Phylogenetic biogeography and taxonomy of

0 downloads 0 Views 743KB Size Report
May 18, 2017 - (Szweykowski et al., 2005). This ..... Fiodorow P, Odrzykoski I, Szweykowski J, Szweykowska- ... Acta Botanica Hungarica 29: 127–. 171.
See discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/230774164

Phylogenetic biogeography and taxonomy of disjunctly distributed bryophytes Article in Journal of Systematics and Evolution · September 2009 DOI: 10.1111/j.1759-6831.2009.00028.x

CITATIONS

READS

71

337

5 authors, including: Jochen Heinrichs

Kathrin Feldberg

Ludwig-Maximilians-University of Munich

Ludwig-Maximilians-University of Munich

235 PUBLICATIONS 3,717 CITATIONS

68 PUBLICATIONS 1,102 CITATIONS

SEE PROFILE

SEE PROFILE

Harald Schneider Xishuangbanna Tropical Botanical Garden 328 PUBLICATIONS 8,595 CITATIONS SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Pteridophyte Phylogeny Group View project

Australian Plagiochilaceae: who what where and how. View project

All content following this page was uploaded by Harald Schneider on 18 May 2017. The user has requested enhancement of the downloaded file. All in-text references underlined in blue are added to the original document and are linked to publications on ResearchGate, letting you access and read them immediately.

Journal of Systematics and Evolution

47 (5): 497–508 (2009)

doi: 10.1111/j.1759-6831.2009.00028.x

Phylogenetic biogeography and taxonomy of disjunctly distributed bryophytes 1

Jochen HEINRICHS∗ 1

1

J¨orn HENTSCHEL 1 Kathrin FELDBERG 2 Harald SCHNEIDER

1

Andrea BOMBOSCH

(Department of Systematic Botany, Albrecht von Haller Institute of Plant Sciences, Georg-August-University, D-37073 G¨ottingen, Germany) 2

(Botany Department, Natural History Museum, London SW7 5BD, UK)

Abstract More than 200 research papers on the molecular phylogeny and phylogenetic biogeography of bryophytes have been published since the beginning of this millenium. These papers corroborated assumptions of a complex genetic structure of morphologically circumscribed bryophytes, and raised reservations against many morphologically justified species concepts, especially within the mosses. However, many molecular studies allowed for corrections and modifications of morphological classification schemes. Several studies reported that the phylogenetic structure of disjunctly distributed bryophyte species reflects their geographical ranges rather than morphological disparities. Molecular data led to new appraisals of distribution ranges and allowed for the reconstruction of refugia and migration routes. Intercontinental ranges of bryophytes are often caused by dispersal rather than geographical vicariance. Many distribution patterns of disjunct bryophytes are likely formed by processes such as short distance dispersal, rare long distance dispersal events, extinction, recolonization and diversification. Key words bryophytes, cryptic speciation, disjunctions, divergence time estimates, Diversity Arrays Technology, DNA sequence variation, isozymes, molecular phylogeny. Bryophytes (liverworts, mosses and hornworts) comprise the three lineages of land plants with a life cycle in which the haploid gametophyte is the dominant photosynthetic active generation. In contrast to other land plants, the sporophyte is unbranched and not autonomously viable (Schofield, 1985). Bryophytes disperse frequently both by spores and by propagules that descend from the gametophyte, or by unspecialized gametophyte fragments with a high potential of regeneration (Correns, 1899). Bryophytes are the progeny of the first plants that successfully colonized terrestrial habitats (Qiu, 2008). Their evolution in space and time is still insufficiently known. In the nineteenth and early twentieth centuries, bryologists preferred to use a geographical or typological species concept where species were defined as largely invariant units. Many species were known only from type material (e.g., Stephani, 1898–1925; Warnstorf, 1911). More recently, authors accepted intraspecific morphological variation and lowered numerous local taxa to synonyms of widespread bryophyte species (Gradstein, 1994; Buck, 1998; Heinrichs, 2002). Consequently, broad geographical ranges that often span



 C

Received: 11 November 2008 Accepted: 26 February 2009 Author for correspondence. E-mail: [email protected]; Tel.: +49551-39-22220; Fax: +49-551-39-22329.

2009 Institute of Botany, Chinese Academy of Sciences

several continents were assigned to many morphologically circumscribed bryophyte species (Herzog, 1926; Grolle, 1969; Gradstein et al., 1983; Schofield, 1992; Tan & P´ocs, 2000). The resulting disjunct ranges of bryophyte species have frequently been explained by ancient vicariance and slow rates of morphological evolution (Herzog, 1926; Crum, 1972; Frey et al., 1999) but other authors provided experimental evidence for the alternative scenario of successful long distance dispersal of bryophytes by spores (van Zanten, 1978; van Zanten & Gradstein, 1988). The increasing availability of DNA sequence data now enables the testing of morphology-based taxonomic and biogeographic concepts and the disclosure of the genotype structure of species. DNA sequence data also allow for an evaluation of different hypotheses concerning biogeographical patterns and processes.

1 Morphological species concepts in the light of molecular phylogenies: indications of nonmonophyly and need for a revised taxonomy An increasing body of published work points to many taxonomic incongruences of widespread morphologically circumscribed bryophyte species and phylogenies derived from molecular markers. Shaw & Allen (2000) resolved populations of morphologically

498

Journal of Systematics and Evolution

Vol. 47

No. 5

circumscribed species of the aquatic moss genus Fontinalis Hedw. in widely diverging clades, as did Vanderpoorten et al. (2004) for species of Hygroamblystegium Loeske and Shaw et al. (2008) for representatives of the Sphagnum subsecundum complex. Stech & Wagner (2005) provided evidence for the polyphyly of several species of Campylopus Brid. Werner & Guerra (2004) resolved Tortula vahliana (Schultz) Mont. nested within Tortula muralis Hedw. Draper et al. (2007) showed the non-monophyly of the pleurocarpous moss species Isothecium alopecuroides (Dubois) Isov., Isothecium holtii Kindb. and Isothecium myosuroides Brid. Vanderpoorten & Goffinet (2006) identified several species of the moss Brachytheciastrum Ignatov & Huttunen as polyphyletic and documented parallel morphological evolution within this genus. Similarly, Cano et al. (2005) found incongruences between current classification schemes of the Tortula subulata complex, and a molecular topology. Incongruences between morphology-based classifications and molecular topologies have also been shown for several liverwort genera including Bryopteris (Nees) Lindenb. (Hartmann et al., 2006), Chiloscyphus Corda (Hentschel et al., 2006), Herbertus Gray (Feldberg et al., 2004), Lophozia (Dumort.) Dumort. (Vilnet et al., 2008), Plagiochila (Dumort.) Dumort. (Rycroft et al., 2004), and Porella L. (Hentschel et al., 2007b). As a consequence of the numerous observations of species polyphyly, Vanderpoorten & Goffinet (2006) raised reservations regarding current species definitions. However, in many cases the molecular topologies allowed for a modified appraisal of morphological evidence including new circumscription of taxa or changes of rank. Vanderpoorten (2004) solved the problem of non-monophyletic Hygroamblystegium species by introducing a wide species concept for Hygroamblystegium varium (Hedw.) M¨onk. Cano et al. (2005) proposed the binomen Tortula schimperi Cano et al. for a taxon that was usually treated as a variety of T. subulata Hedw. Rycroft et al. (2004) reinstated the leafy liverwort Plagiochila maderensis Steph. that was earlier placed in the synonymy of Plagiochila spinulosa (Dicks.) Dumort. Attempts to establish monophyletic entities may be hampered by reticulate evolution (Shaw & Goffinet, 2000; Natcheva & Cronberg, 2004, 2007; Shaw et al., 2008). Introgression, hybridization and incomplete lineage sorting may contradict a taxonomy that is strictly based on the monophyly concept. Redefinitions of species are often connected with changes of distribution range concepts. Pfeiffer et al. (2004) showed phylogeographic structure within the Australasian–South American simple thalloid liverwort Hymenophyton flabellatum (Labill.) Trev. Based on the

2009

outcome of their molecular phylogenetic analyses and accompanying morphological studies, they restricted the range of H. flabellatum to Australasia, and reinstated the New Zealand–Tasmanian Hymenophyton leptopodum (Hook.f. & Taylor) A.Evans as well as the southern South American Hymenophyton pedicellatum Steph. Based on molecular evidence, Feldberg et al. (2004, 2007) excluded Herbertus dicranus (Taylor) Trevis. from tropical America and proposed the application of the name Herbertus sendtneri (Nees) A.Evans, a taxon that was previously assigned to European and Asian populations only. Heinrichs et al. (1998, 2004, 2005a, b) lowered the European Plagiochila killarniensis Pearson to a synonym of the Neotropical Plagiochila bifaria (Sw.) Lindenb., included several African and Neotropical binomia in the European taxon Plagiochila punctata (Taylor) Taylor, and extended the range of the Neotropical Plagiochila boryana Steph. to tropical Africa. Heden¨as (2008a) excluded the African Antitrichia kilimandscharica Broth. and the western North American Antitrichia gigantea (Sull. & Lesq.) Kindb. from the synonymy of Antitrichia curtipendula (Hedw.) Brid. It is assumed that ongoing studies into the molecular phylogeny of bryophytes will lead to numerous new appraisals of ranges.

2 Internal structure of bryophyte species: molecular variation versus morphological stasis Sequencing of variable nuclear or chloroplast markers of multiple accessions of bryophytes usually reveals a phylogenetic structure that follows a geographical rather than a morphological pattern (Shaw & Allen, 2000; Skotnicki et al., 2004; Grundmann et al., 2006; Hartmann et al., 2006; Vanderpoorten & Long, 2006; Feldberg et al., 2007; Hentschel et al., 2007b; Heden¨as, 2008a, b; Heden¨as & Elden¨as, 2007; Huttunen et al., 2008). Genetic variation without concordant morphological variation has often been regarded as an indication of cryptic speciation (Shaw, 2001; Fernandez et al., 2006; Bickford et al., 2007). However, it is not yet clear if the molecular variation that has been documented for many morphospecies of bryophytes always reflects genetically incompatible units. An apparent lack of interchange of genetic material could also be a result of a geographic or ecological separation of populations that still hold the potential to interbreed successfully. If future studies allow for a more definite decision on hybridization capability of  C

2009 Institute of Botany, Chinese Academy of Sciences

HEINRICHS et al.: Phylogenetic biogeography of bryophytes

bryophyte populations, a refined taxonomy including a partial return to the geographical species concept seems possible, especially in taxa in which different clades can be assigned to clear-cut ranges. Application of the geographical species concept is contradicted by indications of infrequent long distance dispersal events in many bryophyte lineages (Skotnicki et al., 2001; McDaniel & Shaw, 2005; Hentschel et al., 2007b; Huttunen et al., 2008). Even more difficult is the formal recognition of different sympatric genotypes that show no or very limited morphological differences. The complex thalloid liverwort Conocephalum F.H.Wigg. is possibly the most exhaustively studied example of a genetically heterogeneous bryophyte species. Isozyme studies in the Holarctic–temperate Asian Conocephalum conicum (L.) Dumort. s.l. revealed the presence of six partly sympatric taxa that were informally named using a letter system (Odrzykoski & Szweykowski, 1991). Subsequent detailed study of morphological and ecological traits led to the formal recognition of one of them as Conocephalum salebrosum Szweyk. et al. (Szweykowski et al., 2005). This species occurs sympatrically with C. conicum s.str. but tends to grow in dryer habitats. Boisselier-Dubayle et al. (1998) reported the presence of a morphologically indistinct Mediterranean sibling species besides the European–Asian–American–African complex thalloid liverwort Reboulia hemisphaerica (L.) Raddi s.str. Similarly, the subcosmopolitan simple thalloid liverwort Aneura pinguis (L.) Dumort. s.l. seems to include at least three reproductively isolated, sympatric cryptic species (Wachowiak et al., 2007). Evidence for cryptic speciation was also provided for the Holarctic simple thalloid liverworts Pellia epiphylla (L.) Corda and Pellia endiviifolia (Dicks.) Dumort. (Pacak & SzweykowskaKuli´nska, 2000; Fiodorow et al., 2001). Similar findings are available for several mosses. Bijlsma et al. (2000) provided evidence for the presence of two reproductively isolated cryptic species in the acrocarpous moss Polytrichum commune Hedw. [P. commune s.str., Polytrichum uliginosum (Wallr.) Schriebl, see also van der Velde & Bijlsma, 2004)]. McDaniel & Shaw (2003) recovered a deep split between the two morphologically weakly separated subspecies of the trans-Antarctic moss Hymenodontopsis mnioides (Hook.) N.E.Bell et al. Fernandez et al. (2006) analyzed amplified fragment length polymorphisms of Californian populations of the cosmopolitan species Grimmia laevigata (Bridel) Bridel. They identified two distinct geographically overlapping clades. Shaw (2000) published a nrITS phylogeny of Mielichhoferia elongata (Hoppe & Hornsch.) Nees & Hornsch. Based on the outcome of his analyses he proposed two cryptic species  C

2009 Institute of Botany, Chinese Academy of Sciences

499

within the morphologically uniform taxon, one with a European–North American range, and the other strictly North American. Heden¨as & Elden¨as (2007) investigated nrITS and chloroplast DNA haplotype variation of the pleurocarpous moss Hamatocaulis vernicosus (Mitt.) Heden¨as. Based on their topologies Heden¨as & Elden¨as (2007) postulated the existence of two cryptic species, of which one is widespread in Europe, in addition to a few North American records. The other cryptic species was found south of the boreal zone in Europe, in northern-most Asiatic Russian Federation, and Peru.

3 Phylogeographic patterns, migration routes, and modes of reproduction as revealed from molecular data Studies of haplotype variation do not only allow for the recognition of putative cryptic taxa but also for a reconstruction of the spatial structure of genetic diversity, potential bottleneck events, and modes of reproduction. Grundmann et al. (2007) studied DNA sequence and allozyme variation to resolve the spatial structure of Mediterranean accessions of the dioecious Holarctic moss Pleurochaete squarrosa (Brid.) Lindb. These authors observed a decline of intraspecific variation from west to east but no difference in gene diversity among populations from islands and mainland areas. Based on the latter observation, Grundmann et al. (2007) concluded that the large Mediterranean islands might function as “mainland” for bryophytes. Vanderpoorten et al. (2008) arrived at a similar conclusion when analyzing chloroplast markers of the moss Grimmia montana Bruch & Schimp. Madeiran and Mediterranean haplotypes of G. montana were identical or closely related to European or North American ones. Grundmann et al. (2008) analyzed diversity patterns of European P. squarrosa using nuclear and chloroplast DNA sequences, and enzyme electrophoresis. These authors provided evidence for sexual reproduction, that is, recombination, of P. squarrosa in the Mediterranean Basin and the Kaiserstuhl Mountains in southwestern Germany, a region that is well known for its unusual climate with high monthly average temperatures and short, mild winter. In other regions of central and northwest Europe P. squarrosa disperses predominantly by vegetative propagules, a finding that is corroborated by a lack of evidence of recombination. Grundmann et al. (2008) also postulated a postglacial recolonization of central Europe from the Iberian Peninsula and the Balkans. Cronberg (2000) observed quite similar patterns for the epiphytic moss Leucodon

500

Journal of Systematics and Evolution

Vol. 47

No. 5

sciuroides (Hedw.) Schw¨agr. Mediterranean populations reproduce sexually and are genetically diverse whereas more northern populations reproduce vegetatively and are genetically quite uniform. This pattern coincides with the expectation of a loss of genetic variation in populations at the northern limit of the glacial refugia. Glacial survival in southern Europe is obviously not a general pattern in bryophytes. Hedderson & Nowell (2006) recognized several unique Homalothecium sericeum (Hedw.) Schimp. haplotypes in the British Isles and adjacent mainland. Based on this observation Hedderson & Nowell (2006) assumed a survival of Homalothecium sericeum in this region during the last glacial period. Sz¨ov´eni et al. (2006) presented a chloroplast phylogeographic analysis of Sphagnum fimbriatum Wilson and Sphagnum squarrosum Crome. Their haplotype distribution patterns seem to support different dispersal scenarios for these species. S. fimbriatum seems to have survived the last glacial period along the Atlantic coast of Europe, and rapidly colonized Europe after the last glacial maximum. In contrast, S. squarrosum obviously had numerous scattered refugia throughout Europe. Although most studies referring to the internal structure of widespread bryophyte species revealed molecular variation, a few examples contradicted this tendency. James et al. (2008) introduced the Diversity Arrays Technology, a hybridization-based genotyping method, to reproducibly detect largely low-copy genomic variation in ferns and mosses. Their study revealed a lack of phylogenetic pattern in the Australian moss Garovaglia elegans (Dozy & Molk.) Bosch & Sande Lac. Similarly, van der Velde & Bijlsma (2003) found nearly no genetic structure among European populations of several Polytrichum species (Polytrichum commune, Polytrichum uliginosum, Polytrichum formosum Hedw., and Polytrichum piliferum Hedw.). The lack of allozyme and microsatellite variation pointed to extensive spore dispersal and contradicted the hypothesis of a recolonization of Europe from southern refugia after the last glacial period. Inter- or intraspecific variation of molecular markers might allow for the reconstruction of range expansion directions. Based on the recent distribution of taxa and their position in a phylogenetic framework, conclusions can be drawn as to the ranges of their ancestors. Heinrichs et al. (2005a) resolved an African accession of Plagiochila sect. Hylacoetes Carl within several tropical American accessions. Based on this topology, these authors proposed a Neotropical origin of the African Plagiochila sect. Hylacoetes populations. Hartmann et al. (2006) arrived at the same conclusion for the Madagascar–R´eunion endemic Bryopteris gau-

2009

dichaudii Gottsche. Feldberg et al. (2007) studied the phylogenetic biogeography of Herbertus and provided evidence for a colonization of Africa from Neotropical and Asian populations (Fig. 1). Huttunen et al. (2008) discovered evidence for a western North American origin of a Holarctic Homalothecium clade. Taxonomic work combined with the reconstruction of migration routes provided new insights into the relationships between floristic regions. Contrary to the earlier belief, the Atlantic European and Macaronesian Plagiochila taxa are connected with Neotropical rather than Asiatic taxa (Heinrichs et al., 2006). Vanderpoorten & Long documented relationships of Macaronesian and Neotropical Leptoscyphus Mitt. Similarly, Stech et al. (2007) showed the close relationships of the Campylopus flora of tropical America and Madeira. This trend is contradicted by the example Porella with the Macaronesian endemic Porella inaequalis Perss. closely related to the Asian Porella grandiloba Lindb. rather than to Neotropical species (Hentschel et al., 2007b). Clearly more case studies are necessary to decide whether each species has its own history or whether there are recurring patterns.

4 Intercontinental range expansion versus conservation Wegener’s (1915) reconstruction of continental drift has provided a theory to explain disjunct distribution patterns of plants. Some bryologists (Stotler & Crandall-Stotler, 1974; Gradstein et al., 1983; Schuster, 1979; Frey et al., 1999) linked ranges of bryophytes to continental movement. The underlying assumption of geographical vicariance predicts an origin of many disjunctly distributed bryophyte species/genera at least in the late Mesozoic. However, accurate morphology-based insights into the historical biogeography of bryophytes would require a continuous fossil record that is not available (Krassilov & Schuster, 1984). The long-distance dispersal ability of bryophytes (van Zanten, 1978; van Zanten & Gradstein, 1988) contradicts strict vicariance scenarios, as does the frequent occurrence of widespread bryophyte species on oceanic islands (Heinrichs et al., 2006; Vanderpoorten et al., 2007). Molecular phylogenies enable us to scrutinize evidence for dispersal or vicariance. One possibility concerning this matter is a critical comparison of phylogenies with breakup events of landmasses. Congruence of ancestral geographical distributions of clades and the sequence of breakup events rather supports vicariance. However, deviation of a phylogeny and a breakup  C

2009 Institute of Botany, Chinese Academy of Sciences

HEINRICHS et al.: Phylogenetic biogeography of bryophytes

501

Fig. 1. Molecular phylogeny of the leafy liverwort Herbertus with the reconstruction of putative migration routes and dispersal events. The distribution of accessions within “clade A” indicates a dispersal event from tropical America to Africa. An ancestral area reconstruction points to an Asian origin of “clade B” (Reproduced from Feldberg et al., 2007 with permission from Wiley-Blackwell).

 C

2009 Institute of Botany, Chinese Academy of Sciences

502

Journal of Systematics and Evolution

Vol. 47

No. 5

sequence supports dispersal (Sanmartin & Ronquist, 2004). The application of the molecular clock hypothesis allows for a transformation of a phylogram into a chronogram. Sequences usually do not exactly evolve at a constant rate but the tempo of mutations within a molecular marker seems to move in a more or less determined range. Accordingly, a comparison of sequence variation with that of dated phylogenies may shed some light on the likelihood of different scenarios (Les et al., 2003). Divergence time estimates based on sequence variation and the fossil record will provide more reliable insights into the historical biogeography of lineages by enabling differentiation between coinciding events in time and pseudo-congruent patterns (Donoghue & Moore, 2003; Renner, 2005). The poor Paleozoic and Mesozoic fossil record of bryophytes (Oostendorp, 1987) is a serious challenge to any study on the timescale of bryophyte diversification. Cenozoic fossils are often preserved in the form of amber inclusions (Grolle & Meister, 2004; Frahm & Newton, 2005). Amber has not been formed continuously throughout the history of plants on land (Grimaldi, 1996). The infrequent generation of amber inclusions makes it difficult to use them as calibration points for dating approaches (Hartmann et al., 2006). Therefore we rely heavily on the variation of substitution rates. Supraspecific intercontinental ranges are generally accepted for bryophytes and have been confirmed on many occasions (Meissner et al., 1998; Groth et al., 2003). Several recent studies also confirmed the monophyly of intercontinentally distributed bryophyte species. Neotropical–African ranges have been supported by molecular data, for example, for species of the liverwort genera Chiloscyphus (Hentschel et al., 2007a), Herbertus (Feldberg et al., 2007), Pallavicinia Gray (Schaumann et al., 2005), Plagiochila (Heinrichs et al., 2005a), and Porella (Hentschel et al., 2007b), and the moss Campylopus (Stech & Wagner, 2005). Similar results are available for North American– European disjunctions, for example, for species of the moss genera Anacolia Schimp. (Werner et al., 2003), Claopodium (Lesq. & James) Renauld & Cardot, Dicranoweisa Milde, Scleropodium Bruch & Schimp. (Shaw et al., 2003), Homalothecium (Huttunen et al., 2008), and species of the liverwort genera Frullania Raddi (Hentschel et al., 2009), Herbertus (Feldberg et al., 2007), Pallavicinia (Schaumann et al., 2005), Porella (Hentschel et al., 2007b), and Scapania (Dumort.) Dumort. (De Roo et al., 2007). Neotropical–European disjunctions have been supported for species of the liverwort genera Lophozia (De Roo et al., 2007) and

2009

Plagiochila (Heinrichs et al., 2004), and Asian–North American ranges for Porella species (Hentschel et al., 2007b). The monophyly of even broader species ranges has been shown, for example, for the moss Tortella humilis (Hedw.) Jenn. (Neotropics, North America, Europe) (Werner et al., 2005) or the liverwort Plagiochila punctata (North America, Neotropics, Africa, Europe) (Davison et al., 2006). The observed sequence analogies of different accessions of intercontinentally distributed bryophytes led many authors to assume long distance dispersal as a feasible explanation for the disjunct ranges (Shaw et al., 2003, 2008; Forrest et al., 2005; Heinrichs et al., 2005a; Feldberg et al., 2007; Hentschel et al., 2007b; Huttunen et al., 2008). A few authors insisted on geographical vicariance and explained similar sequences from different parts of a disjunct range with “stenoevolution”, that is, slow rates of molecular evolution (Frey et al., 1999). Shaw et al. (2003) tested the likelihood of a Madrean–Tethyan origin of several western North American–Mediterranean disjunctions of mosses. These authors stated that “no plausible mutation rate” would link the disjunctions to early Miocene times, and favored recent dispersal as an explanation of the observed distribution ranges. Hentschel et al. (2006) verified the assumption of a recent introduction of the southern hemispheric Chiloscyphus semiteres (Lehm.) Lehm. & Lindenb. into Europe (Paton, 1965) by demonstrating nrITS sequence similarities of accessions from the Netherlands and Australia. Wall (2005) documented a clock-like behaviour of the nuclear glyceraldehyde 3phosphate dehydrogenase gene of the moss Mitthyridium H.Rob., and identified clades that were related to oceanic archipelagos. Using island ages as calibration points, he provided evidence for a diversification of Mitthyridium within less than ten million years. Huttunen et al. (2008) used a mean nrITS substitution rate to deduce a late Miocene–Pliocene age of a split between North American and Eurasian Homalothecium. Hartmann et al. (2006) published a chronogram for the liverwort Bryopteris and provided hypothetical timescales based on assumptions of different scenarios, and nrITS mutation rates. These authors could clearly reject western Gondwanan vicariance for the Neotropical–African range of Bryopteris, and proposed a dispersal scenario. Heinrichs et al. (2006) reconstructed the molecular phylogeny of the cosmopolitan leafy liverwort Plagiochila and presented timescales based on putative fossil assignments. The results did not contradict a Gondwanan origin of Plagiochila, but the geographical distribution of clades (Fig. 2) and divergence time estimates rendered Gondwanan vicariance unlikely. Heinrichs et al. (2006) explained the observed  C

2009 Institute of Botany, Chinese Academy of Sciences

HEINRICHS et al.: Phylogenetic biogeography of bryophytes

503

Fig. 2. Molecular phylogeny of the leafy liverwort Plagiochila. Distribution of species is indicated at branches. 1, Australasia; 2, Southern South America; 3, Subantarctics; 4, Neotropics; 5, Asia; 6, Western Holarctics; 7, Eastern Holarctics; 8, Africa; 9, Hawaiian Islands. Modified from Heinrichs et al. (2006). BT, bootstrap.

 C

2009 Institute of Botany, Chinese Academy of Sciences

504

Journal of Systematics and Evolution

Vol. 47

No. 5

2009

distribution patterns as a result of short distance dispersal, rare long distance dispersal events, extinction, recolonization and diversification. It is not yet clear if the situation in Plagiochila represents a general pattern but several other studies seem to support this combination of mechanisms (Feldberg et al., 2007; Huttunen et al., 2008). Support also comes from comparisons of Southern Hemispheric ranges of bryophytes and main wind directions. Mu˜noz et al. (2004) found a stronger correlation of floristic similarities with wind connectivity than with geographic proximities, and therefore favored wind as a dispersal vector for many Southern Hemispheric biota. However, today’s disjunct ranges of some Southern Hemispheric taxa such as Monoclea Hook. (Meissner et al., 1998) could be a result of short distance dispersal before the final disassembly of Gondwana, and subsequent range fragmentation as a result of climate changes (Schuster, 1979). Dated chronograms, based on sequence variation plus the bryophyte fossil record, have been published for bryophytes in general (Newton et al., 2007, with a strong focus on the pleurocarpous moss lineage), the leafy liverworts (Heinrichs et al., 2007) and the leafy liverwort family Lejeuneaceae (Wilson et al., 2007). However, sampling within these studies was not sufficient to decide on species level disjunctions, and the results supported the idea of a reformation of bryophyte diversity throughout the Cretaceaous and Early Tertiary. Many recent genera seem to have originated not before the Late Cretaceous, rendering Gondwanan vicariance rather unlikely. Although most current authors favor the adoption of infrequent long distance dispersal for disjunct ranges, this hypothesis needs to be tested by further studies that should focus both on a better understanding of the bryophyte fossil record and a more comprehensive taxon sampling.

The relationships of many deep clades of bryophytes have been clarified using molecular phylogenetic approaches plus morphology (Renzaglia et al., 2007). In contrast, boundaries and relationships of many species and genera of bryophytes are still unclear, as is the genetic structure of most species. Future studies should focus strongly on genus or species level relationships, and shed more light on reproduction modes of populations that are still insufficiently known. Population genetic studies using isozymes and microsatellites might result in reliable reconstructions of migration routes and refugia of bryophytes. These studies could also shed more light on the justification of assumptions of intercontinental ranges of bryophyte species. Phylogenetic studies based on variable molecular markers and using divergence time estimates usually support dispersal scenarios rather than geographical vicariance as the preferred explanation of disjunct ranges of bryophytes. The number of studies is still insufficient for a general pattern to emerge, or to determine if there are also patterns that are indicative of geographical vicariance. Future studies should focus not only on sequence variation but also on the fossil record of bryophytes. The search of bryophyte inclusions in Cretaceous amber deposits may be a promising approach to our understanding of the origin of extant bryophyte diversity. Progress in the interpretation of the bryophyte fossil record is essential to achieve more reliable insights into the chronology of bryophyte diversification and distribution range formation.

5

References

State of the art and perspectives

A three-digit number of papers on the molecular phylogeny and phylogenetic biogeography of bryophytes have been published since Shaw’s review on biogeographical patterns and cryptic speciation of bryophytes (Shaw, 2001). On the one hand, these papers corroborated assumptions of a complex genetic structure of bryophytes with a uniform morphology, and raised reservations against many morphologically justified species concepts, especially within the mosses. On the other hand, many molecular topologies allowed for corrections and modifications of morphological classification schemes.

Acknowledgements We thank Jun Wen (Washington DC), Yin-Long Qiu (Ann Arbor, Michigan) and Yan Liang (Beijing) for comments on the manuscript. Financial support from the German Research Foundation (grants HE 3584/1-4) is gratefully acknowledged.

Bickford D, Lohmann DJ, Sodhi NS, Ng PKL, Meier R, Winker K, Ingram KK, Das I. 2007. Cryptic species as a window on diversity and conservation. Trends in Ecology and Evolution 22: 148–155. Bijlsma R, Van Der Velde M, van de Zande L, Boerema AC, van Zanten BO. 2000. Molecular markers reveal cryptic species within Polytrichum commune (common hair-cap moss). Plant Biology 2: 408–414. Boisselier-Dubayle MC, Lambourdiere J, Bischler H. 1998. Taxa delimitation in Reboulia investigated with morphological, cytological, and isozyme markers. The Bryologist 101: 61– 69. Buck, WR. 1998. Pleurocarpous mosses of the West Indies. New York: The New York Botanical Garden.  C

2009 Institute of Botany, Chinese Academy of Sciences

HEINRICHS et al.: Phylogenetic biogeography of bryophytes Cano MJ, Werner O, Guerra J. 2005. A morphometric and molecular study in Tortula subulata complex (Pottiaceae, Bryophyta). Botanical Journal of the Linnean Society 149: 333–350. Correns C. 1899. Untersuchungen u¨ ber die Vermehrung der Laubmoose durch Brutorgane und Stecklinge. Jena: Fischer. Cronberg N. 2000. Genetic diversity of the epiphytic bryophyte Leucodon sciuroides in formerly glaciated versus nonglaciated parts of Europe. Heredity 84: 710– 720. Crum HA. 1972. The geographic origins of the mosses of North America’s eastern deciduous forest. Journal of the Hattori Botanical Laboratory 35: 269–298. Davison PG, Smith D, Feldberg K, Lindner M, Heinrichs J. 2006. Plagiochila punctata (Jungermanniopsida: Plagiochilaceae) in Tennessee, new to North America. The Bryologist 109: 242–246. De Roo RT, Hedderson TA, S¨oderstr¨om L. 2007. Molecular insights into the phylogeny of the leafy liverwort family Lophoziaceae Cavers. Taxon 56: 301–314. Donoghue MJ, Moore BR. 2003. Toward an integrative historical biogeography. Integrative and Comparative Biology 43: 261–270. Draper I, Heden¨as L, Grimm GW. 2007. Molecular and morphological incongruence in European species of Isothecium (Bryophyta). Molecular Phylogenetics and Evolution 42: 700–716. Feldberg K, Groth H, Wilson R, Sch¨afer-Verwimp A, Heinrichs J. 2004. Cryptic speciation in Herbertus (Herbertaceae, Jungermanniopsida): Range and morphology of Herbertus sendtneri inferred from nrITS sequences. Plant Systematics and Evolution 249: 247–261. Feldberg K, Hentschel J, Wilson R, Rycroft DS, Glenny D, Heinrichs J. 2007. Phylogenetic biogeography of the leafy liverwort Herbertus (Jungermanniales, Herbertaceae) based on nuclear and chloroplast DNA sequence data: correlation between genetic variation and geographical distribution. Journal of Biogeography 34: 688–698. Fernandez CC, Shevock JR, Glazer AN, Thomspon JN. 2006. Cryptic species within the cosmopolitan desiccationtolerant moss Grimmia laevigata. Proceedings of the National Academy of Sciences USA 103: 637–642. Fiodorow P, Odrzykoski I, Szweykowski J, SzweykowskaKuli´nska Z. 2001. Phylogeny of the European species of the genus Pellia (Hepaticae; Metzgeriales) based on the molecular data from nuclear tRNALeu (CAA) intergenic sequences. Gene 262: 309–315. Forrest LL, Schuette SW, Crandall-Stotler BJ, Stotler RE. 2005. A molecular study of the simple thalloid liverwort Jensenia (Marchantiophyta, Pallaviciniaceae). The Bryologist 108: 204–211. Frahm JP, Newton AE. 2005. A new contribution to the moss flora of Dominican amber. The Bryologist 108: 526–536. Frey W, Stech M, Meissner K. 1999. Chloroplast DNA relationship in palaeoaustral Lopidium concinnum (Hypopterygiaceae, Musci.), an example of stenoevolution in mosses. Studies in austral temperate rain forest bryophytes 2. Plant Systematics and Evolution 218: 67–75. Gradstein SR. 1994. Lejeuneaceae: Ptychantheae. Brachiolejeuneae. Flora Neotropica Monograph 62: 1–225.  C

2009 Institute of Botany, Chinese Academy of Sciences

505

Gradstein SR, P´ocs T, V´anˇ a J. 1983. Disjunct Hepaticae in tropical America and Africa. Acta Botanica Hungarica 29: 127– 171. Grimaldi DA. 1996. Amber: window to the past. New York: American Museum of National History. Grolle R. 1969. Großdisjunktionen in Artarealen lateinamerikanischer Lebermoose. In: Fittkau EJ ed. Biogeography and ecology in South America. The Hague: Junk. 562–582 Grolle R, Meister K. 2004. The Liverworts in Baltic and Bitterfeld Amber. Jena: Weissdorn. Groth H, Lindner M, Wilson R, Hartmann FA, Schmull M, Gradstein SR, Heinrichs J. 2003. Biogeography of Plagiochila (Hepaticae): natural species groups span several floristic kingdoms. Journal of Biogeography 30: 965–978. Grundmann M, Schneider H, Russel SJ, Vogel JC. 2006. Phylogenetic relationships of the moss genus Pleurochaete Lindb. (Bryales: Pottiaceae) based on chloroplast and nuclear genomic markers. Organisms, Diversity and Evolution 6: 33–45. Grundmann M, Ansell SW, Russell SJ, Koch MA, Vogel JC. 2007. Genetic structure of the widespread and common Mediterranean bryophyte Pleurochaete squarrosa (Brid.) Lindb. (Pottiaceae) – evidence from nuclear and plastidic DNA sequence variation and allozymes. Molecular Ecology 16: 709–722. Grundmann M, Ansell SW, Russell SJ, Koch MA, Vogel JC. 2008. Hotspots of diversity in a clonal world – the Mediterranean moss Pleurochaete squarrosa in Central Europe. Molecular Ecology 17: 825–838. Hartmann FA, Wilson R, Gradstein SR, Schneider H, Heinrichs J. 2006. Testing hypotheses on species delimitations and disjunctions in the liverwort Bryopteris (Jungermanniopsida: Lejeuneaceae). International Journal of Plant Sciences 167: 1205–1214. Heden¨as L. 2008a. Molecular variation and speciation in Antitrichia curtipendula s.l. (Leucodontaceae, Bryophyta). Botanical Journal of the Linnean Society 156: 341–354. Heden¨as L. 2008b. Molecular variation in Drepanocladus aduncus s.l. does not support recognition of more than one species in Europe. Journal of Bryology 30: 108–120. Heden¨as L, Elden¨as P. 2007. Cryptic speciation, habitat differentation, and geography in Hamatocaulis vernicosus. Plant Systematics and Evolution 268: 131–145. Hedderson TA, Nowell TL. 2006. Phylogeography of Homalothecium sericeum (Hedw.) Br. Eur.; toward a reconstruction of glacial survival and postglacial migration. Journal of Bryology 28: 283–292. Heinrichs J. 2002. A taxonomic revision of Plagiochila sect. Hylacoetes, sect. Adiantoideae and sect. Fuscoluteae in the Neotropics with a preliminary subdivision of Neotropical Plagiochilaceae into nine lineages. Bryophytorum Bibiliotheca 58: 1–184, App. 1–5. Heinrichs J, Grolle R, Drehwald U. 1998. The conspecificity of Plagiochila killarniensis Pearson and P. bifaria (Sw.) Lindenb. (Hepaticae). Journal of Bryology 20: 495–528. Heinrichs J, Groth H, Lindner M, Feldberg K, Rycroft DS. 2004. Molecular, morphological and phytochemical evidence for a broad species concept of Plagiochila bifaria (Sw.) Lindenb. (Hepaticae). The Bryologist 107: 28–40.

506

Journal of Systematics and Evolution

Vol. 47

No. 5

Heinrichs J, Hentschel J, Wilson R, Feldberg K, Schneider H. 2007. Evolution of leafy liverworts (Jungermanniidae, Marchantiophyta): estimating divergence times from chloroplast DNA sequences using penalized likelihood with integrated fossil evidence. Taxon 56: 31–44. Heinrichs J, Lindner M, Gradstein SR, Groth H, Buchbender V, Solga A, Fischer E. 2005a. Origin and subdivision of Plagiochila (Jungermanniidae: Plagiochilaceae) in tropical Africa based on evidence from nuclear and chloroplast DNA sequences and morphology. Taxon 54: 317–333. Heinrichs J, Lindner M, Groth H, Renker C. 2005b. Distribution and synonymy of Plagiochila punctata (Taylor) Taylor, with hypotheses on the evolutionary history of Plagiochila sect. Arrectae (Plagiochilaceae, Hepaticae). Plant Systematics and Evolution 250: 105–117. Heinrichs J, Lindner M, Groth H, Hentschel J, Feldberg K, Renker C, Engel JJ, von Konrat M, Long DG, Schneider H. 2006. Goodbye or welcome Gondwana? Insights into the phylogenetic biogeography of the leafy liverwort Plagiochila with a description of Proskauera, gen. nov. (Plagiochilaceae, Jungermanniales). Plant Systematics and Evolution 258: 211–226. Hentschel J, Feldberg K, Z¨undorf HJ, Hellwig FH, Schneider H, Heinrichs J. 2007a. The systematic position of Pachyglossa and Clasmatocolea (Jungermanniopsida: Lophocoleaceae) inferred from nrDNA ITS sequences and morphology. Taxon 56: 1136–1142. Hentschel J, von Konrat M, P´ocs T, Sch¨afer-Verwimp A, Shaw AJ, Schneider H, Heinrichs J. 2009. Molecular insights into the phylogeny and subgeneric classification of Frullania Raddi (Frullaniaceae, Porellales). Molecular Phylogenetics and Evolution 52: 142–156. Hentschel J, Zhu RL, Long DG, Davison PG, Schneider H, Gradstein SR, Heinrichs J. 2007b. A phylogeny of Porella (Porellaceae, Jungermanniopsida) based on nuclear and chloroplast DNA sequences. Molecular Phylogenetics and Evolution 45: 693–795. Hentschel J, Z¨undorf HJ, Hellwig FH, Sch¨afer-Verwimp A, Heinrichs J. 2006. Taxonomic studies in Chiloscyphus Corda (Jungermanniales: Lophocoleaceae) based on nrITS sequences and morphology. Plant Systematics and Evolution 262: 125–137. Herzog T. 1926. Geographie der Moose. Jena: Fischer. Huttunen S, Heden¨as L, Ignatov M, Devos N, Vanderpoorten A. 2008. Origin and evolution of the northern hemisphere disjunction in the moss genus Homalothecium. American Journal of Botany 95: 720–730. James KE, Schneider H, Ansell SW, Evers M, Robba L, Uszynski G, Pedersen N, Newton AE, Russell SJ, Vogel JC, Kilian A. 2008. Diversity Arrays Technology (DArT) for pan-genomic evolutionary studies of non-model organisms. PLoS ONE 3: e1682. doi:10.1371/journal.pone.0001682. Krassilov VA, Schuster RM. 1984. Paleozoic and mesozoic fossils. In: Schuster RM ed. New manual of bryology, vol. 2. Nichinan: The Hattori Botanical Laboratory. 1172– 1193. Les DH, Crawford DJ, Kimball RT, Moody ML, Landolt E. 2003. Biogeography of discontinuously distributed hydrophytes: a molecular appraisal of intercontinental disjunctions. International Journal of Plant Sciences 164: 917–932.

2009 McDaniel SF, Shaw AJ. 2003. Phylogeographic structure and cryptic speciation in the trans-antarctic moss Pyrrhobryum mnioides. Evolution 57: 205–215. McDaniel SF, Shaw AJ. 2005. Selective sweeps and intercontinental migration in the cosmopolitan moss Ceratodon purpureus (Hedw.) Brid. Molecular Ecology 14: 1121– 1132. Meissner K, Frahm JP, Stech M, Frey W. 1998. Molecular divergence patterns and infragenetic relationship of Monoclea (Monocleales, Hepaticae) – studies in austral temperate rain forest bryophytes 1. Nova Hedwigia 67: 289– 302. Mu˜noz J, Felicisimo AM, Cabezas F, Burgaz AR, Mart´ınez I. 2004. Wind as a long-distance dispersal vehicle in the Southern hemisphere. Science 304: 1144–1147. Natcheva R, Cronberg N. 2004. What do we know about hybridization among bryophytes in nature. Canadian Journal of Botany 82: 1687–1704. Natcheva R, Cronberg N. 2007. Recombination and introgression of nuclear and chloroplast genomes between the peat mosses, Sphagnum capillifolium and S. quinquefarium. Molecular Ecology 16: 811– 818. Newton AE, Wikstr¨om N, Bell N, Forrest LL, Ignatov MS. 2007. Dating the diversification of the pleurocarpous mosses. In: Newton AE, Tangney R eds. Pleurocarpous mosses. Systematics and Evolution. The Systematics Association Special Volume Series 71. Boca Raton: CRC Press. 337– 366. Odrzykoski IJ, Szweykowski J. 1991. Genetic differentiation without concordant morphological divergence in the thallose liverwort, Conocephalum conicum. Plant Systematics and Evolution 178: 135–151. Oostendorp C. 1987. The bryophytes of the Paleozoic and Mesozoic. Bryophytorum Bibliotheca 34: 5–112, pl. I–XLIX. Pacak A, Szweykowska-Kuli´nska Z. 2000. Molecular data concerning alloploid character and the origin of the chloroplast and mitochondrial genomes in liverwort species Pellia borealis. Journal of Plant Biotechnology 2: 101–108. Paton J. 1965. Lophocolea semiteres (Lehm.) Mitt. and Telaranea murphyae sp. nov. established on Tresco. Transactions of the British Bryological Society 5: 232–236. Pfeiffer T, Schaumann F, H¨assel de Men´endez GG, Frey W. 2004. Inter- and intrapsecific relationships in the Gondwanan liverwort genus Hymenophyton (Hymenophytaceae, Hepaticophytina). Studies in austral temperate rainforest bryophytes 23. Australian Systematic Botany 17: 407–421. Qiu YL. 2008. Phylogeny and evolution of charophytic algae and land plants. Journal of Systematics and Evolution 46: 287–306. Renner SS. 2005. Relaxed molecular clocks for dating historical plant dispersal events. Trends in Plant Science 10: 550– 558. Renzaglia KS, Schuette S, Duff RJ, Ligrone R, Shaw AJ, Mishler BD, Duckett JG. 2007. Bryophyte phylogeny: Advancing the molecular and morphological frontiers. The Bryologist 110: 179–213. Rycroft DS, Groth H, Heinrichs J. 2004. Reinstatement of Plagiochila maderensis (Jungermanniopsida: Plagiochilaceae) based on chemical evidence and nrDNA ITS sequences. Journal of Bryology 26: 37–45.  C

2009 Institute of Botany, Chinese Academy of Sciences

HEINRICHS et al.: Phylogenetic biogeography of bryophytes Sanmartin I, Ronquist F. 2004. Southern hemisphere biogeography inferred by event-based models: Plant versus animal patterns. Systematic Biology 53: 216–243. Schaumann F, Frey W, Pfeiffer T, Stech M. 2005. Molecular circumscription, intrafamilial relationships and biogeography of the Gondwanan liverwort family Pallaviciniaceae (Hepaticophytina, Bryophyta) – studies in austral temperate rain forest bryophytes 27. Plant Systematics and Evolution 252: 27–48. Schofield WB. 1985. Introduction to bryology. New York: MacMillan. Schofield WB. 1992. Bryophyte distribution patterns. In: Bates JW, Farmer AM eds. Bryophytes and lichens in a changing environment. Oxford: Clarendon. 103–130. Schuster RM. 1979. On the persistence and dispersal of transantarctic Hepaticae. Canadian Journal of Botany 57: 2179–2225. Shaw, AJ. 2000. Molecular phylogeography and cryptic speciation in the mosses, Mielichhoferia elongata and M. mielichhoferiana (Bryaceae). Molecular Ecology 9: 595–608. Shaw AJ. 2001. Biogeographic patterns and cryptic speciation in bryophytes. Journal of Biogeography 28: 253–261. Shaw AJ, Allen B. 2000. Phylogenetic relationships, morphological incongruence, and geographic speciation in the Fontinalaceae (Bryophyta). Molecular Phylogenetics and Evolution 16: 225–237. Shaw AJ, Boles S, Shaw B. 2008. A phylogenetic delimitation of the Sphagnum subsecundum complex (Sphagnaceae, Bryophyta). American Journal of Botany 95: 731–744. Shaw AJ, Goffinet B. 2000. Molecular evidence of reticulate evolution in the peatmosses (Sphagnum), including S. ehyalinum, sp. nov. The Bryologist 103: 357–374. Shaw AJ, Werner O, Ros RM. 2003. Intercontinental mediterranean disjunct mosses: morphological and molecular patterns. American Journal of Botany 90: 540–550. Skotnicki ML, Mackenzie AM, Ninham JA, Selkirk PM. 2004. High levels of genetic variability in the moss Ceratodon purpureus from continental Antarctica, subantarctic Herad and Macquarie Islands, and Australasia. Polar Biology 27: 687–698. Skotnicki ML, Selkirk PM, Broady P, Adam KD, Ninham JA. 2001. Dispersal of the moss Campylopus pyriformis on geothermal ground near the summits of Mount Erebus and Mount Melbourne, Victoria Land, Antarctica. Antarctic Science 13: 280–285. Stech M, Sim-Sim M, Frahm JP. 2007. Campylopus on Madeira island – molecular relationships and biogeographic affinities. Nova Hedwigia, suppl. 131: 91–100. Stech M, Wagner D. 2005. Molecular relationships, biogeography, and evolution of Gondwanan Campylopus species (Dicranaceae, Bryopsida). Taxon 54: 377–382. Stephani F. 1898–1925. Species Hepaticarum I–VI. Geneva: Georg & Cie or Universit´e de Gen`eve. Stotler RE, Crandall-Stotler B. 1974. A monograph of the genus Bryopteris. Bryophytorum Bibiliotheca 3: 1–159, figs. 1– 219. Sz¨ov´enyi P, Hock Z, Urmi E, Schneller JJ. 2006. Contrasting phylogeographic patterns in Sphagnum fimbriatum and Sphagnum squarrosum (Bryophyta, Sphagnopsida) in Europe. New Phytologist 172: 784–794.  C

2009 Institute of Botany, Chinese Academy of Sciences

507

Szweykowski J, Buczkowska K, Odrzykoski IJ. 2005. Conocephalum salebrosum (Marchantiopsida, Conocephalaceae)—a new holarctic liverwort species. Plant Systematics and Evolution 253: 133–158. Tan BC, P´ocs T. 2000. Bryogeography and conservation of bryophytes. In: Shaw AJ, Goffinet B eds. Bryophyte biology. Cambridge: Cambridge University Press. 403–448. Vanderpoorten A. 2004. A simple taxonomic treatment for a complicated evolutionary story: the genus Hygroamblystegium (Hypnales, Amblystegiaceae). Monographs in Systematic Botany from the Missouri Botanical Garden 98: 321– 327. Vanderpoorten A, Devos N, Goffinet B, Hardy OJ, Shaw AJ. 2008. The barriers to oceanic island radiation in bryophytes: insights from the phylogeography of the moss Grimmia montana. Journal of Biogeography 35: 654–663. Vanderpoorten A, Goffinet B. 2006. Mapping uncertainty and phylogenetic uncertainty in ancestral character state reconstruction: an example in the moss genus Brachytheciastrum. Systematic Biology 55: 957–971. Vanderpoorten A, Long DG. 2006. Budding speciation and neotropical origin of the Azorean endemic liverwort, Leptoscyphus azoricus. Molecular Phylogenetics and Evolution 40: 73–83. Vanderpoorten A, Rumsey FJ, Carine MA. 2007. Does Macaronesia exist? Conflicting signal in the bryophyte and pteridophyte floras. American Journal of Botany 94: 625–639. Vanderpoorten A, Shaw AJ, Cox CJ. 2004. Evolution of multiple paralogous adenosine kinase genes in the moss genus Hygroamblystegium: phylogenetic implications. Molecular Phylogenetics and Evolution 31: 505–516. Van Der Velde M, Bijlsma R. 2003. Phylogeography of five Polytrichum species within Europe. Biological Journal of the Linnean Society 78: 203–213. Van Der Velde M, Bijlsma R. 2004. Hybridization and asymmetric reproductive isolation between the closely related bryophyte taxa Polytrichum commune and P. uliginosum. Molecular Ecology 13: 1447–1454. van Zanten BO. 1978. Experimental studies on trans-oceanic long-range dispersal of moss spores in the Southern Hemisphere. Journal of the Hattori Botanical Laboratory 44: 455– 482. van Zanten BO, Gradstein SR. 1988. Experimental dispersal geography of Neotropical liverworts. Beihefte zur Nova Hedwigia 90: 41–94. Vilnet AA, Konstantinova NA, Troitsky AV. 2008. Phylogeny and systematics of the genus Lophozia s. str. (Dumort.) Dumort. (Hepaticae) and related taxa from nuclear ITS1–2 and chloroplast trnL-F sequences. Molecular Phylogenetics and Evolution 47: 403–418. A, Chudzi´nska E, Buczkowska K. Wachowiak W, Baczkiewicz  2007. Cryptic speciation in liverworts – a case study in the Aneura pinguis complex. Botanical Journal of the Linnean Society 155: 273–282. Wall DP. 2005. Origin and rapid diversification of a tropical moss. Evolution 59: 1413–1424. Warnstorf C. 1911. Sphagnales-Sphagnaceae (Sphagnologia Universalis). Das Pflanzenreich, Vol. 5. Leipzig: Engelmann. Wegener A. 1915. Die Entstehung der Kontinente und Ozeane. Braunschweig: Viehweg.

508

Journal of Systematics and Evolution

Vol. 47

No. 5

Werner O, Guerra J. 2004. Molecular phylogeography of the moss Tortula muralis Hedw. (Pottiaceae) based on chloroplast rps4 gene sequence data. Plant Biology 6: 147– 157. Werner O, Ros RM, Grundmann M. 2005. Molecular phylogeny of Trichostomoideae (Pottiaceae, Bryophyta) based on nrITS sequence data. Taxon 54: 361–368.

2009 Werner O, Ros RM, Guerra J, Shaw AJ. 2003. Molecular data confirm the presence of Anacolia menziesii (Bartramiaceae, Musci) in Southern Europe and its separation from Anacolea webbii. Systematic Botany 28: 483–489 Wilson R, Heinrichs J, Hentschel J, Gradstein SR, Schneider H. 2007. Steady diversification of derived liverworts under Tertiary climatic fluctuations. Biology Letters 3: 566–569.

 C

View publication stats

2009 Institute of Botany, Chinese Academy of Sciences