lee 2002). S. anglica displays enhanced mechanisms to transport O2 and exhibits five times greater H2S removal than its progenitor species S. alterniflora (the other parental species, S. maritima, was not investigated; Lee 2003). An interesting function seems to have accompanied the formation of the hexaploid lineage of Spartina: the ability to produce dimethylsulfoniopropionate (DMSP), an osmoprotectant and anti-stress molecule (Larher et al. 1977). DMSP is environmentally important as the main biogenic precursor of atmospheric dimethyl sulfide (DMS), which has roles in the biogeochemical sulfur cycle, in cloud formation and in acid precipitation (Kocsis et al. 1998). DMSP is commonly produced by many marine algae, but this capacity is rare in angiosperms, where it has been found only in three genera, one in the Asteraceae (Wollastonia) and two in the Poaceae (Saccharum and Spartina) (Otte et al. 2004). Asteraceae and Poaceae have independently developed different metabolic pathways to achieve this synthesis (Kocsis et al. 1998; Kocsis and Hanson 2000). High DMSP concentrations are found in leaves of the hexaploids S. alterniflora, S. foliosa and S. maritima (Otte et al. 2004), providing these plants with a characteristic ‘‘unpleasant sulphurous odor’’ noticed by early taxonomists (Mobberley 1956). In contrast, no DMSP is detected in the tetraploid species analyzed to date. Although the evolutionary steps giving rise to this physiological novel capacity are yet to be elucidated, the comparative data generated to date indicate that it may have been enabled by the transition to the hexaploid condition. If so, this would represent an example of a presumably important ecological adaptation arising from polyploidy. Physiological (and more generally phenotypic) evolution following hybridization and polyploidy is a direct outcome of the genomic consequences of genome merger and duplication. Although we are still far from obtaining an exhaustive knowledge of the complex gene and regulatory networks involved in most phenotypic traits of adaptive importance, significant progress has been made in recent years regarding genetic and genomic processes accompanying polyploid evolution. EC 272 233 CO RR 270 271 Book ISBN: 978-3-642-31441-4 Page: 233/242 Polyploid Evolution in Spartina 12.4 Genome Evolution Following Hybridization and Allopolyploid Speciation in Spartina UN Editor Proof 12 Book ID: 272454_1_En Date: 16-8-2012 It is now well established that hybrid and allopolyploid genomes are not simply additive with respect to their parental genomes and that myriad novel interactions at both the structural and functional levels may lead to rapid evolution and Layout: T1 Standard SC Chapter No.: 12 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 234/242 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 F 317 PR OO 316 D 314 315 TE 313 evolutionary novelties. It also may be that these phenomena play a critical role in the evolutionary success of newly formed species in both short-term and long-term evolutionary time (e.g., Wendel 2000; Chen 2007; Doyle et al. 2008; Van de Peer et al. 2009). The genetic context of the hybrid or allopolyploid species formation is an important parameter to consider for subsequent evolution. Recurrent hybridization events in natural populations provide new lineages with an expanded genetic base, as observed in the introgressant populations resulting from multiple hybridization and recurrent backcrosses involving different S. alterniflora and S. foliosa genotypes in California (Ayres et al. 2007). In contrast, low genetic diversity is encountered in the parental populations of S. maritima and S. alterniflora that hybridized in southern England and southern France (Baumel et al. 2003; Yannic et al. 2004; Ainouche et al. Ainouche et al. 2004a). As a result, the F1 hybrids S. x townsendii and S. x neyrautii share very similar parental genotypes and exhibit similar, additive genetic composition. A strong genetic bottleneck seems to have affected the new allopolyploid S. anglica as a result of a unique origin from S. x townsendii: a very low interindividual genetic diversity is encountered within and among populations in both its native and introduced ranges (Guénégou 1988; Raybould et al. 1991b; Baumel et al. 2001; Ainouche et al. 2004a), although some genetic variants may be encountered (Ayres et al. 2001). Contrasting with most other young or experimentally resynthesized allopolyploid systems that exhibit rapid genome structural evolution (Ozkan et al. 2001; Skalika et al. 2005; Gaeta et al. 2007; Lim et al. 2008; Tate et al. 2009; Szadowski et al. 2010; Buggs et al. 2012; Chester et al. 2012), the new allododecaploid S. anglica exhibits relative genome stability (Baumel et al. 2002b; Ainouche et al. 2004a). Most of the early evolutionary changes following allopolyploid speciation in Spartina seem to affect the regulation of gene expression, including epigenetic and transcriptome changes (Fig. 12.3). DNA methylation alterations revealed by Methylation Sensitive AFLP (MSAP) appear triggered by hybridization in both S. x townsendii and S. x neyrautii (Salmon et al. 2005). Genome duplication does not entail significant additional changes, as S. anglica has inherited most of the changes observed in S. x towsendii but exhibits few specific methylation alterations. Parisod et al. (2009) have shown that an important fraction of these methylation changes affect regions flanking transposable elements, which agrees with the general view of methylation having evolved to control transposable elements in eukaryotic genomes (Slotkin and Martienssen 2007) and with the fact that no burst of transposition was detected following allopolyploid speciation (Baumel et al. 2002b; Parisod et al. 2010). The hexaploid parental species S. maritima and S. alterniflora have a genome size of 2C = 3.8 pg and 4.3 pg, respectively (Fortune et al. 2008), which suggest that the basic haploid genome ranges between 600 and 700 MB in these species. A preliminary investigation from 454 pyrosequencing of genomic DNA in S. maritima revealed that about 27 % of the analyzed sequences were recognized as repetitive, with Gypsy-like elements being mostly represented. EC 312 CO RR 311 M. Ainouche et al. UN Editor Proof 234 Layout: T1 Standard SC Chapter No.: 12 Book ISBN: 978-3-642-31441-4 Page: 235/242 Polyploid Evolution in Spartina 235 TE D PR OO F Editor Proof 12 Book ID: 272454_1_En Date: 16-8-2012 357 358 359 360 361 362 363 364 365 366 367 368 369 Transcriptome evolution was first investigated in Spartina using oligo-microarrays. As no Expressed Sequence Tag (EST) database was then available for Spartina, heterologous hybridization was performed using the related model species Oryza sativa (Chelaifa et al. 2010a, b). The hexaploid species S. maritima and S. alterniflora that exhibit high exon sequence identity (94.0–99.7 %) at homologous loci displayed 1,247 differentially expressed genes in leaves from plants grown in the same controlled conditions (Chelaifa et al. 2010a). Most of these genes were found to be up-regulated in S. alterniflora. Similar levels of nonadditive parental patterns of gene expression were observed in both of the hybrids S. x townsendii and S. x neyrautii (6.1 and 6.4 % of the analyzed genes, respectively, Fig. 12.3), including parental (mostly maternal) gene expression dominance and transgressively expressed genes. However, the dominance of maternal expression appeared more pronounced in S. x townsendii than in S. x neyrautii. About 8.7 % of the analyzed genes were found differentially expressed between these two F1 hybrids (Fig. 12.3), and interestingly, most transgressively expressed UN 355 356 CO RR EC Fig. 12.3 Genome evolution following recent hybridization and allopolyploidy in Spartina. Percentages in red represent transcriptome changes evaluated using microarrays as described in Chelaifa et al. (2010a, b). Effects of hybridization were estimated by comparing expression profiles in the two independently formed natural F1 hybrids to a theoretical mid-parent value representing additive parental expression. Effects of genome duplication were evaluated by comparing S. anglica to S. x townsendii. Percentages in black represent DNA methylation alterations evaluated using methylation sensitive AFLP, following the procedure employed by Salmon et al. (2005) and Parisod et al. (2009) Layout: T1 Standard SC Chapter No.: 12 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 236/242 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 F 375 PR OO 374 D 373 TE 372 genes were different, with genes up-regulated in S. x townsendii being related to transport, detoxification and stress, and genes up-regulated in S. x neyrautii being related to cellular growth and development. The two independent hybridization events involving the same parental species appear to have generated differential consequences in terms of gene expression. The functions of these differentially expressed genes are consistent with the phenotypic differences previously mentioned between the two hybrids. Genome duplication in S. anglica entailed additional transcriptome changes (Fig. 12.3), consisting of the attenuation of the maternal dominance observed in the F1 hybrid and an increased number of transgressively overexpressed genes (Chelaifa et al. 2010b). Thus, both hybridization and genome duplication appear to have important, though different, effects on the Spartina transcriptome, occurring shortly after genome merger and polyploidization. For the first time, these decoupled effects were analyzed during the allopolyploid speciation process, by comparing the actual (naturally formed) F1 hybrid to its immediately derived allopolyploid that formed and survived in natural conditions. Interestingly, our findings seem to parallel the conclusions emerging from similar comparisons involving natural, more or less recent allopolyploids and/or synthetic F1 hybrids (Hegarty et al. 2006; Flagel et al. 2008; Flagel and Wendel 2010; Buggs et al. 2011). The microarray analyses allowed the first large-scale investigation of the Spartina transcriptome. However, it should be kept in mind that the expression changes reported in these studies are most likely underestimated: only global expression of the genes that hybridized on the rice microarrays were examined, and homoeologous gene expression could not be distinguished. Similar levels of expression might be attained via biased parental expression, which represents an important component of the functional plasticity of polyploid genomes (Adams et al. 2003; Chaudhary et al. 2009; Flagel et al. 2009; Buggs et al. 2011). Distinction of homoeologous gene expression has been studied in several allopolyploid models where diploid representatives of the parental species are identified (e.g. cotton: Udall et al. 2006; Flagel et al. 2008; Arabidopis: Chang et al. 2010; Tragopogon: Buggs et al. 2010; Buggs et al. 2011). In Spartina, this task is particularly challenging as the parents of the young allopolyploid S. anglica are hexaploids (expected to have retained up to three more-or-less divergent duplicated homoeologs per locus) and because no diploid species is known in the genus (Fig. 12.4a). Next-Generation Sequencing (NGS) technologies offer unique avenues to distinguish homoeologous copies in highly redundant genomes from natural, nonmodel species that have experienced successive polyploidization events. This procedure is being developed for Spartina as follows (Fig. 12.4b): (1) construction of a Spartina reference transcriptome for the hexaploid parental species using 454 Roche GS-FLX pyrosequencing; (2) Single Nucleotide Polymorphism (SNP) detection among reads per annotated contig; (3) haplotype assembly to discriminate homoeoalleles; and (4) Illumina RNA-Sequencing (RNA-Seq) to explore variation of homoeolog expression in the parental species. Using the procedure outlined above, 38,000 contigs representing *17,000 unigenes were annotated for S. maritima and S. alterniflora from leaf and root EC 371 CO RR 370 M. Ainouche et al. UN Editor Proof 236 Layout: T1 Standard SC Chapter No.: 12 Book ISBN: 978-3-642-31441-4 Page: 237/242 Polyploid Evolution in Spartina 237 D PR OO F Editor Proof 12 Book ID: 272454_1_En Date: 16-8-2012 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 cDNA libraries, which represent the first reference transcriptome for the hexaploid Spartina species (Ferreira et al. submitted ). An example of SNP detection among reads is presented in Fig. 12.4b for an aligned portion of two homologous contigs annotated in S. maritima and S. alterniflora as HECT-domain-containing protein (Oryza sativa annotation LOC_Os12g24080.1|13112.m02448|cDNA). Reads were assembled using a 95 % identity threshold, to avoid potential comparisons involving paralogs. Six polymorphic sites are detected in this region, including four polymorphic sites shared between S. maritima and S. alterniflora and two species-specific polymorphic sites. The shared polymorphisms allow distinction of two divergent haplotypes (1 and 2, Fig. 12.4b) present in both hexaploids, and one (in S. maritima) or two (in S. alterniflora) additional minor variants. Screening of a larger number of polymorphic sites and loci will provide information about the number and divergence of (homoeo) alleles encountered in the hexaploids and shed light on the evolutionary history of this lineage. A variable number of retained copies per homologous locus may be expected. For instance, Fortune et al. (2007) analyzed the low-copy nuclear gene Waxy that is present in two paralogs (A and B) in Spartina. Only one B copy was encountered in S. maritima, whereas three distinct B copies were found in S. alterniflora. These two species have CO RR 416 UN 415 EC TE Fig. 12.4 a Expected nuclear homoeologous gene copies per locus in the hexaploid parents, F1 hybrid and allododecaploid Spartina species b Procedure for sequence heterogeneity detection at homologous nuclear coding loci; example from two homologous contigs in the hexaploid S. maritima (contig length = 4,294, number of read = 127) and S. alterniflora (contig length = 3,961, number of reads = 85) assembled from 454 Roche pyrosequencing of cDNA libraries. Arrows of similar color (green or pink) represent divergent alleles (1 and 2) present in the two hexaploids and detected from shared polymorphic sites between the two species, in the central zone of the compared region. Other haplotypes represent slight variants of these two copies Layout: T1 Standard SC Chapter No.: 12 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 238/242 M. Ainouche et al. 464 465 466 467 468 Acknowledgments This work benefited from the financial support of the Centre National de la Recherche Scientifique (CNRS), University of Rennes 1, the French National Research Agency (ANR), Region Bretagne, Genoscope, and from the Biogenouest (Transcriptomics and Environmental Genomics) platform facilities. P. Wincker, J. Poulain, Corinne Da Silva, O. Lima, S. Coudouel, D. Naquin, A. Deillhy are thanked for their contribution to the 454 pyrosequencing 469 References 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 470 471 472 473 PR OO 438 D 437 TE 436 EC 435 CO RR 434 F 463 apparently lost the A copy that is still retained in the hexaploid S. foliosa (Fortune et al. 2007). The large-scale detection of homoeologs from massive parallel sequencing will provide a genome-wide view of the retention-loss process at various evolutionary time scales (in the hexaploid parents and the nascent allododecaploid). Transcript loss might result from either homoeolog silencing as observed in the various cases of subfunctionalization reported in allopolyploids (reviewed in Osborn et al. 2003; Chen 2007; Doyle et al. 2008), physical loss of the duplicated copies that may occur more or less rapidly following polyploid speciation (e.g. Gaeta et al. 2007; Tate et al. 2009; Koh et al. 2010; Buggs et al. 2012) or from homoeologous recombination (Cifuentes et al. 2010; Salmon et al. 2010; Gaeta et al. 2010). High-throughput sequencing of genomic DNA and targeted sequencing (e.g. Grover et al. 2012) offer new possibilities to differentiate the effects of these alternatives. Finally, distinction among Spartina homoeologs will allow more accurate analysis of homoeologous expression in the F1 hybrids and the allododecaploid using RNA-Seq data for various organs from plants collected in different ecological conditions. In conclusion, the well-established framework now available for Spartina offers important opportunities to elucidate the phenotypic, ecological, and genomic consequences of recurrent hybridization and polyploidy. Based on the dramatic increase in knowledge that has accumulated for various polyploid systems in recent years (Ainouche and Jenczewski 2010), it has become clear that hybridization and polyploidy generate a range of possible responses that vary among genera. Recent allopolyploidy in S. anglica was not accompanied by rapid restructuring of the parental genomes as has occurred in other polyploid systems. The major evolutionary events in S. anglica appear to affect the regulation of gene expression (including epigenetic regulation); these appear profoundly altered by the merger of different genomes. The rapid advances of NGS technologies will allow more exhaustive exploration of highly redundant genomes (which until now suffered from severe technical limitations) such as those of Spartina. These will provide a better understanding of the genetic and epigenetic mechanisms underlying expression plasticity, and their effect on adaptive and ecologically relevant functions. 433 UN Editor Proof 238 Adams KL, Cronn R, Percifield R, Wendel JF (2003) Genes duplicated by polyploidy show unequal contributions to the transcriptome and organ-specific reciprocal silencing. Proc Natl Acad Sc USA 100(8):4649–4654 Ainouche ML, Jenczewski E (2010) Focus on polyploidy. New Phytol 186:1–4 Layout: T1 Standard SC Chapter No.: 12 239 EC TE D PR OO F Ainouche ML, Baumel A, Salmon A (2004a) Spartina anglica Schreb. A natural model system for analysing early evolutionary changes that affect allopolyploid genomes. Biol J Linn Soc 82:475–484 Ainouche ML, Baumel A, Salmon A, Yannic G (2004b) Hybridization, polyploidy and speciation in Spartina Schreb (Poaceae). New Phytol 161:165–172 Ainouche ML, Fortune PM, Salmon A, Parisod C, Grandbastien M-A, Fukunaga K, Ricou M, Misset M-T (2009) Hybridization, polyploidy and invasion: lessons from Spartina (Poaceae). Biol Invasion. doi:10.1007s10530-0089383-2 Antilla CK, King RA, Ferris C, Ayres DR, Strong DR (2000) Reciprocal hybrid formation of Spartina in San Francisco Bay. Mol Ecol 9:765–770 Ayres DR, Strong DR (2001) Origin and genetic diversity of Spartina anglica (Poaceae) using nuclear DNA markers. Am J Bot 88:1863–1867 Ayres DR, Garcia-Rossi D, Davis HG, Strong DR (1999) Extent and degree of hybridization between exotic (Spartina alterniflora) and native (S. foliosa) cordgrass (Poaceae) in California, USA determined by randomly amplified polymorphic DNA (RAPDs). Mol Ecol 8:1179–1186 Ayres DR, Smith DL, Zaremba K, Klohr S, Strong DR (2004) Spread of exotic cordgrass and hybrids (Spartina sp) in the tidal marshes of San-Francisco Bay CA, USA. Biol Invasions 6:221–231 Ayres DA, Zaremba K, Sloop CM, Strong DR (2007) Sexual reproduction of cordgrass hybrids (Spartina foliosa 9 alterniflora) invading tidal marshes in San Francisco Bay. Divers Distrib 14:187–195 Ayres DR, Grotkopp E, Zaremba C, Sloop CM, Bloom MJ, Bailey JP, Anttila CK, Strong DR (2008) Hybridization between invasive Spartina densiflora (Poaceae) and native S. foliosa in San Francisco Bay. Am J Bot 95:713–719 Baumel A, Ainouche ML, Levasseur JE (2001) Molecular investigations in populations of Spartina anglica C.E. Hubbard (Poaceae) invading coastal Brittany (France). Mol Ecol 10:1689–1701 Baumel A, Ainouche ML, Bayer RJ, Ainouche AK, Misset M-T (2002a) Molecular phylogeny of hybridizing species from the genus Spartina Schreb. (Poaceae). Mol Phylogenet Evol 22:303–314 Baumel A, Ainouche ML, Kalendar R, Schulman AH (2002b) Retrotransposons and genomic stability in populations of the young allopolyploid species Spartina anglica C.E. Hubbard (Poaceae). Mol Biol Evol 19:1218–1227 Baumel A, Ainouche ML, Misset MT, Gourret JP, Bayer RJ (2003) Genetic evidence for hybridization between the native Spartina maritima and the introduced Spartina alterniflora (Poaceae) in South-West France: Spartina x neyrautii re-examined. Plant Syst Evol 237:87–97 Bortolus A (2006) The austral cordgrass Spartina densiflora Brong.: its taxonomy, biogeography and natural history. J Biogeogr 33:158–168 Buggs RJA, Chamala S, Wu W, Gao L, May GD, Schnable PS, et al. (2010) Characterization of duplicate gene evolution in the recent natural allopolyploid Tragopogon miscellus by next generation sequencing and Sequenom iPLEX MassARRAY genotyping. Mol Ecol 19:132–146 Buggs RJA, Zhang L, Miles N, Gao L, Wu W, Schnable P, Barbazuk WB, Soltis PS, Soltis DE (2011) Transcriptomic shock generates evolutionary novelty in a newly formed, natural allopolyploid plant. Curr Biol 21:551–556 Buggs RJA, Chamala S, Wu W, Tate JA, Schnable PS, Soltis DS, Soltis PS, Barbazuk WB (2012) Rapid, repeated, and clustered loss of duplicated genes in allopolyploid Tragopogon populations of independent origin. Curr Biol 22:248–252 Castillo JM, Ayres DR, Leira-Doce1 P, Bailey J, Blum M, Strong DR, Luque T and Figueroa E (2010) The production of hybrids with high ecological amplitude between exotic Spartina densiflora and native S. maritima in the Iberian Peninsula. Diversity and Distributions, 1–12 doi: 10.1111/j.1472-4642.2010.00673.x Chalupska D, Lee HY, Faris JD, Evrard A, Chalhoub B, Kaselkorn R, Gorniki P (2008) Acc Homeoloci and the evolution of wheat genomes. Proc Natl Acad Sci U S A 105:9691–9696 CO RR 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 Book ISBN: 978-3-642-31441-4 Page: 239/242 Polyploid Evolution in Spartina UN Editor Proof 12 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 12 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 240/242 EC TE D PR OO F Chang PL, Dilkes B, McMahon M, Comai L, Nuzhdin SV (2010) Homoeolog-specific retention and use in allotetraploid Arabidopsis suecica depends on parent of origin and network partners. Genome Biol 11:R125. doi:10.1186/gb-2010-11-12-r125 Chaudhary B, Flagel L, Stupar RM, Udall JA, Verma N, Springer NM, Wendel JF (2009) Reciprocal silencing, transcriptional bias and functional divergence of homeologs in polyploid cotton (Gossypium). Genetics 182: 503–517 Chelaifa H, Mahe F, Ainouche M (2010a) Transcriptome divergence between the hexaploid saltmarsh sister species Spartina maritima and Spartina alterniflora (Poaceae). Mol Ecol 19:2050–2063 Chelaifa H, Monnier A, Ainouche M (2010b) Transcriptomic changes following recent natural hybridization and allopolyploidy in the salt marsh species Spartina x townsendii and Spartina anglica (Poaceae). New Phytol 186:161–174 Chen ZJ (2007) Genetic and epigenetic mechanisms for gene expression and phenotypic variation in plant polyploids. Annu Rev Plant Biol 58:377–406 Chester M, Gallagher JP, Symonds VV, Veruska Cruz da Silva A, da Silva A, Mavrodiev EV, Leitch AR, Soltis PS, Soltis DE (2012) Extensive chromosomal variation generated in a recently formed natural allopolyploid species, Tragopogon miscellus (Asteraceae). Proc Natl Acad Sci U S A 109:1176–1181 Christin P-A, Besnard G, Samaritani E, Duvall MR, Hodkinson TR, Savolainend V, Salamini N (2008) Oligocene CO2 Decline Promoted C4 Photosynthesis in Grasses. Curr Biol 18:37–43 Cifuentes M, Grandont L, Moore G, Chèvre AM, Jenczewski E (2010) Genetic regulation of meiosis in polyploid species: new insights into an old question. New Phytol 186:37–45 Civille JC, Sayce K, Smith SD, Strong DR (2005) Reconstructing a century of Spartina alterniflora invasion with historical records and contemporary remote sensing. Ecoscience 12:330–338 Columbus JT, Cerros-Tlatilpa R, Kinney MS, Siqueiros-Delgado M-E, Bell HL, Griffith MP, Refulio-Rodrigez NF (2007) Phylogenetics of Chloridoideae (Gramineae): a preliminary study based on nuclear internal transcribed spacer and chloroplast trnL-F sequences. Aliso 23:565–579 Corbière L (1926) La Spartine de Townsend en Normandie. Bulletin de la société Linéenne de Normandie, 7e série 9:92–117 Crooks JA (2002) Characterizing ecosystem-level consequences of biological invasions: the role of ecosystem engineers. Oikos 97:153–166 Daehler CC, Strong DR (1997) Hybridization between introduced smooth cordgrass (Spartina alterniflora; Poaceae) and native California cordgrass (S. foliosa) in San Francisco Bay, California. U S A Am J Bot 81:307–313 Doyle JJ, Negan EN (2009) Dating the origins of polyploidy events. New Phytol 186:73–85 Doyle JJ, Flagel LE, Paterson AH, Rapp RA, Soltis DE, Soltis PS, Wendel JF (2008) Evolutionary genetics of genome merger and doubling in plants. Annu Rev Genet 42:443–461 Ferreira de Carvalho J, Poulain J, Da Silva C, Wincker P, Michon-Coudouel S, Dheilly A, Naquin D, Boutte J, Salmon A, Ainouche M. Transcriptome de novo assembly and comparative analysis of the hexaploid salt marsh species Spartina maritima and Spartina alterniflora (Poaceae) using high-throughput 454 Roche pyrosequencing. Heredity, submitted Flagel LE, Wendel JF (2010) Evolutionary rate variation, genomic dominance and duplicate gene expression evolution during allotetraploid cotton speciation. New Phytol 186:184–193 Flagel L, Udall J, Nettleton D, Wendel JF (2008) Duplicate gene expression in allopolyploid Gossypium reveals two temporally distinct phases of expression evolution. BMC Biol 6:16 Flagel LE, Chen L, Chaudhary B, Wendel JF (2009) Coordinated and fine-scale control of homoeologous gene expression in allotetraploid cotton. J Hered 100:487–490 Fortuné PM, Schierenbeck K, Ainouche A, Jacquemin J, Wendel JF, Ainouche ML (2007) Evolutionary dynamics of waxy and the origin of hexaploid Spartina species. Mol Phylogenet Evol 43:1040–1055 Fortuné PM, Schierenbeck K, Ayres D, Bortolus A, Clatrice O, Ainouche ML (2008) The enigmatic invasive Spartina densiflora: a history of hybridizations in a polyploidy context. Mol Ecol 17:4304–4316 CO RR 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 M. Ainouche et al. UN Editor Proof 240 Layout: T1 Standard SC Chapter No.: 12 241 EC TE D PR OO F Foucaud (1897) Un Spartina inédit. Ann Soc Sci Nat Char Inf 32:220–222 Gaeta RT, Pires J (2010) Homoeologous recombination in allopolyploids: the polyploidy rachet. New Phytol 186:18–27 Gaeta RT, Pires JC, Iniguez-Luy F, Leon E, Osborn TC (2007) Genomic changes in resynthesized Brassica napus and their effect on gene expression and phenotype. Plant Cell 19:3403–3417 Gaut BS (2002) Evolutionary dynamics of grass genomes. New Phytol 154:15–28 Gedye K, Gonzalez-Hernandez J, Ban Y, Thimmapuram J, Sun F, Wright C, Ali S, Boe A, Owens V (2010) Investigation of the transcriptome of prairie cordgrass, a new cellulosic biomass crop. The Plant Genome 3, 2:69–80 Grass Phylogeny Working Group (GPWG) (2001) Phylogeny and subfamilial classification of the grasses (Poaceae). Ann Missouri Bot Gard 88:373–457 Gray AJ, Benham PEM, Raybould AF (1990) Spartina anglica-the evolutionary and ecological background. In: Gray AJ, Benham PEM (eds) Spartina anglica-a research review. Institute of terrestrial ecology, Natural environment research council, pp 5–10 Grover CE, Salmon A, and Wendel JF (2012) Targeted sequence capture as a powerful tool for evolutionary analysis. Am J Bot (in press) Groves H, Groves J (1880) Spartina townsendii nobis. Rep Bot Soc Exch Club Br Id 1:37 Guénégou MC, Levasseur JE (1993) La nouvelle espèce amphidiplo Spartina anglica C.E. Hubbard: son origine, argumentation et implications. Biogeographica 69:125–133 Guénégou MC, Citharel J, Levasseur JE (1988) The hybrid status of Spartina anglica (Poaceae). Enzymatic analysis of the species and the presumed parents. Can J Bot 66:1830–1833 Hegarty MJ, Barker GL, Wilson ID, Abbott RJ, Edwards KJ, Hiscock SJ (2006) Transcriptome shock after Interspecific hybridization in Senecio is ameliorated by genome duplication. Curr Biol 16:1652–1659 Hilu KW, Alice LA (2001) A phylogeny of Chloridoideae (Poaceae) based on matK sequences. Syst Bot 26:386–405 Hubbard JCE (1968) Grasses, 2nd edn. Penguin Books, London Huskin CL (1930) The origin of S. x townsendii. Genetica 12:531–538 IUCN (2000) World’s worst invasive alien species. In: IUCN The World Conservation Union. http://iucn.org Jovet P (1941) Notes systématiques et écologiques sur les Spartines du Sud-Ouest. Bull Soc Bot Fr 88:115–123 Kim S, Rayburn AL, Lee DK (2010) Genome size and chromosome analyses in prairie cordgrass. Crop Sci 50:2277–2282 Kocsis MG, Hanson AD (2000) Biochemical evidence for two novel enzymes in the biosynthesis of 3-dimethylsulphoniopropionate in Spartina alterniflora. Plant Physiol 123:1153–1161 Kocsis MG, Nolte KD, Rhodes D, Shen TL, Gage DA, Hanson AD (1998) Dimethylsulfoniopropionate biosynthesis in Spartina alterniflora. Plant Physiol 117:273–281 Koh J, Soltis P, Soltis DE (2010) Homoeolog loss and expression changes in natural populations of the recently formed allotetraploid Tragopogon mirus (Asteraceae). BMC Genomics 11:97 Lambrinos G (2008) Managing invasive ecosystem engineers: the case of Spartina in Pacific estuaries. Theor Ecol Ser 4:299–322 Larher F, Hamelin J, Steward GR (1977) L’acide diméthylsulphonium-3-propano de Spartina anglica. Phytochemistry 16:2019–2020 Le Comber SC, Ainouche ML, Kovarik A, Leitch AR (2010) Making a functional diploid: from polysomic to disomic inheritance. New Phytol 186:113–122 Lee RW (2003) Physiological adaptations of the invasive cordgrass Spartina anglica to reducing sediments: rhizome metabolic gas fluxes and enhanced O2 and H2S transport. Mar Biol 143:9–15 Lim KY, Soltis DE, Soltis PS, Tate J, Matyasek R, Srubarova H, Kovarik A, Pires JC, Xiong Z, Leitch AR (2008) Rapid chromosome evolution in recently formed polyploids in Tragopogon (Asteraceae). PLoS ONE 3(10):e3353. doi:10.1371/journal.pone.0003353 CO RR 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 Book ISBN: 978-3-642-31441-4 Page: 241/242 Polyploid Evolution in Spartina UN Editor Proof 12 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 12 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 242/242 EC TE D PR OO F Marchant CJ (1963) Corrected chromosome numbers for Spartina x townsendii and its parent species. Nature 199:929 Marchant CJ (1967) Evolution in Spartina (Gramineae): I. the history and morphology of the genus in Britain. Bot J Linn Soc 60(381):1–24 Marchant CJ (1968a) Evolution in Spartina (Graminae). III species chromosome numbers and their taxonomic signifiance. Bot J Linn Soc 60(383):411–417 Marchant CJ (1968b) Evolution in Spartina (Graminae) II. chromosome basic relationships and the problem of S. x towsendii agg. Bot J Linn Soc 60(383):381–409 Marchant CJ (1977) Hybrid characteristics in Spartina x neyrautii Fouc., a taxon rediscovered in northern Spain. Bot J Lin Soc. 74:289–296 Maricle BR, Lee RW (2002) Aerenchyma development and oxygen transport in the estuarine cordgrasses Spartina alterniflora and S. anglica. Aquat Bot 74:109–120 Maricle BR, Crosier JJ, Bussiere BC, Lee RW (2006) Respiratory enzyme activities correlate with anoxia tolerance in saltmarsh grasses. J Exp Mar Biol Ecol 337:30–37 Maricle BR, Koteyeva NK, Voznesenskaya EV, Thomasson JR, Edwards GE (2009) Diversity in leaf anatomy, and stomatal distribution and conductance, between salt marsh and freshwater species in the C4 genus Spartina (Poaceae). New Phytol 184:216–233 Mobberley DG (1956) Taxonomy and distribution of the genus Spartina. Iowa State Coll J Sci 30:471–574 Nasrallah JB, Liu P, Sherman-Broyles S, Schmidt R, Nasrallah ME (2007) Epigenetic mechanisms for breakdown of self-incompatibility in interspecific hybrids. Genetics 175:1965–1973 Osborn TC, Pires JC, Birchler JA, Auger DL, Chen ZJ, Lee HS, Comai L, Madlung A, Doerge RW, Colot V, Martienssen RA (2003) Understanding mechanisms of novel gene expression in polyploids. Trends Genet 19:141–147 Otte ML, Wilson G, Morris JT, Moran BM (2004) Dimethylsulphoniopropionate (DMSP) and related compounds in higher plants. J Exp Bot 55:919–925 Otto SP (2007) The evolutionary consequences of polyploidy. Cell 131:452–462 Ozkan H, Levy AA, Feldman M (2001) Allopolyploidy-induced rapid genome evolution in the wheat (Aegilops-Triticum) group. Plant Cell 13:1735–1747 Parisod C, Salmon A, Zerjal T, Tenaillon M, Grandbastien M-A, Ainouche M (2009) Rapid structural and epigenetic reorganization near transposable elements in hybrid and allopolyploid genomes in Spartina. New Phytol 184:1003–1015 Parisod C, Alix K, Just J, Petit M, Sarilar V, Mhiri C, Ainouche M, Chalhoub B, Grandbastien MA (2010) Impact of transposable elements on the organization and function of allopolyploid genomes. New Phytol 186:37–45 Prasad V, Stroemberg CAE, Alimohammadian H, Sahni A (2005) Dinosaur coprolites and the early evolution of grasses and grazers. Science 310:1177–1180 Prasad V, Stroemberg CAE, Leaché AD, Samant B, Patnaik R, Tang L, Mohabey DM, Ge S, Sahni A (2011) Late Cretaceous origin of the rice tribe provides evidence for early diversification in Poaceae. Nat Communication 2:480. doi:10.1038/ncomms1482 Raybould AF, Gray AJ, Lawrence MJ, Marshall DF (1991a) The evolution of Spartina anglica C.E. Hubbard (Gramineae): genetic variation and status of the parental species in Britain. Biol J Linn Soc 44:369–380 Raybould AF, Gray AJ, Lawrence MJ, Marshall DF (1991b) The evolution of S. anglica C.E. Hubbard (Gramineae): origin and genetic variability. Biol J Linnean Soc 43:111–126 Renny-Byfield S, Ainouche M, Leitch IJ, Lim KY, Le Comber SC, Leitch AR (2010) Flow cytometry and GISH reveal mixed ploidy populations and Spartina nonaploids with genomes of S. alterniflora and S. maritima origin. Ann Bot 105:527–533 Saint-Yves A (1932) Monographia Spartinarum. Cand 5:19–100 Salmon A, Ainouche ML, Wendel JF (2005) Genetic and epigenetic consequences of recent hybridization and polyploidy in Spartina (Poaceae). Mol Ecol 14:1163–1175 Salmon A, Flagel L, Ying B, Udall JA, Wendel JF (2010) Homoeologous nonreciprocal recombination in polyploid cotton. New Phytol 186:123–134 CO RR 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 M. Ainouche et al. UN Editor Proof 242 Layout: T1 Standard SC Chapter No.: 12 243 EC TE D PR OO F Skalická K, Lim KY, Matyasek R, Matzke M, Leitch AR, Kovarik A (2005) Preferential elimination of repeated DNA sequences from the paternal, Nicotiana tomentosiformis genome donor of a synthetic, allotetraploid tobacco. New Phytol 166:291–303 Sloop C, Ayres DR, Strong DR (2009) The rapid evolution of self-fertility in Spartina hybrids (Spartina alterniflora 9 foliosa) invading San Francisco Bay, CA. Biol Invasions 11:1131–1144 Slotkin K, Martienssen R (2007) Transposable elements and the epigenetic regulation of the genome. Nat Rev Genet 8:272–285 Soltis DE et al (2013) Polyploidy and genome evolution. In: Soltis PS, Soltis DE (eds) The early stages of polyploidy: rapid and repeated evolution in Tragopogon. Springer, Heidelberg Stebbins GL (1950) Variation and evolution in plants. Columbia University Press, New York Straub SCK, Pfeil BE, Doyle JJ (2003) Testing the polyploid past of soybean using a low-copy nuclear gene-Is Glycine (Fabaceae: Papilionoideae) an auto- or allopolyploid? Mol Phylogenet Evol 39:580–584 Szadkowski E, Eber F, Huteau V, Lode M, Huneau C, Belcram H, Coriton O, ManzanaresDauleux MJ, Delourme R, King GJ et al (2010) The first meiosis of resynthesized Brassica napus, a genome blender. New Phytol 186:102–112 Tate JA, Joshi P, Soltis KA, Soltis PS, Soltis DE (2009) On the road to diploidization? Homoeolog loss in independently formed populations of the allopolyploid Tragopogon miscellus (Asteraceae). BMC Plant Biol 9:80. doi:10.1186/1471-2229-9-80 Thompson JD, McNeilly T, Gray AJ (1991) Population variation in Spartina anglica C. E. Hubbard. I. Evidence from a common garden experiment. New Phytol 117:115–128 Triplet P, Gallicé A (2008) Les plantes envahissantes du littoral atlantique: le cas de la Spartine anglaise (Spartina anglica). Aestuaria 13 Aestuarium– Le Forum des Marais Atlantiques eds Udall JA, Swanson JM, Nettleton D, Percifield RJ, Wendel JF (2006) A novel approach for characterizing expression levels of genes duplicated by polyploidy. Genetics 173:1823–1827 Van de Peer Y, Maere S, Meyer A (2009) OPINION the evolutionary significance of ancient genome duplications. Nat Rev Genet 10:725–732 Warner DA, Edwards GE (1993) Effects of polyploidy on photosynthesis. Photosynthetis research 35:135–147 Wendel JF (2000) Genome evolution in polyploids. Plant Mol Biol 42:225–249 Wolfe KH, Gouy M, Yang Y-W, Sharpt PM, Li W-H (1989) Date of the monocot-dicot divergence estimated from chloroplast DNA sequence data. Proc Natl Acad Sci U S A 86:6201–6205 Yannic G, Baumel A, Ainouche ML (2004) Uniformity of the nuclear and chloroplast genomes of Spartina maritima (Poaceae) a salt marshes species in decline along the Western European Coast. Heredity 93:182–188 CO RR 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 Book ISBN: 978-3-642-31441-4 Page: 243/242 Polyploid Evolution in Spartina UN Editor Proof 12 Book ID: 272454_1_En Date: 16-8-2012 Metadata of the chapter that will be visualized in SpringerLink Book Title Polyploidy and Genome Evolution Series Title Chapter Title Allopolyploid Speciation in Action: the Origins and Evolution of Senecio cambrensis Copyright Year 2012 Copyright HolderName Springer-Verlag Berlin Heidelberg Corresponding Author Family Name Hegarty Particle Given Name Matthew J. Suffix Author Division Institute of Biological, Environmental and Rural Sciences Organization Aberystwyth University Address Penglais Campus, SY23 3DA, Aberystwyth, Ceredigion, UK Email ayh@aber.ac.uk Family Name Abbott Particle Given Name Richard J. Suffix Author Division School of Biology Organization University of St. Andrews Address Harold Mitchell Building, KY16 9TH, St. Andrews, Fife, UK Email rja@st-andrews.ac.uk Family Name Hiscock Particle Given Name Simon J. Suffix Abstract Division School of Biological Sciences Organization University of Bristol Address Woodland Road, BS8 1UG, Bristol, UK Email simon.hiscock@bristol.ac.uk Senecio cambrensis is one of a few allopolyploid plant species known to have originated in the recent past and, therefore, provides excellent material for analysing allopolyploid speciation. This allohexaploid species originated in the UK within the last 100 years following hybridization between diploid S. squalidus and tetraploid S. vulgaris. In this chapter, we first describe the events leading up to hybridization between these two species, focusing mainly on the origin and spread of S. squalidus in the UK. We then consider alternative pathways by which S. cambrensis might have originated and conclude that current evidence suggests an origin via formation of the triploid hybrid (S. x baxteri) followed by chromosome doubling. We next review our investigations into levels of genetic diversity and also changes to gene expression and the possible causes of this (epigenetic effects) during the origin of S. cambrensis. High levels of genetic diversity, assessed by surveys of allozyme and AFLP variation, have been recorded in S. cambrensis, and it is likely that intergenomic recombination was an important generator of this diversity. Our studies of ‘resynthesized’ S. cambrensis have shown that the initial genome merger (hybridization) producing S. x baxteri generates genome-wide, non-additive alterations to parental patterns of gene expression and DNA methylation, with genome duplication resulting in a secondary burst of both transcriptional and epigenetic modification. In synthetic allohexaploid lines of S. cambrensis phenotypic changes become apparent from the second to fifth generations, possibly as a consequence of recombination or epigenetic effects; these include changes in ray flower form and emergence of self-incompatible individuals. We conclude by considering the future of S. cambrensis from the standpoint of it being a model species for further study of allopolyploid speciation, and second its long-term success in the wild. Ongoing work to produce a draft reference genome for S. squalidus will underpin future research in S. cambrensis, enabling a more thorough survey of changes to DNA methylation, small RNA activity and promoter binding in the hybrids, as well as comparison with the related allotetraploid S. eboracensis to determine the effects of genome dosage. The future of the species in the wild is currently uncertain. The population in Edinburgh that represented a separate origin of the species in the wild during the 1970s is now extinct, and there has been a marked decline in the number of populations and individuals of the species in its heartland, North Wales, since the 1980s. An analysis of how its ecology compares with those of its parents is lacking. However, it appears to share the same habitats in the wild with its parents, which might have contributed to its decline. Although S. cambrensis may become extinct in the wild in the near future, the potential will remain for it to originate again in the UK providing that conditions prevail for its parents to hybridize. Book ISBN: 978-3-642-31441-4 Page: 245/270 Chapter 13 5 Matthew J. Hegarty, Richard J. Abbott and Simon J. Hiscock 9 10 11 12 13 14 15 16 17 18 19 20 21 22 D 8 Abstract Senecio cambrensis is one of a few allopolyploid plant species known to have originated in the recent past and, therefore, provides excellent material for analysing allopolyploid speciation. This allohexaploid species originated in the UK within the last 100 years following hybridization between diploid S. squalidus and tetraploid S. vulgaris. In this chapter, we first describe the events leading up to hybridization between these two species, focusing mainly on the origin and spread of S. squalidus in the UK. We then consider alternative pathways by which S. cambrensis might have originated and conclude that current evidence suggests an origin via formation of the triploid hybrid (S. x baxteri) followed by chromosome doubling. We next review our investigations into levels of genetic diversity and also changes to gene expression and the possible causes of this (epigenetic effects) during the origin of S. cambrensis. High levels of genetic diversity, assessed by surveys of allozyme and AFLP variation, have been recorded in S. cambrensis, and it is likely that intergenomic recombination was an important generator of this diversity. Our studies of ‘resynthesized’ S. cambrensis have shown that the initial genome merger (hybridization) producing S. x baxteri generates genome-wide, non-additive alterations to parental patterns of gene expression and DNA TE 7 EC 6 CO RR 3 PR OO 4 Allopolyploid Speciation in Action: the Origins and Evolution of Senecio cambrensis 2 F 1 Book ID: 272454_1_En Date: 16-8-2012 M. J. Hegarty (&) Institute of Biological, Environmental and Rural Sciences, Aberystwyth University, Penglais Campus, Aberystwyth, Ceredigion SY23 3DA, UK e-mail: ayh@aber.ac.uk R. J. Abbott School of Biology, University of St. Andrews, Harold Mitchell Building, St. Andrews, Fife KY16 9TH, UK e-mail: rja@st-andrews.ac.uk UN Editor Proof Layout: T1 Standard SC Chapter No.: 13 S. J. Hiscock School of Biological Sciences, University of Bristol, Woodland Road, Bristol BS8 1UG, UK e-mail: simon.hiscock@bristol.ac.uk P. S. Soltis and D. E. Soltis (eds.), Polyploidy and Genome Evolution, DOI: 10.1007/978-3-642-31442-1_13,  Springer-Verlag Berlin Heidelberg 2012 245 Layout: T1 Standard SC Chapter No.: 13 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 246/270 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 F 26 PR OO 25 methylation, with genome duplication resulting in a secondary burst of both transcriptional and epigenetic modification. In synthetic allohexaploid lines of S. cambrensis phenotypic changes become apparent from the second to fifth generations, possibly as a consequence of recombination or epigenetic effects; these include changes in ray flower form and emergence of self-incompatible individuals. We conclude by considering the future of S. cambrensis from the standpoint of it being a model species for further study of allopolyploid speciation, and second its long-term success in the wild. Ongoing work to produce a draft reference genome for S. squalidus will underpin future research in S. cambrensis, enabling a more thorough survey of changes to DNA methylation, small RNA activity and promoter binding in the hybrids, as well as comparison with the related allotetraploid S. eboracensis to determine the effects of genome dosage. The future of the species in the wild is currently uncertain. The population in Edinburgh that represented a separate origin of the species in the wild during the 1970s is now extinct, and there has been a marked decline in the number of populations and individuals of the species in its heartland, North Wales, since the 1980s. An analysis of how its ecology compares with those of its parents is lacking. However, it appears to share the same habitats in the wild with its parents, which might have contributed to its decline. Although S. cambrensis may become extinct in the wild in the near future, the potential will remain for it to originate again in the UK providing that conditions prevail for its parents to hybridize. D 24 TE 23 M. J. Hegarty et al. 44 13.1 Introduction 46 13.1.1 General Introduction 48 49 50 51 52 53 54 55 56 57 58 59 60 61 Polyploidization (genome duplication) is an important evolutionary process in plants (Grant 1981) that appears to have accompanied major transitions in land plant evolution, including the evolution of the seed habit and the evolution of angiosperms (Jiao et al. 2011). All angiosperms are thought to have a polyploid ancestry (Jiao et al. 2011) but most of these are paleopolyploids that now function essentially as diploids. Nevertheless, there are also numerous examples of recently formed polyploids (Adams and Wendel 2005; Wood et al. 2009). Most recent polyploids have formed in association with interspecific hybridization (Grant 1981; Leitch and Bennett 1997; Soltis and Soltis 1999; Otto and Whitton 2000; Leitch and Leitch 2008; Hegarty and Hiscock 2008)—allopolyploidy, which is now recognised as perhaps the most important mechanism of abrupt speciation in plants (Grant 1981; Leitch and Leitch 2008; Hegarty and Hiscock 2008). Allopolyploidy confers rapid fertility to hybrids because duplicated parental chromosomes pair ‘normally’ during meiosis; it also confers reproductive isolation of the hybrids from their parental species because of aberrant or failed chromosome pairing (Soltis and Soltis 1999; Rieseberg et al. 2003; AQ1 CO RR 47 EC 45 UN Editor Proof 246 AQ2 Layout: T1 Standard SC Chapter No.: 13 Book ISBN: 978-3-642-31441-4 Page: 247/270 Allopolyploid Speciation in Action 247 73 13.1.2 The Genus Senecio and the Origins of UK Allopolyploids 71 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 PR OO 70 The genus Senecio (Asteraceae), which includes ragworts and groundsels, is one of the largest and most morphologically diverse genera of flowering plants. With between 1000 and 3000 species, the genus has a worldwide distribution with species described from almost every land mass on Earth (Vincent 1996). Recent revisions, following molecular phylogenetic analysis, of the genus have assigned many former Senecio species to other genera leaving a conservative 1200 species of Senecio sensu stricto (Pelser et al. 2007). Within the genus there are numerous examples of hybridization and polyploidy, and many species have been proposed to have an allopolyploid origin (Abbott and Lowe 1996; Coleman et al. 2001; Kadereit et al. 2006; Pelser et al. 2012). In the UK, three new polyploid taxa have arisen within the last 100 years as a consequence of hybridization between native tetraploid Senecio vulgaris (common groundsel) and the introduced invasive diploid species S. squalidus. The origins of these new polyploid taxa (see below) provide one of the best examples of ‘evolution in action’ (Abbott and Lowe 1996, Hegarty and Hiscock 2008). The introduction and rapid spread of alien S. squalidus across the UK was the main catalyst for this burst of hybrid speciation, which was further facilitated by the obligate outcrossing mating system of S. squalidus increasing the frequency of interspecies pollinations (Brennan and Hiscock 2010). D 68 69 TE 66 67 EC 64 65 CO RR 63 F 72 Rieseberg and Willis 2007). Numerous examples of allopolyploid speciation have been described in the literature, including at least six new plant species that have arisen in the last century: Tragopogon mirus and T. miscellus (Novak et al. 1991, Chap. 14, this volume); Spartina anglica (Gray et al. 1991, Chap. 12, this volume), Senecio cambrensis and S. eboracensis (Ashton and Abbott 1992, Abbott and Lowe 1996, 2004) and Cardamine schulzii (Urbanska et al. 1997). In this chapter, we review the origins of the Senecio allopolyploid species S. cambrensis (Welsh groundsel). We show how resynthesised forms of this allopolyploid can reveal important insights into the genetic and genomic consequences of allopolyploidization and how parental traits, both morphological and physiological, recombine in neopolyploids. 62 13.1.3 Oxford Ragwort in the British Isles: Introduction and Spread UN Editor Proof 13 Book ID: 272454_1_En Date: 16-8-2012 Senecio squalidus (2n = 2x = 20), commonly known as Oxford ragwort, is itself a recently evolved (homoploid) hybrid species (James and Abbott 2005; Abbott et al. 2010; Brennan et al. 2012) that originated on Mount Etna, Sicily, as a result of hybridization between S. aethnensis (2n = 2x = 20, a Mount Etna endemic of higher altitudes) and S. chrysanthemifolius (2n = 2x = 20, a native Sicilian AQ3 Layout: T1 Standard SC Chapter No.: 13 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 248/270 106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 F PR OO 105 D 104 TE 103 EC 101 102 species of lower altitudes). At mid-altitudes on the volcano the distribution of these species overlaps, leading to the formation of a stable hybrid zone (Brennan et al. 2009). Material from this hybrid zone was introduced to the Oxford Botanic Garden in the early 1700s, but records of the collection locality or localities, methods of cultivation, and numbers of plants cultivated in the garden have been lost. Plants subsequently escaped from the Botanic Garden and colonised the masonry of the old college walls from the end of the eighteenth century (Harris 2002). During the industrial revolution of the nineteenth and early twentieth centuries S. squalidus moved rapidly out of Oxford by colonising the clinker beds of the expanding UK railway, of which Oxford was a key hub. The chronology of this famous plant invasion is meticulously documented in a set of papers by Kent (reviewed in Abbott et al. 2009), which record that it began to spread northwards in the late nineteenth century, reaching different parts of northern England during the middle of the twentieth century, before becoming established in the Central Belt of Scotland by the mid-1950s. Today, S. squalidus continues to spread north in Scotland and across Northern Ireland. The rapidity of this invasion is intriguing because S. squalidus is self-incompatible (Abbott and Forbes 1993; Hiscock 2000a, b, and obligate outcrossers are generally thought not to make good colonizers or invasives (Baker 1967). Most invasive species tend to have uniparental reproduction, either sexual (selfing) or asexual (apomixis or vegetative reproduction), although there are exceptions, most notably other species of Asteraceae, such as yellow star thistle, Centaurea solstitialis (Sun and Ritland 1998). According to Baker’s ‘rule’ (Baker 1967), successful colonising and invasive plants are usually self-fertile (self-compatible [SC]) (Stebbins 1957; Baker 1967). Studies of the mating system of S. squalidus, however, have shown that individuals exhibit strong self-incompatibility (SI) across its entire British range, and, as in other species of Asteraceae, this SI is regulated sporophytically by a single polymorphic S locus (Hiscock 2000a, b; Brennan et al. 2002, 2005, 2006). The finding of strong SI in S. squalidus is intriguing, particularly in the light of the extreme population bottleneck that its ancestors must have experienced during its introduction and early colonisation. Following a population bottleneck, allelic diversity at the S locus will be lowered and opportunities for mating (between individuals carrying different S alleles) correspondingly reduced (Hiscock 2000b; Brennan et al. 2002). An extensive survey of SI in S. squalidus across the UK showed that a combination of substantial between-population sharing of the seven S alleles contained in the entire UK population and low levels of selfing (‘pseudo-self-compatibility’) were the most likely cause of S. squalidus’ reproductive success as a colonizer (Brennan et al. 2005, 2006). CO RR 99 100 M. J. Hegarty et al. UN Editor Proof 248 13.1.4 The Origins of Senecio cambrensis During its spread across the UK, S. squalidus (2n = 2x = 20) hybridized with the self-compatible native groundsel, S. vulgaris (2n = 4x = 40), resulting in the recent Layout: T1 Standard SC Chapter No.: 13 Book ISBN: 978-3-642-31441-4 Page: 249/270 Allopolyploid Speciation in Action 249 PR OO F Editor Proof 13 Book ID: 272454_1_En Date: 16-8-2012 146 147 148 149 150 151 152 153 154 155 156 157 158 159 160 161 162 163 164 165 TE EC 144 145 CO RR 142 143 origin of three hybrid taxa. These are the allohexaploid S. cambrensis (2n = 6x = 60), the recombinant tetraploid S. eboracensis (2n = 4x = 40), and the stabilized introgressant radiate form of S. vulgaris, S. vulgaris var. hibernicus (2n = 4x = 40). Interestingly, all three new hybrid taxa are self-compatible, suggesting that this SC mating system, inherited from S. vulgaris, is ‘dominant’ over the SI mating system present in S. squalidus. Detailed descriptions of these new taxa and what is known about their origins are presented elsewhere (see Abbott et al. 1992; Lowe and Abbott 2003; Abbott and Lowe, 2004; Kim et al. 2008). Here, we briefly summarise the information available on the origin of S. cambrensis (Fig. 13.1) focussing particularly on the possible pathways of its origin. Knowing the pathway of origin of a polyploid taxon is helpful in regard to understanding the species’ potential to generate genetic diversity during its initial stages of development, and also for accurate production of synthetic forms of the polyploid used to study possible genetic and epigenetic changes that occurred in the taxon immediately following its origin in the wild (Hegarty et al. 2006, 2008, 2011; Lukens et al. 2006; Buggs et al. 2009). Senecio cambrensis was described by Rosser (1955) from material provided by H.E. Green, who first observed the plant growing at Ffrith and Ceffn-y-bedd, North Wales, UK, in 1948. The plant was described as an annual or short-lived perennial herb that was hexaploid with flower heads (capitula) containing ray florets having short ligules (*4.8 mm in length). Rosser (in Crisp 1972) later determined a herbarium specimen, collected in Denbigh, North Wales in 1925 and originally named as S. squalidus x S. vulgaris, to be S. cambrensis. However, in the absence of a chromosome count there remains some doubt as to whether this specimen is S. cambrensis or, alternatively, a fertile hybrid of S. squalidus and S. vulgaris (see below). Fertile, hexaploid plants with similar morphology to the wild form of UN 140 141 D Fig. 13.1 The neoallohexaploid Senecio cambrensis (centre) flanked by its parents, tetraploid S. vulgaris (left) and diploid S. squalidus (right). Here, S. cambrensis has flower heads (capitula) of intermediate type to its parents; however, its ray florets can vary in length, and occasionally nonradiate forms (lacking ray florets) are also found in the wild Layout: T1 Standard SC Chapter No.: 13 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 250/270 174 175 176 177 178 179 180 181 182 183 184 185 186 187 188 189 190 191 192 193 194 195 196 197 198 199 200 201 202 203 204 205 206 207 208 209 210 F PR OO 172 173 D 170 171 TE 169 EC 168 S. cambrensis can be produced by treating synthetic triploid hybrids between S. squalidus and S. vulgaris with colchicine (Harland 1955, Weir and Ingram 1980, Hegarty et al. 2005). On this basis, Rosser (1955) concluded that S. cambrensis was a new species that originated by hybridization between native S. vulgaris and introduced S. squalidus followed by chromosome doubling. The species is likely to have originated shortly before it was first recorded in North Wales and after S. squalidus had spread to the region in the early part of the twentieth century (Kent 1963). In 1982, S. cambrensis was found growing in Edinburgh, UK (Abbott et al. 1983), and subsequent molecular analysis involving surveys of allozyme and chloroplast DNA variation showed that it had originated independently in Edinburgh rather than being dispersed there from North Wales (Ashton and Abbott 1992; Harris and Ingram 1992). Herbarium records indicate that the Edinburgh lineage may have existed since at least 1974; however, it is now thought to be extinct as the species has not been recorded in the Edinburgh area or nearby since 1993 (Abbott and Forbes 2002). Because S. cambrensis is readily synthesised by treating the triploid hybrid between S. squalidus and S. vulgaris with colchicine, it has been assumed that chromosome doubling of the triploid hybrid was the likely pathway of origin of the allopolyploid species in the wild (Rosser 1955). In theory, however, the species could have originated along several possible pathways (Table 13.1) with the first step involving formation of a triploid, tetraploid, pentaploid or hexaploid hybrid. Of these alternatives, the formation of a triploid hybrid is more likely in that it results from fusion of normal haploid gametes produced by each parent. In contrast, formation of higher ploidy hybrids relies on the production of unreduced gametes, which will be generated at a much lower frequency in each parent species. There are many records of the triploid hybrid (Senecio x baxteri Druce) occurring in the wild (Crisp 1972; Benoit et al. 1975; Marshall and Abbott 1980). It is easily recognised because of its intermediate morphology and its almost complete seed sterility. Progeny tests of S. vulgaris plants have shown that this hybrid is generated regularly but at very low frequencies in the wild where S. vulgaris and S. squalidus co-occur (Marshall and Abbott 1980). In contrast, the tetraploid hybrid has never been reported unequivocally in the wild, although Crisp (1972) described a plant likely to have been such a hybrid based on morphology and an analysis of its offspring. Although no chromosome count was made of the plant, all of its offspring were approximately tetraploid and segregated for a range of morphological, reproductive and disease resistance traits. Crisp (1972) suggested that such offspring could become stabilized in the wild to form distinct taxa, and it is feasible that the tetraploid S. eboracensis originated in this way (Lowe and Abbott 2000). Whether such a hybrid might have contributed to the origin of S. cambrensis as detailed in Table 13.1 remains unknown, but seems less likely than an origin involving the triploid hybrid, given the apparent rarity of the tetraploid hybrid in the wild. Similarly, origins involving the formation of either a pentaploid or hexaploid hybrid are less parsimonious than one involving the triploid hybrid, although cannot be ruled out entirely. Further support for the hypothesis that formation of a triploid hybrid was the first step in the origin of CO RR 166 167 M. J. Hegarty et al. UN Editor Proof 250 Layout: T1 Standard SC Chapter No.: 13 13 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 251/270 Allopolyploid Speciation in Action 251 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 S. cambrensis comes from reports by Vosa (in Crisp 1972) and Ingram (1978) that rare allohexaploid offspring were produced spontaneously by natural selfing of the synthetic triploid hybrids they made. However, Weir and Ingram (1980) also reported the production of an allohexaploid plant directly from a cross between S. vulgaris and S. squalidus. This could have been formed by fusion of unreduced gametes or alternatively by chromosome doubling of a triploid hybrid early in its development. We shall never know exactly how the different lineages of S. cambrensis originated in the UK, but given that the triploid hybrid is regularly encountered in the wild and is capable of producing allohexaploid offspring spontaneously, an origin involving doubling of the chromosome number of a triploid hybrid seems the most likely route of origin. Because S. cambrensis is self-fertile, one newly formed individual of the species would have been able to reproduce sexually and successfully following the species’ origin, i.e. without need of a mate. Moreover, if the species had originated through chromosome doubling of a triploid hybrid, it would be expected initially to be homozygous at all loci within its parental genomes, but to exhibit frequent fixed heterozygosity at duplicated loci among parental genomes. Somewhat CO RR 212 UN 211 EC TE D PR OO F Editor Proof Table 13.1 Some possible pathways of origin for Senecio cambrensis in the wild (1) Via Triploid hybrid (2n=30) • Step 1 - Formation of triploid hybrid (2n=30) through fusion of haploid gametes produced by each parent species. • Step 2 - Chromosome doubling of triploid hybrid by: (i) fusion of ‘unreduced’ triploid gametes (n=30); or (ii) doubling of chromosome number of a somatic cell ancestral to a floret or flower head producing ‘reduced’ triploid gametes (n=30). (2) Via Tetraploid hybrid (2n=40) • Step 1 - Formation of tetraploid hybrid (2n=40) by: (i) fusion of a haploid gamete (n=20) of S. vulgaris and a diploid (unreduced) gamete (n=20) of S. squalidus; or (ii) fusion of an ‘unreduced’ triploid gamete of triploid hybrid (n=30) and haploid gamete of S. squalidus (n=10). • Step 2 - Production of hexaploid hybrid by: (i) fusion of an ‘unreduced’ tetraploid gamete (n=40) generated by tetraploid hybrid and a ‘reduced’ diploid gamete (n=20) of same hybrid or of S. vulgaris or an unreduced gamete of S. squalidus; or (ii) fusion of ‘balanced’ triploid gametes (n=30) produced by same hybrid. (3) Via Pentaploid hybrid (2n=50) • Step 1 - Formation of pentaploid hybrid by: (i) fusion of a haploid gamete of S. squalidus (n=10) and an unreduced gamete of S. vulgaris (n=40); or (ii) fusion of unreduced gamete of triploid hybrid (n=30) with haploid gamete of S. vulgaris (n=20) or unreduced gamete of S. squalidus (n=20). • Step 2 - Production of hexaploid hybrid by: (i) fusion of gametes with same or different ‘balanced’ chromosome numbers (i.e. n=10, n=20, n=30, n=40, n=50) generated by pentaploid hybrid such that the zygote produced is hexaploid (2n=60); (ii) fusion of tetraploid ‘balanced’ gamete (2n=40) produced by pentaploid hybrid with reduced gamete (n=20) of S. vulgaris or unreduced gamete of S. squalidus. (iii) fusion of diploid ‘balanced’ gamete (n=20) produced by pentaploid hybrid with unreduced gamete (n=40) of S. vulgaris. 4) Direct formation of Hexaploid hybrid (2n=60) • Step 1 - Formation of hexaploid hybrid by fusion of unreduced gametes from both S. squalidus (n=20) and S. vulgaris (n=40). Layout: T1 Standard SC Chapter No.: 13 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 252/270 234 235 236 237 238 239 240 241 242 243 244 245 246 247 248 249 250 F 233 PR OO 232 D 230 231 surprisingly, however, the species has been shown through surveys of allozyme variation (Ashton and Abbott 1992) and particularly AFLP variation (Abbott et al. 2007) to contain high levels of genetic diversity, indicating that it rapidly generated this diversity following its origin. Abbott et al. (2007) considered the ways in which such genetic diversity was produced and concluded that intergenomic recombination would most likely have been an important mechanism, although other mechanisms such as aneuploidy, gene conversion, activation of transposons and retroelements, other forms of mutation and gene flow from parental species could not be ruled out. The occurrence of radiate and non-radiate forms of S. cambrensis as well as variation in the ligule length of ray florets have been attributed to intergenomic recombination (Ingram and Noltie 1984), while the observation of multivalent formation occurring at low frequency in meiotic cells of the species (Ingram and Noltie, 1989) provides a mechanism for such recombination to occur. Although not reported by Ingram and Noltie (1989), Crisp (1972) observed up to eight chromosomes with subterminal centromeres in the somatic complement of S. cambrensis plants (based on root tip squashes). He pointed out that as neither parent species possessed such chromosomes they were probably the products of chromosome rearrangements following meiotic abnormalities. In addition to generating genetic diversity, intergenomic recombination resulting in chromosome rearrangements could lead to the formation of reproductive barriers between divergent lineages of S. cambrensis. Clearly, further work on the frequency of intergenomic recombination and its possible effects in S. cambrensis would be worthwhile. TE 229 EC 228 M. J. Hegarty et al. 13.2 Consequences of Hybridization and Polyploidy in Natural and Resynthesised Senecio cambrensis 253 13.2.1 Transcriptome Shock 254 255 256 257 258 259 260 261 262 263 264 265 266 CO RR 252 251 In common with many other allopolyploid species, the merger of two divergent genomes during the formation of Senecio cambrensis has had a dramatic impact at the level of gene expression. As part of our investigation of the allopolyploid origins of S. cambrensis, we conducted gene expression analysis using a custom cDNA microarray platform (Hegarty et al. 2005) to survey the transcript levels of floral genes in both the intermediate triploid hybrid S. x baxteri and wild S. cambrensis, relative to their progenitor species. The experimental design of this comparison is shown in Fig. 13.2. This experiment revealed an initial large change in floral gene expression in S. x baxteri, with approximately 475 cDNA clones showing up- or down-regulation relative to its parental taxa or, also importantly, relative to natural S. cambrensis, from which it differs primarily by a change in ploidal level (Hegarty et al. 2005). Thus, the greatest changes in gene expression relative to the parents appeared to be associated with the hybridization step to form UN Editor Proof 252 Layout: T1 Standard SC Chapter No.: 13 Book ISBN: 978-3-642-31441-4 Page: 253/270 Allopolyploid Speciation in Action 253 PR OO F Editor Proof 13 Book ID: 272454_1_En Date: 16-8-2012 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 TE EC 271 272 CO RR 269 270 S. x baxteri. This initial burst of altered gene expression we termed ‘‘transcriptome shock’’, after the phenomenon of ‘‘genome shock’’ described by McClintock in her seminal work on transposable elements in plant hybrids (McClintock 1984). The ‘transcriptome shock’ effect in S. x baxteri was confirmed in our later analysis of resynthesised S. cambrensis, which further showed that the polyploidization event (here induced by colchicine) had an immediate calming (ameliorating) effect on altered patterns of gene expression detected in S. x baxteri (Hegarty et al. 2006). Importantly, this altered pattern of gene expression, apparent in first-generation allopolyploids, was preserved in four successive generations of the synthetic allopolyploids and in wild S. cambrensis (Hegarty et al. 2006). Previous research in resynthesised wheat (Feldman and Levy 2005) identified separate effects of hybridization and polyploidization on the genome and transcriptome, but our findings in S. cambrensis represented one of the first indications that these changes in gene expression were genome-wide. Interestingly, the putative functional classes of genes affected by hybridization and allopolyploidization were remarkably similar, with no functional class of genes being overly affected by hybridization or allopolyploidization (Fig. 13.3). However, perhaps not surprisingly, when compared with functional classes of genes not affected by either process, there was a greater representation of genes potentially involved in flower/inflorescence and pollen developments, which may reflect the transitions in floral phenotypes observed after hybridization and allopolyploidization. We later reassessed the data (Hegarty et al. 2008) in light of a new approach used by Wang et al. (2006a) in their studies of allotetraploid Arabidopsis suecica. In this study, they focused on the identification of genes whose expression in hybrids differed from the additive expression midpoint of the two different parental UN 267 268 D Fig. 13.2 Experimental design employed in microarray comparisons of gene expression between the allopolyploid S. cambrensis and its progenitor taxa, S. vulgaris and S. squalidus, and their sterile triploid F1 hybrid, S. x baxteri. Experimental details can be found in Hegarty et al. (2005), but, briefly, mature flower bud tissue was harvested from a mixed population of approximately 30 plants and pooled prior to RNA extraction to create an ‘average’ for each taxon. Labelled cDNA for each taxon was hybridized to a custom floral cDNA microarray. Two taxa were differentially labelled and compared per array hybridization (with 10 replicate hybridizations performed per comparison) using dye swaps to account for any bias in labelling efficiency. Each taxon was compared with the other three, for a total of 30 array hybridizations per taxon. Raw expression data for each taxon were extracted from these 30 replicates and imported separately into the GENESPRING microarray analysis software (Silicon Genetics) to enable comparison between all four taxa. Figure reproduced from Hegarty et al. (2008) AQ5 Layout: T1 Standard SC Chapter No.: 13 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 254/270 M. J. Hegarty et al. D PR OO F Editor Proof 254 294 295 296 297 298 299 300 301 302 303 304 305 306 307 gene copies. A similar approach had also been used to analyse gene expression change in maize hybrids (Stupar et al. 2007). Such an approach provides a consistent and unified methodology for identifying genes affected by hybridization and/or polyploidization in different model study systems. We therefore reanalysed the microarray data from our original study to identify specific genes and classes of genes affected by hybridization and polyploidization. Using methods similar to Stupar et al. (2007), we tested whether changes in gene expression observed in synthetic S. x baxteri and wild allohexaploid S. cambrensis were additive or nonadditive (Hegarty et al. 2008). By averaging the parental expression values for each feature on the array showing differential expression in the hybrids, a parental midpoint expression value (MPV) was obtained. The derived midpoint values were then used to calculate a ratio of hybrid and parental expression values compared to the MPV for each array feature. A ratio of -0.33 indicates additive gene expression, whilst ratios below -1 or above 1 represent expression in the hybrid outside the range of either parent. Statistical analysis of differentially expressed genes from our previous microarray experiment (Hegarty et al. 2005) showed that, CO RR 293 UN 292 EC TE Fig. 13.3 Functional classes of genes affected by allopolyploidization and hybridization. Basic gene ontologies for cDNA clones displaying a conserved expression changes in both wild and synthetic Senecio cambrensis relative to S. x baxteri (genes affected by allopolyploidy, 540 clones), b expression changes relative to the parental taxa S. squalidus and S. vulgaris in both hybrid taxa (genes affected by hybridization, 99 clones) and c genes showing no expression difference between the parental and hybrid taxa (unaffected by hybridization or polyploidy, 289 clones). With the exception of a higher proportion of floral/pollen-related genes in (a, b) compared with (c), there are no substantial differences between the classes of affected genes (adapted with permission from Hegarty et al. 2006) Layout: T1 Standard SC Chapter No.: 13 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 F 314 PR OO 313 D 311 312 for both hybrids, the median ratio was significantly different from -0.33, allowing us to reject the null hypothesis of largely additive gene expression changes. Instead, for both hybrids, the majority of the data were skewed towards one of the parents; in the case of S. x baxteri, expression was skewed towards that of the lower expressing parent, whereas in S. cambrensis it was skewed towards that of the higher expressing parent. Further analysis of the data showed that for both hybrids S. vulgaris was the lower expressing parent in 70 % of cases. Expression outside the parental range was observed in a substantial proportion of cases in both hybrids: 7.42 % in S. x baxteri and 3.03 % S. cambrensis (Hegarty et al. 2008). Having identified a pool of cDNA clones displaying non-additive changes to gene expression in both hybrid taxa, we then tested these clones for evidence of expression beyond the parental ranges, i.e. transgressive gene expression. In S. x baxteri, 80.4 % of non-additively expressed clones differed from the MPV by [1.5-fold, with 42.2 % of clones in S. cambrensis showing the same effect. Within both of these groups, the majority of cases involved upregulation compared with the MPV (66.9 and 70.4 % in S. x baxteri and S. cambrensis, respectively). Aside from the genes for which no functional class could be ascribed (49.2 %), the major functional groups affected in S. x baxteri were genes involved in development (6.6 %), nucleotide binding (6.1 %), mitochondrial activity (4.76 %) and cell wall function (3.97 %). Within the development category, a high proportion of clones (32 %) were found to encode tubulins, profilins or senescence-associated proteins. Of the clones involved in nucleotide binding, 34 % were transcription factors. In S. cambrensis, the majority of clones (58 %) could not be assigned to a functional category. Of the remainder, the largest categories were defense (11.1 %) and cell wall-related genes (6.17 %). Our reanalysis of the microarray data therefore revealed a relatively high proportion of non-additive gene expression change in the hybrids relative to their parental expression levels (Hegarty et al. 2008). In addition, the degree of nonadditive gene expression was lower in allohexaploid S. cambrensis compared with its triploid intermediate S. x baxteri. This finding was consistent with our previous observation that the ‘‘transcriptome shock’’ resulting from allopolyploidization is largely due to hybridization, with polyploidization resulting in a distinct secondary shift (amelioration) in gene expression (Hegarty et al. 2006). The fairly diverse nature of the genes affected was consistent with other findings in Arabidopsis (Wang et al. 2006b), cotton (Adams et al. 2004) and maize (Stupar et al. 2007) that non-additive changes to gene expression are genome-wide. Interestingly, similar functional classes of genes were affected by hybridization in Senecio, Arabidopsis and maize (Fig. 13.4), suggesting that certain gene networks may be particularly susceptible to perturbation by hybridization; the functional categories of nucleotide binding, defense and mitochondria being good examples. In terms of the classes of genes affected in Senecio, it is also noteworthy that one of the major affected groups in S. x baxteri was nucleotide binding. In addition to a number of (primarily down-regulated) transcription factors that have shown similar alterations in expression pattern in the polyploid Arabidopsis suecica (Wang et al. 2006b), this category also contained clones encoding cytidine TE 310 255 EC 309 CO RR 308 Book ISBN: 978-3-642-31441-4 Page: 255/270 Allopolyploid Speciation in Action UN Editor Proof 13 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 13 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 256/270 M. J. Hegarty et al. PR OO F Editor Proof 256 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 deaminase (CDA) and 8-oxoG-DNA glycosylase (OGG1). OGG1 has been implicated in DNA base excision repair (García-Ortiz et al. 2001), while CDA has been suspected to be involved in RNA editing, although it now appears that pentatricopeptide repeat proteins that contain CDA-like domains are the more likely candidates (Salone et al. 2007). These genes were of interest, given that we also observed a relatively high number of clones encoding proteins involved in either DNA modification or cell division. In addition to cytidine deaminase and OGG1 (both of which were upregulated compared with the parental midpoint in S. x baxteri), there was also up-regulation of adenosylhomocysteinase and adenosyl kinase, the genes involved in S-adenosylmethionine (SAM) dependent methylation (Moffatt et al. 2002; Mull et al. 2006). The expression of SAM synthetase was also increased relative to both parents. SAM-dependent methylation is used in gene silencing and also in pectin methylation (Pereira et al. 2006), and, indeed, we observed an increase in the expression of pectin methylesterase as well as another SAM-dependent enzyme, caffeic acid 3-O-methyltransferase (Hegarty et al. 2008). It is clear from these analyses that hybridization and polyploidy have separate, distinct effects on gene expression in Senecio. These changes in gene expression CO RR 354 UN 353 EC TE D Fig. 13.4 Comparison of non-additive expression changes resulting from hybridization in maize and hybridization/polyploidy in Senecio and Arabidopsis. The formation of the hybrids is shown in each case, together with the level of non-additive gene expression in the hybrids expressed as a percentage of the features on the microarray platform used. Finally, the top five functional gene classes affected (ignoring unknowns) for each hybrid are displayed for comparison. Red indicates a functional gene class affected in all four hybrid systems, blue indicates a functional class affected in at least one of the Senecio hybrids and one of the other two hybrid taxa, and green indicates a functional class affected in both Senecio hybrids but not in either Arabidopsis or maize. For Arabidopsis suecica, gene function data were taken from Wang et al. (2006b), and for maize from an extrapolation of the supplementary data given in Stupar et al. (2007). Sb, S. x baxteri; Sc, S. cambrensis. Reproduced from Hegarty et al. (2008) Layout: T1 Standard SC Chapter No.: 13 Book ISBN: 978-3-642-31441-4 Page: 257/270 Allopolyploid Speciation in Action 257 376 13.2.2 Epigenetic Modification 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 PR OO 377 While we were unable to demonstrate complete silencing of any clones in our microarray analyses, it is important to note that the cDNA-based microarrays used could not distinguish between different parental homeologues. Therefore, in the cases where a hybrid showed down-regulation of a gene relative to its parental taxa, it may have been due to either homeologue loss or silencing as a consequence of DNA methylation. Both phenomena were observed using cleaved amplified polymorphic sequence (CAPS) analysis in natural and resynthesised allotetraploid Tragopogon miscellus (Buggs et al. 2009) and suggested that homeologue loss was not an immediate consequence of hybridization or polyploidy but occurs slightly later due to recombinational events. Silencing was also not observed to occur immediately in Tragopogon synthetics, but observations in natural populations suggest that silencing of one homoeologue is more prevalent than sequence loss. Studies in other allopolyploid systems such as wheat, Arabidopsis and Spartina have demonstrated that methylation status in F1 hybrids and first-generation allopolyploids displays similar changes to those observed for gene expression. To determine whether methylation in Senecio x baxteri and S. cambrensis was affected in a similar manner, we undertook a methylation-sensitive AFLP (MSAP) analysis of the parental taxa, three triploid lines and their S0–S1 allohexaploid derivatives (Hegarty et al. 2011). The MSAP technique (Xiong et al. 1999) involves digestion of genomic DNA with a standard, rare-cutter enzyme (i.e., EcoRI) and one of a pair of isoschizomeric enzymes that share a restriction site but are either sensitive or insensitive to cytosine methylation. In this case, we employed HpaII, which is sensitive to methylation of either cytosine in its recognition site (CCGG) and MspI, which is sensitive only to methylation of the external cytosine. Comparing the two AFLP profiles produced from each isoschizomer enables identification of the methylation status of a given locus, as described in Table 13.2. Five selective primer combinations were used to screen 33 plants as detailed in Table 13.3. We successfully amplified MSAP products from all 33 individuals and obtained a total of 408 reliable MSAP loci, which were then surveyed to determine levels of non-additivity in the hybrids. Of the 408 loci, 264 (64.7 %) showed polymorphisms between the parental taxa S. squalidus and S. vulgaris. In the remaining 144 loci that were monomorphic in the parents, 75.7 % were monomorphic across all hybrid samples tested. Surveying the polymorphic loci, it was found that the D 374 TE 373 EC 372 CO RR 371 F 375 affect a wide spectrum of transcripts from a number of functional groups and are non-additive in nature. A good proportion of these changes are also transgressive—outside the range of either parent species—representing a possible source of heterotic success in interspecific hybrids. To investigate possible mechanisms for these changes in gene expression, we focused our attention on epigenetic modification in Senecio hybrids. 370 UN Editor Proof 13 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 13 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 258/270 258 M. J. Hegarty et al. F Editor Proof Table 13.2 HpaII% MspI banding patterns and locus methylation state (reproduced from Hegarty et al, 2011) HpaII MspI Methylation status + CCGG (unmethylated) + CmCGG (methylation of internal cytosine) m CCGG or mCmCGG (methylation of external cytosine) hm + CCGG (hemimethylation of external cytosine) PR OO Note: In the case of methylation of the external cytosine, it is not possible from MSAP alone to determine whether the internal cytosine is also methylated 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 CO RR 413 triploid hybrids each displayed similar overall methylation patterns (Table 13.4), with a strong bias in favour of the S. vulgaris (maternal) methylation state for each locus (an average of 57.4 % of loci between the three triploid lines). In addition, the triploids also displayed approximately equivalent levels of non-additive methylation (13.4 % on average). To determine whether methylation patterns were maintained following genome duplication, the methylation status of the S0 allohexaploids was compared to the triploid lines from which they were derived. In the vast majority of cases, the synthetic allohexaploids retained the methylation state observed in the triploid. On average, 78.2 % of these cases involved additive methylation, whereas 9.7 % of loci involved retention of a non-additive methylation profile. An average of 10.1 % of loci displayed a shift relative to the triploid in the specific parental methylation state favoured (4.5 % shift to S. squalidus, 3.8 % shift to S. vulgaris, 1.8 % shift to both). Finally, an average of 2 % of loci showed novel non-additive methylation not observed in the triploids. The analysis was then extended to the S1 allohexaploids, comparing them to both their triploid ancestors and the preceding S0 generation. As with the S0 allohexaploids, overall methylation state was highly similar among the three lines. Again, the most common result was retention of methylation status compared to both the triploid and the S0 allohexaploid, with an average of 70.3 % of loci showing retention of additive methylation patterns and 7.3 % of non-additive UN 411 412 EC TE D Table 13.3 Oligonucleotides for methylation-sensitive AFLP and number of loci (reproduced from Hegarty et al, 2011) Oligo Sequence (5’-3’ Primer Combination # Loci EcoRI adaptor 1 CTCGTAGACTGCGTACC Eco+AGC/Hpa+CTG 118 EcoRI adaptor 2 AATTGGTACGCAGTC Eco+AAC/Hpa+CTT 88 HpaII adaptor 1 GATCATGAGTCCTGCT Eco+AGC/Hpa+CTT 80 HpaII adaptor 2 CGAGCAGGACTCATGA Eco+AAC/Hpa+AAG Failed Eco+A GACTGCGTACCAATTCA Eco+ACG/Hpa+AAG 69 Hpa ATCATGAGTCCTGCTCGG Eco+AAC/Hpa+CTT 53 Eco+AAC (NED) GACTGCGTACCAATTCAAC Eco+ACG (FAM) GACTGCGTACCAATTCACG Hpa+CTG ATCATGAGTCCTGCTCGGCTG Hpa+CTT ATCATGAGTCCTGCTCGGCTT Hpa+AAG ATCATGAGTCCTGCTCGGAAG Layout: T1 Standard SC Chapter No.: 13 13 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 259/270 Allopolyploid Speciation in Action 259 TE D PR OO F Editor Proof Table 13.4 Summarised methylation status of hybrid lines (reproduced from Hegarty et al. 2011) Methylation Percentage of Loci (triploids) State Line 1 Line 2 Line 3 Additive (SS) 24.58% 24.58% 22.26% Additive (SV) 56.15% 57.14% 58.80% Additive (monomorphic) 4.65% 6.98% 4.65% Nonadditive 14.62% 11.30% 14.29% Percentage of Loci (S0 allohexaploids) Same as triploid (additive) 76.17% 78.86% 79.53% Same as triploid (nonadditive) 9.73% 9.73% 9.73% Differs from triploid (additive SS) 5.03% 4.36% 4.03% Differs from triploid (additive SV) 5.37% 3.69% 2.35% Differs from triploid (monomorphic) 2.01% 0.67% 2.68% Novel nonadditive methylation 1.68% 2.68% 1.68% Percentage of Loci (S1 allohexaploids) Same as triploid+S0 (additive) 72.06% 72.38% 66.35% Same as triploid+S0 (nonadditive) 7.62% 7.94% 6.35% 7.62% 5.71% 3.49% Same as S0 not triploid (additive) Same as S0 not triploid (nonadditive) 0.32% 1.27% 0.63% Same as triploid not S0 (additive) 3.81% 3.17% 1.90% 0.95% 0.00% 2.54% Same as triploid not S0 (nonadditive) Differs from triploid+S0 (additive SS) 2.54% 4.76% 7.62% Differs from triploid+S0 (additive SV) 3.49% 2.22% 7.62% Novel nonadditive methylation 1.59% 2.54% 3.49% 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 methylation. A further 5.6 % of loci showed retention of additive methylation in the two allohexaploid generations, whereas the triploid had displayed non-additivity. Methylation status was not always consistent between the two allohexaploid generations; however, in 4.1 % of cases, loci displayed a shift in methylation state to that seen in the triploid but not in the S0 allohexaploid. Furthermore, the S1 generation showed a return to additivity in 9.4 % of cases, where both the triploid and the S0 allohexaploids had been non-additive. As with the S0 generation, though, there was a small (2.5 % average) degree of novel non-additivity. The results were in accordance with our previous studies of gene expression (Hegarty et al. 2006, 2008): we found that, while cytosine methylation in both hybrid taxa was largely additive between the two parental patterns, a significant degree of non-additivity also exists. Overall methylation status was well conserved between different hybrid lines; while individual loci displayed differences, the global percentages of different methylation states were highly similar between lines (Hegarty et al. 2011). In all three triploid lines, approximately 13.4 % of loci showed non-additive methylation, although the precise type of methylation was not identical between lines in all cases. Levels of non-additive methylation observed in other allopolyploid systems are variable: 8.3 % in Arabidopsis CO RR 432 UN 431 EC Note: SS = S. squalidus, SV = S. vulgaris. ‘‘Monomorphic’’ refers to loci with a common methlyation state in the parents, rather than across hybrid lines Layout: T1 Standard SC Chapter No.: 13 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 260/270 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 F 455 PR OO 454 D 453 TE 451 452 (Madlung et al. 2002), 9 % in Brassica (Gaeta et al. 2007), 13–20 % in wheat (Dong et al. 2005; Pumphrey et al. 2009) and as high as 30 % in Spartina (Salmon et al. 2005). It has been speculated that the higher genome copy number in wheat and Spartina might explain their greater levels of methylation, although Doyle et al. (2008) point out that both species are monocots, which tend to possess a higher GC content (and thus greater potential for methylation) than eudicots. The fact that conserved methylation changes between the Senecio hybrids are more on a par with the levels seen in Arabidopsis and Brassica suggests that this latter hypothesis may be correct, as S. cambrensis exhibits the same ploidy as wheat. We should note, however, that the wheat genome is significantly larger than that of Senecio (2C genome sizes of 33.96 Gbp wheat; 5.05 Gbp S. cambrensis) and is known to contain a significant amount of repetitive DNA including large numbers of retroelements. It is therefore probable that alterations to methylation are more necessary to prevent widespread activation of these genetic regions in wheat and similar polyploid monocots. Indeed, studies of methylation in Spartina (Parisod et al. 2010) showed that such changes frequently occur in the vicinity of transposable elements and, perhaps as a result, no transposition burst was detected in the Spartina hybrids analysed. Methylation change therefore appears to play a frequent role in genome mergers, but there are exceptions: despite significant differential gene expression in the allotetraploid Gossypium hirsutum, almost no differences in methylation could be observed between the hybrid and its parental taxa (Liu et al. 2001) nor do the parental genomes undergo any significant rearrangement. In this situation, it appears that subfunctionalization of the two genomes is the primary cause of phenotypic variation (Adams et al. 2004; Liu and Adams 2007), although a recent study by Chaudhary et al. (2009) demonstrated that neofunctionalization and divergence in parental cis-regulatory sequences also play a significant role. Exactly what factors determine the response of the parental genomes to hybridization are largely unknown, although the degree of parental divergence is speculated to play a large role (Chapman and Burke 2007; Buggs et al. 2008; Paun et al. 2009). In further accordance with our expression analyses, we observed that nonadditive methylation in S. x baxteri triploids was maintained, on average, in only 73.6 and 55.6 % of cases in the S0 and S1 allohexaploids, respectively (Hegarty et al. 2011). In approximately 73 % of cases observed in our microarray studies, the resynthesised allohexaploid lines (S0–S4) displayed either an immediate or gradual shift towards an expression pattern similar to that of wild S. cambrensis (Hegarty et al. 2006). It seems likely, therefore, that our previous observation that non-additivity results from hybridization but can be partially reduced by genome duplication, also holds true when applied to DNA methylation. This again matches observations from MSAP analysis of Spartina allopolyploids (Salmon et al. 2005), which showed that the allopolyploid S. anglica retained 71.4 % of the non-additive methylation patterns observed in the F1 hybrid. These findings were again confirmed when assessing methylation changes associated with transposable elements (Parisod et al. 2010), with the additional observation that many of the changes seen in the F1 hybrid involved loss of parental markers (usually in the maternal EC 450 CO RR 449 M. J. Hegarty et al. UN Editor Proof 260 Layout: T1 Standard SC Chapter No.: 13 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 F PR OO 499 500 D 497 498 genome), indicating that such changes involved structural rearrangements to the parental genomes. A similar process may be at play in S. x baxteri, because MSAP markers also detect structural changes: indeed, wild populations of S. cambrensis show evidence of intergenomic recombination (Abbott et al. 2007, and above), again favouring the S. vulgaris genome as with our triploid lines here. By contrast, most of the differences observed in allopolyploid Spartina involved alterations to methylation status, rather than structural changes. Similar findings have also been identified in Brassica, where most of the methylation changes identified in the S0 allotetraploid were maintained in S5 lines, but with a number of revertants and novel changes present as well (Gaeta et al. 2007). A key finding from our analysis was that the global patterns of DNA methylation change observed in our experiments strongly mirror the global changes in gene expression observed in our earlier microarray analyses, indicating a possible underlying causation. Whilst further investigation of specific loci showing methylation differences is required to make a definitive case, the similarities between changes in gene expression and DNA methylation are nevertheless striking. For example, we noted that a number of loci displayed novel non-additivity in both the S0 and S1 allohexaploids (2.01 % on average in the S0 lines, 2.54 % in the S1), again a point of consistency between the methylation study and our earlier microarray expression analysis (Hegarty et al. 2008). The overall level of nonadditive methylation may therefore not actually decrease as a consequence of genome duplication, but instead the level of conserved methylation may be lessened. A proportion of loci also displayed unstable methylation patterns across generations in the hexaploids, with an average of 16.08 % of loci showing differences between the S0 and S1 lines (including the aforementioned novel nonadditive methylation). Of these, the majority consisted of cases where the S1 allohexaploids revert to additivity or favoured a different parental methylation state to the S0 line. This reversion to an additive profile was also observed between the triploids and the S0 plants and agreed with observations from the microarray data that wild S. cambrensis often showed an opposing expression pattern to S. x baxteri. However, approximately one-quarter of loci also displayed a shift relative to the S0 allohexaploids to favour the methylation state seen in the triploid. These findings suggest that the methylation state of some loci may vary as a consequence of segregation. This may explain the novel changes observed by Gaeta et al. (2007) in their S5 allopolyploids of Brassica napus. Similarly, an analysis of natural populations of the allopolyploid Tragopogon miscellus (Buggs et al. 2009), where hybridization occurred at least 40 generations ago, identified a random loss of one parental homoeologue at a rate of 3.2 % across 10 loci in 57 natural hybrids from five populations. In addition, a further 6.8 % of loci showed evidence of gene silencing in one parental copy. The loci lost/silenced were not consistent across populations or individuals, although within populations, there was some conservation in the loci affected. Whilst Buggs et al. (2009) did not note any homoeologue loss/silencing in resynthesised S0 hybrids, the variability in silencing after such an extended period of time since hybrid formation suggests that TE 496 261 EC 495 CO RR 494 Book ISBN: 978-3-642-31441-4 Page: 261/270 Allopolyploid Speciation in Action UN Editor Proof 13 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 13 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 262/270 M. J. Hegarty et al. 559 13.2.3 Phenotypic Change 546 547 548 549 550 551 552 553 554 555 556 557 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 PR OO 545 D 544 TE 542 543 EC 541 Senecio x baxteri F1 hybrids generated by crossing S. vulgaris, as female, with S. squalidus were all self-sterile in contrast to previous studies which reported some self-fertility (Crisp 1972; Ingram 1978). Treating shoots of S. x baxteri plants with colchicine produced ‘chimeric’ plants with allohexaploid branches that produced flower heads that were fully self-fertile. Seed from these flower heads was then used to found the synthetic S. cambrensis lines (S0–S5) used in the transcriptomic and epigenetic analyses described above. The first wholly allohexaploid plants generated from this seed (S0 lines), and their progeny (S1 lines), showed similar ray flower structure and self-compatibility (Hiscock and Hegarty, unpublished). However, in subsequent lines, from the S2 onwards, variation in ray flower form was detected between individuals within and between the nine independent lineages of synthetic S. cambrensis. Some individuals were observed with no ray flowers, some had short partially tubular ray flowers, while in others ray flowers were observed of different lengths and number (Fig. 13.5.) Observations on the progeny of these different individuals showed that these various forms of ray flower are heritable (Hiscock, unpublished). Comparable variation in S. cambrensis ray flower form was previously attributed to intergenomic recombination (Ingram and Noltie 1984), but here we suggest another possibility for such abrupt changes to ray flower phenotype—epigenetic effects associated with the epigenetic instability observed in early-generation synthetic S. cambrensis CO RR 540 F 558 independently formed hybrids can still display significant levels of epigenetic variation. This study, as well as that of Gaeta et al. (2007), was based on a survey of a limited number of loci using cDNA-AFLP or CAPS assay. A later study by Buggs et al. (2011) used the Sequenom MassARRAY allelotyping technology to survey a much larger set of loci in Tragopogon miscellus and confirmed that natural hybrids displayed altered patterns of tissue-specific gene expression, whilst resynthesised hybrids demonstrated relaxed control of tissue-specificity. This latter finding suggests a possible mechanism for the ‘transcriptome shock’ effect we observed in Senecio x baxteri and S. cambrensis, resulting from a loss of tissuespecific expression patterns seen in the parent taxa. Further work will be needed to confirm if this is the case in Senecio. The Sequenom assay used by Buggs et al. (2011) shows the benefits of new molecular tools for studies of allopolyploid systems: with the advent of new technologies for global analysis of DNA methylation such as MSAP and next-generation sequencing (Salmon and Ainouche 2010), it would also be interesting to analyse our later-generation allohexaploid derivatives at a global scale to investigate the longer term changes in methylation as hybrid genomes undergo recombination and adaptation. Such studies may therefore provide further insights into which epigenetic changes are mandated by hybridity, and which may vary between populations and serve as a source of novelty upon which selection may act. 539 UN Editor Proof 262 Layout: T1 Standard SC Chapter No.: 13 Book ISBN: 978-3-642-31441-4 Page: 263/270 Allopolyploid Speciation in Action 263 PR OO F Editor Proof 13 Book ID: 272454_1_En Date: 16-8-2012 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 EC 583 CO RR 582 (Hegarty et al. 2011). Genes shown to be involved in ray flower development in Senecio (RAY1 and RAY2) are orthologues of CYCLOIDEA (CYC) (Kim et al. 2008), which has been shown to occur as a stable mutant (hypermethylated) epiallele that in Linaria (toadflax, Plantaginaceae) manifests itself in a change in flower symmetry from bilateral to radial symmetry (Cubas et al. 1999). We therefore hypothesise that the observed changes in ray flower form may in part be associated with epigenetic modifications to RAY1 and/or RAY2. Testing this hypothesis will be a focus for future work. Another unexpected observation in the synthetic S. cambrensis lines was the appearance of self-sterile individuals in otherwise SC lines, again from the S2 generation onwards. Subsequent analyses using controlled self- and crosspollinations confirmed that these self-sterile individuals possessed functional sporophytic self-incompatibility (Brennan and Hiscock 2010). This is the first time that sporophytic SI has been shown to be inherited and expressed in an allopolyploid and raises intriguing questions about the mechanism regulating this important trait. All S0 and S1 lines of synthetic S. cambrensis were highly selffertile (SC) indicating that the SI system, present in parental S. squalidus, was repressed in these allopolyploids, only to be reactivated/derepressed later. The emergence of SI individuals may be a consequence of recombination or might also be associated with epigenetic changes observed in the early-generation synthetic allopolyploids. Most observations of wild S. cambrensis have reported it to be SC (Abbott and Lowe 2004), but the finding of SI in synthetic S. cambrensis prompted a re-examination of the mating system of wild S. cambrensis. An analysis of selfing rates in 41 wild S. cambrensis individuals from Edinburgh and North Wales UN 580 581 TE D Fig. 13.5 Variation in ray flower morphology observed in individuals of synthetic lines of allohexaploid S. cambrensis from the S2 to S5 generation Layout: T1 Standard SC Chapter No.: 13 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 264/270 M. J. Hegarty et al. 608 identified five SI individuals (Brennan and Hiscock 2010) implying, albeit from a relatively small sample size, that SI may be present in wild S. cambrensis at a frequency of *12 %. This important finding means that S. cambrensis should now be considered as possessing a mixed mating system that has the potential to evolve towards either outcrossing or selfing. 609 13.3 Future Prospects 604 605 606 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 D 614 We are currently engaged in the generation of a draft reference sequence of the Senecio squalidus ‘‘gene-space’’ (low-copy, non-methylated regions of the genome). Nextgeneration sequencing (NGS) platforms enable a variety of potential experiments to examine the consequences of polyploidy and hybridization in Senecio further. For example, we intend to identify promoter regions using chromatin-immunoprecipitation sequencing (ChIPseq) to enrich for DNA fragments bound by enzymes involved in transcription. We can then determine whether hybrids and polyploids display alterations in promoter sequence/binding that may explain the altered patterns of expression observed in our microarray experiments. Further, once a reference sequence is available, we can consider bisulphite sequencing to identify genomic regions which show differential methylation in hybrids relative to their progenitors. One key target for bisulphite sequencing will be the RAY1 and RAY2 genes (Kim et al. 2009) which we suspect may show differential methylation associated with the observed variation in ray flower morphology that appears in synthetic S. cambrensis lines. Identification of small interfering RNAs (siRNAs) and their targets will enable analysis of changes to epigenetic regulation of gene expression in hybrids. At the structural level, resequencing of hybrid ‘‘gene-space’’ and comparative sequence analysis may allow us to detect genomic rearrangements and sequence loss, plus the activity of transposable elements. The increasing capability of genotyping-by-sequencing (GBS) approaches such as restriction-associated DNA (RAD) sequencing (Baird et al. 2008) may also prove useful in detecting structural rearrangements in hybrid genomes. Finally, comparative sequencing of Senecio cambrensis and the two other hybrid derivatives of S. vulgaris and S. squalidus, i.e. S. vulgaris var. hibernicus and S. eboracensis, may enable identification of dosage effects, since the hybrids share parental genomes but differ in the specific combinations thereof. TE 612 613 13.3.1 Next-Generation Approaches to Studying Evolution in S. cambrensis EC 611 CO RR 610 PR OO F 607 UN Editor Proof 264 13.3.2 S. cambrensis in the Wild Senecio cambrensis now exists in the wild only in North Wales, UK, following extinction of the Edinburgh population in 1993 (Abbott and Forbes 2002). AQ6 Layout: T1 Standard SC Chapter No.: 13 Editor Proof 13 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 265/270 Allopolyploid Speciation in Action (a) 265 40 30 25 F 20 15 10 5 0 1982 1984 1987 Year 3500 3000 2500 2000 2002 2003 2004 2003 2004 D 1500 1000 500 0 1982 1983 TE Number of individuals (b) 1983 PR OO Number of sites 35 1984 1987 2002 Year 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 Although the species was recorded at numerous sites and in relatively high numbers in North Wales in the early 1980s (Ingram and Noltie 1995; Fig. 13.6), it has undergone a significant decline in this region since then. In the most recent census of the species undertaken in 2004, it was found at only nine sites in North Wales with a total of 349 individuals recorded across all sites (Abbott et al. 2007; Fig. 13.6). It has been speculated that this decline in numbers has been caused by a reduction of available sites for colonisation by the species (e.g., waste ground) and also to increased use of herbicide to control weed populations (Abbott et al. 2007). Its disappearance from the thin layer of soil that collects along roads between the road edge and verge, where it was found often in the past, is most likely due to increased use of herbicides on plants growing along road sides. However, it has also been noted that plants of the species are frequently heavily infected with the rust Puccinia lagenophorae. Although this rust also infects both parent species, it is possible that its effects on S. cambrensis are more dramatic in terms of production of offspring for colonising new sites, given that numbers of S. squalidus and S. vulgaris that reproduce each year in North Wales and elsewhere in the UK are vast relative to those of S. cambrensis. Although no detailed analysis has been CO RR 641 UN 640 EC Fig. 13.6 Decline in the number of a sites and b flowering individuals across sites, recorded for S. cambrensis in North Wales in different years from 1982 to 2004. Records are from Abbott et al. (2007) and were made in June each year except in 1987 when they were collected in May Layout: T1 Standard SC Chapter No.: 13 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 266/270 M. J. Hegarty et al. 679 680 681 682 683 684 685 Acknowledgments Our research on Senecio cambrensis and its parents has been supported by grants from the NERC, BBSRC and Leverhulme Trust to whom we are grateful. We would also like to acknowledge our long-term collaboration with Keith Edwards and the considerable efforts of former research students, technicians and postdocs who contributed to the research over the past 25 years. These include Alexandra Allen, Paul Ashton, Garry Barker, Tom Batstone, Adrian Brennan, Mark Chapman, David Forbes, Amanda Gillies, Helen Ireland, Juliet James, Joanna Jones, Andrew Lowe, David Tabah, and Ian Wilson. 686 References 687 688 689 690 691 692 693 694 695 696 697 698 699 Abbott RJ, Forbes DG (1993) Outcrossing rate and self-incompatibility in the colonizing species Senecio squalidus L. Heredity 71:155–159 Abbott RJ, Forbes DG (2002) Extinction of the edinburgh lineage of the allopolyploid neospecies, Senecio cambrensis Rosser (Asteraceae). Heredity 88:267–269 Abbott RJ, Lowe AJ (1996) A review of hybridization and evolution in British Senecio In:Hind DJN, Beentje HJ (eds.) Compositae: systematics. Proceedings of the international compositae conference Kew 1994, Royal Botanic Gardens, Kew, pp 679–689 Abbott RJ, Lowe AJ (2004) Origins, establishment and evolution of new polyploid species: Senecio cambrensis and S. eboracensis in the British isles. Biol J Linn Soc 82:467–474 Abbott RJ, Ashton PA, Forbes DG (1992) Introgressive origin of the radiate groundsel Senecio vulgaris L. var. hibernicus Syme: Aat-3 evidence. Heredity 68:425–435 Abbott RJ, Ingram R, Noltie HJ (1983) Discovery of Senecio cambrensis Rosser in edinburgh. Watsonia 14:407–408 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 PR OO 662 D 661 TE 660 EC 659 CO RR 658 F 678 undertaken on whether S. cambrensis is ecologically divergent from its two parents, it tends to grow in sympatry with one or both parent species in the wild. Thus, there is likely competition between the three species for occupation of available open sites, and this might place S. cambrensis at a disadvantage if the seed it produces is less numerous relative to that of S. vulgaris and/or S. squalidus occurring in the same area. Whatever the cause of its marked decline in numbers in North Wales over the last 25 years or so, it is clear that S. cambrensis has reached the stage where its presence in the wild is endangered and that a conservation plan is required to prevent it from possibly becoming extinct in the near future. Given the decline in numbers of S. cambrensis in recent years it is important to gain a better understanding of the nature of its mating system and mating dynamics, especially in the light of our findings of SI in wild populations (Brennan and Hiscock 2010). If there has been a recent shift in mating system from predominantly SC towards SI, it is possible that mating potential has become compromised due to obligate outcrossing. Given the possibility of a single hybrid origin for S. cambrensis in Wales, it is likely that wild populations possess very few S alleles. Whilst a small number of shared S alleles are not necessarily a problem when populations are large, it becomes a problem when populations are in decline and when stochastic effects can lead to S allele loss (Brennan et al. 2003; Pickup and Young 2008). This can then exacerbate decline, leading to an uncontrollable spiral of extinction. A reappraisal of the mating system of S. cambrensis is thus urgently needed. 657 UN Editor Proof 266 Layout: T1 Standard SC Chapter No.: 13 267 EC TE D PR OO F Abbott RJ, Ireland HE, Rogers HJ (2007) Population decline despite high genetic diversity in the new allopolyploid species Senecio cambrensis (Asteraceae). Mol Ecol 16:1023–1033 Abbott RJ, Brennan AC, James JK, Forbes DG, Hegarty MJ, Hiscock SJ (2009) Recent hybrid origin and invasion of the British isles by a self-incompatible species, Oxford ragwort (Senecio squalidus L., Asteraceae). Biol Invasions 11:1145–1158 Abbott RJ, Hegarty MJ, Hiscock SJ, Brennan AC (2010) Homoploid hybrid speciation in action. Taxon 59:1375–1386 Adams KL, Percifield R, Wendel JF (2004) Organ-specific silencing of genes in a newly synthesized cotton allotetraploid. Genetics 168:2217–2226 Adams KL, Wendel JF (2005) Polyploidy and genome evolution in plants. Curr Opin Plant Biol 8:135–141 Ashton PA, Abbott RJ (1992) Multiple origins and genetic diversity in the newly arisen allopolyploid species, Senecio cambrensis Rosser (Compositae). Heredity 68:25–32 Baird NA, Etter PD, Atwood TS, Currey MC, Shiver AL, Lewis ZA, Selker WA, Cresko WA, Johnson EA (2008) Rapid SNP discovery and genetic mapping using sequenced RAD markers. PLoS ONE 3:e3376 Baker HG (1967) Support for Baker’s’ Law—as a rule. Evolution 21:853–856 Benoit PM, Crisp PC, Jones BMG (1975) Senecio L. In: Stace CA (ed) Hybridization and the flora of the British isles. Academic, London, pp 404–410 Brennan AC, Hiscock SJ (2010) Expression and inheritance of sporophytic self-incompatibility in synthetic allohexaploid Senecio cambrensis (Asteraceae). New Phytol 186:251–261 Brennan ACE, Harris SA, Tabah DA, Hiscock SJ (2002) The population genetics of sporophytic self-incompatibility in Senecio squalidus L. (Asteraceae) I: S allele diversity in a natural population. Heredity 89:430–438 Brennan ACE, Harris SA, Hiscock SJ (2003) The population genetics of sporophytic selfincompatibility in Senecio squalidus L. (Asteraceae): avoidance of mating constraints imposed by low S-allele number. Phil Trans R. Soc Lond B 358:1047–1050 Brennan AC, Bridle JR, Wang A-L, Hiscock SJ, Abbott RJ (2009) Adaptation and selection in the Senecio (Asteraceae) hybrid zone on Mount Etna, Sicily. New Phytol 183:702–717 Brennan ACE, Harris SA, Hiscock SJ (2006) The population genetics of sporophytic selfincompatibility in Senecio squalidus L. (Asteraceae): S allele diversity across the British range. Evolution 60:213–224 Brennan ACE, Harris SA, Hiscock SJ (2005) Modes and rates of selfing and associated inbreeding depression in the self-incompatible plant Senecio squalidus (Asteraceae): a successful colonizing species in the British Isles. New Phytol 168:475–486 Brennan AC, Barker D, Hiscock SJ, Abbott RJ (2012) Molecular genetic and quantitative trait divergence associated with recent homoploid hybrid speciation: a study of Senecio squalidus (Asteraceae). Heredity 109:87–95 Buggs RJA, Soltis PS, Mavrodiev EV, Symonds VV, Soltis DE (2008) Does phylogenetic distance between parental genomes govern the success of polyploids? Castanea 73:74–93 Buggs RJA, Doust AN, Tate JA, Koh J, Soltis K, Feltus FA, Paterson A, Soltis PS, Soltis DE (2009) Gene loss and silencing in Tragopogon miscellus (Asteraceae): comparison of natural and synthetic allotetraploids. Heredity 103:73–81 Buggs RJA, Zhang L, Miles N, Tate JA, Gao L, Wei W, Schnable PS, Barbazuk WB, Soltis PS, Soltis DE (2011) Transcriptomic shock generates evolutionary novelty in a newly formed natural allopolyploid plant. Curr Biol 21:551–556 Chaudhary B, Flagel L, Stupar RM, Udall JA, Verma N, Springer NM, Wendel JH (2009) Reciprocal silencing, transcriptional bias and functional divergence of homeologs in polyploid cotton (gossypium). Genetics 182:503–517 Chapman MA, Burke JM (2007) Genetic divergence and hybrid speciation. Evolution 61:1773–1780 Coleman M, Forbes DG, Abbott RJ (2001) A new subspecies of Senecio mohavensis (Compositae) reveals an old–new World species disjunction. Edinburgh J Botany 58:384–403 Crisp PC (1972) Cytotaxonomic studies in the section Annui of Senecio. Ph.D Thesis, University of London CO RR 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 Book ISBN: 978-3-642-31441-4 Page: 267/270 Allopolyploid Speciation in Action UN Editor Proof 13 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 13 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 268/270 EC TE D PR OO F Cubas P, Vincent C, Coen E (1999) An epigenetic mutation responsible for natural variation in floral symmetry. Nature 401:157–161 Dong YZ, Liu ZL, Shan XH, Qiu T, He MY, Liu B (2005) Allopolyploidy in wheat induces rapid and heritable alterations in DNA methylation patterns of cellular genes and mobile elements. Genetika 41:1089–1095 Doyle JJ, Flagel LE, Paterson AH, Rapp RA, Soltis DE, Soltis PS, Wendel JF (2008) Evolutionary genetics of genome merger and doubling in plants. Annu Rev Genet 42:443–461 Feldman M, Levy AA (2005) Allopolyploidy—a shaping force in the evolution of wheat genomes. Cytogenet Genome Res 109:250–258 Garcı0 a-Ortiz MV, Ariza RR, Rolda0 n-Arjona T (2001) An OGG1 orthologue encoding a functional 8-oxoguanine DNA glycosylase/lyase in Arabidopsis thaliana. Plant Mol Biol 47:795–804 Gaeta RT, Pires JC, Iniguez-Luy F, Leon E, Osborn TC (2007) Genomic changes in resynthesized Brassica napus and their effect on gene expression and phenotype. Plant Cell 19:3403–3417 Gray AJ, Marshall DF, Raybould AF (1991) A century of evolution in Spartina anglica. Adv Ecol Res 21:1–62 Grant V (1981) Plant speciation. Columbia University Press, New York Harland SC (1955) The experimental approach to the species problem. In: Lousley JE (ed) Species studies in the British flora. Botanical Society of the British Isles, London, pp 16–20 Harris SA (2002) Introduction of Oxford ragwort, Senecio squalidus L. (Asteraceae), to the United Kingdom. Watsonia 24:31–43 Harris SA, Ingram R (1992) Molecular systematics of the genus Senecio L. I. hybridization in a British polyploid complex. Heredity 69:1–10 Hegarty MJ, Hiscock SJ (2008) Genomic clues to the evolutionary success of polyploid plants. Curr Biol 18:R435–R444 Hegarty MJ, Jones JM, Wilson ID, Barker GL, Coghill JA, Sanchez-Baracaldo P, Liu G, Buggs RJA, Abbott RJ, Edwards KJ, Hiscock SJ (2005) Development of anonymous cDNA microarrays to study changes to the Senecio floral transcriptome during hybrid speciation. Mol Ecol 14:2493–2510 Hegarty MJ, Barker GL, Wilson ID, Abbott RJ, Edwards KJ, Hiscock SJ (2006) Transcriptome shock after interspecific hybridization in Senecio is ameliorated by genome duplication. Curr Biol 16:1652–1659 Hegarty MJ, Barker GL, Brennan AC, Edwards KJ, Abbott RJ, Hiscock SJ (2008) Changes to gene expression associated with hybrid speciation in plants: further insights from transcriptomic studies in Senecio. Philos Trans R Soc Lond B 363:3055–3069 Hegarty MJ, Batstone T, Barker GL, Edwards KJ, Abbott RJ, Hiscock SJ (2011) Nonadditive changes to cytosine methylation as a consequence of hybridization and genome duplication in Senecio (Asteraceae). Mol Ecol 20:105–113 Hiscock SJ (2000a) Genetic control of self-incompatibility in Senecio squalidus L. (Asteraceae) – a successful colonising species. Heredity 84:10–19 Hiscock SJ (2000b) Self-incompatibility in Senecio squalidus L. (Asteraceae). Ann Bot 85(Supplement A):181–190 Ingram R (1978) The genomic relationship of Senecio squalidus L. and Senecio vulgaris L. and the significance of genomic balance in their hybrid S. x baxteri Druce. Heredity 40:459–462 Ingram R, Noltie HJ (1984) Ray floret morphology and the origin of variability in Senecio cambrensis Rosser, a recently established allopolyploid species. New Phytol 96:601–607 Ingram R, Noltie HJ (1989) Early adjustment of patterns of metaphase association in the evolution of polyploid species. Genetica 78:21–24 Ingram R, Noltie HJ (1995) Biological flora of the British isles: Senecio cambrensis Rosser. J Ecol 83:537–546 James JK, Abbott RJ (2005) Recent, allopatric, homoploid hybrid speciation: the origin of Senecio squalidus (Asteraceae) in the British Isles from a hybrid zone on Mount Etna, Sicily. Evolution 59:2533–2547 CO RR 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768 769 770 771 772 773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792 793 794 795 796 797 798 799 800 801 802 803 804 805 806 807 M. J. Hegarty et al. UN Editor Proof 268 Layout: T1 Standard SC Chapter No.: 13 269 EC TE D PR OO F Jiao L, Wickett NJ, Ayyampalayam S, Chanderbali et al (2011) Ancestral polyploidy in seed plants and angiosperms. Nature 473:97–100 Kadereit JW, Uribe-Convers S, Westberg E, Comes HP (2006) Reciprocal hybridization at different times between Senecio flavus and Senecio glaucus gave rise to two polyploidy species in north Africa and south-west Asia. New Phytol 169:431–441 Kent DH (1963) Senecio squalidus L. in the British isles. 7 Wales. Nat Wales 8:175–178 Kim M, Cui M-L, Cubas P, Gillies A, Lee K, Chapman MA, Abbott RJ, Coen E (2008) Regulatory genes control a key morphological and ecological trait transferred between species. Science 322:1116–1119 Leitch IJ, Bennett MD (1997) Polyploid in angiosperms. Trends Plant Sci 2:470–476 Leitch AR, Leitch IJ (2008) Genomic plasticity and diversity of polyploid plants. Science 320:481–483 Liu Z, Adams KL (2007) Expression partitioning between genes duplicated by polyploidy under abiotic stress and during organ development. Curr Biol 17:1669–1674 Liu B, Brubaker CL, Mergeai G, Cronn RC, Wendel JF (2001) Polyploid formation in cotton is not accompanied by rapid genomic changes. Genome 44:321–330 Lowe AJ, Abbott RJ (2000) Routes of origin of two recently evolved hybrid taxa: Senecio vulgaris var. hibernicus and York radiate groundsel (Asteraceae). Am J Bot 87:1159–1167 Lowe AJ, Abbott RJ (2003) A new British species, Senecio eboracensis (Asteraceae), another hybrid derivative of S. vulgaris L. and S. squalidus L. Watsonia 24:375–388 Lukens LN, Pires JC, Leon E, Vogelzang R, Oslach L, Osborn T (2006) Patterns of sequence loss and cytosine methylation within a population of newly resynthesized Brassica napus allopolyploids. Plant Physiol 140:336–348 McClintock B (1984) The significance of responses of the genome to challenge. Science 226:792–801 Madlung A, Masuelli RW, Watson B, Reynolds SH, Davison J, Comai L (2002) Remodeling of DNA methylation and phenotypic and transcriptional changes in synthetic Arabidopsis allotetraploids. Plant Physiol 129:733–746 Marshall DF, Abbott RJ (1980) On the frequency of introgression of the radiate (Tr) allele from Senecio squalidus L. into Senecio vulgaris L. Heredity 45:133–135 Moffatt BA, Allen M, Snider S, Pereira LA, Todorova M, Summers PS, Weretilnyk EA, MartinMcCaffrey L, Wagner C (2002) Adenosine kinase deficiency is associated with developmental abnormalities and reduced transmethylation. Plant Physiol 128:812–821 Mull L, Ebbs ML, Bender J (2006) A histone methylation-dependent DNA methylation pathway is uniquely impaired by deficiency in Arabidopsis S-adenosylhomocysteine hydrolase. Genetics 174:1161–1171 Novak SJ, Soltis DE, Soltis PS (1991) Ownbey’s Tragopogons: 40 years later. Am J Bot 78:1586–1600 Otto SP, Whitton J (2000) Polyploidy: incidence and evolution. Annu Rev Genet 34:401–437 Parisod C, Salmon A, Zerjal T, Tenaillon M, Grandbastien MA, Ainouche M (2010) Rapid structural and epigenetic reorganization near transposable elements in hybrid and allopolyploid genomes in Spartina. New Phytol 184:1003–1015 Paun O, Forest F, Fay MF, Chase MW (2009) Hybrid speciation in angiosperms: parental divergence drives ploidy. New Phytol 182:507–518 Pelser PB, Nordenstam B, Kadereit JW, Watson LE (2007) An ITS phylogeny of tribe Senecioneae (Asteraceae) and a new delimitation of Senecio L. Taxon 56:1062–1077 Pelser PB, Abbott RJ, Comes HP, Milton JJ, Möller M, Looseley ME, Cron GV, Barcelona JF, Kennedy AH, Watson LE, Barone R, Hernández F, Kadereit JW (2012) The genetic ghost of an invasion past: colonization and extinction revealed by historical hybridization in Senecio. Mol Ecol 21:369–387 Pereira LA, Schoor S, Goubet F, Dupree P, Moffatt BA (2006) Deficiency of adenosine kinase activity affects the degree of pectin methyl-esterification in cell walls of Arabidopsis thaliana. Planta 224:1401–1414 Pickup M, Young AG (2008) Population size, self-incompatibility and genetic rescue in diploid and tetraploid races of Rutidosis leptorrhynchoides (Asteraceae). Heredity 100:268–274 CO RR 808 809 810 811 812 813 814 815 816 817 818 819 820 821 822 823 824 825 826 827 828 829 830 831 832 833 834 835 836 837 838 839 840 841 842 843 844 845 846 847 848 849 850 851 852 853 854 855 856 857 858 859 860 861 Book ISBN: 978-3-642-31441-4 Page: 269/270 Allopolyploid Speciation in Action UN Editor Proof 13 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 13 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 270/270 EC TE D PR OO F Pumphrey M, Bai J, Laudencia-Chingcuanco D, Anderson O, Gill BS (2009) Nonadditive expression of homoeologous genes is established upon polyploidization in hexaploid wheat. Genetics 181:1147–1157 Rieseberg LH, Raymond O, Rosenthal DM, Lai Z, Livingstone K, Nakazato T, Durphy JL, Schwarzbach AE, Donovan LA, and Lexer C (2003) Major ecological transitions in wild sunflowers facilitated by hybridization. Science 301:1211–1216 Rieseberg LH, Willis JH (2007) Plant speciation. Science 317:910–914 Rieseberg LH, Archer MA, Wayne RK (1999) Transgressive segregation, adaptation and speciation. Heredity 83:363–372 Rosser EM (1955) A new British species of Senecio. Watsonia 3:228–232 Salone V, Rudinger M, Polsakiewicz M, Hoffmann B, Groth-Malonek M, Szurek B, Small I, Knoop V, Lurin C (2007) A hypothesis on the identification of the editing enzyme in plant organelles. FEBS Lett 581:4132–4138 Salmon A, Ainouche ML (2010) Polyploidy and DNA methylation: new tools available. Mol Ecol 19:213–215 Salmon A, Ainouche ML, Wendel JF (2005) Genetic and epigenetic consequences of recent hybridization and polyploidy in Spartina (Poaceae). Mol Ecol 14:1163–1175 Soltis PS, Soltis DE (1999) Polyploidy: recurrent formation and genome evolution. Trends Ecol Evol 14:348–352 Stebbins GL (1957) Self fertilization and population variability in higher plants. Am Nat 91:337–354 Stupar RM, Hermanson PJ, Springer NM (2007) Nonadditive expression and parent-of-origin effects identified by microarray and allele-specific expression profiling of maize endosperm. Plant Physiol 145:411–425 Sun M, Ritland K (1998) Mating system of yellow starthistle (Centaurea solstitialis), a successful colonizer in North America. Heredity 80:225–232 Urbanska KM, Hurka H, Landolt E, Neuffer B, Mummenhoff K (1997) Hybridization and evolution in Cardamine (Brassicaceae) at Urnerboden, central Switzerland: biosystematics and molecular evidence. Plant Syst Evol 204:233–256 Vincent PLD (1996) Progress on clarifying the generic concept of Senecio based on an extensive world-wide sample of taxa. In Hind DJN, Beentje HJ (eds) Compositae: systematics. proceedings of the international compositae conference Kew 1994, vol 1, Royal Botanic Gardens, Kew, pp 597–611 Wang J, Tian L, Lee H-Y, Chen ZJ (2006a) Nonadditive regulation of FRI and FLC loci mediates flowering-time variation in Arabidopsis allopolyploids. Genetics 173:965–974 Wang J, Tian L, Lee H-S, Wei NE, Jiang H, Watson B, Madlung A, Osborn TC, Doerge RW, Comai L, Chen ZJ (2006b) Genomewide nonadditive gene regulation in Arabidopsis allotetraploids. Genetics 172:507–517 Weir J, Ingram R (1980) Ray morphology and cytological investigations of Senecio cambrensis Rosser. Heredity 86:237–241 Wood TE, Takebayashi N, Barker MS, Mayrose I, Greenspoon PB, Rieseberg LH (2009) The frequency of polyploid speciation in flowering plants. Proc Nat Acad Sci USA 106:13875–13879 Xiong LZ, Xu CG, Saghai Maroof MA, Zhang Q (1999) Patterns of cytosine methylation in an elite rice hybrid and its parental lines, detected by a methylation-sensitive amplification polymorphism technique. Mol Gen Genet 261:439–446 CO RR 862 863 864 865 866 867 868 869 870 871 872 873 874 875 876 877 878 879 880 881 882 883 884 885 886 887 888 889 890 891 892 893 894 895 896 897 898 899 900 901 902 903 904 905 906 M. J. Hegarty et al. UN Editor Proof 270 13 F Chapter No.: PR OO Editor Proof Author Queries Details Required Author’s Response AQ1 Please provide the index term for this chapter. AQ2 ‘Rieseberg et al (2003)’ has been changed to ‘Rieseberg et al. (1999)’ so that this citation matches the list. AQ3 References ’ Hegarty and Hiscock (2002); James and Abbott (2005); Ingram and Noltie (1989); Hegarty et al. 2009; Hiscock and Hegarty, unpublished are cited in the text but not provided in the list please provide it or delete these citations AQ5 McClintock 1986 has been changed to McClintock 1984 so that this citation matches the list. AQ6 ‘Baird et al. (2008)’ has been changed to ‘Baird et al. (2009)’ so that this citation matches the list. AQ7 Kindly check and confirm the updated volume id, issue id and page range for the reference ’Brennan et al (2012)’ AQ8 Kindly check and confirm the updated year, volume, issue id and page range for the reference ‘Pelser etal. (2011)’. UN CO RR EC TE D Query Refs. Metadata of the chapter that will be visualized in SpringerLink Book Title Polyploidy and Genome Evolution Series Title Chapter Title The Early Stages of Polyploidy: Rapid and Repeated Evolution in Tragopogon Copyright Year 2012 Copyright HolderName Springer-Verlag Berlin Heidelberg Corresponding Author Family Name Soltis Particle Given Name Douglas E. Suffix Author Division Department of Biology Organization University of Florida Address 32611, Gainesville, FL, USA Email dsoltis@botany.ufl.edu Family Name Buggs Particle Given Name Richard J. A. Suffix Division Department of Biology Organization University of Florida Address 32611, Gainesville, FL, USA Division Organization School of Biological and Chemical Sciences, Queen Mary University of London Address E1 4NS, London, UK Email Author Family Name Barbazuk Particle Given Name W. Brad Suffix Division Department of Biology Organization University of Florida Address 32611, Gainesville, FL, USA Email Author Family Name Chamala Particle Given Name Srikar Suffix Division Department of Biology Organization University of Florida Address 32611, Gainesville, FL, USA Email Author Family Name Chester Particle Given Name Michael Suffix Division Department of Biology Organization University of Florida Address 32611, Gainesville, FL, USA Email Author Family Name Gallagher Particle Given Name Joseph P. Suffix Division Department of Biology Organization University of Florida Address 32611, Gainesville, FL, USA Division Department of Ecology, Evolution, and Organismal Biology Organization Iowa State University Address 50011, Ames, IA, USA Email Author Family Name Schnable Particle Given Name Patrick S. Suffix Division Organization Center for Plant Genomics, Iowa State University Address 50011, Ames, IA, USA Email Author Family Name Soltis Particle Given Name Pamela S. Suffix Division Organization Florida Museum of Natural History, University of Florida Address 32611, Gainesville, FL, USA Email Abstract Elucidating the causes and consequences of polyploidy (whole-genome duplication; WGD) is arguably central to understanding the evolution of most eukaryotic lineages. However, much of what we know about these processes is derived from the study of crops and synthetic polyploids. Tragopogon provides the unique opportunity to investigate the genetic and genomic changes that occur across an evolutionary series from F1 hybrids, synthetic allopolyploids, independently formed natural populations of T. mirus and T. miscellus that are 60–80 years post-formation, to older Eurasian polyploids that are dated by molecular clocks at several million years old, and finally to a putative ancient polyploidization thought to have occurred prior to or early in the history of the Asteraceae (40–43 mya). Tragopogon joins other well-studied natural polyploid systems (e.g., Glycine, Nicotiana, Gossypium, Spartina, Senecio), but presents a range of research possibilities that is not available in any other system. We have shown in T. mirus and T. miscellus that upon allopolyploidization, massive gene loss occurs in patterns that are repeated across populations of independent origin and with a bias against genes derived from T. dubius, the diploid parent shared by both new allotetraploids. We have also shown significant changes in gene expression (transcriptomic shock) in the early generations of allopolyploidy in these species. Massive and repeated patterns of chromosomal variation (intergenomic translocations and aneuploidy) have been revealed by fluorescence in situ hybridization. Aneuploidy results in substitutions between homeologous chromosomes, through reciprocal monosomytrisomy (1:3 copies) or nullisomy-tetrasomy (0:4 copies). We propose that substantial chromosomal instability results in karyotype restructuring, a likely common process following WGD and a driver of allopolyploid speciation, which has largely unexplored implications for gene losses, gains, and expression patterns. But gene loss and expression changes as well as karyotypic changes are ongoing in T. mirus and T. miscellus, in that no population is fixed for any of these events; thus, we have literally caught evolution in the act. 6 7 8 9 10 11 12 13 14 15 16 17 18 F Douglas E. Soltis, Richard J. A. Buggs, W. Brad Barbazuk, Srikar Chamala, Michael Chester, Joseph P. Gallagher, Patrick S. Schnable and Pamela S. Soltis PR OO 5 Abstract Elucidating the causes and consequences of polyploidy (whole-genome duplication; WGD) is arguably central to understanding the evolution of most eukaryotic lineages. However, much of what we know about these processes is derived from the study of crops and synthetic polyploids. Tragopogon provides the unique opportunity to investigate the genetic and genomic changes that occur across an evolutionary series from F1 hybrids, synthetic allopolyploids, independently formed natural populations of T. mirus and T. miscellus that are 60–80 years post-formation, to older Eurasian polyploids that are dated by molecular clocks at several million years old, and finally to a putative ancient polyploidization thought to have occurred prior to or early in the history of the Asteraceae (40–43 mya). Tragopogon joins other wellstudied natural polyploid systems (e.g., Glycine, Nicotiana, Gossypium, Spartina, Senecio), but presents a range of research possibilities that is not available in any other D 4 The Early Stages of Polyploidy: Rapid and Repeated Evolution in Tragopogon TE 3 Chapter 14 EC 2 Book ISBN: 978-3-642-31441-4 Page: 271/291 D. E. Soltis (&)  R. J. A. Buggs  W. B. Barbazuk  S. Chamala  M. Chester  J. P. Gallagher Department of Biology, University of Florida, Gainesville, FL 32611, USA e-mail: dsoltis@botany.ufl.edu CO RR 1 Book ID: 272454_1_En Date: 16-8-2012 Present Address: J. P. Gallagher Department of Ecology, Evolution, and Organismal Biology, Iowa State University, Ames, IA 50011, USA R. J. A. Buggs School of Biological and Chemical Sciences, Queen Mary University of London, London, E1 4NS, UK P. S. Schnable Center for Plant Genomics, Iowa State University, Ames, IA 50011, USA UN Editor Proof Layout: T1 Standard SC Chapter No.: 14 P. S. Soltis Florida Museum of Natural History, University of Florida, Gainesville, FL 32611, USA P. S. Soltis and D. E. Soltis (eds.), Polyploidy and Genome Evolution, DOI: 10.1007/978-3-642-31442-1_14,  Springer-Verlag Berlin Heidelberg 2012 271 Layout: T1 Standard SC Chapter No.: 14 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 272/291 23 24 25 26 27 28 29 30 31 32 33 F 21 22 system. We have shown in T. mirus and T. miscellus that upon allopolyploidization, massive gene loss occurs in patterns that are repeated across populations of independent origin and with a bias against genes derived from T. dubius, the diploid parent shared by both new allotetraploids. We have also shown significant changes in gene expression (transcriptomic shock) in the early generations of allopolyploidy in these species. Massive and repeated patterns of chromosomal variation (intergenomic translocations and aneuploidy) have been revealed by fluorescence in situ hybridization. Aneuploidy results in substitutions between homeologous chromosomes, through reciprocal monosomy-trisomy (1:3 copies) or nullisomy-tetrasomy (0:4 copies). We propose that substantial chromosomal instability results in karyotype restructuring, a likely common process following WGD and a driver of allopolyploid speciation, which has largely unexplored implications for gene losses, gains, and expression patterns. But gene loss and expression changes as well as karyotypic changes are ongoing in T. mirus and T. miscellus, in that no population is fixed for any of these events; thus, we have literally caught evolution in the act. PR OO 19 20 D. E. Soltis et al. 35 14.1 Introduction 36 14.1.1 General Introduction 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 Polyploidy, or whole-genome duplication (WGD), is currently recognized as a major evolutionary force in eukaryotes (e.g., Mable 2003; Gregory and Mable 2005; Mable et al. 2011). Polyploidy generally results in instant speciation, increasing biodiversity and providing new genetic material for evolution (e.g., Levin 1983, 2002). Some evidence suggests that two polyploid events occurred in ancestors of vertebrates (Ohno 1970; Panapoulou and Poustka 2005; see Chap. 16 of this volume), with subsequent polyploidy in amphibians and fish (Mable et al. 2011; see Chaps. 18 and 17, respectively, this volume). The genomes of yeast and other Saccharomyces also appear to be anciently duplicated (Wolfe and Shields 1997; Kellis et al. 2004; Dujon et al. 2004; see Chap. 15 of this volume). Researchers have long recognized that polyploidy is an inseparable part of angiosperm biology; the polyploidy process has, in fact, been studied in plants for a little over a century. Early reviews of polyploidy in plants include now classic papers by Müntzing (1936), Darlington (1937), Clausen et al. (1945), Löve and Löve (1949), and Stebbins (1950, 1971). However, these seminal works did not anticipate the huge role for polyploidy in evolution that genomic studies now suggest. During the past decade there has been a tremendous resurgence of interest in polyploidy, stimulated in large part by the development of increasingly powerful genetic and genomic tools. The result has been numerous new insights into the genomic and genetic consequences of polyploidy (other chapters of this volume). Recent discoveries have dramatically reshaped traditional views and concomitantly revealed that polyploidy is a highly dynamic and ubiquitous process. For example, studies of many duplicated genes across genomes suggest that all EC 38 CO RR 37 TE D 34 UN Editor Proof 272 Layout: T1 Standard SC Chapter No.: 14 Book ISBN: 978-3-642-31441-4 Page: 273/291 The Early Stages of Polyploidy 273 82 14.1.2 Introduction to Tragopogon 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 PR OO 65 D 64 TE 63 EC 61 62 Tragopogon now provides a well-known, textbook example of recent allopolyploid speciation; two new allotetraploid species originated within the last 80 years (Soltis et al. 2004, 2009b) (Fig. 14.1). Tragopogon mirus and T. miscellus (Fig. 14.1) formed repeatedly following the introduction of three diploids from Europe into the Palouse region of North America in the early 1900s (Ownbey 1950; Soltis et al. 2004); the tetraploids have not formed in Europe. The parentage of T. mirus (T. dubius and T. porrifolius) and T. miscellus (T. dubius and T. pratensis) is well documented (Figs. 14.1 and 14.2) and confirmed with multiple markers and approaches (Soltis et al. 1995, 2004, 2009b). Ownbey (1950) described the few populations of the newly formed allotetraploids (each consisting of fewer than 100 individuals) as ‘‘small and precarious’’, but noted that they had ‘‘attained a degree of success’’ and they appeared to be ‘‘competing successfully’’ with their diploid parents. He also stated that it would be ‘‘important to follow the ecological development of the newly formed polyploids’’ through time. Significantly, both tetraploids have been highly successful since their formation. Novak et al. (1991) conducted a survey to determine the distributions of the two polyploids 40 years after Ownbey’s discovery. One or CO RR 60 F 81 angiosperms have undergone at least one round of genome doubling. Significantly, genomic and phylogenetic analyses also associate polyploidy with major diversifications (e.g., within Poales, Solanaceae, Fabaceae; Soltis et al. 2009a), and the origin of angiosperms and seed plants (Jiao et al. 2011). As a result of a diverse array of studies, we have learned a great deal about the interactions that occur among the diploid genomes forced together via allopolyploidy. Because of its apparent prevalence, elucidating the causes and consequences of polyploidy is arguably central to understanding the origin and diversification of most major lineages of eukaryotes. Significantly, however, most of what we know about the genetic and genomic consequences of polyploidy is derived from the study of synthetic polyploids, crops (e.g., cotton, wheat, Brassica, Nicotiana), and genetic models (Arabidopsis) (see refs above, Chen et al. 2004; Soltis and Soltis 2009). To understand better how polyploidization impacts genome evolution and gene function in natural populations, we must extend from a few crops, genetic models, and synthetics to naturally occurring polyploids. Three systems are known that permit insights into the early stages of polyploidy in nature: Spartina anglica (Ainouche et al. 2004, 2009; Salmon et al. 2005; see Chap. 12 of this volume), Senecio cambrensis (Ashton and Abbott 1992; Abbott and Lowe 2004; Hegarty et al. 2005, 2006; see Chap. 13, this volume), and Tragopogon (T. mirus, T. miscellus, Soltis et al. , 2009b). Research has progressed on all, and they are complementary. In addition, a2004s stressed here, Tragopogon also affords unique opportunities to investigate polyploidy over a continuum of ages, as well as the consequences of recent and frequently repeated polyploidy. 59 UN Editor Proof 14 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 14 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 274/291 D. E. Soltis et al. 106 14.1.3 Tragopogon as a Unique Evolutionary Model 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 PR OO 109 110 For many reasons Tragopogon is a novel system that affords the opportunity to examine the early stages of polyploidization in nature. The natural populations are approximately 80 years (40 generations in these biennials) old—this time frame and the fact that they have experienced natural selection provide a window into polyploidization that cannot be matched via the study of synthetic polyploids (e.g., crops and genetic models). Furthermore, molecular studies suggest that the diploid parents diverged *2.5–5 MYA and that there may be as many as 21 lineages of separate origin of T. miscellus and 11 of T. mirus just in the Palouse (Soltis et al. 1995, 2004, 2009b; Symonds et al. 2010); the polyploids have also formed in Arizona, Oregon, Wyoming, and Montana (Soltis et al. 2012; Ownbey unpublished data). Recent studies employing microsatellite markers (Symonds et al. 2010) reveal multiple origins on a small geographic scale (see Sect. 14.1.4). Given that multiple polyploidizations are common in plants (Soltis and Soltis 1993, 1999, 2000, 2009), Tragopogon represents in microcosm what occurs in other polyploids over much larger geographic areas and longer time frames. These repeated origins in a small geographic area and a narrow time frame also provide the unique opportunity to ask if evolution repeats itself across these many lineages. Adding to the utility of the Tragopogon system as an evolutionary model is the recent production of multiple synthetic lines of both T. mirus and T. miscellus (Tate et al. 2009a), providing the added opportunity of examining both species from polyploidization onward. Morphologically, the synthetics resemble the natural polyploids with short- and long-liguled forms of T. miscellus resulting when T. pratensis and T. porrifolius are reciprocally crossed (Tate et al. 2009a). In nature, all formations of T. mirus have T. porrifolius as the maternal parent and T. dubius as the paternal parent, but we have synthesized T. mirus reciprocally. We also produced allotetraploids between T. porrifolius and T. pratensis, which are not known from nature (Fig. 14.1). All of these synthetic lines are now in the fourth generation and offer the unique opportunity for comparative study of repeated formations of both natural and synthetic polyploids. Further adding to the allure of Tragopogon are Old World polyploids that are much older than the recently formed New World polyploids. Tragopogon comprises *150 species, 12 of which are Eurasian polyploids (Mavrodiev et al. 2008a). Of several Eurasian polyploids for which we clarified parentage (Mavrodiev et al. 2008a, b, c), T. castellanus (2n = 24) from Spain (Blanca and Díaz de la Guardia 1996) has emerged as a promising new model. Estimating the D 107 108 TE 104 EC 102 103 CO RR 101 F 105 both polyploids were found in most towns of the Palouse with populations ranging from small (fewer than 100 individuals) to many thousands of individuals. Tragopogon miscellus is now one of the most common weeds in and around Spokane, WA, as well as in Moscow, ID, and Spangle, WA. Populations of T. mirus and T. miscellus often form dense stands and are, in fact, displacing their parents, particularly T. pratensis and T. porrifolius. 100 UN Editor Proof 274 Layout: T1 Standard SC Chapter No.: 14 Book ISBN: 978-3-642-31441-4 Page: 275/291 The Early Stages of Polyploidy 275 EC TE D PR OO F Editor Proof 14 Book ID: 272454_1_En Date: 16-8-2012 142 143 144 145 146 147 148 date of origin of a polyploid is difficult, but given well-known caveats, we used DNA sequence data to estimate the age of T. castellanus as 0.8–2.8 million years (Mavrodiev et al. 2008a, b, c). This agrees with other evidence in that T. castellanus occurs in well-known Pleistocene glacial refugia that harbor other paleoendemics (e.g., Petit et al. 2003). Thus, in Tragopogon we have the opportunity to extend our analyses to older Eurasian polyploids, providing a continuum of ages from F1 hybrids and raw synthetics, to 80-year-old natural polyploids, to natural polyploids that are perhaps several million years old. UN 141 CO RR Fig. 14.1 Summary of parentage of tetraploid Tragopogon species comparing natural with synthetic allopolyploids. The diploid parents (with 2n = 12) are at the corners of the triangle; polyploids (2n = 24) are in between the corners. Synthetic polyploids are on the outside of the triangle (connected by white lines); those polyploids forming naturally are to the inside of the triangle. Polyploids are placed closer to the maternal parent. In nature, T. miscellus has formed reciprocally, and T. mirus has formed only with T. porrifolius as the maternal parent. However, we have made reciprocal synthetic lines of both and have also made reciprocal polyploids of T. pratensis 9 T. porrifolius; this polyploid has not formed in nature. Note that populations of T. miscellus of reciprocal origin differ in morphology. Those with T. pratensis as the maternal parent have short ligules, and those with T. dubius as the maternal parent have long ligules Layout: T1 Standard SC Chapter No.: 14 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 276/291 149 D. E. Soltis et al. 14.1.4 Origins of Species 174 14.2 Does Evolution Repeat Itself? 156 157 158 159 160 161 162 163 164 165 166 167 168 169 170 171 172 175 176 177 178 179 180 181 182 183 184 185 186 187 PR OO 155 D 154 TE 153 EC 152 CO RR 151 F 173 The Tragopogon system is comparable in some ways to an island biogeography scenario. The diploids and the derivative polyploids only occur in small towns that are scattered across the Palouse area of eastern Washington and adjacent Idaho and not in the intervening areas, which are large tracts of agricultural land. This begged the question—how did the polyploids spread so quickly to many towns? Certainly, seed dispersal is one possibility given the wind-dispersed nature of the achenes. Molecular data indicate instead that repeated formation has played the major role in range expansion. This was suspected by Ownbey and McCollum (1953), who surmised that T. miscellus had formed reciprocally with the long-liguled form (found only in Pullman, WA) having T. dubius as the maternal parent and all other populations having T. pratensis as the maternal parent; this reciprocal parentage results in distinctive morphologies (Fig. 14.1) and was later confirmed by molecular methods (Soltis and Soltis 1989). A suite of molecular markers including allozymes, AFLPs, sequence data, and microsatellites has now revealed that most populations of the allotetraploids are of distinct origin (Soltis et al. 2004; Symonds et al. 2010). Furthermore, microsatellites have documented multiple origins on a fine geographic scale, revealing in several cases that distinct populations in the same town separated by only 1–2 km are of separate origin (Symonds et al. 2010). Interestingly, microsatellite data reveal that of the many genotypes of T. dubius currently in nature, only three general types appear to have contributed to the repeated formations of both polyploids, and there are no exact matches to present-day T. dubius genotypes. Hence, the genotypes detected in the two polyploid species appear to represent a snapshot of the historical population structure in the diploid progenitors, rather than modern diploid genotypes. 150 Evolutionary biologists have long wondered if evolution would repeat itself, given the chance. Gould (1994) suggested that, on a broad evolutionary scale, if we could replay the evolutionary tape of life on Earth, it would play differently— ‘‘history involves too much chaos,’’ and too many chance events are involved for the evolutionary process to be repetitive. He stated that ‘‘chains of historical events are so intricate, so imbued with random and chaotic elements, so unrepeatable in encompassing such a multitude of unique objects, that standard models of simple prediction and replication do not apply.’’ In contrast, other researchers have argued that ‘‘within certain limits the outcome of evolutionary processes might be rather predictable’’ (Morris 1998; see also Stern et al. 2009). However, is this true on a finer scale? Are certain aspects of the polyploidy process actually ‘‘hard-wired’’? Preservation of duplicated gene copies following genome duplication appears far from random, with specific functional categories preferentially retained (Blanc and UN Editor Proof 276 Layout: T1 Standard SC Chapter No.: 14 Book ISBN: 978-3-642-31441-4 Page: 277/291 The Early Stages of Polyploidy 277 TE D PR OO F Editor Proof 14 Book ID: 272454_1_En Date: 16-8-2012 190 191 192 193 194 195 196 197 198 199 200 201 202 Wolfe 2004; Freeling 2009) and reduplicated in subsequent polyploidizations (Paterson et al. 2006). Independent WGDs in the ancestors of Arabidopsis, Oryza (rice), Saccharomyces (yeast), and Tetraodon (pufferfish) have been followed by convergent fates of many gene families (Paterson et al. 2006). Collectively, these observations indicate that at deep timescales there may exist certain ‘‘principles’’ that govern the fates of gene and genome duplications. But is this true in the early stages following polyploid formation? In the sections that follow we review data from different sets of molecular markers as well as chromosomal data across polyploid populations of separate origin and ask if aspects of polyploid evolution are indeed hard-wired or if stochastic processes prevail. CO RR 189 UN 188 EC Fig. 14.2 Maps showing the location of Tragopogon allotetraploid populations that are found within the Palouse region of the Pacific Northwest of the U.S.A. Populations of T. miscellus (squares) and T. mirus (circles) located in towns are indicated in the relevant counties of Washington and Idaho. A contiguous urbanized area containing multiple T. miscellus populations is indicated (shaded) 14.2.1 rDNA Loci/Concerted and Repeated Evolution Concerted evolution, which results in the homogenization of gene sequences to one type, is a common feature of ribosomal RNA genes (e.g., Zimmer et al. 1980). In both T. mirus and T. miscellus, concerted evolution is ongoing, but incomplete Layout: T1 Standard SC Chapter No.: 14 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 278/291 D. E. Soltis et al. 225 14.2.2 Homeolog Loss and Gene Silencing 226 14.2.2.1 One Gene at a Time 211 212 213 214 215 216 217 218 219 220 221 222 223 227 228 229 230 231 232 233 234 235 236 237 238 239 240 241 242 PR OO 209 210 D 207 208 TE 206 EC 205 Although Tragopogon affords unique opportunities for evolutionary study, it is not a genetic model organism; hence, until recently, genetic resources have not been available (but see Sect. 14.2.2.2), which has slowed research progress. As a result, genetic and genomic changes in the newly formed tetraploids were initially examined one gene at a time for upto 29 loci (Tate et al. 2006, 2009b; Buggs et al. 2009, 2010b; Koh et al. 2010). Initially, we used AFLP-cDNA display to screen plants of T. miscellus, T. mirus, and parental diploids to look for promising candidate genes—that is, fragments that did not show additivity in the allopolyploids as would be expected (Tate et al. 2006; Koh et al. 2010). Additional genes were surveyed because they were orthologous to genes that were singletons in other Asteraceae species (Buggs et al. 2009; Koh et al. 2010); the fate of such genes seemed of particular interest in new polyploids. The results of these one-gene-at-a-time surveys are presented in detail elsewhere (Tate et al. 2006, 2009b; Buggs et al. 2009, 2010b; Koh et al. 2010). Significantly, most of the changes observed in populations of both young polyploids are homeolog loss events, which far outnumber gene-silencing events in all CO RR 204 F 224 (Kovarik et al. 2005). In contrast to 80-year-old natural polyploids, F1 hybrids have equal contributions of the diploid parents, as do raw (S0) synthetic polyploids, as well as the earliest natural populations of T. mirus and T. miscellus (based on DNA from herbarium specimens). But in all modern-day natural populations except one, each representing a distinct origin, the rDNA type of T. dubius is consistently in very low abundance, with either the T. pratensis rDNA type (in T. miscellus) or T. porrifolius rDNA type (in T. mirus) in much greater abundance. In only one population of T. mirus are the parental contributions balanced (Malinska et al. 2011). Thus, concerted evolution has consistently occurred in these new polyploid lines of separate origin, and it has repeatedly operated ‘‘against’’ T. dubius, homogenizing those copies in the direction of the other parent. Surprisingly, despite being the least abundant in terms of rDNA gene copy number, T. dubius is by far the most abundant transcript in natural polyploid populations (Matyasek et al. 2007). This bias against the rDNA cistron in natural polyploid populations is already apparent in the S1 generation of synthetic polyploids (Malinska et al. 2010, 2011). For example, in four lines of synthetic T. miscellus, only three individuals (4 %) had balanced parental gene ratios while 65 individuals (92 %) inherited more T. pratensisorigin units than would be expected under additivity. In seven lines of synthetic T. mirus, 32 individuals (29 %) exhibited balanced rDNA genotypes, 69 individuals (63 %) showed more 35S rDNA of T. porrifolius origin than expected, and only 9 plants (8 %) had more T. dubius-origin rDNA (Malinska et al. 2010, 2011). 203 UN Editor Proof 278 Layout: T1 Standard SC Chapter No.: 14 Book ISBN: 978-3-642-31441-4 Page: 279/291 The Early Stages of Polyploidy 279 D PR OO F Editor Proof 14 Book ID: 272454_1_En Date: 16-8-2012 TE Fig. 14.3 Considerable variation from plant to plant within and among populations in the amount of homeolog loss detected in natural populations of T. miscellus. Population and plant designations are given on the x-axis; proportion of loci on the y-axis. Colors designate genotype as indicated in the figure. D = T. dubius allele; P = T. pratensis allele 247 248 249 250 251 252 253 254 255 256 257 258 259 260 261 262 EC 245 246 populations examined of both polyploids. Furthermore, most of the homeolog losses in both polyploids were from T. dubius, the diploid parent that is shared by both T. mirus and T. miscellus. It is also noteworthy that the same suite of genes consistently shows additivity of the parental gene copies (no loss or silencing) in polyploid populations of separate origin, whereas some of the genes analyzed consistently show some evidence of loss across at least some of the populations surveyed. Thus, these early analyses of Tragopogon polyploids are also in agreement with the hypothesis that there may be some underlying ‘‘principles’’ to polyploidization at the genetic or biochemical level. But stochasticity is operating as well in these young polyploids. Although homeolog loss is present in the polyploid populations, the process is ongoing, and appears to be stochastic within individual populations. In no population examined has silencing or loss been complete (i.e., observed in all individuals of a population). The amount of gene loss within populations varies (Fig. 14.3). Furthermore, these losses and gene silencing events were not detected in F1 hybrids or early-generation synthetic lines (S1). Hence, loss of homeologs and gene silencing are not immediate consequences of hybridization or polyploidization in Tragopogon, but appear to occur several generations after polyploid formation and for certain genes. CO RR 244 UN 243 Layout: T1 Standard SC Chapter No.: 14 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 280/291 280 D. E. Soltis et al. 263 F PR OO Editor Proof Fig. 14.4 Chromosome diagram for the alloetraploid T. miscellus (representative of other allotetraploids) illustrating terminology used throughout this chapter— chromosome, homeolog, and allele 14.2.2.2 Tragopogon Goes Genomic 285 14.2.2.3 Genomic Insights 269 270 271 272 273 274 275 276 277 278 279 280 281 282 283 286 287 288 289 290 291 TE 268 EC 267 CO RR 266 Using Sequenom genotyping, we examined patterns of homeolog presence/ absence in 70 sets of homeologs in 59 plants from five independently formed populations of allotetraploid T. miscellus (Buggs et al. 2012). Extensive gene loss occurred in the B40 generations since polyploidization in T. miscellus. An average of *20 % of the 70 loci investigated in each plant of T. miscellus was missing one or both alleles of one homeolog (excluding assays where neither homeolog was UN 265 D 284 Although important insights were obtained in Tragopogon, as well as other nonmodel systems, using the methods described above, these approaches have shortcomings. These analyses were based on examination of specific homeolog pairs using cleaved amplified polymorphic sequence analysis (CAPS). This approach uses restriction enzymes and hence has limitations, as discussed by Gaeta and Pires (2010). In addition, such surveys are slow and labor-intensive (reviewed in Soltis et al. 2009b; Buggs et al. 2009). Because of these limitations, we have sought ways to quickly and inexpensively build a framework for addressing genome-scale questions in Tragopogon. Newly developed genomic resources quickly facilitated the use of Tragopogon as a model for the study of recent and repeated polyploidization. Using a combination of 454 and Illumina sequencing of genomic and cDNA, we identified SNPs (singlenucleotide polymorphisms) between the parental species T. dubius and T. pratensis for analysis of the allotetraploids (Buggs et al. 2010a, b). Thousands of SNPs distinguish the three parental diploids. Following SNP discovery, we designed primers for Sequenom analysis (see Buggs et al. 2010a). Briefly, Sequenom iPLEx genotyping uses mass spectrometry to carry out high-throughput and highly accurate genotyping with multiple SNPs multiplexed in one reaction (Jurinke et al. 2005). The assays permitted detection of gene loss and changes in allele number (‘allele’ is used here to designate alleles at both homologous and homeologous loci; Fig. 14.4) within polyploids versus changes in gene expression (see Sect. 14.2.3 below). 264 Layout: T1 Standard SC Chapter No.: 14 298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 F 297 PR OO 296 D 295 TE 294 detected). Approximately one-third of these cases were single-allele absences. There was an overall bias toward loss of paternally derived homeologs. Assuming that immediately after polyploidization each individual had two alleles from T. dubius and two from T. pratensis at each pair of homeologous loci, on average at least 7.7 % of the original 280 allele copies have been lost in an individual plant’s lineage since polyploidization occurred. In only one case did we find that absence of a homeolog was fixed in a population. The fact that few gene losses were fixed in populations, and many individuals were missing single alleles, indicates that allele loss is ongoing in these populations (Buggs et al. 2012). These genomic data confirmed and extended our previous results and provided more insight into the processes involved. Certain loci were repeatedly missing in independently formed populations; the loci studied could be grouped into 12 clusters that followed recurrent patterns of presence/absence in populations with unique origins. Therefore, evolution is repeated in separate lineages of independent origin. Homeolog loss is also repeated at deeper phylogenetic scales. We compared patterns of gene loss found in T. miscellus with patterns found in 12 other species of Asteraceae that are considered ancient polyploids (Barker et al. 2008). Eighteen genes in our study had GO categories that tended to be lost in Asteraceae (Barker et al. 2008), and these had median homeolog absence of 7.0 % in T. miscellus. A comparison of these 18 genes versus all other genes (i.e., those that tend to be retained in Asteraceae, plus those with no preferential loss or retention: 45 genes, with a median absence of 4.0 %) showed significantly higher homeolog absence in the 18 genes. Thus, gene loss in T. miscellus, a very young Asteraceae polyploid, appears to repeat patterns occurring in older Asteraceae polyploids. One hypothesis for the repeated patterns of duplicate gene retention is the gene balance hypothesis (Freeling 2009; Freeling and Thomas 2006; Birchler et al. 2005; Papp et al. 2003). This hypothesis maintains that genes coding for products that are highly connected (i.e., within protein complexes or biochemical pathways) are dosage-sensitive in that they must be present in the nucleus in the same number of copies as the genes for products with which they interact. Thus, ‘‘connected’’ genes are hypothesized to be retained together as duplicate copies so as to preserve stoichiometry rather than reverting to singleton status one by one over time. In contrast, genes whose products are less connected are dosageinsensitive and are expected to revert gradually to singleton status. Our analyses so far suggest that dosage sensitivity may in fact be playing a role in Tragopogon (Buggs et al. 2012). EC 293 281 CO RR 292 Book ISBN: 978-3-642-31441-4 Page: 281/291 The Early Stages of Polyploidy UN Editor Proof 14 Book ID: 272454_1_En Date: 16-8-2012 14.2.3 Tissue-Specific Silencing Divergence of duplicate gene expression patterns among tissues has been suggested as a precursor of future evolution (Ohno 1970). Expression of a gene duplicate in a tissue where the progenitor copy was not expressed may indicate Layout: T1 Standard SC Chapter No.: 14 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 282/291 340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 F 339 PR OO 338 D 336 337 TE 335 neofunctionalization (Ohno 1970; Duarte et al. 2006; see Chap. 1 of this volume), while division of ancestral patterns of tissue-specific expression among duplicates suggests subfunctionalization (Lynch and Conery 2000; Rodin and Riggs 2003; Duarte et al. 2006) and silencing of a gene duplicate in all tissues points to nonfunctionalization (Duarte et al. 2006). Both neofunctionalization and subfunctionalization will lead to long-term retention of duplicated genes, whereas silencing/nonfunctionalization will generally lead to loss of a duplicate. Tissuespecific expression of duplicated genes has been studied in older gene duplicates in model organisms and crops (Adams et al. 2003; Duarte et al. 2006; Ganko et al. 2007; Semon and Wolfe 2008; Chaudhary et al. 2009), but in these species the ancestral patterns of gene expression and the ages of duplicates are not known precisely. Cases of tissue-specific expression patterns of very young gene duplicates are restricted to a few synthetic polyploids (Adams et al. 2003, 2004; Wang et al. 2004; Chaudhary et al. 2009). We examined the expression of 13 homeolog pairs in seven tissues of 10 plants of T. mirus from two natural populations of independent origin (Buggs et al. 2010b). Of the 910 assays in T. mirus, 63 % showed expression of both homeologs, 7 % showed no expression of either homeolog, 20 % showed non-expression of one homeolog across all tissues of a plant, and 8 % showed non-expression of a homeolog in a particular tissue within a plant. We found two cases of reciprocal tissue-specific expression between homeologs, potentially indicative of subfunctionalization. This study therefore showed that tissue-specific silencing, and even apparent subfunctionalization, can arise rapidly in the early generations of natural allopolyploidy. Similar results were found for 18 homeolog pairs using the same approaches in T. miscellus populations (Buggs et al. 2011b). In T. miscellus we also examined tissue-specific gene expression using the previously described Sequenom assays (Buggs et al. 2011b). Tissue-specific expression of 144 homeolog pairs in two natural populations was compared with patterns of allelic expression in both in vitro ‘‘hybrids’’ and hand-crossed F1 hybrids between the parental diploids T. dubius and T. pratensis, and with patterns of homeolog expression in synthetic (S1) allotetraploids. Tissue-specific homeolog expression was frequent in natural allopolyploids, but F1 hybrids and S1 allopolyploids showed less tissue-specific homeolog expression than the natural allopolyploids and the in vitro ‘‘hybrids’’ of diploid parents. These results suggest that ‘‘transcriptomic shock’’ upon hybridization (McClintock 1984) includes the activation of allele/homeolog expression in all tissues, causing a loss of tissue-specific expression patterns seen in the diploid parents. Such activation has seldom been considered in terms of the tissue-specific activation of protein-coding genes. Activation of homeologs has also been found in cotton F1 hybrids and allopolyploids (Chaudhary et al. 2009), who termed it ‘‘transcriptional neofunctionalization’’. We showed this to be widespread in Tragopogon. This may fit a scenario in which activity of small interfering RNA molecules, which influence gene expression, is temporarily lost in F1 hybrids and early allopolyploids, but restored subsequently. EC 334 CO RR 333 D. E. Soltis et al. UN Editor Proof 282 Layout: T1 Standard SC Chapter No.: 14 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 F 382 PR OO 381 D 380 Despite enormous progress in our understanding of many aspects of polyploidy, little attention has been paid to chromosomal constitution, structure, and organization. The New World Tragopogon allotetraploids illustrate the valuable role of cytology in examining hybridization and polyploidy in plants. Our recent studies have revealed high levels of chromosomal variation in *80-year-old natural populations, as well as the newly synthesized allotetraploids T. mirus and T. miscellus. The earliest studies of T. mirus and T. miscellus identified mitotic complements of 2n = 24 and typically 12 bivalents at meiosis (Ownbey 1950; Ownbey and McCollum 1954). This supported the expectation that allotetraploids were chromosomally additive of the diploid progenitors, which are both 2n = 12. Lim et al. (2008) used genomic and fluorescence in situ hybridization (GISH/ FISH) in a preliminary survey of several T. mirus and T. miscellus plants and found a few plants of each species that were not chromosomally additive of the diploid parents. Several natural T. mirus and T. miscellus plants were found to be aneuploid, with intergenomic translocations. Although most of the aneuploid individuals examined were 2n = 24, not all chromosomes were present in two copies as expected. GISH conducted on synthetic T. mirus pollen mother cells at the diplotene stage of meiosis showed allosyndetic pairing within multivalents. Meiotic instability, in the form of anaphase bridges and lagging chromosomes, was also observed by Tate et al. (2009a) in the first synthetic polyploid generation (S1) in T. mirus and T. miscellus lines. Thus, gametes with aneuploid and/or rearranged chromosome complements can be potentially generated as early as the first meiosis following genome doubling, which has been shown to be the case for synthetic allotetraploid B. napus (Szadkowski et al. 2010). The preliminary work of Lim et al. (2008) prompted a detailed examination using GISH and FISH of the chromosomal variation generated in six T. miscellus populations of independent origin (Chester et al. 2012). In all six populations, both aneuploidy and translocations were common (Fig. 14.5). Only 3 of the 58 plants exhibited the expected additivity of the diploid parental karyotypes (with neither aneuploidy nor translocations). Although approximately 70 % of polyploid plants were aneuploid for one or more chromosomes, variation in copy number appears to be constrained. Most plants were 2n = 24, and the total copy number for each homeologous group of chromosomes was typically four as a result of aneuploidy being reciprocal between homeologous chromosomes (Fig. 14.5). Thus, most deviations from disomy were in the form of monosomy-trisomy or nullisomytetrasomy, between homeologous chromosomes. This pattern of extensive aneuploidy while maintaining the overall copy number closely resembles cytological changes in synthetic neoallotetraploid Brassica napus (Xiong et al. 2011). Gene dosage has been implicated as a major factor constraining chromosomal changes such that imbalances, which arise, require compensation by chromosomes (compensatory aneuploidy) or homeologous segments (compensatory translocations). TE 379 EC 378 283 14.2.4 Cytogenetic Insights CO RR 377 Book ISBN: 978-3-642-31441-4 Page: 283/291 The Early Stages of Polyploidy UN Editor Proof 14 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 14 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 284/291 D. E. Soltis et al. PR OO F Editor Proof 284 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 TE 422 423 EC 421 In T. miscellus, genome rearrangements via intergenomic translocations were common; 76 % of the plants showed evidence of at least one translocation. The two largest chromosomes (groups A and B) showed the highest incidence of translocations. Across all T. miscellus plants at least six different translocation breakpoint positions were observed along the A group chromosomes. In one T. miscellus population in Spokane, WA, an A-chromosome intergenomic translocation appeared close to fixation, being homozygous in 8 of the 10 individuals analyzed. Individuals from six populations also appear to have undergone reciprocal translocations on the B-group chromosomes, but they mostly occur at a similar position, near the end of the long arm. The observed chromosomal changes in the recently formed Tragopogon allotetraploids provide two possible explanations for the loss of DNA that has been observed using molecular SNP-based assays. (1) If a translocation was in a nonreciprocal state and homozygous, this could lead to the loss of DNA from one of the parental diploids in the translocated region. (2) Nullisomy would lead to the complete loss of DNA from one of the parental chromosomes. This may have contributed to the clustering of patterns of gene loss and retention found using Sequenom assays (see above). However, the diverse patterns of homeolog losses that have been detected are not readily explainable by only these large-scale chromosomal changes detectable with GISH. CO RR 420 UN 419 D Fig. 14.5 Mitotic GISH karyotypes of three aneuploid T. miscellus individuals. Chromosomes derived from each parental genome are shown, i.e., the T. dubius genome (D-subgenome) and the T. pratensis genome (P-subgenome). Chromosomes deviating from disomy are as follows: top, monosomy: trisomy for E chromosomes; middle, nullisomy: tetrasomy for E chromosomes; bottom, four cases of monosomy: trisomy for chromosomes B, C, D, and E. All three individuals are 2n = 24 and have four copies of each homeologous group (A–F). Scale bar: 5 lm 14.3 Comparing Tragopogon to Other Well-Studied Systems Genome evolution in other well-studied polyploid systems (e.g., Gossypium, Triticum, Nicotiana, Arabidopsis, Brassica, Senecio, Spartina) exhibits important similarities and differences. Tragopogon is noteworthy in that initial studies show Layout: T1 Standard SC Chapter No.: 14 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 F 448 PR OO 447 D 445 446 that while true expression changes play a major role, homeolog losses are also very prominent in these young polyploids (Tate et al. 2006, 2009b; Koh et al. 2010; Buggs et al. 2009, 2010a, b, 2011b, 2012). Of the initial 23 genes analyzed in T. miscellus, 15 showed homeolog loss in one or more plants from nature, and 8 showed true expression changes (results for T. mirus are comparable); these patterns were confirmed via genomic analyses of many more genes. In contrast, in synthetic wheat (Triticum) and synthetic Arabidopsis thaliana and A. suecica polyploids, as well as in tetraploid cotton (Gossypium), expression changes dominate (Adams et al. 2003, 2004; Hovav et al. 2008a, b; Udall et al. 2006; Flagel et al. 2008; Madlung et al. 2004; Kashkush et al. 2002; Wang et al. 2006). Tragopogon may be most similar to the allotetraploid B. napus, in which most of the apparent gene silencing events observed in later generations were due to gene losses, most likely resulting from genomic rearrangements (Song et al. 1995; Gaeta et al. 2007). Across 50 lines of B. napus, genetic changes are equally distributed between the parental diploid genomes (Gaeta et al. 2007). The system is dynamic— some lines become more ‘‘oleracea like’’ and others more ‘‘rapa like’’ in terms of losses and corresponding expression differences. In Arabidopsis suecica allopolyploids, silencing of homeologs from one parent (A. thaliana) was observed more frequently than the silencing of homeologs from the other parent, Arabidopsis arenosa (Wang et al. 2006). In synthetic polyploids in Triticale, the contribution of the rye genome (Secale cereale) is preferentially silenced (Ma et al. 2004; Ma et al. 2006). In cotton, tissue-specific subfunctionalization occurs for some loci (Adams et al. 2003, 2004; Adams and Wendel 2004), but overall gene expression is biased toward one parent (Udall et al. 2006; Flagel et al. 2008; Rapp et al. 2009; Flagel and Wendel 2010). In both T. miscellus and T. mirus, homeologs of one diploid genome (T. dubius) are more often lost or not expressed; that is, T. dubius is often the ‘‘loser genome’’ based on the set of genes surveyed to date. Tragopogon also exhibits cytogenetic similarities to Brassica. Both T. mirus and T. miscellus exhibit numerous translocations as well as extensive aneuploidy while maintaining the overall copy number expected in an allotetraploid (reciprocal monosomy: trisomy and nullisomy: tetrasomy); this closely resembles cytological changes in synthetic neoallotetraploid B. napus (Xiong et al. 2011). It is also now possible to compare results for Tragopogon with several of the other recently formed natural polyploids. Great research progress has now been made on four of the five polyploids known to have formed in the past two centuries. In addition to T. mirus and T. miscellus, numerous insights into recent polyploidy have been obtained for Senecio cambrensis (also in the Asteraceae) and Spartina anglica (Poaceae). Both these model systems are reviewed elsewhere in this volume (see Chaps. 13 and 12, respectively). Interestingly, all of these recent polyploids have formed following introductions of one or both progenitors into a completely new geographic area. Some of these new polyploids have been successful in nature, particularly Spartina anglica, which is now distributed worldwide with major ecological impact (Ainouche et al. 2004, 2009; see Chap. 12 of this volume). The Tragopogon polyploids are now major weeds in the Palouse region of northwestern USA. In contrast, following initial range expansion, TE 444 285 EC 443 CO RR 442 Book ISBN: 978-3-642-31441-4 Page: 285/291 The Early Stages of Polyploidy UN Editor Proof 14 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 14 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 286/291 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 F 493 PR OO 492 D 491 TE 489 490 Senecio cambrensis now seems to be disappearing from parts of its original range (Abbott et al. 2007; see Chap. 13 of this volume). Senecio cambrensis formed from parents of different ploidal levels (one parent is itself a polyploid), whereas in Spartina anglica both parents are polyploids but of the same ploidal level. In Senecio cambrensis there is an additional level of complexity in that the diploid parent is itself of homoploid hybrid origin. In contrast, both parents of the recent Tragopogon polyploids are diploid. The multiple layers of recent hybridization and polyploidization in Senecio and Spartina could influence the genetic and expression changes detected in the recent polyploids in those genera. Molecular studies have revealed striking genetic and genomic changes in all of these recently formed polyploids. Transcriptomic shock has now been shown following hybridization in both Senecio (Hegarty et al. 2005, 2006) and Tragopogon (Buggs et al. 2011b), and substantial expression changes have also been reported in F1 hybrids and polyploids in Spartina (Cheilafa et al. 2010a, b) Thus, in recent polyploids in all three genera, hybridization seems to result in a major burst of altered gene expression. Changes in gene expression have been shown to be important in all of the systems, but have been best characterized in Spartina (Ainouche et al. 2009, Chap. 12 of this volume) and Senecio (Hegarty et al. 2008, Chap. 13 of this volume). Methylation alterations have been detected in both Spartina (Salmon et al. 2005) and Senecio (Hegarty et al. 2011), with epigenetic changes now considered to play a major role in these new polyploids (Parisod et al. 2009; Ainouche et al. 2009; Hegarty et al. 2011). However, epigenetic changes have not yet been analyzed in Tragopogon. Dramatic changes in gene expression have been documented in Spartina anglica, Senecio cambrensis, and the recently formed Tragopogon polyploids. However, homeolog loss seems to have played a more prominent role in young Tragopogon polyploids than in these other recent polyploids. In both the recently formed Tragopogon polyploids and in S. cambrensis, genomic studies reveal that one parental genome predominates over the other in terms of global patterns of expression and gene retention. In the allohexaploid S. cambrensis, gene expression is more similar to that of the tetraploid parent (S. vulgaris) than to the diploid parent, S. squalidus (Hegarty et al. 2006). In the young Tragopogon polyploids, homeologs of the diploid parent T. dubius are more often lost or not expressed. That is, following polyploidy there are clear winner and loser parental genomes, and these patterns are established quickly and repeatedly. Chromosomal changes have been investigated in detail in both Tragopogon and Spartina using FISH/GISH; recent polyploids in both genera exhibit substantial change, but different types of variation are evident. In Spartina, nonaploid plants (2n = ca. 90) were detected (rather than the expected 2n = 120–124), most likely resulting from backcrosses between S. anglica and its maternal parent S. alterniflora (Renny-Byfield et al. 2010). In Tragopogon, frequent translocations were detected as well as frequent reciprocal monosomy: trisomy and nullisomy: tetrasomy—these have not been previously reported from nature (Chester et al. 2012). Chromosomal pairing abnormalities (multivalents) have been detected in both Tragopogon EC 488 CO RR 487 D. E. Soltis et al. UN Editor Proof 286 Layout: T1 Standard SC Chapter No.: 14 Book ISBN: 978-3-642-31441-4 Page: 287/291 The Early Stages of Polyploidy 287 539 14.4 Conclusions 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 PR OO 540 A diverse array of experimental approaches has helped us to understand the evolution of young, natural Tragopogon polyploids at the genomic and transcriptomic levels. These have demonstrated: (1) the allopolyploids have multiple origins; (2) substantial gene loss occurs within the first 40 generations in nature, and is still ongoing; (3) patterns of gene loss are repeated among independent origins of the allopolyploids; (4) transcriptomic shock occurs upon allopolyploidization, involving a reduction in tissue-specific expression; (5) major chromosomal changes have occurred and are ongoing in the natural allopolyploids. The dynamic evolutionary processes that appear to be underway in Tragopogon polyploids may be representative of those that have occurred in other groups, and after past polyploidization events. Care is needed when extrapolating from a single system, because patterns of change may be influenced by the genetic background of the polyploid, such as the level of genetic differentiation between the two parental species of an allopolyploid (reviewed in Buggs et al. 2011a). Studies of polyploids in many plant families, of different ages and of different parental combinations, are needed to provide a comprehensive understanding of the evolution of allopolyploids. Tragopogon has great potential for further development as a model system for polyploidy. Older Eurasian polyploids will allow placement of another time point in the evolutionary series. Research is also underway on natural hybrids between T. pratensis and T. porrifolius (T. 9 mirabilis) that occur in the United Kingdom. Further development of genetic resources—sequencing of the T. dubius genome and production of a genetic map—will allow new questions to be addressed. In recent years, progress in understanding the genetic consequences of polyploidy has outstripped progress in understanding the ecological background in which evolution occurs (but see Ramsey 2011). As a genus that occurs almost exclusively in the wild, Tragopogon affords a system in which relevant ecological studies can be carried out, comparing the fitness of polyploids and diploids, and the different chromosomal variants of the polyploids. Such studies will contribute to a comprehensive view of polyploid evolution. D 537 TE 535 536 EC 534 CO RR 533 F 538 allopolyploids as well as in synthetic polyploids (Ownbey and McCollum 1953; Tate et al. 2009a). FISH/GISH has been problematic in Senecio, but traditional cytogenetic analysis suggests possible structural chromosomal changes, as well as some meiotic pairing irregularities (Crisp 1972). In addition to generating genetic diversity, intergenomic recombination resulting in chromosome rearrangements could lead to the formation of reproductive barriers between lineages of separate or independent origin, which is under investigation in Tragopogon. 532 UN Editor Proof 14 Book ID: 272454_1_En Date: 16-8-2012 Acknowledgments Funding for this research was provided by the University of Florida and NSF grants MCB-0346437, DEB-0614421, DEB-0919254, DEB-0922003, and DEB-0919348. R.J.A.B. has been supported since March 2010 by NERC Fellowship NE/G01504X/1. Layout: T1 Standard SC Chapter No.: 14 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 288/291 D. E. Soltis et al. References 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 Abbott RJ, Ireland HE, Rogers HJ (2007) Population decline despite high genetic diversity in the new allopolyploid species Senecio cambrensis (Asteraceae). Mol Ecol 16:1023–1033 Abbott RJ, Lowe AJ (2004) Origins, establishment, and evolution of new polyploids species: Senecio cambrensis and S. eboracensis in the British Isles. Biol J Linn Soc 82:467–474 Adams KL, Cronn R, Percifield R, Wendel JF (2003) Genes duplicated by polyploidy show unequal contributions to the transcriptome and organ-specific reciprocal silencing. Proc Nat Acad Sci USA 100:4649–4654 Adams KL, Percifield R, Wendel JF (2004) Organ-specific silencing of duplicated genes in a newly synthesized cotton allotetraploid. Genetics 168:2217–2226 Adams KL, Wendel JF (2004) Exploring the genomic mysteries of polyploidy in cotton. Biol J Linn Soc 82:573–581 Ainouche ML, Baumel A, Salmon A (2004) Spartina anglia C. E. Hubbard: a natural model system for analyzing early evolutionary changes that affect allopolyploid genomes. Biol J Linn Soc 82:475–484 Ainouche ML, Fortune PM, Salmon A, Parisod C, Grandbastien M-A, Ricou K, Fukunaga M, Misset M-T (2009) Hybridization, polyploidy and invasion: Lessons from Spartina (Poaceae). Biol Invasion. doi:10.1007s10530-0089383-2 Ashton PA, Abbott RJ (1992) Multiple origins and genetic diversity in the newly arisen allopolyploid species Senecio cambrensis Rosser (Compositae). Heredity 68:25–32 Barker MS, Kane NC, Matvienko M, Kozik A, Michelmore RW, Knapp SJ, Rieseberg LH (2008) Multiple Paleopolyploidizations during the evolution of the compositae reveal parallel patterns of duplicate gene retention after millions of years. Mol Biol Evol 25:2445–2455 Birchler JA, Riddle NC, Auger DL, Veitia R (2005) Dosage balance in gene regulation: biological implications. Trends Genet 21:219–226 Blanc G, Wolfe KH (2004) Functional divergence of duplicated genes formed by polyploidy during Arabidopsis divergence. Plant Cell 16:1679–1691 Blanca G, Díaz de la Guardia C (1996) Sinopsis del género Tragopogon L. (Asteraceae) en la Peninsula Ibérica. Anales del Jardín Botánico de Madrid 54:358–363 Buggs RJA, Doust AN, Tate JA, Koh J, Soltis K, Feltus FA, Paterson AH, Soltis PS, Soltis DE (2009) Gene loss and silencing in Tragopogon miscellus (Asteraceae): comparison of natural and synthetic allotetraploids. Heredity 103:73–81 Buggs RJA, Chamala S, Wu W, Gao L, May GD, Schnable PS, Soltis DE, Soltis PS, Barbazuk WB (2010a) Characterization of duplicate gene evolution in the recent natural allopolyploid Tragopogon miscellus by next-generation sequencing and Sequenom iPLEX genotyping. Mol Ecol 19(1):1–15 Buggs RJA, Elliott NM, Zhang L, Koh J, Viccini LF, Soltis DE, Soltis PS (2010b) Tissue-specific silencing of homoeologs in natural populations of the recent allopolyploid Tragopogon mirus. New Phytol 186:175–183 Buggs RJA, Soltis PS, Soltis DE (2011a) Biosystematic relationships and the formation of polyploids. Taxon 60:324–332 Buggs RJA, Zhang L, Miles N, Tate JA, Gao L, Schnable PS, Barbazuk WB, Soltis PS, Soltis DE (2011b) Genomic and transcriptomic shock generate evolutionary novelty in a newly formed, natural allopolyploid plant. Curr Biol 21:1–6 Buggs RJA, Gao L, Wu W, Chamala S, Tate JA, Schnable PS, Soltis DE, Soltis PS, Barbazuk WB (2012) Rapid and repeated gene loss in a young polyploidy species. Curr Biol 22:248–252 Chaudhary B, Flagel L, Stupar RM, Udall JA, Verma N, Springer NM, Wendel JF (2009) Reciprocal silencing, transcriptional bias and functional divergence of homoeologs in polyploid cotton (Gossypium). Genetics 182:503–517 Chelaifa H, Mahe F, Ainouche M (2010a) Transcriptome divergence between the hexaploid saltmarsh sister species Spartina maritima and Spartina alterniflora (Poaceae). Mol Ecol 19:2050–2063 CO RR EC TE D PR OO F 573 UN Editor Proof 288 Layout: T1 Standard SC Chapter No.: 14 289 EC TE D PR OO F Chelaifa H, Monnier A, Ainouche M (2010b) Transcriptomic changes following recent natural hybridization and allopolyploidy in the salt marsh species Spartina x townsendii and Spartina anglica (Poaceae). New Phytol 186:161–174 Chen ZJ, Wang J, Tian L, Lee HS, Wang JJ, Chen M, Lee JJ, Josefsson C, Madlung A, Watson B, Pires JC, Lippman Z, Vaughn M, Colot V, Birchler JA, Doerge RW, Martienssen RA, Comai L, Osborn TC (2004) The development of an Arabidopsis model system for genome-wide analysis of polyploidy effects. Biol J Linn Soc 82:689–700 Chester M, Gallagher JP, Symonds VV, da Veruska Cruz Silva A, Mavrodiev EV, Leitch AR, Soltis PS, Soltis DE (2012) Extensive and repeated patterns of chromosomal variation in natural populations of a recently formed polyploid plant species. Proc Nat Acad Sci USA 109:1176–1181 Clausen J, Keck DD, Hiesey WM. 1945. Experimental studies on the nature of species II. Plant evolution through amphiploidy and autopolyploidy, with examples from the Madiinae. Publication 564, Carnegie Institute of Washington, Washington, DC Crisp PC (1972) Cytotaxonomic studies in the section Annui of Senecio. Ph. D Thesis, University of London Darlington CD (1937) Recent advances in cytology, 2nd edn. The Blakiston Company, Philadelphia Duarte JM, Cui L, Wall PK, Zhang Q, Zhang X, Leebens-Mack J, Ma H, Altman N, dePamphilis CW (2006) Expression pattern shifts following duplication indicative of subfunctionalization and neofunctionalization in regulatory genes of Arabidopsis. Mol Biol Evol 23:469–478 Dujon B, Sherman D, Fischer G, Durrens P, Casaregola S, Lafontaine I, De Montigny J, Marck C, Neuvéglise C, Talla E et al (2004) Genome evolution in yeasts. Nature 430:35–44 Flagel L, Udall J, Nettleton D, Wendel J (2008) Duplicate gene expression in allopolyploid Gossypium reveals two temporally distinct phases of expression evolution. BMC Biol 6:11 Flagel LE, Wendel JF (2010) Evolutionary rate variation, genomic dominance and duplicate gene expression during allotetraploid cotton speciation. New Phytol 186:184–193 Freeling M (2009) Bias in plant gene content following different sorts of duplication: tandem, whole-genome, segmental, or by transposition. Annu Rev Plant Biol 60:433–453 Freeling M, Thomas BC (2006) Gene-balanced duplications, like tetraploidy, provide predictable drive to increase morphological complexity. Genome Res 16:805–814 Gaeta RT, Pires JC (2010) Homoeologous recombination in allopolyploids: the polyploid ratchet. New Phytol 186:18–28 Gaeta RT, Pires JC, Iniguez-Luy F, Leon E, Osborn TC (2007) Genomic changes in resynthesized Brassica napus and their effect on gene expression and phenotype. Plant Cell 19:3403–3417 Ganko EW, Meyers BC, Vision TJ (2007) Divergence in expression between duplicated genes in Arabidopsis. Mol Biol Evol 24:2298–2309 Gould SJ (1994) The evolution of life on Earth. Sci Am 271:85–86 Gregory TR, Mable BK (2005) Polyploidy in animals. In: Gregory TR (ed) The evolution of the Genome. Elsevier/Academic, San Diego, pp 428–501 Hegarty MJ, Jones JM, Wilson ID, Barker GL, Coghill JA, Sanchez-Baracaldo P, Liu G, Buggs RJA, Abbott RJ, Edwards KJ, Hiscock SJ (2005) Development of anonymous cDNA microarrays to study changes to the Senecio floral transcriptome during hybrid speciation. Mol Ecol 14:2493–2510 Hegarty MJ, Barker GL, Wilson ID, Abbott RJ, Edwards KJ, Hiscock SJ (2006) Transcriptome shock after interspecific hybridization in Senecio is ameliorated by genome duplication. Curr Biol 16:1652–1659 Hegarty MJ, Barker GL, Brennan AC, Edwards KJ, Abbott RJ, Hiscock SJ (2008) Changes to gene expression associated with hybrid speciation in plants: further insights from transcriptomic studies in Senecio. Philos Trans R Soc London B series 363:3055–3069 Hegarty MJ, Batstone T, Barker GL, Edwards KJ, Abbott RJ, Hiscock SJ (2011) Nonadditive changes to cytosine methylation as a consequence of hybridization and genome duplication in Senecio (Asteraceae). Mol Ecol 20:105–113 CO RR 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 Book ISBN: 978-3-642-31441-4 Page: 289/291 The Early Stages of Polyploidy UN Editor Proof 14 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 14 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 290/291 EC TE D PR OO F Hovav R, Udall J, Chaudhary B, Flagel L, Rapp R, Wendel J (2008a) Partitioned expression of duplicated genes during development and evolution of a single cell in a polyploid plant. Proc Nat Acad Sci USA 105:6191 Hovav R, Udall JA, Chaudhary B, Hovav E, Flagel L, Hu G, Wendel JF (2008b) The evolution of spinnable cotton fiber entailed prolonged development and a novel metabolism. PLoS Genet 4:e2 Jiao Y, Wickett N, Ayyampalayam S, Chanderbali A, Landherr L, Ralph PE, Soltis PS, Soltis DE, Clifton SE, Ma H, Leebens-Mack J, dePamphilis CW (2011) Phylogenomic analysis reveals ancient genome duplications in seed plant and angiosperm history. Nature 473:97–100 Jurinke C, Denissenko MF, Oeth P, Ehrich M, van den Boom D, Cantor CR (2005) A single nucleotide polymorphism based approach for the identification and characterization of gene expression modulation using MassARRAY. Mutat Res 573:83–95 Kashkush K, Feldman M, Levy AA (2002) Gene loss, silencing, and activation in a newly synthesized wheat allotetraploid. Genetics 160:1651–1659 Kellis M, Birren BW, Lander ES (2004) Proof and evolutionary analysis of ancient genome duplication in the yeast Saccharomyces cerevisiae. Nature 428:617–624 Koh J, Tate JA, Soltis PS, Soltis DE (2010) Genomic and expression differences in natural populations of the recently formed allotetraploid Tragopogon mirus (Asteraceae). BMC Genomics 11:97 Kovarik A, Pires JC, Leitch AR, Lim KY, Sherwood A, Matyasek R, Rocca J, Soltis DE, Soltis PS (2005) Rapid concerted evolution in two allopolyploids of recent and recurrent origin. Genetics 169:931–944 Levin DA (1983) Polyploidy and novelty in flowering plants. Am Nat 122:1–25 Levin DA (2002) The role of chromosomal change in plant evolution. Oxford University Press, New York Lim KY, Soltis DE, Soltis PS, Tate JA, Matyasek R, Srubarova H, Kovarik A, Pires JC, Xiong ZY, Leitch AR (2008) Rapid chromosome evolution in recently formed polyploids in Tragopogon (Asteraceae). PLoS One 3:e3353 Löve A, Löve D (1949) The geobotanical significance of polyploidy. I. Polyploidy and latitude. Portugaliae Acta Biologica Serie A, Suppl vol. pp 273–352 Lynch M, Connery JS (2000) The evolutionary fate and consequences of duplicate genes. Science 290:1151–1155 Ma XF, Fang P, Gustafson JP (2004) Polyploidization-induced genome variation in Triticale. Genome 47:839–848 Ma XF, Fang P, Gustafson JP (2006) Timing and rate of genome variation in Triticale following allopolyploidization. Genome 49:950–958 Mable B (2003) Breaking down taxonomic barriers in polyploidy research. Trends Plant Sci 8:582–590 Mable BK, Alexandrou MA, Taylor MI (2011) Genome duplications in amphibians and fish: an extended synthesis. J Zool 284:151–182 Madlung A, Comai L (2004) The effect of stress on genome regulation and structure. Ann Bot 94:481 495 Malinska H, Tate JA, Matyasek R, Leitch AR, Soltis DE, Soltis PS, Kovarik A (2010) Similar patterns of rDNA evolution in synthetic and recently formed natural populations of Tragopogon (Asteraceae) allotetraploids. BMC Evol Biol 10:291 Malinska H, Tate JA, Mavrodiev E, Matyasek R, Lim KY, Leitch AR, Soltis DE, Soltis PS, Kovarik A (2011) Ribosomal RNA genes evolution in Tragopogon: A story of New and Old World allotetraploids and synthetic lines. Taxon 60:348–354 Matyasek R, Tate J, Lim YK, Srubaraova H, Koh J, Leitch A, Soltis DE, Soltis PS (2007) Concerted evolution of rDNA in recently formed Tragopogon allotetraploids is typically associated with an inverse correlation between gene copy number and expression. Genetics 176:2509–2519 Mavrodiev E, Soltis PS, Soltis DE (2008a) Parentage of six Old World polyploids in Tragopogon L. (Asteraceae: Scorzonerinae) based on ITS. ETS and plastid sequence data. Taxon 57:1217–1232 CO RR 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 D. E. Soltis et al. UN Editor Proof 290 Layout: T1 Standard SC Chapter No.: 14 291 EC TE D PR OO F Mavrodiev E, Nawchoo I, Soltis DE, Soltis PS (2008b) Molecular data reveal that the tetraploid Tragopogon kashmirianus Singh, a narrow endemic of Kashmir, is distinct from the North American T. mirus M. Ownbey. Bot J Linn Soc 158:391–398 Mavrodiev EV, Albach DC, Speranza P (2008c) A new polyploid species of the genus Tragopogon (Asteraceae, Cichorieae) from Russia. Novon 18:229–232 McClintock B (1984) The significance of responses of the genome to challenge. Science 226:792–801 Morris SC (1998) The crucible of creation: the Burgess shale and the rise of animals. Oxford University Press, Oxford Müntzing A (1936) The evolutionary significance of autopolyploidy. Hereditas 21:263–378 Novak SJ, Soltis DE, Soltis PS (1991) Ownbey Tragopogons—40 Years later. Am J Bot 78:1586–1600 Ohno S (1970) Evolution by gene duplication. Springer, Berlin Ownbey M (1950) Natural hybridization and amphiploidy in the genus Tragopogon. Am J Bot 37:487–499 Ownbey M, McCollum G (1954) The chromosomes of Tragopogon. Rhodora 56:7–21 Ownbey M, McCollum GD (1953) Cytoplasmic inheritance and reciprocal amphiploidy in Tragopogon. Am J Bot 40:788–796 Panopoulou G, Poustka AJ (2005) Timing and mechanism of ancient vertebrate genome duplications—the adventure of a hypothesis. Trends Genet 10:559–567 Papp B, Pal C, Hurst LD (2003) Dosage sensitivity and the evolution of gene families in yeast. Nature 424:194–197 Parisod C, Salmon A, Zerjal T, Tenaillon M, Grandbastien M-A, Ainouche M (2009) Rapid structural and epigenetic reorganization near transposable elements in hybrid and allopolyploid genomes in Spartina. New Phytol 184:1003–1015 Paterson AH, Chapman BA, Kissinger J, Bowers JE, Feltus FA, Estill JC, Marler BS (2006) Many gene and domain families have convergent fates following independent whole-genome duplication events in Arabidopsis, Oryza, Saccharomyces and Tetraodon. Trends Genet 22:597–602 Petit RJ, Aguinagalde, JL de Beaulieu, C Bittkau, S Brewer, R Cheddadi, R Ennos, S Fineschi, D Grivet, M Lascoux, A Mohanty, G Müller-Starck, B Musch, A Palmé, S Rendell, GG. Vendramin (2003) Glacial refugia: hotspots but not melting pots of genetic diversity. Science 300:1563–1565 Ramsey J (2011) Polyploidy and ecological adaptation in wild yarrow. Proc Nat Acad Sci USA 108:6697–6698 Rapp RA, Udall JA, Wendel JF (2009) Genomic expression dominance in allopolyploids. BMC Biol. 7:18 Renny-Byfield S, Ainouche M, Leitch IJ, Lim KY, Le Comber SC, Leitch AR (2010) Flow cytometry and GISH reveal mixed ploidy populations and Spartina nonaploids with genomes of S. alterniflora and S. maritima origin. Ann Bot 105:527–533 Rodin SN, Riggs AD (2003) Epigenetic silencing may aid evolution by gene duplication. J Mol Evol 56:718–729 Salmon A, Ainouche ML, Wendel JF (2005) Genetic and epigenetic consequences of recent hybridization and polyploidy in Spartina (Poaceae). Mol Ecol 14:1163–1175 Semon M, Wolfe KH (2008) Preferential subfunctionalization of slow-evolving genes after allopolyploidization in Xenopus laevis. Proc Nat Acad Sci USA 105:8333–8338 Soltis DE, Soltis PS (1995) The dynamic nature of polyploid genomes. Proc Nat Acad Sci USA 92:8089–8091 Soltis DE, Soltis PS, Pires JC, Kovarik A, Tate JA, Mavrodiev E (2004) Recent and recurrent polyploidy in Tragopogon (Asteraceae): cytogenetic, genomic and genetic comparisons. Biol J Linn Soc 82:485–501 Soltis DE, Albert VA, Leebens-Mack J, Bell CD, Paterson A, Zheng C, Sankoff D, Wall PK, Soltis PS (2009a) Polyploidy and angiosperm diversification. Am J Bot 96:336–348 CO RR 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768 769 770 771 772 773 774 775 776 777 778 779 780 781 782 783 784 785 Book ISBN: 978-3-642-31441-4 Page: 291/291 The Early Stages of Polyploidy UN Editor Proof 14 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 14 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 292/291 EC TE D PR OO F Soltis DE, Buggs RJA, Barbazuk WB, Schnable PS, Soltis PS (2009b) On the origins of species: does evolution repeat itself in polyploid populations of independent origin? Cold spring harbor symposia on quantitative biology, Vol. LXXIV Soltis DE, Mavrodiev EV, Meyers, SC, Severns PM, Zhang L, Gitzendanner MA, Ayers T, Chester M, Soltis PS (2012) Additional origins of Ownbey’s Tragopogon mirus. Bot Linn Soc 169:297–311 Soltis DE, Soltis PS (1989) Allopolyploid speciation in Tragopogon: Insights from chloroplast DNA. Am J Bot. 76:1119–1124 Soltis DE, Soltis PS (1993) Molecular data and the dynamic nature of polyploidy. Crit Rev Plant Sci 12:243–273 Soltis DE, Soltis PS (1999) Polyploidy: recurrent formation and genome evolution. Trends Ecol Evol 14:348–352 Soltis PS, Soltis DE (2000) The role of genetic and genomic attributes in the success of polyploids. Proc Nat Acad Sci (USA) 97:7051–7057 Soltis PS, Soltis DE (2009) The role of hybridization in plant speciation. Annu Rev Plant Biol 60:561–588 Soltis PS, Plunkett GM, Novak SJ, Soltis DE (1995) Genetic variation in Tragopogon species: additional origins of the allotetraploids T. mirus and T. miscellus (Compositae). Am J Bot 82:1329–1341 Song KM, Lu P, Tang KL, Osborn TC (1995) Rapid genome change in synthetic polyploids of Brassica and its implications for polyploid evolution. Proc Nat Acad Sci USA 92:7719–7723 Stebbins GL (1950) Variation and evolution in plants. Columbia, New York Stebbins GL (1971) Chromosomal evolution in higher plants. Addison-Wesley, London Stern DL, Orgogozo V (2009) Is genetic evolution predictable? Science 323:746–751 Symonds VV, Soltis PS, Soltis DE (2010) Dynamics of polyploid formation in Tragopogon (Asteraceae): recurrent formation, gene flow, and population structure. Evolution 64:1984–2003 Szadkowski E, Eber F, Huteau V, Lod M, Huneau C, Belcram H, Coriton O, ManzanaresDauleux M, Delourme R, King G (2010) The first meiosis of resynthesized Brassica napus, a genome blender. New Phytol 186:102–112 Tate JA, Ni Z, Scheen AC, Koh J, Gilbert CA, Lefkowitz D, Chen ZJ, Soltis PS, Soltis DE (2006) Evolution and expression of homoeologous loci in Tragopogon miscellus (Asteraceae), a recent and reciprocally formed allopolyploid. Genetics 173:1599–1611 Tate JA, Symonds VV, Doust AN, Buggs RJA, Mavrodiev EV, Soltis PS, Soltis DE (2009a) Synthetic polyploids of Tragopogon miscellus and T. mirus (Asteraceae): 50 ? years after Ownbey’s discovery. Am J Bot 96:979–988 Tate JA, Joshi P, Soltis K, Soltis PS, Soltis DE (2009b) On the road to diploidization? Homoeolog loss in independently formed populations of the allopolyploid Tragopogon miscellus (Asteraceae). BMC Plant Biol 9:80 Udall JA, Swanson JM, Nettleton D, Percifield RJ, Wendel JF (2006) A novel approach for characterizing expression levels of genes duplicated by polyploidy. Genetics 173(3):1823–1827 Xiong Z, Gaeta RT, Pires JC (2011) Homoeologous shuffling and chromosome compensation maintain genome balance in resynthesized allopolyploid Brassica napus. Proc Nat Acad Sci USA 108:7908–7913 Wang JL, Tian L, Madlung A, Lee HS, Chen M, Lee JJ, Watson B, Kagochi T, Comai L, Chen ZJ (2004) Stochastic and epigenetic changes of gene expression in Arabidopsis polyploids. Genetics 167:1961–1973 Wang JL, Tian L, Lee HS, Wei NE, Jiang HM, Watson B, Madlung A, Osborn TC, Doerge RW, Comai L, Chen ZJ (2006) Genomewide nonadditive gene regulation in Arabidopsis allotetraploids. Genetics 172:507–517 Wolfe KH, Shields DC (1997) Molecular evidence for an ancient duplication of the entire yeast genome. Nature 387:708–713 Zimmer EA, Martin SL, Beverley SM, Kan YW, Wilson AC (1980) Rapid duplication and loss of genes coding for the q chains of hemoglobin. Proc Nat Acad Sci USA 77:2158–2162 CO RR 786 787 788 789 790 791 792 793 794 795 796 797 798 799 800 801 802 803 804 805 806 807 808 809 810 811 812 813 814 815 816 817 818 819 820 821 822 823 824 825 826 827 828 829 830 831 832 833 834 835 836 837 838 D. E. Soltis et al. UN Editor Proof 292 Metadata of the chapter that will be visualized in SpringerLink Book Title Polyploidy and Genome Evolution Series Title Chapter Title Yeast as a Window into Changes in Genome Complexity Due to Polyploidization Copyright Year 2012 Copyright HolderName Springer-Verlag Berlin Heidelberg Corresponding Author Family Name Conant Particle Given Name Gavin C. Suffix Author Division MU Informatics Institute Organization University of Missouri Address 65211, Columbia, MO, USA Division Division of Animal Sciences Organization University of Missouri Address Columbia, MO, USA Email conantg@missouri.edu Family Name Hudson Particle Given Name Corey M. Suffix Division MU Informatics Institute Organization University of Missouri Address 65211, Columbia, MO, USA Email Abstract Due to the long history of genetic analyses in yeasts and their experimental tractability, the yeast genome duplication provides important perspectives on the genome and population-level processes that follow wholegenome duplication (WGD). We discuss the history of the discovery of the Saccharomyces cerevisiae WGD, with special emphasis on the role of comparative genomics in its analysis. We then explore models of the WGD shaped population and species divergence, both at a gene level (e.g., Dobzhansky-Muller incompatibility) and from the perspective of recent work on secondary allopolyploidy in Saccharomyces pastorianus. Finally, we explore the selective forces that act on the WGD-produced paralogs and shape their patterns of loss and retention. In addition to discussing the dosage balance hypothesis as it applies to the yeast WGD, we explore the role of the WGD in shaping several complex metabolic and regulatory phenotypes. Book ISBN: 978-3-642-31441-4 Page: 293/307 Chapter 15 5 Corey M. Hudson and Gavin C. Conant 9 10 11 12 13 14 15 16 17 D 8 Abstract Due to the long history of genetic analyses in yeasts and their experimental tractability, the yeast genome duplication provides important perspectives on the genome and population-level processes that follow whole-genome duplication (WGD). We discuss the history of the discovery of the Saccharomyces cerevisiae WGD, with special emphasis on the role of comparative genomics in its analysis. We then explore models of population and species divergence, both at a gene level (e.g., Dobzhansky-Muller incompatibility) and from the perspective of recent work on secondary allopolyploidy in Saccharomyces pastorianus. Finally, we explore the selective forces that act on the WGD-produced paralogs and shape their patterns of loss and retention. In addition to discussing the dosage balance hypothesis as it applies to the yeast WGD, we explore the role of the WGD in shaping several complex metabolic and regulatory phenotypes. TE 6 7 EC 3 20 21 22 23 24 CO RR 18 19 PR OO 4 Yeast as a Window into Changes in Genome Complexity Due to Polyploidization 2 F 1 Book ID: 272454_1_En Date: 16-8-2012 15.1 Introduction Researchers have found remnants of ancient whole-genome duplications (WGDs) preserved in the genomes of many and diverse eukaryotes. This book, in fact, is a testament to that diversity, illustrating the sheer number of independent events in plants as well as the evolutionarily basal events in vertebrates and other, more recent WGDs in teleost fishes and frogs. Although we have still not fully validated C. M. Hudson  G. C. Conant MU Informatics Institute, University of Missouri, Columbia, MO 65211, USA UN Editor Proof Layout: T1 Standard SC Chapter No.: 15 G. C. Conant (&) Division of Animal Sciences, University of Missouri, Columbia, MO, USA e-mail: conantg@missouri.edu P. S. Soltis and D. E. Soltis (eds.), Polyploidy and Genome Evolution, DOI: 10.1007/978-3-642-31442-1_15,  Springer-Verlag Berlin Heidelberg 2012 293 Layout: T1 Standard SC Chapter No.: 15 Book ID: 272454_1_En Date: 16-8-2012 C. M. Hudson and G. C. Conant 48 15.2 Evidence for WGD in Yeast 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 PR OO 30 D 29 TE 28 EC 27 Although early analyses of genes potentially created by the vertebrate 2R events used phylogenetic approaches (Hughes 1999; Furlong and Holland 2002), most current studies of WGD rely on one or both of two methods: (1) finding numerous blocks of paralogous genes in multiple chromosomes with similar gene orders and (2) clustering homologs into groups by measuring the rate of synonymous substitutions (Ks or dS). This second method assumes that the gene pairs created by WGD cluster about some mean Ks value (Lynch and Conery 2000). The simultaneous application of both methods has been used to group multiple WGD events within species (e.g., Arabidopsis thaliana and Tetraodon nigroviridis; Jaillon et al. 2004; Van de Peer et al. 2009). However, the yeast genomes present an interesting challenge in this respect because the synonymous substitutions between yeast paralogs produced by WGD (hereafter ohnologs; Wolfe 2000) are often saturated (Byrne and Wolfe 2007). In other words, identical synonymous positions between two ohnologs occur almost as often due to repeated convergent substitutions as due to common ancestry, a fact pointed out by Smith (1987), who attempted to date histone gene duplicates in yeast. While the genomic structure of the core histone genes suggested that they were all duplicated simultaneously, these genes show CO RR 26 F 47 Susumo Ohno’s claim for the primacy of polyploidization in the generation of new adaptations (Ohno 1970), it is clear that WGD events have had a massive influence on the content and structure of the genomes of their possessors. The next step in exploring Ohno’s hypothesis is to link genome evolution to known changes in function. This goal, however, remains challenging, primarily because our knowledge of how genotype links to phenotype remains woefully incomplete (Pigliucci 2010). However, one group of organisms in which we can at least begin to make such associations is in the polyploid yeasts. Our knowledge of the functional genomics of yeast is drawn primarily from Saccharomyces cerevisiae, which has a well-annotated genome, decades of biochemical, genetic, and cell biology research, a relatively small genome, and a life cycle that lends itself to scalable laboratory analyses. Taken together, these facts have allowed yeast researchers not only to understand the structure of the genome following WGD but also to experimentally evaluate hypotheses regarding the evolution of particular complex phenotypes. As we will stress throughout this chapter, one of the themes that emerges from all of these analyses is the degree to which the outcome of a WGD depends as much on the interactions between genes as on the role of any particular locus. Of equal importance evolutionarily, we now also have data from other polyploid yeast species, which are valuable both as a point of comparison to S. cerevisiae and for their own sakes. This wealth of data affords us insight into the mechanisms that drive the preferential loss and preservation of gene duplicates after polyploidy, lead to the functional divergence of genes, and are behind the evolutionary origins of complex phenotypes. 25 UN Editor Proof 294 Book ISBN: 978-3-642-31441-4 Page: 294/307 Layout: T1 Standard SC Chapter No.: 15 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 F 71 72 PR OO 70 15.2.1 Synteny-Based Evidence for WGD in Saccharomyces cerevisiae D 69 TE 68 295 considerable variation in the numbers of synonymous substitutions separating them. This observation led Smith (1987) to hypothesize that S. cerevisiae underwent a WGD ancient enough that the duplicates surviving from it had saturated. However, it was not until the genome sequence of S. cerevisiae became available that this speculation could be confirmed (see below). Another similar but subtler problem in using Ks as a means of dating duplicate genes is the issue of gene conversion. Gene conversion was presumed to be quite common in yeast (Petes and Hill 1988), even prompting some authors (Gao and Innan 2004) to suggest that estimates of duplication rates based on duplicate divergences were inapplicable due to the homogenization of duplicate loci by conversion. Fortunately, although gene conversion is very common among yeast ribosomal proteins, it does not appear to be a general characteristic of the genome (Evangelisti and Conant 2010). Nonetheless, these various issues collectively meant that comparisons of paralogous sequence divergence were deemed unhelpful as a means to detect WGD in yeast. Given that paralogous sequence comparisons were generally unhelpful in finding WGD relics, another tactic was to consider gene order. In fact, even before the S. cerevisiae genome sequence was completed in 1996, it was clear to many researchers that it contained numerous, long, homologous clusters of ordered genes (Goffeau et al. 1996). Melnick and Sherman (1993) found ordered homologous gene clusters in chromosomes V and X covering 7.5 kb. Lalo et al. (1993) similarly found ordered homologous gene clusters in chromosomes XIV and III covering 15 kb. When the genome was sequenced, researchers found 18 ordered homologous genes in chromosomes IV and II that covered 120 and 170 kb, respectively (Goffeau et al. 1996). Just how to interpret these redundant regions remained a challenge at that time (Goffeau et al. 1996; Oliver 1996), and, in spite of Smith’s prior hypothesis (Smith 1987), few, if any, of the contemporaneous explanations included an ancient WGD. However, opinions changed the next year when Wolfe and Shields (1997) presented a thorough, genome sequence-based, analysis that gave strong evidence for WGD in S. cerevisiae. To find syntenic regions, they conducted a BLASTP search of amino acid sequences throughout the yeast genome and made a dot plot of the results. They then created gene blocks from these data, where each block was required to have at least three homologous pairs with intergenic distances B 50 kb and conservation of gene order and orientation. This analysis yielded 55 duplicated regions containing a total of 376 pairs of ohnologs. The large number of duplicated regions led Wolfe and Shields to posit two explanations: (1) successive independent gene duplications, and (2) a single duplication of the entire genome, EC 67 CO RR 66 Book ISBN: 978-3-642-31441-4 Page: 295/307 Yeast as a Window into Changes in Genome Complexity UN Editor Proof 15 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 15 Book ID: 272454_1_En Date: 16-8-2012 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 F 112 PR OO 110 111 15.2.2 Comparative Genomics and Proof of WGD in S. cerevisiae D 109 A number of researchers disputed the claims of Wolfe and Shields (1997), arguing that, because the syntenic regions identified made up only a small part of the genome, independent duplications better explained the genomic structure of S. cerevisiae (Coissac et al. 1997; Mewes et al. 1997; Hughes et al. 2000; Llorente et al. 2000a; Llorente et al. 2000b; Friedman and Hughes 2001; Piskur 2001; Koszul et al. 2004). However, this independent duplication hypothesis became untenable following the genome sequencing of other yeasts that proved to lack these syntenic paralog blocks. These sequences were described by three independent groups. The comparison of S. cerevisiae with Kluyveromyces waltii (Kellis et al. 2004) and the comparison of S. cerevisiae with Ashbya gossypii (Dietrich et al. 2004) involved different genomes, but effectively made the same argument: that the 2:1 mapping of blocks of paralogs from S. cerevisiae to homologous single-copy genes in K. waltii/A.gossypii could best be explained by WGD. This explanation was particularly striking because the doubly conserved synteny blocks cover 90 % of the genome in K. waltii (Kellis et al. 2004) and 96 % of that in A. gossypii (Dietrich et al. 2004). Furthermore, both studies found a large number of 2:1 pairing of centromeres in the species-respective chromosomes. There were 16:8 such pairings between S. cerevisiae and K. waltii and 14:7 between S. cerevisiae and A. gossypii with a subsequent break at the expected centromere position in S. cerevisiae chromosomes X and XII that are syntenic with regions in A. gossypii chromosomes I and III. Finally, and perhaps most strikingly, both groups also showed that the single-copy orthologs of genes from A. gossypii or K. waltii in the genome of S. cerevisiae are interleaved between two paralogous chromosomes in S. cerevisiae that nonetheless retain the relative gene order of the single chromosome in the non-WGD yeast (see Fig. 15.1). Such a pattern is only explicable under the hypothesis of a WGD event followed by massive gene losses. TE 108 followed by massive gene loss. There were two lines of evidence discounting the first possibility. First, 90 % (50/55) of the gene regions shared the same orientation with respect to the centromeres of the duplicated regions when we would expect independent duplications to be instead randomly distributed about the centromeres. Second, there were no examples of triplicated regions in the S. cerevisiae genome. If the duplications involved several distinct events separated in time, such a pattern would be highly unlikely, because it would require that later duplication events never overlapped with prior ones. Given these arguments, Wolfe and Shields (1997) argued for a single ancient WGD, which they dated to be hundreds of millions of years old (note that attempts to conclusively date this event have been difficult, due to a lack of fossils and the previously mentioned saturation of substitutions; see Taylor and Berbee 2006; Rolland and Dujon 2011). EC 107 CO RR 106 C. M. Hudson and G. C. Conant UN Editor Proof 296 Book ISBN: 978-3-642-31441-4 Page: 296/307 Layout: T1 Standard SC Chapter No.: 15 Book ISBN: 978-3-642-31441-4 Page: 297/307 Yeast as a Window into Changes in Genome Complexity 297 CO RR EC TE D PR OO F Editor Proof 15 Book ID: 272454_1_En Date: 16-8-2012 UN Fig. 15.1 Yeast gene order browser (YGOB) screenshots with a window size of six. Each box represents a gene; each color, a chromosome. The gene in focus, the A. gossypii gene ABR086W, is highlighted by an orange border. Each vertical column (‘‘pillar’’) represents a single gene prior to the WGD (hence, all genes in a column are homologs, and the paired upper and lower genes, when present, are paralogs). The ancestral order of these genes (pink boxes) just prior to the WGD has also been exhaustively inferred (Gordon et al. 2011). Connectors join nearby genes: a solid bar for adjacent genes, two bars for loci less than five genes apart, and one bar for loci \20 genes apart. The connectors are extended in gray over intervening space. The end of a chromosome or contig is denoted by a brace. Arrows denote transcriptional orientation. The browser also includes a control panel that allows users to select the window size and the gene to focus on. This panel also has buttons for running BLAST searches against YGOB’s database, outputting YGOB data in tabular format, obtaining pairwise Ka and Ks values among genes, and computing multiple sequence alignments and phylogenetic trees of individual pillars. Species names for each track are labeled at right (Byrne and Wolfe 2005) Layout: T1 Standard SC Chapter No.: 15 Book ID: 272454_1_En Date: 16-8-2012 C. M. Hudson and G. C. Conant 157 15.2.3 Yeast Gene Order Browser 148 149 150 151 152 153 154 155 PR OO 147 F 156 The argument of Dujon et al. (2004) is subtly different. They sequenced and analyzed four other genomes. One genome that of Candida glabrata, shares the genome duplication with S. cerevisiae. This was determined by comparing syntenic blocks in S. cerevisiae and C. glabrata with the other three sequenced genomes, Kluveromyces lactis, Debaryomyces hansenii, and Yarrowia lipolytica. Dujon et al. (2004) found 20 distinct blocks of paralogs shared by both S. cerevisiae and C. glabrata. These blocks allowed them to map the WGD onto a phylogeny, rather than do a simple pairwise comparison. Mapping this WGD phylogenetically creates distinct hypotheses as to where in the tree we expect to find polyploid yeasts (c.f., Fig. 15.1), predictions that have been confirmed with each of the subsequently sequenced genomes of known phylogenetic position (Wapinski et al. 2007; Scannell et al. 2011). 146 170 15.2.4 Additional Non-Saccharomyces-Specific WGDs 162 163 164 165 166 167 168 TE 161 EC 160 CO RR 159 D 169 One of the major benefits of studying the yeast WGD is that the relatively slow rates of gene order change in yeast genomes and the compactness of their genomes means that an exhaustive enumeration of all WGD-produced ohnologs is possible. Just such a project was carried out, with the results presented as the web-based Yeast Gene Order Browser (YGOB) (Byrne and Wolfe 2005), which illustrates a number of non- and post-WGD yeasts in a graphical framework (Fig. 15.2). This work has been followed by a reconstruction of the set of genes and their relative orders that existed just prior to the WGD (Gordon et al. 2009) and by a likelihood-based model of post-WGD duplicate loss that attempts to quantify the orthology inferences made by YGOB (Conant and Wolfe 2008a). On the basis of these three projects, the post-WGD evolutionary history of virtually every locus in the S. cerevisiae genome can be traced (Fig. 15.1 is thus illustrative of the predominant pattern seen across the genome). 158 174 In addition to the ancient WGD that characterizes the Saccharomyces clade (Fig. 15.2), several cases of allopolyploidy have been discovered in yeasts. Some of these occur in species within the Saccharomyces sensu stricto clade (Scannell et al. 2011), while others are independent. 175 15.2.4.1 Secondary Allopolyploidy in Saccharomyces pastorianus 171 172 173 176 177 178 179 UN Editor Proof 298 Book ISBN: 978-3-642-31441-4 Page: 298/307 Several cases of allopolyploidy are known from within S. sensu stricto (Dequin and Casaregola 2011). One of the most well studied is that of the lager yeast, S. pastorianus (syn. S. carlsbergensis). It has long been known that the polyploid S. pastorianus and other members of the complex of related lager yeasts are Layout: T1 Standard SC Chapter No.: 15 Book ISBN: 978-3-642-31441-4 Page: 299/307 Yeast as a Window into Changes in Genome Complexity 299 PR OO F Editor Proof 15 Book ID: 272454_1_En Date: 16-8-2012 184 185 186 187 188 189 190 191 192 193 194 195 196 197 198 199 200 201 EC 183 CO RR 182 allotetraploids of diploid S. cerevisiae and some other unknown diploid species (Martini and Kurtzman 1985; Kielland-Brandt et al. 1995). However, aside from the general difficulties facing anyone interested in identifying the origins of hybrid genomes, the debate surrounding the origin of the second parental diploid species was further complicated by a difficulty in delimiting species within these groups (Rainieri et al. 2006). The tetraploid S. pastorianus belongs to a group of yeast species which, until recently, was represented as a phylogenetically unresolved species complex including S. pastorianus, S. monacensis (S. pastorianus strain CBS 1503), S. bayanus, and S. bayanus var. uvarum (Casaregola et al. 2001; Rainieri et al. 2006). This taxonomic confusion has recently been partially resolved through the sequencing of the genomes of both S. pastorianus and one of its presumed parental diploid species, S. eubayanus (Nakao et al. 2009). The genome history that has emerged is a complicated story of allopolyploidy followed by genomic transformation forming the related species S. bayanus (Libkind et al. 2011). As summarized by Libkind et al. (2011), S. cerevisiae hybridized with S. eubayanus (a species recently recovered in Patagonia) with subsequent genome doubling producing the allotetraploid progenitor of modern S. pastorianus. Following domestication, smaller regions of the S. pastorianus genome were then apparently transferred into the genome of the diploid parent S. eubayanus (which is nonetheless a descendant of the ancient polyploidy). This hybrid form of S. eubayanus, with contributions from S. pastorianus, then proceeded to interbreed with diploid S. uvarum to produce the modern, diploid, S. bayanus (Fig. 15.3). UN 180 181 TE D Fig. 15.2 Consensus view of the evolutionary relationships between the yeast taxa discussed. Black branches indicate relationships described by both Kurtzman and Robnett (2003) and Fitzpatrick et al. (2006). Red branches indicate conflicts between the two phylogenies, in which case Fitzpatrick et al. (2006) is presented. Curved grey branches illustrate the allopolyploidy events between two species (S. cerevisiae and S. bayanus, Z. rouxii, and Z. pseudorouxii). Taxa in blue are reported in text (e.g., S. pastorianus and Z. rouxii ATCC 42981). Stars mark wholegenome duplications. Note that genus names are an imperfect guide to the relationships Layout: T1 Standard SC Chapter No.: 15 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 300/307 Editor Proof 300 C. M. Hudson and G. C. Conant natural environments brewing environment early S. pastorianus 1 50:50 hybrid of S. eubayanus and S.cerevisiae modern S. pastorianus 3 S. pastorianus cells lyse and release large DNA fragments 4 transformation of S. eubayanus by S. pastorianus 2 SUL1 inactivation, loss-of-heterozygosity on the right arm of chromosome VII adds an extra copy of ScerIMA1, aneuploidy and chromosome rearrangements S. bayanus F S. eubayanus brewing contaminants domestication PR OO ale-type S. cerevisiae NBRC 1948 CBS 380 S. uvarum 5 independent hybridizations with S. uvarum functional SUL1 inactive SUL1 202 TE D Fig. 15.3 Genome evolution in S. pastorianus. A model of the formation of S. pastorianus and the hybrid strains of S. bayanus. First, wild S. eubayanus and ale-type S. cerevisiae hybridized to form an allotetraploid that became the ancestor of the modern (doubly paleopolyploid) S. pastorianus. Second, domestication imposed strong selective pressure for strains with the most desirable brewing properties. Third, in the brewing vats with high densities of S. pastorianus, cell lysis releases large DNA fragments that occasionally transform, fourth, contaminating wild strains of S. eubayanus (which possesses only the ancient WGD shared with S. cerevisiae) because of the lack of pure culture techniques. Fifth, multiple hybridization events between S. eubayanus and wild strains of S. uvarum gave rise to CBS 380T and NBRC 1948. This model does not exclude prior or parallel involvement of S. uvarum in brewing or contamination. Reprinted from Libkind et al. (2011) 15.2.4.2 Allopolyploidy in Zygosaccharomyces rouxii 215 15.3 WGD and Speciation 206 207 208 209 210 211 212 213 216 217 218 219 220 CO RR 205 UN 204 EC 214 The spoilage agent and industrial yeast Zygosaccharomyces rouxii strain ATCC 42981 was identified as another allopolyploid by James et al. (2005) and Gordon and Wolfe (2008). This hybridization/polyploidy event is significant for two reasons. First, unlike all of the previous examples, it occurs outside of S. sensu stricto. Second, Gordon and Wolfe (2008) determined that most of the paralogs produced by WGD are still present, presumably due to the recentness of the event. Thus, while other yeast genome duplications are ancient and show considerable gene loss and rearrangement (Wolfe and Shields 1997), the Z. rouxii genome retains most of the ‘‘new’’ genes produced by its WGD. Since the survival time of ohnologs has been modeled to follow a power law, most of the duplicates are expected to be lost very rapidly (Maere et al. 2005), suggesting that Z. rouxii represents an example of the early features of genome evolution following WGD. 203 An important potential outcome of polyploidy is in altering patterns of speciation. This change can happen in at least two ways. First, the WGD can relax selective constraints resulting in an adaptive radiation by means of ecological speciation. Another, more neutral mechanism, is a special case of the Dobzhansky-Muller (DM) process of speciation, in which species lose reciprocal paralogs following Layout: T1 Standard SC Chapter No.: 15 227 228 229 230 231 232 233 234 235 236 237 238 239 240 241 242 243 244 245 246 247 248 249 250 251 252 253 254 255 256 257 258 259 260 261 262 263 264 F 226 PR OO 225 D 224 TE 223 301 some period of isolation (Lynch and Force 2000; Werth and Windham 1991; see also Chap. 1, this volume). WGD potentially increases the probability of this simply by increasing the number of paired genes in a genome. The fertility of hybrids is 0.75n, where n is the number of reciprocal losses of essential genes among populations (Werth and Windham 1991). Clearly, for any significant number of reciprocal losses (such as occur after WGD), the number of viable, fertile offspring of a crossing of two such populations is negligible. Both phylogenetic and experimental studies of the DM process after WGD have been carried out in yeast. Scannell et al. (2006) showed that the number of reciprocal gene losses in several species of yeast sharing the S. cerevisiae WGD was sufficient to induce such inviability. This observation suggests that a DM mechanism was partly responsible for the multiple speciation events among the Saccharomyces species (e.g., S. cerevisiae, S. bayanus, and C. glabrata) following WGD. An advantage of studying the DM process in yeast is the ability to experimentally create and cross artificial polyploids. This possibility has been highlighted in experimental studies of reproductive isolation. Polyploid yeasts have been allowed to evolve in different selective environments (Dettman et al. 2007) and in neutral environments subject to random mutagenesis (Maclean and Greig 2011). These two experiments have shown that moderate reproductive isolation, coupled with reciprocal gene loss, results in a clear loss of fitness when independently derived polyploids are crossed. Similarly, Lee et al. (2008) showed that hybrids of S. cerevisiae and S. bayanus were less fit than their parental phenotypes, due primarily to incompatibility between their nuclear and mitochondrial genomes. Chou et al. (2010) extended this analysis, providing another pair of mitochondrial and nuclear genes and posited nuclear-mitochondrial incompatibility as a common mechanism in species formation. In another twist, Anderson et al. (2010) demonstrated the existence of alleles with depressed hybrid fitness in lowglucose environments, which argues for a model in which neutral changes in paired genes are followed by strong selection, a sequence of events that promotes rapid reproductive isolation. Kao et al. (2010), however, argue against the existence of a small number of so-called speciation genes, instead claiming that genome scans provide no evidence of any single paired dominant or recessive genic incompatibilities. They instead argue that following WGD, many changes in loci of little effect resulted in lowered fitness due, in part, to the rewiring of transcriptional and metabolic networks. Another debate that has emerged in this field is whether these changes are due primarily to the decrease in the fertility of hybrids (Xu and He 2011) or a decrease in their viability (Greig 2008). This question ultimately amounts to a debate about what stage in the yeast life cycle the genetic incompatibilities occur—sporulation or clonal growth, and whether the decrease in fitness is the result of competition for resources or offspring. The discontinuity between these ideas likely represents an opportunity to explain speciation as a process across different genomic and temporal scales, and we would speculate that the process of DM incompatibility induces selection for the evolution of some form of prezygotic barrier. EC 222 CO RR 221 Book ISBN: 978-3-642-31441-4 Page: 301/307 Yeast as a Window into Changes in Genome Complexity UN Editor Proof 15 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 15 Book ID: 272454_1_En Date: 16-8-2012 271 272 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 293 294 295 296 297 298 299 300 301 302 303 304 305 306 307 F 270 PR OO 269 D 268 Duplicate retention and evolutionary models. In addition to such population processes as speciation, WGD also altered many other aspects of the S. cerevisiae lifestyle. For instance, several pairs of ohnologs have been shown to have undergone various types of functional divergence, allowing the study of some of the proposed mechanisms of duplicate divergence after duplication (Conant and Wolfe 2008b). In an elegant series of experiments, van Hoof (2005) showed that two ohnologs, ORC1 and SIR3, have distinct and non-overlapping functions (in DNA replication and gene silencing, respectively). Strikingly, however, the mutual ortholog of these genes from the non-WGD yeast S. kluyveri is able to complement both functions, constituting a clear example of subfunctionalization. An apparently similar case, involving the S. cerevisiae ohnolog pair GAL1 and GAL3, which presently functions, respectively, as an enzyme and as a transcriptional regulator, was complicated by the discovery of an adaptive conflict between the shared regulator and enzymatic function of their ortholog in the non-WGD K. lactis. Thus, although the K. lactis GAL1 gene does indeed serve the functions of both GAL1 and GAL3 in S. cerevisiae, it does so in a suboptimal way, being unable to tune its expression to both roles simultaneously (Hittinger and Carroll 2007). This conflict illustrates an important point about subfunctionalization, namely that the original neutral model of subfunctionalization proposed by Force and coauthors (Force et al. 1999) is not the only possible mechanism for such functional partitioning (Des Marais and Rausher 2008). Other examples of divergence among ohnologs where the mechanism of that divergence is less clear include ribosomal proteins (Ni and Snyder 2001; Komili et al. 2007; Kim et al. 2009), glucose sensors (Özcan et al. 1998), and glycolysis enzymes (Boles et al. 1997). The dosage balance hypothesis (DBH). In addition to facilitating the above work, the wealth of functional data from S. cerevisiae also provides an excellent opportunity to test hypotheses explaining the differences in gene retention patterns after WGD and small-scale duplications (hereafter SSD). Chief among these is probably the DBH (Papp et al. 2003; Freeling and Thomas 2006; Birchler and Veitia 2007; Freeling 2009), which states that, in eukaryotes, there is selection operating to disfavor duplications of central network genes due to the imbalance in network stoichiometry that results. This situation is reversed for WGD because in that case, the loss of a second copy of a gene introduces imbalances relative to the remaining duplicated genes. In keeping with the DBH, several classes of genes are over-retained after several evolutionarily ancient WGD events, including that in yeast. They include ribosomal proteins, protein kinases and transcription factors (Seoighe and Wolfe 1999; Blanc and Wolfe 2004; Maere et al. 2005; Aury et al. 2006; Conant and Wolfe 2008a). Similarly, genes that tend to have been fixed by WGD are less likely to have undergone SSD in other yeast species (Wapinski et al. 2007). However, duplicates produced by WGD have more protein interactions (Guan et al. 2007; Hakes et al. 2007), more phosphorylation sites (Amoutzias et al. 2010), and tend to be highly expressed (Seoighe and Wolfe 1999) than those from TE 267 EC 266 15.4 Changes in Genome Content and Complexity Post-WGD CO RR 265 C. M. Hudson and G. C. Conant UN Editor Proof 302 Book ISBN: 978-3-642-31441-4 Page: 302/307 Layout: T1 Standard SC Chapter No.: 15 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 F 313 PR OO 312 D 311 TE 310 303 SSD. Although genes retained in duplicate after WGD are rarely essential on an individual basis (Guan et al. 2007), this dispensability appears to be due to functional compensation by the other ohnolog (DeLuna et al. 2008). Thus, it appears that while ohnologs are less likely to be essential than their SSD counterparts today, their ancestral genes were actually at least as essential as current single-copy genes (DeLuna et al. 2008). System-level changes produced by WGD. Of course, one of the unique features of polyploidy relative to SSD is the possibility of coordinated changes among multiple sets of ohnologs. At the simplest level, we have previously illustrated examples of what appears to be network subfunctionalization where a number of ohnologs collectively divided two expression domains among themselves (Conant and Wolfe 2006). A more complex and interesting example is the role of the WGD (Piškur et al. 2006) in shaping S. cerevisiae’s propensity for aerobic glucose fermentation (the Crabtree effect; Geladé et al. 2003; Johnston and Kim 2005), a novel and somewhat paradoxical phenotype. There is a general association between the presence of the WGD and the Crabtree effect across yeast species (Merico et al. 2007). As a result, we and others have argued that dosage effects among the glycolysis enzymes post-WGD helped to increase flux through glycolysis (Blank et al. 2005; Kuepfer et al. 2005; Conant and Wolfe 2007; Merico et al. 2007; van Hoek and Hogeweg 2009). Such increased flux likely could only be accommodated through fermentation pathways, given the complex spatial organization of the competing respiratory pathway (Conant and Wolfe 2007). Supporting this hypothesis is an elegant computational analysis by van Hoek and Hogeweg (2009) showing that future WGD events in modern S. cerevisiae could also be expected to provide a selective advantage in glucose-rich environments through the preferential retention of duplicated glycolysis enzymes. Note that the apparently ‘‘wasteful’’ fermentation can actually be selectively advantageous in the context of rich but ephemeral resource patches (Pfeiffer et al. 2001; Pfeiffer and Schuster 2005), a phenomenon that has been experimentally confirmed in yeast (MacLean and Gudelj 2006). Such a change in the yeast lifestyle likely led to other, later changes in the genome. One suggestive example concerns the decoupling of cytosolic and mitochondrial ribosomal protein expression post-WGD (Ihmels et al. 2005). Prior to WGD, bakers’ yeast was likely similar to other yeasts in having a strong association in the expression of the two types of ribosomal proteins. After WGD, however, cisregulatory element evolution diverged in the two groups of genes (Ihmels et al. 2005), allowing S. cerevisiae to express only cytosolic proteins at high levels during fermentation, an important refinement in a fermentative lifestyle. Connecting the DBH to large-scale evolutionary changes following WGD, Conant (2010) and Fusco et al. (2010) found transcriptional regulatory motifs to be over-retained in ohnologs. Modeling network evolution after WGD, these authors find the network enriched for transcription factors and particular network motifs. Duplicated transcription factors still show some relics of the WGD, being more likely to share targets than are random transcription factors, but on the whole show considerable divergence post-WGD (Conant 2010). Given this rapid regulatory evolution, it may not be easy to ascertain the role of WGD in the evolution of EC 309 CO RR 308 Book ISBN: 978-3-642-31441-4 Page: 303/307 Yeast as a Window into Changes in Genome Complexity UN Editor Proof 15 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 15 Book ID: 272454_1_En Date: 16-8-2012 C. M. Hudson and G. C. Conant 357 the modern S. cerevisiae regulatory network. Nonetheless, the retention of many transcription factors that have acquired distinct sets of target genes may imply that the WGD served to ‘‘relax’’ the regulatory complexity of this organism, which may have implications for its future ability to adapt (as seen for the GAL1/GAL3 example). 358 15.5 Conclusions 353 354 355 PR OO F 356 370 371 372 373 Acknowledgments We would like to thank Michaël Bekaert, Patrick Edger, and Chris Pires for helpful discussions. This work was supported by the National Library of Medicine Biomedical and Health Informatics Training Fellowship [LM007089-19] (CMH) and the Reproductive Biology Group of the Food for the twenty-first century program at the University of Missouri (GCC). 374 References 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 Amoutzias GD, He Y, Gordon J, Mossialos D, Oliver SG, Van de Peer Y (2010) Posttranslational regulation impacts the fate of duplicated genes. Proc Natl acad sci U S A 107:2967–2971 Anderson JB, Funt J, Thompson DA et al (2010) Determinants of divergent adaptation and Dobzhansky-Muller interaction in experimental yeast populations. Curr Biol 20:1383–1388 Aury JM, Jaillon O, Duret L et al (2006) Global trends of whole-genome duplications revealed by the ciliate Paramecium tetraurelia. Nature 444:171–178 Birchler JA, Veitia RA (2007) The gene balance hypothesis: from classical genetics to modern genomics. Plant Cell 19:395–402 Blanc G, Wolfe KH (2004) Functional divergence of duplicated genes formed by polyploidy during Arabidopsis evolution. Plant Cell 16:1679–1691 Blank LM, Lehmbeck F, Sauer U (2005) Metabolic-flux and network analysis of fourteen hemiascomycetous yeasts. FEMS Yeast Res 5:545–558 Boles E, Schulte F, Miosga T, Freidel K, Schlüter E, Zimmermann FK, Hollenberg CP, Heinisch JJ (1997) Characterization of a glucose-repressed pyruvate kinase (Pyk2p) in Saccharomyces cerevisiae that is catalytically insensitive to fructose-1-6-biphosphate. J Bacteriol 179:2987–2993 Byrne KP, Wolfe KH (2005) The yeast gene order browser: combining curated homology and syntenic context reveals gene fate in polyploid species. Genome Res 15:1456–1461 364 365 366 367 368 TE 362 363 EC 361 CO RR 360 D 369 The S. cerevisiae WGD has been implicated in a number of evolutionarily complex events. At a minimum, a set of duplicated genes of identical age is a powerful system for exploring duplicate gene evolution (van Hoof 2005; Conant and Wolfe 2006; Fares et al. 2006; Kim and Yi 2006). However, we also suggest that, as with the GAL1/GAL3 example, we will not fully understand the biology of S. cerevisiae until we account for how the WGD has altered both the individual roles of particular genes and their relationships to each other. We have outlined some of the areas of yeast biology that we think were altered by this genome-doubling event: there remain others yet to be discovered. Similarly, the presence of other WGD events, of varying ages, allows us to study how these events unfold over various timescales, including, potentially, on the timescale of laboratory experiments in evolution. 359 UN Editor Proof 304 Book ISBN: 978-3-642-31441-4 Page: 304/307 Layout: T1 Standard SC Chapter No.: 15 305 EC TE D PR OO F Byrne KP, Wolfe KH (2007) Consistent patterns of rate asymmetry and gene loss indicate widespread neofunctionalization of yeast genes after whole-genome duplication. Genetics 175:1341–1350 Casaregola S, Nguyen HV, Lapathitis G, Kotyk A, Gaillardin C (2001) Analysis of the constitution of the beer yeast genome by PCR, sequencing and subtelomeric sequence hybridization. Int J Syst Evol Microbiol 51:1607–1618 Chou J-Y, Hung Y-S, Lin K-H, Lee H-Y, Leu J-Y (2010) Multiple molecular mechanisms cause reproductive isolation between three yeast species. PLoS Biol 8:e1000432 Coissac E, Maillier E, Netter P (1997) A comparative study of duplications in bacteria and eukaryotes: the importance of telomeres. Mol Biol Evol 14:1062–1074 Conant GC (2010) Rapid reorganization of the transcriptional regulatory network after genome duplication in yeast. Proc R Soc B 277:869–876 Conant GC, Wolfe KH (2006) Functional partitioning of yeast co-expression networks after genome duplication. PLoS Biol 4:e109 Conant GC, Wolfe KH (2007) Increased glycolytic flux as an outcome of whole-genome duplication in yeast. Mol Syst Biol 3:129 Conant GC, Wolfe KH (2008a) Probabilistic cross-species inference of orthologous genomic regions created by whole-genome duplication in yeast. Genetics 179:1681–1692 Conant GC, Wolfe KH (2008b) Turning a hobby into a job: how duplicated genes find new functions. Nat Rev Genet 9:938–950 DeLuna A, Vetsigian K, Shoresh N, Hegreness M, Colón-González M, Chao S, Kishony R (2008) Exposing the fitness contribution of duplicated genes. Nat Genet 40:676–681 Dequin S, Casaregola S (2011) The genomes offermentative Saccharomyces. CR Biol 334:687–693 Des Marais DL, Rausher MD (2008) Escape from adaptive conflict after duplication in an anthocyanin pathway gene. Nature 454:762–765 Dettman JR, Sirjusingh C, Kohn LM, Anderson JB (2007) Incipient speciation by divergent adaptation and antagonistic epistasis in yeast. Nature 447:585–588 Dietrich FS, Voegeli S, Brachat S et al (2004) The Ashbya gossypii genome as a tool for mapping the ancient Saccharomyces cerevisiae genome. Science 304:304–307 Dujon B, Sherman D, Fischer G et al (2004) Genome evolution in yeasts. Nature 430:35–44 Evangelisti AM, Conant GC (2010) Nonrandom survival of gene conversions among yeast ribosomal proteins duplicated through genome doubling. Genome Biol Evol 2:826–834 Fares MA, Byrne KP, Wolfe KH (2006) Rate asymmetry after genome duplication causes substantial long-branch attraction artifacts in the phylogeny of Saccharomyces species. Mol Biol Evol 23:245–253 Fitzpatrick D, Logue M, Stajich J, Butler G (2006) A fungal phylogeny based on 42 complete genomes derived from supertree and combined gene analysis. BMC Evol Biol 6:99 Force A, Lynch M, Pickett FB, Amores A, Yan Y, Postlethwait J (1999) Preservation of duplicate genes by complementary, degenerative mutations. Genetics 151:1531–1545 Freeling M (2009) Bias in plant gene content following different sorts of duplication: tandem, whole-genome, segmental, or by transposition. Annu Rev Plant Biol 60:433–453 Freeling M, Thomas BC (2006) Gene-balanced duplications, like tetraploidy, provide predictable drive to increase morphological complexity. Genome Res 16:805–814 Friedman R, Hughes AL (2001) Gene duplication and the structure of eukaryotic genomes. Genome Res 11:373–381 Furlong RF, Holland PWH (2002) Were vertebrates octoploid? Philos Trans R Soc Lond B 357:531–544 Fusco D, Grassi L, Bassetti B, Caselle M, Lagomarsino MC (2010) Ordered structure of the transcription network inherited from the yeast whole-genome duplication. BMC Syst Biol 4:77 Gao LZ, Innan H (2004) Very low gene duplication rate in the yeast genome. Science 306: 1367–1370 Geladé R, Van De Velde S, Van Dijck P, Thevelein JM (2003) Multi-level response of the yeast genome to glucose. Genome Biol 4:233 Goffeau A, Barrell B, Russey H et al (1996) Life with 6000 genes. Science 274:562–567 CO RR 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 Book ISBN: 978-3-642-31441-4 Page: 305/307 Yeast as a Window into Changes in Genome Complexity UN Editor Proof 15 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 15 Book ID: 272454_1_En Date: 16-8-2012 EC TE D PR OO F Gordon JL, Byrne KP, Wolfe KH (2009) Additions, losses and rearrangements on the evolutionary route from a reconstructed ancestor to the modern Saccharomyces cerevisiae genome. PLoS Genet 5:e1000485 Gordon JL, Byrne KP, Wolfe KH (2011) Mechanisms of chromosome number evolution in yeast. PLoS Genet 7:e1002190 Gordon JL, Wolfe KH (2008) Recent allopolyploid origin of Zygosaccharomyces rouxii strain ATCC 42981. Yeast 25:449–456 Greig D (2008) Reproductive isolation in Saccharomyces. Heredity 102:39–44 Guan Y, Dunham MJ, Troyanskaya OG (2007) Functional analysis of gene duplications in Saccharomyces cerevisiae. Genetics 175:933–943 Hakes L, Pinney JW, Lovell SC, Oliver SG, Robertson DL (2007) All duplicates are not equal: the difference between small-scale and genome duplication. Genome Biol 8:R209 Hittinger CT, Carroll SB (2007) Gene duplication and the adaptive evolution of a classic genetic switch. Nature 449:677–681 Hughes AL (1999) Phylogenies of developmentally important proteins do not support the hypothesis of two rounds of genome duplication early in vertebrate history. J Mol Evol 48:565–576 Hughes TR, Roberts CJ, Dai H et al (2000) Widespread aneuploidy revealed by DNA microarray expression profiling. Nat Genet 25:333–337 Ihmels J, Bergmann S, Gerami-Nejad M, Yanai I, McClellan M, Berman J, Barkai N (2005) Rewiring of the yeast transcriptional network through the evolution of motif usage. Science 309:938–940 Jaillon O, Aury J-M, Brunet F et al (2004) Genome duplication in the teleost fish Tetraodon nigroviridis reveals the early vertebrate proto-karyotype. Nature 431:946–957 James SA, Bond CJ, Stratford M, Roberts IN (2005) Molecular evidence for the existence of natural hybrids in the genus Zygosaccharomyces. FEMS Yeast Res 5:747–755 Johnston M, Kim J-H (2005) Glucose as a hormone: Receptor-mediated glucose sensing in the yeast Saccharomyces cerevisiae. Biochem Soc Trans 33:247–252 Kao KC, Schwartz K, Sherlock G (2010) A genome-wide analysis reveals no nuclear Dobzhansky-Muller pairs of determinants of speciation between S. cerevisiae and S. paradoxus, but suggests more complex incompatibilities. PLoS Genet 6:e1001038 Kellis M, Birren BW, Lander ES (2004) Proof and evolutionary analysis of ancient genome duplication in the yeast Saccharomyces cerevisiae. Nature 428:617–624 Kielland-Brandt MC, Nilsson-Tillgren T, Gjermansen C, Holmberg S, Pedersen MB (1995) Genetics of brewing yeasts. In: Rose AH, Wheals AE, Harrison JS (eds) The Yeasts, vol 6, 2nd edn. Academic, London, pp 223–254 Kim S-H, Yi SV (2006) Correlated asymmetry of sequence and functional divergence between duplicate proteins of Saccharomyces cerevisiae. Mol Biol Evol 23:1068–1075 Kim T-Y, Ha CW, Huh W-K (2009) Differential subcellular localization of ribosomal protein L7 paralogs in Saccharomyces cerevisiae. Mol Cells 27:539–546 Komili S, Farny NG, Roth FP, Silver PA (2007) Functional specificity among ribosomal proteins regulates gene expression. Cell 131:557–571 Koszul R, Caburet S, Dujon B, Fischer G (2004) Eucaryotic genome evolution through the spontaneous duplication of large chromosomal segments. EMBO J 23:234–243 Kuepfer L, Sauer U, Blank LM (2005) Metabolic functions of duplicate genes in Saccharomyces cerevisiae. Genome Res 15:1421–1430 Kurtzman C, Robnett C (2003) Phylogenetic relationships among yeasts of the ‘Saccharomyces complex’ determined from multigene sequence analyses. FEMS Yeast Res 3:417–432 Lalo D, Stettler S, Mariotte S, Slominski PP, Thuriaux P (1993) Two yeast chromosomes are related by a fossil duplication of their centromeric regions. C R Acad Sci 316:367–373 Lee H-Y, Chou J-Y, Cheong L, Chang N-H, Yang S-Y, Leu J-Y (2008) Incompatibility of nuclear and mitochondrial genomes causes hybrid sterility between two yeast species. Cell 135:1065–1073 CO RR 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 C. M. Hudson and G. C. Conant UN Editor Proof 306 Book ISBN: 978-3-642-31441-4 Page: 306/307 Layout: T1 Standard SC Chapter No.: 15 307 EC TE D PR OO F Libkind D, Hittinger CT, Valério E, Gonçalves C, Dover J, Johnston M, Gonçalves P, Sampaio JP (2011) Microbe domestication and the identification of the wild genetic stock of lagerbrewing yeast. Proc Nat Acad Sci 108:14539–14544 Llorente B, Durrens P, Malpertuy A et al (2000a) Genomic exploration of the hemiascomycetous yeasts: 20. Evolution of gene redundancy compared to Saccharomyces cerevisiae. FEBS Lett 487:122–133 Llorente B, Malpertuy A, Neuvéglise C et al (2000b) Genomic exploration of the hemiascomycetous yeasts: 18. Comparative analysis of chromosome maps and synteny with Saccharomyces cerevisiae. FEBS Lett 487:101–112 Lynch M, Conery JS (2000) The evolutionary fate and consequences of duplicate genes. Science 290:1151–1155 Lynch M, Force AG (2000) The origin of interspecific genomic incompatibility via gene duplication. Am Nat 156:590–605 Maclean CJ, Greig D (2011) Reciprocal gene loss following experimental whole-genome duplication causes reproductive isolation in yeast. Evolution 65:932–945 MacLean RC, Gudelj I (2006) Resource competition and social conflict in experimental populations of yeast. Nature 441:498–501 Maere S, De Bodt S, Raes J, Casneuf T, Van Montagu M, Kuiper M, Van de Peer. Y (2005) Modeling gene and genome duplications in eukaryotes. Proc Nat Acad Sci U S A 102:5454–5459 Martini AV, Kurtzman CP (1985) Deoxyribonucleic acid relatedness among species of the genus Saccharomyces sensu stricto. Int J Syst Bacteriol 35:508–511 Melnick L, Sherman F (1993) The gene clusters ARC and COR on chromosomes 5 and 10, respectively, of Saccharomyces cerevisiae share a common ancestry. J Mol Biol 233:372–388 Merico A, Sulo P, Piškur J, Compagno C (2007) Fermentative lifestyle in yeasts belonging to the Saccharomyces complex. FEBS J 274:976–989 Mewes H, Albermann K, Bähr M et al (1997) Overview of the yeast genome. Nature 387:7–65 Nakao Y, Kanamori T, Itoh T, Kodama Y, Rainieri S, Nakamura N, Shimonaga T, Hattori M, Ashikari T (2009) Genome sequence of the lager brewing yeast, an interspecies hybrid. DNA Res 16:115–129 Ni L, Snyder M (2001) A genomic study of the bipolar bud site selection pattern in Saccharomyces cerevisiae. Mol Biol Cell 12:2147–2170 Ohno S (1970) Evolution by gene duplication. Springer, New York Oliver SG (1996) From DNA sequence to biological function. Nature 379:597–600 Özcan S, Dover J, Johnston M (1998) Glucose sensing and signaling by two glucose receptors in the yeast Saccharomyces cerevisiae. EMBO J 17:2566–2573 Papp B, Pal C, Hurst LD (2003) Evolution of cis-regulatory elements in duplicated genes of yeast. Trends Genet 19:417–422 Petes T, Hill C (1988) Recombination between repeated genes in microorganisms. Annu Rev Genet 22:147–168 Pfeiffer T, Schuster S (2005) Game-theoretical approaches to studying the evolution of biochemical systems. Trends Biochem Sci 30:20–25 Pfeiffer T, Schuster S, Bonhoeffer S (2001) Cooperation and competition in the evolution of ATP-producing pathways. Science 292:504–507 Pigliucci M (2010) Genotype-phenotype mapping and the end of the ‘genes as blueprint’ metaphor. Philos Trans R Soc Lond B 365:557–566 Piskur J (2001) Origin of the duplicated regions in the yeast genomes. Trends Genet 17:302–303 Piškur J, Rozpedowska E, Polakova S, Merico A, Compagno C (2006) How did Saccharomyces evolve to become a good brewer? Trends Genet 22:183–186 Rainieri S, Kodama Y, Kaneko Y, Mikata K, Nakao Y, Ashikari T (2006) Pure and mixed genetic lines of Saccharomyces bayanus and Saccharomyces pastorianus and their contribution to the lager brewing strain genome. Appl Environ Microbiol 72:3968–3974 Rolland T, Dujon B (2011) Yeasty clocks: dating genomic changes in yeasts. CR Biol 334:620–628 Scannell DR, Byrne KP, Gordon JL, Wong S, Wolfe KH (2006) Multiple rounds of speciation associated with reciprocal gene loss in polyploid yeasts. Nature 440:341–345 CO RR 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 Book ISBN: 978-3-642-31441-4 Page: 307/307 Yeast as a Window into Changes in Genome Complexity UN Editor Proof 15 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 15 Book ID: 272454_1_En Date: 16-8-2012 EC TE D PR OO F Scannell DR, Zill OA, Rokas A, Payen C, Dunham MJ, Eisen MB, Rine J, Johnston M, Hittinger CT (2011) The awesome power of yeast evolutionary genetics: new genome sequences and strain resources for the Saccharomyces sensu stricto genus. G3: Genes, Genomes, Genetics 1:11–25 Seoighe C, Wolfe KH (1999) Yeast genome evolution in the post-genome era. Curr Opin Microbiol 2:548–554 Smith M (1987) Molecular evolution of the Saccharomyces cerevisiae histone gene loci. J Mol Evol 24:252–259 Taylor JW, Berbee ML (2006) Dating divergences in the fungal tree of life: review and new analyses. Mycologia 98:838–849 Van de Peer Y, Fawcett J, Proost S, Sterk L, Vandepoele K (2009) The flowering world: a tale of duplications. Trends Plant Sci 14:680–688 van Hoek MJ, Hogeweg P (2009) Metabolic adaptation after whole genome duplication. Mol Biol Evol 26:2441–2453 van Hoof A (2005) Conserved functions of yeast genes support the duplication, degeneration and complementation model for gene duplication. Genetics 171:1455–1461 Wapinski I, Pfeffer A, Friedman N, Regev A (2007) Natural history and evolutionary principles of gene duplication in fungi. Nature 449:54–61 Werth CR, Windham MD (1991) A model for divergent, allopatric speciation of polyploid pteridophytes resulting from silencing of duplicate-gene expression. Am Nat 137:515–526 Wolfe KH (2000) Robustness-it’s not where you think it is. Nat Genet 25:3–4 Wolfe KH, Shields D (1997) Molecular evidence for an ancient duplication of the entire yeast genome. Nature 387:708–713 Xu M, He X (2011) Genetic incompatibility dampens hybrid fertility more than hybrid viability: yeast as a case study. PLoS ONE 6:e18341 CO RR 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 C. M. Hudson and G. C. Conant UN Editor Proof 308 Book ISBN: 978-3-642-31441-4 Page: 308/307 Metadata of the chapter that will be visualized in SpringerLink Book Title Polyploidy and Genome Evolution Series Title Chapter Title Two Rounds of Whole-Genome Duplication: Evidence and Impact on the Evolution of Vertebrate Innovations Copyright Year 2012 Copyright HolderName Springer-Verlag Berlin Heidelberg Corresponding Author Family Name Cañestro Particle Given Name Cristian Suffix Abstract Division Departament de Genètica, Facultat de Biologia Organization Universitat de Barcelona Address Av. Diagonal, 643, edifici Prevosti, 2a planta, 08028, Barcelona, Spain Email canestro@ub.edu The origin and evolution of the vertebrates have been linked to the study of genome duplications since Susumo Ohno ventured the 2R-hypothesis, suggesting that the successful diversification of complex vertebrates was facilitated by polyploidization in the stem vertebrate ancestor due to two rounds of whole-genome duplication (2R-WGD). This chapter first reviews evidence supporting Ohno’s 2R-hypothesis and gathers information about the timing and mechanisms underlying the 2R-WGD. Second, this chapter describes the impact of the 2R-WGD on the evolution of the vertebrate genome structure, gene number, and the evolutionary dynamics of the functional fate of vertebrate ohnologs (paralogous genes that originated by WGD) in comparison with non-vertebrate chordate gene homologs. Finally, this review discusses the functional consequences of the 2R-WGD on the origin and evolution of vertebrate innovations compared with urochordates and cephalochordates, paying special attention to the origin of neural crest cells, placodes, and the big complex brain, key features that probably facilitated the transition from ancestral filter-feeding non-vertebrate chordates to voracious vertebrate predators. Currently available data, however, seem to suggest that these putative key features were present to at least some extent in stem Olfactores; hence, the impact of the 2RWGD may not be related to the immediate origin of vertebrate innovations, but to the subsequent diversification of a wide variety of complex structures that facilitated the successful radiation of vertebrates. Book ISBN: 978-3-642-31441-4 Page: 309/338 Chapter 16 6 Cristian Cañestro 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 D 9 10 Abstract The origin and evolution of the vertebrates have been linked to the study of genome duplications since Susumo Ohno ventured the 2R-hypothesis, suggesting that the successful diversification of complex vertebrates was facilitated by polyploidization in the stem vertebrate ancestor due to two rounds of whole-genome duplication (2R-WGD). This chapter first reviews evidence supporting Ohno’s 2R-hypothesis and gathers information about the timing and mechanisms underlying the 2R-WGD. Second, this chapter describes the impact of the 2R-WGD on the evolution of the vertebrate genome structure, gene number, and the evolutionary dynamics of the functional fate of vertebrate ohnologs (paralogous genes that originated by WGD) in comparison with non-vertebrate chordate gene homologs. Finally, this review discusses the functional consequences of the 2R-WGD on the origin and evolution of vertebrate innovations compared with urochordates and cephalochordates, paying special attention to the origin of neural crest cells, placodes, and the big complex brain, key features that probably facilitated the transition from ancestral filter-feeding non-vertebrate chordates to voracious vertebrate predators. Currently available data, however, seem to suggest that these putative key features were present to at least some extent in stem Olfactores; hence, the impact of the 2R-WGD may not be related to the immediate origin of vertebrate innovations, but to the subsequent diversification of a wide variety of complex structures that facilitated the successful radiation of vertebrates. TE 7 8 EC 4 CO RR 3 PR OO 5 Two Rounds of Whole-Genome Duplication: Evidence and Impact on the Evolution of Vertebrate Innovations 2 F 1 Book ID: 272454_1_En Date: 16-8-2012 UN Editor Proof Layout: T1 Standard SC Chapter No.: 16 C. Cañestro (&) Departament de Genètica, Facultat de Biologia, Universitat de Barcelona, Av. Diagonal, 643, edifici Prevosti, 2a planta, 08028 Barcelona, Spain e-mail: canestro@ub.edu P. S. Soltis and D. E. Soltis (eds.), Polyploidy and Genome Evolution, DOI: 10.1007/978-3-642-31442-1_16,  Springer-Verlag Berlin Heidelberg 2012 309 Layout: T1 Standard SC Chapter No.: 16 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 310/338 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 F 33 PR OO 32 D 31 The study of the origin and evolution of vertebrates, the subphylum to which we belong, has stood at the crossroad between genome evolution and molecular developmental biology since the late 1960s, when Susumo Ohno published his famous work on Evolution by Gene Duplication and proposed his hypothesis about the pivotal role of genome duplication in the origin of vertebrates and their diversification (Ohno et al. 1968; Ohno 1970). Vertebrates comprise all animals that have a backbone and include mammals, birds, reptiles, amphibians, fishes, and agnathans—the jawless lampreys and hagfishes. Vertebrates together with urochordates (tunicates) form the Olfactores, which together with cephalochordates (amphioxus or lancelets) constitute the chordates (Fig. 16.1). All chordates share a common basic body plan at least during the larval stage of their life cycle, consisting of a notochord running through a post-anal tail, with a dorsal hollow nerve cord, longitudinal blocks of muscle along the notochord, and ciliated pharyngeal gill slits (Brusca and Brusca 2002). Recent phylogenomic analyses have dethroned cephalochordates from the long-assumed position as sister group of the vertebrates; this position is now occupied by urochordates, which include ascidians, larvaceans and thaliaceans (Wada et al. 2006; Oda et al. 2002; Bourlat et al. 2006; Delsuc et al. 2006; Putnam et al. 2008) (Fig. 16.1). Ohno’s 2R-hypothesis was based on comparative analyses of genome sizes and isozyme complexity among chordate taxa. Ohno found that basally divergent chordate subphyla had smaller genomes and less isozyme complexity than vertebrate lineages. This observation led him to suggest that the combination of tandem gene duplication and in particular an octoploidization event involving two rounds of whole-genome duplication were key to the invertebrate–vertebrate transition, and for the subsequent successful vertebrate diversification (Ohno et al. 1968). Ohno was one of the pioneers in conceiving the evolutionary significance of whole-genome duplication (see also Chap. 1, this volume). Ohno emphasized the importance of gene duplication as probably the main source of raw genetic material for the evolution of new gene functions (reviewed in Taylor and Raes 2004). In Ohno’s classical model, one of the duplicated genes retains the original function whereas its duplicate either disappears by accumulation of detrimental mutations (called pseudogenization or nonfunctionalization) or it is preserved after gaining advantageous mutations that confer positively selected novel functions (neofunctionalization) (Ohno 1970; Nowak et al. 1997; Force et al. 1999). The duplication, degeneration, complementation hypothesis (or DDC model) suggests a third possibility for duplicate gene preservation: subfunctionalization, the complementary partitioning of ancestral structural and/or regulatory subfunctions between two duplicate genes, so that the sum of their functions provides at least that of the original pre-duplication gene (Force et al. 1999). The DDC model predicts that subfunctionalized genes will have lower pleiotropy than the original pre-duplicated gene and lower evolutionary constraints, and thereby will be more permissive to the TE 30 EC 29 16.1 Introduction CO RR 28 C. Cañestro UN Editor Proof 310 Layout: T1 Standard SC Chapter No.: 16 Editor Proof 16 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 311/338 Two Rounds of Vertebrate Whole-Genome Duplication 311 CHORDATES OLFACTORES UROCHORDATES teleost fish sharks - rays amphibians birds mammals F CEPHALOCHORDATES lamprey VERTEBRATES PR OO amphioxus ascidianslarvaceans tetrapods cartilaginous vertebrates agnathans (jawless vert.) R1 R2 R1+R2 gnathostomes (jawed vert.) R1+R2 3rd. Scenario: R1+R2 after the agnathan-gnathostome split ? 2nd. Scenario: (Panvertebrate Quadruplication PV4) R1+R2 prior to agnathan-gnathostome split 1st. Scenario: R1 prior to agnathan-gnathostome split R2 in stem jawed vert. prior to cartilagenous-bony vert. split D Vertebrate Innovations: Big Brain Neural Crest Cells Placodes bony vertebrates 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 accumulation of mutations that might confer novel functions. The acquisition of new functions is favored if the duplication affects the entire genome at once, as opposed to multiple individual gene duplications, because when entire gene networks are duplicated, gene stoichiometry is maintained, and therefore deleterious gene dosage effects can be counteracted (Birchler and Veitia 2007, 2010; Van de Peer et al. 2009; Makino and McLysaght 2010; see also Chap. 2, this volume). Genes that originated by gene duplication are called paralogs. Genes that have been duplicated via genome duplication, however, are a special type of paralogs referred as ohnologs, a term suggested by Wolfe (2000) in honor of Ohno’s contribution. This term is useful because of the special properties that ohnologs possess at their birth compared to duplications that arise by other local mechanisms such as unequal crossing-over, tandem gene duplication, or retrotransposition. The sudden creation and evolution of ohnologs by the 2R-WGD that occurred in the stem vertebrate lineage has been suggested as one of the potential key events underlying the increase of morphological complexity, facilitating the acquisition of genetic and developmental innovations of vertebrates (Shimeld and Holland 2000; Aburomia et al. 2003). Genome duplication doubles the number of genes, CO RR 72 UN 71 EC TE Fig. 16.1 Evolutionary tree of chordates representing the three possible scenarios for the timing of the 2R-WGD and the origins for vertebrate innovations. Evidence suggests that the 2R-WGD might have had a significant impact on the diversification (black star) of vertebrate innovations, including structures derived from neural crest cells and placodes, as well as the development of a big complex brain in the stem vertebrate. But whether the 2R-WGD was crucial for the evolutionary origin of these structures remains unclear, and that the hypothesis that the origin of these innovations dated back to at least stem olfactores (gray star) cannot be dismissed. (Lamprey picture courtesy of Juan Pascual-Anaya) Layout: T1 Standard SC Chapter No.: 16 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 312/338 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 F 94 95 PR OO 93 D 92 TE 90 91 many of which have the chance of evolving ‘novel’ functions that might provide new selectable advantages promoting species diversification (Lynch et al. 2001; Van de Peer et al. 2009). Some vertebrate species that have undergone recent polyploidization, such as the frog Xenopus laevis that experienced tetraploidization *40 million years ago (Hellsten et al. 2007), show a higher adaptability to a variety of different environments, such as drought, salt, cold, and disease resistance, than closely related diploid species, such as Silurana tropicalis (for further details on frog polyploidization see also Chap. 18, this volume). Interestingly, only a limited number of retained ohnologs present evidence of neofunctionalization or subfunctionalization in X. laevis, suggesting that additional selective mechanisms might act on preserving gene duplicates that could promote species diversification (Chain and Evans 2006; Semon and Wolfe 2008). In addition to the evolutionary significance of novel fates of duplicate genes on species biology, another mechanism that might contribute to species diversification after genome duplication is reciprocal ohnolog loss between different populations, a concept known as ‘divergent resolution’ that can lead to reproductive isolation (Lynch and Force 2000). Reciprocal ohnolog loss is likely to occur in the period of relaxed selection that duplicate genes experience while they are functionally redundant (Werth and Windham 1991; Lynch and Conery 2000; Lynch and Force 2000; Scannell et al. 2006; Taylor et al. 2001; Semon and Wolfe 2007). The divergent resolution of gene redundancies, such that one population loses one ohnolog copy while the second population loses the other ohnolog copy, leads to chromosomal restructuring such that gametes produced by hybrid individuals can be completely lacking in functional genes for a duplicate pair. In addition to the isolation due to reciprocal gene losses, this model can be further expanded to isolation due to independent processes of gene duplicate subfunctionalization between two populations, in which hybrids will lack one or more subfunctions (Force et al. 1999; Lynch and Force 2000). Hence, large-scale reciprocal ohnolog loss and independent subfunctionalization of ohnologs can be the cause of reproductive isolation of two populations after polyploidization, favoring genetic divergence of these newly incipient future species. This hypothesis is supported by the analyses of both fish and angiosperm lineages that have undergone polyploidization and include more species diversity (e.g. salmonids, catostomids, eudicots, grasses) than their sister groups that did not go through polyploidization and include a lower number of species (Nelson 1994; Ferris et al. 1979; Soltis et al. 2009). Recent integrated approaches of comparative genomics and gene expression analyses in teleosts, however, provide limited evidence supporting the significance of differential ohnolog loss in reproductive isolation and diversification (Kassahn et al. 2009) (see also Chap. 17, this volume). Many studies have tackled the central question of whether or not the 2R-WGD had a significant impact on the origin of vertebrate innovations and their subsequent diversification. This chapter first reviews evidence supporting the 2R-hypothesis and information regarding the timing and potential mechanisms underlying the 2R-WGD in vertebrates. Second, this chapter examines the impact that the 2R-WGD may have had on the evolution of vertebrate genome structure, number of EC 89 CO RR 88 C. Cañestro UN Editor Proof 312 Layout: T1 Standard SC Chapter No.: 16 Book ISBN: 978-3-642-31441-4 Page: 313/338 Two Rounds of Vertebrate Whole-Genome Duplication 313 142 16.2 Supporting Evidence for the 2R-Hypothesis 139 140 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 159 160 161 162 163 164 165 166 167 168 169 170 171 172 173 PR OO 138 If two rounds of genome duplication (2R-WGD) have occurred, we would expect the presence of many paralogs (ohnologs) in conserved, syntenic genomic regions, which are known as paralogons (Coulier et al. 2000) (or ohnologons (Gout et al. 2009)). Paralogons, therefore, consist of series of linked (but frequently functionally and phylogenetically unrelated) genes on one chromosome region, many of which have linked paralogs on at least one other chromosome region. The discovery of paralogy groups made of four paralogons in the genome of human and mouse was interpreted as remnants of the two events of tetraploidization that occurred early during vertebrate evolution and therefore provided the earliest strong evidence supporting Ohno’s 2R-hypothesis (Lundin 1979, 1993; Pebusque et al. 1998). Possibly one of the best and first examples of a paralogy group supporting the 2R-hypothesis is the case of the four Hox-bearing regions on human chromosomes Hsa2, Hsa7, Hsa12, and Hsa17 (Fig. 16.2) (Kappen et al. 1989; Bailey et al. 1997; Larhammar et al. 2002; Lundin et al. 2003). In contrast, only a single Hox cluster is present in the cephalochordate amphioxus (Garcia-Fernàndez and Holland 1994). This 4:1 ratio is consistent with Ohno’s hypothesis of two tetraploidization events after the cephalochordate–vertebrate split (Holland et al. 1994; Sidow 1996; Garcia-Fernandez 2005). In addition to the Hox paralogy group, several other similar examples have been identified (e.g. MHC (Katsanis et al. 1996), Tbx (Ruvinsky and Silver 1997), G-protein-coupled receptors (Fredriksson et al. 2003), ParaHox clusters (Ferrier et al. 2005), linked receptor tyrosine kinases (Siegel et al. 2007), endothelin ligands and receptors (Braasch et al. 2009), Fox cluster (Wotton and Shimeld 2006), and the EGF ligand paralogons (Laisney et al. 2010)). Several databases of conserved syntenic chromosomal regions between different species are available, and these provide additional evidence of two rounds of genome duplication. Popovici et al. (2001), for instance, identified 14 paralogons containing more than 1600 genes assembled in a human genome paralogy map (http://u119.marseille.inserm.fr/Db/paralogy.html). The ParaDB (http://abi.mars eille.inserm.fr/paradb) predicted that the human genome far exceeds 1000 paralogons that contain more than three pairs of duplicated genes (Leveugle et al. 2003). D 137 TE 136 EC 135 CO RR 134 F 141 genes, and the fate of retained ohnologs in comparison with non-vertebrate chordate paralogs. Finally, this chapter discusses how the 2R-WGD might have affected the origin and evolution of vertebrate innovations, with special emphasis on the vertebrate big complex brain, and structures derived from neural crest cells and placodes. Recent data, however, suggest that neural crest cells and placodes could already have been present in stem chordates. Hence, the impact of the 2R-WGD cannot be related to the origin of neural crest cells and placodes, but it could be related to their subsequent diversification and development of a wide variety of complex structures. 133 UN Editor Proof 16 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 16 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 314/338 Editor Proof 314 C. Cañestro (a) Genomic Synteny Conservation A F B C (b) Hsa2 R2 HOX-D cluster THSD7B UPP2 ITGB6FIGN COBLL1 GRB14 HOXD1 OSBPL6 HOXD9 PLEKHA3 HNRNPA3 CDCA7HOXD10 DFNB59 FKBP7 PDE1A SP9SP3 CHN1 FRZB NEUROD1 ATF2 EVX2 SCRN3 STK17B IGFBP2 VIL1 ERBB4 ABCA12 IGFBP5 TNS1 BZW1 HECW2 FAM126BICA1L HOXB1 HOXB3 HOXB5 HOX-B cluster HOXB6 ADAP2 CCT6B MYO1D ERBB2RAPGEFL1 CDK12GRB7 IGFBP4 NEUROD2 FKBP10 NT5C3L HOXB7 ARL4D NFE2L1 HOXB8 TMEM106A CRHR1ITGB3SP2 HOXB9 FAM117A OSBPL7 ETV4SOSTHDAC5 SNX11 SCRN2 MPP2 CALCOCO2 MMD PRKAR1A AC113554.1 AMZ2 TTYH2 NACA2 GNA13WIPI1 SDK2 FAM20A ITGB4 DNAH17 USP36 CYTH1 FOXK2 FSCN2 RAC3 ACTG1 RFNG MAFG Hsa7 TE D Hsa17 D PR OO 2R-WGD: R1 TTYH3 FOXK1 SP4 DFNA5 AMZ1 CREB5PLEKHA8 FSCN1 TMEM106BTSPAN13 DNAH11 NFE2L3 CDK13 STK17APPIAIGFBP1ABCA13 EGFR ARL4ABZW2 LFNG TAX1BP1 RAC1GLCCI1 FAM20C MAFK SDK1 WIPI2USP42 SCIN ITGB8 MPP6SNX10 EVX1 CHN2 AQP1 POU6F2 HECW1 ADCY1 UPP1 CCT6A PRKAR1B GNA12 ACTB ETV1 SCRN1 CRHR2 PDE1C DPY19L1 INHBA FAM126A HOXA1 NACADTNS3FIGNL1 DGKB NT5C3 SFRP4 SOSTDC1 RAPGEF5 HNRNPA2B1 HOXA3 FKBP14NEUROD6 COBL AMPH GLI3 MYO1G MMD2 CYTH3ICA1 ADAP1 IGFBP3 GRB10 HOXA4 SP8 OSBPL3 THSD7A HDAC9 CDCA7L HOXA5 HOXA6 HOXA7 HOXA9 HOXA10 HOXA11 HOXC11 HOXC9 HOXC8 GLI1 HOXC6 AQP5 ERBB3 INHBE TSPAN31 AQP2 SP1 HOXC5 PDE1B DGKA INHBC NFE2 MIP AVIL HDAC7ADCY6 BIN2 CALCOCO1 NACA DPY19L2 NEUROD4 TENC1 POU6F1 ITGB7 HNRNPA1 RAPGEF3 SP7 ATF7 IGFBP6 Hsa12 EC HOX-A cluster CO RR HOX-C cluster Fig. 16.2 Conserved synteny in the vertebrate genome generated by the 2R-WGD. a Simplified representation of a genomic region that has been amplified by the 2R-WGD (R1 and R2) showing conserved genes (colored boxes) in four syntenic regions (a–d), which have suffered genomic rearrangements and gene loss (white boxes) and different degrees of conservation (green and red lines label ohnologs preserved in two or more than two regions, respectively). b Representation of the four human paralogons containing Hox A-D clusters in chromosomes Hsa2, Hsa7, Hsa12, and Hsa17, displaying high amounts of conserved synteny between ohnologs in two (green lines) or at least three (red lines) different paralogons. This representation has been generated using a 100gene sliding window in the Synteny Database (Catchen et al. 2009) 175 176 177 178 The Genomicus database v60.01 (http://www.dyogen.ens.fr/genomicus) predicted that 18,228 ancestral vertebrate genes were grouped in 2,642 conserved ancestral synteny blocks with a median N50 size of 5 genes (Muffato et al. 2010). The Synteny Database (http://teleost.cs.uoregon.edu/synteny_db) predicted 231 paralogy clusters with more than 5 genes, and 102 paralogy clusters with more than 10 genes, in a UN 174 Layout: T1 Standard SC Chapter No.: 16 185 186 187 188 189 190 191 192 193 194 195 196 197 198 199 200 201 202 203 204 205 206 207 208 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 F 184 PR OO 183 D 182 TE 181 315 more rigorous count using a 100-gene sliding window and taking amphioxus genes as the outgroup for paralogy assignment in human (Catchen et al. 2009, 2011). Until recently, however, available evidence did not permit us to discard the possibility that these groups of paralogous genes originated by multiple independent block duplications, rather than two duplications of the entire genome (Skrabanek and Wolfe 1998; Wolfe 2001; Larhammar et al. 2002). An initial hypothesis, based on extensive phylogenetic analysis and dating of the duplications that produced hundreds of vertebrate gene families, proposed a ‘big-bang mode’ of sudden large-scale gene origin resulting from two waves of gene duplications, rather than the alternative hypothesis of a constant generation by small-scale duplications (Gu et al. 2002). Wave-I was suggested to consist of tandem or segmental duplications that occurred after the mammalian radiation, and wave-II was interpreted as a rapid increase of paralogs in the early stage of vertebrate evolution after their split from non-vertebrate chordates, consistent with one round of whole-genome duplication (Gu et al. 2002). The first analyses of the human genome draft led to the conclusion that the most parsimonious explanation of the current structure of the human paralogons was a ‘big-bang’ expansion event by a paleopolyploidy that included the whole genome or substantial sections of it. However, no specific evidence was found for two rounds of polyploidy as opposed to one (Venter et al. 2001; McLysaght et al. 2002; Panopoulou et al. 2003). Recently, however, several analyses have provided definitive support for the 2R-hypothesis (reviewed in Kasahara 2007). Dehal and Boore (2005) developed an elegant, compelling approach to test the 2R-hypothesis by plotting the genomic map position of only those genes that were duplicated prior to the fish–tetrapod split, which rendered a clear global physical pattern of four-way paralogon organization covering most of the human genome. Dehal and Boore’s work therefore provided unmistakable evidence of two distinct rounds of genome duplication during early vertebrate evolution. Furthermore, the recent sequencing of the whole genome of the chordate amphioxus Branchiostoma floridae (sister to all other chordates) provided even more indisputable evidence supporting the 2R-hypothesis (Putnam et al. 2008). Despite the fact that small-region comparison between human, chicken, teleost fish, and amphioxus genomes revealed low gene-order conservation at the local level (microsynteny), striking extensive gene linkage conservation was observed when entire chromosomes were considered (macrosynteny). Syntenic analysis reconstructed 17 chordate linkage groups (CLG) that might represent the protochromosomes of the last common chordate ancestor (Putnam et al. 2008). Exhaustive evaluation of the 17 CLGs revealed that most of the human genome (112 segments spanning 2.68 Gb, which is the equivalent of 95 % of the euchromatic genome) was affected by large-scale duplication events that occurred on the stem vertebrate lineage before the teleost/tetrapod split. Analysis of the distribution of the human segments among the 17 CLGs showed that nearly all ancient chordate chromosomes were quadruplicated (Putnam et al. 2008) (Fig. 16.3). This result robustly demonstrated the occurrence of two rounds of genome duplication, corroborating previous lines of evidence based on analysis of EC 180 CO RR 179 Book ISBN: 978-3-642-31441-4 Page: 315/338 Two Rounds of Vertebrate Whole-Genome Duplication UN Editor Proof 16 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 16 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 316/338 C. Cañestro 231 16.2.1 Timing of the Vertebrate 2R-WGD 232 233 234 235 236 237 238 239 240 241 242 243 244 245 246 247 248 249 250 251 252 253 254 255 256 257 258 259 260 261 262 263 264 PR OO 229 Analyses of the completely sequenced genome of the cephalochordate amphioxus (Putnam et al. 2008; Holland et al. 2008) and the genomes from various urochordates (Dehal et al. 2002; Small et al. 2007; Denoeud et al. 2010) validated Ohno’s hypothesized lower-bound timing for the 2R-WGD as after the split between vertebrates and non-vertebrate chordates. Regarding the upper-bound timing, extensive analysis of gene duplicates (Robinson-Rechavi et al. 2004) and the identification of the four clusters in the genome of the elephant shark suggested that the 2R-WGD took place before the cartilaginous/bony vertebrate split (Venkatesh et al. 2007; Ravi et al. 2009). Within this time window, the most prevalent hypothesis suggests a scenario in which the first round (R1) occurred before the split between gnathostome and jawless vertebrates, and the second (R2) occurred in the stem jawed vertebrates after their divergence from jawless vertebrates (Fig. 16.1). However, a second scenario proposes that both rounds (R1 ? R2) of genome duplication took place before the split between gnathostome and jawless vertebrates (pan-vertebrate quadruplication (PV4) hypothesis (Kuraku et al. 2009)) (Fig. 16.1). Comparative analysis of 55 gene families revealed a common expansion in both jawless and jawed vertebrates, which has been interpreted as evidence supporting this second scenario (Kuraku et al. 2009). Available information from sea lampreys and hagfish does not permit us to discern between these two hypothetical scenarios, because these organisms also appear to have suffered lineage-specific duplications and reciprocal gene losses compared to vertebrates, which together obscure the assessments of orthology/paralogy (reviewed in Kuraku 2008, 2010). For instance, multiple Hox gene surveys in different species of sea lampreys and hagfish suggested that extensive independent duplications of Hox genes might have occurred during the evolution of jawless vertebrates (Pendleton et al. 1993; Sharman and Holland 1998; Takio et al. 2004; Force et al. 2002; Irvine et al. 2002; Fried et al. 2003; Stadler et al. 2004; Kuraku et al. 2009). Some of the jawless Hox clusters might have disintegrated, casting doubt as to the usefulness of Hox genes as reliable markers to trace duplications during genome evolution in stem vertebrates (Kuraku 2011). Finally, recent phylogenetic analysis of the degenerated ParaHox cluster in hagfish has opened the possibility of a third scenario, in which both rounds (R1 ? R2) occurred in stem jawed vertebrates after their divergence D 228 TE 227 EC 226 CO RR 225 F 230 specific regions of interest, such as the Hox-bearing regions (Garcia-Fernàndez and Holland 1994; Larhammar et al. 2002) and the major histocompatibility regions (Vienne et al. 2003; Danchin and Pontarotti 2004). Spring (1997) proposed the term ‘‘tetralogs’’ to refer groups of quadruplicated vertebrate genes at four different chromosomal locations formed by the 2R-WGD corresponding to a single invertebrate gene, with all four more similar to each other than to members of the other tetralogy group. 224 UN Editor Proof 316 Layout: T1 Standard SC Chapter No.: 16 Book ISBN: 978-3-642-31441-4 Page: 317/338 Two Rounds of Vertebrate Whole-Genome Duplication 317 PR OO F Human Chromosomes X Y 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 Bfl_V2_12 0Mb 1Mb D Amphioxus contig Bfl_V2_12 CO RR EC Human Chromosomes TE X Y 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 Bfl_V2_21 0Mb 2Mb 4Mb 6Mb 8Mb 10Mb Amphioxus contig Bfl_V2_21 Fig. 16.3 Quadruplicated conserved syntenic pattern between the amphioxus and the human genome as a result of the 2R-WGD. Dot-plots display the distribution throughout human chromosomes (y-axes) of human orthologs (blue dots) of amphioxus genes (black dots) located in two arbitrarily selected genomic regions of approximately 1 Mb (a) and 10 Mb (b) (x-axes). The dot-plots reveal four major human chromosomes (yellow shadow) of conserved synteny as the product of the two rounds of whole-genome duplication. In panel (a), the four paralogons coincide with the Hox-cluster bearing chromosomes 2, 7, 12, and 17, whereas in panel B the four paralogons coincide with the endothelin receptors and ParaHox-cluster bearing chromosomes 2/5, 4, 13, and X. The dot-plots were generated as described in Canestro et al. (2009) using the Synteny Database (Catchen et al. 2009) UN Editor Proof 16 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 16 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 318/338 C. Cañestro 275 16.2.2 Mechanisms Underlying the Vertebrate 2R-WGD 272 273 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 293 294 295 296 297 298 299 300 301 302 303 304 305 PR OO 270 271 A question that still remains is how did the stem vertebrate genome become octoploid by two rounds of tetraploidization. There are two main mechanisms of tetraploidization observed in many species of plants and animals (Van de Peer et al. 2009). The first mechanism is allotetraploidy, which occurs when two related but not identical genomes are combined by hybridization of closely related species and associated (often subsequent) genome duplication. In the case of allotetraploidy, the pairs of distinct ‘homologous’ chromosomes that are sufficiently different due to their separated origin are called homeologs. The second mechanism is autotetraploidy, which occurs when the genomes are not sufficiently diverged into homeologous sets; autotetraploidy therefore ranges from the combination of genomes of two conspecific individuals (perhaps from different populations) to the combination of identical genomes from a single individual. The genetic attributes of allo- and autotetraploids differ and may have substantial effects at individual, population, and species levels (see also Chap. 2, this volume). Both allotetraploidy and autotetraploidy could be generated by several processes such as: (i) an abnormal non-disjunction of sister chromatids at meiosis; (ii) the uncoupling of mitotic DNA replication and cell division during early development of the germ line (this process, for instance, occurs normally during the endoreplication of megakaryocytic bone-marrow precursors of blood platelets, or during the development of the oikoblastic epithelia that secrete the house in basal urochordate larvaceans); (iii) potential cell fusion during early embryo development or in germline precursors in syncytial gametogenesis (cell fusion is observed naturally, for instance, in skeletal muscle cells and placenta) (reviewed in Storchova and Pellman 2004; Shemer and Podbilewicz 2000). In the case of allotetraploids, each pair of homologous chromosomes should segregate normally during meiosis, and genetic interchange between homeologous chromosomes is rare. If two consecutive events of allotetraploidization occurred in stem vertebrates, we would predict that in the ideal situation in the absence of gene losses, a phylogenetic tree of homeologs will render a symmetrical (A,B) (C,D) topology (Furlong and Holland 2002). D 269 TE 268 EC 267 CO RR 266 F 274 from jawless vertebrates (Furlong et al. 2007) (Fig. 16.1). The validation of this third scenario could have a significant impact on our understanding of vertebrate evolution, because it would imply that the 2R-WGD would have not been important for the origin of vertebrate innovations (i.e. big brain, neural crest cells, and placodes, which clearly exist in jawless vertebrates (Kuratani and Ota 2008; Kuratani 2009)). According to this third scenario, however, the 2R-WGD would have been important for the radiation of gnathostomes into cartilaginous fish, bony fishes, and tetrapods. A solid answer about the timing of the 2R-WGD may have to wait until larger-scale comparisons of the whole-genome organization of hagfish and lampreys are available. 265 UN Editor Proof 318 Layout: T1 Standard SC Chapter No.: 16 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 F 311 PR OO 310 16.3 Consequences of the 2R-WGD on the Evolution of Vertebrate Genome Structure D 309 It has been suggested that polyplodization events, at least in plants, can trigger genomic stress associated with major genomic rearrangements, in many cases mediated by a burst of mobilization of transposable elements (Matzke and Matzke 1998; Comai 2000). Transposable elements can be substrates for unequal and illegitimate recombination and can be responsible for a variety of genome reorganizations associated with the transposition, including chromosomal insertions, deletions, inversions, translocations, and duplications. Lineage-specific genome rearrangements mediated by transposable elements might facilitate rapid evolution, reproductive isolation of different populations, and consequently species diversification (Parisod et al. 2010). Contrary to possible genome reorganization after polyploidization, as noted above for plants, in stem vertebrates, recent work based on proximate gene pair methods and measurement of syntenic clustering conservation found that the 2RWGD in vertebrates were not followed by an increase of genome rearrangement (Hufton et al. 2008). Unexpectedly, this work measured massive genome rearrangements prior to the 2R-WGD, which has been interpreted as a pre-existing ‘disposition’ toward genomic structural change (Hufton et al. 2008). Interestingly, in contrast to the archetypal condition that has been described in the organization of particular genomic regions (e.g. Hox-cluster region (GarciaFernàndez and Holland 1994)), the amphioxus genome structure is not exceptionally well conserved, evolving its own particular type of repetitive elements (e.g. ‘mirage’ minisatellites (Cañestro et al. 2002b; Ebner et al. 2010)), undergoing extensive local tandem gene duplications (see section below), and experiencing a moderate rate of synteny loss similar to that of sea urchin or sea TE 308 319 In autopolyploids, however, meiotic pairing might occur between any of the four identical chromosomes at meiosis I, facilitating genetic interchanges freely among the four alleles, and leading to ‘tetrasomic inheritance’. Eventually the alleles, and chromosomes, might diverge, starting a process of diploidization that reestablishes diploidy. Randomly one of the chromosomes will diverge first and no longer form homologous structures, while the other three will keep pairing until another further divergence. Hence, if two consecutive events of autotetraploidy occurred in quick succession (pseudo-octoploidy) in stem vertebrates, we would predict that in the absence of gene losses, gene family phylogenetic trees will likely render asymmetrical (((A,B),C),D) topologies (Furlong and Holland 2002). Because many vertebrate gene families do render asymmetrical tree topologies (Friedman and Hughes 2001; Hughes 1999; Hughes and Friedman 2003), two quick consecutive events of autotetraploidy have been considered a likely mechanism for the 2R-WGD in stem vertebrates (Furlong and Holland 2002; Lynch and Wagner 2009). EC 307 CO RR 306 Book ISBN: 978-3-642-31441-4 Page: 319/338 Two Rounds of Vertebrate Whole-Genome Duplication UN Editor Proof 16 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 16 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 320/338 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 F PR OO 353 D 351 352 TE 349 350 EC 348 anemone (Hufton et al. 2008). Therefore, the amphioxus genome structure cannot be considered a fossil genome representing the pre-duplication condition, at least in terms of genome structure (Garcia-Fernàndez et al. 2001; Hufton et al. 2008), although it is far less divergent from the vertebrate genome structure than is any known urochordate genome (Dehal et al. 2002; Denoeud et al. 2010; Louis et al. 2012). There have been several attempts to infer the karyotype and genome structure from common chordate ancestors and to reconstruct the evolutionary history leading to present chromosome structures. The first comparisons of conserved syntenic associations in different vertebrate karyotypes, using an in silico chromosome painting approach, allowed reconstructions of the ancestral vertebrate genome containing 10–13 ancestral proto-chromosomes (Kohn et al. 2006; Nakatani et al. 2007). Recently, the sequencing of the amphioxus genome has allowed researchers to reconstruct the ancestral chordate genome as consisting of 17 conserved syntenic blocks, which might represent the ancient chordate protochromosomes (Putnam et al. 2008). After the 2R-WGD, under the naive assumption of absence of loss or fusions of chromosomes, we would expect 68 (17 9 4) proto-vertebrate segments, but parsimonious reconstruction of chromosome history revealed that numerous chromosomal fusions and translocations have occurred. These reconstructions predict at least 20 fusions that led to 37–49 chromosomes in the bony vertebrate ancestor, which became 12 chromosomes in the stem teleost ancestor due to many additional fusions, and 33–45 chromosomes in the stem tetrapod ancestor due to at least 4 fusions shared between human and chicken genomes (Putnam et al. 2008; Naruse et al. 2004; Nakatani et al. 2007). An excellent example of chromosomal rearrangement after the 2R-WGD has been recently provided by a phylogenetic analysis of members of the four Hox paralogons that resulted in a (B(A(C,D))) topology. These results suggest that two chromosomal rearrangements between protochromosomes 11 and 4, and 7 and 5 occurred after the clusters duplicated but before the diversification of extant vertebrates 450 million years ago (Lynch and Wagner 2009). These chromosomal rearrangements resolve conflicting data regarding the order of linked genes and support the hypothesis that the 2R-WGD occurred by two consecutive events of autotetraploidy, and thereby the ancestral vertebrate might have been ‘‘pseudo-octoploid’’. Interestingly, the asymmetrical (B(A(C,D))) topology of the vertebrate Hox cluster (Lynch and Wagner 2009) contrasts with the symmetrical (A,B) (C,D) topology inferred from the cartilaginous elephant shark using the amphioxus Hox cluster as the outgroup (Ravi et al. 2009). Further extensive analyses including HoxA-D clusters from a broader representation of cartilaginous and bony vertebrates will be required to resolve these conflicting topologies, which could suggest that the Hox-cluster rearrangement took place after the cartilaginous/bony vertebrate split and not immediately subsequent to the 2R-WGD. CO RR 346 347 C. Cañestro UN Editor Proof 320 Layout: T1 Standard SC Chapter No.: 16 Book ISBN: 978-3-642-31441-4 Page: 321/338 Two Rounds of Vertebrate Whole-Genome Duplication 390 16.4.1 Function of Gene Duplicates After 2R-WGD 321 F 389 16.4 Consequences of the 2R-WGD on the Evolution of Vertebrate Gene Fate 388 415 16.4.2 Gene Network Rewiring by Tranposons After 2R-WGD 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 416 417 418 419 420 421 422 423 424 425 D 395 TE 394 EC 393 CO RR 392 PR OO 414 After polyploidization, a period of transilience may follow in which genes might enjoy extra ‘degrees of freedom’ to mutate without selective penalty (reviewed in Soltis and Soltis 1999; Otto 2007). Understanding the processes by which genome duplication might influence the fate of duplicated genes is crucial to evaluate how the 2R-WGD might have impacted the evolution of vertebrate innovations. Neofunctionalization and subfunctionalization are the two main processes driving the functional fate of newly generated ohnologs after the 2R-WGD and have been extensively discussed in the literature (Hughes 1994; Force et al. 1999; Lynch and Conery 2000; Durand 2003; Postlethwait et al. 2004; Hoekstra and Coyne 2007; Conant and Wolfe 2008; Semon and Wolfe 2007, 2008; Jimenez-Delgado et al. 2009). A prominent example of neofunctionalization related to the 2R-WGD occurred during the expansion of the vertebrate retinoic acid receptor (RAR) family, which acquired new functions in both their expression domains and in their structural protein activities (Escriva et al. 2006). There are also examples, however, in which neofunctionalization and subfunctionalization are related to both 2R-WGD and local tandem duplications (e.g. the expansion of the vertebrate globin superfamily, which promotes the vertebrate innovation related to oxygen transport and storage (Hoffmann et al. 2011)), or merely related to local tandem duplications, and not the 2R-WGD, such as the expansion of the vertebrate Alcohol Dehydrogenase (Adh) family, which promotes the acquisition of new enzymes for the synthesis of retinoic acid (Cañestro et al. 2000, 2002a, 2003b). Therefore, not all vertebrate innovations can be exclusively attributed to the 2R-WGD, and the global weight of the impact of the 2R-WGD on the evolution of vertebrate gene functions remains unknown. 391 In many cases, neofunctionalization and subfunctionalization can be due to alterations in cis-regulatory elements that might lead to adaptative changes in duplicated genes (Force et al. 1999). Many of the cis-regulatory elements appear to be embedded in distinct repeat families, especially in transposable elements (TE) (Thornburg et al. 2006; Polak and Domany 2006; Bourque et al. 2008). Analysis of the distribution of 10,000 TEs in the human genome, for instance, revealed that most TEs are concentrated under strong purifying selection near regulatory and developmental genes (Lowe et al. 2007). Most of the described examples of TE mobilization and rewiring of gene regulatory networks have been associated with relatively recent events of TE mobilizations. For instance, TE-mediated rewiring UN Editor Proof 16 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 16 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 322/338 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 F 432 PR OO 431 D 430 TE 428 429 for neofunctionalization after gene duplication has been recently described for the sex-determining gene dmrt1bY in medaka fish, in which a novel regulatory element driving a negative feedback on dmrt1bY has been acquired due to the insertion of an Izanagi transposon (Herpin et al. 2010), or for the origin of a novel gene regulatory network dedicated to pregnancy in placental mammals, which was due to a transposition of the MER20 TE (Lynch et al. 2011). A massive expansion of TEs appears, therefore, as a powerful mechanism that could boost a vast redeployment of cis-regulatory elements into new gene regulatory networks (Feschotte 2008), promoting large-scale events of neofunctionalization and subfunctionalization (van de Lagemaat et al. 2003; Bennetzen 2005; Bejerano et al. 2006). Polyploidization can trigger the mobilization of transposable elements (Matzke and Matzke 1998; Parisod et al. 2010), because recently duplicated genomes contain many redundant genes and substantial repetitive DNA, which serve as buffer against TE insertional mutagenesis (Matzke et al. 2000). According to this expectation, bursts of TE mobilization have been described after polyploidization in different organisms (Matzke and Matzke 1998; Comai 2000; SanMiguel et al. 1996, 1998). A question that remains unclear is whether there was or was not a massive TE mobilization after the 2R-WGD that could have favored a significant redeployment of cis-regulatory elements into new gene regulatory networks in the stem vertebrate lineage. Recent comparison of the diversity and content of TEs between vertebrates and amphioxus has provided some insights that might help to answer this question (Canestro and Albalat 2012). The dynamics of the TE content within a genome follows a competition model in which the expansion of a particular TE might cause the reduction of other types of TEs, consequently reducing the TE diversity, until a new equilibrium that preserves the functionality of the genome is reached (Abrusán and Krambeck 2006). According to this model, if a massive expansion of TEs occurred after the 2R-WGD, we expect that the diversity of TEs shared among vertebrates should be smaller than in cephalochordates. Consistent with this prediction, a recent comparative study reveals that the shared TE diversity of vertebrates (14 superfamilies in lampreys, 28 in ray-finned fishes, 20 in amphibians, 14 in reptiles, 10 in birds, and 15 in mammals) is lower than the TE diversity in amphioxus (33 superfamilies), which makes plausible the hypothesis that a TE burst could have occurred after the 2R-WGD in the stem vertebrate lineage (Canestro and Albalat 2012). Further comparative genomic analysis between different vertebrates and cephalochordates will be required to test this hypothetical burst of TEs, and especially to evaluate its putative impact on the evolution of gene functions after the 2R-WGD. EC 427 CO RR 426 C. Cañestro UN Editor Proof 322 Layout: T1 Standard SC Chapter No.: 16 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 F 469 PR OO 468 While several studies focus on the functional fate of retained gene duplicates, less attention has been paid to how losses of paralogs or ohnologs might impact the evolution of the functions of other genes (reviewed in Cañestro et al. 2007). Loss of one copy of two fully redundant gene duplicates should not usually have significant impact, but loss of one of the paralogs after functional divergence likely has evolutionary consequences. Recent analyses of gene losses by comparative genomics have led to the unexpected finding that significant components of the developmental toolkit might be lost without major changes to the body plan (Cañestro and Postlethwait 2007; Holland 2007), which suggests the presence of compensatory mechanisms or the acquisition of innovations that have preserved unaltered the ancestral condition (Cañestro et al. 2007). Tracing the evolution of gene families throughout ancestral proto-chromosomes using blocs of conserved synteny has become a powerful tool to clarify uncertain phylogenies, to detect ohnologs gone missing (OGM) (Postlethwait 2007; Catchen et al. 2009, 2011), to provide robust assessments of orthology and paralogy between different species, and to discern evolutionary innovations from losses of ancestral features in sister lineages (Canestro et al. 2009). There are cases in which different ohnologs in different species acquire the same expression pattern, which has been called function shuffling (McClintock et al. 2001) and synfunctionalization (Gitelman 2007), and in some cases the convergence of expression patterns between paralogs can be related to OGM (Postlethwait 2007; Canestro et al. 2009). The evolution of the vertebrate retinaldehyde dehydrogenease Aldh1a family provides a paradigmatic example of how uncovering the evolution of gene family members through the 2R-WGD has been fundamental to illuminating how gene functions evolve among newly generated paralogs after genome duplications in the face of loss of ohnologs (Canestro et al. 2009). For instance, analysis of conserved synteny revealed that the presence of Aldh1a1 in tetrapods and its absence in teleost fish was not due to a tetrapod innovation, but to an OGM in the teleost stem lineage, which was accompanied by a re-acquisition of ancestral functions by surviving paralogs (Canestro et al. 2009). Medaka provides a more radical example in which aldh1a2, the only survivor of the aldh1a family in this species, recapitulates the expression pattern of all other aldh1a paralogs that have been lost in medaka. This result is in agreement with a model of functional evolution in which surviving genes re-acquire ancestral gene family roles in the face of loss of ohnologs. Other examples that illustrate the importance of identifying OGMs ohnologs are shown in the endothelin and agouti systems, in which the exclusive presence of endothelin 4 (edn4) and the agouti-signaling protein 2 genes (asip2a/ b) in teleost fish was not due to a fish innovation related to the teleost-specific wholegenome duplication, but instead to a loss of ohnologs that originated in the 2R-WGD in the tetrapod lineage (Braasch et al. 2009; Braasch and Postlethwait 2011). To understand acquisition of functions of vertebrate ohnologs that were generated by D 467 TE 466 EC 465 323 16.4.3 Ohnologs Gone Missing After 2R-WGD and Impact on Surviving Ohnologs CO RR 464 Book ISBN: 978-3-642-31441-4 Page: 323/338 Two Rounds of Vertebrate Whole-Genome Duplication UN Editor Proof 16 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 16 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 324/338 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 F PR OO 511 16.5 Consequences of the 2R-WGD on Vertebrate Gene Number and Functional Evolution How many of the genes that were part of the original fourfold increase in genes generated by 2R-WGD in the stem vertebrate have actually survived nonfunctionalization? And importantly, how significant have the functional consequences of those retained genes been for promoting the origin and evolution of vertebrate innovations? Estimates on gene retention in other organisms that have experienced a WGD have reported *13 % retention over *100 million years (MY) in yeast (Wolfe and Shields 1997), *72 % in maize over *11 MY (Ahn and Tanksley 1993; Gaut and Doebley 1997), and *77 % in Xenopus over *40 MY (Hellsten et al. 2007). In vertebrates, a *33 % retention of divergent functional genes after the 2R-WGD over *500 MY was inferred initially based on theoretical models applied to 270 gene families of the human genome (Nadeau and Sankoff 1997). More recent and broader analyses based in the complete catalog of human ohnologs estimated a rate of retention between 20 and 30 % (Putnam et al. 2008; Huminiecki and Heldin 2010; Makino and McLysaght 2010). But how can we assess the impact of the 2R-WGD on the origin of vertebrate complex features? A naive approach to estimating this impact could be to perform a comparison of the total number of retained paralogs and their distribution among functional categories in vertebrates and non-vertebrate chordates that did not undergo any WGD since their split from our last common chordate ancestor. Comparison of the gene catalog of the three chordate subphyla (i.e. cephalochordates, urochordates, and vertebrates) has allowed us to identify a lower bound of 8,437 gene families with members that descend from a single gene in the last common chordate ancestor (Putnam et al. 2008). Through subsequent genome or local duplication, these families account for 13,610 amphioxus genes, 13,401 human genes, and 7,216 ascidian genes, the latter being a significantly lower number due to the extensive gene losses that have occurred in urochordate lineages (Dehal et al. 2002; Cañestro et al. 2003a; Edvardsen et al. 2005; Denoeud et al. 2010). Although it is frequently true that the multiple ohnologs of a vertebrate gene family are represented by a single gene in amphioxus, the total number of paralogs derived from a single-copy gene in the last common ancestor is surprisingly similar between amphioxus (13,610) and human (13,401) (Putnam et al. 2008). Therefore the mere total numbers of retained genes after the 2R-WGD duplicates might not be the key to explain the gain of complexity during the evolution of the vertebrate lineage in comparison with amphioxus. In vertebrates, analysis of the functional categories of the gene families that have expanded after the 2R-WGD revealed that cell signalers and transcriptional D 510 TE 509 the 2R-WGD, both the impact of the retention of neo- or subfunctionalized ohnologs, as well as the impact of OGM, on the functions of other survivor gene family members should be studied. EC 508 CO RR 507 C. Cañestro UN Editor Proof 324 Layout: T1 Standard SC Chapter No.: 16 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 F 552 553 PR OO 551 D 550 TE 549 325 regulators of developmental pathways are generally retained as multiple ohnologs (Roux and Robinson-Rechavi 2008; Putnam et al. 2008; Hufton et al. 2008; Huminiecki and Heldin 2010). Genes associated with basic cellular functions (i.e. translation, replication, splicing, and recombination, with the important exception of cell cycle), however, have been less successfully retained after the 2R-WGD (Huminiecki and Heldin 2010) (although see Gout et al. (2009) for different results in other organisms that have also undergone WGD). Analysis of the human genome reveals that dosage-balance constraints act on the retention of ohnologs, resulting in an enrichment of dosage-balanced genes, an observation predicted following WGD (Birchler and Veitia 2007, 2010) and also reported for other vertebrates, plants, and yeast (e.g., Paterson et al. 2006). Interestingly, many of these retained ohnologs in humans are refractory to copy number variation, have rarely experienced subsequent small-scale duplication, and are frequently associated with diseases related to dosage-imbalance such as down syndrome (Makino and McLysaght 2010). Analysis of retained genes that have originated in vertebrates by local duplications revealed a strong underrepresentation of genes related to cell communication, cell cycle, and embryo development (Huminiecki and Heldin 2010). In amphioxus, although a thorough analysis of amphioxus-specific gene family expansions has not been performed, Table 16.1 shows an extensive list of amphioxus-specific duplicated genes reported in the literature (this list is probably not complete, and may be biased toward the research with which I am most familiar). This list shows numerous retained duplicates from a broad array of functional categories, including metabolic enzymes, members of transduction and signaling cascades, members of the immunity system, as well as pivotal transcription factors of developmental pathways. Awaiting a more exhaustive analysis, including different amphioxus species to infer the ancestral cephalochordate condition, the list in Table 16.1 shows no obvious bias toward any particular functional category, although it is noticeable that duplicated developmental transcription factors do not account for more than two paralogs (with the exception of the eight hairy amphioxus paralogs (Minguillon et al. 2003)). Remarkably, the main difference between the newly acquired paralogs in amphioxus and ohnologs in vertebrates is the mechanism of duplication. While approximately 25 % of the ancestral chordate gene families have two or more ancient vertebrate ohnologs generated by the 2R-WGD, there is strong evidence that most amphioxus paralogs originated by local tandem duplications rather than largescale chromosomal duplications (Table 16.1). Therefore, considering the functional bias of retention of genes duplicated by WGD or local duplication, it is reasonable to speculate that the key influence of the 2R-WGD promoting the successful diversification of vertebrate features resides in the fact that whole networks were duplicated, in contrast to local duplications such as those that occurred in amphioxus, an organism that seems to have maintained morphological and genetic stasis during the last 200 million years (Garcia-Fernàndez and Holland 1994; Cañestro et al. 2002a; Somorjai et al. 2008; Canestro and Albalat 2012; Paps et al. 2012). Duplication of whole gene networks is dosage-balanced and increases the EC 548 CO RR 547 Book ISBN: 978-3-642-31441-4 Page: 325/338 Two Rounds of Vertebrate Whole-Genome Duplication UN Editor Proof 16 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 16 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 326/338 C. Cañestro Table 16.1 List of paralogs originated independently in the amphioxus lineage Functional category Gene Number of References amphioxus paralogs 28 39 406 EC Developmental transcription factors F Toll-like receptors C1q-like LRR-containing gene models Hairy Brachyury bHLH Emx Hox13–14 Evx Mnx Uncx Pou3 Lhx2/9 Iro CO RR Immunity system Cañestro et al. (2006) Albalat (2009) Albalat et al. (2011) PR OO 5 3 22 12 14 22 6 5 10 8 7 8 3 20 18 3 5 12 TE Transduction and signaling cascades Aldh1a Cyp26 Hsd11b1 Rdh cluster Bdh1 Rdh11/12 Hsd17b8 ApoD Cyp2 Cyp11/24/27 Hsd3b Fabp Nos Opsins GPR54 GnRH CRHR Somatostatine receptors LGR7/8 PTHR Secretin receptors Estrogen receptors NOK ACK TIE MARTK EXTK HNK-Ras Calmoduline-like Andreakis et al. (2011) Holland et al. (2008) D Metabolic proteins UN Editor Proof 326 6 4 5 2 22 3 7 8 47 2 3 8 2 2 2 2 2 2 2 2 2 2 Bridgham et al. (2008) D’Aniello et al. (2008) Bertrand et al. (2009) Karabinos and Bhattacharya (2000) Holland et al. (2008) Minguillon et al. (2003) Holland et al. (1995) Araki et al. (1996) Minguillon et al. (2002) Ferrier et al. (2000) Ferrier et al. (2001) Holland et al. (2008) Layout: T1 Standard SC Chapter No.: 16 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 F 597 PR OO 596 D 595 TE 594 327 evolvability to generate novel functions, which in the case of the vertebrate 2RWGD could have led to an increase in complexity of the signaling and developmental regulatory networks that facilitated the acquisition of innovations. In addition to the evolutionary role of coding genes in the acquisition of innovations, microRNAs (miRNAs) also play crucial roles during development and have been postulated as important players for the evolution of organismal complexity (Lee et al. 2007; Sempere et al. 2006). Analysis of miRNAs in chordate species showed that the 2R-WGD has increased the diversity of the inventory of miRNAs in vertebrates, which correlated with the increase of complex patterns of tissue specificity of miRNAs (Heimberg et al. 2008; Campo-Paysaa et al. 2011). However, the finding of 41 vertebrate-specific miRNA families, absent in non-vertebrate chordates, suggests that their origin must have occurred in stem vertebrates after their separation from urochordates and is not explained by the 2R-WGD (Heimberg et al. 2008). The appearance of these 41 vertebrate-specific miRNA families has been proposed as a potential key evolutionary force lying behind the dramatic increase of vertebrate complexity (Heimberg et al. 2008). Future exhaustive analysis of the expression patterns of the members of these 41 vertebrate-specific families, and an understanding of their roles, will allow a reevaluation of the importance that this innovation could have had on the origin of vertebrate features. 16.6 Consequences of the 2R-WGD on the Innovation of Vertebrate Features A major question not resolved yet is the precise impact of the 2R-WGD on the innovation of particular vertebrate features. Three vertebrate features are perhaps the most prominent innovations: derivatives from neural crest cells, sensory organs concentrated in the head derived from ectodermal placodes, and a big complex brain. When taken together, these features probably allowed the transition from ancestral, peaceful, filter-feeding, non-vertebrate chordates to active, voracious, vertebrate predators (Northcutt and Gans 1983; Gans and Northcutt 1983); reviewed in Yu et al. 2008; Holland 2009) (Fig. 16.1). Vertebrate neural crest cells are a transient population of developmental cells that delaminate at the border of the neural plate through an epithelial–mesenchymal transition, migrate, and differentiate at their final destination into a variety of structures such as sensory neurons, glial cells, peripheral nervous system, pigment cells, smooth muscle cells, connective tissue, cranio-facial cartilage, skeletal bones, and teeth (Weston 1970). Vertebrate crest development depends on four crucial sets of genes that form what is called the neural crest gene regulatory network (NC-GRN): (1) patterning signal genes establish the expression of (2) neural plate border specifier genes, which activate (3) crest specifier genes, which turn on (4) neural crest effector genes that provide differentiated products (Meulemans and Bronner-Fraser 2004, 2005; Ota and Kuratani 2007; Sauka-Spengler et al. 2007). Analysis of the amphioxus genome has revealed the presence of cephalochordate EC 593 CO RR 592 Book ISBN: 978-3-642-31441-4 Page: 327/338 Two Rounds of Vertebrate Whole-Genome Duplication UN Editor Proof 16 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 16 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 328/338 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 F 638 PR OO 637 D 636 TE 635 orthologs from all of these four sets of genes, including (1): Fgf, Wnt, Bmp, Notch Dlx, AP2, SoxB, Zic, and islet; (2): Pax3/7, Msx, Dlx5, and Zic; (3): Snail, SoxE, AP2, Twist, Id, FoxD, and Myc; and (4): Rho, cRet, Erbb3, Mitf, tyrosinase, and tyrosinase-related genes, with the remarkable exception of the tyrosine kinase c-Kit essential for migration and survival of crest cells, and the gene for myelin protein P0, consistent with the notion that the glial myelin sheath is a vertebrate innovation (Meulemans and Bronner-Fraser 2007; Holland et al. 2008; Holland and Short 2008; Nikitina et al. 2009). The fact that most of the specifier genes are present as single copy in amphioxus, but multiple paralogs in vertebrates, presumably due to the 2R-WGD, has led to the hypothesis that neofunctionalization and subfunctionalization of paralogs may have facilitated the co-option of ancestral genes into the NC-GRN (Sauka-Spengler et al. 2007; Meulemans and Bronner-Fraser 2007; Holland et al. 2008; Holland and Short 2008). Gene ontology (GO) analysis estimates that 91 % of the neural crest genes in vertebrates have been co-opted from genes already present in basal metazoans, while the remaining 9 % of the neural crest genes are vertebrate innovations (Martinez-Morales et al. 2007), including the assembly of new signaling pathways like the endothelin system (Braasch et al. 2009). The evolution of the vertebrate NC-GRN, therefore, appears as the result of a combination of ancestral gene cooption, newly evolving genes, and amplification of these components by the 2RWGDs (Braasch et al. 2009). Interestingly, however, the recent description in urochordates of neural crest-like cells that express typical vertebrate crest marker orthologs, migrate, and differentiate into pigments challenges the idea that neural crest cells are a vertebrate innovation (Jeffery et al. 2004, 2008; Jeffery 2007). Thus, it cannot be discounted that some types of neural crest cells might have been present in the last common ancestor of olfactores (urochordates ? vertebrates), followed by losses during the significant morphological and genetic simplification suffered by urochordate lineages (Seo et al. 2004; Cañestro et al. 2005; Cañestro and Postlethwait 2007; Holland 2007). Therefore, it seems plausible to consider that the 2R-WGD might have not been crucial for the origin of the neural crest cells, but the 2R-WGD might have been important for increasing the evolvability of the NC-GRN and the diversification of derivative structures. Similar conclusions have been reached through studies of the gene regulatory network underlying placode and brain development. Analysis of placode-marking genes (e.g. Eye, Pitx, Six, and Pax) in ascidian and larvacean urochordates suggested that the last common olfactore ancestor already presented multiple placode derivatives, such as olfactory and adenohypophyseal. Additional and independent proliferation and loss of a variety of placodes probably occurred in both urochordate and vertebrate lineages (Bassham and Postlethwait 2005; Mazet et al. 2005), in some cases recruiting paralogs that had been independently duplicated in both urochordates and vertebrates (Bassham et al. 2008). Despite the fact that non-vertebrate chordates have a simple brain lacking a midbrain and a midbrain–hindbrain organizer (MHB), most brain-making gene orthologs are present in non-vertebrate chordates, suggesting that vertebrate brain EC 634 CO RR 633 C. Cañestro UN Editor Proof 328 Layout: T1 Standard SC Chapter No.: 16 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 F 683 PR OO 682 D 681 TE 680 329 features were built on a foundation already present in the ancestral chordate probably facilitated by the new ohnologs created by the vertebrate 2R-WGD (reviewed in Holland 2009). Recent analysis of developmental genes in the ascidian brain revealed that the expression of Fgf8 can reorganize the expression of other brain genes and transform hindbrain structures into an expanded mesencephelon, recapitulating the organizing activity of the vertebrate MHB and therefore suggesting that the MHB was already present at least in the last common ancestor of olfactores (Imai et al. 2009). Analysis of urochordate genomes revealed that important genes (i.e. Gbx) for the positioning of the MHB have been lost in stem urochordates (Cañestro et al. 2005), as has the retinoic acid dependent anterior–posterior axial patterning of the central nervous system (Cañestro et al. 2006), making plausible the hypothesis that the absence of midbrain in urochordates is not due to a vertebrate innovation of a midbrain, but a simplification in urochordates of an ancestral tripartite brain structure (Cañestro and Postlethwait 2007; Cañestro et al. 2007). Evolutionary analysis of the origin of the complex Nova-regulated splice variants of the vertebrate brain genes revealed that many of these variants were already present in the last common olfactore ancestor (Irimia et al. 2011). It is possible that the 2R-WGD promoted the increase of the complexity of Nova-dependent splice variants in the vertebrate brain, although a simplification of this system during urochordate evolution cannot be discarded. In conclusion, it is likely that the origin of vertebrate features such as neural crest cells, placodes, and a complex tripartite brain are not related to the 2R-WGD, but that these features were already present to some extent in stem non-vertebrate chordates (Fig. 16.1) (reviewed in Donoghue et al. 2008). However, it is likely that the subsequent evolution of these three features has been strongly influenced by the new ohnologs that originated after the 2R-WGD, due to processes of neofunctionalization, subfunction partitioning and subsequent refinement, recruitment of cis-regulatory elements driven by genome rearrangement and transposable element activity, inventions of novel miRNA families, and evolution of novel splice variants, which overall increased the complexity of duplicated developmental gene regulatory networks after the 2R-WGD. Future integrative analysis of comparative genomics, functional evo-devo, and examinations of gene regulatory networks in a wide variety of non-vertebrate chordates as well as basally divergent jawless vertebrates will help to narrow down the precise timing of the 2R-WGD and evaluate its actual impact on the origin and evolution of vertebrate features. Probably the new ‘2R, or not 2R’ question (Hughes and Friedman 2003) is now to ascertain whether the origins of vertebrate innovations were, or were not, the consequence of the 2R-WGD, and to understand the mechanisms by which the 2R-WGD increased the evolvability of developmental gene regulatory networks that facilitated the diversification of complex vertebrate features. EC 679 CO RR 678 Book ISBN: 978-3-642-31441-4 Page: 329/338 Two Rounds of Vertebrate Whole-Genome Duplication UN Editor Proof 16 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 16 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 330/338 C. Cañestro Acknowledgments This work has been supported by grant BFU2010-14875 from the Ministerio de Ciencia e Innovación (Spain). I would like to thank Julian Catchen for his generous support on the Synteny Database, and Ingo Braasch, John H. Postlethwait, Ricard Albalat, and Adriana Rodriguez for their valuable comments on the chapter, and cheerful and endless discussions on ‘‘2R, or not 2R, that is the question … on vertebrate innovations’’. 724 References 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767 Abrusán G, Krambeck H-J (2006) Competition may determine the diversity of transposable elements. Theor Popul Biol 70(3):364–375 Aburomia R, Khaner O, Sidow A (2003) Functional evolution in the ancestral lineage of vertebrates or when genomic complexity was wagging its morphological tail. J Struct Funct Genomics 3(1–4):45–52 Ahn S, Tanksley SD (1993) Comparative linkage maps of the rice and maize genomes. Proc Natl Acad Sci U S A 90(17):7980–7984 Albalat R (2009) The retinoic acid machinery in invertebrates: ancestral elements and vertebrate innovations. Mol Cell Endocrinol 313(1–2):23–35 Albalat R, Brunet F, Laudet V, Schubert M (2011) Evolution of retinoid and steroid signaling: vertebrate diversification from an amphioxus perspective. Genome Biol Evol 3:985–1005 Andreakis N, D’Aniello S, Albalat R, Patti FP, Garcia-Fernandez J, Procaccini G, Sordino P, Palumbo A (2011) Evolution of the nitric oxide synthase family in metazoans. Mol Biol Evol 28(1):163–179 Araki I, Terazawa K, Satoh N (1996) Duplication of an amphioxus myogenic bHLH gene is independent of vertebrate myogenic bHLH gene duplication. Gene 171(2):231–236 Bailey W, Kim J, Wagner G, Ruddle F (1997) Phylogenetic reconstruction of vertebrate Hox cluster duplications. Mol Biol Evol 14:843–853 Bassham S, Cañestro C, Postlethwait JH (2008) Evolution of developmental roles of Pax2/5/8 paralogs after independent duplication in urochordate and vertebrate lineages. BMC Biol 6:35 Bassham S, Postlethwait JH (2005) The evolutionary history of placodes: a molecular genetic investigation of the larvacean urochordate Oikopleura dioica. Development 132(19):4259– 4272 Bejerano G, Lowe CB, Ahituv N, King B, Siepel A, Salama SR, Rubin EM, Kent WJ, Haussler D (2006) A distal enhancer and an ultra conserved exon are derived from a novel retroposon. Nature 441(7089):87–90 Bennetzen JL (2005) Transposable elements, gene creation and genome rearrangement in flowering plants. Curr Opin Genet Dev 15(6):621–627 Bertrand S, Campo-Paysaa F, Camasses A, Garcia-Fernandez J, Escriva H (2009) Actors of the tyrosine kinase receptor downstream signaling pathways in amphioxus. Evol Dev 11(1):13–26 Birchler JA, Veitia RA (2007) The gene balance hypothesis: from classical genetics to modern genomics. Plant Cell 19(2):395–402 Birchler JA, Veitia RA (2010) The gene balance hypothesis: implications for gene regulation, quantitative traits and evolution. New Phytol 186(1):54–62 Bourlat SJ, Juliusdottir T, Lowe CJ, Freeman R, Aronowicz J, Kirschner M, Lander ES, Thorndyke M, Nakano H, Kohn AB, Heyland A, Moroz LL, Copley RR, Telford MJ (2006) Deuterostome phylogeny reveals monophyletic chordates and the new phylum Xenoturbellida. Nature 444(7115):85–88 Bourque G, Leong B, Vega VB, Chen X, Lee YL, Srinivasan KG, Chew JL, Ruan Y, Wei CL, Ng HH, Liu ET (2008) Evolution of the mammalian transcription factor binding repertoire via transposable elements. Genome Res 18(11):1752–1762 Braasch I, Postlethwait JH (2011) The teleost agouti-related protein 2 gene is an ohnolog gone missing from the tetrapod genome. Proc Natl Acad Sci U S A 108(13):E47–E48 CO RR EC TE D PR OO F 719 720 721 722 723 UN Editor Proof 330 Layout: T1 Standard SC Chapter No.: 16 331 EC TE D PR OO F Braasch I, Volff JN, Schartl M (2009) The endothelin system: evolution of vertebrate-specific ligand–receptor interactions by three rounds of genome duplication. Mol Biol Evol 26(4):783–799 Bridgham JT, Brown JE, Rodriguez-Mari A, Catchen JM, Thornton JW (2008) Evolution of a new function by degenerative mutation in cephalochordate steroid receptors. PLoS Genet 4(9):e1000191 Brusca RC, Brusca GJ (2002) Invertebrates. Sinauer Associates, Sunderland Campo-Paysaa F, Semon M, Cameron RA, Peterson KJ, Schubert M (2011) microRNA complements in deuterostomes: origin and evolution of microRNAs. Evol Dev 13(1):15–27 Cañestro C, Albalat R (2012) Transposon diversity is higher in amphioxus than in vertebrates: functional and evolutionary inferences. Brief Funct Genomics 11(2):131–141 Cañestro C, Albalat R, Hjelmqvist L, Godoy L, Jornvall H, Gonzalez-Duarte R (2002a) Ascidian and amphioxus Adh genes correlate functional and molecular features of the ADH family expansion during vertebrate evolution. J Mol Evol 54(1):81–89 Cañestro C, Bassham S, Postlethwait JH (2003a) Seeing chordate evolution through the Ciona genome sequence. Genome Biol 4(3):208–211 Cañestro C, Bassham S, Postlethwait JH (2005) Development of the central nervous system in the larvacean Oikopleura dioica and the evolution of the chordate brain. Dev Biol 285(2):298– 315 Cañestro C, Catchen JM, Rodriguez-Mari A, Yokoi H, Postlethwait JH (2009) Consequences of lineage-specific gene loss on functional evolution of surviving paralogs: ALDH1A and retinoic acid signaling in vertebrate genomes. PLoS Genet 5(5):e1000496 Cañestro C, Godoy L, Gonzàlez-Duarte R, Albalat R (2003b) Comparative expression analysis of Adh3 during arthropod, urochordate, cephalochordate and vertebrate development challenges its predicted housekeeping role. Evol Dev 5(2):157–162 Cañestro C, Gonzàlez-Duarte R, Albalat R (2002b) Minisatellite instability at the Adh locus reveals somatic polymorphism in amphioxus. Nucleic Acids Res 30(13):2871–2876 Cañestro C, Hjelmqvist L, Albalat R, Garcia-Fernàndez J, Gonzàlez-Duarte R, Jörnvall H (2000) Amphioxus alcohol dehydrogenase is a class 3 form of single type and of structural conservation but with unique developmental expression. Eur J Biochem 267:6511–6518 Cañestro C, Postlethwait JH (2007) Development of a chordate anterior–posterior axis without classical retinoic acid signaling. Dev Biol 305(2):522–538 Cañestro C, Postlethwait JH, Gonzàlez-Duarte R, Albalat R (2006) Is retinoic acid genetic machinery a chordate innovation? Evol Dev 8(5):394–406 Cañestro C, Yokoi H, Postlethwait JH (2007) Evolutionary developmental biology and genomics. Nat Rev Genet 8(12):932–942 Catchen JM, Braasch I, Postlethwait JH (2011) Conserved synteny and the zebrafish genome. Methods Cell Biol 104:259–285 Catchen JM, Conery JS, Postlethwait JH (2009) Automated identification of conserved synteny after whole-genome duplication. Genome Res 19(8):1497–1505 Chain FJ, Evans BJ (2006) Multiple mechanisms promote the retained expression of gene duplicates in the tetraploid frog Xenopus laevis. PLoS Genet 2(4):e56 Comai L (2000) Genetic and epigenetic interactions in allopolyploid plants. Plant Mol Biol 43 (2–3):387–399 Conant GC, Wolfe KH (2008) Turning a hobby into a job: how duplicated genes find new functions. Nat Rev Genet 9(12):938–950 Coulier F, Popovici C, Villet R, Birnbaum D (2000) MetaHox gene clusters. J Exp Zool 288(4):345–351 D’Aniello S, Irimia M, Maeso I, Pascual-Anaya J, Jimenez-Delgado S, Bertrand S, GarciaFernandez J (2008) Gene expansion and retention leads to a diverse tyrosine kinase superfamily in amphioxus. Mol Biol Evol 25(9):1841–1854 Danchin EG, Pontarotti P (2004) Towards the reconstruction of the bilaterian ancestral pre-MHC region. Trends Genet 20(12):587–591 CO RR 768 769 770 771 772 773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792 793 794 795 796 797 798 799 800 801 802 803 804 805 806 807 808 809 810 811 812 813 814 815 816 817 818 819 820 Book ISBN: 978-3-642-31441-4 Page: 331/338 Two Rounds of Vertebrate Whole-Genome Duplication UN Editor Proof 16 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 16 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 332/338 EC TE D PR OO F Dehal P, Boore JL (2005) Two rounds of whole genome duplication in the ancestral vertebrate. PLoS Biol 3(10):e314 Dehal P, Satou Y, Campbell RK, Chapman J, Degnan B, De Tomoso A, Davidson B, Di Gregorio A, Gelpke M, Goodstein DM, Harafuji N, Hastings KEM, Ho I, Hotta K, Huang W, Kawashima T, Lemaire P, Martinez D, Meinertzhagen IA, Necula S, Nonaka M, Putnam N, Rash S, Saiga H, Satake M, Terry A, Yamada L, Wang H-G, Awazu S, Azumi K, Boore J, Branno M, Chin-bow S, DeSantis R, Doyle S, Francino P, Keys DN, Haga S, Hayashi H, Hino K, Imai KS, Inaba K, Kano S, Kobayashi K, Kobayashi M, Lee B-I, Makabe KW, Manohar C, Matassi G, Medina M, Mochizuki Y, Mount S, Morishita T, Miura S, Nakayama A, Nishizaka S, Nomoto H, Ohta F, Oishi K, Rigoutsos I, Sano M, Sasaki A, Sasakura Y, Shoguchi E, Shin-i T, Spagnuolo A, Stainier D, Suzuki MM, Tassy O, Takatori N, Tokuoka M, Yagi K, Yoshizaki F, Wada S, Zhang C, Hyatt PD, Larimer F, Detter C, Doggett N, Glavina T, Hawkins T, Richardson P, Lucas S, Kohara Y, Levine M, Satoh N, Rokhsar DS (2002) The draft genome of Ciona intestinalis: insights into chordate and vertebrate origins. Science 298(5601):2157–2167 Delsuc F, Brinkmann H, Chourrout D, Philippe H (2006) Tunicates and not cephalochordates are the closest living relatives of vertebrates. Nature 439(7079):965–968 Denoeud F, Henriet S, Mungpakdee S, Aury JM, Da Silva C, Brinkmann H, Mikhaleva J, Olsen LC, Jubin C, Cañestro C, Bouquet JM, Danks G, Poulain J, Campsteijn C, Adamski M, Cross I, Yadetie F, Muffato M, Louis A, Butcher S, Tsagkogeorga G, Konrad A, Singh S, Jensen MF, Cong EH, Eikeseth-Otteraa H, Noel B, Anthouard V, Porcel BM, Kachouri-Lafond R, Nishino A, Ugolini M, Chourrout P, Nishida H, Aasland R, Huzurbazar S, Westhof E, Delsuc F, Lehrach H, Reinhardt R, Weissenbach J, Roy SW, Artiguenave F, Postlethwait JH, Manak JR, Thompson EM, Jaillon O, Du Pasquier L, Boudinot P, Liberles DA, Volff JN, Philippe H, Lenhard B, Roest Crollius H, Wincker P, Chourrout D (2010) Plasticity of animal genome architecture unmasked by rapid evolution of a pelagic tunicate. Science 330(6009):1381–1385 Donoghue PC, Graham A, Kelsh RN (2008) The origin and evolution of the neural crest. BioEssays 30(6):530–541 Durand D (2003) Vertebrate evolution: doubling and shuffling with a full deck. Trends Genet 19(1):2–5 Ebner B, Panopoulou G, Vinogradov SN, Kiger L, Marden MC, Burmester T, Hankeln T (2010) The globin gene family of the cephalochordate amphioxus: implications for chordate globin evolution. BMC Evol Biol 10:370 Edvardsen RB, Seo HC, Jensen MF, Mialon A, Mikhaleva J, Bjordal M, Cartry J, Reinhardt R, Weissenbach J, Wincker P, Chourrout D (2005) Remodelling of the homeobox gene complement in the tunicate Oikopleura dioica. Curr Biol 15(1):R12–R13 Escriva H, Bertrand S, Germain P, Robinson-Rechavi M, Umbhauer M, Cartry J, Duffraisse M, Holland L, Gronemeyer H, Laudet V (2006) Neofunctionalization in vertebrates: the example of retinoic acid receptors. PLoS Genet 2(7):e102 Ferrier DE, Dewar K, Cook A, Chang JL, Hill-Force A, Amemiya C (2005) The chordate ParaHox cluster. Curr Biol 15(20):R820–R822 Ferrier DE, Minguillon C, Holland PW, Garcia-Fernandez J (2000) The amphioxus Hox cluster: deuterostome posterior flexibility and Hox14. Evol Dev 2(5):284–293 Ferrier DEK, Minguillón C, Cebrián C, Garcia-Fernàndez J (2001) Amphioxus Evx genes: implications for the evolution of the midbrain–hindbrain boundary and the chordate tailbud. Dev Biol 237(2):270–281 Ferris SD, Portnoy SL, Whitt GS (1979) The roles of speciation and divergence time in the loss of duplicate gene expression. Theor Popul Biol 15(1):114–139 Feschotte C (2008) Transposable elements and the evolution of regulatory networks. Nat Rev Genet 9(5):397–405 Force A, Amores A, Postlethwait JH (2002) Hox cluster organization in the jawless vertebrate Petromyzon marinus. J Exp Zool 294(1):30–46 CO RR 821 822 823 824 825 826 827 828 829 830 831 832 833 834 835 836 837 838 839 840 841 842 843 844 845 846 847 848 849 850 851 852 853 854 855 856 857 858 859 860 861 862 863 864 865 866 867 868 869 870 871 872 873 C. Cañestro UN Editor Proof 332 Layout: T1 Standard SC Chapter No.: 16 333 EC TE D PR OO F Force A, Lynch M, Pickett FB, Amores A, Yan Y-L, Postlethwait J (1999) Preservation of duplicate genes by complementary, degenerative mutations. Genetics 151(4):1531–1545 Fredriksson R, Lagerstrom MC, Lundin LG, Schioth HB (2003) The G-protein-coupled receptors in the human genome form five main families. Phylogenetic analysis, paralogon groups, and fingerprints. Mol Pharmacol 63(6):1256–1272 Fried C, Prohaska SJ, Stadler PF (2003) Independent Hox-cluster duplications in lampreys. J Exp Zoolog Part B Mol Dev Evol 299(1):18–25 Friedman R, Hughes AL (2001) Pattern and timing of gene duplication in animal genomes. Genome Res 11(11):1842–1847 Furlong R, Holland P (2002) Were vertebrates octoploid? Philos T Roy Soc B 357:531–544 Furlong RF, Younger R, Kasahara M, Reinhardt R, Thorndyke M, Holland PW (2007) A degenerate ParaHox gene cluster in a degenerate vertebrate. Mol Biol Evol 24(12): 2681–2686 Gans C, Northcutt RG (1983) Neural crest and the origin of vertebrates: a new head. Science 220:268–274 Garcia-Fernandez J (2005) The genesis and evolution of homeobox gene clusters. Nat Rev Genet 6(12):881–892 Garcia-Fernàndez J, Ferrier DE, Minguillón C, Cebrián C (2001) The amphioxus genome in evodevo: archetype or ‘‘cul de sac’’. Int J Dev Biol 45(S1):S137–S138 Garcia-Fernàndez J, Holland PW (1994) Archetypal organization of the amphioxus Hox gene cluster. Nature 370(6490):563–566 Gaut BS, Doebley JF (1997) DNA sequence evidence for the segmental allotetraploid origin of maize. Proc Natl Acad Sci U S A 94(13):6809–6814 Gitelman I (2007) Evolution of the vertebrate twist family and synfunctionalization: a mechanism for differential gene loss through merging of expression domains. Mol Biol Evol 24(9): 1912–1925 Gout JF, Duret L, Kahn D (2009) Differential retention of metabolic genes following wholegenome duplication. Mol Biol Evol 26(5):1067–1072 Gu X, Wang Y, J G (2002) Age distribution of human gene families shows significant roles of both large- and small-scale duplications in vertebrate evolution. Nat Genet 31:205–209 Heimberg AM, Sempere LF, Moy VN, Donoghue PC, Peterson KJ (2008) microRNAs and the advent of vertebrate morphological complexity. Proc Natl Acad Sci U S A 105(8):2946–2950 Hellsten U, Khokha MK, Grammer TC, Harland RM, Richardson P, Rokhsar DS (2007) Accelerated gene evolution and subfunctionalization in the pseudotetraploid frog Xenopus laevis. BMC Biol 5:31 Herpin A, Braasch I, Kraeussling M, Schmidt C, Thoma EC, Nakamura S, Tanaka M, Schartl M (2010) Transcriptional rewiring of the sex determining dmrt1 gene duplicate by transposable elements. PLoS Genet 6(2):e1000844 Hoekstra HE, Coyne JA (2007) The locus of evolution: evo devo and the genetics of adaptation. Evolution 61(5):995–1016 Hoffmann FG, Opazo JC, Storz JF (2012) Whole-genome duplications spurred the functional diversification of the globin gene superfamily in vertebrates. Mol Biol Evol 29:303-312 Holland LZ (2007) Developmental biology: a chordate with a difference. Nature 447(7141):153– 155 Holland LZ (2009) Chordate roots of the vertebrate nervous system: expanding the molecular toolkit. Nat Rev Neurosci 10(10):736–746 Holland LZ, Albalat R, Azumi K, Benito-Gutierrez E, Blow MJ, Bronner-Fraser M, Brunet F, Butts T, Candiani S, Dishaw LJ, Ferrier DE, Garcia-Fernandez J, Gibson-Brown JJ, Gissi C, Godzik A, Hallbook F, Hirose D, Hosomichi K, Ikuta T, Inoko H, Kasahara M, Kasamatsu J, Kawashima T, Kimura A, Kobayashi M, Kozmik Z, Kubokawa K, Laudet V, Litman GW, McHardy AC, Meulemans D, Nonaka M, Olinski RP, Pancer Z, Pennacchio LA, Pestarino M, Rast JP, Rigoutsos I, Robinson-Rechavi M, Roch G, Saiga H, Sasakura Y, Satake M, Satou Y, Schubert M, Sherwood N, Shiina T, Takatori N, Tello J, Vopalensky P, Wada S, Xu A, Ye Y, Yoshida K, Yoshizaki F, Yu JK, Zhang Q, Zmasek CM, de Jong PJ, Osoegawa K, Putnam CO RR 874 875 876 877 878 879 880 881 882 883 884 885 886 887 888 889 890 891 892 893 894 895 896 897 898 899 900 901 902 903 904 905 906 907 908 909 910 911 912 913 914 915 916 917 918 919 920 921 922 923 924 925 926 927 Book ISBN: 978-3-642-31441-4 Page: 333/338 Two Rounds of Vertebrate Whole-Genome Duplication UN Editor Proof 16 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 16 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 334/338 EC TE D PR OO F NH, Rokhsar DS, Satoh N, Holland PW (2008) The amphioxus genome illuminates vertebrate origins and cephalochordate biology. Genome Res 18(7):1100–1111 Holland LZ, Short S (2008) Gene duplication, co-option and recruitment during the origin of the vertebrate brain from the invertebrate chordate brain. Brain Behav Evol 72(2):91–105 Holland PWH, Garcia-Fernàndez J, Williams NA, Sidow A (1994) Gene duplications and the origins of vertebrate development. Development (Suppl.):125–133 Holland PWH, Koschorz B, Holland LZ, Herrmann BG (1995) Conservation of Brachyury (T) genes in amphioxus and vertebrates: developmental and evolutionary implications. Development 121:4283–4291 Hufton AL, Groth D, Vingron M, Lehrach H, Poustka AJ, Panopoulou G (2008) Early vertebrate whole genome duplications were predated by a period of intense genome rearrangement. Genome Res 18(10):1582–1591 Hughes AL (1994) The evolution of functionally novel proteins after gene duplication. Proc Roy Soc Lond B 256:119–124 Hughes AL (1999) Phylogenies of developmentally important proteins do not support the hypothesis of two rounds of genome duplication early in vertebrate history. J Mol Evol 48(5):565–576 Hughes AL, Friedman R (2003) 2R or not 2R: testing hypotheses of genome duplication in early vertebrates. J Struct Funct Genomics 3(1–4):85–93 Huminiecki L, Heldin CH (2010) 2R and remodeling of vertebrate signal transduction engine. BMC Biol 8:146 Imai KS, Stolfi A, Levine M, Satou Y (2009) Gene regulatory networks underlying the compartmentalization of the Ciona central nervous system. Development 136(2):285–293 Irimia M, Denuc A, Burguera D, Somorjai I, Martin-Duran JM, Genikhovich G, JimenezDelgado S, Technau U, Roy SW, Marfany G, Garcia-Fernandez J (2011) Stepwise assembly of the Nova-regulated alternative splicing network in the vertebrate brain. Proc Natl Acad Sci U S A 108(13):5319–5324 Irvine SQ, Carr JL, Bailey WJ, Kawasaki K, Shimizu N, Amemiya CT, Ruddle FH (2002) Genomic analysis of Hox clusters in the sea lamprey Petromyzon marinus. J Exp Zool 294:47–62 Jeffery WR (2007) Chordate ancestry of the neural crest: new insights from ascidians. Semin Cell Dev Biol 18:481-491 Jeffery WR, Chiba T, Krajka FR, Deyts C, Satoh N, Joly JS (2008) Trunk lateral cells are neural crest-like cells in the ascidian Ciona intestinalis: insights into the ancestry and evolution of the neural crest. Dev Biol 324(1):152–160 Jeffery WR, Strickler AG, Yamamoto Y (2004) Migratory neural crest-like cells form body pigmentation in a urochordate embryo. Nature 431(7009):696–699 Jimenez-Delgado S, Pascual-Anaya J, Garcia-Fernandez J (2009) Implications of duplicated cisregulatory elements in the evolution of metazoans: the DDI model or how simplicity begets novelty. Brief Funct Genomic Proteomic 8(4):266–275 Kappen C, Schughart K, Ruddle FH (1989) Two steps in the evolution of Antennapedia-class vertebrate homeobox genes. Proc Natl Acad Sci U S A 86(14):5459–5463 Karabinos A, Bhattacharya D (2000) Molecular evolution of calmodulin and calmodulin-like genes in the cephalochordate Branchiostoma. J Mol Evol 51(2):141–148 Kasahara M (2007) The 2R hypothesis: an update. Curr Opin Immunol 19(5):547–552 Kassahn KS, Dang VT, Wilkins SJ, Perkins AC, Ragan MA (2009) Evolution of gene function and regulatory control after whole-genome duplication: comparative analyses in vertebrates. Genome Res 19(8):1404–1418 Katsanis N, Fitzgibbon J, Fisher EMC (1996) Paralogy mapping: identification of a region in the human MHC triplicated onto human chromosomes 1 and 9 allows the prediction and isolation of novel PBX and NOTCH loci. Genomics 35:101–108 Kohn M, Hogel J, Vogel W, Minich P, Kehrer-Sawatzki H, Graves JA, Hameister H (2006) Reconstruction of a 450-My-old ancestral vertebrate protokaryotype. Trends Genet 22(4):203–210 CO RR 928 929 930 931 932 933 934 935 936 937 938 939 940 941 942 943 944 945 946 947 948 949 950 951 952 953 954 955 956 957 958 959 960 961 962 963 964 965 966 967 968 969 970 971 972 973 974 975 976 977 978 979 980 981 C. Cañestro UN Editor Proof 334 Layout: T1 Standard SC Chapter No.: 16 335 EC TE D PR OO F Kuraku S (2008) Insights into cyclostome phylogenomics: pre-2R or post-2R. Zoolog Sci 25(10):960–968 Kuraku S (2010) Palaeophylogenomics of the vertebrate ancestor—impact of hidden paralogy on hagfish and lamprey gene phylogeny. Integr Comp Biol 50(1):124–129 Kuraku S (2011) Hox gene clusters of early vertebrates: do they serve as reliable markers for genome evolution? Genomics Proteomics Bioinformatics 9(3):97–103 Kuraku S, Meyer A, Kuratani S (2009) Timing of genome duplications relative to the origin of the vertebrates: did cyclostomes diverge before or after? Mol Biol Evol 26(1):47–59 Kuratani S (2009) Insights into neural crest migration and differentiation from experimental embryology. Development 136(10):1585–1589 Kuratani S, Ota KG (2008) Hagfish (cyclostomata, vertebrata): searching for the ancestral developmental plan of vertebrates. BioEssays 30(2):167–172 Laisney JA, Braasch I, Walter RB, Meierjohann S, Schartl M (2010) Lineage-specific co-evolution of the Egf receptor/ligand signaling system. BMC Evol Biol 10:27 Larhammar D, Lundin L, Hallbook F (2002) The human Hox-bearing chromosome regions did arise by block or chromosome (or even genome) duplications. Genome Res 12(12): 1910–1920 Lee CT, Risom T, Strauss WM (2007) Evolutionary conservation of microRNA regulatory circuits: an examination of microRNA gene complexity and conserved microRNA-target interactions through metazoan phylogeny. DNA Cell Biol 26(4):209–218 Leveugle M, Prat K, Perrier N, Birnbaum D, Coulier F (2003) ParaDB: a tool for paralogy mapping in vertebrate genomes. Nucleic Acids Res 31(1):63–67 Louis A, Roest Crollius H, Robinson-Rechavi M (2012) How much does the amphioxus genome represent the ancestor of chordates? Brief Funct Genomics 11(2):89–95 Lowe CB, Bejerano G, Haussler D (2007) Thousands of human mobile element fragments undergo strong purifying selection near developmental genes. Proc Natl Acad Sci U S A 104(19):8005–8010 Lundin LG (1979) Evolutionary conservation of large chromosomal segments reflected in mammalian gene maps. Clin Genet 16(2):72–81 Lundin LG (1993) Evolution of the vertebrate genome as relected in paralogous chromosomal regions in man and the house mouse. Genomics 16:1–19 Lundin LG, Larhammar D, Hallbook F (2003) Numerous groups of chromosomal regional paralogies strongly indicate two genome doublings at the root of the vertebrates. J Struct Funct Genomics 3(1–4):53–63 Lynch M, Conery J (2000) The evolutionary fate and consequences of gene duplication. Science 290(5494):1151–1155 Lynch M, Force A (2000) The origin of interspecific genomic incompatibility via gene duplication. Am Nat 156:590–605 Lynch M, O’Hely M, Walsh B, Force A (2001) The probability of preservation of a newly arisen gene duplicate. Genetics 159(4):1789–1804 Lynch VJ, Leclerc RD, May G, Wagner GP (2011) Transposon-mediated rewiring of gene regulatory networks contributed to the evolution of pregnancy in mammals. Nat Genet 43:1154-1159 Lynch VJ, Wagner GP (2009) Multiple chromosomal rearrangements structured the ancestral vertebrate Hox-bearing protochromosomes. PLoS Genet 5(1):e1000349 Makino T, McLysaght A (2010) Ohnologs in the human genome are dosage balanced and frequently associated with disease. Proc Natl Acad Sci U S A 107(20):9270–9274 Martinez-Morales JR, Henrich T, Ramialison M, Wittbrodt J (2007) New genes in the evolution of the neural crest differentiation program. Genome Biol 8(3):R36 Matzke MA, Matzke AJ (1998) Polyploidy and transposons. Trends Ecol Evol 13(6):241 Matzke MA, Mette MF, Matzke AJ (2000) Transgene silencing by the host genome defense: implications for the evolution of epigenetic control mechanisms in plants and vertebrates. Plant Mol Biol 43(2–3):401–415 CO RR 982 983 984 985 986 987 988 989 990 991 992 993 994 995 996 997 998 999 1000 1001 1002 1003 1004 1005 1006 1007 1008 1009 1010 1011 1012 1013 1014 1015 1016 1017 1018 1019 1020 1021 1022 1023 1024 1025 1026 1027 1028 1029 1030 1031 1032 1033 1034 Book ISBN: 978-3-642-31441-4 Page: 335/338 Two Rounds of Vertebrate Whole-Genome Duplication UN Editor Proof 16 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 16 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 336/338 EC TE D PR OO F Mazet F, Hutt JA, Milloz J, Millard J, Graham A, Shimeld SM (2005) Molecular evidence from Ciona intestinalis for the evolutionary origin of vertebrate sensory placodes. Dev Biol 282(2):494–508 McClintock JM, Carlson R, Mann DM, Prince VE (2001) Consequences of Hox gene duplication in the vertebrates: an investigation of the zebrafish Hox paralogue group 1 genes. Development 128(13):2471–2484 McLysaght A, Hokamp K, Wolfe KH (2002) Extensive genomic duplication during early chordate evolution. Nat Genet 31:200–204 Meulemans D, Bronner-Fraser M (2004) Gene-regulatory interactions in neural crest evolution and development. Dev Cell 7(3):291–299 Meulemans D, Bronner-Fraser M (2005) Central role of gene cooption in neural crest evolution. J Exp Zoolog B Mol Dev Evol 304(4):298–303 Meulemans D, Bronner-Fraser M (2007) Insights from amphioxus into the evolution of vertebrate cartilage. PLoS One 2(8):e787 Minguillon C, Ferrier DE, Cebrian C, Garcia-Fernandez J (2002) Gene duplications in the prototypical cephalochordate amphioxus. Gene 287(1–2):121–128 Minguillon C, Jimenez-Delgado S, Panopoulou G, Garcia-Fernandez J (2003) The amphioxus Hairy family: differential fate after duplication. Development 130(24):5903–5914 Muffato M, Louis A, Poisnel CE, Roest Crollius H (2010) Genomicus: a database and a browser to study gene synteny in modern and ancestral genomes. Bioinformatics 26(8):1119–1121 Nadeau JH, Sankoff D (1997) Comparable rates of gene loss and functional divergence after genome duplications early in vertebrate evolution. Genetics 147(3):1259–1266 Nakatani Y, Takeda H, Kohara Y, Morishita S (2007) Reconstruction of the vertebrate ancestral genome reveals dynamic genome reorganization in early vertebrates. Genome Res 17(9):1254–1265 Naruse K, Tanaka M, Mita K, Shima A, Postlethwait J, Mitani H (2004) A medaka gene map: the trace of ancestral vertebrate proto-chromosomes revealed by comparative gene mapping. Genome Res 14(5):820–828 Nelson JS (1994) Fishes of the world, 3rd edn. Wiley-Interscience, New York Nikitina N, Sauka-Spengler T, Bronner-Fraser M (2009) Chapter 1. Gene regulatory networks in neural crest development and evolution. Curr Top Dev Biol 86:1–14 Northcutt RG, Gans C (1983) The genesis of neural crest and epidermal placodes: a reinterpretation of vertebrate origins. Quart Rev Biol 58:1–28 Nowak MA, Boerlijst MC, Cooke J, Smith JM (1997) Evolution of genetic redundancy. Nature 388(6638):167–171 Oda H, Wada H, Tagawa K, Akiyama-Oda Y, Satoh N, Humphreys T, Zhang S, Tsukita S (2002) A novel amphioxus cadherin that localizes to epithelial adherens junctions has an unusual domain organization with implications for chordate phylogeny. Evol Dev 4(6):426–434 Ohno S (1970) Evolution by gene duplication. Springer, New York Ohno S, Wolf U, Atkins NB (1968) Evolution from fish to mammals by gene duplication. Hereditas 59(1):169–187 Ota KG, Kuratani S (2007) Cyclostome embryology and early evolutionary history of vertebrates. Integr Comp Biol 47(3):329–337 Otto SP (2007) The evolutionary consequences of polyploidy. Cell 131(3):452–462 Panopoulou G, Hennig S, Groth D, Krause A, Poustka AJ, Herwig R, Vingron M, Lehrach H (2003) New evidence for genome-wide duplications at the origin of vertebrates using an amphioxus gene set and completed animal genomes. Genome Res 13(6A):1056–1066 Paps J, Holland PW, Shimeld SM (2012) A genome-wide view of transcription factor gene diversity in chordate evolution: less gene loss in amphioxus? Brief Funct Genomics 11(2):177–186 Parisod C, Alix K, Just J, Petit M, Sarilar V, Mhiri C, Ainouche M, Chalhoub B, Grandbastien MA (2010) Impact of transposable elements on the organization and function of allopolyploid genomes. New Phytol 186(1):37–45 CO RR 1035 1036 1037 1038 1039 1040 1041 1042 1043 1044 1045 1046 1047 1048 1049 1050 1051 1052 1053 1054 1055 1056 1057 1058 1059 1060 1061 1062 1063 1064 1065 1066 1067 1068 1069 1070 1071 1072 1073 1074 1075 1076 1077 1078 1079 1080 1081 1082 1083 1084 1085 1086 1087 C. Cañestro UN Editor Proof 336 Layout: T1 Standard SC Chapter No.: 16 337 EC TE D PR OO F Paterson AH, Chapman BA, Kissinger JC, Bowers JE, Feltus FA, Estill JC (2006) Many gene and domain families have convergent fates following independent whole-genome duplication events in Arabidopsis, Oryza, Saccharomyces and Tetraodon. Trends Genet 22(11):597–602 Pebusque MJ, Coulier F, Birnbaum D, Pontarotti P (1998) Ancient large-scale genome duplications: phylogenetic and linkage analyses shed light on chordate genome evolution. Mol Biol Evol 15(9):1145–1159 Pendleton JW, Nagai BK, Murtha MT, Ruddle FH (1993) Expansion of the Hox gene family and the evolution of chordates. Proc Natl Acad Sci U S A 90:6300–6304 Polak P, Domany E (2006) Alu elements contain many binding sites for transcription factors and may play a role in regulation of developmental processes. BMC Genomics 7:133 Popovici C, Leveugle M, Birnbaum D, Coulier F (2001) Coparalogy: physical and functional clusterings in the human genome. Biochem Biophys Res Commun 288(2):362–370 Postlethwait J, Amores A, Cresko W, Singer A, Yan YL (2004) Subfunction partitioning, the teleost radiation and the annotation of the human genome. Trends Genet 20(10):481–490 Postlethwait JH (2007) The zebrafish genome in context: ohnologs gone missing. J Exp Zoolog B Mol Dev Evol 308(5):563–577 Putnam NH, Butts T, Ferrier DE, Furlong RF, Hellsten U, Kawashima T, Robinson-Rechavi M, Shoguchi E, Terry A, Yu JK, Benito-Gutierrez EL, Dubchak I, Garcia-Fernandez J, GibsonBrown JJ, Grigoriev IV, Horton AC, de Jong PJ, Jurka J, Kapitonov VV, Kohara Y, Kuroki Y, Lindquist E, Lucas S, Osoegawa K, Pennacchio LA, Salamov AA, Satou Y, Sauka-Spengler T, Schmutz J, Shin IT, Toyoda A, Bronner-Fraser M, Fujiyama A, Holland LZ, Holland PW, Satoh N, Rokhsar DS (2008) The amphioxus genome and the evolution of the chordate karyotype. Nature 453(7198):1064–1071 Ravi V, Lam K, Tay BH, Tay A, Brenner S, Venkatesh B (2009) Elephant shark (Callorhinchus milii) provides insights into the evolution of Hox gene clusters in gnathostomes. Proc Natl Acad Sci U S A 106(38):16327–16332 Robinson-Rechavi M, Boussau B, Laudet V (2004) Phylogenetic dating and characterization of gene duplications in vertebrates: the cartilaginous fish reference. Mol Biol Evol 21(3): 580–586 Roux J, Robinson-Rechavi M (2008) Developmental constraints on vertebrate genome evolution. PLoS Genet 4(12):e1000311 Ruvinsky I, Silver LM (1997) Newly identified paralogous groups on mouse chromosomes 5 and 11 reveal the age of a T-box cluster duplication. Genomics 40(2):262–266 SanMiguel P, Gaut BS, Tikhonov A, Nakajima Y, Bennetzen JL (1998) The paleontology of intergene retrotransposons of maize. Nat Genet 20(1):43–45 SanMiguel P, Tikhonov A, Jin YK, Motchoulskaia N, Zakharov D, Melake-Berhan A, Springer PS, Edwards KJ, Lee M, Avramova Z, Bennetzen JL (1996) Nested retrotransposons in the intergenic regions of the maize genome. Science 274(5288):765–768 Sauka-Spengler T, Meulemans D, Jones M, Bronner-Fraser M (2007) Ancient evolutionary origin of the neural crest gene regulatory network. Dev Cell 13(3):405–420 Scannell DR, Byrne KP, Gordon JL, Wong S, Wolfe KH (2006) Multiple rounds of speciation associated with reciprocal gene loss in polyploid yeasts. Nature 440(7082):341–345 Semon M, Wolfe KH (2007) Consequences of genome duplication. Curr Opin Genet Dev 17(6):505–512 Semon M, Wolfe KH (2008) Preferential subfunctionalization of slow-evolving genes after allopolyploidization in Xenopus laevis. Proc Natl Acad Sci U S A 105(24):8333–8338 Sempere LF, Cole CN, McPeek MA, Peterson KJ (2006) The phylogenetic distribution of metazoan microRNAs: insights into evolutionary complexity and constraint. J Exp Zool B Mol Dev Evol 306(6):575–588 Seo HC, Edvardsen RB, Maeland AD, Bjordal M, Jensen MF, Hansen A, Flaat M, Weissenbach J, Lehrach H, Wincker P, Reinhardt R, Chourrout D (2004) Hox cluster disintegration with persistent anteroposterior order of expression in Oikopleura dioica. Nature 431(7004):67–71 Sharman AC, Holland PWH (1998) Estimation of hox gene cluster number in lampreys. Int J Dev Biol 42:617–620 CO RR 1088 1089 1090 1091 1092 1093 1094 1095 1096 1097 1098 1099 1100 1101 1102 1103 1104 1105 1106 1107 1108 1109 1110 1111 1112 1113 1114 1115 1116 1117 1118 1119 1120 1121 1122 1123 1124 1125 1126 1127 1128 1129 1130 1131 1132 1133 1134 1135 1136 1137 1138 1139 1140 1141 Book ISBN: 978-3-642-31441-4 Page: 337/338 Two Rounds of Vertebrate Whole-Genome Duplication UN Editor Proof 16 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 16 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 338/338 EC TE D PR OO F Shemer G, Podbilewicz B (2000) Fusomorphogenesis: cell fusion in organ formation. Dev Dyn 218(1):30–51 Shimeld SM, Holland PW (2000) Vertebrate innovations. Proc Natl Acad Sci U S A 97(9): 4449–4452 Sidow A (1996) Gen(om)e duplications in the evolution of early vertebrates. Curr Opin Genet Dev 6:715–722 Siegel N, Hoegg S, Salzburger W, Braasch I, Meyer A (2007) Comparative genomics of ParaHox clusters of teleost fishes: gene cluster breakup and the retention of gene sets following whole genome duplications. BMC Genomics 8:312 Skrabanek L, Wolfe KH (1998) Eukaryote genome duplication—where’s the evidence? Curr Opin Genet Dev 8:694–700 Small KS, Brudno M, Hill MM, Sidow A (2007) A haplome alignment and reference sequence of the highly polymorphic Ciona savignyi genome. Genome Biol 8(3):R41 Soltis DE, Albert VA, Leebens-Mack J, Bell CD, Paterson AH, Zheng C, Sankoff D, Depamphilis CW, Wall PK, Soltis PS (2009) Polyploidy and angiosperm diversification. Am J Bot 96(1):336–348 Soltis DE, Soltis PS (1999) Polyploidy: recurrent formation and genome evolution. Trends Ecol Evol 14(9):348–352 Somorjai I, Bertrand S, Camasses A, Haguenauer A, Escriva H (2008) Evidence for stasis and not genetic piracy in developmental expression patterns of Branchiostoma lanceolatum and Branchiostoma floridae, two amphioxus species that have evolved independently over the course of 200 Myr. Dev Genes Evol 218(11–12):703–713 Spring J (1997) Vertebrate evolution by interspecific hybridization—are we polyploid? Fed Eur Biol Soc Lett 400:2–8 Stadler PF, Fried C, Prohaska SJ, Bailey WJ, Misof BY, Ruddle FH, Wagner GP (2004) Evidence for independent Hox gene duplications in the hagfish lineage: a PCR-based gene inventory of Eptatretus stoutii. Mol Phylogenet Evol 32(3):686–694 Storchova Z, Pellman D (2004) From polyploidy to aneuploidy, genome instability and cancer. Nat Rev Mol Cell Biol 5(1):45–54 Takio Y, Pasqualetti M, Kuraku S, Hirano S, Rijli FM, Kuratani S (2004) Evolutionary biology: lamprey Hox genes and the evolution of jaws. Nature 429(6989):1–262 Taylor JS, Raes J (2004) Duplication and divergence: the evolution of new genes and old ideas. Annu Rev Genet 38:615–643 Taylor JS, van de Peer Y, Meyer M (2001) Genome duplication, divergent resolution and speciation. Trends Genet 17:299–301 Thornburg BG, Gotea V, Makalowski W (2006) Transposable elements as a significant source of transcription regulating signals. Gene 365:104–110 van de Lagemaat LN, Landry JR, Mager DL, Medstrand P (2003) Transposable elements in mammals promote regulatory variation and diversification of genes with specialized functions. Trends Genet 19(10):530–536 Van de Peer Y, Maere S, Meyer A (2009) The evolutionary significance of ancient genome duplications. Nat Rev Genet 10(10):725–732 Venkatesh B, Kirkness EF, Loh YH, Halpern AL, Lee AP, Johnson J, Dandona N, Viswanathan LD, Tay A, Venter JC, Strausberg RL, Brenner S (2007) Survey sequencing and comparative analysis of the elephant shark (Callorhinchus milii) genome. PLoS Biol 5(4):e101 Venter JC, Adams MD, Myers EW, Li PW, Mural RJ, Sutton GG, Smith HO, Yandell M, Evans CA, Holt RA et al (2001) The sequence of the human genome. Science 291(5507):1304–1351 Vienne A, Shiina T, Abi-Rached L, Danchin E, Vitiello V, Cartault F, Inoko H, Pontarotti P (2003) Evolution of the proto-MHC ancestral region: more evidence for the plesiomorphic organisation of human chromosome 9q34 region. Immunogenetics 55(7):429–436 Wada H, Okuyama M, Satoh N, Zhang S (2006) Molecular evolution of fibrillar collagen in chordates, with implications for the evolution of vertebrate skeletons and chordate phylogeny. Evol Dev 8(4):370–377 CO RR 1142 1143 1144 1145 1146 1147 1148 1149 1150 1151 1152 1153 1154 1155 1156 1157 1158 1159 1160 1161 1162 1163 1164 1165 1166 1167 1168 1169 1170 1171 1172 1173 1174 1175 1176 1177 1178 1179 1180 1181 1182 1183 1184 1185 1186 1187 1188 1189 1190 1191 1192 1193 1194 C. Cañestro UN Editor Proof 338 Layout: T1 Standard SC Chapter No.: 16 339 EC TE D PR OO F Werth CR, Windham MD (1991) A model for divergent, allopatric speciation of polypoid pteridophytes resulting from silencing of duplicate-gene expression. Am Nat 137:515–526 Weston JA (1970) The migration and differentiation of neural crest cells. Adv Morphog 8:41–114 Wolfe K (2000) Robustness—it’s not where you think it is. Nat Genet 25(1):3–4 Wolfe KH (2001) Yesterday’s polyploids and the mystery of diploidization. Nat Rev Genet 2(5):333–341 Wolfe KH, Shields DC (1997) Molecular evidence for an ancient duplication of the entire yeast genome. Nature 387(6634):708–713 Wotton KR, Shimeld SM (2006) Comparative genomics of vertebrate Fox cluster loci. BMC Genomics 7:271 Yu JK, Meulemans D, McKeown SJ, Bronner-Fraser M (2008) Insights from the amphioxus genome on the origin of vertebrate neural crest. Genome Res 18(7):1127–1132 CO RR 1195 1196 1197 1198 1199 1200 1201 1202 1203 1204 1205 1206 Book ISBN: 978-3-642-31441-4 Page: 339/338 Two Rounds of Vertebrate Whole-Genome Duplication UN Editor Proof 16 Book ID: 272454_1_En Date: 16-8-2012 Metadata of the chapter that will be visualized in SpringerLink Book Title Polyploidy and Genome Evolution Series Title Chapter Title Polyploidy in Fish and the Teleost Genome Duplication Copyright Year 2012 Copyright HolderName Springer-Verlag Berlin Heidelberg Corresponding Author Family Name Postlethwait Particle Given Name John H. Suffix Author Division Institute of Neuroscience Organization University of Oregon Address 97403, Eugene, OR, USA Email jpostle@uoneuro.uoregon.edu Family Name Braasch Particle Given Name Ingo Suffix Abstract Division Institute of Neuroscience Organization University of Oregon Address 97403, Eugene, OR, USA Email ibraasch@uoneuro.uoregon.edu Multiple rounds of whole-genome duplications (WGDs) punctuated the evolution of rayfin fish, a speciesrich group comprising about half of all vertebrates. Rayfin fish, along with lobefin vertebrates including humans, derive from early vertebrate ancestors that evolved through two rounds of polyploidization (the first and second rounds of vertebrate genome duplication, VGD1 and VGD2) at the dawn of the vertebrate lineage. Furthermore, teleost fish underwent an additional round of polyploidization in their stem lineage, the teleost genome duplication (TGD). Additional WGD events occurred independently in numerous species and higher level taxa of teleosts and other rayfin fish, for example in salmonids, carp, and sturgeon, so that some fish lineages experienced at least four rounds of WGD since the origin of vertebrates. This chapter provides an overview of these polyploidization events in the fish lineage and focuses on the impact these genome duplications (GD) had on genome evolution in selected fish taxa. We then review evidence for the TGD and discuss its consequences for the evolution of gene content, order, and functions in the teleost lineage. We argue that, although evidence remains sparse, the TGD may have had a profound influence on the evolutionary success and the biodiversity of teleosts. Importantly, an in-depth understanding of the causes and consequences of the TGD and other teleost GD events will help to inform us about the evolution of our own paleopolyploid genome. 1 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 341/383 Chapter 17 4 Ingo Braasch and John H. Postlethwait 11 12 13 14 15 16 17 18 19 20 21 22 23 24 PR OO D 9 10 TE 8 EC 7 Abstract Multiple rounds of whole-genome duplications (WGDs) punctuated the evolution of rayfin fish, a species-rich group comprising about half of all vertebrates. Rayfin fish, along with lobefin vertebrates including humans, derive from early vertebrate ancestors that evolved through two rounds of polyploidization (the first and second rounds of vertebrate genome duplication, VGD1 and VGD2) at the dawn of the vertebrate lineage. Furthermore, teleost fish underwent an additional round of polyploidization in their stem lineage, the teleost genome duplication (TGD). Additional WGD events occurred independently in numerous species and higher level taxa of teleosts and other rayfin fish, for example in salmonids, carp, and sturgeon, so that some fish lineages experienced at least four rounds of WGD since the origin of vertebrates. This chapter provides an overview of these polyploidization events in the fish lineage and focuses on the impact these genome duplications (GD) had on genome evolution in selected fish taxa. We then review evidence for the TGD and discuss its consequences for the evolution of gene content, order, and functions in the teleost lineage. We argue that, although evidence remains sparse, the TGD may have had a profound influence on the evolutionary success and the biodiversity of teleosts. Importantly, an in-depth understanding of the causes and consequences of the TGD and other teleost GD events will help to inform us about the evolution of our own paleopolyploid genome. CO RR 5 6 F 3 Polyploidy in Fish and the Teleost Genome Duplication 2 I. Braasch  J. H. Postlethwait (&) Institute of Neuroscience, University of Oregon, Eugene, OR 97403, USA e-mail: jpostle@uoneuro.uoregon.edu UN Editor Proof Layout: T1 Standard SC Chapter No.: 17 I. Braasch e-mail: ibraasch@uoneuro.uoregon.edu P. S. Soltis and D. E. Soltis (eds.), Polyploidy and Genome Evolution, DOI: 10.1007/978-3-642-31442-1_17,  Springer-Verlag Berlin Heidelberg 2012 341 Layout: T1 Standard SC Chapter No.: 17 Book ID: 272454_1_En Date: 16-8-2012 I. Braasch and J. H. Postlethwait The problems of this world are only truly solved in two ways: by extinction or duplication. Susan Sontag 25 26 27 28 29 30 31 32 33 34 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 D 38 In common usage, ‘‘fish’’ applies to any aquatic vertebrate that possesses gills and fins (if any appendages) (Nelson 2006). Following this phenotypic definition, ‘‘fish’’ include jawless fish such as lampreys and hagfish, cartilaginous fish, such as sharks and rays, lobefin fish including coelacanth and lungfish, and rayfin fish such as teleosts. ‘‘Fish’’, however, are a paraphyletic assemblage because it excludes tetrapod vertebrates, which share a more recent common ancestor with lungfish than lungfish share with teleost fish. Until we tend to call tetrapods (including ourselves) lobefin fish and accept that all living vertebrates are fish, the term ‘‘fish’’ should be used with caution (see Fig. 17.1). This chapter focuses on the monophyletic group of rayfin fish, or Actinopterygii, and, in more detail, their largest subgroup, teleost fish. The jawless, cartilaginous, and lobefin fish, however, are important for understanding evolution of the vertebrate genome by whole-genome duplications (WGDs) at the dawn of the vertebrate lineage and are thus featured in Chap. 12 (this volume). Among actinopterygian fish, the teleosts are by far the largest subgroup in terms of extant species numbers. Nelson (2006) lists 26,891 living rayfin fish species (453 families), of which only 51 species (5 families) are not teleosts. The teleosts, in contrast, comprise 26,840 living species (448 families), an impressive 99.8 % of all rayfins. Even more impressive, adding all living vertebrates to the calculation gives around 50 % of all vertebrates being teleost fish. Draft genome assemblies of ten rayfin species (all of them clupeocephalan teleosts) are currently publicly available (Fig. 17.1): zebrafish (Danio rerio), threespined stickleback (Gasterosteus aculeatus), medaka (Oryzias latipes) (Kasahara et al. 2007), fugu (Takifugu rubripes) (Aparicio et al. 2002), spotted green pufferfish (Tetraodon nigroviridis) (Jaillon et al. 2004), Atlantic cod (Gadus morhua) (Star et al. 2011), and four species of East African cichlids, including tilapia (Oreochromis niloticus). More species such as platyfish (Xiphophorus maculatus), Atlantic salmon (Salmo salar), rainbow trout (Oncorhynchus mykiss), and the nonteleost spotted gar (Lepisosteus oculatus; see below) are soon to be added to this list (Fig. 17.1), and numerous additional rayfins will be sequenced as part of the GENOME 10K Project (Haussler et al. 2009). TE 37 EC 36 17.1 A Brief Introduction to ‘‘Fish’’ CO RR 35 PR OO F In their great numbers and degree of anatomical diversity, the modern ray-finned fishes may be considered the most successful of all vertebrates. Robert L. Carroll UN Editor Proof 342 Book ISBN: 978-3-642-31441-4 Page: 342/383 Layout: T1 Standard SC Chapter No.: 17 Book ISBN: 978-3-642-31441-4 Page: 343/383 Polyploidy in Fish and the Teleost Genome Duplication Editor Proof 17 Book ID: 272454_1_En Date: 16-8-2012 343 Spotted green pufferfish Tetraodon nigroviridis Medaka Oryzias latipes Platyfish Xiphophorus maculatus Atlantic cod Gadus morhua SaGD Salmonids Zebrafish Danio rerio TGD F rayfin fish PR OO East African cichlids clupeocephala teleosts Three-spined stickleback Gasterosteus aculeatus percomorpha Japanese pufferfish Takifugu rubripes Carps + Goldfish CGD D Goldeye Hiodon alosoides TE Bowfin Amia calva Spotted gar Lepisosteus oculatus AGDs VGD1 700 CO RR VGD2 600 500 400 300 200 Bichirs Tetrapods Lungfish lobefin fish EC Sturgeons Coelacanth Cartilaginous fish Jawless fish 100 0 MYA UN Fig. 17.1 A phylogeny of all kinds of ‘‘fish’’. Tree topology and timing of divergence events follows data from Setiamarga et al. (2009) and www.timetree.org. The figure shows the timing of genome duplication (GD) events (VGD1/VGD2 the vertebrate GDs, TGD the teleost GD, AGDs acipenserid GDs, SaGD the salmonid GD, CGD the carp GD). Whether the second vertebrate genome duplication (VGD2) occurred before or after the divergence of jawless fish remains controversial. MYA million years ago Layout: T1 Standard SC Chapter No.: 17 Book ID: 272454_1_En Date: 16-8-2012 I. Braasch and J. H. Postlethwait 71 The species richness of rayfin and teleost fish poses the question of the secret of their evolutionary success. A recurring pattern of genome evolution in rayfin fish is their propensity to polyploidization. In line with this idea, a positive correlation of genome size and species richness has been reported among actinopterygians as well as for specific teleost clades (Mank and Avise 2006a). 72 17.2 Duplications in the Crown: Polyploidy in Rayfin Fish 67 68 69 PR OO F 70 91 17.2.1 Twigs and Leaves: Polyploidization Events in Fish 79 80 81 82 83 84 85 86 87 88 89 92 93 94 95 96 97 98 99 100 101 102 103 TE 77 78 EC 75 76 CO RR 74 D 90 In comparison to plants, polyploidizations are rare in the animal kingdom (Muller 1925), yet the reasons for this genomic discrepancy between fauna and flora remain controversial (Mable 2004). In the vertebrate lineage, polyploids are most often found among amphibians and rayfin fish (Mable et al. 2011; Otto 2007; Comai 2005). Factors that may favor the proclivity of rayfin fish to generate surviving polyploid lineages compared to other groups of vertebrates (e.g., high production rate of unreduced gametes, their propensity for hybridization, genomic flexibility, etc.) have been extensively discussed recently elsewhere, yet a clear explanation remains elusive (Mable et al. 2011). Given that polyploid events also appear to have occurred in some jawless, cartilaginous, and lobefin fish (see Leggatt and Iwama 2003 and references therein), it may be worth considering that a paucity of genome duplication (GD) is an amniote vertebrate-specific phenomenon. Several works provide comprehensive lists of known cases of polyploidy in fish (Otto and Whitton 2000; Leggatt and Iwama 2003; Le Comber and Smith 2004; Mable et al. 2011). Here, we give an overview of the few general trends that can be inferred from the phylogenetic distribution of polyploidization events and will then discuss several specific cases in more detail. 73 Despite large variations in genome size, rayfin fish have surprisingly conserved karyotypes, with the majority of diploid genomes having 48 or 50 chromosomes (Ohno et al. 1968; Mank and Avise 2006b); thus, chromosome counts have been used as indicators of polyploidizations in fish. A study based on chromosome counts of 615 rayfin species concluded that at least 7–20 polyploidization events occurred in the evolution of extant rayfins (Mank and Avise 2006b); this estimate is likely an underestimate because less than 3 % of all living rayfin species were included in this study. Polyploidization events in fish are phylogenetically restricted, i.e., are unevenly distributed across the actinopterygian tree. Some groups, especially those occupying early diverging branches in rayfin phylogeny, and among teleosts particularly the Ostariophysi (including carps and suckers and many others), seem to be UN Editor Proof 344 Book ISBN: 978-3-642-31441-4 Page: 344/383 Layout: T1 Standard SC Chapter No.: 17 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 F 109 PR OO 107 108 D 106 particularly prone to polyploidization in their lineage, while other groups, including more derived lineages such as the percomorphs (perch and relatives, like pufferfish, stickleback, cichlids and many more), have few or no GDs after the teleost genome duplication (TGD) (Leggatt and Iwama 2003; Mable et al. 2011). Polyploidizations have occurred in many individual fish genera, species, and populations (i.e., the ‘‘leaves’’ of the rayfin tree), as for example in multiple species of barbs (genus Barbus, family Cyprinidae) (Chenuil et al. 1999), in the cyprinid Squalius alburnoides species complex (Alves et al. 2001), in the pond loach (Misgurnus anguillicaudatus) (Li et al. 2011) and other loach species (Cobitidae, order Cypriniformes) (Ferris and Whitt 1977a), and in the stinging catfish (Heteropneustes fossilis, order Siluriformes) (Pandian and Koteeswaran 1999). On the other hand, piscine GD events have also occurred in the last common ancestor of higher level taxa (some of the tree’s ‘‘twigs’’). All members are polyploid in the three families Salmonidae (salmonids) (Allendorf and Thorgaard 1984), Callichthyidae (armored catfish) (Oliveira et al. 1992), and Catostomidae (suckers) (Ferris 1984; Uyeno and Smith 1972). Common carp and goldfish (subfamily Cyprininae) also arise from a shared tetraploid origin (Ohno et al. 1967; David et al. 2003; Schultz 1980; Larhammar and Risinger 1994). Among rayfins diverging before teleosts, the Acipenseridae (sturgeons) are famous for their multiple ploidy levels, up to at least hexadecaploid (Birstein et al. 1997; Ludwig et al. 2001). In addition, artificial polyploids, particularly triploids, are routinely produced for multiple commercially important fish species to induce sterility and thereby improve somatic growth rate (Piferrer et al. 2009). TE 105 345 EC 104 Book ISBN: 978-3-642-31441-4 Page: 345/383 Polyploidy in Fish and the Teleost Genome Duplication 17.2.2 Piscine Polyploidy and the Evolution of Genome Function: Salmon, Carp, and Calandinos 133 17.2.2.1 Salmonids 130 131 134 135 136 137 138 139 140 CO RR 132 Traditional studies of polyploidy in fish generally involve the analysis of karyotypes and allozymes. Only recently, with the advent of genome sequencing techniques, we have begun to understand some of the evolutionary footprints left by polyploidization events in fish genomes. 129 Salmonids (salmon, trout, whitefish, grayling) are among the fish of highest economic importance (Davidson et al. 2010). Salmonids have been suspected to be polyploids since the 1940s (Svärdson 1945; Kupka 1948), but it was several decades until it was generally accepted that all living salmonids are of autotetraploid origin due to a GD event that occurred in their stem lineage 25–100 million years ago (Ohno et al. 1968; Allendorf and Thorgaard 1984). This WGD is unlikely to have occurred by segmental duplications because all segments of each UN Editor Proof 17 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 17 Book ID: 272454_1_En Date: 16-8-2012 148 149 150 151 152 153 154 155 156 157 158 159 160 161 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 184 185 F 146 147 PR OO 145 D 144 TE 143 chromosome appear to be in duplicate (although of course not every gene in each segment is still present in duplicate); if segmental duplications had been responsible, then some chromosome segments would be present in three or five copies. Salmonids still show multivalent chromosomes during meiosis as well as tetrasomic inheritance of some loci (Allendorf and Thorgaard 1984; Phillips and Rab 2001) and are thus considered pseudotetraploids, i.e., the process of diploidization has not yet concluded in these fish (Danzmann et al. 2008). The autotetraplodization in salmonids is sometimes referred to as ‘‘4R’’, because it is the fourth round of WGD in this lineage since the rise of the vertebrate lineage (see later this chapter and Chap. 16, this volume). This numerating terminology, however, imposes difficulties to distinguish independent GD events in various lineages of rayfins as discussed above. Therefore, we suggest using the term salmonid genome duplication (SaGD). Recent progress in salmonid genomics (Davidson et al. 2010; Miller et al. 2011) has provided early insights into the evolutionary dynamics of the salmonid genome after the SaGD. In line with tetraploidization in a salmonid ancestor, the diploid genomes of most salmonids contain 96–104 chromosome arms. The Atlantic salmon is a special case among salmonids, because its more derived karyotype has been secondarily reduced to 72–74 chromosome arms by chromosome fusions (Phillips and Rab 2001; Phillips et al. 2009). Conserved synteny analysis comparing genetic linkage maps of Atlantic salmon and rainbow trout (Danzmann et al. 2008; Lien et al. 2011) and analyzing the genomic distribution of conserved non-coding elements (CNEs) (Moghadam et al. 2009) show first that gene order in salmonids generally reflects the inferred ancestral rayfin karyotypes (Nakatani et al. 2007; Kasahara et al. 2007), and second that two salmonid homeologous chromosomes share conserved synteny with one teleost outgroup chromosome (e.g., medaka, zebrafish, and stickleback), as would be expected from an additional round of GD. The duplication of hox gene clusters, generally a good indicator for GDs in animals (see Chap. 16, this volume), has been studied in detail in salmonids. Salmonid genomes contain at least 13 hox clusters, nearly twice as many as most diploid teleost genomes (Moghadam et al. 2005a, b; Mungpakdee et al. 2008a). Given that gene loss (nonfunctionalization, or pseudogenization) is the most frequent fate of gene duplicates (Lynch and Conery 2000), it is surprising that the salmonid hox cluster repertoire, as judged by collapsing duplicated clusters, shows no loss of hox paralogy groups between the ancestral teleost and today’s Atlantic salmon genome, which has retained more hox paralogy groups than any other studied teleost (Mungpakdee et al. 2008a). After the SaGD, several hox paralogs evolved asymmetrically, as manifested by pseudogenization, elevated rates of molecular evolution, and/or divergence of non-coding regions in one of the two hox gene duplicates. This asymmetry is also apparent over entire hox clusters (Mungpakdee et al. 2008a). Overall, however, expression domains of duplicated Atlantic salmon hox genes are conserved with their unduplicated orthologs in other teleosts, although salmon hox paralogs show evidence for quantitative subfunctionalization (Mungpakdee et al. 2008b), a predicted type of functional evolution of duplicate genes (Force et al. 1999). EC 142 CO RR 141 I. Braasch and J. H. Postlethwait UN Editor Proof 346 Book ISBN: 978-3-642-31441-4 Page: 346/383 Layout: T1 Standard SC Chapter No.: 17 193 194 195 196 197 198 199 200 201 202 203 204 205 206 207 208 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 F 192 PR OO 190 191 17.2.2.2 Carp and Goldfish: Facilitated Domestication by Genome Duplication? An allotetraploidization occurred in a shared ancestor of common carp (Cyprinus carpus; 2n = 100) and goldfish (Cyprinus carpio; 2n = 100). The carp genome duplication (CaGD) occurred between 11 and 21 million years ago (Risinger and Larhammar 1993; Ohno et al. 1967; Larhammar and Risinger 1994; David et al. 2003; Yuan et al. 2010). Evidence for the hybrid (i.e., allotetraploid) origin of these polyploids comes from the absence of quadrivalents (Ohno et al. 1967) and the disomic inheritance of genetic markers in view of the short time period since the GD event (David et al. 2003). Independent chromosomal rearrangements took place in goldfish and carp after the allotetraploidization (Ohno et al. 1967). Around 52–60 % of genomic loci of the common carp are still duplicated (David et al. 2003; Ferris and Whitt 1977b), but duplicate retention seems to be considerably lower in goldfish (Woods and Buth 1984). As in salmonids, several duplicates of hox cluster genes have become pseudogenes after the CaGD (Luo et al. 2007; Yuan et al. 2010). Like salmonids, the common carp and its relatives are important economic species in aquaculture as food sources and as ornamental species. Common carp and goldfish are the oldest domesticated fish species, bred in central Europe for *2,000 years and in China for *1,000 years, respectively (reviewed in Balon 2004). During domestication of carp and goldfish, significant morphological change has been fixed in populations compared to their wild ancestors in a short period of time, as exemplified by the multiple, independent loss of scales in the carp (Balon 2004), the color morphs of koi carp (David et al. 2004), and the selection for monstrosities such as telescope eyes, lion-head, or fin loss in ornamental goldfish (Komiyama et al. 2009). The molecular basis of the strong response to artificial selection in carp and goldfish has rarely been explored. Based on the multigenic inheritance of coloration in ornamental koi carp, it has been suggested that the propensity to generate the many color morphs may be related to the tetraploidization of carps (David et al. 2004). Is it possible that allopolyploidization provided carp and goldfish with a high degree of genomic flexibility similar to domestic polyploid plants (see Chaps. 7 and 10, this volume)—made them particularly responsive to morphological selection D 189 TE 188 347 Overall, around 50 % of SaGD-duplicated loci have been retained in salmonids (Allendorf and Thorgaard 1984). Although purifying selection appears to be the prevalent mode of salmonid gene duplicate evolution, asymmetric divergence of salmonid gene duplicates was also observed in an analysis of EST sequences when compared to the diploid Northern pike (Esox lucius) as outgroup (Koop et al. 2008; Leong et al. 2010). The asymmetric pattern of molecular evolution among paralogs, observed particularly often for DNA binding proteins, is thought to be caused by relaxed functional constraints on one of the two paralogs after the SaGD (Leong et al. 2010; Mungpakdee et al. 2008a) as for GDs in general. EC 187 CO RR 186 Book ISBN: 978-3-642-31441-4 Page: 347/383 Polyploidy in Fish and the Teleost Genome Duplication UN Editor Proof 17 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 17 Book ID: 272454_1_En Date: 16-8-2012 I. Braasch and J. H. Postlethwait 234 17.2.2.3 Gene Expression Regulation in Triploid Calandino 230 231 232 PR OO 229 F 233 and domestication? A recent study of the common carp suggests that the duplication of the fgfr1a gene during the course of the CaGD may have permitted the breeding of ‘‘mirror’’ forms that lack almost all scales (Rohner et al. 2009). Importantly, two different domesticated forms of mirror carp have been bred by selection of two different loss-of-function mutations in the same paralog, fgfr1a1 (Rohner et al. 2009). 228 261 17.3 The Doubled Trunk: The Teleost Genome Duplication 239 240 241 242 243 244 245 246 247 248 249 250 251 252 253 254 255 256 257 258 259 262 263 264 265 TE 238 EC 237 CO RR 236 D 260 Although polyploidization events in salmonids and carp are fairly recent events on the evolutionary timescale, they are nevertheless too old to study changes in gene expression pattern following soon after polyploidization and hybridization in fish and the interplay of these two factors. The relationship of polyploidization and hybridization, however, has been analyzed in naturally occurring populations of an Iberian cyprinid, the calandino (S. alburnoides). The allopolyploid S. alburnoides species complex consists of di-, tri-, and tetraploid fish of different genomic compositions derived from interspecific hybridizations of a paternal ancestor (the A genome) and different, geographically separated, maternal genome contributors, the southern S. pyraneicus (P genome) or the northern S. caroliterii (C genome) (Alves et al. 2001; Pala et al. 2010). Expression studies of seven genes in different adult tissues revealed that dosage compensation occurs in polyploids, so that overall gene expression is reduced to the diploid level (Pala et al. 2008, 2010). In S. alburnoides, gene dosage compensation is accomplished by a complex pattern of gene copy silencing in a genespecific and tissue-specific manner: for example, in southern triploids of the PAA genome composition, some genes are expressed only from the A genome, others from P and A genomes; some tissues express only A genes, while others express both P and A genes; some genes are expressed from A in one tissue, but from P and A in another tissue (Pala et al. 2008, 2010). The gene expression patterns also depend on genomic composition: while the A genome allele is dominant in southern individuals, C and A alleles are co-dominant in northern polyploids (CAA, CCA, CCAA, or C-A-) (Pala et al. 2010). Although the exact molecular mechanism leading to dosage compensation in the calandino remains elusive at this point, these studies illustrate the potential these animals provide for the study of gene regulation in diverging polyploid populations. 235 UN Editor Proof 348 Book ISBN: 978-3-642-31441-4 Page: 348/383 Although recent polyploidization events appear to be comparatively common in rayfin fish compared to other vertebrate clades, the most important polyploidization event for genome evolution in fish took place on the branch leading to the teleosts, the largest clade of rayfin fish. Initial arguments for and against the Layout: T1 Standard SC Chapter No.: 17 Book ISBN: 978-3-642-31441-4 Page: 349/383 Polyploidy in Fish and the Teleost Genome Duplication 349 269 17.3.1 Evidence for the Teleost Genome Duplication 270 17.3.1.1 Expansion of Vertebrate Gene Families in Rayfin Fish PR OO 267 F 268 occurrence of such a third round of vertebrate WGD (alias 3R, fish-specific genome duplication, FSGD, or teleost genome duplication, TGD) have been overcome by sequencing the genomes of several teleost species. 266 287 17.3.1.2 Teleost Hox Gene Clusters 276 277 278 279 280 281 282 283 284 285 288 289 290 291 292 293 294 295 296 297 298 299 300 301 302 TE 275 EC 273 274 CO RR 272 D 286 Initially, the branch leading to the teleosts did not raise suspicion of polyploidization because the ancestral vertebrate karyotype, like that of teleosts, was initially inferred to be 48 chromosomes (Ohno et al. 1968). A general observation since the rise of allozyme data, however, was that many enzyme loci with a single locus in tetrapods appeared to have multiple copies in fish (see Morizot 1990 and references therein). With the advent of the zebrafish and other teleosts as model organisms for developmental and genetic studies in the 1990s and the accompanying cloning of teleost DNA sequences, this impression of ‘‘more genes in fish’’ (Wittbrodt et al. 1998) was substantially reinforced. Three alternative explanations were initially considered to explain these observations: (1) higher frequencies of individual or local or tandem gene duplication in teleosts than in tetrapods; (2) higher retention of gene duplicates from the ancestral vertebrate genome duplications (VGD) in teleosts than in tetrapods; or (3) a WGD in the rayfin fish lineage after the divergence from tetrapods (Morizot 1990; Postlethwait et al. 1998; Wittbrodt et al. 1998; Meyer 1998; RobinsonRechavi et al. 2001). 271 Sequencing the hox gene clusters, which had already suggested the earlier VGDs (see Chap. 16, this volume), settled the argument toward the teleost-specific GD hypothesis: since the cloning of the first teleost hox genes in the late 1980s (Eiken et al. 1987; Njolstad et al. 1988), it became apparent that hox genes were no exception to the generalization of ‘‘more genes in fish’’ and additional paralogs of hox genes were found in different teleosts (e.g., Misof and Wagner 1996; Aparicio et al. 1997; Prince et al. 1998). The genomes of tetrapods and other non-teleost vertebrates generally have four hox gene clusters (Graham et al. 1989). The discovery that hox genes are actually organized in seven gene clusters in the zebrafish, that two copies of the hoxA cluster were present in fugu, and that each duplicated hox cluster in fugu was orthologous to a single duplicated copy in zebrafish provided the first significant support that a GD had occurred in the lineage leading to zebrafish and that the event was shared by fugu (Amores et al. 1998; Aparicio et al. 1997). Importantly, the hox clusters and genes closely linked to the hox clusters were found to be distributed over eight different chromosomes UN Editor Proof 17 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 17 Book ID: 272454_1_En Date: 16-8-2012 303 304 305 306 in zebrafish, as would be expected if a WGD increased the hox cluster number from four to eight, followed by the loss of one of the seven hox complements (Amores et al. 1998). Seven or eight hox gene clusters have since been identified in several other teleosts (see e.g. Prohaska and Stadler 2004; Hoegg and Meyer 2005; Hoegg et al. 2007, and references therein). 308 309 PR OO F 307 I. Braasch and J. H. Postlethwait 17.3.1.3 More than just Hox: Duplication of Other Multigene Families in Fish 337 17.3.1.4 Global Evidence from Teleost Genome Projects 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 338 339 340 341 TE 315 EC 314 CO RR 312 313 D 336 Although the genomic location of hox gene clusters and linked gene families could be best explained by a WGD in the lineage leading to teleosts (Meyer and Schartl 1999), several alternative possibilities were discussed, including: (1) that the seven hox clusters in zebrafish were due to polyploidization specific to the zebrafish lineage (Stellwag 1999); (2) that hox cluster-containing chromosomal segments had been amplified in teleosts by local duplications (Elgar et al. 1999); and/or (3) that teleosts had an unusually high gene duplication rate (Robinson-Rechavi et al. 2001). Phylogenetic analyses of numerous duplicated zebrafish genes, however, including those not related or linked to the hox genes and those distributed all over the genome, provided evidence for their origin after the divergence of zebrafish and tetrapod lineages (Taylor et al. 2001; Van de Peer et al. 2001). Additional work on specific gene families was also in line with this hypothesis [e.g. mitf (Altschmied et al. 2002), midkines (Winkler et al. 2003), sox9 (Cresko et al. 2003), egfr (Gomez et al. 2004), pomc (de Souza et al. 2005), and receptor tyrosine kinases (Braasch et al. 2006)]. At the same time, putting more and more genes on the genetic maps of zebrafish (Gates et al. 1999; Postlethwait et al. 2000; Woods et al. 2000, 2005) and medaka (Naruse et al. 2004) provided additional evidence for a teleost WGD beyond just the duplication of hox cluster-bearing chromosomes. Importantly, the inclusion of gene sequences from the Japanese pufferfish (T. rubripes) revealed that these fish-specific gene duplications within multigene families usually date back to a point during rayfin fish evolution before the divergence of zebrafish and pufferfish (Amores et al. 1998; Taylor et al. 2003). Therefore, it became apparent that the GD shared by these teleost fish most likely occurred somewhere along the branch to the teleost fish; but did it occur after the origin of teleosts? Did it include non-teleost rayfin fish? Or did it include all teleosts and only teleosts? 310 311 UN Editor Proof 350 Book ISBN: 978-3-642-31441-4 Page: 350/383 The sequence of the fugu (T. rubripes) genome was the second vertebrate and the first teleost fish draft genome assembly (Aparicio et al. 2002). The authors were cautious about making conclusions from the Takifugu genome assembly with respect to the TGD hypothesis, stating that the distribution of gene duplicates in Layout: T1 Standard SC Chapter No.: 17 Book ISBN: 978-3-642-31441-4 Page: 351/383 Polyploidy in Fish and the Teleost Genome Duplication 351 371 17.3.1.5 Phylogenetic Timing of the Teleost Genome Duplication 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 372 373 374 375 376 377 378 379 380 381 382 PR OO 348 D 346 347 TE 345 EC 344 CO RR 343 F 370 the Takifugu genome indicated segmental or large-scale, but not tandem, duplications (Aparicio et al. 2002). Two follow-up large-scale analyses, however, came to the conclusion that the distribution and age of gene duplicates in the Takifugu genome were indeed consistent with the TGD hypothesis (Vandepoele et al. 2004; Christoffels et al. 2004). Both studies found that the majority of gene duplicates in Takifugu were located in paralogons, i.e., chromosomal blocks that share paralogous syntenies within the genome. Also, molecular clock-based analyses of gene families containing fish-specific gene duplicates helped to date the duplication event and hinted at their occurrence around the origin of the teleost lineage. In contrast to the genome assembly of Takifugu, the genome assembly of the spotted green pufferfish (T. nigroviridis) was anchored onto chromosomes, which made possible a more detailed analysis of conserved synteny blocks within a teleost genome (Jaillon et al. 2004). The publication of the Tetraodon genome assembly put the TGD beyond doubt, by showing that internal paralogy and chromosomal blocks of doubled conserved synteny extended genome-wide within this pufferfish genome. Doubled conserved synteny refers to a human chromosomal region sharing conserved synteny with two pufferfish paralogons (see Sect. 17.3.2.2). These patterns of conserved synteny also led to reconstructions of ancestral karyotypes (Jaillon et al. 2004). With the publication of the medaka genome assembly, patterns of conserved synteny within and between the genomes of medaka, pufferfish, and zebrafish, using human as an outgroup, allowed more refined reconstruction of ancestral, preWGD karyotypes (see Sect. 17.3.2.5) (Kasahara et al. 2007). Formal publications of the zebrafish and stickleback genome assemblies are pending, but analyses of these two species (as exemplified in Sect. 17.3.2) overall confirm observations made from the genomes of pufferfish and medaka: the occurrence of a WGD in the lineage leading to the teleosts is the most parsimonious explanation of all available genome data. 342 Genome analyses of zebrafish, pufferfish, medaka, and stickleback pointed to a GD event somewhere in the rayfin fish lineage prior to the divergence of these teleosts, but when exactly did the event occur? Is it, for example, a trait shared only by clupeocephalan teleosts, to which the aforementioned sequenced species belong (Fig. 17.1)? Or did it happen in an ancestor of all teleosts? Or even earlier, before the rise of the teleost lineage and as such represents a trait shared with earlier diverging rayfin branches, such as those leading to bowfin, gar, sturgeon, or bichir? Of course, the phylogenetic timing of the TGD is essential for understanding any possible causal relationships between the TGD and the teleost radiation. This question is further complicated by difficulties in establishing historical relationships of various rayfin fish lineages diverging basal to the teleosts as well as of the UN Editor Proof 17 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 17 Book ID: 272454_1_En Date: 16-8-2012 I. Braasch and J. H. Postlethwait 418 17.3.2 The TGD and Gen(om)e Evolution in Teleosts 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 419 420 421 422 PR OO 389 D 388 TE 387 EC 385 386 CO RR 384 F 417 basal relationships within teleosts themselves based on morphological and molecular data. The sequencing of four hox clusters from the bichir (Polypterus senegalus), a representative of the most basally branching extant lineage of rayfin fish, the polypteriforms, indicated that the GD did not occur in an ancestor of all rayfins (Chiu et al. 2004; Raincrow et al. 2011). Subsequently, several studies cloned and sequenced homeobox genes and several other nuclear markers from different species representing major phylogenetic lineages of rayfins (Hoegg et al. 2004; Crow et al. 2006; Mulley et al. 2006; Hurley et al. 2007). These studies did not find any evidence for the presence of paralogs derived from the TGD in the genomes of extant nonteleosts, i.e., Lepisosteiformes (gars), Amiiformes (bowfin), and Acipenseriformes (sturgeons and paddlefish, which have, however, experienced their own lineage-specific polyploidizations; see above). In contrast, paralogs were found in all teleost lineages analyzed, including representatives of the early branching teleost lineages, i.e., Osteoglossomorpha (bonytongues and mooneyes), Elopomorpha (tarpons and eels), and Clupeomorpha (herrings and shads) (see Fig. 17.1). Finally, Amores et al. (2011) recently presented the first genome-wide synteny comparison between a non-teleost and teleost rayfins through the generation of a high-density genetic map for the spotted gar (L. oculatus). A clear pattern of double conserved synteny became apparent between the gar and teleost genomes showing that the gar lineage diverged from the teleost lineage before the GD shared among teleosts (Amores et al. 2011). To summarize, the GD initially detected in zebrafish, medaka, and pufferfish was an event shared by all teleosts but not by non-teleost rayfin fish, and it occurred (depending on the use of different molecular clocks) around 226–350 million years ago (Vandepoele et al. 2004; Christoffels et al. 2004; Hoegg et al. 2004; Hurley et al. 2007) (Fig. 17.1). It thus seems better to call this tetraploidization event the TGD (The often-used term Fish-Specific Genome Duplication, or FSGD, is ambiguous due to the many GD events in multiple fish lineages, such as salmonids, sturgeons, and cyprinids, and the paraphyletic term ‘‘fish’’.). Reviewing morphological characters defining the teleost lineage, de Pinna (1996) articulated: Intriguingly, molecular data have yet to provide consistent support for teleostean monophyly. Fifteen years later, apparently it is the TGD that provides the best synapomorphy, a molecular character supporting the monophyly of teleosts and probably the best we will ever obtain from extant molecules. 383 UN Editor Proof 352 Book ISBN: 978-3-642-31441-4 Page: 352/383 17.3.2.1 The TGD and Conserved Syntenies A pair of genes that are syntenic are on the same chromosome in one species. Conserved syntenies are situations in which a pair of genes that are syntenic in one species have orthologs that are syntenic in another species. People often use the Layout: T1 Standard SC Chapter No.: 17 Polyploidy in Fish and the Teleost Genome Duplication (b) 1 hbbe1 Dre Chromosomes 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 Hsa16 cdh1 hbbe2 si:dkey-30c15.12 cen HBA 0Mb 10Mb 20Mb 30Mb CDH1 40Mb 50Mb 60Mb 70Mb 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 Hsa17 hoxba hoxbb slc43a2a SLC43A2 10Mb (d) XY cen 6Mb 12Mb 18Mb 24Mb hoxbb 30Mb cen TRAF4 20Mb 30Mb 36Mb Dre12 Orthologs D Hsa Chromosomes HOXB 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 Dre3 TE 0Mb slc16a5b HOXB 40Mb 50Mb SLC16A5 60Mb 70Mb Hsa17 Orthologs (c) X Y 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 Dre12 traf4a traf4b 0Mb 80Mb slc16a5a slc43a2b Hsa16 Orthologs Hsa Chromosomes 353 F (a) Dre Chromosomes Book ISBN: 978-3-642-31441-4 Page: 353/383 PR OO Editor Proof 17 Book ID: 272454_1_En Date: 16-8-2012 42Mb 0Mb HOXB cen hoxba 10Mb 20Mb 30Mb 40Mb 50Mb 60Mb Dre3 Orthologs 424 425 426 427 428 429 430 431 432 433 434 435 term ‘syntenic’ erroneously to mean ‘conserved synteny’. Conserved syntenies give strong evidence for the TGD. The Synteny Database (http://teleost.cs.uoregon.edu/ synteny_db/) uses BLAST scores to identify groups of paralogous genes in a genome rather than a single pair of genes showing the ‘‘best hit’’; the algorithm then anchors paralogy groups to an ortholog in another genome, and finally plots orthologs and paralogs along chromosomes (Catchen et al. 2009, 2011). Results show that segments of human chromosomes generally have orthologs and coorthologs in paralogons on two zebrafish chromosomes. For example, the horizontal axis of Fig. 17.2 displays genes ordered along human chromosome 16 (Hsa16) and on the vertical axis shows the zebrafish orthologs and co-orthologs of each Hsa16 gene plotted on the appropriate linkage groups directly above the human gene. The short arm of Hsa16 (Hsa16p) has orthologs and co-orthologs mainly on zebrafish chromosomes Dre3 and Dre12, including duplicated hemoglobin loci, and the long UN 423 CO RR EC Fig. 17.2 Conserved syntenies are as predicted by GD. a Genes on human chromosome 16 are aligned along the horizontal axis and their zebrafish orthologs and co-orthologs are marked directly above each human gene on the zebrafish chromosome on which it resides. Results reveal duplicate ohnologons on Dre3 and Dre12 for Hsa16p and Dre7 and Dre18 for Hsa16q. A gray circle marks the location of the centromere on Hsa16. b Zebrafish ohnologons for Hsa17 are mainly on Dre3 and Dre12. c Human orthologs of Dre12 genes occupy single portions of human chromosomes, mainly Hsa10 and Hsa17. d Human orthologs of Dre12 are widely scattered on mostly Hsa16 and Hsa17. Individual genes are located on each plot Layout: T1 Standard SC Chapter No.: 17 Book ID: 272454_1_En Date: 16-8-2012 I. Braasch and J. H. Postlethwait 456 17.3.2.2 Doubled Conserved Synteny 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 PR OO 439 D 438 TE 437 F 455 arm (Hsa16q) has orthologs and co-orthologs on Dre7 and Dre18, including co-orthologs of CDH1 and many other genes. Other zebrafish chromosomes also have strings of genes that appear on the plot, including Dre24 and Dre25, but most of these genes are more distantly related paralogs resulting from the VGDs. The content of other human chromosomes appears to have involved more rearrangements than Hsa16 since the divergence of the human and zebrafish lineages, either in the rayfin or the lobefin lineage or both. For example, in zebrafish, the genetic content now on Hsa17 occupies duplicates of at least seven distinct large chromosome segments, mostly on Dre3 and Dre12 (Fig. 17.2b), which indicates several major translocations with respect to human, some of which occurred before the TGD because they are shared by both copies, and some after the TGD, because they differ between the two paralogons (or ohnologons, paralogons arising from GD). Comparisons in the other direction, zebrafish to human, generally display a one-to-one relationship with a substantial number of translocations and inversions (Fig. 17.2c). For example, about half of the human orthologs of Dre12 genes are located on Hsa17, and about half on Hsa10, each with several rearrangements; minor numbers are located on four other zebrafish chromosomes. Other chromosomes, like Dre3, appear to have experienced an even greater number of chromosome rearrangements (Fig. 17.2d). 436 467 17.3.2.3 An Example: The Hox Clusters 459 460 461 462 463 464 465 468 469 470 471 472 473 CO RR 458 EC 466 The most frequent fate of a pair of ohnologs after WGD is nonfunctionalization (Lynch and Conery 2000); hence, ohnologons often experience reciprocal gene loss, like the regions containing two zebrafish co-orthologs of SLC16A 5 on Dre3 and Dre12 (Figs. 17.2b and 17.3). Of the string of four genes on Hsa17 including human SLC16A 5 , different sets are missing from each duplicated zebrafish segment, so that the content of both zebrafish ohnologons must be summed to get the orthologous human gene content in this region (Fig. 17.3). In the extreme, all WGD duplicates will have resolved to singletons, and so conserved syntenies will be maintained but with no co-orthologs to anchor the segments in a doubled conserved synteny. 457 Circle plots can display conserved paralogs on a genome-wide scale. For example, the Synteny Database shows that paralogs of genes within 10 Mb of the four human HOX clusters preferentially occupy regions near each of the other clusters (Fig. 17.4a) (For nomenclature, see https://wiki.zfin.org/display/general/ZFIN+ Zebrafish+Nomenclature+Guidelines.). Note stronger links between the pairs HOXA/HOXB and HOXC/HOXD, which may reflect the order of their origin, with UN Editor Proof 354 Book ISBN: 978-3-642-31441-4 Page: 354/383 Layout: T1 Standard SC Chapter No.: 17 17 Book ID: 272454_1_En Date: 16-8-2012 Polyploidy in Fish and the Teleost Genome Duplication Fig. 17.3 Doubled conserved synteny. The region at about 73 Mb on Hsa17 including SLC16A5 shares syntenies with small regions of Dre3 and Dre12. Zebrafish has two coorthologs of SLC16A5 (boxed), but some of its neighbors were reciprocally lost from one or the other ohnologon. In addition, an inversion on Dre12 reversed gene orders and caused other genes to disrupt the ancestral group. This pattern of doubled conserved synteny with disrupting inversions is quite common 355 LOC794371 Dre3 F atp5h PR OO ARMC7 SLC16A5 KCTD2 Hsa17 ATP5H Dre12 zgc:153947 LOC565514 zgc:153278 TE D Editor Proof Book ISBN: 978-3-642-31441-4 Page: 355/383 486 17.3.2.4 Symmetry of Gene Loss Between Members of Ohnologons 478 479 480 481 482 483 484 487 488 489 490 491 492 493 494 CO RR 476 477 Immediately after GD, both ohnologons have identical gene compositions. As nonfunctionalization events occur, ohnologons begin to diverge from each other. Gene loss events raise the question: Do both members of a pair of ohnologons tend to lose genes at the same rate? Or do genes tend to disappear more frequently from one duplicated chromosome segment than the other? According to one hypothesis, gene losses would be purely stochastic. According to an alternative hypothesis, expression of a number of genes in the same neighborhood might be regulated by a central element, for example, an element that helped to regulate local chromosome UN 475 EC 485 VGD1 giving rise to HOXA/B and HOXC/D clusters (Amores et al. 1998) and VGD2 providing all four clusters. A circle plot of orthologs and paralogs of genes surrounding zebrafish hox clusters (Fig. 17.4b) shows that zebrafish has eight hox cluster-containing ohnologons even though only seven of them contain proteincoding hox genes (Amores et al. 1998). While zebrafish has two copies of the hoxa, hoxb, and hoxc clusters, it has a single hoxd cluster that contains proteincoding genes. The hoxdb cluster, while lacking protein-coding genes, retains a microRNA-10 paralog, which is embedded in Hox clusters from flies to fish to humans (Woltering and Durston 2006), as well as paralogs of many surrounding genes easily detected in the plot (Fig. 17.4b). Likewise, stickleback lacks a second hoxc cluster, but the location where it ‘should’ go is apparent from the circle plot (Fig. 17.4c). 474 Layout: T1 Standard SC Chapter No.: 17 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 356/383 Editor Proof 356 (a) I. Braasch and J. H. Postlethwait HOXC (b) HOXB (c) hoxaa hoxda hoxab hoxca hoxca (hoxcb) hoxba hoxcb hoxba hoxda PR OO hoxdb hoxbb F hoxbb HOXA hoxab hoxaa HOXD hoxdb Fig. 17.4 Hox cluster ohnologons. a Paralogs within 10 Mb of human HOX clusters mapped on the human genome. b Paralogs within 5 Mb of zebrafish hox clusters mapped on the zebrafish genome. Although only a microRNA gene remains in the hoxdb cluster, the paralogy of this segment and its relationship to hoxda is evident. c Paralogs within 3 Mb of stickleback hox clusters mapped on the stickleback genome 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 D 500 TE 498 499 EC 497 structure, and loss of that element would then lead to rapid loss of the several genes that element controls and hence asymmetric gene loss. Analysis of a few individual chromosome regions suggests that asymmetries often occur in teleosts (Canestro et al. 2009; Braasch et al. 2006; Siegel et al. 2007), but this question has not been explored sufficiently on a genome-wide scale. To study this question, we used the Synteny Database to plot the zebrafish and stickleback chromosomes on which orthologs, co-orthologs, and paralogs of each human gene lie for all human chromosomes. Figure 17.5 displays results for two human chromosomes (Hsa7 and Hsa18) for both zebrafish and stickleback. Figure 17.5a shows genes on Hsa7 as gray dots along the bottom with the zebrafish orthologs and paralogs directly above; note that gene order in these plots reflects gene arrangements in the human genome, not the fish genome. Results revealed several principles (Fig. 17.5a–d). (1) Many regions of ohnology are clear, as marked by rectangles. For example, the region from 18 to 38 Mb on Hsa7 has two clear ohnologons in both zebrafish and stickleback, with orthologous chromosome segments marked by color (Fig. 17.5a, b). (2) In contrast, many regions have ambiguous ohnologons, such as the region from the left telomere of Hsa7 to about 18 Mb for both zebrafish and stickleback. (3) Some ohnolog pairs are clear in one species but ambiguous in the other; for example, from 38 to 64 Mb of Hsa7, where three or four zebrafish chromosomes have strings of ‘hits’ compared to two clear regions in stickleback. Regions like these (also note for stickleback region 101–142 Mb the ‘ghost’ paralogons on groupXII and groupXVIII) are often due to paralogs arising in the VGD1 and VGD2 events rather than the TGD. (4) Most chromosome translocations occurred before the divergence of the zebrafish and stickleback lineages. Evidence supporting this conclusion is that, with respect to the human chromosome, orthologous blocks tend to have the same termini in both fish species. And (5), a minority of translocations occurred in one or the other fish lineage after divergence; for example, the region from 77 to 88 Mb and the region from 101 to 142 Mb are syntenic (on the same chromosome) in human (Hsa7) and in both ohnologons in stickleback CO RR 496 UN 495 Layout: T1 Standard SC Chapter No.: 17 Book ISBN: 978-3-642-31441-4 Page: 357/383 Polyploidy in Fish and the Teleost Genome Duplication (a) 357 2 4 6 Dre Chromosomes 8 10 12 F 14 16 18 PR OO 20 22 24 Hsa7 (b) groupII Gac Chromosomes groupIV groupVI groupVIII groupX groupXII groupXIV groupXVI groupXVIII groupXX Hsa7 20Mb 40Mb 60Mb 80Mb 100Mb D 0Mb 120Mb 140Mb (c) TE Hsa7 Orthologs 2 4 6 Dre Chromosomes 8 10 12 EC 14 16 18 20 22 24 (d) CO RR Hsa18 groupII groupIV Gac Chromosomes groupVI groupVIII groupX groupXII groupXIV groupXVI groupXVIII groupXX Hsa18 0Mb UN Editor Proof 17 Book ID: 272454_1_En Date: 16-8-2012 10Mb 20Mb 30Mb 40Mb 50Mb 60Mb 70Mb Hsa18 Orthologs Fig. 17.5 Ohnologons in zebrafish (a, c) and stickleback (b, d) for Hsa7 (a, b) and Hsa18 (c, d). These plots show doubled conserved synteny as well as the pattern of chromosome translocations in various lineages Layout: T1 Standard SC Chapter No.: 17 Book ID: 272454_1_En Date: 16-8-2012 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 F 531 PR OO 530 D 528 529 TE 527 (groupIV and groupXX)—thus representing the ancestral condition—but for zebrafish, these two regions are syntenic for only one of the copies (Dre4) but not the other (Dre18 and Dre25). These results demonstrate a translocation in the zebrafish lineage after it diverged from the stickleback lineage (Fig. 17.5a, b). Note that the effects of inversions in the fish lineage are invisible in these plots because gene orders in the fish genomes are displayed according to their order in the human genome. Having identified ohnologons, we can now pose the question of gene loss asymmetries. Over the human genome, we identified 49 unambiguous ohnologon pairs in zebrafish or stickleback or both, and counted the number of fish orthologs and paralogs of human genes occupying each segment. For zebrafish, 7 of 39 (17.9 %), and for stickleback, 13 of 44 (29.5 %) ohnologon pairs had significantly different number of genes [v2 test, 1 df, p \ 0.01; 10/39 (25.6 %), which increased to 18/44 (40.9 %) if p \ 0.05]. We conclude: (1) that most ohnologons appear to lose genes approximately randomly between the two duplicated copies, but (2) that about 20–40 % of ohnologons in these fish species have resolved TGD ohnologs asymmetrically between the two duplicated chromosome segments, with one member of the pair retaining significantly more genes than the other. (3) Of 13 segments with significance at the 0.01 level in one or the other fish, 6 are significant at 0.01 or 0.05 level in both species, which diverged deeply in teleost phylogeny (Fig. 17.1), and for all of those pairs, chromosome segments retaining the most genes in zebrafish are orthologs of those retaining the most in stickleback. This result is consistent with the idea that the basis for most asymmetric gene losses occurred before the divergence of zebrafish and stickleback lineages. A question that remains is the mechanism that led to asymmetric resolution of gene duplicates after the TGD. One hypothesis is that, if a long-range regulatory function that controls the expression of many genes in a neighborhood disappears from one ohnologon, then nonfunctionalization of the genes that the regulatory function controls can follow without penalty given the maintenance of that longrange function in the sister ohnologon. That hypothesis is yet to be seriously investigated. An example of resolution asymmetries is shown in Fig. 17.6 for Hsa15 in zebrafish and stickleback, along with chromosome diagrams comparing genes in their order along a segment of Hsa15 and their fish orthologs and coorthologs in zebrafish (52 ortholog pairs in one ohnologon vs. 4 ortholog pairs in the other) and stickleback (54 ortholog pairs vs. 16 ortholog pairs) (Fig. 17.6c, d). A related question is whether ohnologs residing in ohnologons that suffer asymmetric rates of gene loss also experience asymmetric rates of gene evolution. Asymmetric rates of molecular evolution have been found for a substantial number of TGD paralogs (Van de Peer et al. 2001; Steinke et al. 2006; Brunet et al. 2006), sometimes correlating with asymmetric loss of genes from ohnologons (Braasch et al. 2006; Siegel et al. 2007). Available evidence from other WGDs, however, suggests that gene evolutionary rate asymmetry does not strongly depend on the conservation of syntenies (Bu et al. 2011). EC 526 CO RR 525 I. Braasch and J. H. Postlethwait UN Editor Proof 358 Book ISBN: 978-3-642-31441-4 Page: 358/383 Layout: T1 Standard SC Chapter No.: 17 Editor Proof 17 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 359/383 Polyploidy in Fish and the Teleost Genome Duplication (a) 359 2 4 6 Dre Chromosomes 8 10 12 F 14 16 18 PR OO 20 22 24 Hsa15 (b) groupII Gac Chromosomes groupIV groupVI groupVIII groupX groupXII groupXIV groupXVI groupXVIII groupXX Hsa15 0Mb 20Mb 40Mb (c) PAK6(1of2) sb:cb730 srp14 zgc:136872 FMN1 GJD2 MEIS2 SRP14 C15orf57 C15orf24 ACTC1 PLCB2 AQR C15orf52 ATPBD4 EXD1 EHD4 LRRC57 RPAP1 PLA2G4F CDAN1 VPS39 TTBK2 ZFP106 UBR1 spint1b cx35 zgc:86709 meis2.1 SCG5 ENSGACG00000007697 ENSGACG00000007661 ENSGACG00000005610 GJD2(2of2) CASC5 GacgroupXVIII 54 pairs im:7147183 80Mb GREM1 SCG5CHRM5 ARHGAP11A RYR3 Hsa15 16 pairs GacgroupXV FMN1 AC012652.1 EIF2AK4 CASC5 CHAC1 MAPKBP1 FAM98B RPUSD2 SPINT1 TYRO3 SPRED1 BUB1B DISP2 ZFYVE19 NUSAP1 JMJD7 C15orf41 THBS1 PAK6 IVD GCHFR DLL4 RTF1 MGA GJD2 MEIS2 FSIP1 PLCB2 DNAJC17 C15orf57 AVEN ACTC1 RASGRP1 GPR176 C15orf24 AQR SRP14 C15orf29 ATPBD4 MAPKBP1(2of2) ENSGACG00000003916 TYRO3 LRRC57 JMJD7 100Mb NDUFAF1 RPAP1 CHAC1(1of2) DLL4(2of2) LOC100334937 vps39 LOC100331590 si:ch211-283f6.8 zgc:63611 zgc:154068 zgc:110758 meis2.2 MEIS2(2of2) fam98b spred1 zgc:154061 CAPN3(2of2) VPS39 CDAN1 TMEM87A TTBK2 ZFP106 UBR1 LRRC57 TMEM87A ZFP106 capn3 LOC100331516 ubr1 ENSGACG00000009824 FMN1(2of2) CHRM5(2of2) THBS1(1of2) RYR3(2of2) STARD9 HAUS2 SNAP23 CAPN3 GANC TMEM62 CAPN3(1of2) VPS39 GANC ENSGACG00000005683 ENSDARG00000091548 LOC565314 zgc:77419 PAK6(1of2) FMN1(1of2) GCHFR BUB1B FSIP1 EIF2AK4 SRP14 RASGRP1 CHRM5(1of2) DNAJC17 ZFYVE19 THBS1(2of2) SPINT1(1of2) RYR3(1of2) PLCB2 GREM1 ENSGACG00000006109 C15orf57 DISP2 UBR1 RPUSD2 IVD RPAP1 FAM98B SPRED1 C15orf41 GJD2(1of2) MEIS2 ENSGACG00000010421 AQR ATPBD4 TMEM62 SNAP23 ENSGACG00000011127 CDAN1 TTBK2(1of2) RTF1 NUSAP1 ENSGACG00000012183 NDUFAF1 EC HAUS2 C15orf24 C15orf29 GPR176 AVEN zgc:123218 zgc:56072 LOC565088 ivd LOC564712 RPAP1 si:dkey-170l10.1 TE 4 pairs Dre20 NP_001185682.1 disp2 itpka MAPKBP1 jmjd7 mga tyro3 JMJD7 STARD9 RPUSD2 MAPKBP1 HAUS2 FAM98B PAK6 BAHD1 MGA SNAP23 SPRED1 BUB1B IVD SPINT1 TYRO3 CAPN3 C15orf41 THBS1 DISP2 ZFYVE19 ITPKA GANC TMEM62 CHRM5 GREM1 ARHGAP11A Hsa15 LOC565386 ENSDARG00000056535 lrrc57 ENSDARG00000079066 actc1b PLCB2 chrm5a spint1a LOC100005728 zgc:92683 arhgap11a LOC796966 52 pairs D ENSDARG00000088538 THBS1(2of2) zgc:195050 Dre17 (d) 60Mb Hsa15 Orthologs 569 570 571 572 573 574 575 576 577 578 579 580 581 17.3.2.5 Reconstruction of the Ancestral Karyotype and Teleost Chromosome Rearrangements As we discussed earlier, the ancestral haploid teleost (post-TGD) genome was inferred to contain 24 chromosomes based on karyotype data (Ohno et al. 1968; Mank and Avise 2006b). Using several teleost genetic maps as well as the genome assemblies of Tetraodon and medaka, several attempts have been made to reconstruct the ancestral pre-TGD protokaryotype (Naruse et al. 2004; Jaillon et al. 2004; Woods et al. 2005; Kohn et al. 2006; Kasahara et al. 2007; Nakatani et al. 2007). All studies largely agree that the teleost karyotype is derived from 11 to 12 pre-TGD protochromosomes, which is in agreement with the inferred number of 24 post-TGD chromosomes (Mank and Avise 2006b). There are, however, some differences in the assignment of chromosomal blocks from different extant teleosts to the protochromosomes (Woods et al. 2000; Kasahara et al. 2007). The reconstruction of the ancestral teleost karyotype has so far relied on using tetrapod UN 568 CO RR Fig. 17.6 Asymmetrically resolved ohnologons for Hsa15 in zebrafish (a) and stickleback (b). Chromosome contents displayed with proper gene orders for zebrafish duplicates (c, 52 vs. 4 ortholog pairs) and stickleback (d, 54 vs. 16 pairs) Layout: T1 Standard SC Chapter No.: 17 Book ID: 272454_1_En Date: 16-8-2012 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 F 587 PR OO 586 D 585 TE 584 genomes as outgroups, but the genome sequence of gar or bowfin, rayfin fish that can serve as outgroup to the TGD, will help to provide a more detailed picture of the ancestral teleost genome. Nevertheless, the current state of karyotype reconstruction is helpful to compare and cross-refer different types of data on duplicated genes in teleosts. As an example, for zebrafish and stickleback we downloaded all pairs of predicted duplicated genes and their genomic locations from the EnsemblCompara GeneTree set (Vilella et al. 2009) and parsed them for the ‘Clupeocephala’ duplication node, which indicates a duplication on the branch leading to teleosts, after divergence from tetrapods, but before the divergence of teleosts. We then plotted the location of these pairs as Oxford grids for zebrafish and stickleback genomes (Fig. 17.7) and overlaid them with the 12 ancestral pre-TGD protochromosomes a–m inferred from the analysis of the medaka genome (Kasahara et al. 2007). This analysis clearly shows that paralogs are nonrandomly distributed over the genome; for example, of the 228 genes in the analysis that reside on zebrafish chromosome Dre3 and have TGD duplicates, the vast majority, 105 genes, are on Dre12 and a sizable minority on Dre1, and of the 167 genes on stickleback groupXIV that are duplicated and assigned to chromosomes, 152 are on groupXIII. These results show that the vast majority of ‘Clupeocephala-duplicated’ genes obtained from the tree-only method used by Ensembl supports a TGD origin of the paralogs and is in line with conserved synteny data and ancestral karyotype reconstructions. Teleost genomes are rearranged with respect to tetrapod genomes, which may be due to chromosome rearrangements that were facilitated by the TGD, for example through illegitimate recombination between homeologous (paralogous) chromosomes. This hypothesis, however, is controversial (Comai 2005; Semon and Wolfe 2007a; Hufton et al. 2008), because chromosome restructuring may have likewise occurred before the TGD, on the long branch separating teleosts and tetrapods after the divergence of rayfin fish from lobefin fish. Using the genetic map of the spotted gar, which contains nearly 1,000 coding markers, Amores et al. (2011) recently showed unexpectedly high conservation of synteny between human and gar when compared to zebrafish and stickleback. This suggests that chromosome rearrangements and the loss of ancestral syntenies accelerated after the TGD, but before the divergence of stickleback from zebrafish, and supports the hypothesis that WGD can facilitate syntenic rearrangements. The zebrafish genome appears to be more rearranged than the genome of percomorph teleosts (stickleback, medaka, pufferfish). Chromosomal blocks of the inferred ancestral pre-TGD karyotype (Kasahara et al. 2007; Nakatani et al. 2007) are generally distributed over many more chromosomes in zebrafish than in percomorphs. For example, as shown in Fig. 17.7, chromosomal blocks of the ancestral pre-TGD chromosome m are found on four stickleback chromosomes (GacI, GacIII, GacVIII, GacXXI), but on six chromosomes in zebrafish (Dre2, Dre6, Dre8, Dre11, Dre22, Dre24); pieces of ancestral chromosome j are found on two chromosomes in stickleback (GacII, GacXIX), but on three in zebrafish (Dre7, Dre18, Dre25), and so forth. A genome sequence of another representative of the ostariophysians, which, after percomorphs, is the next most species-rich group of EC 583 CO RR 582 I. Braasch and J. H. Postlethwait UN Editor Proof 360 Book ISBN: 978-3-642-31441-4 Page: 360/383 Layout: T1 Standard SC Chapter No.: 17 Polyploidy in Fish and the Teleost Genome Duplication chr 1 2 3 4 5 6 7 8 9 10 11 12 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 Un 8 1 51 0 0 2 18 1 49 0 0 0 25 47 0 0 7 0 1 4 2 1 15 0 1 1 10 0 0 0 30 22 17 0 0 24 9 0 0 20 0 0 6 0 25 0 48 7 86 0 3 8 0 0 29 0 0 0 0 4 105 0 0 0 1 0 0 1 0 0 7 0 17 1 4 4 0 2 0 2 0 1 0 0 0 0 0 7 3 26 0 1 0 0 0 0 44 1 48 1 7 14 1 48 3 3 0 18 36 1 0 3 0 0 36 0 2 7 0 3 12 0 7 67 0 22 1 4 0 0 0 0 0 0 0 0 2 60 1 0 2 8 0 1 13 1 0 1 2 0 0 0 41 0 0 18 0 0 0 70 2 16 0 45 33 0 1 0 0 0 0 0 0 0 25 13 14 0 0 2 12 1 8 2 0 2 1 1 0 0 1 0 0 17 0 0 0 2 2 0 0 0 14 54 0 0 2 0 0 1 0 0 4 2 1 2 0 6 1 0 0 1 0 2 0 0 0 58 0 1 3 I II III IV V VI VII VIII IX X XI 10 1 1 49 0 35 0 0 0 0 2 7 0 1 0 77 1 0 6 6 63 0 0 1 0 125 0 3 0 1 0 0 0 1 2 0 6 13 14 15 16 6 49 1 0 0 27 0 3 0 0 0 0 0 1 3 4 0 0 0 0 0 1 1 1 8 0 18 0 0 36 8 0 0 1 0 0 3 0 2 1 8 0 0 2 11 0 36 0 0 0 0 0 14 3 0 0 0 174 4 4 132 0 1 0 2 3 2 1 0 0 1 2 2 XII XIII XV XVI XVII XVIII XIX 2 8 (b) 2 0 0 48 0 1 0 1 0 2 9 0 2 0 0 0 9 32 1 28 0 0 0 1 1 2 0 1 0 0 0 7 19 8 1 0 1 0 0 0 17 0 0 0 0 0 1 2 17 18 19 20 21 22 23 24 XIV 2 0 0 8 1 0 0 24 1 2 1 0 0 1 0 2 1 4 0 0 0 1 0 4 10 0 0 1 0 3 9 7 0 0 2 XX XXI Un 28 0 0 2 5 1 5 0 0 6 107 0 0 0 0 0 0 0 0 14 0 0 0 0 0 4 0 0 1 147 1 0 22 2 4 1 1 0 0 0 0 0 0 0 0 30 25 Un 48 0 6 0 152 0 0 0 0 103 10 0 0 1 0 0 0 0 0 70 12 0 0 0 0 1 0 0 0 14 2 3 0 97 0 1 2 10 0 0 0 0 0 0 14 8 0 0 0 0 40 4 8 8 ancestral chromosomes: D 6 1 2 16 0 4 0 2 1 0 0 0 0 0 1 123 0 8 3 2 150 1 2 0 1 0 0 1 0 0 0 0 0 0 0 0 0 1 0 3 1 0 74 20 11 0 0 0 4 11 0 3 126 1 2 2 0 0 0 56 9 1 33 TE I II III IV V VI VII VIII IX X XI XII XIII XIV XV XVI XVII XVIII XIX XX XXI Un EC group 361 F (a) Book ISBN: 978-3-642-31441-4 Page: 361/383 PR OO Editor Proof 17 Book ID: 272454_1_En Date: 16-8-2012 2 0 14 0 8 8 0 0 19 a b c d e f g h i j k l m 0 0 23 2 7 51 628 629 630 631 632 633 teleosts and includes zebrafish and carps (Nelson 2006), will be necessary to clarify whether this enhanced rearrangement rate is specific to the zebrafish genome or a more general phenomenon on this major branch of teleost diversity. UN 627 CO RR Fig. 17.7 Oxford grid of duplicated genes in zebrafish (a) and stickleback (b). Chromosomal location of paralogs determined by EnsemblCompara GeneTrees (Vilella et al. 2009) to be generated at the ‘Clupeocephala’ node. Each box in the grid shows the number of duplicated genes shared between the indicated chromosomes. The distribution of duplicated genes is nonrandom as expected from the TGD. Most clusters of duplicated genes can be traced back to the ancestral protochromosome (Kasahara et al. 2007), from which they are derived (see color code) 17.3.2.6 Global Patterns of Gene Retention and Loss After the TGD The TGD obviously had a major impact on the evolution of genome structure in teleosts. We still do not know, however, how many paralogs from the TGD have escaped nonfunctionalization and have been retained in duplicate in extant Layout: T1 Standard SC Chapter No.: 17 Book ID: 272454_1_En Date: 16-8-2012 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 F 640 PR OO 639 D 637 638 TE 636 teleosts. Genome-wide estimates range between 12 and 24 % depending on the approach and dataset (Postlethwait et al. 2000; Jaillon et al. 2004; Woods et al. 2005; Brunet et al. 2006; Kassahn et al. 2009) and do not differ significantly among sequenced teleost genomes (Kassahn et al. 2009). Between *1,800 and *2,200 TGD-derived pairs of gene duplicates can be identified in the five sequenced teleost genomes (I. Braasch, ‘‘unpublished data’’). Loss of paralogs after the TGD appears to be time-dependent, with most losses occurring soon after the GD event (Braasch et al. 2009a; Sato et al. 2009). A similar trend can be observed when comparing paralog loss after different GD events in teleosts, from the CaGD *11–21 million years ago (mya) with a paralog retention rate of *60 % (David et al. 2003), over the sucker genome duplication (SuGD) and SaGD, which occurred *50 mya and 25–100 mya, respectively, and which both have estimated paralog retention rates of *50 % (Ferris 1984; Allendorf and Thorgaard 1984), to the TGD that happened *226–350 mya and that has an estimated paralog retention of *12–24 % (see above). Retained paralogs from the TGD are enriched for transcription factors, developmental genes, and cell communication proteins (Brunet et al. 2006; Kassahn et al. 2009), similar to results for the earlier rounds of WGD in vertebrates (see Putnam et al. 2008; Huminiecki and Heldin 2010; Chap. 16, this volume). Some gene classes appear to have a high rate of TGD paralog retention, such as genes encoding components of pigmentation pathways (Braasch et al. 2007, 2009a) and of signaling pathways involved in long-term potentiation of synaptic transmission, and olfactory and taste transduction (Sato et al. 2009). However, a more detailed analysis of gene family evolution after the TGD awaits the genome assembly of a rayfin outgroup to the TGD, the spotted gar (see below). EC 635 17.3.3 Functional Divergence of Gene Duplicates: Examples in Fish CO RR 634 I. Braasch and J. H. Postlethwait After a polyploidization event, ohnolog pairs are initially identical, both in regulatory elements and in protein-coding areas, so they are fully redundant. Ohnologs, however, soon begin to accumulate mutations that become fixed in populations and distinguish the a copy from the b copy; these mutations can occur in both regulatory and protein coding regions. Some mutations can cause one copy of a duplicated pair of genes to become nonfunctional, either because they inactivate the protein by premature stop mutation or destructive amino acid substitution, or because they eliminate gene expression. Nonfunctionalization is the most frequent fate of gene duplicates (Lynch and Conery 2000). For genes that are not dosage-sensitive, nonfunctionalization of one copy is likely to carry little selective penalty (and may provide in some cases an advantage), reflecting the fact that most mutations are recessive in a diploid. UN Editor Proof 362 Book ISBN: 978-3-642-31441-4 Page: 362/383 Layout: T1 Standard SC Chapter No.: 17 Book ISBN: 978-3-642-31441-4 Page: 363/383 Polyploidy in Fish and the Teleost Genome Duplication 363 684 17.3.3.1 Subfunctionalization 680 681 682 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 PR OO 679 The classic example that led to the idea of subfunctionalization—duplicate gene preservation by the reciprocal loss of ancestral gene subfunctions—is the zebrafish pair of engrailed-1 co-orthologs (Force et al. 1999). The mouse En1 gene is expressed in both the pectoral appendage bud and in a specific set of hindbrain interneurons (Joyner and Martin 1987; Davis et al. 1991; Gardner and Barald 1992). In contrast, the zebrafish co-orthologs eng1a and eng1b partitioned these two expression domains between them, with eng1a expressed in the pectoral appendage bud and, reciprocally, eng1b expressed in the hindbrain interneurons, a case of spatial subfunctionalization (Force et al. 1999). The reciprocal loss of ancestral regulatory domains like this, which can occur purely by neutral evolution, makes both copies essential, with eng1a essential to provide Engrailed function to the pectoral fin bud and eng1b necessary to deliver Engrailed protein to interneurons. Many genes evolved like eng1 genes to show spatial subfunctionalization, but some genes evolved by temporal subfunctionalization, duplicate gene expression in the same tissue but at multiple developmental stages. For example, Sox1 is expressed in tetrapods at several developmental stages in the lens, but after the TGD, sox1a came to be expressed several hours earlier in the lens than is sox1b (Okuda et al. 2006). Subfunctionalization is pervasive: a global analysis of TGD ohnologs showed that nearly all that were examined had diverged in their patterns of spatial and/or temporal expression during embryogenesis and that such regulatory changes were more frequent than changes at the level of protein function (Kassahn et al. 2009). Tetrapod enhancers, which represent a type of regulatory subfunction, and CNEs, many of which may be enhancers, evolved divergent sequences in teleosts by altering a few specific bases that underlie divergent functions, as for medaka and fugu ohnologs derived from Hoxa2 (Tumpel et al. 2006) or zebrafish and stickleback Fgf8 co-orthologs (Canestro et al. 2007); this result would be expected to occur by subfunctionalization. Cases of preserved ohnologs that have no documented differences in embryonic expression domains may display differences at developmental stages or tissues that have not yet been examined. D 678 TE 677 EC 676 CO RR 675 F 683 Besides nonfunctionalization of one member of a duplicate pair, reciprocal changes to both members of the pair cause both ohnologs to become essential. Reciprocal changes can be of two main sorts, either reciprocal loss of ancestral gene subfunctions that occur purely by neutral processes (subfunctionalization), or by mutation in one or both copies that result in both ohnologs being positively selected for new functions (neofunctionalization) (Force et al. 1999). Mutations that lead to the partitioning of subfunctions between ohnologs do not preclude the later origin of novel subfunctions [called subneofunctionalization (He and Zhang 2005; Rastogi and Liberles 2005)] and vice versa. Here, we note a few examples for each of several fates of TGD ohnologs. 674 UN Editor Proof 17 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 17 Book ID: 272454_1_En Date: 16-8-2012 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 F 722 PR OO 721 D 719 720 TE 718 Besides spatial or temporal differences in expression domains, some ohnologs may achieve quantitative subfunctionalization, the accumulation of activityreducing mutations in both ohnologs, so that neither gene by itself can provide sufficient product to achieve normal function (Force et al. 1999; Stoltzfus 1999). Mutations that partially reduce gene function should be relatively common, and if a population fixes a partial loss-of-function allele at one copy, then it is unlikely to fix a null allele in the gene’s ohnolog. Although quantitative subfunctionalization may be common after WGD in yeast (Scannell and Wolfe 2008), few examples exist from the TGD, probably because of the lack of appropriate analyses. We do know that quantitative subfunctionalization occurred in many of the hox gene duplicates after the SaGD (Mungpakdee et al. 2008b). The finding that most pairs of CNEs in teleost ohnologs experience substantial degeneration in element length might also reflect cases of quantitative subfunctionalization (Woolfe and Elgar 2007). The evolutionary significance of quantitative subfunctionalization may be that it preserves both duplicates long enough to increase the opportunity for later mutations that might secondarily increase gene activity, as appears to have happened for the medaka copy of aldh1a2, a gene originating in VGD2 (Canestro et al. 2009). Protein structure evolution: Individual tetrapod genes can encode proteins with multiple functions. In some cases, a single protein sequence performs several functions; in other cases, a single gene encodes different protein sequences, often with different functions, due to alternative splicing or start sites. Amino acid substitutions can cause proteins encoded by a pair of ohnologs to diverge in sequence in functionally important ways. For example, the chemokine Cxcl12 guides gastrulation and the migration of primordial germ cells (PGCs) in several vertebrates (Nair and Schilling 2008; Richardson and Lehmann 2010). Zebrafish has co-orthologs of Cxcl12 that appear to have arisen in the TGD (Doitsidou et al. 2002). Cxcl12 functions by binding Cxcr4 receptor, which is also duplicated in zebrafish. Cxcl12a, but not Cxcl12b, guides PGCs to their target, the gonad, and a single amino acid substitution causes the relative affinity of Cxcl12 ligands to switch from one of the duplicated Cxcr4 receptors to the other, thereby controlling whether the Cxcl12:Cxcr4 ligand:receptor system controls PGC migration or gastrulation (Boldajipour et al. 2011). These experiments are a particularly deeply studied case study supporting the concept of protein subfunctionalization. Alternative splicing can generate different proteins with variant functions. For example, the human STAT1 gene, which mediates interferon signaling, makes two different proteins by alternative splicing: STAT1-alpha, which contains a TAZ2binding domain, and STAT1-beta, which lacks that domain and may act as a negative regulator of STAT1-alpha (Bromberg et al. 1996). The TGD provided teleosts with duplicate stat1 genes, and, due to reciprocal mutations that affect transcript splicing, the zebrafish stat1a gene came to encode a protein that contains the TAZ2-binding domain like the human STAT1-alpha isoform, while stat1b evolved to encode a protein that lacks this domain like STAT1-beta (Song et al. 2011). Thus, due to subfunctionalization of splicing signals, two genes in EC 717 CO RR 716 I. Braasch and J. H. Postlethwait UN Editor Proof 364 Book ISBN: 978-3-642-31441-4 Page: 364/383 Layout: T1 Standard SC Chapter No.: 17 Book ISBN: 978-3-642-31441-4 Page: 365/383 Polyploidy in Fish and the Teleost Genome Duplication 365 772 17.3.3.2 Neofunctionalization 768 769 770 773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792 793 794 795 796 797 798 799 800 801 802 PR OO 767 In the race between nonfunctionalization, subfunctionalization, and neofunctionalization, nonfunctionalization generally rides the fastest horse. Subfunctionalization appears to place second, and neofunctionalization—the preservation of ohnologs by the origin of new functions (Ohno 1970)—shows in third. As with subfunctionalization, neofunctionalization can occur at the level of either gene regulation or protein structure. Preservation by neofunctionalization differs from subfunctionalization in two main ways: first, it involves positive selection rather than neutral events, and second, it involves the origin of gene subfunctions that did not exist in the ancestral gene rather than the partitioning of ancestral subfunctions. Genome-wide studies suggest that the acquisition of novel protein domains might have occurred in a quarter of duplicate pairs after the TGD and happened more frequently in duplicates than in single-copy genes (Kassahn et al. 2009). Although several specific cases of neofunctionalization after the TGD have been suggested (Yao and Ge 2010; Sha et al. 2008; Howarth et al. 2008), conclusions are often difficult due to assumptions about the ancestral state, which is usually based on tetrapods, species that are far removed from the teleost lineage. To distinguish neofunctionalization from subfunctionalization, one must accurately identify the ancestral state. Inferences about the ancestral state can be difficult with WGD but are often easier with tandem duplications, which have led to neofunctionalization of vitellogenin and aquaporin genes in teleost egg hydration (Finn and Kristoffersen 2007; Zapater et al. 2011; Cerda 2009), the evolution of antifreeze proteins by Antarctic teleosts (Deng et al. 2010), and the preservation of fatty acid-binding protein genes (Karanth et al. 2009). Several examples of neofunctionalization after the TGD have been suggested. (1) Phosphoglucose isomerase (PGI) ohnologs from the TGD, after the partitioning of tissue-specific regulatory elements for expression, diverged so that the copy expressed in organs of the body cavity became more negative in charge while the muscle isoform became more positively charged, although the functional significance of these changes is not yet well understood (Sato and Nishida 2007). (2) After the TGD, cyprinid fish, like zebrafish, retained only one copy of the D 766 TE 765 EC 763 764 CO RR 762 F 771 zebrafish encode two protein variants that are encoded by a single gene in human. Another example is the synapsin gene in fugu (Yu et al. 2003). Alternative start sites can also result in alternative exon usage. The human MITF gene controls pigment production both in melanocytes and in the pigmented retina, with different transcription start sites employed by the two different cell types (Udono et al. 2000). Zebrafish and other teleosts have two MITF co-orthologs, with mitfa expressed in melanophores and mitfb in the pigmented retina; correspondingly, evolution of the two ohnologs resulted in mitfa using one transcription start site and mitfb using the other (Lister et al. 2001; Altschmied et al. 2002). Again, two different tissue-specific proteins encoded by one gene in mammals are formed from sister ohnologs in teleosts due to subfunctionalization. 761 UN Editor Proof 17 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 17 Book ID: 272454_1_En Date: 16-8-2012 810 811 812 813 814 815 816 817 818 819 820 821 822 823 824 825 826 827 828 829 830 831 832 833 834 835 836 837 838 839 840 841 842 843 844 F 809 PR OO 807 808 D 806 TE 805 androgen receptor gene (ar), but percomorph fish, like pufferfish and medaka, retained both ohnologs, one of which (arb) experienced substantial sequence alterations that are consistent with neofunctionalization; further, these changes occurred after the divergence of basal teleost lineages, including eels, but before the radiation of percomorphs (Douard et al. 2008). (3) Some TGD duplicates show elevated evolutionary rates, like csf1rb and its neighbor pdgfrbb in cichlids, which could be due either to neofunctionalization or to relaxed selection; it has been suggested that such behavior could be a causal factor in the evolution of the spectacular coloration patterns of some teleost fish (Braasch et al. 2006). To learn whether neofunctionalization is at play in these examples and many others, the duplicates must be compared to each other by function in addition to sequence and compared to the structure and function of the ancestral genes in a rayfin fish that diverged before the TGD. In two cases of neofunctionalization—a hox gene and a sodium channel gene— TGD duplicates have been investigated functionally in a phylogenetic context that leaves little doubt about the functions of the preduplication gene. Zebrafish has coorthologs of Hoxa13 derived from the TGD (Amores et al. 1998), and, as determined by a broad phylogenetic study, these ohnologs began to diverge asymmetrically long after their initial preservation subsequent to the TGD (Crow et al. 2009). Ohnolog knockdown experiments in zebrafish, coupled with rescue experiments, showed that hoxa13a, the faster evolving ohnolog, contains many derived a clade-specific amino acid replacements that are necessary for development of the yolk sac extension (the hind yolk), a derived feature in zebrafish and its relatives (Crow et al. 2009). A great example of neofunctionalization after the TGD occurred in electric fish. The distantly related African elephantfish (Mormyridae; Osteoglossiformes) and South American knifefish (Gymnotideae; Gymnotiformes) independently evolved a muscle-derived electric organ, which facilitates electrical communication and utilizes the voltage-gated sodium channel subunit Scn4a. Teleosts have two scn4a genes from the TGD, both expressed in skeletal muscle in most teleosts; in two lineages of electric fish, however, scn4aa independently lost its expression in skeletal muscle and gained expression in the electric organ (Arnegard et al. 2010). Phylogenetic analyses showed that this neofunctionalization event occurred more than 100 million years after the TGD with a signal of positive selection (Arnegard et al. 2010). Meanwhile, scn4ab maintained the ancestral expression in skeletal muscle. These results for Scn4a and Hoxa13 both show that, long after the preservation of ohnologs by subfunctionalization, one copy can still provide the genetic variation necessary to lead to the innovation of new morphologies and functions. EC 804 CO RR 803 I. Braasch and J. H. Postlethwait 17.3.3.3 Spotted Gar as a Rayfin Outgroup for Functional Studies of TGD Paralogs UN Editor Proof 366 Book ISBN: 978-3-642-31441-4 Page: 366/383 As pointed out above, knowledge of the functions of the preduplication gene is essential for inferring subfunctionalization versus neofunctionalization. Investigators often infer the properties of the pre-TGD gene from the phenotype of well- Layout: T1 Standard SC Chapter No.: 17 851 852 853 854 855 856 857 858 859 860 861 862 863 864 865 866 867 868 869 870 871 872 873 874 875 876 877 878 879 880 881 882 883 884 F 850 PR OO 849 D 848 studied tetrapods, especially mouse. The tetrapod lineage, however, diverged from the teleost lineage about 450 million years ago and involved the evolution of substantial differences in morphologies from ancestral bony fish. In addition, rayfin fish followed an independent lineage for about 200 million years before the TGD; thus, what appears to involve neofunctionalization when comparing teleost co-orthologs to tetrapod genes (Fig. 17.8a) could instead be due to the tetrapod lineage having lost a subfunction originally present in the single-copy gene in the last common ancestor of teleosts and tetrapods (Fig. 17.8b). A third alternative is the reciprocal loss of ancestral VGD ohnologs, which could also mimic neofunctionalization, pointing out the importance of careful distinction between orthologs and ohnologs (Fig. 17.8c). Finally, the novel function thought to have arisen after the TGD might have actually originated in ancestral rayfin fish, pointing out the importance of identifying a fish representing this node (Fig. 17.8d). In addition, distinguishing among the possibilities shown in Fig. 17.8b–d requires examination of an outgroup diverging basal to the diagrammed tree, such as a cartilaginous fish. Recently, the spotted gar L. oculatus has been identified as an unduplicated rayfin outgroup to the TGD (see Sect. 17.3.1.5 and Amores et al. 2011). The spotted gar occupies a lineage of ‘ancient fish’ that includes not only lepisosteids (gars), but also bowfin (Amia calva) and possibly acipenseriforms (sturgeons, e.g. Acipenser transmontanus, and paddlefish, e.g. Polyodon spathula) (Inoue et al. 2003; Kikugawa et al. 2004; and Fig. 17.1). Among these ‘ancient fish’, spotted gar is the species that has the smallest genome and can most readily be fertilized in vitro, providing embryos that are amenable to in situ hybridization studies and that grow to adults in laboratory aquaria with little care. Furthermore, no lineagespecific polyplodization has occurred in gar, in contrast to acipenseriforms. Thus, the spotted gar makes an ideal outgroup for the investigation of the mechanisms of evolution of gene function after GD. TE 847 367 EC 846 CO RR 845 Book ISBN: 978-3-642-31441-4 Page: 367/383 Polyploidy in Fish and the Teleost Genome Duplication 17.3.4 Mind the Gap: The TGD and the Teleost Radiation—Is There a Connection? The preceding paragraphs demonstrate that rayfin species experienced many WGD events and that teleosts are the most species-rich group of vertebrates. Importantly, many lineages experiencing WGD, such as barbs, armored catfish, and salmonids, appear to be particularly species-rich (Le Comber and Smith 2004). These considerations raise the question of whether WGD facilitates lineage diversification in fish. Ever since its discovery, the TGD has been suggested to have had a major impact on the radiation and biodiversity of the teleost lineage as a whole (Amores et al. 1998; Wittbrodt et al. 1998; Meyer and Schartl 1999; Taylor et al. 2001, 2003; Postlethwait et al. 2004; Volff 2005; Meyer and Van de Peer 2005; Froschauer et al. 2006; Ravi and Venkatesh 2008; Volff et al. 2011; Christoffels UN Editor Proof 17 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 17 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 368/383 I. Braasch and J. H. Postlethwait (a) Neofunctionalization and (b) Subfunction loss in tetrapods; subfunctionalization after TGD g1 at ele os t g1 bt ele os t g1 pre -TG D ray G1 fin tet rap od s g1 at ele os t g1 bt ele os t g1 pre -TG D ray G1 fin tet rap od s subfunctionalization sub sub TGD loss TGD (c) Reciprocal ohnologs gone missing; (d) Origin of new function in ancestral rayfin; subfunctionalization after TGD s ray tet rap od -TG D G1 g1 D pre os ele at g1 TGD sub TE VGD ohnologs t t os ele rap tet G1 reciprocal OGM g1 s od -TG D pre os sub g1 ele bt g1 g1 at ele os t t ray fin fin subfunctionalization after TGD TGD PR OO F neo bt Editor Proof 368 new 886 887 888 889 890 891 et al. 2004; Van de Peer et al. 2009), but likewise, this has been questioned by others based on the fossil record and speciation rate analyses (Donoghue and Purnell 2005; Hurley et al. 2007; Santini et al. 2009). UN 885 CO RR EC Fig. 17.8 Functional analysis of a pre-TGD diverging rayfin fish is important to distinguish neofunctionalization from subfunctionalization. a Preservation of one ohnolog by neofunctionalization (the function symbolized by a star) and preservation of the other by subfunctionalization. b Preservation of teleost ohnologs by subfunctionalization; the ‘star’ function was ancestral, not newly originated after the TGD, but was lost in the tetrapod lineage, as demonstrated by the pre-TGD-diverging rayfin outgroup. c The ‘star’ function that appears to be newly originated after the TGD was ancestral in a VGD ohnolog that was lost in the tetrapod lineage, while the rayfin lineage reciprocally lost the other ohnolog. d. A novel function not present in the ancestral gene could have originated in stem rayfins and then have been lost in one TGD ohnolog. Distinguishing among b, c and d requires an additional outgroup, such as a cartilaginous fish 17.3.4.1 Testing Speciation Models in the Fish World Genetic principles argue that a WGD event that is followed by population isolation, lineage-specific reciprocal nonfunctionalization events, and later hybridization would tend to promote speciation (Lynch and Force 2000; Werth and Layout: T1 Standard SC Chapter No.: 17 369 ancestral population AA whole genome duplication ohnolog divergence Aa Aa ; Ab Ab lineage divergence reciprocal nonfunctionalization Aa Aa ; A- A- A- A- ; Ab Ab populations hybridize Aa A- ; Ab Ahybrids mate 9 Aa - ; Ab survives F Polyploidy in Fish and the Teleost Genome Duplication Fig. 17.9 How WGD could facilitate lineage diversification. ‘A’ indicates the original gene. ‘Aa’ and ‘Ab’ indicate functional but diverged ohnologs. ‘A-’ indicates a nonfunctionalized allele. A ‘-’ indicates an allele that could be either functional or nonfunctional. The analysis assumes that non-functional alleles are recessive to functional alleles Editor Proof Book ISBN: 978-3-642-31441-4 Page: 369/383 PR OO 17 Book ID: 272454_1_En Date: 16-8-2012 3 Aa - ; A- Asurvives 3 A- A- ; Ab survives 1 A- A- ; A- Alethal 915 17.3.4.2 TGD-Based Morphological Evolution in Teleosts 898 899 900 901 902 903 904 905 906 907 908 909 910 911 912 913 916 917 918 919 TE 897 EC 896 CO RR 895 UN 894 D 914 Windham 1991). As Fig. 17.9 shows, the fertility of hybrids would be expected to be somewhat reduced if even a single duplicated gene had reciprocal nonfunctionalization, but if reciprocal nonfunctionalization (or reciprocal gene loss) occurred at least once on several or many chromosomes, which is likely, then hybrids would have greatly reduced fertility, thus strengthening lineage divergence. Regarding the TGD, while one study estimated *7 % reciprocal gene loss rate between zebrafish and Tetraodon, which would amount to *1,700 genes (Semon and Wolfe 2007b), more recent work including more teleost species found little support for reciprocal gene loss coinciding with speciation events in teleosts (Kassahn et al. 2009). The latter is more in line with our observation that differences in asymmetric gene loss from paralogons are rare between zebrafish and stickleback (see Sect. 17.3.2.4). Similar to reciprocal nonfunctionalization, speciation after WGD may also be promoted by reciprocal subfunctionalization that results in genomic incompatibilities in post-WGD hybrids (Lynch and Force 2000; Werth and Windham 1991). Differential functional evolution of TGD paralogs in divergent teleost lineages is being found in more and more cases (see below), yet associations with genetic incompatibilities and speciation in teleosts remain to be shown. Finally, it has also been suggested that the TGD may have caused reduced probability of extinction in teleosts (Crow and Wagner Crow and Wagner 2006). The extinction rate of teleost lineages, however, is apparently higher than in nonteleost rayfin lineages (Santini et al. 2009). 892 893 Besides their species richness, teleosts are also extremely diverse in terms of morphology, physiology, behavior, ecology, and biogeography. If the TGD laid the groundwork for divergence in such traits, one would expect to find differential function of TGD paralogs in various teleost lineages. Testing this proposition Layout: T1 Standard SC Chapter No.: 17 Book ID: 272454_1_En Date: 16-8-2012 928 929 930 931 932 933 934 935 936 937 938 939 940 941 942 943 944 945 946 947 948 949 950 951 952 953 954 955 956 957 958 959 960 961 962 F 926 927 PR OO 924 925 D 923 TE 922 involves experimentally challenging studies in different fish species, and only a handful of examples are currently available. For teleost sox9 co-orthologs, several ancient expression domains are differentially retained in zebrafish, stickleback, and medaka, indicating lineage-specific subfunctionalization (Cresko et al. 2003; Kluver et al. 2005; Yokoi et al. 2009); similar lineage-specific partitioning of expression domains have been reported for fgf8 (Jovelin et al. 2007) and tyrp1 (Braasch et al. 2009b) co-orthologs. A better way to compare gene functions across teleosts is to compare mutant phenotypes. Unfortunately, few genes have yet been mutated in teleosts other than zebrafish and medaka, and still fewer have been mutated in both species. Most teleosts have two TGD ohnologs encoding FGF receptor-1; in zebrafish and carp, both cyprinids, mutation of fgfr1a causes a scaleless phenotype (Rohner et al. 2009). In contrast, mutation of fgfr1a in medaka, a perciform fish, deletes the trunk and tail (the headfish mutant) (Yokoi et al. 2007; Shimada et al. 2008). This finding shows that subfunction partitioning in fgfr1 ohnologs occurred after the divergence of medaka and zebrafish lineages. Such divergence in subfunction partitioning after the TGD may provide the genomic basis for some morphological divergence in teleosts. And what about the TGD and morphological complexity? The identification of the seven hox clusters in teleosts, generally not considered to be more complex than other vertebrate groups, called into question the idea that morphological complexity along the body axis and the number of hox clusters and genes may be directly linked (Amores et al. 1998; Prince 2002). The high retention rate of TGD paralogs for developmentally important genes (Brunet et al. 2006; Kassahn et al. 2009), however, argues for an involvement of at least some TGD paralogs in teleost morphological evolution. Also, some aspects of teleost physiology and morphology are unique among vertebrates. To these belong, for example, the complexity and diversity of teleost body pigmentation and color patterning (Braasch et al. 2008); some functional modules of pigment cell development and differentiation have been retained in large part in duplicate in teleosts after the TGD (Braasch et al. 2009a). Another question that has so far been rarely addressed but will need to be investigated in more detail is the possible involvement of TGD duplicates in morphological and physiological novelties. For example, is functional divergence of TGD paralogs causing some of the synapomorphies of the teleost lineage (de Pinna 1996), such as the truly symmetric (homocercal) tail fin? Or are functional shifts in TGD paralogs involved in lineage-specific key innovations? We met one intriguing case earlier, the parallel, independent gain of electric organs used for electrical communication in knifefish and elephantfish by neofunctionalization of the scn4aa paralog (Arnegard et al. 2010). A more comprehensive picture, however, can only be gathered by studying TGD paralogs in more than just a few teleost representatives and adding an appropriate rayfin outgroup to the analyses. EC 921 CO RR 920 I. Braasch and J. H. Postlethwait UN Editor Proof 370 Book ISBN: 978-3-642-31441-4 Page: 370/383 Layout: T1 Standard SC Chapter No.: 17 963 Book ISBN: 978-3-642-31441-4 Page: 371/383 Polyploidy in Fish and the Teleost Genome Duplication 371 17.3.4.3 The Significance of the TGD for the Teleost Radiation 996 17.4 Outlook 971 972 973 974 975 976 977 978 979 980 981 982 983 984 985 986 987 988 989 990 991 992 993 994 997 998 999 1000 1001 1002 PR OO 970 D 969 TE 968 EC 966 967 CO RR 965 F 995 Based on the fossil record, a major temporal delay of about 150 million years separated the TGD event and the major radiation of the teleost lineage in the percomorphs (Donoghue and Purnell 2005; Hurley et al. 2007). Recent diversification rate shift analyses lead to the recognition that three incidents of accelerated speciation rates occurred within rayfin fish: on the branches leading (1) to the teleosts, (2) to the ostariophysians (zebrafish and relatives), and (3) to the percomorphs (perch-related fish, including medaka, stickleback, and pufferfish), respectively (Santini et al. 2009; Alfaro et al. 2009). Diversification rate shifts shortly after the TGD accounted for about 10 % of species diversity in teleosts (Santini et al. 2009). As noted above, however, TGD duplicates can alter their functions hundreds of millions of years after the GD itself. Whether individual fish taxa with additional WGD events (salmonids, barbs, etc.) are actually more species rich than diploid taxa has not been critically tested. For plants, it has recently been shown that recent polyploids had lower rates of speciation and higher rates of extinction than diploid control lineages (Mayrose et al. 2011). In contrast to recent polyploids, the number of ancient polyploidization events that provided today’s flora is higher than expected if diversification rates of diploids and polyploids were equal (Mayrose et al. 2011). Furthermore, ancient WGDs appear to be associated with several major radiations in the angiosperms (Soltis et al. 2009). In contrast, the ratio of species death rate (extinction) to species birth rate is 1.5 times higher in teleosts than in non-teleost rayfin fish (Santini et al. 2009). Thus, crucial questions are, when and under which circumstances do we expect a polyploidization event to deploy its full potential for lineage diversification and radiation? Given the deep influence of the TGD on genome evolution in teleosts illustrated in this chapter and given the described potential for ongoing non-, sub-, and neofunctionalization long after the TGD event, its major impact may not be apparent immediately after the GD event. Obviously, the TGD event was not an evolutionary dead-end, but quite the contrary should be considered a genomic exaptation that generated countless ‘‘spandrels’’ (Gould and Lewontin 1979) in the teleost genome upon which selection could act long after the actual polyploidization event, when ‘‘opportunity meets preparation’’. 964 As we have seen—although recognized since early studies of vertebrate genomes—the apparent propensity for polyploidization in the rayfin fish lineage, and among them, of cyprinid teleosts in particular, still requires an explanation. Since the first genomic evidence was published in 1998, the TGD hypothesis has overcome initial criticism, and the TGD is so far the best-studied GD in ‘‘fish’’. We are confident that many questions regarding the evolution of genome structure, UN Editor Proof 17 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 17 Book ID: 272454_1_En Date: 16-8-2012 I. Braasch and J. H. Postlethwait CO RR EC TE D PR OO F Editor Proof 372 Book ISBN: 978-3-642-31441-4 Page: 372/383 Fig. 17.10 The cover of Susumo Ohno’s book (1970) underlines the importance of fish for understanding our own past 1004 1005 1006 1007 1008 1009 1010 1011 1012 1013 gene families, and regulatory circuitry after the TGD can be solved by adding the sequences of basal rayfins (such as the spotted gar) and of basal teleosts to the genomic pond. The even harder question about the significance of the TGD for the teleost radiation will require extensive work on gene function evolution at all levels of teleost diversity, including numerous non-model fish species. The advent of next-generation sequencing techniques and the application of developmental biology studies to an increasing arsenal of fish species are promising for the acceleration of important research opportunities in this direction. Yet other questions, for example whether the TGD was an auto- or an allopolyploidization event, may never be solved due to the millions of years that elapsed since the event. UN 1003 Layout: T1 Standard SC Chapter No.: 17 Book ISBN: 978-3-642-31441-4 Page: 373/383 Polyploidy in Fish and the Teleost Genome Duplication 373 1024 1025 1026 1027 Acknowledgments We would like to thank Cristian Canestro for extensive discussions as well as Irene Pala for discussions of gene expression in the calandino. This work was supported by a grant from the Volkswagen Foundation Germany (IB) and National Institutes of Health grant R01 RR020833 (JHP). 1028 References 1029 1030 1031 1032 1033 1034 1035 1036 1037 1038 1039 1040 1041 1042 1043 1044 1045 1046 1047 1048 1049 1050 1051 1052 1053 1054 1055 1056 1057 1058 Alfaro ME, Santini F, Brock C, Alamillo H, Dornburg A, Rabosky DL, Carnevale G, Harmon LJ (2009) Nine exceptional radiations plus high turnover explain species diversity in jawed vertebrates. Proc Nat Acad Sci USA 106(32):13410–13414. doi:10.1073/pnas.0811087106 0811087106 [pii] Allendorf FW, Thorgaard GH (1984) Tetraploidy and the evolution of salmonid fishes. In: Turner BT (ed) Evolutionary genetics of fishes. Plenum Press, New York, pp 1–53 Altschmied J, Delfgaauw J, Wilde B, Duschl J, Bouneau L, Volff JN, Schartl M (2002) Subfunctionalization of duplicate mitf genes associated with differential degeneration of alternative exons in fish. Genetics 161(1):259–267 Alves MJ, Coelho MM, Collares-Pereira MJ (2001) Evolution in action through hybridisation and polyploidy in an Iberian freshwater fish: a genetic review. Genetica 111(1–3):375–385 Amores A, Force A, Yan YL, Joly L, Amemiya C, Fritz A, Ho RK, Langeland J, Prince V, Wang YL, Westerfield M, Ekker M, Postlethwait JH (1998) Zebrafish hox clusters and vertebrate genome evolution. Science 282(5394):1711–1714 Amores A, Catchen J, Ferrara A, Fontenot Q, Postlethwait JH (2011) Genome evolution and meiotic maps by massively parallel DNA sequencing: spotted gar, an outgroup for the teleost genome duplication. Genetics 188(4):799–808. doi:10.1534/genetics.111.127324 188/4/799 [pii] Aparicio S, Hawker K, Cottage A, Mikawa Y, Zuo L, Venkatesh B, Chen E, Krumlauf R, Brenner S (1997) Organization of the Fugu rubripes hox clusters: evidence for continuing evolution of vertebrate hox complexes. Nat Genet 16(1):79–83. doi:10.1038/ng0597-79 Aparicio S, Chapman J, Stupka E, Putnam N, Chia JM, Dehal P, Christoffels A, Rash S, Hoon S, Smit A, Gelpke MD, Roach J, Oh T, Ho IY, Wong M, Detter C, Verhoef F, Predki P, Tay A, Lucas S, Richardson P, Smith SF, Clark MS, Edwards YJ, Doggett N, Zharkikh A, Tavtigian SV, Pruss D, Barnstead M, Evans C, Baden H, Powell J, Glusman G, Rowen L, Hood L, Tan YH, Elgar G, Hawkins T, Venkatesh B, Rokhsar D, Brenner S (2002) Whole-genome shotgun assembly and analysis of the genome of Fugu rubripes. Science 297(5585):1301–1310. doi:10.1126/science.1072104 1072104[pii] Arnegard ME, Zwickl DJ, Lu Y, Zakon HH (2010) Old gene duplication facilitates origin and diversification of an innovative communication system-twice. P Nat Acad Sci USA 107(51):22172–22177. doi:10.1073/Pnas.1011803107 1021 1022 PR OO 1019 1020 D 1018 TE 1017 EC 1016 CO RR 1015 F 1023 Studying other, more recent GDs in rayfins, such as the salmonid and carp polyloidizations, will help to reveal in more detail the immediate epigenetic and genomic changes that follow a polyploidization. Luckily, the multitude and distribution of piscine GD events over evolutionary time will contribute to studies of the gradual evolution of vertebrate animals after different types of GD. Finally, as illustrated so marvelously by the cover of Ohno’s book (Ohno 1970) (Fig. 17.10), the study of polyploidizations in fish will not only allow us to better understand our own fishy heritage, but importantly, will help to inform us about the evolutionary processes that have shaped the vertebrate genome after VGD1 and VGD2, which had such a profound influence on our own, human evolution. 1014 UN Editor Proof 17 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 17 Book ID: 272454_1_En Date: 16-8-2012 EC TE D PR OO F Balon EK (2004) About the oldest domesticates among fishes. J Fish Biol 65:1–27. doi:10.1111/ j.1095-8649.2004.00563.x Birstein VJ, Hanner R, DeSalle R (1997) Phylogeny of the Acipenseriformes: cytogenetic and molecular approaches. Environ Biol Fish 48(1–4):127–156 Boldajipour B, Doitsidou M, Tarbashevich K, Laguri C, Yu SR, Ries J, Dumstrei K, Thelen S, Dorries J, Messerschmidt EM, Thelen M, Schwille P, Brand M, Lortat-Jacob H, Raz E (2011) Cxcl12 evolution—subfunctionalization of a ligand through altered interaction with the chemokine receptor. Development 138(14):2909–2914. doi:10.1242/Dev.068379 Braasch I, Salzburger W, Meyer A (2006) Asymmetric evolution in two fish-specifically duplicated receptor tyrosine kinase paralogons involved in teleost coloration. Mol Biol Evol 23(6):1192–1202. doi:10.1093/molbev/msk003 msk003 [pii] Braasch I, Schartl M, Volff JN (2007) Evolution of pigment synthesis pathways by gene and genome duplication in fish. BMC Evol Biol 7:74. doi:10.1186/1471-2148-7-74 1471-2148-774 [pii] Braasch I, Volff JN, Schartl M (2008) The evolution of teleost pigmentation and the fish-specific genome duplication. J Fish Biol 73(8):1891–1918. doi:10.1111/J.1095-8649.2008.02011.X Braasch I, Brunet F, Volff JN, Schartl M (2009a) Pigmentation pathway evolution after wholegenome duplication in fish. Genome Biol Evol 1:479–493. doi:10.1093/gbe/evp050 Braasch I, Liedtke D, Volff JN, Schartl M (2009b) Pigmentary function and evolution of tyrp1 gene duplicates in fish. Pigment Cell Melanoma Res 22(6):839–850. doi:10.1111/j.1755148X.2009.00614.x PCR614 [pii] Bromberg JF, Horvath CM, Wen ZL, Schreiber RD, Darnell JE (1996) Transcriptionally active Stat1 is required for the antiproliferative effects of both interferon alpha and interferon gamma. P Nat Acad Sci USA 93(15):7673–7678 Brunet FG, Crollius HR, Paris M, Aury JM, Gibert P, Jaillon O, Laudet V, Robinson-Rechavi M (2006) Gene loss and evolutionary rates following whole-genome duplication in teleost fishes. Mol Biol Evol 23(9):1808–1816. doi:10.1093/Molbev/Mls049 Bu L, Bergthorsson U, Katju V (2011) Local synteny and codon usage contribute to asymmetric sequence divergence of Saccharomyces cerevisiae gene duplicates. BMC Evol Biol 11:279. doi:10.1186/1471-2148-11-279 1471-2148-11-279 [pii] Canestro C, Yokoi H, Postlethwait JH (2007) Evolutionary developmental biology and genomics. Nat Rev Genet 8(12):932–942. doi:10.1038/nrg2226 nrg2226 [pii] Canestro C, Catchen JM, Rodriguez-Mari A, Yokoi H, Postlethwait JH (2009) Consequences of lineage-specific gene loss on functional evolution of surviving paralogs: ALDH1A and retinoic acid signaling in vertebrate genomes. PLoS Genet 5(5):e1000496. doi:10.1371/ journal.pgen.1000496 Catchen JM, Conery JS, Postlethwait JH (2009) Automated identification of conserved synteny after whole-genome duplication. Genome Res 19(8):1497–1505. doi:10.1101/Gr.090480.108 Catchen JM, Braasch I, Postlethwait JH (2011) Conserved synteny and the zebrafish genome. Method Cell Biol 104:259–285. doi:10.1016/B978-0-12-374814-0.00015-X Cerda J (2009) Molecular pathways during marine fish egg hydration: the role of aquaporins. J Fish Biol 75(9):2175–2196. doi:10.1111/j.1095-8649.2009.02397.x JFB2397 [pii] Chenuil A, Galtier N, Berrebi P (1999) A test of the hypothesis of an autopolyploid vs. allopolyploid origin for a tetraploid lineage: application to the genus Barbus (Cyprinidae). Heredity (Edinb) 82(Pt 4):373–380. doi:her489 [pii] Chiu CH, Dewar K, Wagner GP, Takahashi K, Ruddle F, Ledje C, Bartsch P, Scemama JL, Stellwag E, Fried C, Prohaska SJ, Stadler PF, Amemiya CT (2004) Bichir HoxA cluster sequence reveals surprising trends in ray-finned fish genomic evolution. Genome Res 14(1):11–17. doi:10.1101/gr.1712904 14/1/11[pii] Christoffels A, Koh EG, Chia JM, Brenner S, Aparicio S, Venkatesh B (2004) Fugu genome analysis provides evidence for a whole-genome duplication early during the evolution of rayfinned fishes. Mol Biol Evol 21(6):1146–1151. doi:10.1093/molbev/msh114 msh114 [pii] Comai L (2005) The advantages and disadvantages of being polyploid. Nat Rev Genet 6(11):836– 846. doi:10.1038/nrg1711 CO RR 1059 1060 1061 1062 1063 1064 1065 1066 1067 1068 1069 1070 1071 1072 1073 1074 1075 1076 1077 1078 1079 1080 1081 1082 1083 1084 1085 1086 1087 1088 1089 1090 1091 1092 1093 1094 1095 1096 1097 1098 1099 1100 1101 1102 1103 1104 1105 1106 1107 1108 1109 1110 1111 1112 I. Braasch and J. H. Postlethwait UN Editor Proof 374 Book ISBN: 978-3-642-31441-4 Page: 374/383 Layout: T1 Standard SC Chapter No.: 17 375 EC TE D PR OO F Cresko WA, Yan YL, Baltrus DA, Amores A, Singer A, Rodriguez-Mari A, Postlethwait JH (2003) Genome duplication, subfunction partitioning, and lineage divergence: sox9 in stickleback and zebrafish. Dev Dyn 228(3):480–489. doi:10.1002/dvdy.10424 Crow KD, Wagner GP (2006) Proceedings of the SMBE tri-national young investigators’ workshop. What is the role of genome duplication in the evolution of complexity and diversity? Mol Biol Evol 23(5):887–892. doi:10.1093/molbev/msj083 msj083 [pii] Crow KD, Stadler PF, Lynch VJ, Amemiya C, Wagner GP (2006) The ‘‘fish-specific’’ Hox cluster duplication is coincident with the origin of teleosts. Mol Biol Evol 23(1):121–136. doi:10.1093/molbev/msj020 msj020 [pii] Crow KD, Amemiya CT, Roth J, Wagner GP (2009) Hypermutability of HoxA13A and functional divergence from its paralog are associated with the origin of a novel developmental feature in zebrafish and related taxa (cypriniformes). Evolution 63(6):1574–1592. doi:10.1111/j.15585646.2009.00657.x EVO657 [pii] Danzmann RG, Davidson EA, Ferguson MM, Gharbi K, Koop BF, Hoyheim B, Lien S, Lubieniecki KP, Moghadam HK, Park J, Phillips RB, Davidson WS (2008) Distribution of ancestral proto-actinopterygian chromosome arms within the genomes of 4R-derivative salmonid fishes (rainbow trout and Atlantic salmon). BMC Genomics 9:557. doi:10.1186/ 1471-2164-9-557 David L, Blum S, Feldman MW, Lavi U, Hillel J (2003) Recent duplication of the, common carp (Cyprinus carpio L.) genome as revealed by analyses of microsatellite loci. Mol Biol Evol 20(9):1425–1434. doi:10.1093/molbev/msg173 David L, Rothbard S, Rubinstein I, Katzman H, Hulata G, Hillel J, Lavi U (2004) Aspects of red and black color inheritance in the Japanese ornamental (Koi) carp (Cyprinus carpio L.). Aquaculture 233(1–4):129–147. doi:10.1016/j.aquaculture.2003.10.033 Davidson WS, Koop BF, Jones SJM, Iturra P, Vidal R, Maass A, Jonassen I, Lien S, Omholt SW (2010) Sequencing the genome of the Atlantic salmon (Salmo salar). Genome Biol 11(9):403. doi:10.1186/gb-2010-11-9-403 Davis CA, Holmyard DP, Millen KJ, Joyner AL (1991) Examining pattern formation in mouse, chicken and frog embryos with an en-specific antiserum. Development 111(2):287–298 de Pinna MCC (1996) Teleostean monophyly. In: Stiassny MLJ, Parenti LR, Johnson GD (eds) Interrelationships of fishes. Academic, San Diego, pp 147–162 de Souza FS, Bumaschny VF, Low MJ, Rubinstein M (2005) Subfunctionalization of expression and peptide domains following the ancient duplication of the proopiomelanocortin gene in teleost fishes. Mol Biol Evol 22(12):2417–2427. doi:10.1093/molbev/msi236 msi236 [pii] Deng C, Cheng CH, Ye H, He X, Chen L (2010) Evolution of an antifreeze protein by neofunctionalization under escape from adaptive conflict. Proc Nat Acad Sci USA 107(50):21593–21598. doi:10.1073/pnas.1007883107 1007883107 [pii] Doitsidou M, Reichman-Fried M, Stebler J, Koprunner M, Dorries J, Meyer D, Esguerra CV, Leung T, Raz E (2002) Guidance of primordial germ cell migration by the chemokine SDF-1. Cell 111(5):647–659 S0092867402011352[pii] Donoghue PC, Purnell MA (2005) Genome duplication, extinction and vertebrate evolution. Trends Ecol Evol 20(6):312–319. doi:10.1016/j.tree.2005.04.008 S0169-5347(05)00108-4 [pii] Douard V, Brunet F, Boussau B, Ahrens-Fath I, Vlaeminck-Guillem V, Haendler B, Laudet V, Guiguen Y (2008) The fate of the duplicated androgen receptor in fishes: a late neofunctionalization event? BMC Evol Biol 8:336. doi:10.1186/1471-2148-8-336 Artn 336 Eiken HG, Njolstad PR, Molven A, Fjose A (1987) A zebrafish homeobox-containing gene with embryonic transcription. Biochem Biophys Res Commun 149(3):1165–1171 Elgar G, Clark MS, Meek S, Smith S, Warner S, Edwards YJ, Bouchireb N, Cottage A, Yeo GS, Umrania Y, Williams G, Brenner S (1999) Generation and analysis of 25 Mb of genomic DNA from the pufferfish Fugu rubripes by sequence scanning. Genome Res 9(10):960–971 Ferris SD (1984) Tetraploidy and the evolution of the catostomid fishes. In: Turner BT (ed) Evolutionary genetics of fishes. Plenus Press, New York, pp 54–93 Ferris SD, Whitt GS (1977a) Duplicate gene expression in diploid and tetraploid loaches (cypriniformes, Cobitidae). Biochem Genet 15(11–12):1097–1112 CO RR 1113 1114 1115 1116 1117 1118 1119 1120 1121 1122 1123 1124 1125 1126 1127 1128 1129 1130 1131 1132 1133 1134 1135 1136 1137 1138 1139 1140 1141 1142 1143 1144 1145 1146 1147 1148 1149 1150 1151 1152 1153 1154 1155 1156 1157 1158 1159 1160 1161 1162 1163 1164 1165 1166 Book ISBN: 978-3-642-31441-4 Page: 375/383 Polyploidy in Fish and the Teleost Genome Duplication UN Editor Proof 17 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 17 Book ID: 272454_1_En Date: 16-8-2012 EC TE D PR OO F Ferris SD, Whitt GS (1977b) Evolution of duplicate gene-expression in carp (Cyprinus carpio). Experientia 33(10):1299–1301 Finn RN, Kristoffersen BA (2007) Vertebrate vitellogenin gene duplication in relation to the ‘‘3R hypothesis’’: correlation to the pelagic egg and the oceanic radiation of teleosts. PLoS One 2(1):e169. doi:10.1371/journal.pone.0000169 Force A, Lynch M, Pickett FB, Amores A, Yan YL, Postlethwait J (1999) Preservation of duplicate genes by complementary, degenerative mutations. Genetics 151(4):1531–1545 Froschauer A, Braasch I, Volff JN (2006) Fish genomes, comparative genomics and vertebrate evolution. Curr Genomics 7(1):43–57. doi:10.2174/138920206776389766 Gardner CA, Barald KF (1992) Expression patterns of engrailed-like proteins in the chick embryo. Dev Dyn 193(4):370–388. doi:10.1002/aja.1001930410 Gates MA, Kim L, Egan ES, Cardozo T, Sirotkin HI, Dougan ST, Lashkari D, Abagyan R, Schier AF, Talbot WS (1999) A genetic linkage map for zebrafish: comparative analysis and localization of genes and expressed sequences. Genome Res 9(4):334–347 Gomez A, Volff JN, Hornung U, Schartl M, Wellbrock C (2004) Identification of a second egfr gene in Xiphophorus uncovers an expansion of the epidermal growth factor receptor family in fish. Mol Biol Evol 21(2):266–275. doi:10.1093/molbev/msh017 msh017[pii] Gould SJ, Lewontin RC (1979) Spandrels of San-Marco and the Panglossian paradigm—a critique of the adaptationist program. Proc R Soc Lond B Biol Sci 205(1161):581–598 Graham A, Papalopulu N, Krumlauf R (1989) The murine and Drosophila homeobox gene complexes have common features of organization and expression. Cell 57(3):367–378. doi:0092-8674(89)90912-4 [pii] Haussler D, O’Brien SJ, Ryder OA, Barker FK, Clamp M, Crawford AJ, Hanner R, Hanotte O, Johnson WE, McGuire JA, Miller W, Murphy RW, Murphy WJ, Sheldon FH, Sinervo B, Venkatesh B, Wiley EO, Allendorf FW, Amato G, Baker CS, Bauer A, Beja-Pereira A, Bermingham E, Bernardi G, Bonvicino CR, Brenner S, Burke T, Cracraft J, Diekhans M, Edwards S, Ericson PGP, Estes J, Fjelsda J, Flesness N, Gamble T, Gaubert P, Graphodatsky AS, Graves JAM, Green ED, Green RE, Hackett S, Hebert P, Helgen KM, Joseph L, Kessing B, Kingsley DM, Lewin HA, Luikart G, Martelli P, Moreira MAM, Nguyen N, Orti G, Pike BL, Rawson DM, Schuster SC, Seuanez HN, Shaffer HB, Springer MS, Stuart JM, Sumner J, Teeling E, Vrijenhoek RC, Ward RD, Warren WC, Wayne R, Williams TM, Wolfe ND, Zhang YP, Graph-Odatsky A, Johnson WE, Felsenfeld A, Turner S, Scientists GKC, Grp M, Grp B, Grp AR, Grp F, Grp GP, Grp A (2009) Genome 10 K: a proposal to obtain whole-genome sequence for 10,000 vertebrate species. J Hered 100(6):659–674. doi:10.1093/jhered/esp086 He X, Zhang J (2005) Rapid subfunctionalization accompanied by prolonged and substantial neofunctionalization in duplicate gene evolution. Genetics 169(2):1157–1164. doi:10.1534/ genetics.104.037051 genetics.104.037051 [pii] Hoegg S, Meyer A (2005) Hox clusters as models for vertebrate genome evolution. Trends Genet 21(8):421–424. doi:10.1016/j.tig.2005.06.004 S0168-9525(05)00165-4 [pii] Hoegg S, Brinkmann H, Taylor JS, Meyer A (2004) Phylogenetic timing of the fish-specific genome duplication correlates with the diversification of teleost fish. J Mol Evol 59(2):190– 203. doi:10.1007/s00239-004-2613-z Hoegg S, Boore JL, Kuehl JV, Meyer A (2007) Comparative phylogenomic analyses of teleost fish hox gene clusters: lessons from the cichlid fish Astatotilapia burtoni. BMC Genomics 8:317. doi:10.1186/1471-2164-8-317 1471-2164-8-317 [pii] Howarth DL, Law SHW, Barnes B, Hall JM, Hinton DE, Moore L, Maglich JM, Moore JT, Kullman SW (2008) Paralogous vitamin D receptors in teleosts: transition of nuclear receptor function. Endocrinology 149(5):2411–2422. doi:10.1210/En.2007-1256 Hufton AL, Groth D, Vingron M, Lehrach H, Poustka AJ, Panopoulou G (2008) Early vertebrate whole genome duplications were predated by a period of intense genome rearrangement. Genome Res 18(10):1582–1591. doi:10.1101/gr.080119.108 gr.080119.108 [pii] Huminiecki L, Heldin CH (2010) 2R and remodeling of vertebrate signal transduction engine. BMC Biol 8:146. doi:10.1186/1741-7007-8-146 1741-7007-8-146 [pii] CO RR 1167 1168 1169 1170 1171 1172 1173 1174 1175 1176 1177 1178 1179 1180 1181 1182 1183 1184 1185 1186 1187 1188 1189 1190 1191 1192 1193 1194 1195 1196 1197 1198 1199 1200 1201 1202 1203 1204 1205 1206 1207 1208 1209 1210 1211 1212 1213 1214 1215 1216 1217 1218 1219 I. Braasch and J. H. Postlethwait UN Editor Proof 376 Book ISBN: 978-3-642-31441-4 Page: 376/383 Layout: T1 Standard SC Chapter No.: 17 377 EC TE D PR OO F Hurley IA, Mueller RL, Dunn KA, Schmidt EJ, Friedman M, Ho RK, Prince VE, Yang Z, Thomas MG, Coates MI (2007) A new time-scale for ray-finned fish evolution. Proc Biol Sci 274(1609):489–498 Inoue JG, Miya M, Tsukamoto K, Nishida M (2003) Basal actinopterygian relationships: a mitogenomic perspective on the phylogeny of the ‘‘ancient fish’’. Mol Phylogenet Evol 26(1):110–120 Jaillon O, Aury JM, Brunet F, Petit JL, Stange-Thomann N, Mauceli E, Bouneau L, Fischer C, Ozouf-Costaz C, Bernot A, Nicaud S, Jaffe D, Fisher S, Lutfalla G, Dossat C, Segurens B, Dasilva C, Salanoubat M, Levy M, Boudet N, Castellano S, Anthouard V, Jubin C, Castelli V, Katinka M, Vacherie B, Biemont C, Skalli Z, Cattolico L, Poulain J, De Berardinis V, Cruaud C, Duprat S, Brottier P, Coutanceau JP, Gouzy J, Parra G, Lardier G, Chapple C, McKernan KJ, McEwan P, Bosak S, Kellis M, Volff JN, Guigo R, Zody MC, Mesirov J, Lindblad-Toh K, Birren B, Nusbaum C, Kahn D, Robinson-Rechavi M, Laudet V, Schachter V, Quetier F, Saurin W, Scarpelli C, Wincker P, Lander ES, Weissenbach J, Roest Crollius H (2004) Genome duplication in the teleost fish Tetraodon nigroviridis reveals the early vertebrate proto-karyotype. Nature 431(7011):946–957. doi:10.1038/nature03025 nature03025 [pii] Jovelin R, He X, Amores A, Yan YL, Shi R, Qin B, Roe B, Cresko WA, Postlethwait JH (2007) Duplication and divergence of fgf8 functions in teleost development and evolution. J Exp Zool B Mol Dev Evol 308(6):730–743. doi:10.1002/jez.b.21193 Joyner AL, Martin GR (1987) En-1 and En-2, two mouse genes with sequence homology to the Drosophila engrailed gene: expression during embryogenesis. Genes Dev 1(1):29–38 Karanth S, Lall SP, Denovan-Wright EM, Wright JM (2009) Differential transcriptional modulation of duplicated fatty acid-binding protein genes by dietary fatty acids in zebrafish (Danio rerio): evidence for subfunctionalization or neofunctionalization of duplicated genes. BMC Evol Biol 9:219. doi:10.1186/1471-2148-9-219 1471-2148-9-219 [pii] Kasahara M, Naruse K, Sasaki S, Nakatani Y, Qu W, Ahsan B, Yamada T, Nagayasu Y, Doi K, Kasai Y, Jindo T, Kobayashi D, Shimada A, Toyoda A, Kuroki Y, Fujiyama A, Sasaki T, Shimizu A, Asakawa S, Shimizu N, Hashimoto S, Yang J, Lee Y, Matsushima K, Sugano S, Sakaizumi M, Narita T, Ohishi K, Haga S, Ohta F, Nomoto H, Nogata K, Morishita T, Endo T, Shin IT, Takeda H, Morishita S, Kohara Y (2007) The medaka draft genome and insights into vertebrate genome evolution. Nature 447(7145):714–719. doi:10.1038/nature05846 nature05846 [pii] Kassahn KS, Dang VT, Wilkins SJ, Perkins AC, Ragan MA (2009) Evolution of gene function and regulatory control after whole-genome duplication: comparative analyses in vertebrates. Genome Res 19(8):1404–1418. doi:10.1101/Gr.086827.108 Kikugawa K, Katoh K, Kuraku S, Sakurai H, Ishida O, Iwabe N, Miyata T (2004) Basal jawed vertebrate phylogeny inferred from multiple nuclear DNA-coded genes. BMC Biol 2:3. doi:10.1186/1741-7007-2-3 1741-7007-2-3 [pii] Kluver N, Kondo M, Herpin A, Mitani H, Schartl M (2005) Divergent expression patterns of Sox9 duplicates in teleosts indicate a lineage specific subfunctionalization. Dev Genes Evol 215(6):297–305. doi:10.1007/s00427-005-0477-x Kohn M, Hogel J, Vogel W, Minich P, Kehrer-Sawatzki H, Graves JA, Hameister H (2006) Reconstruction of a 450-My-old ancestral vertebrate protokaryotype. Trends Genet 22(4):203–210. doi:10.1016/j.tig.2006.02.008 S0168-9525(06)00063-1 [pii] Komiyama T, Kobayashi H, Tateno Y, Inoko H, Gojobori T, Ikeo K (2009) An evolutionary origin and selection process of goldfish. Gene 430(1–2):5–11. doi:10.1016/j.gene.2008.10.019 S0378-1119(08)00543-X [pii] Koop BF, von Schalburg KR, Leong J, Walker N, Lieph R, Cooper GA, Robb A, Beetz-Sargent M, Holt RA, Moore R, Brahmbhatt S, Rosner J, Rexroad CE, McGowan CR, Davidson WS (2008) A salmonid EST genomic study: genes, duplications, phylogeny and microarrays. BMC Genomics 9:545. doi:10.1186/1471-2164-9-545 Kupka E (1948) Chromosomale Verschiedenheiten bei schweizerischen Coregonen (Felchen). Rev Suisse Zool 55:293–295 CO RR 1220 1221 1222 1223 1224 1225 1226 1227 1228 1229 1230 1231 1232 1233 1234 1235 1236 1237 1238 1239 1240 1241 1242 1243 1244 1245 1246 1247 1248 1249 1250 1251 1252 1253 1254 1255 1256 1257 1258 1259 1260 1261 1262 1263 1264 1265 1266 1267 1268 1269 1270 1271 1272 Book ISBN: 978-3-642-31441-4 Page: 377/383 Polyploidy in Fish and the Teleost Genome Duplication UN Editor Proof 17 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 17 Book ID: 272454_1_En Date: 16-8-2012 EC TE D PR OO F Larhammar D, Risinger C (1994) Molecular genetic aspects of tetraploidy in the common carp Cyprinus carpio. Mol Phylogenet Evol 3(1):59–68. doi:10.1006/mpev.1994.1007 S10557903(84)71007-4 [pii] Le Comber SC, Smith C (2004) Polyploidy in fishes: patterns and processes. Biol J Linn Soc 82(4):431–442 Leggatt RA, Iwama GK (2003) Occurrence of polyploidy in the fishes. Rev Fish Biol Fisher 13(3):237–246 Leong JS, Jantzen SG, von Schalburg KR, Cooper GA, Messmer AM, Liao NY, Munro S, Moore R, Holt RA, Jones SJM, Davidson WS, Koop BF (2010) Salmo salar and Esox lucius fulllength cDNA sequences reveal changes in evolutionary pressures on a post-tetraploidization genome. BMC Genomics 11:279. doi:10.1186/1471-2164-11-279 Li YJ, Yu Z, Zhang MZ, Qian C, Abe S, Arai K (2011) The origin of natural tetraploid loach Misgurnus anguillicaudatus (Teleostei: Cobitidae) inferred from meiotic chromosome configurations. Genetica 139(6):805–811. doi:10.1007/s10709-011-9585-x Lien S, Gidskehaug L, Moen T, Hayes BJ, Berg PR, Davidson WS, Omholt SW, Kent MP (2011) A dense SNP-based linkage map for Atlantic salmon (Salmo salar) reveals extended chromosome homeologies and striking differences in sex-specific recombination patterns. BMC Genomics 12(1):615. doi:10.1186/1471-2164-12-615 1471-2164-12-615 [pii] Lister JA, Close J, Raible DW (2001) Duplicate mitf genes in zebrafish: complementary expression and conservation of melanogenic potential. Dev Biol 237(2):333–344 Ludwig A, Belfiore NM, Pitra C, Svirsky V, Jenneckens I (2001) Genome duplication events and functional reduction of ploidy levels in sturgeon (Acipenser, Huso and Scaphirhynchus). Genetics 158(3):1203–1215 Luo J, Stadler PF, He S, Meyer A (2007) PCR survey of Hox genes in the goldfish Carassius auratus auratus. J Exp Zool B Mol Dev Evol 308(3):250–258. doi:10.1002/jez.b.21144 Lynch M, Conery JS (2000) The evolutionary fate and consequences of duplicate genes. Science 290(5494):1151–1155 8976 [pii] Lynch M, Force AG (2000) The origin of interspecific genomic incompatibility via gene duplication. Am Nat 156(6):590–605 Mable BK (2004) ‘Why polyploidy is rarer in animals than in plants’: myths and mechanisms. Biol J Linn Soc 82(4):453–466 Mable BK, Alexandrou MA, Taylor MI (2011) Genome duplication in amphibians and fish: an extended synthesis. J Zool 284(3):151–182. doi:10.1111/j.1469-7998.2011.00829.x Mank JE, Avise JC (2006a) Cladogenetic correlates of genomic expansions in the recent evolution of actinopterygiian fishes. Proc R Soc Lond B Biol Sci 273(1582):33–38. doi:10.1098/rspb.2005.3295 Mank JE, Avise JC (2006b) Phylogenetic conservation of chromosome numbers in actinopterygiian fishes. Genetica 127(1–3):321–327. doi:10.1007/s10709-005-5248-0 Mayrose I, Zhan SH, Rothfels CJ, Magnuson-Ford K, Barker MS, Rieseberg LH, Otto SP (2011) Recently formed polyploid plants diversify at lower rates. Science 333(6047):1257. doi:10.1126/Science.1207205 Meyer A (1998) Hox gene variation and evolution. Nature 391 (6664):225, 227–228. doi:10.1038/34530 Meyer A, Schartl M (1999) Gene and genome duplications in vertebrates: the one-to-four (-toeight in fish) rule and the evolution of novel gene functions. Curr Opin Cell Biol 11(6):699– 704. doi:S0955-0674(99)00039-3 [pii] Meyer A, Van de Peer Y (2005) From 2R to 3R: evidence for a fish-specific genome duplication (FSGD). BioEssays 27(9):937–945. doi:10.1002/bies.20293 Miller MR, Brunelli JP, Wheeler PA, Liu S, Rexroad CE 3rd, Palti Y, Doe CQ, Thorgaard GH (2011) A conserved haplotype controls parallel adaptation in geographically distant salmonid populations. Mol Ecol. doi:10.1111/j.1365-294X.2011.05305.x Misof BY, Wagner GP (1996) Evidence for four hox clusters in the killifish Fundulus heteroclitus (teleostei). Mol Phylogenet Evol 5(2):309–322. doi:10.1006/mpev.1996.0026 S10557903(96)90026-3 [pii] CO RR 1273 1274 1275 1276 1277 1278 1279 1280 1281 1282 1283 1284 1285 1286 1287 1288 1289 1290 1291 1292 1293 1294 1295 1296 1297 1298 1299 1300 1301 1302 1303 1304 1305 1306 1307 1308 1309 1310 1311 1312 1313 1314 1315 1316 1317 1318 1319 1320 1321 1322 1323 1324 1325 1326 I. Braasch and J. H. Postlethwait UN Editor Proof 378 Book ISBN: 978-3-642-31441-4 Page: 378/383 Layout: T1 Standard SC Chapter No.: 17 379 EC TE D PR OO F Moghadam HK, Ferguson MM, Danzmann RG (2005a) Evidence for Hox gene duplication in rainbow trout (Oncorhynchus mykiss): a tetraploid model species. J Mol Evol 61(6):804–818. doi:10.1007/s00239-004-0230-5 Moghadam HK, Ferguson MM, Danzmann RG (2005b) Evolution of Hox clusters in Salmonidae: a comparative analysis between Atlantic salmon (Salmo salar) and rainbow trout (Oncorhynchus mykiss). J Mol Evol 61(5):636–649. doi:10.1007/s00239-004-0338-7 Moghadam HK, Ferguson MM, Danzmann RG (2009) Comparative genomics and evolution of conserved noncoding elements (CNE) in rainbow trout. BMC Genomics 10:278. doi:10.1186/ 1471-2164-10-278 Artn 278 Morizot DC (1990) Use of fish gene maps to predict ancestral vertebrate genome organization. In: Ogita Z-I, Markert CL (eds) Isozymes: structure, function, and use in biology, and medicine. Wiley-Liss, Inc., New York, pp 207–234 Muller HJ (1925) Why polyploidy is rarer in animals than in plants. Am Nat 59(663):346–353 Mulley JF, Chiu CH, Holland PW (2006) Breakup of a homeobox cluster after genome duplication in teleosts. Proc Nat Acad Sci USA 103(27):10369–10372. doi:10.1073/ pnas.0600341103 0600341103 [pii] Mungpakdee S, Seo HC, Angotzi AR, Dong XJ, Akalin A, Chourrout D (2008a) Differential evolution of the 13 Atlantic salmon Hox clusters. Mol Biol Evol 25(7):1333–1343. doi:10.1093/molbev/msn097 Mungpakdee S, Seo HC, Chourrout D (2008b) Spatio-temporal expression patterns of anterior Hox genes in Atlantic salmon (Salmo salar). Gene Expr Patterns 8(7–8):508–514. doi:10.1016/j.gep.2008.06.004 Nair S, Schilling TF (2008) Chemokine signaling controls endodermal migration during zebrafish gastrulation. Science 322(5898):89–92. doi:10.1126/science.1160038 1160038 [pii] Nakatani Y, Takeda H, Kohara Y, Morishita S (2007) Reconstruction of the vertebrate ancestral genome reveals dynamic genome reorganization in early vertebrates. Genome Res 17(9):1254–1265. doi:10.1101/gr.6316407 Naruse K, Tanaka M, Mita K, Shima A, Postlethwait J, Mitani H (2004) A medaka gene map: the trace of ancestral vertebrate proto-chromosomes revealed by comparative gene mapping. Genome Res 14(5):820–828. doi:10.1101/gr.2004004 2004004 [pii] Nelson JS (2006) Fishes of the world, 4th edn. Wiley, Hoboken Njolstad PR, Molven A, Hordvik I, Apold J, Fjose A (1988) Primary structure, developmentally regulated expression and potential duplication of the zebrafish homeobox gene ZF-21. Nucleic Acids Res 16(19):9097–9111 Ohno S (1970) Evolution by gene duplication. Springer, Berlin Ohno S, Muramoto J, Christia L, Atkin NB (1967) Diploid-tetraploid relationship among oldworld members of fish family Cyprinidae. Chromosoma 23(1):1 Ohno S, Wolf U, Atkin NB (1968) Evolution from fish to mammals by gene duplication. Hereditas-Genetisk A 59(1):169 & Okuda Y, Yoda H, Uchikawa M, Furutani-Seiki M, Takeda H, Kondoh H, Kamachi Y (2006) Comparative genomic and expression analysis of group B1 sox genes in zebrafish indicates their diversification during vertebrate evolution. Dev Dyn 235(3):811–825. doi:10.1002/ dvdy.20678 Oliveira C, Almeidatoledo LF, Mori L, Toledofilho SA (1992) Extensive chromosomal rearrangements and nuclear-DNA content changes in the evolution of the armored catfishes genus Corydoras (pisces, siluriformes, callichthyidae). J Fish Biol 40(3):419–431 Otto SP (2007) The evolutionary consequences of polyploidy. Cell 131(3):452–462. doi:10.1016/ j.cell.2007.10.022 Otto SP, Whitton J (2000) Polyploid incidence and evolution. Annu Rev Genet 34:401–437. doi:10.1146/annurev.genet.34.1.401 34/1/401 [pii] Pala I, Coelho MM, Schartl M (2008) Dosage compensation by gene-copy silencing in a triploid hybrid fish. Curr Biol 18(17):1344–1348. doi:10.1016/j.cub.2008.07.096 CO RR 1327 1328 1329 1330 1331 1332 1333 1334 1335 1336 1337 1338 1339 1340 1341 1342 1343 1344 1345 1346 1347 1348 1349 1350 1351 1352 1353 1354 1355 1356 1357 1358 1359 1360 1361 1362 1363 1364 1365 1366 1367 1368 1369 1370 1371 1372 1373 1374 1375 1376 1377 1378 Book ISBN: 978-3-642-31441-4 Page: 379/383 Polyploidy in Fish and the Teleost Genome Duplication UN Editor Proof 17 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 17 Book ID: 272454_1_En Date: 16-8-2012 EC TE D PR OO F Pala I, Schartl M, Brito M, Vacas JM, Coelho MM (2010) Gene expression regulation and lineage evolution: the north and south tale of the hybrid polyploid Squalius alburnoides complex. Proc R Soc B Biol Sci 277(1699):3519–3525. doi:10.1098/rspb.2010.1071 Pandian TJ, Koteeswaran R (1999) Natural occurrence of monoploids and polyploids in the Indian catfish, Heteropneustes fossilis. Curr Sci India 76(8):1134–1137 Phillips R, Rab P (2001) Chromosome evolution in the Salmonidae (pisces): an update. Biol Rev Camb Philos Soc 76(1):1–25 Phillips RB, Keatley KA, Morasch MR, Ventura AB, Lubieniecki KP, Koop BF, Danzmann RG, Davidson WS (2009) Assignment of Atlantic salmon (Salmo salar) linkage groups to specific chromosomes: conservation of large syntenic blocks corresponding to whole chromosome arms in rainbow trout (Oncorhynchus mykiss). BMC Genet 10:46. doi:10.1186/1471-2156-10-46 Piferrer F, Beaumont A, Falguiere JC, Flajshans M, Haffray P, Colombo L (2009) Polyploid fish and shellfish: production, biology and applications to aquaculture for performance improvement and genetic containment. Aquaculture 293(3–4):125–156. doi:10.1016/j.aquaculture. 2009.04.036 Postlethwait JH, Yan YL, Gates MA, Horne S, Amores A, Brownlie A, Donovan A, Egan ES, Force A, Gong Z, Goutel C, Fritz A, Kelsh R, Knapik E, Liao E, Paw B, Ransom D, Singer A, Thomson M, Abduljabbar TS, Yelick P, Beier D, Joly JS, Larhammar D, Rosa F, Westerfield M, Zon LI, Johnson SL, Talbot WS (1998) Vertebrate genome evolution and the zebrafish gene map. Nat Genet 18(4):345–349. doi:10.1038/ng0498-345 Postlethwait JH, Woods IG, Ngo-Hazelett P, Yan YL, Kelly PD, Chu F, Huang H, Hill-Force A, Talbot WS (2000) Zebrafish comparative genomics and the origins of vertebrate chromosomes. Genome Res 10(12):1890–1902 Postlethwait J, Amores A, Cresko W, Singer A, Yan YL (2004) Subfunction partitioning, the teleost radiation and the annotation of the human genome. Trends Genet 20(10):481–490. doi:10.1016/j.tig.2004.08.001 S0168-9525(04)00213-6 [pii] Prince VE (2002) The Hox paradox: more complex(es) than imagined. Dev Biol 249(1):1–15. doi:10.1006/Dbio.2002.0745 Prince VE, Joly L, Ekker M, Ho RK (1998) Zebrafish Hox genes: genomic organization and modified colinear expression patterns in the trunk. Development 125(3):407–420 Prohaska SJ, Stadler PF (2004) The duplication of the Hox gene clusters in teleost fishes. Theory Biosci 123(1):89–110. doi:10.1016/j.thbio.2004.03.004 Putnam NH, Butts T, Ferrier DE, Furlong RF, Hellsten U, Kawashima T, Robinson-Rechavi M, Shoguchi E, Terry A, Yu JK, Benito-Gutierrez EL, Dubchak I, Garcia-Fernandez J, GibsonBrown JJ, Grigoriev IV, Horton AC, de Jong PJ, Jurka J, Kapitonov VV, Kohara Y, Kuroki Y, Lindquist E, Lucas S, Osoegawa K, Pennacchio LA, Salamov AA, Satou Y, Sauka-Spengler T, Schmutz J, Shin IT, Toyoda A, Bronner-Fraser M, Fujiyama A, Holland LZ, Holland PW, Satoh N, Rokhsar DS (2008) The amphioxus genome and the evolution of the chordate karyotype. Nature 453(7198):1064–1071. doi:10.1038/nature06967 nature06967 [pii] Raincrow JD, Dewar K, Stocsits C, Prohaska SJ, Amemiya CT, Stadler PF, Chiu CH (2011) Hox clusters of the bichir (Actinopterygii, Polypterus senegalus) highlight unique patterns of sequence evolution in gnathostome phylogeny. J Exp Zool B Mol Dev Evol 316(6):451–464. doi:10.1002/jez.b.21420 Rastogi S, Liberles DA (2005) Subfunctionalization of duplicated genes as a transition state to neofunctionalization. BMC Evol Biol 5:28. doi:10.1186/1471-2148-5-28 1471-2148-5-28 [pii] Ravi V, Venkatesh B (2008) Rapidly evolving fish genomes and teleost diversity. Curr Opin Genet Dev 18(6):544–550. doi:10.1016/j.gde.2008.11.001 S0959-437X(08)00151-2 [pii] Richardson BE, Lehmann R (2010) Mechanisms guiding primordial germ cell migration: strategies from different organisms. Nat Rev Mol Cell Biol 11(1):37–49. doi:10.1038/ nrm2815 nrm2815 [pii] Risinger C, Larhammar D (1993) Multiple loci for synapse protein SNAP-25 in the tetraploid goldfish. Proc Nat Acad Sci USA 90(22):10598–10602 CO RR 1379 1380 1381 1382 1383 1384 1385 1386 1387 1388 1389 1390 1391 1392 1393 1394 1395 1396 1397 1398 1399 1400 1401 1402 1403 1404 1405 1406 1407 1408 1409 1410 1411 1412 1413 1414 1415 1416 1417 1418 1419 1420 1421 1422 1423 1424 1425 1426 1427 1428 1429 1430 I. Braasch and J. H. Postlethwait UN Editor Proof 380 Book ISBN: 978-3-642-31441-4 Page: 380/383 Layout: T1 Standard SC Chapter No.: 17 381 EC TE D PR OO F Robinson-Rechavi M, Marchand O, Escriva H, Bardet PL, Zelus D, Hughes S, Laudet V (2001) Euteleost fish genomes are characterized by expansion of gene families. Genome Res 11(5):781–788. doi:10.1101/gr.165601 Rohner N, Bercsenyi M, Orban L, Kolanczyk ME, Linke D, Brand M, Nusslein-Volhard C, Harris MP (2009) Duplication of fgfr1 permits Fgf signaling to serve as a target for selection during domestication. Curr Biol 19(19):1642–1647. doi:10.1016/j.cub.2009.07.065 S09609822(09)01542-5 [pii] Santini F, Harmon LJ, Carnevale G, Alfaro ME (2009) Did genome duplication drive the origin of teleosts? A comparative study of diversification in ray-finned fishes. BMC Evol Biol 9:194. doi:10.1186/1471-2148-9-194 1471-2148-9-194 [pii] Sato Y, Nishida M (2007) Post-duplication charge evolution of phosphoglucose isomerases in teleost fishes through weak selection on many amino acid sites. BMC Evol Biol 7:204. doi:10.1186/1471-2148-7-204 Artn 204 Sato Y, Hashiguchi Y, Nishida M (2009) Temporal pattern of loss/persistence of duplicate genes involved in signal transduction and metabolic pathways after teleost-specific genome duplication. BMC Evol Biol 9:127. doi:10.1186/1471-2148-9-127 1471-2148-9-127 [pii] Scannell DR, Wolfe KH (2008) A burst of protein sequence evolution and a prolonged period of asymmetric evolution follow gene duplication in yeast. Genome Res 18(1):137–147. doi:10.1101/gr.6341207 gr.6341207 [pii] Schultz RJ (1980) Role of polyploidy in the evolution of fishes. In: Lewis WH (ed) Polyploidy— biological relevance. Basic life science, vol 13. Plenum Press, New York, pp 313–340 Semon M, Wolfe KH (2007a) Rearrangement rate following the whole-genome duplication in teleosts. Mol Biol Evol 24(3):860–867. doi:10.1093/molbev/msm003 msm003 [pii] Semon M, Wolfe KH (2007b) Reciprocal gene loss between Tetraodon and zebrafish after whole genome duplication in their ancestor. Trends Genet 23(3):108–112. doi:10.1016/ j.tig.2007.01.003 S0168-9525(07)00021-2 [pii] Setiamarga DH, Miya M, Yamanoue Y, Azuma Y, Inoue JG, Ishiguro NB, Mabuchi K, Nishida M (2009) Divergence time of the two regional medaka populations in Japan as a new time scale for comparative genomics of vertebrates. Biol Lett 5(6):812–816. doi:10.1098/ rsbl.2009.0419 rsbl.2009.0419 [pii] Sha Z, Yu P, Takano T, Liu H, Terhune J, Liu Z (2008) The warm temperature acclimation protein Wap65 as an immune response gene: its duplicates are differentially regulated by temperature and bacterial infections. Mol Immunol 45(5):1458–1469. doi:10.1016/ J.Molimm.2007.08.012 Shimada A, Yabusaki M, Niwa H, Yokoi H, Hatta K, Kobayashi D, Takeda H (2008) Maternalzygotic medaka mutants for fgfr1 reveal its essential role in the migration of the axial mesoderm but not the lateral mesoderm. Development 135(2):281–290. doi:10.1242/ dev.011494 135/2/281 [pii] Siegel N, Hoegg S, Salzburger W, Braasch I, Meyer A (2007) Comparative genomics of ParaHox clusters of teleost fishes: gene cluster breakup and the retention of gene sets following whole genome duplications. BMC Genomics 8:312. doi:10.1186/1471-2164-8-312 1471-2164-8-312 [pii] Soltis DE, Albert VA, Leebens-Mack J et al (2009) Polyploidy and angiosperm diversification. Am J Bot 96(1):336–348 Song H, Yan YL, Titus T, He XJ, Postlethwait JH (2011) The role of stat1b in zebrafish hematopoiesis. Mech Develop 128(7–10):442–456. doi:10.1016/J.Mod.2011.08.004 Star B, Nederbragt AJ, Jentoft S, Grimholt U, Malmstrom M, Gregers TF, Rounge TB, Paulsen J, Solbakken MH, Sharma A, Wetten OF, Lanzen A, Winer R, Knight J, Vogel JH, Aken B, Andersen O, Lagesen K, Tooming-Klunderud A, Edvardsen RB, Tina KG, Espelund M, Nepal C, Previti C, Karlsen BO, Moum T, Skage M, Berg PR, Gjoen T, Kuhl H, Thorsen J, Malde K, Reinhardt R, Du L, Johansen SD, Searle S, Lien S, Nilsen F, Jonassen I, Omholt SW, Stenseth NC, Jakobsen KS (2011) The genome sequence of Atlantic cod reveals a unique immune system. Nature 477(7363):207–210. doi:10.1038/nature10342 nature10342 [pii] CO RR 1431 1432 1433 1434 1435 1436 1437 1438 1439 1440 1441 1442 1443 1444 1445 1446 1447 1448 1449 1450 1451 1452 1453 1454 1455 1456 1457 1458 1459 1460 1461 1462 1463 1464 1465 1466 1467 1468 1469 1470 1471 1472 1473 1474 1475 1476 1477 1478 1479 1480 1481 1482 1483 Book ISBN: 978-3-642-31441-4 Page: 381/383 Polyploidy in Fish and the Teleost Genome Duplication UN Editor Proof 17 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 17 Book ID: 272454_1_En Date: 16-8-2012 EC TE D PR OO F Steinke D, Salzburger W, Braasch I, Meyer A (2006) Many genes in fish have species-specific asymmetric rates of molecular evolution. BMC Genomics 7:20. doi:10.1186/1471-2164-7-20 1471-2164-7-20 [pii] Stellwag EJ (1999) Hox gene duplication in fish. Semin Cell Dev Biol 10(5):531–540. doi:10.1006/scdb.1999.0334 S1084-9521(99)90334-8 [pii] Stoltzfus A (1999) On the possibility of constructive neutral evolution. J Mol Evol 49(2):169– 181. doi:JME1907 [pii] Svärdson G (1945) Chromosome studies on Salmonidae. Rep Swed State Inst Fresh Fish Res 23:1–151 Taylor JS, Van de Peer Y, Braasch I, Meyer A (2001) Comparative genomics provides evidence for an ancient genome duplication event in fish. Philos Trans R Soc Lond B Biol Sci 356(1414):1661–1679. doi:10.1098/rstb.2001.0975 Taylor JS, Braasch I, Frickey T, Meyer A, Van de Peer Y (2003) Genome duplication, a trait shared by 22,000 species of ray-finned fish. Genome Res 13(3):382–390. doi:10.1101/ gr.640303 Tumpel S, Cambronero F, Wiedemann LM, Krumlauf R (2006) Evolution of cis elements in the differential expression of two Hoxa2 coparalogous genes in pufferfish (Takifugu rubripes). Proc Nat Acad Sci USA 103(14):5419–5424. doi:10.1073/pnas.0600993103 0600993103 [pii] Udono T, Yasumoto K, Takeda K, Amae S, Watanabe K, Saito H, Fuse N, Tachibana M, Takahashi K, Tamai M, Shibahara S (2000) Structural organization of the human microphthalmia-associated transcription factor gene containing four alternative promoters. Biochim Biophys Acta 1491(1–3):205–219 Uyeno T, Smith GR (1972) Tetraploid origin of karyotype of catostomid fishes. Science 175(4022):644 Van de Peer Y, Taylor JS, Braasch I, Meyer A (2001) The ghost of selection past: rates of evolution and functional divergence of anciently duplicated genes. J Mol Evol 53(4–5):436– 446. doi:10.1007/s002390010233 Van de Peer Y, Maere S, Meyer A (2009) The evolutionary significance of ancient genome duplications. Nat Rev Genet 10(10):725–732. doi:10.1038/nrg2600 nrg2600 [pii] Vandepoele K, De Vos W, Taylor JS, Meyer A, Van de Peer Y (2004) Major events in the genome evolution of vertebrates: paranome age and size differ considerably between rayfinned fishes and land vertebrates. Proc Nat Acad Sci USA 101(6):1638–1643. doi:10.1073/ pnas.0307968100 0307968100[pii] Vilella AJ, Severin J, Ureta-Vidal A, Heng L, Durbin R, Birney E (2009) EnsemblCompara GeneTrees: complete, duplication-aware phylogenetic trees in vertebrates. Genome Res 19(2):327–335. doi:10.1101/Gr.073585.107 Volff JN (2005) Genome evolution and biodiversity in teleost fish. Heredity (Edinb) 94(3):280– 294. doi:10.1038/sj.hdy.6800635 6800635 [pii] Volff JN, Brunet F, Böhne A, Galiana-Arnoux D (2011) Evolution of fish genomes. In: Farrell AP, Cech JJ, Richards JG, Stevens ED (eds) Encyclopedia of fish physiology: from genome to environment. Elsevier Inc., San Diego Werth CR, Windham MD (1991) A model for divergent, allopatric speciation of polyploid pteridophytes resulting from silencing of duplicate-gene expression. Am Nat 137(4):515–526 Winkler C, Schafer M, Duschl J, Schartl M, Volff JN (2003) Functional divergence of two zebrafish midkine growth factors following fish-specific gene duplication. Genome Res 13(6A):1067–1081. doi:10.1101/gr.1097503 GR-10975R [pii] Wittbrodt J, Meyer A, Schartl M (1998) More genes in fish? BioEssays 20(6):511–515. doi:10.1002/(sici)1521-1878(199806)20:6\511:aid-bies10[3.0.co;2-3 Woltering JM, Durston AJ (2006) The zebrafish hoxDb cluster has been reduced a single microRNA. Nat Genet 38(6):601–602. doi:10.1038/Ng0606-601 Woods TD, Buth DG (1984) High-level of gene silencing in the tetraploid goldfish. Biochem Syst Ecol 12(4):415–421 Woods IG, Kelly PD, Chu F, Ngo-Hazelett P, Yan YL, Huang H, Postlethwait JH, Talbot WS (2000) A comparative map of the zebrafish genome. Genome Res 10(12):1903–1914 CO RR 1484 1485 1486 1487 1488 1489 1490 1491 1492 1493 1494 1495 1496 1497 1498 1499 1500 1501 1502 1503 1504 1505 1506 1507 1508 1509 1510 1511 1512 1513 1514 1515 1516 1517 1518 1519 1520 1521 1522 1523 1524 1525 1526 1527 1528 1529 1530 1531 1532 1533 1534 1535 1536 1537 I. Braasch and J. H. Postlethwait UN Editor Proof 382 Book ISBN: 978-3-642-31441-4 Page: 382/383 Layout: T1 Standard SC Chapter No.: 17 383 EC TE D PR OO F Woods IG, Wilson C, Friedlander B, Chang P, Reyes DK, Nix R, Kelly PD, Chu F, Postlethwait JH, Talbot WS (2005) The zebrafish gene map defines ancestral vertebrate chromosomes. Genome Res 15(9):1307–1314. doi:10.1101/gr.4134305 gr.4134305 [pii] Woolfe A, Elgar G (2007) Comparative genomics using fugu reveals insights into regulatory subfunctionalization. Genome Biol 8(4):R53. doi:10.1186/gb-2007-8-4-r53 gb-2007-8-4-r53 [pii] Yao K, Ge W (2010) Kit system in the zebrafish ovary: evidence for functional divergence of two isoforms of kit (kita and kitb) and kit ligand (kitlga and kitlgb) during folliculogenesis. Biol Reprod 82(6):1216–1226. doi:10.1095/biolreprod.109.082644 biolreprod.109.082644 [pii] Yokoi H, Shimada A, Carl M, Takashima S, Kobayashi D, Narita T, Jindo T, Kimura T, Kitagawa T, Kage T, Sawada A, Naruse K, Asakawa S, Shimizu N, Mitani H, Shima A, Tsutsumi M, Hori H, Wittbrodt J, Saga Y, Ishikawa Y, Araki K, Takeda H (2007) Mutant analyses reveal different functions of fgfr1 in medaka and zebrafish despite conserved ligandreceptor relationships. Dev Biol 304(1):326–337. doi:10.1016/j.ydbio.2006.12.043 S00121606(06)01502-8 [pii] Yokoi H, Yan YL, Miller MR, BreMiller RA, Catchen JM, Johnson EA, Postlethwait JH (2009) Expression profiling of zebrafish sox9 mutants reveals that sox9 is required for retinal differentiation. Dev Biol 329(1):1–15. doi:10.1016/j.ydbio.2009.01.002 S00121606(09)00020-7 [pii] Yu WP, Brenner S, Venkatesh B (2003) Duplication, degeneration and subfunctionalization of the nested synapsin-timp genes in fugu. Trends Genet 19(4):180–183 S0168952503000489[pii] Yuan J, He Z, Yuan X, Jiang X, Sun X, Zou S (2010) Speciation of polyploid Cyprinidae fish of common carp, crucian carp, and silver crucian carp derived from duplicated Hox genes. J Exp Zool B Mol Dev Evol 314(6):445–456. doi:10.1002/jez.b.21350 Zapater C, Chauvigne F, Norberg B, Finn RN, Cerda J (2011) Dual neofunctionalization of a rapidly evolving aquaporin-1 paralog resulted in constrained and relaxed traits controlling channel function during meiosis resumption in teleosts. Mol Biol Evol 28(11):3151–3169. doi:10.1093/molbev/msr146 msr146 [pii] CO RR 1538 1539 1540 1541 1542 1543 1544 1545 1546 1547 1548 1549 1550 1551 1552 1553 1554 1555 1556 1557 1558 1559 1560 1561 1562 1563 1564 1565 Book ISBN: 978-3-642-31441-4 Page: 383/383 Polyploidy in Fish and the Teleost Genome Duplication UN Editor Proof 17 Book ID: 272454_1_En Date: 16-8-2012 Metadata of the chapter that will be visualized in SpringerLink Book Title Polyploidy and Genome Evolution Series Title Chapter Title Polyploidization and Sex Chromosome Evolution in Amphibians Copyright Year 2012 Copyright HolderName Springer-Verlag Berlin Heidelberg Corresponding Author Family Name Evans Particle Given Name Ben J. Suffix Author Division Department of Biology Organization McMaster University Address Life Sciences Building Room 328, 1280 Main Street West, L8S 4K1, Hamilton, ON, Canada Email evansb@mcmaster.ca Family Name Alexander Pyron Particle Given Name R. Suffix Division Department of Biological Sciences Organization The George Washington University Address 2023 G St. NW, 20052, Washington, DC, USA Email Author Family Name Wiens Particle Given Name John J. Suffix Division Department of Ecology and Evolution Organization Stony Brook University Address 11794-5245, Stony Brook, NY, USA Email Abstract Genome duplication, including polyploid speciation and spontaneous polyploidy in diploid species, occurs more frequently in amphibians than mammals. One possible explanation is that some amphibians, unlike almost all mammals, have young sex chromosomes that carry a similar suite of genes (apart from the genetic trigger for sex determination). These species potentially can experience genome duplication without disrupting dosage stoichiometry between interacting proteins encoded by genes on the sex chromosomes and autosomal chromosomes. To explore this possibility, we performed a permutation aimed at testing whether amphibian species that experienced polyploid speciation or spontaneous polyploidy have younger sex chromosomes than other amphibians. While the most conservative permutation was not significant, the frog genera Xenopus and Leiopelma provide anecdotal support for a negative correlation between the age of sex chromosomes and a species’ propensity to undergo genome duplication. This study also points to more frequent turnover of sex chromosomes than previously proposed, and suggests a lack of statistical support for male versus female heterogamy in the most recent common ancestors of frogs, salamanders, and amphibians in general. Future advances in genomics undoubtedly will further illuminate the relationship between amphibian sex chromosome degeneration and genome duplication. Book ISBN: 978-3-642-31441-4 Page: 385/410 Chapter 18 4 Ben J. Evans, R. Alexander Pyron and John J. Wiens 11 12 13 14 15 16 17 18 19 20 21 22 23 D 9 10 TE 8 EC 7 Abstract Genome duplication, including polyploid speciation and spontaneous polyploidy in diploid species, occurs more frequently in amphibians than mammals. One possible explanation is that some amphibians, unlike almost all mammals, have young sex chromosomes that carry a similar suite of genes (apart from the genetic trigger for sex determination). These species potentially can experience genome duplication without disrupting dosage stoichiometry between interacting proteins encoded by genes on the sex chromosomes and autosomal chromosomes. To explore this possibility, we performed a permutation aimed at testing whether amphibian species that experienced polyploid speciation or spontaneous polyploidy have younger sex chromosomes than other amphibians. While the most conservative permutation was not significant, the frog genera Xenopus and Leiopelma provide anecdotal support for a negative correlation between the age of sex chromosomes and a species’ propensity to undergo genome duplication. This study also points to more frequent turnover of sex chromosomes than previously proposed, and suggests a lack of statistical support for male versus female heterogamy in the most recent common ancestors of frogs, salamanders, and amphibians in general. Future advances in genomics undoubtedly will further illuminate the relationship between amphibian sex chromosome degeneration and genome duplication. CO RR 5 6 PR OO 3 Polyploidization and Sex Chromosome Evolution in Amphibians 2 F 1 Book ID: 272454_1_En Date: 16-8-2012 B. J. Evans (&) Department of Biology, McMaster University, Life Sciences Building Room 328, 1280 Main Street West, Hamilton, ON L8S 4K1, Canada e-mail: evansb@mcmaster.ca R. Alexander Pyron Department of Biological Sciences, The George Washington University, 2023 G St. NW, Washington, DC 20052, USA UN Editor Proof Layout: T1 Standard SC Chapter No.: 18 J. J. Wiens Department of Ecology and Evolution, Stony Brook University, Stony Brook, NY 11794-5245, USA P. S. Soltis and D. E. Soltis (eds.), Polyploidy and Genome Evolution, DOI: 10.1007/978-3-642-31442-1_18,  Springer-Verlag Berlin Heidelberg 2012 385 Layout: T1 Standard SC Chapter No.: 18 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 386/410 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 F PR OO 28 29 D 27 Why polyploidization is more common in plants than in animals is a central question in biology (Mable 2004; Muller 1925; Orr 1990), and multiple explanations have been put forward (reviewed in Gregory and Mable 2005; Mable 2004; Orr 1990; Otto and Whitton 2000). One possibility is that the propensity for a species to undergo polyploidization is related to the extent of sex chromosome degeneration. Sex chromosome degeneration is the evolution of differences in gene content that goes beyond the fundamental difference in the presence or absence of a genetic trigger for sex determination. Degenerate sex chromosomes could present problems during ‘‘diploidization’’ of a polyploid genome. Diploidization refers to the phenomenon by which a polyploid species transitions to a mode of chromosomal inheritance that is similar to a diploid species. A key feature of diploidization is the switch from polysomic inheritance, where multivalents form during cell division, to disomic inheritance, where only bivalents form (Wolfe 2001). This phenomenon is probably achieved via divergence between duplicated pairs of homologous chromosomes. Diploidization therefore could be instantaneous when polyploidization occurs via allopolyploidization (genome duplication associated with hybridization among diverged species) because duplicated homologous chromosome pairs already diverged from one another in the ancestral species. When a polyploid genome with duplicated sex chromosomes becomes diploidized, one pair of sex chromosomes presumably begins to segregate autosomally. With a degenerate Y-chromosome, for instance, the nascent autosomal pair that was previously a pair of sex chromosomes would initially have three possible genotypes: AXAX, AX0, and 00 where AX refers to an autosomal allele derived from an ancestral X-chromosome and 0 refers to a missing allele that was lost on the ancestral Y-chromosome. If the 00 genotype is deleterious or lethal, there would be reproductive incompatibilities in the early stages of diploidization until the degenerate chromosome is lost. This fitness cost could be mitigated if a functional paralogous allele were still present on the sex chromosomes, as would be expected in an autopolyploid (formed from genome duplication within a species). In allopolyploids, however, sex chromosomes could degenerate in unique ways in each ancestral species, giving rise to diverged gene content, so homozygous null genotypes would be a bigger problem. In both types of polyploids, dosage balance requirements—natural selection favoring a specific relative expression level of interacting genes (Papp et al. 2003; Qian and Zhang 2008)— could impose a fitness cost on a homozygous or heterozygous null autosomal genotype in a polyploid. Polyploidization could also present challenges to species with degenerate sex chromosomes that have also evolved mechanisms for dosage compensation (Orr 1990). Dosage compensation equilibrates expression levels of genes that have one allele in one sex and two alleles in the other sex (for example, X-linked genes have one allele in XY males but two alleles in XX females). In this way the stoichiometry of expression of X-linked and autosomal genes is constant, or ‘‘balanced’’, in males TE 26 EC 25 18.1 Introduction CO RR 24 B. J. Evans et al. UN Editor Proof 386 Layout: T1 Standard SC Chapter No.: 18 Book ISBN: 978-3-642-31441-4 Page: 387/410 Polyploidization and Sex Chromosome Evolution in Amphibians 387 101 18.1.1 Sex Chromosome Evolution 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 102 103 104 105 106 107 PR OO 73 D 72 TE 71 EC 70 CO RR 69 F 100 and females. Orr (1990) argued that a polyploid lineage would initially be established via backcrossing a new polyploid individual to diploid individuals, and that this would disrupt this balance. The evolution of differences in gene content is a combined consequence of the migration of genes, especially genes with sex-specific function, to one or the other sex chromosome, and also the loss of genes from the region of suppressed recombination on the sex-specific sex chromosome (for example, the Y-chromosome). The disparity in gene content between the sex chromosomes is thought to increase over time as a consequence of natural selection (Bergero and Charlesworth 2009; Charlesworth et al. 2005). Substantial disparity in gene content between the sex chromosomes could be coupled with selective pressure favoring the evolution of mechanisms of dosage compensation. For this reason, the proposal that degenerate sex chromosomes deter polyploidization (including species that lack dosage compensation) is not independent of Orr’s (1990) proposal that dosage compensation deters polyploidization. In either case, if sex chromosome degeneration acts as a barrier to polyploidization, this would predict that polyploid species or species with polymorphism in ploidy levels would have relatively ‘‘young’’, minimally degenerate sex chromosomes as compared with other species. In this chapter we briefly review polyploidization in frogs and salamanders and general features of sex chromosome evolution. Using previously published information, we then use a maximum likelihood approach to analyze the evolution of new sex-determining mechanisms in frogs and salamanders in a phylogenetic context, where new mechanisms are inferred either from a change in which chromosomes are the sex chromosomes, the evolution of a new trigger for sex determination, or from observed polymorphism in sex-determining mechanisms. Following this, we explore whether polyploidization occurs more frequently soon after a new sex-determining mechanism evolves using a permutation test that accommodates uncertainty in ancestral reconstruction. We conclude that novel sex-determining mechanisms have evolved in amphibians even more frequently than previously proposed, and that amphibians with young sex chromosomes may be more likely to experience genome duplication, resulting either in polyploid speciation or in spontaneous polyploidy of individuals of an otherwise diploid species. A major caveat to the latter result is that information on the age of sexdetermining mechanisms of most polyploid amphibians is lacking. 67 68 Sex chromosomes originate from autosomes (Ohno 1967) but differ in carrying genetic information that (a) differs between the sexes and (b) triggers or represses sex-specific gonadal differentiation. The ‘‘heterogametic’’ sex produces two types of gametes, each type with a different sex chromosome, and the ‘‘homogametic’’ sex produces only one type of gamete with respect to the sex chromosomes. The sex chromosomes of species with male heterogamy are called ‘‘X’’ and ‘‘Y’’ (females UN Editor Proof 18 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 18 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 388/410 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 F 113 PR OO 112 D 111 TE 110 have two X chromosomes and males have an X and a Y), and the sex chromosomes of species with female heterogamy are called ‘‘Z’’ and ‘‘W’’ (males have two Z chromosomes, and females have a Z and a W). In species with genetic sex determination, gonadal differentiation—also known as primary sex determination—is achieved either using a sex chromosome-specific genetic trigger or by gene dosage, where the homogametic sex carries two doses of an activator of that sex, or a repressor of the heterogametic sex. The age of sex chromosomes influences important aspects of their evolution and divergence, including divergence in gene content, and the origin of dosage compensation. For example, the sex chromosomes of therian (placental and marsupial) mammals are extremely old ([180 million years; Graves 2008), and the Y chromosome is much smaller than the X and carries fewer genes. This disparity in size and gene content arose due to Y chromosome ‘‘degeneration’’ as a consequence of suppressed recombination with the X chromosome (Charlesworth and Charlesworth 2000). Suppressed recombination ensures that male progeny inherit an intact copy of the genetic trigger for testis formation, which in therians is the SRY gene. But this also permits deleterious mutations to accumulate in Y-linked genes (Muller 1964), leading to loss of function and deletion. ‘‘Muller’s ratchet’’, the stochastic loss of the least deleterious allele in a population (Felsenstein 1974), leads to a decline in fitness. This decline occurs more quickly in regions of the genome that do not recombine. Hill-Robertson effects, background selection, and hitchhiking of deleterious alleles also contribute to fitness declines of non-recombining portions of the genome (reviewed in Charlesworth and Charlesworth 2000). Degeneration of the therian Y-chromosome occurred in a stepwise fashion as the region of suppressed recombination expanded in large increments (Skaletsky et al. 2003). In therians the disparity in gene content increased over time after the origin of SRY and associated suppression of recombination between the X and Y chromosomes. Most angiosperm plants that have separate sexes (dioecy), in contrast, have comparatively young sex chromosomes that are not substantially differentiated, although exceptions exist (Bergero and Charlesworth 2011; Charlesworth 2002). Polyploid species are prevalent in angiosperms (Otto and Whitton 2000) but absent in therians (Svartman et al. 2005), and these observations thus provide anecdotal support for the contention that the extent of sex chromosome degeneration is negatively correlated with the incidence of polyploid speciation. Amphibians offer an interesting focal group with which to further evaluate this hypothesis because some features of amphibian genome evolution resemble plants more than other animal groups such as therian mammals. In particular, unlike therian mammals, sex chromosomes in many amphibians are relatively young, chromosome degeneration is modest or absent, and polyploidization is fairly common. Species that determine sex exclusively using environmental triggers do not have genomic differences between the sexes and therefore have no sex chromosomes. In amphibians, sex determination is genetic so all species are expected to have sex chromosomes. In addition, temperature has been reported to influence offspring sex ratios of various species of the salamander genera Pleurodeles and EC 109 CO RR 108 B. J. Evans et al. UN Editor Proof 388 Layout: T1 Standard SC Chapter No.: 18 159 160 161 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 184 185 186 187 188 189 190 191 192 193 F 158 PR OO 157 D 156 Hynobius and the frog genera Bufo, Rana, and laboratory-generated polyploids of the genus Xenopus (Hayes 1998; Kobel 1996; Schmid and Steinlein 2001). Sex chromosomes are cytologically distinct in some amphibian species (Schmid et al. 2010). Differences in gene content between the sex chromosomes, which is suggested by cytologically distinct sex chromosomes, led to the independent evolution of dosage compensation mechanisms in placental mammals, birds, and other groups such as Drosophila and Caenorhabditis (Arnold et al. 2008; Straub and Becker 2007). However, in amphibians evidence of dosage compensation has not been found (Hayes 1998; Schmid et al. 1986; Schmid and Steinlein 2001). One possible reason for this is that we do not yet know the identity of any amphibian genes that are hemizygous in the heterogametic sex, so a rigorous test for dosage compensation in amphibians is not yet possible. These genes would be restricted to the portion of the X or Z chromosome that does not recombine with the Y or W chromosome, respectively. A recent study of European tree frogs identified one way that amphibians circumvent sex chromosome degeneration (Stöck et al. 2011). In three species, no recombination occurred between the sex chromosomes in males generated from intraspecific crosses, yet no intraspecific sex chromosome divergence was observed (Stöck et al. 2011). This suggests that Muller’s ratchet is periodically reset in these species by infrequent recombination between the sex chromosomes. Sex chromosome degeneration can also be circumvented by genomic translocation of the sexdetermining locus to another chromosomal pair, or by re-assignment of the sexdetermining function to a gene located elsewhere in the genome. Both of these phenomena result in a change in which chromosomes are the sex chromosomes (hereafter ‘‘sex chromosome turnover’’). Ancient examples of sex chromosome turnover are evinced in amniotes by homology between the sex chromosomes of platypuses and those of birds but not those of therian mammals (Graves 2008). In amphibians, sex chromosome turnover is common and is suggested by variation among and within species in male versus female heterogamy (Ezaz et al. 2006; Hillis and Green 1990). Using maximum parsimony, Hillis and Green (1990) analyzed variation in male and female heterogamy in amphibians in a phylogenetic context and concluded that sex chromosome turnover occurred at least seven times. TE 155 389 EC 154 CO RR 153 Book ISBN: 978-3-642-31441-4 Page: 389/410 Polyploidization and Sex Chromosome Evolution in Amphibians 18.1.2 How Many Frog and Salamander Species are Polyploid? Comprehensive reviews of polyploidization in amphibians are available in Bogart (1980), Kawamura (1984), Duellman and Trueb (1994), Beçak and Beçak (1998), Otto and Whitton (2000), Gregory and Mable (2005), Schmid et al. (2010), and Mable et al. (2011). The two most recent of these reviews have up-to-date lists of known polyploid species and associated citations that document polyploidy. Schmid et al. (2010) also summarize male and female heterogamy in frogs and salamanders, including information on species with cytologically detectable sex UN Editor Proof 18 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 18 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 390/410 201 202 203 204 205 206 207 208 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 236 237 238 F PR OO 200 D 198 199 TE 197 EC 196 chromosome divergence (their Table 8, pp 160–161). In their Supplementary Information, Mable et al. (2011) provide data on confirmed diploid species that are closely related to the polyploids. A key difference between these two reviews is that Mable et al. (2011) include only bisexually reproducing polyploids whereas Schmid et al. (2010) also include unisexual polyploids. We have attempted to compile this information as inclusively as possible in Table 18.1, including some minor corrections and a few additional species and associated citations. Thus, not all polyploids listed in this table are bisexual, and some are diploid species in which polyploid individuals occur spontaneously or by induction due to laboratory manipulation. Fifty polyploid frog species have been described, including seven triploids, 30 tetraploids, 11 octoploids, and two dodecaploids derived from 15 families and 20 genera (Table 18.1). Three tetraploids and two dodecaploids have been reported from the genus Xenopus but not yet formally described as species (Evans 2007, 2008; Evans et al. 2004a, 2005a; Tymowska 1991). Stable triploids are known from three frog genera (Bufo, Eupsophus, and Rana), tetraploids from 16 (Aphantophryne, Astylosternus, Bufo, Chiasmocleis, Dicroglossus, Eleuthrodactylus, Hyla, Neobatrachus, Odontophrynus, Phyllomedusa, Pleurodema, Pyxicephalus, Scaphiophryne Silurana, Tomopterna, and Xenopus), octoploids from three (Ceratophrys, Pleurodema, and Xenopus), and dodecaploids only from Xenopus. Spontaneous or experimentally induced polyploidy has been reported in at least five frog species. Six polyploid species of salamander, including four triploids and two tetraploids, are known from only two genera (Ambystoma and Siren) from two families (Table 18.1). Spontaneous or experimentally induced triploidy or tetraploidy has been reported in eight salamander species. The origin of polyploidy necessarily is preceded by the existence of one or more diploid ancestors. Interestingly, a number of polyploid frog species are inferred to have originated from ancestral diploid species that do not have known extant diploid descendants. In Xenopus and Silurana, for example, three currently unknown diploid species contributed their genomes to extant tetraploid species (reviewed in Evans 2008). There are also three currently unknown tetraploid species that contributed their genomes to extant octoploid and dodecaploid Xenopus species (reviewed in Evans 2008). The tetraploid Hyla versicolor is thought to be derived from multiple independent allopolyploidization events between three diploid species, two of which are currently unknown, and probably extinct given that the region in which they occur (temperate North America) is well studied (Holloway et al. 2006). In Ceratophrys, there are no known tetraploid species even though three species in this genus are octoploid (Table 18.1). Similarly, the tetraploid species Bufo pewzowi is thought to be derived from the diploid B. turanensis and another unidentified diploid (Stöck et al. 2009), and various tetraploid species of Neobatrachus are derived from diploid ancestors whose diploid descendants are currently unknown (Mable and Roberts 1997). It is tempting to speculate from these observations that polyploidization contributed to the long-term survival of these lineages, given that the diploid ancestors of extant polyploids seem to have gone extinct in many cases. However, we lack information on how frequently CO RR 194 195 B. J. Evans et al. UN Editor Proof 390 Layout: T1 Standard SC Chapter No.: 18 Book ISBN: 978-3-642-31441-4 Page: 391/410 Polyploidization and Sex Chromosome Evolution in Amphibians 391 Table 18.1 A list of known polyploid amphibians compiled primarily from Schmid et al. (2010) , Mable et al. (2011) and citations therein Family and species Ploidy Family and species Ploidy (Frogs) (Frogs continued) Tetraploid Tetraploid Tetraploid Tetraploid F Myobatrachidae Neobatrachus aquilonius Neobatrachus centralis Arthroleptidae Neobatrachus sudelli Astylosternus diadematus Tetraploid Neobatrachus kunapalari Microhylidae Scaphiophryne gottlebei Bufonidaea Bufo asmaerae / Tetraploid Aphantophryne(Cophixalus) Amietophrynus pansa asmaerae Bufo baturae Triploid Chiasmocleis leucosticta Bufo kerinyagae Tetraploid Odontophrynidae Bufo oblongus and Tetraploid Odontophrynus americanus subspecies (synonym of ‘‘B. danatensis’’) Bufo pewzowi and Tetraploid Pipidae subspecies (also a synonym of ‘‘B. danatensis’’) Bufo poweri Triploid Silurana epitropicalis Triploid ‘‘Silurana new tetraploid 1’’ Bufo pseudoraddeia (‘‘Silurana sp. Nov VII’’, ‘‘Silurana paratropicalis’’)b,c Bufo viridis Triploid ‘‘Silurana new tetraploid 2’’b a Triploid Xenopus borealis Bufo zugmayeri Xenopus clivii Ceratophryidae Xenopus fraseri Tetraploid Xenopus gilli Craugastoridae Eleutherodactylus (Haddadus) binotatus Ceratophrys dorsata / Octoploid Xenopus laevisd Ceratophrys aurita Ceratophrys joazeirensis Octoploid Xenopus muelleri Ceratophrys ornata Octoploid Xenopus pygmaeus Xenopus largeni (‘‘Xenopus sp. Nov. III’’)e Triploid PR OO Alsodidae Eupsophus vertebralis Tetraploid Tetraploid Tetraploid Tetraploid D TE EC CO RR UN Editor Proof 18 Book ID: 272454_1_En Date: 16-8-2012 Tetraploid Tetraploid Tetraploid Tetraploid Tetraploid Tetraploid Tetraploid Tetraploid Tetraploid Tetraploid Tetraploid (continued) Layout: T1 Standard SC Chapter No.: 18 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 392/410 B. J. Evans et al. Family and species (Frogs continued) Dicroglossus (Hoplobotrachus) occipitalis Tetraploid ‘‘Xenopus new tetraploid 1’’ (‘‘Xenopus sp. Nov. VI’’,’’ Xenopus muelleri west’’)e,f Xenopus amieti Tetraploid Tetraploid Xenopus andrei Xenopus boumbaensis Xenopus itombwensisg Xenopus lenduensish Xenopus vestitus Xenopus wittei ‘‘Xenopus sp. nov. X’’e,i Hylidae Ploidy Octoploid Octoploid Octoploid Octoploid Octoploid Octoploid Octoploid Octoploid Dodecaploid Dodecaploid Dodecaploid Dodecaploid Tetraploid Tetraploid Ranidae Rana esculenta Rana japonica*k Triploid Triploid Triploid / tetraploid Rana nigromaculata** Triploid / Triploid Triploid Triploid Rana pipiens*l Rana rugosa*m Triploid Triploid Triploid CO RR Leiopelmatidae Leiopelma hochstetteri** Family and species (Salamanders) Ambystomatidae Ambystoma jeffersonianum Ambystoma mexicanum** tetraploid Ambystoma nothagenesn Ambystoma platineumn Ambystoma tremblayin Plethodontidae Eurycea bislineata** Xenopus longipes Xenopus ruwenzoriensis Tetraploid ‘‘X. cf. boumbaensis’’j Octoploid ‘‘Xenopus sp. Nov. VIII’’e,i Tetraploid Pyxicephalidae Pyxicephalus (Tomopterna) delalandii Tomopterna tandyi Triploid D Leptodactylidae Pleurodema bibroni Pleurodema cordobae Pleurodema kriegi EC Hyla versicolor Phyllomedusa tetraploidea Tetraploid PR OO Dicroglossidae Ploidy F Ploidy TE Table 18.1 (continued) Family and species (Frogs) UN Editor Proof 392 Triploid / tetraploid (continued) Layout: T1 Standard SC Chapter No.: 18 18 Book ISBN: 978-3-642-31441-4 Page: 393/410 Polyploidization and Sex Chromosome Evolution in Amphibians Table 18.1 (continued) Family and species (Frogs) Editor Proof Book ID: 272454_1_En Date: 16-8-2012 Ploidy Family and species (Frogs continued) 393 Ploidy Salamandridae F Triploid Triploid Triploid Triploid PR OO Pleurodeles waltl* Ichthyosaura alpestris* Cynops pyrrhogaster**o Notophtalamus viridescens**p Lissotriton vulgaris**q Sirenidaer Siren intermedia Siren lacertina Triploid Tetraploid Tetraploid CO RR EC TE D Additional citations are provided for examples not included in these references, and for unnamed species * Experimentally induced ** Spontaneously observed, in some cases also experimentally induced a Following taxonomy of Stöck et al. (2009). Further work is needed to confirm stable triploid ploidy of B. pseudoraddei and B. zugmayeri (Stöck et al. 2009) b Evans et al. (2004b) c ‘‘Silurana paratropicalis’’ is a nomen nudem; see Blackburn DC and Beier M (2011) d Here we consider as X. laevis all diverged populations within this clade as identified by Evans et al. (2004). This includes Xenopus sp. Nov. IX (a.k.a. X. congo 3) from Tymowska (1991), X. petersi, and X. victorianus e Tymowska (1991) f Kobel et al. (1996) g Table 13 of Schmid et al. (2010) incorrectly lists Xenopus itombwensis as a dodecaploid h Evans et al. (2011) i Species status requires further investigation j Evans (2007) k Kawamura and Tokunaga (1952) l Briggs (1947) m Kashiwagi (1993) n These names refer to individuals that are not species in the sense of being reproductively isolated lineages that persist through time. They are unisexual progeny resulting from hybridization of other species o Fankhauser et al. (1942) p Fankhauser (1941); Fankhauser and Watson (1942) q Litvinchuk et al. (1998) r Morescalchi and Olmo (1974) found Pseudobranchus striatus to be polyploid but this result was not supported by the analysis of Moler and Kezer (1993) 240 241 242 243 polyploidization occurs and how frequently diploids outcompete polyploids, so it is difficult to test this. It is also plausible, for example, that variation in ploidy level is a neutral phenomenon influenced by stochastic survival and extinction of polyploids and diploids, or by variation among lineages, including polyploids (Mayrose et al. 2011), in their ability to speciate by polyploidization. UN 239 Layout: T1 Standard SC Chapter No.: 18 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 394/410 250 251 252 253 254 255 256 257 258 259 260 261 262 263 264 265 266 267 268 269 270 271 272 273 274 275 276 277 278 279 280 281 282 283 284 285 286 F 249 PR OO 248 About one third of the described polyploid frog species belong to the genus Xenopus. At least six independent instances of genome duplication gave rise to the ploidy levels seen among extant species in this group, including multiple episodes that generated the highest ploidy level of any vertebrate—dodecaploidy (reviewed in Evans 2008). Tetraploid Xenopus evolved at least once, octoploid Xenopus evolved independently at least three times, and dodecaploid Xenopus evolved independently at least two times (and possibly more depending on the species status of Xenopus cf. boumbaensis and of Xenopus sp. nov. VIII; Table 1, Evans 2007, 2008; Evans et al. 2008a, 2011, 2005a; Tymowska 1991). Tetraploidy also occurred independently in Silurana (Evans 2007; Evans et al. 2005a). With respect to genome duplication, something is clearly special about Xenopus— but what? One possible clue emerges from the recent discovery of the first known genetic trigger of sex determination in amphibians by Yoshimoto et al. (2008). These researchers identified a female-specific gene called DMW in the tetraploid species Xenopus laevis. DMW is a W-chromosome linked gene that evolved via gene duplication from another important regulator of sexual differentiation called DMRT1 (Yoshimoto et al. 2008). DMW may function by blocking DMRT1 induction of testis differentiation (Yoshimoto et al. 2010, 2008). Potentially relevant to the high incidence of polyploidization in Xenopus is the discovery that DMW originated extremely recently in amphibian evolution—after divergence of Silurana and Xenopus, but before diversification of most or all extant species of Xenopus (Bewick et al. 2011). Not surprisingly, the sex chromosomes of Xenopus are not cytologically distinct (Tymowska 1991). Gene contents of the W and Z chromosomes of Xenopus are therefore probably very similar, and Xenopus species presumably lack mechanisms of dosage compensation operating over most sex-linked genes because both sexes have two alleles at most loci on the sex chromosomes. The preponderance of polyploids in Xenopus is therefore consistent with the proposal that polyploidization is more likely to occur in lineages with young, minimally degenerate sex chromosomes. Another possible link between sex chromosome evolution and polyploidization is provided by Leiopelma hochstetteri. This species has intraspecific variation in the presence of a recently evolved univalent W chromosome that governs sex determination in females (Green 1988). Leiopelma hochstetteri is diploid but also has spontaneous triploidy (that is, polyploidy without speciation; Green et al. 1984). It is not clear whether novel mechanisms for sex determination are more likely to evolve and persist in species that have nondegenerate sex chromosomes, but this seems plausible under the same reasoning discussed above with respect to the propensity for lineages to experience polyploidization. More specifically, if a new pair of sex chromosomes appears in a population then the old ones would segregate as a newly established autosomal pair. For this reason, ancestral sex chromosomes with similar gene content would lack or have few null alleles when they segregate autosomally. While the observation of D 247 TE 246 EC 245 18.1.3 Examples of polyploidy in species with demonstrably young sex chromosomes CO RR 244 B. J. Evans et al. UN Editor Proof 394 Layout: T1 Standard SC Chapter No.: 18 Book ISBN: 978-3-642-31441-4 Page: 395/410 Polyploidization and Sex Chromosome Evolution in Amphibians 395 291 18.2 Evolution of Sex Determination Systems in Amphibians 292 18.2.1 Methods 296 297 298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 PR OO 295 Changes in the heterogametic sex, evolution of new triggers for sex determination, and polymorphism in sex chromosomes mark the origin of novel features in genetic pathways for sex determination. In order to quantify how many times this has happened in amphibians, we began with the large amphibian phylogeny reported by Pyron and Wiens (2011). We trimmed from this tree all species except those for which we had information on either heterogamy or polyploidy, or both, and retained the original maximum likelihood branch lengths among the retained species. For illustrative purposes, we also retained diploid species (confirmed or presumed) from the phylogeny of Pyron and Wiens (2011) that are closely related to polyploid species. In many cases diploidy has been confirmed in these species or other closely related species (see Supplementary Information of Mable et al. 2011). We had heterogamy information for Physalaemus (Engystomops) petersi, but this species was not included in the phylogeny. Therefore, we used a closely related species (Physalaemus cuvieri) that is included in the phylogeny to represent Physalaemus petersi. Physalaemus petersi was the only species from this genus that was analyzed, so this substitution should be uncontroversial (note that placing some Physalaemus in Engystomops makes no difference as Engystomops and Phsalaemus are sister taxa). To better illustrate the phylogenetic distribution of polyploid species in Fig. 18.1, we also substituted Chiasmocleis hudsoni, which was present in the phylogeny of Pyron and Wiens (2011), with the tetraploid species C. leucosticta, which was not present in the phylogeny of Pyron and Wiens (2011). However, C. leucosticta was not included in any of the analyses described below because we lack data on heterogamy for this species. A total of 143 species (97 frogs, 45 salamanders, and one caecilian as an outgroup) were included. We then converted this tree to a chronogram (a timecalibrated phylogeny) using the penalized likelihood approach (Sanderson 2002), implemented in r8s version 1.71 (Sanderson 2003). We used the calibration points detailed in Wiens (2011) but some had to be excluded given the more limited taxon sampling used here, and some were added or modified given the differences in taxon sampling (e.g., we added two calibration points within the genus Hyla, given our more extensive sampling of species in that genus relative to Wiens (2011). We used the following 17 calibration points. The first 16 were treated as D 294 TE 293 EC 289 CO RR 288 F 290 spontaneous triploidy suggests a tolerance of polyploidy, L. hochstetteri is not a polyploid species, so a direct link between the age of the sex chromosomes and polyploid speciation (as opposed to the toleration of polyploidy) is not established by this species. 287 UN Editor Proof 18 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 18 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 396/410 B. J. Evans et al. Necturus punctatus Necturus alabamensis Necturus maculosus Necturus beyeri Necturus lewisi Aneides ferreus Hydromantes flavus Hydromantes supramontis Hydromantes imperialis Hydromantes italicus Hydromantes ambrosii Eurycea junaluska Eurycea bislineata Nototriton abscondens Nototriton picadoi Nototriton richardi Oedipina parvipes Oedipina pseudouniformis Oedipina cyclocauda Oedipina poelzi Cryptotriton veraepacis Dendrotriton rabbi Chiropterotriton dimidiatus Thorius dubitus Pleurodeles waltl Pleurodeles poireti Triturus marmoratus Triturus pygmaeus Triturus dobrogicus Triturus karelinii Triturus cristatus Triturus carnifex Notophthalmus perstriatus Notophthalmus viridescens Ambystoma laterale Ambystoma mexicanum Ambystoma dumerilii Ambystoma tigrinum Pseudobranchus striatus Pseudobranchus axanthus Siren intermedia Siren lacertina Hynobius quelpaertensis Hynobius tokyoensis Hynobius hidamontanus Leiopelma hamiltoni Leiopelma pakeka Leiopelma archeyi Leiopelma hochstetteri Bombina orientalis Discoglossus pictus Xenopus gilli Xenopus laevis Silurana epitropicalis Silurana tropicalis Hymenochirus boettgeri Pelodytes punctatus Rana clamitans Rana catesbeiana Rana pipiens Rana sphenocephala Rana blairi Rana berlandieri Rana tagoi Rana japonica Rana temporaria Rana tsushimensis Rana rugosa Rana esculenta Rana lessonae Rana ridibunda Rana plancyi Rana nigromaculata Buergeria buergeri Hoplobatrachus crassus Hoplobatrachus occipitalis Tomopterna cryptotis Tomopterna tandyi Tomopterna delalandii Pyxicephalus adspersus Pyxicephalus edulis Aphantophryne pansa Cophixalus humicola Scaphiophryne spinosa Scaphiophryne gottlebei Chiasmocleis leucosticta Chiasmocleis shudikarensis Astylosternus batesi Astylosternus diadematus Strabomantis biporcatus Pristimantis pulvinatus Pristimantis shrevei Pristimantis euphronides Eleutherodactylus johnstonei Eleutherodactylus glamyrus Eleutherodactylus cavernicola Eleutherodactylus turquinensis Eleutherodactylus cuneatus Eleutherodactylus casparii Eleutherodactylus emiliae Eleutherodactylus albipes Physalaemus petersi Pleurodema thaul Pleurodema bibroni Pseudopaludicola falcipes Vitreorana antisthenesi Bufo boulengeri Bufo viridis Bufo balearicus Bufo oblongus Bufo pewzowi Bufo bufo Bufo regularis Bufo poweri Eupsophus calcaratus Eupsophus roseus Eupsophus insularis Eupsophus migueli Eupsophus vertebralis Proceratophrys boiei Odontophrynus cultripes Odontophrynus americanus Ceratophrys cranwelli Ceratophrys ornata Phyllomedusa bahiana Phyllomedusa burmeisteri Phyllomedusa tetraploidea Phyllomedusa distincta Hyla femoralis Hyla chrysoscelis Hyla versicolor Hyla arborea Hyla intermedia Hyla sarda Hyla meridionalis Pseudis tocantins Gastrotheca riobambae Gastrotheca pseustes Gastrotheca walkeri Crinia signifera Neobatrachus pictus Neobatrachus sudelli PR OO F Editor Proof 396 Salamanders CO RR EC TE D Frogs UN Fig. 18.1 Heterogamy and polyploidy in salamanders and frogs. The phylogeny is from Pyron and Wiens (2011) with branch lengths proportional to time based on a relaxed molecular clock and calibration points described in the text. A black scale bar indicates 40 million years of evolution. Red and blue rectangles on tips indicate male or female heterogamy respectively; green, orange, and yellow rectangles indicate de novo sex determining systems; missing data on heterogamy have no rectangles in this column. In the right column, black rectangles indicate polyploid species; other species are either confirmed or assumed diploid. Some polyploid species listed in Table 18.1 are not included in this figure due to a lack of phylogenetic information. Pie charts on nodes indicate ancestral reconstructions of heterogametic state Layout: T1 Standard SC Chapter No.: 18 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 F 330 PR OO 329 (1) Most recent common ancestor (MRCA) of extant salamanders, at least 150.8 Mya (Millions of years ago), based on the fossil Iridotriton hechti of the Kimmeridgian/Early Tithonian (Late Jurassic), which is considered to be a crown-group caudate (Evans et al. 2005b). (2) MRCA of Salamandroidea (all salamanders exclusive of cryptobranchids, hynobiids, and sirenids), at least 125.0 Mya (early Barremian, Cretaceous), based on Galverpeton and Valdotriton (Evans and Milner 1996). (3) MRCA of plethodontids and proteids, at least 65.5 Mya. The oldest known amphiumid fossil (Proamphiuma cretacea) is late Maastrichtian or early Paleocene, and thus from approximately 65.5 Mya (Gardner 2003). The split between Plethodontidae and Amphiumidae must be at least this old. We do not have amphiumids included here, but the clade of plethodontids and proteids must be at least this old given the well-supported clade consisting of proteids, rhyacotritonids, amphiumids, and plethodontids; see Pyron, Wiens (2011) and earlier studies. (4) MRCA of Aneides and Hydromantes, at least 19 Mya. Given the presence of an Aneides vertebra in the Arikareean period (Tihen and Wake 1981), the MRCA of the clade containing modern Aneides must be at least 19 Myo (Millions of years old). (5) MRCA of Triturus and Notophthalmus at least 33.9 Mya, based on fossils of Triturus from the Eocene of Europe (33.9–55.8 Mya; Milner 2000) (6) MRCA of Ambystomatidae and Salamandridae, at least 56.8 Mya, based on a fossil dicamptodontid (Paleocene; Tiffanian; 60.2–56.8 Mya; Naylor and Fox 1993), and given that the ambystomatidae is the sister group to the Dicamptodontidae (so the sister group to Ambystomatidae ? Dicamptodontidae must be at least this old). (7) MRCA of frogs and salamanders, at least 245 Mya, based on a fossil anuran (Triadobatrachus) from the Early Triassic (251–245 Mya) of Madagascar (Carroll 1988; Rage and Rocek 1989) (8) MRCA of pipoids and all other frogs, at least 145.5 Mya, given Rhadinosteus parvus, ostensibly a rhinophrynid and clearly a pipoid, from the Late Jurassic (Tithonian, 145.5–150.8 Mya; Rocek 2000). (9) MRCA of Hymenochirus and Xenopus, at least 83.5 Mya, given the pipid Pachybatrachus taqueti from the Upper Cretaceous (Coniacian-Santonian, 83.5–89.3 Mya), which is thought to be closely related to Hymenochirus (Rocek 2000). (10) MRCA of Myobatrachidae (represented here by the limnodynastine Neobatrachus and the myobatrachine Crinia) at least 54.6 Mya, given fossils assigned to the limnodynastine genus Lechriodus (Evans et al., 2008b; Sanchiz, 1998). (11) MRCA of Bufonidae ? Leptodactylidae ? Centrolenidae, at least 55.8 Mya, given putative fossil Bufo from the late Paleocene (55.8–58.7 Mya; Baéz 2000). D 328 TE 327 397 constraints on the minimum age of each clade, and the final calibration point was a fixed age for the root of the tree. EC 326 CO RR 325 Book ISBN: 978-3-642-31441-4 Page: 397/410 Polyploidization and Sex Chromosome Evolution in Amphibians UN Editor Proof 18 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 18 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 398/410 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 408 407 409 410 411 412 F 374 PR OO 373 D 372 TE 371 EC 370 (12) MRCA of Ranidae (sensu Wiens et al. 2009) at least 33.9 Mya, given fossil Rana from the Late Eocene (37.2–33.9 Mya; Rocek and Rage 2000). (13) Crown group of Terrarana (the clade including the families Brachycephalidae, Ceuthomantidae, Craugastoridae, Eleutherodactylidae, and Strabomantidae, or more simply, the clade including Eleutherodactylus and related genera) at least 35 Mya, based on an Eleutherodactylus fossil in amber from the La Toca formation (Dominican Republic) estimated to be *35 Myo (Poinar and Cannatella 1987). (14) Stem group of Ceratophryidae (the clade including the genera Ceratophrys, Chacophrys, and Lepidobatrachus) at least 65.5 Mya based on the late Cretaceous fossil genera Beelzebufo and Baurubatrachus (Evans et al. 2008b). Evans et al. (2008b) considered the Madagascan taxon Beelzebufo to be a ceratophryine. This taxon is of Maastrichtian (Late Cretaceous) age (65.5–70.6 Mya). The South American genus Baurubatrachus is also considered to be a ceratophryine (Evans et al. 2008b; Rocek 2000). Although the exact relationships of these taxa are somewhat uncertain, the presence of seemingly ceratophryine fossils in South America suggests that the stem group age of Ceratophryidae is at least 65.5 Mya. The relationships of ceratophryids are uncertain, but in this molecular analysis, they appear as the sister group to a clade including Odontophrynidae and Alsodidae (Eupsophus). (15) Crown-group age of North American and European Hyla clade, at least 16 Myo; given fossil Hyla similar to extant H. arborea and H. meridionalis in the Lower Miocene of Austria (*16 Myo; Sanchiz 1998). We assume that these Hyla are closely related to Hyla presently extant in Europe. However, we cannot assume that these fossils are younger than the crown-group age of the extant European species. We assume instead that the crown group of the clade of Hyla is at least 16 Myo based on these European fossils. (16) MRCA of H. avivoca-H. chrysocelis-H. versicolor clade; H. miocenica is thought to be closely related to H. chrysocelis and H. versicolor and occurs in the Barstovian of the Middle Miocene (14–16 Myo; Holman 2003). In our phylogeny, H. avivoca, H. chrysocelis, and H. versicolor form a clade. We assume that the stem group age of these three species is at least 14 Myo. (17) We fixed the root age of the tree using the estimated age from Wiens (2011) for the MRCA of lissamphibians (frogs, salamanders, caecilians) of 368.3 Mya, using penalized likelihood. Although the use of a fixed calibration point (rather than a minimum constraint) may seem controversial, it should be noted that at least one node must be given a fixed age. Furthermore, our focus here is not on re-estimating these ages, but providing relative assessments of clade ages (see below). CO RR 368 369 B. J. Evans et al. UN Editor Proof 398 Construction of a chronogram using r8s requires a cross-validation step that identifies a best-fitting value for the ‘‘smoothing parameter’’, which specifies the cost of differing rates of evolution between neighboring branches (Sanderson 2002). Cross-validation considered smoothing parameter values from 100 –105.5 in exponential increments of 0.5. These cross-validation analyses failed until a species with a Layout: T1 Standard SC Chapter No.: 18 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 F 418 PR OO 417 D 416 zero-length branch (Bufo pewzoi) was removed. After removing this species, the cross-validation analyses showed that a value of 101 gave the lowest Chi-squared error. We then generated a chronogram for the 142 remaining species using this smoothing parameter. We then performed a second analysis using this same smoothing parameter but including all 143 species. This second analysis gave identical divergence-date estimates throughout the tree as the first analysis with 142 species. The resulting chronogram (Fig. 18.1) shows species for which we have information on phylogenetic relationships and either heterogamy and/or ploidy. Using the R package ‘‘ape’’ (Paradis et al. 2004), which is a software package for phylogenetic analysis, we tested the fit of alternative models for the evolution of new sex determination systems for a total of 90 species (55 frogs and 35 salamanders) for which we had data on heterogamy or sex chromosome polymorphism. This was done to select an appropriate model for ancestral reconstruction and for use in a permutation test described below. We coded all species as having either male heterogamy (0), female heterogamy (1), or, for each of three species (Xenopus laevis, Leiopelma hochstetteri, and Rana rugosa), a unique ‘‘de novo’’ sex-determining mechanism (2, 3, and 4), in which categories we include species with polymorphisms in mechanisms for sex determination. Xenopus laevis was assigned a unique heterogametic state in order to account for the finding that this species evolved its W-linked sex-determining gene after the split from S. tropicalis (Bewick et al. 2011), which also has female heterogamy. Leiopelma hochstetteri was assigned a unique heterogametic state in order to account for the finding that this species recently evolved a derived W0 sex-determining system that is unique to this lineage (Green et al. 1993; Sharbel et al. 1998). Rana rugosa was assigned a unique heterogamy state in order to accommodate evidence for recent and possibly repeated instances of sex chromosome turnover (Ogata et al. 2008). The three de novo states were coded as separate character states in order to ensure that known instances of novel sex determination mechanisms were included in the analysis even though they did not necessarily involve a change in heterogamy. We note that the newly evolved sex chromosomes of L. hochstetteri and R. rugosa are polymorphisms, and it is not clear whether these new polymorphisms will eventually fix in each species, and thus actually constitute a sex chromosome turnover. However, in both of these examples, at least one of the polymorphic systems for sex determination is species-specific and therefore new. Using the ‘‘ace’’ function of the ‘‘ape’’ package, we then evaluated the following models for evolution of the five heterogamy states: TE 415 399 EC 414 CO RR 413 Book ISBN: 978-3-642-31441-4 Page: 399/410 Polyploidization and Sex Chromosome Evolution in Amphibians (1) All rates equal (one rate for all possible transitions between states, one parameter). (2) One reversible rate between XY and ZW, and one reversible rate to and from any of the de novo states (two rates in total, two parameters). By reversible, we mean, for example, that the rate of change from XY to ZW is equal to the rate of change from ZW to XY. (3) One rate for XY to ZW, another rate for ZW to XY, and one reversible rate to and from any of the de novo states (three rates in total, three parameters). (4) Rates between each of the five heterogamy states are reversible and unique (ten rates in total, ten parameters). UN Editor Proof 18 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 18 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 400/410 B. J. Evans et al. (5) All rates unique (twenty rates in total, twenty parameters). 458 18.2.2 Results and Discussion 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 PR OO 463 D 462 TE 461 The likelihoods of each of these models were compared using the Akaike Information Criterion (Akaike 1974) calculated as 2k–2ln(L) where k is the number of parameters in the model and L is the maximum value of the likelihood function of the model. A P value was generated with a hierarchical likelihood ratio test (hLRT) with degrees of freedom equal to the difference in free parameters of the models under the assumption of a Chi-squared distribution. For the hLRT we evaluated whether adding complexity to the models resulted in a significant increase in model fit. The likelihoods of Models 1, 2, 3, 4 and 5 were -70.74572 (AIC = 143.5), -53.82688 (AIC = 111.7), -53.81533 (AIC = 113.6), -52.31287 (AIC = 124.6), and -49.97701 (AIC = 140.0) respectively. Model 2 thus was favored by the Akaike Information Criterion. According to the hLRT, Model 2 was also preferred over Model 1 (P\0.0001), but Model 3 was not preferred over Model 2 (P = 0.879). Model 4 was not preferred over Model 2 (P = 0.932) or Model 3 (P = 0.885), and Model 5 was not preferred over Model 2 (P = 0.982), or Model 3 (P = 0.999), or Model 4 (P = 0.912). These results suggests that the transition rate from ZW to XY is not significantly higher than the transition rate from XY to ZW. Model 2 was therefore used to reconstruct ancestral heterogamy states and also used for simulations in our permutation test described below. The model used in the ancestral reconstructions differs slightly from the model used in the permutation test in that the rate of reversal from the de novo heterogamy states to other heterogamy states was set to zero for the ancestral reconstructions. This was not possible with the permutation test, which requires a reversible model. We present results from this slightly different version for the ancestral reconstructions for illustrative purposes because with this model there is zero likelihood for all of the de novo states in all of the ancestral reconstructions. Other inferences, such as the likelihood of male and female heterogamy in the most recent common ancestor of frogs, of salamanders, and of frogs and salamanders discussed below, are identical with both of these models. The ancestral state reconstructions estimated from the analysis with 90 species but plotted on the chronogram with 143 species (Fig. 18.1) suggest that there is not strong statistical support to distinguish whether the ancestral hetrogamy state was female or male heterogamy (that is, ZW females and ZZ males) in frogs or salamanders. The marginal likelihood of female heterogamy (ZW) for the most recent common ancestor of frogs is 0.544, for the most recent common ancestor of salamanders is 0.499, and for the most recent common ancestor of frogs and salamanders is 0.513. Thus, the support for ZW versus XY heterogamy as the ancestral state in each group is effectively equivocal. EC 460 CO RR 459 F 457 UN Editor Proof 400 Layout: T1 Standard SC Chapter No.: 18 Book ISBN: 978-3-642-31441-4 Page: 401/410 Polyploidization and Sex Chromosome Evolution in Amphibians 401 497 18.3.1 Methods 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 PR OO 502 D 501 TE 500 In our analysis, we consider the sex-determining system to have changed every time that a change in heterogamy occurred from XY to ZW, from ZW to XY, or from ZW or XY to one of the three de novo sex-determining systems. To quantify the number of times that the system for sex determination changed in amphibians, we used the stochastic character mapping approach proposed by Nielsen (2002) as implemented by the R package ‘‘phytools’’ (Revell 2011). This approach simulates character evolution on a phylogeny, conditioning on the observed character states of the terminals. In this way, one can estimate the number of character-state transitions that occurred and also evaluate where in the phylogeny changes are likely to have occurred. The stochastic mapping simulations were performed using a reversible version of Model 2 described above. As discussed earlier, recent discoveries implicate sex chromosome turnover in facilitating polyploid speciation or the tolerance of polyploidization, at least in Xenopus and Leiopelma. These observations raise the question: is this a general phenomenon in amphibians? Since a change in heterogamy necessarily involves a change in the sex-determining system, we predicted that the time since a change in heterogamy (XY to ZW or ZW to XY) would be lower in species that are polyploid or that tolerate polyploidy than expected by chance, if the same number of polyploid species (or polyploid-tolerant species) were to evolve randomly on the phylogeny. We note that this hypothesis does not involve correlation between polyploidy and a particular heterogamy state, so standard approaches to test for phylogenetic correlation among traits cannot be used to test this prediction. Instead, we developed a novel permutation test that accommodates uncertainty in when and in which lineages sex chromosome turnover occurred during amphibian evolution. From the set of 90 species for which heterogamy information was available, we identified five phylogenetically independent instances of stable or spontaneous polyploidy. We emphasize that this is an underestimate of the number of independent polyplodization events, and we were able to use only five instances of polyploidization because we lack heterogamy and/or phylogenetic information for the other examples listed in Table 18.1. Diploidy has either been confirmed for the other species for which heterogamy information was available (Mable et al. 2011) or was assumed. The five examples of independent polyploidization or tolerance of polyploidization are: EC 499 CO RR 498 F 496 18.3 Is Polyploidy Tolerated to a Greater Degree in Species with Young Sex Chromosomes? 495 (1) Siren intermedia, a tetraploid, which may have descended from a tetraploid ancestor that also gave rise to S. lacertina (Morescalchi and Olmo 1974). We note that the polyploid status of the family Sirenidae has not been confirmed by additional studies (Mable et al. 2011). UN Editor Proof 18 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 18 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 402/410 542 543 544 546 545 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 F PR OO 540 541 For this permutation test, we used as a test statistic the mean time since the origin of the current heterogamy state for the five polyploid lineages. This mean was calculated from 1,000 simulations that are conditioned on the observed heterogamy states, using the stochastic character mapping approach described by Nielsen (2002) and implemented by the R package ‘‘phytools’’ (Revell 2011). Each stochastic mapping simulation provides one possible evolutionary scenario that is consistent with the data. It was necessary to perform many (1,000) stochastic mapping simulations for the observed data in order to accommodate uncertainty in these evolutionary scenarios (that is, to accommodate uncertainty in the ancestral reconstruction of the evolution of heterogamy). If polyploidy tends to occur soon after a change in heterogamy, then the observed test statistic should be smaller than the distribution of statistics calculated after repeatedly randomly selecting five species across the tree to be polyploid, and performing 1,000 stochastic character mapping simulations for each of the randomizations. The observed test statistic was therefore compared to a distribution of statistics generated from 100 randomizations where, in each randomization, five species are selected to be polyploid, with 1,000 stochastic character mapping simulations that were conditioned on the observed heterogamy states performed for each of the randomizations. The difference between the test statistic and the randomizations therefore is that the mean path length (that is, for each polyploid, the path length between the most recent change in heterogamy and the branch tip) was calculated respectively either from real polyploid species (for the test statistic), or from five species selected at random from the 90 species for which we have heterogamy data (for each of the randomizations). This test is conservative in the sense that it does not consider subsequent, phylogenetically independent instances of polyploidization that occurred in Xenopus (three additional independent instances of octoploidization and at least two additional independent instances of dodecaploidization). More specifically, because Xenopus has a relatively young sex-determining system, a test statistic generated by counting each of the independent polyploidization events in Xenopus is even lower than the test statistic that counts polyploidization of Xenopus only once (see below). One limitation of this analysis is that the frequency of transitions in amphibian sex-determining systems is undoubtedly underestimated. It is possible, for example, that changes in heterogamy occurred in other species that were excluded in the D 539 TE 537 538 (2) Ambystoma mexicanum, a species with spontaneous triploidy (Humphrey 1963) that is closely related to the unisexual triploids A. jeffersonianum, A. platineum, and A. tremblayi (see notes on species status of unisexuals in Table 18.1). (3) Xenopus laevis, a tetraploid species (Tymowska 1991). We excluded X. gilli from this analysis even though we have heterogamy data for this species because it shares a polyploid ancestor with X. laevis. (4) Rana esculenta, a naturally occurring diploid/triploid hybridogenic species formed from hybridization of R. lessonae and R. ridibunda (Uzzell et al. 1975). (5) Leiopelma hochstetteri, a diploid species with spontaneous triploidy (Green et al. 1984). EC 536 CO RR 535 B. J. Evans et al. UN Editor Proof 402 Layout: T1 Standard SC Chapter No.: 18 Book ISBN: 978-3-642-31441-4 Page: 403/410 Polyploidization and Sex Chromosome Evolution in Amphibians 403 604 18.3.2 Results and Discussion 589 590 591 592 593 594 595 596 597 598 599 600 601 602 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 PR OO 587 588 D 586 TE 584 585 EC 582 583 Stochastic mapping of heterogamy state, including the independent evolution of three de novo sex chromosomes provides an average estimated number of times that the sex chromosomes turned over of 32 (95% confidence interval: 25–41) based on 1,000 simulations that were conditioned on the observed heterogamy states. This is much higher than the maximum parsimony inference of only seven changes by Hillis and Green (1990). In the example simulation depicted in Fig. 18.2a, for instance, there are 28 changes in heterogamy. These simulations also suggest that more changes occurred from male heterogamy to female heterogamy than the reverse, even though our model comparison suggested that the rate of change in each direction was not significantly different. Out of 1,000 stochastic mapping simulations, the mean number of changes from male to female heterogamy (XY to ZW) was 22 (95 % confidence interval: 13–28) and the mean number of changes from female to male heterogamy (ZW to XY) was 7 (95 % confidence interval: 2–14). The example simulation depicted in Fig. 18.2a is typical of the other simulations in the sense that there are 19 changes CO RR 581 F 603 analysis because differences between the sex chromosomes were not cytologically detectable. It is also possible that some species experienced a change in the sexdetermining system that did not involve a change in heterogamy. Gastrotheca pseustes, for instance, is known to have polymorphism in the morphology of the Y chromosome (Schmid et al. 1990), but this species was coded as XY because these size variants may involve homologous chromosomes and no change in the sexdetermining system. Intraspecific polymorphism in sex-determining mechanisms has also been observed in Rana narina, Eleutherodactylus maussi, Rana japonica and R. narina (Eggert 2005), but phylogenetic information was lacking from these species in the phylogeny of Pyron and Wiens (2011). Furthermore, adding more taxa might influence the inferred timing of the transitions to heterogamy, even if the phylogeny remains the same (e.g., added taxa could subdivide long branches and help clarify where on a given branch the heterogamy transition occurred). Another limitation of this analysis is that we use the total time that the polyploid species have been in the observed heterogamy states for our test statistic, rather than the difference between these times and the age of each polyploid species. The latter difference would be a better metric for this test because it focuses on events prior to polyploidization. However, it is difficult to estimate the age of each polyploid (or of a randomly selected species in the permutation) because in some cases the diploid ancestor of the polyploid is unknown or extinct (see above), or other cases because we lack phylogenetic information from the sister taxon. We note that results are contingent on the phylogeny and evolutionary model, and that this analysis does not accommodate uncertainty in divergence times and phylogenetic relationships. 580 UN Editor Proof 18 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 18 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 404/410 B. J. Evans et al. (a) (b) Salamanders 1 Frogs PR OO 5 2 F Editor Proof 404 XY ZW de novo 1 de novo 2 de novo 3 4 3 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 from male to female heterogamy but only 6 changes from female to male heterogamy. The observed average path length to a change in the sex-determining system for the five polyploids averaged over 1,000 stochastic mapping simulations, was 89.0 million years. This is not to suggest that no sex chromosome degeneration occurred within this period. In fact, if sex chromosome degeneration in amphibians occurred at a similar rate as it did in therian mammals (4.6 genes per million years; Graves 2004), the ancestors of these polyploids probably did have somewhat degenerate sex chromosomes. Rather, this result suggests that the amount of sex chromosome degeneration that typically occurs within this period of time was not of sufficient magnitude to prevent polyploidization or the tolerance of spontaneous polyploidization. The permutation test indicates that this test statistic is not significantly lower than the distribution of statistics calculated when polyploids evolved five times on random branches in this phylogeny (P = 0.059; average path length in permutations was 141.2 million years and the standard deviation was 32.7 million years), although the P value is close to 0.05. If we include six independent polyploidizations in Xenopus (one tetraploidization, three CO RR 621 UN 620 EC TE D Fig. 18.2 Simulations that stochastically map evolution of heterogamy form the basis of the permutation test. Shown here is (a) an example of a stochastic mapping simulation for the evolution of the observed heterogamy states and (b) highlighted simulated paths to each of the five observed polyploid species. Species with missing heterogamy data from Fig. 18.1 have been removed for these analyses and species names are omitted for clarity. In (b) simulated ages of the observed heterogamy states for Ambystoma mexicanum, Siren intermedia, Leiopelma hochstetteri, Xenopus laevis, and Rana esculentia are labeled 1, 2, 3, 4, and 5 respectively. The test statistic is the average age of the observed heterogamy state for these five polyploid species, averaged over 1,000 stochastic mapping simulations. This test statistic is compared to analogous calculations from 100 permutations where five species were randomly selected to be polyploid and 1,000 stochastic mapping simulations were performed for each randomization Layout: T1 Standard SC Chapter No.: 18 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 F 642 PR OO 641 D 640 TE 639 405 octoploidizations, and at least two dodecaploidizations; reviewed in Evans 2008) in the observed test statistic, in addition to the four other examples of polyploidization itemized above, the average observed path length to a change in the sexdetermining system for the ten independent polyploid lineages is 68.1 million years. This test statistic is significantly smaller than statistics calculated from 100 permutations where one of the five randomly selected polyploids is also assumed to undergo six independent polyploidizations (P = 0.020; average path length in permutations was 169.8 million years and the standard deviation was 55.3 million years). Although there are at least six independent polyploidizations in Xenopus, this test suffers from pseudoreplication in that these polyploid lineages may share the same system for sex determination (i.e., DMW). Additional data on whether other polyploid species of amphibians have male or female heterogamy would clearly help illuminate the question of whether species with young sex chromosomes are more tolerant of polyploidization. It is surprising how little is known about heterogamy of polyploid amphibians given that karyotypes of essentially all of these species were inspected in order to identify polyploidy in the first place. One possible reason for this dearth of information on heterogamy of polyploid species is that many of these species may lack morphologically distinct sex chromosomes. This proposal, if accurate, would be consistent with the contention that polyploidization is better tolerated by species with minimally degenerate sex chromosomes. Additional insights into the influence of sex chromosome evolution and polyploidization may be gained from studies of laboratory-generated polyploid Xenopus. Laboratory allopolyploidization in Xenopus duplicates autosomal chromosomes but generates female polyploid individuals with 3 Z chromosomes and 1 W chromosome and male polyploid individuals with 4 Z chromosomes (Kobel and Du Pasquier 1986). Thus, the W chromosome of one of the ancestral diploids is not inherited by Xenopus allopolyploids, and therefore never gets a chance to segregate as an autosome. In this way, the mechanism of Xenopus allopolyploidization in nature may circumvent autosomal segregation of a W chromosome (whether degenerate or not) (reviewed in Evans 2008), and this could account for the unusually high incidence of polyploidization in this genus. Analyses presented here show that changes in the system for sex chromosome turnover were much more common in amphibians than previously proposed [*32 versus 7 as proposed by Hillis and Green (1990)], and that there is not strong support for female versus male heterogamy in the ancestor of salamanders, frogs, or the most recent common ancestor of salamanders and frogs. These changes need not involve the evolution of completely novel systems for sex determination, and some of these inferred changes may be reversals to an ancestral system. We also found that changes from male to female heterogamy occurred more frequently than changes from female to male heterogamy, although the transition rates between each state were not significantly different. This result also contradicts the conclusions of Hillis and Green (1990), who stated that there was a bias in evolution from female heterogamy to male heterogamy. One reason for these differences is that our analysis included new data on heterogamy (reviewed in EC 638 CO RR 637 Book ISBN: 978-3-642-31441-4 Page: 405/410 Polyploidization and Sex Chromosome Evolution in Amphibians UN Editor Proof 18 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 18 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 406/410 B. J. Evans et al. 691 18.4 Conclusions 685 686 687 688 689 PR OO 684 F 690 Schmid et al. 2010), more species, and a more comprehensive phylogeny (Pyron and Wiens 2011). Another reason for our higher estimate in the number of changes is that we considered three de novo changes in sex determination as a new heterogamy state, thereby forcing a sex chromosome turnover in these lineages (but this is only 3 out of *32 changes). Differences in the resolution, relationships, and branch lengths of the phylogenies used in each study are also likely to have played a role in these differing conclusions. Finally, and importantly, the different analytical approaches may have influenced the results (that is, the use of maximum parsimony by Hillis and Green (1990) and maximum likelihood here). 682 683 703 704 705 706 Acknowledgements We are particularly grateful to Barbara Mable who provided a comprehensive critical assessment of an earlier version of this chapter. We also thank Liam Revell for advice and assistance with the R package ‘‘phytools’’ and Jim Bogart and Ben Bolker for comments. 707 References 708 709 710 711 712 713 714 715 716 717 718 719 Akaike H (1974) A new look at the statistical model identification. IEEE Trans Autom Contr 19:716–723 Arnold AP, Itoh Y, Melamed E (2008) A bird’s-eye view of sex chromosome dosage compensation. Annu Rev Genomics Hum Genet 9:109–127 Baéz AM (2000) Tertiary anurans from South America. In: Heatwole H, Carroll RL (eds) Amphibian biology. Surrey Beatty, Chipping Norton, Australia, pp 1388–1401 Beçak ML, Beçak W (1998) Evolution by polyploidy in amphibia: new insights. Cytogenet Cell Genet 80:28–33 Bergero R, Charlesworth D (2009) The evolution of restricted recombination in sex chromosomes. Trends Ecol Evol 24:94–102 Bergero R, Charlesworth D (2011) Preservation of the Y transcriptome in a 10 million-year-old plant sex chromosome system. Curr Biol 21:1470–1474 696 697 698 699 700 701 TE 695 EC 694 CO RR 693 D 702 Polyploidization generates new species and duplicates genes; the resulting genetic redundancy has the capacity to degrade or to undergo innovation. The question of why some lineages frequently undergo polyploidization whereas others do not thus has important implications for evolution and adaptation. Eventually we will have a much more comprehensive understanding of genetic variation in (a) the triggers of sex determination in amphibians, (b) the extent of suppressed recombination that surrounds these genetic triggers, (c) the extent of sex chromosome degeneration that exists in amphibians, and (d) whether or not other lineages of polyploid amphibians have minimally degenerate sex chromosomes. Future discoveries in these areas can undoubtedly be leveraged to provide exciting new insights into the role of sex chromosome degeneration in the propensity of species to tolerate polyploidization. 692 UN Editor Proof 406 Layout: T1 Standard SC Chapter No.: 18 407 EC TE D PR OO F Bewick AJ, Anderson DW, Evans BJ (2011) Evolution of the closely related, sex-related genes DM-W and DMRT1 in African clawed frogs (Xenopus). Evolution 65:698–712 Blackburn DC and Beier M (2011) ‘‘Xenopus paratropicalis’’ is not a valid name. Zootaxa 3035:57–58 Briggs R (1947) The experimental production and development of triploid frog embryos. J Exp Zool 106:237–266 Bogart JP (1980) Evolutionary significance of polyploidy in amphibians and reptiles. In: Lewis WH (ed) Polyploidy, biological relavance. Basic life sciences, New York, pp 341–378 Carroll RL (1988) Vertebrate paleontology and evolution W. H. Freeman and Company, New York Charlesworth B, Charlesworth D (2000) The degeneration of Y chromosomes. Philos Trans R Soc London B 355:1563–1572 Charlesworth D (2002) Plant sex determination and sex chromosomes. Heredity 88:94–101 Charlesworth D, Charlesworth B, Mariais G (2005) Steps in the evolution of heteromorphic sex chromosomes. Heredity 95:118–128 Duellman WE, Trueb L (1994) Biology of amphibians. The Johns Hopkins University Press, Baltimore Eggert C (2005) Sex determination: the amphibian models. Reprod Nutr Dev 44:539–549 Evans BJ (2007) Ancestry influences the fate of duplicated genes millions of years after duplication in allopolyploid clawed frogs (Xenopus). Genetics 176:1119–1130 Evans BJ (2008) Genome evolution and speciation genetics of allopolyploid clawed frogs (Xenopus and Silurana). Front Biosci 13:4687–4706 Evans BJ, Cannatella DC, Melnick DJ (2004a) Understanding the origins of areas of endemism in phylogeographic analyses: a reply to Bridle et al. Evolution 58:1397–1400 Evans BJ, Kelley DB, Tinsley RC, Melnick DJ, Cannatella DC (2004b) A mitochondrial DNA phylogeny of clawed frogs: phylogeography on sub-Saharan Africa and implications for polyploid evolution. Mol Phylogenet Evol 33:197–213 Evans BJ, Carter TF, Hanner R et al (2008a) A new species of clawed frog (genus Xenopus), from the Itombwe Plateau, democratic republic of the congo: implications for DNA barcodes and biodiversity conservation. Zootaxa 1780:55–68 Evans BJ, Greenbaum E, Kusamba C et al (2011) Description of a new octoploid frog species (Anura: Pipidae: Xenopus) from the democratic republic of the congo, with a discussion of the biogeography of African clawed frogs in the Albertine Rift. J Zool London 283:276–290 Evans BJ, Kelley DB, Melnick DJ, Cannatella DC (2005a) Evolution of RAG-1 in polyploid clawed frogs. Mol Biol Evol 22:1193–1207 Evans SE, Jones MEH, Krause DW (2008b) A giant frog with South American affinities from the Late Cretaceous of Madagascar. Proc Nat Acad Sci 105:2951–2956 Evans SE, Lally C, Chure DC, Elder A, Maisano JA (2005b) A new fully metamorphosed salamander from the Late Jurassic of North America. Zool J Linn Soc 143 Evans SE, Milner AR (1996) A metamorphosed salamander from the early Cretaceous of Las Hoyas, Spain. Philos Trans R Soc London B 351:627–646 Ezaz T, Stiglec R, Veyrunes F, Graves JAM (2006) Relationships between vertebrate ZW and XY sex chromosome systems. Curr Biol 16:R736–R743 Felsenstein J (1974) The evolutionary advantage of recombination. Genetics 78:737–756 Fankhauser G, Crotta R, Perrot M (1942) Spontaneous and cold-induced triploidy in the Japanese newt Triturus pyrrhogaster. J Exp Zool 89:167–181 Fankhauser G (1941) The frequency of polyploidy and other spontaneous aberrations of chromosome number among larvae of the newt Triturus viridescens. Proc Nat Acad Sci 27(11):507–512 Fankhauser G, Watson RC (1942) Heat-indiced triploidy in the newt, Triturus viridescens. Proc Nat Acad Sci 28:436–440 Gardner JD (2003) The fossil salamander Proamphiuma cretacea Estes (Caudata: Amphiumidae) and relationships within the Amphiumidae. J Vertebr Paleontol 23:769–782 Graves JAM (2004) The degenerate Y chromosome—can conversion save it? Reprod Fertil Dev 16:527–534 Graves JAM (2008) Weird animal genomes and the evolution of vertebrate sex and sex chromosomes. Annu Rev Genet 42:565–586 CO RR 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768 769 770 771 772 773 Book ISBN: 978-3-642-31441-4 Page: 407/410 Polyploidization and Sex Chromosome Evolution in Amphibians UN Editor Proof 18 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 18 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 408/410 EC TE D PR OO F Green DM (1988) Cytogenetics of the endemic New Zealand frog, Leiopelma hochstetteri: extraordinary supernumerary chromosome variation and a unqie sex-chromosome system. Chromosoma 97:55–70 Green DM, Kezer J, Nussbaum RA (1984) Triploidy in Hochstetter’s frog, Leiopelma hochstetteri, from New Zealand. New Zealand J Zool 11:457–460 Green DM, Zeyl CW, Sharbel TF (1993) The evolution of hypervariable sex and supernumerary (B) chromosomes in the relict New Zealand frog, Leiopelma hochstetteri. J Evol Biol 6:417–441 Gregory TR, Mable BK (2005) Polyploidy in animals. In: Gregory TR (ed) The Evolution of the Genome. Elsevier Academic Press, Burlington, pp 428–517 Hayes TB (1998) Sex determination and primary sex differentiation in amphibians: genetic and developmental mechanisms. J Exp Zool 281:373–399 Hillis DM, Green DM (1990) Evolutionary changes of heterogametic sex in the phylogenetic history of amphibians. J Evol Biol 3:49–64 Holloway AK, Cannatella DC, Gerhardt HC, Hillis DM (2006) Polyploids with different origins and ancestors form a single sexual polyploid species. Am Nat 167:E88–E101 Holman JA (2003) Fossil frogs and toads of North America Indiana University Press. Bloomington and Indianapolis, IN Kashiwagi K (1993) Production of triploids and their reproductive capacity in Rana rugosa. Sci Rep Lab Amphibian Biol Hiroshima Univ 12:23–36 Kawamura T, Tokunaga C (1952) The sex of triploid frogs, Rana japonica Günther. J Sci Hiroshima Univ, Ser B, Div 1 (Zoology) 13 Humphrey RR (1963) Polyploidy in the Mexican axolotl (Ambystoma mexicanum) resulting from multinucleate ova. Proc Nat Acad Sci 50:1122–1127 Kobel HR, Loumont C, Tinsley RC (1996) The extant species. In: Tinsley RC, Kobel HR (eds) The Biology of Xenopus. Clarendon Press, Oxford, pp 9–33 Kawamura T (1984) Polyploidy in amphibians. Zool Sci 1:1–15 Kobel HR (1996) Allopolyploid speciation. In: Tinsley RC, Kobel HR (eds) The Biology of Xenopus. Clarendon Press, Oxford, pp 391–401 Kobel HR, Du Pasquier L (1986) Genetics of polyploid Xenopus. Trends Genet 2:310–315 Litvinchuk SN, Rosanov JM, Borkin LJ (1998) A case of natural triploidy in a smooth newt Triturus vulgaris (Linneaus, 1958), from Russia (Caudata: Salamandridae). Herpetozoa 11:93–95 Mable BK (2004) ‘Why polyploidy is rarer in animals than in plants’: myths and mechanisms. Biol J Linn Soc 82:453–466 Mable BK, Alexandrou MA, Taylor MI (2011) Genome duplication in amphibians and fish: an extended synthesis. J Zool 284:151–182 Mable BK, Roberts JD (1997) Mitochondrial DNA evolution in the genus Neobatrachus (Anura: Myobatrachidae). Copeia 1997:680–689 Mayrose I, Zhan SH, Rothfels CJ et al (2011) Recently formed polyploid plants diversify at lower rates. Science 333:1257 Milner AR (2000) Mesozoic and Tertiary Caudata and Albanerpetontidae. In: Heatwole H, Carrol RL (eds) Amphibian Biology. Surrey Beatty, Chipping Norton, Australia, pp 31–108 Morescalchi A, Olmo E (1974) Sirenids: a family of polyploid urodeles? Experientia 30:491–492 Moler PE, Kezer J (1993) Karyology and systematics of the salamander genus Pseudobranchus (Sirenidae). Copeia 1993:39–47 Muller HJ (1925) Why polyploidy is rarer in animals in plants. Am Nat 59:346–353 Muller HJ (1964) The relation of recombination to mutational advance. Mutat Res 106:2–9 Naylor BG, Fox RC (1993) A new ambystomatid salamander Dicamptodon antiquus n. sp. from the Paleocene of Alberta. Can J Earth Sci 30:814–818 Nielsen R (2002) Mapping mutations on phylogenies. Syst Biol 51:729–739 Ogata M, Hasegawa Y, Ohtani H, Mineyama M, Miura I (2008) The ZZ/ZW sex-determining mechanism originated twice and independently during evolution of the frog, Rana rugosa. Heredity 100:92–99 Ohno S (1967) Sex chromosomes and sex-linked genes. Springer, Berlin Orr HA (1990) ‘Why polyploidy is rarer in animals than in plants’ revisited. Am Nat 136:759–770 CO RR 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792 793 794 795 796 797 798 799 800 801 802 803 804 805 806 807 808 809 810 811 812 813 814 815 816 817 818 819 820 821 822 823 824 825 826 827 B. J. Evans et al. UN Editor Proof 408 Layout: T1 Standard SC Chapter No.: 18 409 EC TE D PR OO F Otto SP, Whitton J (2000) Polyploid incidence and evolution. Annu Rev Genet 34:401–437 Papp B, Pál C, Hurst LD (2003) Dosage sensitivity and the evolution of gene families in yeast. Nature 424:194–197 Paradis E, Claude J, Strimmer K (2004) APE: analysis of phylogenetics and evolution in R language. Bioinf 20:289–290 Poinar GO, Cannatella DC (1987) An upper Eocene frog from the Dominican Republic and its implication for Caribbean biogeography. Science 237:1215–1216 Pyron RA, Wiens JJ (2011) A large-scale phylogeny of amphibia including over 2800 species, and a revised classification of extant frogs, salamanders, and caecilians. Mol Phylogenet Evol 61:543–583 Qian W, Zhang J (2008) Gene dosage and gene duplicability. Genetics 179:2319–2324 Rage JC, Rocek Z (1989) Redescription of Triadobatrachus massinoti (Piveteau, 1936) an anuran amphibian from the early Triassic. Palaeontographica Paleontologica 206:1–16 Revell JJ (2011) Phytools: phylogenetic tools for comparative biology (and other things). (R Package) Rocek Z (2000) Mesozoic anurans. In: Heatwole H, Carrol RL (eds) Amphibian biology. Surrey Beatty, Chipping Norton, Australia, pp 1295–1331 Rocek Z, Rage J-C (2000) Tertiary Anura of Europe, Africa, Asia, North America, and Australia. In: Heatwole H, Carrol RL (eds) Amphibian Biology. Surrey Beatty, Chipping Norton, Australia, pp 1332–1387 Sanchiz FB (1998) Salienta. In: Wellnhofer P (ed) Encyclopedia of paleoherpetology, Part 4, Salienta. Verlag, Pfeil, Munich, pp 1–276 Sanderson MJ (2002) Estimating absolute rates of molecular evolution and divergence times: a penalized likelihood approach. Mol Biol Evol 19:1218–1231 Sanderson MJ (2003) r8s: inferring absolute rates of evolution and divergence times in the absence of a molecular clock. Bioinformatics 19:301–302 Schmid M, Sims SH, Haaf T, Macgregor HC (1986) Chromosome banding in amphibia X. 18S and 28S ribosomal RNA genes, nucleolus organizers and nucleoli in Gastrotheca riobambae. Chromosoma 94:139–145 Schmid M, Steinlein C (2001) Sex chromosomes, sex-linked genes, and sex determination in the vertebrate class Amphibia. In: Scherer G, Schmid M (eds) Genes and mechanisms in vertebrate sex determination. Verlag, Basel, pp 143–176 Schmid M, Steinlein C, Bogart JP et al (2010) The chromosomes of terraranan frogs. Cytogenetic Genome Res 130–131:1–568 Schmid M, Steinlein C, Friedl R et al (1990) Chromosome banding in Amphibia. XV. Two types of Y chromosomes and heterochromatin hypervariability in Gastrotheca pseustes (Anura, Hylidae). Chromosoma 99:413–423 Sharbel TF, Green DM, Houben A (1998) B-chromosome origin in the endemic New Zealand frog Leiopelma hochstetteri through sex chromosome devolution. Genome 41:14–22 Skaletsky H, Kuroda-Kawaguchi T, Minx PJ et al (2003) The male-specific region of the human Y chromosome is a mosaic of discrete sequence classes. Nature 432:823–837 Stöck M, Horn A, Grossen C et al (2011) Ever-young sex chromosomes in European tree frogs. PLoS Biol 9:e1001062 Stöck M, Ustinova J, Lamatsch DK et al (2009) A vertebrate reproductive system involving three ploidy levels: Hybrid origin of triploids in a contact zone of diploid and tetraploid Paleartic green toads (Bufo viridis subgroup). Evolution 64:944–959 Straub T, Becker PB (2007) Dosage compensation: the beginning and end of a generalization. Nat Rev Genet 8:47–57 Svartman M, Stone G, Stanyon R (2005) Molecular cytogenetics discards polyploidy in mammals. Genomics 85:425–430 Tihen JA, Wake DB (1981) Vertebrae of plethodontid salamanders from the Lower Miocene of Montana. J Herpetology 15:35–40 Tymowska J (1991) Polyploidy and cytogenetic variation in frogs of the genus Xenopus. In: Green DS, Sessions SK (eds) Amphibian cytogenetics and evolution. Academic Press, San Diego, pp 259–297 CO RR 828 829 830 831 832 833 834 835 836 837 838 839 840 841 842 843 844 845 846 847 848 849 850 851 852 853 854 855 856 857 858 859 860 861 862 863 864 865 866 867 868 869 870 871 872 873 874 875 876 877 878 879 880 881 Book ISBN: 978-3-642-31441-4 Page: 409/410 Polyploidization and Sex Chromosome Evolution in Amphibians UN Editor Proof 18 Book ID: 272454_1_En Date: 16-8-2012 Layout: T1 Standard SC Chapter No.: 18 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 410/410 EC TE D PR OO F Uzzell T, Berger L, Günther R (1975) Diploid and triploid progeny from a diploid female of Rana esculenta (Amphibia Salientia). Proc Acad Nat Sci Philadelphia 127:81–91 Wiens JJ (2011) Re-evolution of lost mandibular teeth in frogs after more than 200 million years, and re-evaluating Dollo’s Law. Evolution 65:1283–1296 Wiens JJ, Sukumaran J, Pyron RA, Brown RM (2009) Evolutionary and biogeographic origins of high tropical diversity in Old World frogs (Ranidae). Evolution 63:1217–1231 Wolfe KH (2001) Yesterdays’s polyploids and the mystery of diploidization. Nat Rev Genet 2:333–341 Yoshimoto S, Ikeda K, Izutsu Y et al (2010) Opposite roles of DMRT1 and its W-linked paralog, DM-W, in sexual dimorphism of Xenopus laevis: implications of a ZZ/ZW-type sexdetermining system. Development 137:2519–2526 Yoshimoto S, Okada E, Umemoto H et al (2008) A W-linked DM-domain gene, DM-W, participates in primary ovary development in Xenopus laevis. Proc Nat Acad Sci 105:2469–2474 CO RR 882 883 884 885 886 887 888 889 890 891 892 893 894 B. J. Evans et al. UN Editor Proof 410 Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 411/415 Index PR OO F 1 2 19 20 21 22 23 24 25 26 27 28 29 B Background selection, 388 Bichir, 352 Bivalent formation/bivalent, 218, 219 Bivalent(s), 37–41, 43–45 Bottom-up approach (to genome alignment), 97 Bowfin, 351, 360, 367 Brachypodium, 96 Brassica napus, 283, 285 30 31 32 33 34 35 Cephalochordate, 309, 310, 313, 316, 325 Chiasma, 218, 219 Chromosome, 272, 280, 283, 284, 287 non-reduction, 99 number, 78–80 Cod, 343 Comparative expression profiling, 202 Compensating aneuploidy, 283 Concerted evolution, 103, 193 Conserved synteny, 346, 351–355, 360 Cotton, 183, 273, 282, 285 Crossover (co), 34, 39–41 Cytomixis, 42 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 D Detecting whole genome duplication, 77–82, 84, 85, 87, 88 Diakinesis, 34, 42 Diploidization, 386 Diplotene, 34 Disomic, 43 Disomic inheritance, 218–220 Divergent resolution/reciprocal gene loss after polyploidization, 9, 10, 12 DNA replication, 34, 35 Dosage balance, 293, 302 compensation, 386, 387 sensitivity, 281 Dosage/dosage balance and duplicate genes, 2, 5, 14 Double reduction, 43, 44, 46 Double strand breaks (DSB), 34, 39, 40 Duplicate gene expression, 192, 195 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 D A Allopolyploidy, 7, 8, 9, 22, 183, 184, 245–247, 250, 252, 254, 255, 257, 259–263, 272, 282 Amphibians, 385, 387–389, 394, 395, 401, 404–406 Amplification fragment length polymorphism (AFLP), 245, 252, 257, 262 Anaphase, 34, 41, 43, 46 Aneuploidy, 26, 218, 272, 283, 285 Apomixis, 47, 48 Arabidopsis suecica, 285 Arabidopsis thaliana, 285 Autonomous pairing sites (APS), 37 Autopolyploids, 7, 8, 23, 98 Axial element (AE), 35 CO RR EC TE 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 UN Editor Proof Layout: T1 Standard SC Chapter No.: BM C Calandinos, 345 Carp, 341, 345, 347, 348 Catfish, 345, 367 Centromere, 34, 35, 39, 41, 43 Centromeric divergence, 100 P. S. Soltis and D. E. Soltis (eds.), Polyploidy and Genome Evolution, DOI: 10.1007/978-3-642-31442-1,  Springer-Verlag Berlin Heidelberg 2012 411 Layout: T1 Standard SC Chapter No.: BM Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 412/415 Index E Ecological and adaptive consequences of hybridization and polyploidy in spartina, 231 Epigenetic, 216–218, 245, 246, 249, 257, 262–264 Expression, 272, 280–282, 285–287 Expression stoichiometry, 386 79 80 81 82 83 84 85 86 87 88 89 90 F Fertility, 210, 217–219 Fiber, 202 Fish, 341–344, 349, 350, 352 Fluorescence in situ hybridization (fish), 213, 283, 286, 287 Fractionation Biased, 141 Mechanism, 143 Fugu, 342, 349, 350, 363, 365 Functional coherence, 104 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 G Gamete(s), gametogenesis, 33, 34, 42, 43, 46–48 Gametophyte, 42, 47 Gar, 342, 351, 352, 360, 366, 367 Gene balance hypothesis, 281 conversion, 194 copy number, 80, 81 duplication and loss, 85–87 Chordate, 309, 310, 313, 315, 322, 324, 325, 327, 328 Vertebrate, 309, 310, 313, 315, 316, 319, 321, 327, 328 dosage, 283 expression, 27, 245, 252, 256, 257, 261, 262, 264 function evolution, 310, 321, 322 loss, 101, 191 Gene tree parsimony, 82, 83–85 reconciliation, 81–87 Genome dominance, 141 doubling, 127 duplication, 293, 300, 341, 343, 344, 349, 351, 353 duplication and evolution, 309, 310, 313, 316, 323 evolution, 112, 117, 127 evolution following hybridization and allopolyploid speciation in spartina, 233 groups, 187 rearrangement, 139, 140 sequence(s), 94 size, 189 Genomic balance, 27 Genomic dominance, 141, 198 Genomic downsizing, 189 Genomic in situ hybridization (GISH), 211, 283, 284, 286, 287 Genomic plasticity, 109, 127 Ghost duplication, 98 Goldfish, 345, 347 Gossypium, 188, 271, 284 barbadense, 188 hirsutum, 187 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 H Heterodimerization of duplicate genes, 4 Heterogametic;Heterogamy, 387–389, 400, 403, 405 Heterosis, 25 Hexaploids, 28 Hierarchical clustering approach (to genome alignment), 97 Hill–Robertson effect, 388 Hitchhiking, 388 Homeolog, 195, 278–282, 284–286 Homoelogous (nonhomologous chromosomes), 38, 39 Homoeolog expression, 197–200 Homoeologous gene expression, 197 Homoeologous pairing, 188 Homoeologue, 257, 261 Homogametic;Homogamy, 387 Homogenization/intergenomic homogenization, 211, 217, 218 Homologous (chromosomes) homology, 34–40, 45 Hox, 313, 314, 316, 318, 320, 346, 347, 349, 350, 352, 354, 355, 363, 364, 366, 370 Hybridization, 38, 40–43, 46, 47, 245–247, 250, 252–257, 260, 261, 264 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 159 160 161 162 163 164 I Illegitimate recombination, 190 Introduction, 226, 227, 230, 232 Inversion(s)/chromosomal inversion(s), 217, 219 ITS 165 166 167 168 169 170 AQ1 CO RR EC TE D PR OO F 71 72 73 74 75 76 77 78 UN Editor Proof 412 Layout: T1 Standard SC Chapter No.: BM Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 413/415 413 K Karyotype, 272, 283, 284 evolution, 100 restructuring, 210, 217, 218, 220 175 176 L Leptotene, 34, 35 177 178 179 180 181 182 183 184 185 186 187 188 189 190 M Medaka, 342, 350–352, 359, 360, 363, 366, 370, 371 Meiosis, 33 Metaphase, 33, 34 Methylation-sensitive amplification polymorphism (msap), 257, 260–262 Microarray, 197, 252–257, 260, 261, 264 Monosomy, 272, 283–286 Muller’s ratchet, 388 Multivalent(s), 36–39, 41, 43, 44, 46, 47 formation, 7, 8 formation after polyploidization, 211, 218, 219 191 192 193 194 195 196 197 198 199 200 201 202 203 204 205 206 207 208 209 210 N Neofunctionalization, 363, 365–367, 370, 371 of duplicate genes, 3, 11, 14 of pathways, 6 Next-generation sequencing (NGS), 211, 213–215 Nobilization, 99 Non-crossover, 40 Nonfunctionalization, 346, 354, 355, 358, 361–363, 365, 369 Nonfunctionalization of duplicate genes, 2 Nonreciprocal exchange, 194 homoeologous recombination, 191 Nuclear-cytoplasmic interaction hypothesis, 215 Null allele, 394 Nullisomy, 272, 283, 284, 286 211 212 213 214 215 216 P Pachytene, 33, 34, 36, 42 Pairing (of chromosomes), 33 Pairing-partner switch(es) (PPS), 37–39, 41 Paleopolyploidy, 95 Ph1 locus, 38, 39 Placode, 309, 313, 318, 327–329 Platyfish, 342 Poaceae (gramineae), 94 Polyploidization, 341, 344, 345, 348, 350, 362, 371 Polyploidy, 129 Polysomic, 45 Presence absence variation, 141, 142 Prophase, 34, 41 Psuedogenization, 191, 193 Pufferfish, 345, 350–352, 366 Punnett square, 43 Putative ancestral region (PAR), 97 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 Q Quadrivalent(s), 43–45, 218 236 237 D TE EC CO RR O Octoploids, 28 Ohnolog, 309, 311, 312, 314, 321, 323–325, 329, 354, 356, 358, 362–367, 370 Ohnologon, 354–356, 358 PR OO F 171 172 173 174 UN Editor Proof Index R Rabl-configuration, 35, 36 Random-end pairing model, 37 Rates of gene duplication, 1, 2 of speciation after polyploidization, 13 Ray flower/ray floret, 262, 263, 264 rDNA, 277, 278 Recent polyploidy, 273, 274, 285, 286 Recombination, 33, 34, 38–41, 43, 44, 46, 47 Recombination nodules (RNS), 39 Recurrent reticulate evolution and polyploidy in spartina, 226 Repeated polyploidy, 273, 274, 276 Retention rates of duplicate genes, 2, 5, 15 Retroelements/retrotransposons, 210, 211, 212, 214, 215 Rho, 96 Ribosomal DNA (rDNA), 210, 211, 213, 215, 216 Rice, 96 Rice chromosomes 11 and 12, 102 238 239 240 241 242 243 244 245 246 247 248 249 250 251 252 253 254 255 256 257 258 259 S Saccharomyces cerevisiae, 294 Salmon, 342, 345, 346 Satellite repeat(s), 211 Secondary association(s), 96 260 261 262 263 264 Layout: T1 Standard SC Chapter No.: BM Book ID: 272454_1_En Date: 16-8-2012 Book ISBN: 978-3-642-31441-4 Page: 414/415 Index F T Tandem repeat(s), 211, 212 Teleost, 341, 349–352, 359, 360, 366, 369 Telomere, 35, 36, 39 Telomere bouquet (formation), 35, 36 Telophase, 34 Tetraodon, 351, 359, 369 Tetrasomic (inheritance), 43 Tetrasomy, 272, 283–286 Tilapia, 342 PR OO 305 306 307 308 309 310 311 312 313 314 Tissue-specific silencing, 281, 282 Tobacco, 210, 214, 215 Tragopogon, 271, 273–288 castellanus, 274, 275 mirus, 271–275, 277–279, 282, 283 miscellus, 271–274, 276–285 x mirabilis, 287 Transcriptome shock, 252, 253, 255, 262 Transcriptomic shock, 272, 282, 286, 287 Transgressive, 255 Translocation(s)/intergenomic translocation(s), 211, 217, 218, 219 Transposable elements, 123, 189 Transposon, 322 Transposon/transposable element, 252, 253, 260 Trichomes, 182, 198, 202 Tripartite brain, 329 Triploids, 28 Trisomy, 272, 283–286 Triticum, 110, 111, 112, 115, 119, 121, 125, 284, 285 Trout, 342, 346 CO RR EC TE Segmental allopolyploids, 25 Segregation, 33, 34, 40, 41, 43–45, 47, 48 Self-incompatibility, 248, 263 Senecio cambrensis, 245–266, 273, 285, 286 eboracensis, 246, 247, 249, 250, 264 squalidus, 245–251, 253, 254, 257, 258, 262, 263, 266 vulgaris, 248–251, 253–255, 257, 258, 261, 264–266 x baxteri, 245, 250, 252–257, 260–262 Sequenom, 280, 282, 284 Sex chromosome, 385–389, 394, 399, 401, 404, 405 degeneration, 386, 387, 404 turnover, 389, 405 Sigma, 97 Silencing, 192, 196–199 Sister chromatid, 34, 35, 39, 41, 43, 46 Spartina anglica, 273, 285, 286 Speciation, 245–247 Speciation after polyploidization, 9, 11–13 Squalius alburnoides, 345, 348 Stickleback, 342, 345, 351, 355, 356, 358–360, 370, 371 Subfunctionalization, 282, 285, 363–366, 369, 370 Subfunctionalization of duplicate genes, 3, 4, 11, 13, 15 Synapsis, 34, 35, 37–39 Synaptonemal complex (SC), 34, 39 Synonymous substitution, 182, 192 Synteny blocks, 97 conservation, 314 Synthetic nicotiana tabacum, 211, 214–216, 219, 220 Synthetic polyploids, 271, 273–275, 278, 282, 285, 287 D S (cont.) 265 266 267 268 269 270 271 272 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 293 294 295 296 297 298 299 300 301 302 303 304 UN Editor Proof 414 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 336 337 U Unequal crossing over, 194 Unreduced gametes, 33, 42, 46–48 Urochordate evolution, 329 338 339 340 341 V Vertebrate, 341, 342, 344, 348–350, 362, 364, 367, 370, 372 Vertebrate Origin and Innovations Neural Crest, 309, 313, 327, 328 342 343 344 345 346 W Wheat, 109–123, 125–129, 273, 285 Whole-genome duplication, 94, 271, 272, 277 347 348 349 X Xenopus, 391, 393, 394, 399, 401, 402, 404 350 351 Z Zebrafish, 353, 355, 364–366 Zygotene, 34–36 352 353 354 AQ2 356 F Chapter No.: PR OO 357 358 363 361 362 360 364 365 Query Refs. Details Required 370 368 367 369 371 AQ1 Please provide appropriate page number for the index term ‘ITS’. 376 373 374 375 AQ2 Please provide index terms with appropriate page numbers for chapters 4 and 9. Author’s Response CO RR EC TE D 377 UN Editor Proof Author Queries 355 MARKED PROOF Please correct and return this set Please use the proof correction marks shown below for all alterations and corrections. If you wish to return your proof by fax you should ensure that all amendments are written clearly in dark ink and are made well within the page margins. Instruction to printer Leave unchanged Insert in text the matter indicated in the margin Delete Textual mark under matter to remain New matter followed by or through single character, rule or underline or through all characters to be deleted Substitute character or substitute part of one or more word(s) Change to italics Change to capitals Change to small capitals Change to bold type Change to bold italic Change to lower case Change italic to upright type under matter to be changed under matter to be changed under matter to be changed under matter to be changed under matter to be changed Encircle matter to be changed (As above) Change bold to non-bold type (As above) Insert ‘superior’ character Marginal mark through letter or through characters through character or where required or new character or new characters or under character e.g. Insert ‘inferior’ character (As above) Insert full stop Insert comma (As above) Insert single quotation marks (As above) Insert double quotation marks (As above) over character e.g. (As above) or or (As above) Transpose Close up Insert or substitute space between characters or words Reduce space between characters or words linking and/or or or Insert hyphen Start new paragraph No new paragraph or characters through character or where required between characters or words affected and/or