1-s2.0-s1055790304000326-main

Upload: linubinoi

Post on 03-Apr-2018

213 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/28/2019 1-s2.0-S1055790304000326-main

    1/18

    Distribution of introns in the mitochondrial gene nad1 in landplants: phylogenetic and molecular evolutionary implications

    Olena Dombrovska and Yin-Long Qiu*

    Department of Ecology and Evolutionary Biology, University of Michigan, 830 North University Avenue, Ann Arbor, MI 48109-1048, USA

    Institute of Systematic Botany, University of Zurich, Zollikerstrasse 107, 8008 Zurich, Switzerland

    Received 22 July 2003; revised 17 December 2003

    Available online 27 February 2004

    Abstract

    Forty-six species of diverse land plants were investigated by sequencing for their intron content in the mitochondrial gene nad1. A

    total of seven introns, all belonging to group II, were found, and two were newly discovered in this study. All 13 liverworts examined

    contain no intron, the same condition as in green algae. Mosses and hornworts, however, share one intron by themselves and

    another one with vascular plants. These intron distribution patterns are consistent with the hypothesis that liverworts represent the

    basal-most land plants and that the two introns were gained in the common ancestor of mosseshornwortsvascular plants after

    liverworts had diverged. Hornworts also possess a unique intron of their own. A fourth intron was found only in Equisetum L.,

    Marattiaceae, Ophioglossum L., Osmunda L., Asplenium L., and Adiantum L., and was likely acquired in their common ancestor,

    which supports the monophyly of moniliformopses. Three introns that were previously characterized in angiosperms and a few

    pteridophytes are now all extended to lycopods, and were likely gained in the common ancestor of vascular plants. Phylogenetic

    analyses of the intron sequences recovered topologies mirroring those of the plants, suggesting that the introns have all been

    vertically inherited. All seven nad1 group II introns show broad phylogenetic distribution patterns, with the narrowest being in

    moniliformopses and hornworts, lineages that date back to at least the Devonian (345 million years ago) and Silurian (435 million

    years ago), respectively. Hence, these introns must have invaded the genes via ancient transpositional events during the early stage of

    land plant evolution. Potentially heavy RNA editing was observed in nad1 of Haplomitrium Dedecek, Takakia Hatt. & Inoue,

    hornworts, Isoetes L., Ophioglossum, and Asplenium. A new nomenclature is proposed for group II introns.

    2004 Elsevier Inc. All rights reserved.

    Keywords: Genomic structural characters; Group II introns; Intron nomenclature; Land plant phylogeny; Mitochondrial DNA;nad1; RNA editing

    1. Introduction

    Evolution of land plants (embryophytes) fundamen-

    tally altered the terrestrial ecosystem and set the stage

    for evolution of other land-dwelling life (Gensel and

    Edwards, 2001; Graham, 1993; Kenrick and Crane,

    1997a). A large number of paleobotanical, morpholog-

    ical, and molecular studies have been devoted to un-

    derstanding this important event in the history of life on

    earth. However, our knowledge on radiation of early

    land plants remains incomplete, and in particular several

    key questions regarding basal land plant phylogeny are

    still controversial. These include: what represents the

    first lineage of land plants, which bryophyte group is

    sister to vascular plants, and how are several extant

    pteridophyte lineages related to each other (Edwards

    et al., 1995; Hedderson et al., 1998; Kenrick and Crane,

    1997b; Lewis et al., 1997; Mishler et al., 1994; Nickrent

    et al., 2000; Nishiyama and Kato, 1999; Pryer et al.,

    2001; Qiu et al., 1998; Renzaglia et al., 2000; Samigullin

    et al., 2002; Taylor, 1995; Wellman et al., 2003)? Clearly,

    more data are needed to resolve these issues.

    The mitochondrial genome in land plants is a rich

    source of information to investigate plant phylogeny

    (Beckert et al., 1999, 2001; Qiu et al., 1998, 1999; Vange-

    row et al., 1999). Two sets of introns have been found in

    angiosperms (Kubo et al.,2000; Notsu et al.,2002; Unseld

    et al., 1997) and the liverwort Marchantia polymorpha

    (Oda et al., 1992), respectively. Most of the angiosperm

    * Corresponding author. Fax: 1-734-764-0544.

    E-mail address: [email protected] (Y.-L. Qiu).

    1055-7903/$ - see front matter 2004 Elsevier Inc. All rights reserved.

    doi:10.1016/j.ympev.2003.12.013

    Molecular Phylogenetics and Evolution 32 (2004) 246263

    MOLECULAR

    PHYLOGENETICS

    AND

    EVOLUTION

    www.elsevier.com/locate/ympev

    http://mail%20to:%[email protected]/http://mail%20to:%[email protected]/http://mail%20to:%[email protected]/
  • 7/28/2019 1-s2.0-S1055790304000326-main

    2/18

    introns are now known to extend their presence to gym-

    nosperms, pteridophytes, and even hornworts, mosses, or

    liverworts (Beckert et al., 1999, 2001; Gugerli et al., 2001;

    Hashimoto and Sato, 2001; Malek and Knoop, 1998;

    Malek et al., 1997; Pruchner et al., 2001, 2002; Qiu et al.,

    1998). Theintrondistributionpatternscan be used to infer

    phylogenetic relationships among basal lineages of landplants, given that conservative genomic structural

    changes have proven to be informative markers in re-

    solving difficult phylogenetic issues (Manhart and Pal-

    mer, 1990; Raubeson and Jansen, 1992).

    The plant mitochondrial gene nad1 contains four

    group II introns in angiosperms (Chapdelaine and Bo-

    nen, 1991; Conklin et al., 1991; Kubo et al., 2000; Notsu

    et al., 2002; Unseld et al., 1997; Wissinger et al., 1991),

    but none in Marchantia (Oda et al., 1992) nor in several

    green algae that have been sequenced for their mito-

    chondrial genomes: Prototheca wickerhamii(Wolff et al.,

    1994), Chlorogonium elongatum (Kroymann and Zet-

    sche, 1998), Nephroselmis olivacea and Pedinomonas

    minor (Turmel et al., 1999), Scenedesmus obliquus (Kuck

    et al., 2000; Nedelcu et al., 2000), Mesostigma viride

    (Turmel et al., 2002a), Chaetosphaeridium globosum

    (Turmel et al., 2002b), Chara vulgaris (Turmel et al.,

    2003), and Chlamydomonas reinhardtii [M.W. Gray,

    unpublished data (GenBank Accession No. U03843);

    one group I intron has been reported in Chlamydomonas

    eugametos nad1 (Denovan-Wright et al., 1998)]. The first

    three angiosperm introns have been shown to be present

    in a few basal pteridophytes (Gugerli et al., 2001; Malek

    and Knoop, 1998), and the last one is present

    throughout land plants except liverworts (Qiu et al.,1998). An additional group II intron, located upstream

    of the first angiosperm intron, has been found in the

    moss Ceratodon purpureus (Malek and Knoop, 1998).

    However, the information on the full intron content in

    nad1 across land plants, especially in bryophytes, is still

    lacking. It would be desirable to obtain this information

    to trace origins of these introns, to test whether they

    follow vertical inheritance or undergo horizontal trans-

    fer, and to examine their frequency of secondary losses.

    Understanding evolution of these introns will not only

    ensure their proper use as phylogenetic markers, but

    also will help us to gain insight into their role in shaping

    evolution of the mitochondrial genome in land plants.

    In this study, we carry out a broad survey of intron

    distribution in nad1 in land plants by sequencing the

    nearly entire length of the gene, with particularly dense

    taxon sampling in liverworts, mosses, and hornworts.

    We aim to uncover phylogenetically informative intron

    distribution patterns to resolve relationships among

    basal land plant lineages. We also seek to understand

    the tempo and mode of intron evolution in mitochon-

    drial genomes of early land plants, as a relatively large

    body of data is now available from both mitochondrial

    genome sequencing studies and evolutionary surveys.

    2. Materials and methods

    2.1. DNA extraction, gene amplification and sequencing,

    intron secondary structure modeling

    Materials for 46 representatives of all major non-

    flowering land plant lineages, including 12 liverworts, 19mosses, four hornworts, two lycopods, eight monili-

    formopses (sensu Kenrick and Crane, 1997b, including

    Equisetum, Psilotaceae, and ferns), and one gymno-

    sperm were collected in the field or greenhouses, care-

    fully identified, and properly vouchered (Table 1). All

    bryophytes were cleaned and checked under a dissecting

    scope to avoid contamination. The voucher information

    and GenBank accession numbers are given in Table 1.

    The green algae (see Figs. 1 and 2) were used as out-

    groups for comparing intron content and timing the

    evolutionary origins of the introns in nad1. Total cellu-

    lar DNAs were extracted using the cetyltrimethylam-

    monium bromide (CTAB) method (Doyle and Doyle,

    1987), with 2% polyvinyl pyrrolidone, 0.1% diet-

    hyldithiocarbamic acid (sodium salt), and 0.1% ascorbic

    acid added, or alternatively using the Plant DNAeasy

    kit (Qiagen) according to the manufacturers protocol.

    Multiple sets of primers were tested to circumvent any

    potential problems caused by direct and reverse RNA-

    editing at priming sites, and they are listed in Table 2.

    Where possible, a single amplicon was recovered for the

    gene. In some species, where the size of nad1 (including

    both exons and introns) exceeded the capacity of con-

    ventional PCR methods, the gene was amplified and se-

    quenced in several overlapping fragments and the nad1gene contig was assembled subsequently. PCR amplifi-

    cation reactions contained approximately 10 ng template

    DNA, 10 mM Tris/HCl (pH 8.85), 25 mM KCl, 5 mM

    (NH4)2SO4, 2 mM MgSO4, 200 lM each dNTP, 0.20 lg

    each primer, and 2.0 U Taq DNA polymerase (Qiagen) in

    a final volume of 50 ll. The reactions were carried out

    using 30 cycles, each consisting of 30 s denaturation at

    94 C, 30 s annealing at 52 C, and 14 min extension at

    72 C. The first cycle was preceded by an initial denatur-

    ation step (4min, 94 C), and the last one was followed by

    a 12 min extension step at 72 C. Alternatively, for am-

    plifying fragments larger than 3.5 kb, the Expand Long

    Template PCR-System (BoehringerMannheim) was

    used, following the protocol supplied by the manufac-

    turer. PCR fragments, depending on their size, were

    cloned with TA-TOPO or XL-TOPO cloning kits (Invit-

    rogen). Positive clones were sequenced in both directions

    using ABI Prism BigDye Terminator Cycle Sequencing

    Ready Reaction Kits on an ABI377 sequencer (both from

    Applied Biosystems). M13 universal primers and those

    designed to match internal sequences of cloned fragments

    (Table 2) were used during sequencing. Sequence

    processing and contig assembly were done using the

    Sequencher program (Gene Codes Corporation).

    O. Dombrovska, Y.-L. Qiu / Molecular Phylogenetics and Evolution 32 (2004) 246263 247

  • 7/28/2019 1-s2.0-S1055790304000326-main

    3/18

    Table 1

    Plant material used in this study, intron presence (+) or absence ()) in the mitochondrial gene nad1, and GenBank accession number of the sequence

    Species Vouchera Accession No.

    nad1, one

    accession #

    nad1i258 nad1i287 nad1i 348 nad1i394 nad1i

    GymnospermsCycas revoluta Thunb. Qiu94051 + AY

    Moniliformopses

    Asplenium nidus L. Qiu94081 AY354954 + ) ) + )

    Adiantum sp. Qiu95118 AY354953 + ) ) + )

    Osmunda regalis L. Qiu96171 + AY

    Ophioglossum lusitanicum L. Qiu98097 +AY354943 ) AY354943 ) AY354943 ) AY354943 + AY

    AY35

    Marattia attenuata Lab. Qiu96169 +AY354948 ) AY354948 ) AY354948 +AY354948 + AY

    Angiopteris evecta (G.Forst.) Hoffm. Qiu96170 +AY354947 ) AY354947 ) AY354947 +AY354947 + AY

    Equisetum arvense L. Qiu94003 AY354940 + ) ) + )

    Psilotum nudum (L.) P.Beauv. Qiu94002

    Lycopods

    Isoetes sp. Qiu96272 AY354939 ) ) ) + )

    Huperzia lucidula (Michx.) Trevis. Qiu94173 AY354938 ) ) ) + +

    Hornworts

    Notothylas breutelii (Gottsche) Gottsche Qiu97078 AY354981 ) + + ) )

    Phaeoceros carolinianus (Michx.) Prosk. Qiu97002 AY354980 ) + + ) )

    Megaceros tosanus Steph. Qiu97001 AY354979 + ) )

    Anthoceros agrestis Paton Qiu99112 AY354978 ) + + ) )

    Mosses

    Hypnum cupressiforme Hedw. Qiu99079 AY354976 ) + ) ) )

    Brachythecium rutabulum (Hedw.) BSG. Qiu94071 AY354975 ) + ) ) )

    Thuidium recognitum (Hedw.) Lindb. Qiu94174 AY354974 ) + ) ) )

    Anomodon viticulosus (Hedw.) Hook. & Tayl. Qiu99066 AY354973 ) + ) ) )

    Hedwigia ciliata (He dw.) P.Beauv. Qiu99056 AY354972)

    +) ) )Rhizogonium paramattense (Mull. Hal.) Qiu98069 AY354971 ) + ) ) )

    Reichardt

    Bartramia halleriana Hedw. Qiu99062 AY354970 ) + ) ) )

    Leucobryum albidum (Brid.) Lindb. Qiu94075 AY354969 ) + ) ) )

    Dicranum scoparium Hedw. Qiu94073 AY354968 ) + ) ) )

    Mnium sp. Qiu94097 AY354935 ) + ) ) )

    Gymnostomum rucurvirostrum Hedw. Qiu94072 AY354936 ) + ) ) )

    Fissidens dubius P.Beauv. Qiu99069 AY354967 ) + ) ) )

    Tetraphis pellucida Hedw. Qiu00001 AY354956 ) + ) ) )

    Polytrichum juniperum Hedw. Qiu94176 AY354966 ) + ) ) )

  • 7/28/2019 1-s2.0-S1055790304000326-main

    4/18

    Sequence alignment was performed using ClustalX

    program (Thompson et al., 1997). Secondary structure

    models of group II introns were inferred according to

    the consensus model of Michel et al. (1989) on Foldalign

    server (Mathews et al., 1999; Zuker et al., 1999). Manual

    adjustments were made after computer-assisted folding

    analysis.

    2.2. Group II intron nomenclature

    Because there is a great deal of confusion on intron

    names in literature due to lack of a standard nomen-

    clature, we develop a system here for group II introns

    that is similar to what has been proposed for group I

    introns in ribosomal DNA (Johansen and Haugen,

    2001), which has been adopted for archael introns

    (Nakayama et al., 2003). The intron name in our system

    will be composed of: (1) gene name, (2) the letter i for

    intron, and (3) insertion site in the orthologous gene of

    M. polymorpha, and it will always be italicized. For

    example, nad1i728 refers to an intron located in the gene

    nad1 of a given species at the M. polymorpha nucleotide

    position 728 (an alignment of the exon sequence of the

    species and that of M. polymorpha will be necessary to

    identify the intron insertion site).

    Compared to the group I intron nomenclature, our

    system has omitted use of the species name acronym and

    the genome (chloroplast, mitochondrial, or nuclear)

    designation letter as part of the formal intron name. In-

    stead, we recommend that all that information as well as

    other information such as trans-splicing (Malek and

    Knoop, 1998), presence of ORF (Zimmerly et al., 2001),status within a twintron (Hallick et al., 1993), and intron

    phase (Sharp, 1981) be amended to the formal intron

    name in particular studies when the need of incorporating

    such information arises. Thus, this system has advantages

    of being flexible and convenient in communication.

    Marchantia polymorpha is chosen as a reference spe-

    cies because its mitochondrial and chloroplast genomes

    have been sequenced (Oda et al., 1992; Ohyama et al.,

    1986) and they contain most of the genes that have been

    found so far in organellar genomes of land plants and

    other photosynthetic eukaryotes. In cases where

    M. polymorpha lacks an ortholog, or orthologous part

    due to deletion, of the gene in which an intron has been

    found in one or several species, the exon sequence of the

    species in which the intron was first discovered will be

    used as the reference. The insertion site in this species

    will be used with a superscript asterisk followed by

    the species name in a parenthesis. For example,

    ycf66i106(Chaetosphaeridium globosum) refers to an

    intron present at position 106 in the chloroplast gene

    ycf66(which is absent in M. polymorpha) and it was first

    discovered in C. globosum, which has been used as the

    reference species. For introns located in the 50 and 30 un-

    translated regions, a ) and + sign will be usedAtrichumangustatum(Brid.)BSG.

    Qiu94074

    AY354934

    )

    +

    )

    )

    )

    )

    +

    Diphysciumfoliosum(Hedw.)Mohr

    Qiu99083

    AY354933

    )

    +

    )

    )

    )

    )

    +

    SphagnumrecurvumP.Beauv.

    Qiu94175

    AY354932

    )

    +

    )

    )

    )

    )

    +

    Takakiaceratophylla(Mitt.)Grolle

    Qiu97125

    AY354937

    )

    +

    )

    )

    )

    )

    +

    TakakialepidozioidesHatt.&Inoue

    Qiu97126

    AY354977

    )

    +

    )

    )

    )

    )

    +

    Liverworts

    Frullaniadilatata(L.)Dumort.

    Qiu99058

    AY354962

    )

    )

    )

    )

    )

    )

    )

    Radulacomplanata(L.)Dum.

    Qiu99075

    AY354961

    )

    )

    )

    )

    )

    )

    )

    Plagiochilaasplenioides(L.)Dum.

    Qiu99064

    AY354960

    )

    )

    )

    )

    )

    )

    )

    Lophocoleaheterophylla(Schrad.)Dum

    .

    Qiu98007

    AY354959

    )

    )

    )

    )

    )

    )

    )

    Marsupellaemarginata(Ehrh.)Dumor

    t.

    Qiu99077

    AY354958

    )

    )

    )

    )

    )

    )

    )

    Pelliasp.

    Qiu95001

    AY354964

    )

    )

    )

    )

    )

    )

    )

    MetzgeriatemperataKuwahara

    Qiu98008

    AY354931

    )

    )

    )

    )

    )

    )

    )

    MetzgeriaconjugataLindb.

    Qiu99080

    AY354965

    )

    )

    )

    )

    )

    )

    )

    Conocephalumsp.

    Qiu94096

    AY354957

    )

    )

    )

    )

    )

    )

    )

    Ricciocarposnatans(L.)Corda

    Qiu97153

    AY354929

    )

    )

    )

    )

    )

    )

    )

    Blasiapusilla(Micheli)L.

    Qiu99108

    AY354963

    )

    )

    )

    )

    )

    )

    )

    Haplomitriummnioides(Lindb.)Schust.

    Qiu97127

    AY354930

    )

    )

    )

    )

    )

    )

    )

    a

    ThevoucherswithnumbersbetweenQiu94001and97999aredepositedatIndianaUniversityHerbarium(IND,Bloomington

    ,IN,USA),andthosewithnumbersbetwee

    nQiu98001and00999

    atUniversityofZurichHerbarium(Z,Zurich,Switzerland).

    O. Dombrovska, Y.-L. Qiu / Molecular Phylogenetics and Evolution 32 (2004) 246263 249

  • 7/28/2019 1-s2.0-S1055790304000326-main

    5/18

    Fig. 1. The exon phylogeny of the mitochondrial nad1 from 64 species of land plants and green algae (Nephroselmis olivacea (AF110138) was used as

    the outgroup). Shown here is a maximum likelihood tree ()Ln likelihood 14993.51076). The branch length difference is contributed by both the

    divergence level and length of the sequences analyzed. Asterisks mark the nodes collapsed in the strict consensus tree of 25,941 most parsimonious

    trees (length: 3106 steps, CI (consistency index): 0.4858, and RI (retention index): 0.6863). The only difference between the likelihood and parsi-

    monious trees concerns placement ofOphioglossum, Psilotum, and Osmunda. Bootstrap values higher than 50% are shown above the branches. The

    taxa with sequences from the GenBank are underlined:Prototheca (NC_001613), Mesostigma (AF353999), Pedinomonas (NC_000892), Scenedesmus

    (AF204057), Chlorogonium (Y13644), Chlamydomonas (U03843), Chaetosphaeridium (NC_004118), Marchantia (NC_001660), Ginkgo (AF227466,

    AF227470), Gnetum (AF227467, AF227471), Pinus (AF160261), Arabidopsis (Y08501, Y08502), Petunia (X60400-X60403), Oenothera (AH003143),

    Beta (NC_002511), Oryza (AB076665), and Triticum (X57965-X57968).

    250 O. Dombrovska, Y.-L. Qiu / Molecular Phylogenetics and Evolution 32 (2004) 246263

  • 7/28/2019 1-s2.0-S1055790304000326-main

    6/18

    Fig. 2. Intron distribution in the mitochondrial nad1 of land plants and green algae. The species are arranged into phylogenetic groups. The coding

    sequences (gray boxes) are drawn to an approximate scale, with all intron positions marked. Rectangular boxes denote complete coding sequences;

    triangle-edged boxes indicate partial coding sequences.Trans-spliced introns are represented by shifted exon boxes. Species with data from literature

    are underlined. The numbers in parentheses after intron names indicate intron phases. *For nad1i728, (1) all mosses except Takakia (both species)

    and Sphagnum have lost the matR, (2) splicing status in Psilotum and Ophioglossum is unknown due to the partial sequences obtained, (3) presence of

    the cis-spliced intron in the four gymnosperms (except in Pinus where splicing status is unknown) has been demonstrated by Southern hybridizations

    (Qiu and Palmer, in press) and matR sequencing (Qiu et al., 1999), and (4) trans-splicing in Beta, Petunia, Triticum, and Oryza evolved independently

    (Qiu and Palmer, in press).

    O. Dombrovska, Y.-L. Qiu / Molecular Phylogenetics and Evolution 32 (2004) 246263 251

  • 7/28/2019 1-s2.0-S1055790304000326-main

    7/18

    before the insertion site, respectively. Clearly, future

    discoveries will present situations for which amendment

    to this nomenclature will be necessary.

    2.3. Phylogenetic analyses

    Because bryophytes often grow as a mixture of several

    species, the chance of cross-contamination could still exist

    even though we made an effort to check tissue purity

    under a dissecting scope. To confirm that the sequences

    obtained wereindeedfrom the specieswe targeted andnot

    from biological contaminants, we performed phyloge-

    netic analyses on the exon sequences ofnad1, using both

    the parsimony and maximum likelihood methods as im-

    plemented in PAUP* 4.0b10 (Swofford, 2003). In the

    parsimony analysis, we weighted all characters and mu-

    tations equally. For the maximum likelihood analysis, we

    used the HYK85 model with the following settings: two

    Table 2

    Primers used for mitochondrial nad1 gene amplification and sequencing

    Forward primers Reverse primers

    Location Sequence Location Sequence

    Mp nad1-1 50-ATGAGAATTTATCTAATTGG-30

    Ol nad1i258-1327 50-TAGGCCACTCTGAATTCC-30

    Mp nad1-82 CTAGCWGAACGWAAAGTCATGGC Ol nad1i258-2118 CTTTCAGTCGAGTGTCTGGAA

    Mp nad1-153 GTTACAACCTTTDGCWGAYGG Mp nad1-292 GACAATACCATACCATAATCMp nad1-191 AAGARCCTATTTTAYYRAGTAGTGCT Hl nad1i394-1007 GTCGGTTATTCCATCTCC

    Ol nad1i258-763 CARGTGCGCAAGGACCTKWMAC Hl nad1i394-1476 AGCRARGTGCAGARAACRCG

    Ol nad1i258-1213 TTCGGAGAGGTCCTATCG Hl nad1i394-1486 CGATAGAGGAAAGCGAGGTGC

    Mp nad1-264 TTTGGTTGCTTGGGCGGTTATC Mp nad1-395 TCCTGAAAAAGCATATTTGGRAT

    Aa nad1i287-334 ACTATGCGGACATCTTAC Mp nad1-407 ATCGTAATGCTCCYARAAADG

    Mp nad1-292 GATTATGGTATGGTATTGTC At-nad1i477-12 ACCTTCGCGGATCCAAACGCGCTC

    Hl nad1i394-369 TGTAYGCTGTCTTGCTTCCT At-nad1i477-21 TGCTGCGGTTCCCTTCGGTCCGTTA

    Hl nad1i394-705 CTCRAATCYCACGAATCC At nad1i477-479 CTTCCAACGTCAACCTGCG

    Hl nad1i394-1060 MGRTGGAAACAAGCAGCG At nad1i477-720 GTRTTGAGTAGGCAGCGCC

    Hl nad1i394-1341 CTGGGTACTGAGCCCTTTGC Mp nad1-623

    GCGACTRATTCMGCTTCTGCTTCTGGTA Hl nad1i394-1350 GGAACCWTTGTAMCYTAGAKAC

    Mp nad1-625 TTCTGCYTCDGCYTCTGG Mp nad1-404 ATGCTTTTYTRGGAGCATTACGAT

    Hl nad1i669-1185 ATCTAGGTATATGTTGGTAA Mp nad1-408 TTTTTYRGGAGSWTTACGDTCTG

    Hl nad1i669-2391 GTGTGAAGGCAGGACCMS At nad1i477-379 CCTGATCTGGATAACGATAG

    Hl nad1i669-2628 CAGCCGCAATAACAGCAC At nad1i477-481 CMAATGGACTCTACRAGGACCMp nad1-671 AGAGCAAAYCCCATDGAVG At nad1i477-586 GGGCTGTAGGTGATAAYGC

    Mp nad1-703 CTCATTAAGATCATATTAGCATACTC Mp nad1-623

    TACCAGAAGCAGAAGCKGAATYAGTCGC At nad1i728-168 CTTAGATARGTCGTCCGATGGGT

    Mp nad1-636 AGCTGAATTMGTHGCDGG At nad1i728-126 GARGGGTAGCCTGGTGTGTARGC

    Hl nad1i669-538 RAGTATRATGACCTGGTCG At nad1i728-450 TCTCCCAGRACCARAATGATTATC

    Hl nad1i669-984 CACRATCMATATACACACC0 At nad1i728-1712 GGTTTAGACTKGATTYGCTCATA

    Hl nad1i669-1435 TACGCCGTTACAATACGCAGA At nad1i728-2070 TCTTTACTAAGTGTCCCTGGCGGT

    Hl nad1i669-1818 GGAGCTGGTCTCAGATGATG At nad1i728-2087

    GCTATGCRAGATYTCTCTTGATCTTTGC Hl nad1i669-2290 CATTMTTGMTTGGCGAAG

    At nad1i728-2229 TCGTCTGCGTCCTCTTCGTTC Mp nad1-670 TCYTCWATGGGRTTTGCTYTKTT

    At nad1i728-2490 GTRAATATAGCRGACCAGCGRAT Mp nad1-709 GCYAATATGATCTTAATGAG

    At nad1i728-2517 GARGATTTGTGCTTGTKGGC Mp nad1-714 TATGATCTTAATGAGGTGCGGAG

    Mp nad1-762 TAGGAATATCTAGGATGGGCAGC Mp nad1-725

    TGAGGTGCGGAGCYTTGCATCYGACATT Mp nad1-789 CGAGCCCGGAATCACCTGG

    At nad1i728-432 GTAAACTCATTRRACAGG Mp nad1-832 CCCATATATATACMAACAAAAGGGAAt nad1i728-745 CGGTACYCGAATCYATTTACGA Mp nad1-858 MGATATCGWGGAAATGCTGCRC

    At nad1i728-766 AKCCCGAGTTTCYAGACACAT Mp nad1-966 TTAAGRAAGCCATTCAAAGGCT

    At nad1i728-1028 CAGTGTACTACTATCACCCCTAC

    At nad1i728-1249 GCYTACACACCAGGCTACCCYTC

    At nad1i728-1852 TAGAGAAGCGTCGGAGAGTA

    At nad1i728-1852 TRGARAAGCGTCGGAGAGTAAAGCACC

    At nad1i728-1859 GCGTCGGAGAGTAAAGCACCGTATC

    At nad1i728-1870 TAAAGCACCGTATCCATCTCACAG

    At nad1i728-2890 RGCCTCATAGTGCCTCCTGT

    At nad1i728-3128 CAAAAGACTGCTAGTGGCGTCC

    * Numbers correspond to the 50 and 30 positions of the forward and reverse primers in Marchantia polymorpha (Mp) mitochondrial nad1 gene

    (NC_001660, Oda et al., 1992), Arabidopsis thaliana (At) introns nad1i477** and nad1i728 (NC_001284, Unseld et al., 1997), and Anthoceros agrestis

    (Aa) nad1i287, Huperzia lucidula (Hl) nad1i394 and nad1i669, and Ophioglossum lucitanicum (Ol) nad1i258 (all from this study). Amplification

    primers are underlined. Two 50-end nucleotides (TT) from Arabidopsis thaliana nad1i477 (NC_001284) actually belong to the upstream exon. Here we follow the

    corrected exonintron boundary, starting nucleotide counting in the intron from GTGCG. . .

    252 O. Dombrovska, Y.-L. Qiu / Molecular Phylogenetics and Evolution 32 (2004) 246263

  • 7/28/2019 1-s2.0-S1055790304000326-main

    8/18

    substitution types (HYK85 variant), a transition/trans-

    version ratio of 2 (j 4.2901951), empirical base fre-

    quencies (A 0:23129; C 0:17128; G 0:21140; T

    0:38603), no invariable sites, equal rates for all sites,

    number of distinct data patterns under this model 932,

    molecular clock not enforced, starting branch lengths

    obtained using RogersSwofford approximation method,branch-length optimizationone-dimensional Newton

    Raphson with pass limit 20, delta 1e ) 06, and )ln L

    (unconstrained)unavailable due to missing data and/or

    ambiguities. A heuristic search was conducted to find the

    trees of the shortest length or the highest likelihood using

    simple (for the parsimony analysis) or as is (for the

    maximum likelihood analysis) taxon-addition replicates,

    one tree held at each step during stepwise addition, TBR

    branch swapping, steepest descent option in effect, Mul-

    Trees option in effect and no upper limit of MaxTrees.

    A bootstrap analysis was conducted using the parsimony

    method to test robustness of the topology, using 100 re-

    sampling replicates and the same tree search procedure as

    described above except with SPR branch swapping.

    We also carried out phylogenetic analyses on intron

    sequences, to evaluate how intron phylogenies reflect the

    plant phylogeny and to determine whether an intron

    originated via a single acquisition event followed by

    vertical inheritance or through multiple independent

    gains caused by horizontal transfers. Because the par-

    simony and maximum likelihood analyses produced

    virtually identical results on the exon data (see Fig. 1

    legend), we conducted only parsimony analyses for the

    intron sequences. For two introns (nad1i287 and

    nad1i728) with data from a relatively large number ofspecies, the same heuristic tree search method and

    bootstrap analysis procedure as the exon sequence

    analyses were used, except with the TBR branch swap-

    ping option. For other five introns that had data from

    only a small number of species, we were able to use the

    branch-and-bound or exhaustive tree search methods to

    find the most parsimonious tree(s). The bootstrap

    analysis procedure was the same as in the exon analysis.

    3. Results

    Among the 46 species of diverse land plants we exam-

    ined (Table 1), amplification of almost entire nad1 coding

    region encompassing all known intron positions was

    possible for a majority of the species. However, only

    partial coding regions of varyinglengths covering some of

    the intron positions were obtained from Isoetes, Psilotum

    Sw., Angiopteris Hoffm., Marattia Sw., Ophioglossum,

    Osmunda, and Cycas L.. Our analyses of the exon se-

    quences produced a phylogeny (Fig. 1) that is in general

    agreement with the relationships of these plants as re-

    ported in published studies (Hedderson et al., 1998; Lewis

    et al., 1997; Newton et al., 2000; Pryer et al., 2001) and in

    our own unpublished analysis of six genes from over 180

    land plants. However, there were several anomalies in the

    exon phylogeny, e.g., Haplomitrium falling into the clade

    of leafy and simple thalloid liverworts, paraphyly of

    mosses, and Takakia being sister to hornworts. Several

    factors may have contributed to these results: short se-

    quences used for such a broad range of taxa, uneven RNAediting patterns (see below), and insufficient taxon sam-

    pling. Nonetheless, we think that theexon phylogeny does

    suggest that nad1 is a useful phylogenetic marker and that

    all of the sequences obtained are indeed from the species

    we targeted, i.e., not from biological contaminants.

    A total of seven introns were found, and all of them

    belong to group II (Table 1, Fig. 2). Five of them have

    been reported before (Chapdelaine and Bonen, 1991;

    Conklin et al., 1991; Kubo et al., 2000; Malek and

    Knoop, 1998; Notsu et al., 2002; Wissinger et al., 1991),

    and two are newly discovered in this study. They all

    have characteristic group II intron primary and sec-

    ondary structural features: the consensus sequence of

    GBGYG at the 50-end (the second nucleotide consensus

    needs to be expanded; the consensus sequence of AY at

    the 30-end is not being followed by many group II in-

    trons, including three of the seven introns reported here:

    nad1i348, nad1i669, and nad1i728); six double helical

    domains radiating from a central wheel; the bulging A

    residue at the 7th8th position from the 30-end in do-

    main VI, which is required for lariat formation and in-

    tron excision; and the conserved domain V with a

    GNRA terminal tetraloop (for six out of seven introns)

    (Lambowitz and Belfort, 1993; Qin and Pyle, 1998).

    For 12 liverworts, we obtained the full length codingregion ofnad1 using primers that covered start and stop

    codons (Metzgeria conjugata and M. temperata missed

    two codons at the 30-end), and found no introns at all

    (Table 1, Fig. 2). This is the same condition as reported in

    Marchantia (Oda et al., 1992). In other land plants, the

    introns showed lineage-specific distribution patterns as

    described below.

    3.1. nad1i258

    This is a novel intron we discovered in the present

    study; its secondary structure is shown in Fig. 3A. It is

    present in all the moniliformopses for which we could get

    a sequence of this region of the gene: Equisetum, Ophio-

    glossum, Angiopteris, Marattia, Adiantum, and Asple-

    nium, but is absent from green algae, bryophytes and

    other vascular plants (Table 1, Fig. 2). Its typical length is

    approximately 900 bp, with an exception in Ophioglos-

    sum, where the intron contains a remnant ORF, resulting

    in a length of 2598 bp. Our phylogenetic analyses show

    that the evolutionary historyof the intron parallels that of

    the plants (Fig. 4A; Pryer et al., 2001), suggesting that the

    intron was acquired in the common ancestor of monili-

    formopses and has been inherited vertically.

    O. Dombrovska, Y.-L. Qiu / Molecular Phylogenetics and Evolution 32 (2004) 246263 253

  • 7/28/2019 1-s2.0-S1055790304000326-main

    9/18

    Fig. 3. Inferred secondary structure of two newly discovered group II introns in the mitochondrialnad1. (A) nad1i258 from Marattia attenuata and

    (B) nad1i348 from Phaeoceros carolinianus.

    254 O. Dombrovska, Y.-L. Qiu / Molecular Phylogenetics and Evolution 32 (2004) 246263

  • 7/28/2019 1-s2.0-S1055790304000326-main

    10/18

    3.2. nad1i287

    This intron, being orthologous to the one discovered

    in Ceratodon purpureus by Malek and Knoop (1998),

    was found in all 19 mosses and three hornworts we in-

    vestigated (no PCR product was obtained from

    Megaceros D.H. Campbell), and has an average length

    of 900 bp. It is absent in green algae and other lineages

    of land plants (Table 1, Fig. 2). The intron phylogeny

    (Fig. 4B) is in agreement with that of the plants (Newton

    et al., 2000), indicating vertical transmission of the

    intron in the mosses and hornworts.

    Fig. 4. Phylogenies of the seven introns found in the mitochondrial nad1 of land plants. The numbers above the branches are bootstrap values.

    The taxa with sequences from the GenBank are underlined: Ceratodon (Y17810), Isoetes lac. (Y17812), Pinus (AF160261), Osmunda 1 (Y17815),

    Equisetum tel. (Y17811), and Notothylas (AF068932). (A) nad1i258, the single shortest tree found in an exhaustive tree search (length: 1033 steps,

    CI: 0.9555, and RI: 0.8357). (B) nad1i287, the strict consensus tree of 10 shortest trees found in a heuristic search (length: 694 steps, CI: 0.7277, and

    RI: 0.7970). (C) nad1i348, the single shortest tree found in an exhaustive tree search (length: 73 steps, CI: 0.9863, and RI: 0.8000). (D) nad1i394, thesingle shortest tree found in an exhaustive tree search (length: 1057 steps, CI: 0.9574, and RI: 0.9240). (E) nad1i477, the single shortest tree found

    in an exhaustive tree search (length: 2223 steps, CI: 0.9353, and RI: 0.6949). (F) nad1i669, the single shortest tree found in an exhaustive tree search

    (length: 1708 steps, CI: 0.9895, and RI: 0.9669). (G) nad1i728, the strict consensus tree of 12 shortest trees found in a heuristic search (length: 7556

    steps, CI: 0.6546, and RI: 0.5223).

    O. Dombrovska, Y.-L. Qiu / Molecular Phylogenetics and Evolution 32 (2004) 246263 255

  • 7/28/2019 1-s2.0-S1055790304000326-main

    11/18

    3.3. nad1i348

    This is another new intron we discovered, and its

    secondary structure is shown in Fig. 3B. It is present

    in all four hornworts we examined: Anthoceros [Mic-

    heli] L., Megaceros, Phaeoceros Proskauer and Not-

    othylas Sull., but absent in green algae and all otherland plants (Table 1, Fig. 2). The intron has a typical

    length of 600 bp. Our phylogenetic analyses indicate

    that it has been transmitted vertically following its

    acquisition in the common ancestor of hornworts

    (Fig. 4C).

    3.4. nad1i394

    This intron is orthologous to the cis-spliced nad1Ti1

    in Isoetes lacustris reported by Malek and Knoop

    (1998) and the trans-spliced first intron of nad1 in

    angiosperms (Chapdelaine and Bonen, 1991; Conklin

    et al., 1991; Wissinger et al., 1991). According to our

    survey, it is also present, in a cis-spliced form as in

    Isoetes lacustris, in Huperzia Bernh., Isoetes sp., Eq-

    uisetum arvense, Angiopteris, Marattia, Adiantum, and

    Asplenium, but absent entirely in bryophytes and green

    algae (Table 1, Fig. 2). Among the species investi-

    gated, the intron size ranges from 1400 to 2000 bp in

    most species, but in Isoetes it drops dramatically to

    520 bp. This sharp decrease of intron size in Isoetes

    has been observed earlier by Malek and Knoop (1998)

    for this intron as well as for two other introns in nad2

    and nad5. It appears now that the decrease of intron

    size is restricted to only Isoetes; the other lycopod weexamined, Huperzia, has a normal-sized intron. The

    absence of this intron in Ophioglossum (Fig. 2) and

    Equisetum telmateia (Malek and Knoop, 1998) can be

    explained by secondary losses. The intron phylogeny

    (Fig. 4D) essentially agrees with the plant phylogeny

    (Pryer et al., 2001), with the exception of Equisetum

    embedded in Marattiaceae. This topological anomaly

    may be caused by insufficient taxon sampling, rather

    than serving as evidence of intron horizontal transfer.

    Distribution of the intron in land plants (Fig. 2) is

    mostly consistent with the hypothesis of a single origin

    of the intron in the beginning of vascular plant evo-

    lution followed by independent losses. Our failure to

    amplify the portion of the gene covering the intron

    insertion site in Psilotum, Osmunda, and Cycas, if not

    due to technical reasons, could be an indication that

    this intron is present in a trans-spliced state in these

    plants, as multiple independent evolutions of trans-

    splicing have been detected in another nad1 intron,

    nad1i728 (Conklin et al., 1991; Kubo et al., 2000;

    Notsu et al., 2002; Qiu and Palmer, in press; Wis-

    singer et al., 1991). Further investigation in these

    plants would be desirable to find out the exact status

    of the intron.

    3.5. nad1i477

    This intron is orthologous to the second intron in

    nad1 of angiosperms (Chapdelaine and Bonen, 1991;

    Conklin et al., 1991; Wissinger et al., 1991). It has been

    reported to be present in Huperzia lucidula and a large

    number of seed plants by Gugerli et al. (2001) and Wonand Renner (2003) using either sequencing or Southern

    hybridization. In this survey, we found that this intron is

    present in Huperzia lucidula (the same DNA was used

    for both Gugerli et al., 2001, and this study), Angiop-

    teris, Marattia, Ophioglossum, Osmunda, and Cycas, but

    absent in green algae, bryophytes, and four vascular

    plants that were sequenced through the intron insertion

    site, Isoetes, Equisetum, Adiantum, and Asplenium (Table

    1, Fig. 2). Clearly, absence of the intron in the four

    vascular plants represents secondary losses, as has been

    observed earlier in some seed plants (Gugerli et al.,

    2001). The size of this intron shows a full range of

    variation from 896 bp in Arabidopsis to 2844 bp in

    Ophioglossum, unlike nad1i394 which has a bimodal size

    distribution. The intron phylogeny (Fig. 4E) agrees well

    with the plant phylogeny. Judging from the data shown

    here and in Gugerli et al. (2001), the intron was most

    likely acquired in the common ancestor of vascular

    plants and has been vertically inherited ever since, with

    secondary losses in a small number of plant lineages.

    3.6. nad1i669

    This is another trans-spliced intron in angiosperms

    (Chapdelaine and Bonen, 1991; Conklin et al., 1991;Wissinger et al., 1991), and its cis-spliced ortholog has

    been isolated from Equisetum telmateia and Osmunda

    regalis (Malek and Knoop, 1998). In this study, we ob-

    tained sequences ofcis-spliced introns from Huperzia and

    Psilotum as well as from a different species of Equisetum

    (E. arvense) and a different accession ofOsmunda regalis

    than those used by Malek and Knoop (1998) (Table 1,

    Fig. 2). The size of the intron varies from 1561 bp in E.

    arvense to 3090 bp in Huperzia, with size in Psilotum and

    Osmunda being intermediate. The intron phylogeny

    (Fig. 4F) shows no sign of horizontal transfer. Besides its

    entire absence in green algae and bryophytes, the intron is

    also absent in Isoetes, Adiantum, and Asplenium, which

    can be attributed to secondary losses (Fig. 2). Its distri-

    bution pattern is consistent with the idea that the intron

    was gained in the ancestor of vascular plants. Failure of

    obtaining PCR product in Angiopteris, Marattia, Ophio-

    glossum,and Cycas again could indicate presence oftrans-

    spliced introns in these plants.

    3.7. nad1i728

    This intron is orthologous to the nad1.i4 that has

    previously been isolated from Sphagnum L. and

    256 O. Dombrovska, Y.-L. Qiu / Molecular Phylogenetics and Evolution 32 (2004) 246263

  • 7/28/2019 1-s2.0-S1055790304000326-main

    12/18

    Notothylas (Qiu et al., 1998) as well as several angio-

    sperms (Chapdelaine and Bonen, 1991; Conklin et al.,

    1991; Kubo et al., 2000; Notsu et al., 2002; Thomson

    et al., 1994; Unseld et al., 1997; Wahleithner et al., 1990;

    Wissinger et al., 1991). It has also been shown to be

    present in a large number of land plants except liver-

    worts according to the Southern hybridizations (Qiuet al., 1998; Qiu and Palmer, in press) and matR (an

    ORF encoded within the intron) sequencing (Anderberg

    et al., 2002; Qiu et al., 1999). Here, we report sequences

    of this intron from 19 mosses, 3 hornworts (the Not-

    othylas intron sequence was taken from Qiu et al. (1998)

    and hence was not included in this count), and 9 pteri-

    dophytes (Table 1, Fig. 2). The intron in hornworts and

    pteridophytes shows a range of size variation from

    2221 bp in Osmunda to 4037 bp in Marattia. However, in

    mosses, while two Takakia species still have a full-sized

    intron (slightly over 2800 bp), Sphagnum and other

    mosses have lost most or all of their matR (with the

    intron size being 1452bp in Sphagnum and around

    850 bp in all other mosses). The portion of matR still

    present in Sphagnum is the X domain, which is impli-

    cated in intron splicing and is usually retained in any

    group II intron that contains an ORF (Mohr et al.,

    1993). The phylogeny of the intron (Fig. 4G) mirrors

    that of the plants rather well, indicating that the intron

    has been inherited vertically since its acquisition in the

    common ancestor of non-liverwort land plants as sug-

    gested by Qiu et al. (1998). For both Psilotum and

    Ophioglossum, we obtained only partial sequences of the

    intron using primers in the adjacent exons and in 5 0- or

    30

    -end of the intron, and thus cannot confirm splicingstatus of the intron. Isoetes was reported to lack the

    intron in a Southern hybridization survey (Qiu et al.,

    1998), and our attempt to use PCR to check for the

    presence of the intron failed this time.

    4. Discussion

    4.1. Phylogenetic implications of intron distribution in

    nad1 of land plants

    A total of seven introns, all belonging to group II,

    have been discovered in the mitochondrial gene nad1

    through surveys of a large number of land plants by this

    study and many others (Chapdelaine and Bonen, 1991;

    Conklin et al., 1991; Gugerli et al., 2001; Kubo et al.,

    2000; Malek and Knoop, 1998; Notsu et al., 2002; Qiu

    et al., 1998; Qiu and Palmer, in press; Thomson et al.,

    1994; Unseld et al., 1997; Wahleithner et al., 1990;

    Wissinger et al., 1991). Yet, no intron has been found in

    the 12 liverworts sequenced in this study nor in

    Marchantia by Oda et al. (1992), which span most of the

    liverwort diversity: Calobryales, Marchantiales, Metz-

    geriales, and Jungermanniales (Schuster, 1966, 2000).

    This complete absence of introns in liverworts is the

    same condition as in the green algal outgroups (Kroy-

    mann and Zetsche, 1998; K uck et al., 2000; Nedelcu

    et al., 2000; Turmel et al., 1999, 2002a,b, 2003; Wolff

    et al., 1994; M.W. Gray, unpublished data). The sole

    group I intron in nad1 of Chlamydomonas eugametos

    (Denovan-Wright et al., 1998), but not in C. reinhardtii(M.W. Gray, unpublished data (GenBank Accession

    No. U03843)), is irrelevant to the discussion here, as its

    distribution clearly suggests that it was gained via hor-

    izontal transfer.

    Earlier, Qiu et al. (1998) used the hypothesized gains

    of nad1i728 (nad1.i4 in their nomenclature) and two

    other introns in the mitochondrial gene cox2, cox2i373

    and cox2i691, to argue for the basal-most position of

    liverworts among land plants. Nickrent et al. (2000)

    suggested that absence of these introns in the liverworts

    could be caused by secondary losses, as seen in some

    plants reported in Qiu et al. (1998). Further, Nishiyama

    and Kato (1999), Nickrent et al. (2000), and Renzaglia

    et al. (2000) conducted phylogenetic analyses of multi-

    gene data sets (with limited taxon sampling) and sper-

    matogenesis characters and concluded that hornworts

    occupy the basal-most position in land plants. In light of

    the data presented here, we believe that the most par-

    simonious explanation for intron distribution in nad1 is

    still that liverworts, not hornworts, represent the oldest

    land plants, and that both nad1i287 and nad1i728 were

    gained in the common ancestor of mosseshornworts

    vascular plants after liverworts had diverged. While the

    former was lost in the ancestor of vascular plants, the

    latter had been retained. The idea of a single gain ofnad1i287 in a common ancestor of mosses and horn-

    worts can only be entertained if the two groups form a

    clade, but both published studies (Kenrick and Crane,

    1997b; Mishler et al., 1994; Nickrent et al., 2000; Ni-

    shiyama and Kato, 1999; Renzaglia et al., 2000; see also

    Qiu and Lee, 2000) and our unpublished multigene

    analysis of land plant phylogeny have failed to recover

    this topology. If we follow the hornwort basal hypoth-

    esis, both introns would have to be gained in the com-

    mon ancestor of land plants and lost in liverworts, with

    an additional loss ofnad1i287in vascular plants since no

    study so far has suggested a sister relationship of liver-

    worts and vascular plants (Qiu and Lee, 2000). This

    scenario is significantly less parsimonious than the ex-

    planation offered by the liverwort basal hypothesis. The

    hornwort basal hypothesis is also contradicted by the

    distribution of two other mitochondrial group II in-

    trons, one in nad2 (nad2i1282, Pruchner et al., 2002) and

    the other in nad5 (nad5i477, Malek and Knoop, 1998),

    both of which are present in vascular plants and horn-

    worts but not in liverworts and mosses. Meanwhile, two

    introns in nad7, nad7i40, and nad7i209, have now been

    found in mosses, hornworts, and vascular plants but not

    in liverworts, thus adding more evidence to the liverwort

    O. Dombrovska, Y.-L. Qiu / Molecular Phylogenetics and Evolution 32 (2004) 246263 257

  • 7/28/2019 1-s2.0-S1055790304000326-main

    13/18

    basal hypothesis (Hashimoto and Sato, 2001; Pruchner

    et al., 2001; N. Sato, personal communication). There-

    fore, the overall intron distribution in the mitochondrial

    genes (cox2, nad1, nad2, nad5, and nad7) is more con-

    sistent with the liverwort basal hypothesis than with the

    hornwort basal hypothesis.

    The liverwort basal hypothesis received an indirectsupport from recently completed studies of chloroplast

    genome sequencing of Physcomitrella patens (a moss,

    Sugiura et al., 2003), Anthoceros formosae (Kugita et al.,

    2003a), Psilotum nudum (Wakasugi et al., unpublished

    data, GenBank Accession No. AP004638), and Adian-

    tum capillus-veneris (Wolf et al., 2003). In Marchantia

    (Ohyama et al., 1986), Physcomitrella, and Chaetosph-

    aeridium (Turmel et al., 2002b), the inverted repeat in

    the chloroplast genome does not include the genes ndhB,

    rps7, and rps12 (partial). However, Anthoceros cpDNA

    has these three genes as part of its inverted repeat, a

    condition otherwise only seen in Psilotum, Adiantum,

    and other vascular plants such as rice (Hiratsuka et al.,

    1989) and tobacco (Wakasugi et al., 1998). This ex-

    pansion of the inverted repeat to encompass these three

    genes can almost certainly be regarded as a synapo-

    morphy to unite hornworts and vascular plants, a rela-

    tionship alluded to by the aforementioned phylogenetic

    distribution of nad2i1282 and nad5i477 as well as some

    sequence analyses (Lewis et al., 1997; Samigullin et al.,

    2002; see also Qiu and Lee, 2000) (admittedly, taxon

    sampling, in particular of vascular plants, in these se-

    quence analyses is rather thin). Lastly, the most recent

    report of fossil evidence from the Ordovician rocks of

    Oman on sporangium structure that contain sporetetrads characteristic of early land plants also support

    the liverwort basal hypothesis (Wellman et al., 2003).

    A recent study found that certain genomic structural

    changes, such as ribosomal protein gene loss and

    transfer from the mitochondrial to nuclear genomes,

    occur at rather high frequencies (Adams et al., 2002).

    Another study also showed that a group II intron had

    likely experienced horizontal transfer from angiosperms

    to Gnetum (Won and Renner, 2003). Moreover, deep

    losses of introns have clearly happened, as can be seen

    from this study and others that investigated mitochon-

    drial introns in basal land plants (Beckert et al., 1999,

    2001; Pruchner et al., 2002). These studies raise the issue

    of utility of genomic structural characters in inferring

    plant phylogenetic relationships. However, our phylo-

    genetic analyses of intron sequences showed little sign of

    independent gains and horizontal transfer of the group

    II introns investigated here (Fig. 4). Hence, we contend

    that while one certainly cannot use genomic structural

    characters, nor any other kind of characters, blindly for

    phylogenetic inference, it is appropriate to use intron

    distribution patterns, particularly intron gains, if they

    are demonstrated to occur at very low frequencies, to

    evaluate competing phylogenetic hypotheses. Intron

    distribution patterns as well as other genomic structural

    characters such as gene order and content in the cpDNA

    inverted repeat will add a new source of information for

    plant phylogenetic reconstruction.

    Another issue that these intron distribution data may

    help to resolve is the phylogenetic position of the enig-

    matic bryophyte genus Takakia. Though first thought tobe related to liverworts (Hattori and Inoue, 1958;

    Schuster, 1966) or as an independent lineage at the same

    rank as liverworts, mosses, and hornworts (Crandall-

    Stotler, 1986), discovery of antheridia and sporophytes

    has revived an earlier idea that the genus shows affinity

    to mosses (Mizutani, 1967; Smith and Davison, 1993).

    Recent phylogenetic analyses of morphological and

    molecular data seem to support this conclusion, but the

    exact placement of Takakia within mosses is far from

    being certain (Hedderson et al., 1998; Kenrick and

    Crane, 1997b; Mishler et al., 1994; Newton et al., 2000;

    Renzaglia et al., 2000). In this study, we found that the

    intron content in nad1 of both Takakia species is the

    same as 17 diverse mosses we investigated, but different

    from that of liverworts and hornworts (Table 1, Fig. 2).

    Hence, a moss affinity for the genus seems most prob-

    able. Furthermore, our examination of the ORF (matR)

    in nad1i728 indicates that the two Takakia species still

    have an intact matR, whereas Sphagnum has only the X

    domain of the ORF and the other 16 mosses lack the

    ORF entirely. Two lines of evidence support the fol-

    lowing ordering of the three states in these taxa: Takakia

    (ancestral), Sphagnum (intermediate), and all other

    mosses (derived). One is that the X domain has been

    suggested to be the least dispensable element in group IIintron ORFs since it is involved in intron splicing (Mohr

    et al., 1993). The other is that no matR loss has been

    observed in any of the large number of other land plants

    that have been investigated (Anderberg et al., 2002; Qiu

    et al., 1999; Qiu and Palmer, in press). Consequently, the

    matR distribution data would support a topology in

    which Takakia is sister to the clade of Sphagnum plus

    other mosses, as has been recovered in some recent

    sequence analyses (Renzaglia et al., 2000; Qiu et al.,

    unpublished data).

    The data of the present study may also address the

    monophyly of moniliformopses, a group that comprises

    traditionally circumscribed ferns, Equisetum, and Psi-

    lotaceae. The clade was first identified by Kenrick and

    Crane (1997b) after they analyzed both extant and ex-

    tinct pteridophytes, and was subsequently supported by

    both multigene sequence analysis and a shared unique

    insertion in the chloroplast gene rps4 (Pryer et al., 2001).

    In this study, we found that the ferns (both eusporan-

    giate and leptosporangiate) and Equisetum share the

    intron nad1i258. Unfortunately, no sequence was ob-

    tained from Psilotaceae. Our phylogenetic analyses of

    the intron sequences indicate that the intron was most

    likely acquired in a single event in the common ancestor

    258 O. Dombrovska, Y.-L. Qiu / Molecular Phylogenetics and Evolution 32 (2004) 246263

  • 7/28/2019 1-s2.0-S1055790304000326-main

    14/18

    of moniliformopses and has been inherited vertically

    since the acquisition (Fig. 4A). This intron gain would

    then be a synapomorphy to support the monophyly of

    moniliformopses.

    Finally, three introns that were first characterized in

    angiosperms, nad1i394, nad1i477, and nad1i669, are now

    found to be present throughout vascular plants (Fig. 2).Thus, they were most likely gained in the common an-

    cestor of the group. These characters would add to the

    large body of information already in existence that

    supports the monophyly of vascular plants (Kenrick and

    Crane, 1997b).

    4.2. Molecular evolutionary implications of intron distri-

    bution in nad1 of land plants

    One conspicuous aspect of distribution of the seven

    nad1 introns in land plants is that they are all of broad

    phylogenetic occurrence. The intron that shows the

    narrowest phylogenetic range is nad1i258, which is still

    found in Equisetum, and eusporangiate and leptospo-

    rangiate ferns (Table 1; Fig. 2). The common ancestor of

    these plants likely existed between the Silurian and

    Devonian (435345 million years ago) (Kenrick and

    Crane, 1997a,b). The other intron that is restricted to

    only one lineage is nad1i348, which occurs in all four

    genera (belonging to the only two families) of hornworts

    that were examined (Table 1; Fig. 2). The hornworts as a

    lineage must have existed between the mid-Ordovician

    (about 450 million years ago) and Silurian, as fossils of

    liverwort affinity have been reported from the mid-

    Ordovician (Edwards et al., 1995; Taylor, 1995; Well-man et al., 2003) and those of vascular plants from the

    lower Silurian (Kenrick and Crane, 1997a,b). All other

    five introns are found to be shared by at least two major

    lineages of land plants (Table 1; Fig. 2). Hence, these

    data suggest that these introns arrived at their current

    positions at least during the Silurian to Devonian pe-

    riod. There is also evidence suggesting that these and

    other group II introns in the non-liverwort land plant

    mitochondrial genome originated via transposition from

    pre-existing introns in other genes of liverworts [cur-

    rently based on data mostly from Marchantia (Oda

    et al., 1992)] and other organisms (Dombrovska and

    Qiu, in preparation). If these observations hold, it seems

    that group II intron transposition was quite active in the

    mitochondrial genome during the early stage of land

    plant evolution. Meanwhile, among studies that have

    surveyed introns in a broad range of land plants (Ha-

    shimoto and Sato, 2001; Malek and Knoop, 1998;

    Pruchner et al., 2001; Qiu et al., 1998), only one intron

    has been shown to be restricted to a single species, the

    nad5i392 in Huperzia selago (Vangerow et al., 1999).

    This general lack of group II introns with extremely

    narrow phylogenetic distribution leads us to conclude

    that they have not been actively undergoing transposi-

    tion in the recent history of land plant evolution. Fur-

    ther, our phylogenetic analyses of all seven nad1 introns

    uncovered no evidence of horizontal transfer, specifi-

    cally intron homing (Fig. 4). This situation is in stark

    contrast to what was reported by Cho et al. (1998) of

    rampant horizontal transfer (homing) by a group I in-

    tron in the mitochondrial gene cox1 of angiosperms.The only explanation we can offer is that the cox1 intron

    carries an endonuclease ORF whereas most of the group

    II introns in vascular plant mitochondrial genomes have

    lost their ORFs, which contain domains involved in

    intron mobility (Mohr et al., 1993; Zimmerly et al.,

    2001).

    Despite stable inheritance after ancient gains, some of

    the introns did experience losses, possibly very deep

    losses. The distribution pattern of nad1i287 potentially

    suggests a deep loss in the common ancestor of vascular

    plants. The likelihood that this intron was gained once

    in a common ancestor of mosses and hornworts or twice

    in ancestors of each lineage is much less than that of a

    single gain in the ancestor of non-liverwort land plants

    and a subsequent loss in vascular plants, given the in-

    tron phylogeny (Fig. 4B) and current results on land

    plant phylogeny reconstruction (Kenrick and Crane,

    1997b; Mishler et al., 1994; Renzaglia et al., 2000; Qiu

    et al., unpublished data). Three other introns, nad1i394,

    nad1i477, and nad1i669, also showed a few sporadic

    losses (Fig. 2; Gugerli et al., 2001), but they are all

    phylogenetically restricted. Both deep and shallow losses

    have also been observed for two cox2 introns in angio-

    sperms (Joly et al., 2001; Kudla et al., 2002; Rabbi and

    Wilson, 1993; Qiu and Palmer, unpublished data). Theprecise mechanisms for how both deep and shallow in-

    tron losses have occurred remain largely unknown at

    this stage. Among all plants studied, Isoetes is distinct in

    that it has lost two, if not three (the third one possibly

    being nad1i728, the exact status of which still cannot be

    confirmed by sequencing because of our failure to ob-

    tain a PCR product), of the four introns that are usually

    present in nad1 in vascular plants (Fig. 2). Even when

    introns are present in this and other mitochondrial genes

    (nad2 and nad5) in Isoetes, they are unusually small

    (384520bp, this study; Malek and Knoop, 1998;

    Pruchner et al., 2002). These two lines of evidence sug-

    gest that the mitochondrial genome in Isoetes is proba-

    bly under strong selection pressure for genome size

    reduction, as has been seen in some abnormal genomes

    (Douglas et al., 2001).

    4.3. RNA editing in nad1 of land plants

    RNA editing, post-transcriptional modification of

    cytidine to uridine and vice versa in RNAs, has been

    reported to occur commonly in organellar genes and

    genomes of certain land plants (Giege and Brennicke,

    1999; Kugita et al., 2003b; Steinhauser et al., 1999).

    O. Dombrovska, Y.-L. Qiu / Molecular Phylogenetics and Evolution 32 (2004) 246263 259

  • 7/28/2019 1-s2.0-S1055790304000326-main

    15/18

    In this study, we noticed that after aligning the exon

    sequences from all the plants, some lineages seem to

    have numerous aberrant C!T(U) and T(U)!C mu-

    tations in the first or second codon positions. Hence, we

    checked for the possibility of RNA editing by looking

    for whether the mutation has altered a conserved amino

    acid or generated an internal stop codon in some cases.Obviously, this approach is not as conclusive and

    accurate as cDNA analysis, but has been employed and

    proven to be valid by others (Steinhauser et al., 1999).

    With this approach, we found that the nad1 gene in the

    liverwort Haplomitrium, two Takakia species, all four

    hornworts, and several pteridophytes (Isoetes, Ophio-

    glossum, Adiantum, and Asplenium) potentially experi-

    ence heavy editing, with up to 5.11% of all thenucleotide positions in Anthoceros being edited (Fig. 5).

    Fig. 5. Selected codons of the mitochondrialnad1 in land plants showing the extent of RNA editing. Inferred edited nucleotides are underlined, body-

    faced ones having been confirmed by cDNA analysis in the studies that reported editing. Editing frequency (%) equals the number of edited nu-

    cleotides divided by the exon length examined. Boxes mark stop codons that require editing for conversion. Dots indicate nucleotides identical to

    those in Marchantia polymorpha and dashes denote missing data. Marchantia position refers to the position of the first nucleotide in a codon.

    260 O. Dombrovska, Y.-L. Qiu / Molecular Phylogenetics and Evolution 32 (2004) 246263

  • 7/28/2019 1-s2.0-S1055790304000326-main

    16/18

    Previously, high frequencies of RNA editing have been

    noticed in many of these plants in other mitochondrial

    genes, cox3, nad2, nad4, and nad5 (Malek and Knoop,

    1998; Malek et al., 1996; Pruchner et al., 2001, 2002).

    However, questions remain as to why there is RNA

    editing in the first place and why it occurs in certain

    lineages but not in others.

    Acknowledgments

    We thank J.D. Palmer for providing initial ideas on

    starting this project, and V. Knoop for stimulating dis-

    cussion on plant mitochondrial intron evolution, and

    both of them for their advice on developing the intron

    nomenclature. We also thank B. Crandall-Stotler and E.

    Urmi for help with obtaining plant material. This work

    was supported by a Swiss NSF research grant (3100-

    053602) and a NSF CAREER Award (DEB 0093012) to

    Y.Q.

    References

    Adams, K.L., Qiu, Y.-L., Stoutemyer, M., Palmer, J.D., 2002.

    Punctuated evolution of mitochondrial gene content: high and

    variable rates of mitochondrial gene loss and transfer during

    angiosperm evolution. Proc. Natl. Acad. Sci. USA 99, 99059912.

    Anderberg, A.A., Rydin, C., Kallersjo, M., 2002. Phylogenetic

    relationships in the order Ericales s.l.: analyses of molecular data

    from five genes from the plastid and mitochondrial genomes. Am.

    J. Bot. 89, 677687.

    Beckert, S., Muhle, H., Pruchner, D., Knoop, V., 2001. Themitochondrial nad2 gene as a novel marker locus for phylogenetic

    analysis of early land plants: a comparative analysis in mosses.

    Mol. Phylogenet. Evol. 18, 117126.

    Beckert, S., Steinhauser, S., Muhle, H., Knoop, V., 1999. A molecular

    phylogeny of bryophytes based on nucleotide sequences of the

    mitochondrial nad5 gene. Plant Syst. Evol. 218, 179192.

    Chapdelaine, Y., Bonen, L., 1991. The wheat mitochondrial gene for

    subunit I of the NADH dehydrogenase complex: a trans-splicing

    model for this gene-in-pieces. Cell 65, 465472.

    Cho, Y., Qiu, Y.-L., Kuhlman, P., Palmer, J.D., 1998. Explosive

    invasion of plant mitochondria by a group I intron. Proc. Natl.

    Acad. Sci. USA 95, 1424414249.

    Conklin, P.L., Wilson, R.K., Hanson, M.R., 1991. Multiple trans-

    splicing events are required to produce a mature nad1 transcript in

    a plant mitochondrion. Genes Dev. 5, 14071415.Crandall-Stotler, B., 1986. Morphogenesis, developmental anatomy

    and bryophyte phylogenetics: contradictions of monophyly. J.

    Bryol. 14, 123.

    Denovan-Wright, E.M., Nedelcu, A.M., Lee, R.W., 1998. Complete

    sequence of the mitochondrial DNA ofChlamydomonas eugametos.

    Plant Mol. Biol. 36, 285295.

    Douglas, S., Zauner, S., Fraunholz, M., Beaton, M., Penny, S., Deng,

    L.-T., Wu, X., Reith, M., Cavalier-Smith, T., Maier, U.-G., 2001.

    The highly reduced genome of an enslaved algal nucleus. Nature

    410, 10911096.

    Doyle, J.J., Doyle, J.S., 1987. A rapid DNA isolation procedure for

    small quantities of fresh leaf tissue. Phytochem. Bull. 19, 1115.

    Edwards, D., Duckett, J.G., Richardson, J.B., 1995. Hepatic charac-

    ters in the earliest land plants. Nature 374, 635636.

    Gensel, P.G., Edwards, D., 2001. Plants Invade the Land: Evolution-

    ary and Environmental Perspectives. Columbia University Press,

    New York, NY.

    Giege, P., Brennicke, A., 1999. RNA editing in Arabidopsis mitochon-

    dria effects 441 C to U changes in ORFs. Proc. Natl. Acad. Sci.

    USA 96, 1532415329.

    Graham, L.E., 1993. Origin of Land Plants. Wiley, New York, NY.

    Gugerli, F., Sperisen, C., Buchler, U., Brunner, I., Brodbeck, S.,

    Palmer, J.D., Qiu, Y.-L., 2001. The evolutionary split of Pinaceae

    from other conifers: evidence from an intron loss and a multigene

    phylogeny. Mol. Phylogenet. Evol. 21, 167175.

    Hallick, R.B., Hong, L., Drager, R.G., Favreau, M.R., Monfort, A.,

    Orsat, B., Spielmann, A., Stutz, E., 1993. Complete sequence of

    Euglena gracilis chloroplast DNA. Nucleic Acids Res. 21, 3537

    3544.

    Hashimoto, K., Sato, N., 2001. Characterization of the mitochondrial

    nad7 gene in Physcomitrella patens: similarity with angiosperm

    nad7 genes. Plant Sci. 160, 807815.

    Hattori, S., Inoue, H., 1958. Preliminary report on Takakia lepido-

    zioides. J. Hattori Bot. Lab. 19, 133137.

    Hedderson, T.A., Chapman, R., Cox, C.J., 1998. Bryophytes and the

    origins and diversification of land plants: new evidence from

    molecules. In: Bates, J.W., Ashton, N.W., Duckett, J.G. (Eds.),

    Bryology for the Twenty-first Century. W.S. Maney & Son Ltd.,

    Leeds, UK, pp. 6577.

    Hiratsuka, J., Shimada, H., Whittier, R., Ishibashi, T., Sakamoto, M.,

    Mori, M., Kondo, C., Honji, Y., Sun, C.-R., Meng, B.-Y., Li, Y.

    -Q., Kanno, A., Nishizawa, Y., Hirai, A., Shinozaki, K., Sugiura,

    M., 1989. The complete sequence of the rice (Oryza sativa)

    chloroplast genome: intermolecular recombination between distinct

    tRNA genes accounts for a major plastid DNA inversion during

    the evolution of the cereals. Mol. Gen. Genet. 217, 185194.

    Johansen, S., Haugen, P., 2001. A new nomenclature of group I

    introns in ribosomal DNA. RNA 7, 935936.

    Joly, S., Brouillet, L., Bruneau, A., 2001. Phylogenetic implications of

    the multiple losses of the mitochondrial coxII.i3 intron in the

    angiosperms. Int. J. Plant Sci. 162, 359373.

    Kenrick, P., Crane, P.R., 1997a. The origin and early evolution ofplants on land. Nature 389, 3339.

    Kenrick, P., Crane, P.R., 1997b. The Origin and Early Diversification

    of Land Plants: A Cladistic Study. Smithsonian Institution Press,

    Washington, DC.

    Kroymann, J., Zetsche, K., 1998. The mitochondrial genome of

    Chlorogonium elongatum inferred from the complete sequence. J.

    Mol. Evol. 47, 431440.

    Kubo, T., Nishizawa, S., Sugawara, A., Itchoda, N., Estiati, A.,

    Mikami, T., 2000. The complete nucleotide sequence of the

    mitochondrial genome of sugar beet (Beta vulgaris L.) reveals a

    novel gene for tRNACys (GCA). Nucleic Acids Res. 28, 25712576.

    Kuck, U., Jekosch, K., Holzamer, P., 2000. DNA sequence analysis of

    the complete mitochondrial genome of the green alga Scenedesmus

    obliquus: evidence for UAG being a leucine and UCA being a non-

    sense codon. Gene 253, 1318.Kudla, J., Albertazzi, F.J., Blazevic, D., Hermann, M., Bock, R., 2002.

    Loss of the mitochondrial cox2 intron 1 in a family of monocoty-

    ledonous plants and utilization of mitochondrial intron sequences

    for the construction of a nuclear intron. Mol. Genet. Genomics

    267, 223230.

    Kugita, M., Kaneko, A., Yamamoto, Y., Takeya, Y., Matsumoto, T.,

    Yoshinaga, K., 2003a. The complete nucleotide sequence of the

    hornwort (Anthoceros formosae) chloroplast genome: insight into

    the earliest land plants. Nucleic Acids Res. 31, 716721.

    Kugita, M., Yamamoto, Y., Fujikawa, T., Matsumoto, T., Yoshinaga,

    K., 2003b. RNA editing in hornwort chloroplasts makes more than

    half the genes functional. Nucleic Acids Res. 31, 24172423.

    Lambowitz, A.M., Belfort, M., 1993. Introns as mobile genetic

    elements. Annu. Rev. Biochem. 62, 587622.

    O. Dombrovska, Y.-L. Qiu / Molecular Phylogenetics and Evolution 32 (2004) 246263 261

  • 7/28/2019 1-s2.0-S1055790304000326-main

    17/18

    Lewis, L.A., Mishler, B.D., Vilgalys, R., 1997. Phylogenetic relation-

    ships of the liverworts (Hepaticae), a basal embryophyte lineage,

    inferred from nucleotide sequence data of the chloroplast gene

    rbcL. Mol. Phylogenet. Evol. 7, 377393.

    Malek, O., Knoop, V., 1998. Trans-splicing group II introns in plant

    mitochondria: the complete set ofcis-arranged homologs in ferns,

    fern allies, and a hornwort. RNA 4, 15991609.

    Malek, O., Brennicke, A., Knoop, V., 1997. Evolution oftrans-splicing

    plant mitochondrial introns in pre-Permian times. Proc. Natl.

    Acad. Sci. USA 94, 553558.

    Malek, O., Lattig, K., Hiesel, R., Brennicke, A., Knoop, V., 1996.

    RNA editing in bryophytes and a molecular phylogeny of land

    plants. EMBO J. 15, 14031411.

    Manhart, J.R., Palmer, J.D., 1990. The gain of two chloroplast tRNA

    introns marks the green algal ancestors of land plants. Nature 345,

    268270.

    Mathews, D.H., Sabina, J., Zuker, M., Turner, D.H., 1999. Expanded

    sequence dependence of thermodynamic parameters improves

    prediction of RNA secondary structure. J. Mol. Biol. 288, 911940.

    Michel, F., Umesono, K., Ozeki, H., 1989. Comparative and functional

    anatomy of group II catalytic intronsa review. Gene 82, 530.

    Mishler, B.D., Lewis, L.A., Buchheim, M.A., Renzaglia, K.S.,

    Garbary, D.J., Delwiche, C.F., Zechman, F.W., Kantz, T.S.,

    Chapman, R.L., 1994. Phylogenetic relationships of the green

    algae and bryophytes. Ann. Missouri Bot. Gard. 81, 451483.

    Mizutani, M., 1967. A new knowledge of archegonia of Takakia

    lepidozioides. J. Jpn. Bot. 42, 379381.

    Mohr, G., Perlman, P.S., Lambowitz, A.M., 1993. Evolutionary

    relationships among group II intron-encoded proteins and identi-

    fication of a conserved domain that may be related to maturase

    function. Nucleic Acids Res. 21, 49914997.

    Nakayama, H., Morinaga, Y., Nomura, N., Nunoura, T., Sako, Y.,

    Uchida, A., 2003. An archaeal homing endonuclease I-Pog I

    cleaves at the insertion site of the neighboring intron, which has no

    nested open reading frame. FEBS 544, 165170.

    Nedelcu, A.M., Lee, R.W., Lemieux, C., Gray, M.W., Burger, G.,

    2000. The complete mitochondrial DNA sequence ofScenedesmus

    obliquus reflects an intermediate stage in the evolution of the greenalgal mitochondrial genome. Genome Res. 10, 819831.

    Newton, A.E., Cox, C.J., Duckett, J.G., Wheeler, J.A., Goffinet, B.,

    Hedderson, T.A.J., Mishler, B.D., 2000. Evolution of major moss

    lineages: phylogenetic analyses based on multiple gene sequences

    and morphology. Bryologist 103, 187211.

    Nickrent, D.L., Parkinson, C.L., Palmer, J.D., Duff, R.J., 2000.

    Multigene phylogeny of land plants with special reference to

    bryophytes and the earliest land plants. Mol. Biol. Evol. 17, 1885

    1895.

    Nishiyama, T., Kato, M., 1999. Molecular phylogenetic analysis

    among bryophytes and tracheophytes based on combined data of

    plastid coded genes and the 18S rRNA gene. Mol. Biol. Evol. 16,

    10271036.

    Notsu, Y., Masood, S., Nishikawa, T., Kubo, N., Akiduki, G.,

    Nakazono, M., Hirai, A., Kadowaki, K., 2002. The completesequence of the rice (Oryza sativa L.) mitochondrial genome:

    frequent DNA sequence acquisition and loss during the evolution

    of flowering plants. Mol. Genet. Genomics 268, 434445.

    Oda, K., Yamato, K., Ohta, E., Nakamura, Y., Takemura, M.,

    Nozato, N., Akashi, K., Kanegae, T., Ogura, Y., Kohchi, T.,

    Ohyama, K., 1992. Gene organization deduced from the complete

    sequence of liverwort Marchantia polymorpha mitochondrial DNA:

    a primitive form of plant mitochondrial genome. J. Mol. Biol. 223,

    17.

    Ohyama, K., Fukuzawa, H., Kohchi, T., Shirai, H., Sano, T., Sano, S.,

    Umesono, K., Shiki, Y., Takeuchi, M., Chang, Z., Aota, S.,

    Inokuchi, H., Ozeki, H., 1986. Chloroplast gene organization

    deduced from complete sequence of liverwort Marchantia poly-

    morpha chloroplast DNA. Nature 322, 572574.

    Pruchner, D., Beckert, S., Muhle, H., Knoop, V., 2002. Divergent

    intron conservation in the mitochondrial nad2 gene: signatures for

    the three bryophyte classes (mosses, liverworts, and hornworts) and

    the lycophytes. J. Mol. Evol. 55, 265271.

    Pruchner, D., Nassal, B., Schindler, M., Knoop, V., 2001. Mosses

    share mitochondrial group II introns with flowering plants, not

    with liverworts. Mol. Genet. Genomics 266, 608613.

    Pryer, K.M., Schneider, H., Smith, A.R., Cranfill, R., Wolf, P.G.,

    Hunt, J.S., Sipes, S.D., 2001. Horsetails and ferns are a mono-

    phyletic group and the closest relatives to seed plants. Nature 409,

    618622.

    Qin, P.Z., Pyle, A.M., 1998. The architectural organization and

    mechanistic function of group II intron structural elements. Curr.

    Opin. Struct. Biol. 8, 301308.

    Qiu, Y.-L., Cho, Y., Cox, J.C., Palmer, J.D., 1998. The gain of three

    mitochondrial introns identifies liverworts as the earliest land

    plants. Nature 394, 671674.

    Qiu, Y.-L., Lee, J., Bernasconi-Quadroni, F., Soltis, D.E., Soltis, P.S.,

    Zanis, M., Zimmer, E.A., Chen, Z., Savolainen, V., Chase, M.W.,

    1999. The earliest angiosperms: evidence from mitochondrial,

    plastid and nuclear genomes. Nature 402, 404407.

    Qiu, Y.-L., Lee, J., 2000. Transition to a land flora: a molecular

    phylogenetic perspective. J. Phycol. 36, 799802.

    Qiu, Y.-L., Palmer, J.D., in press. Many independent origins oftrans-

    splicing of a plant mitochondrial group II intron. J. Mol. Evol.

    Rabbi, M.F., Wilson, K.G., 1993. The mitochondrial coxII intron has

    been lost in two different lineages of dicots and altered in others.

    Am. J. Bot. 80, 12161223.

    Raubeson, L.A., Jansen, R.K., 1992. Chloroplast DNA evidence on

    the ancient evolutionary split in vascular land plants. Science 255,

    16971699.

    Renzaglia, K.S., Duff, R.J., Nickrent, D.L., Garbary, D.J., 2000.

    Vegetative and reproductive innovations of early land plants:

    implications for a unified phylogeny. Philos. Trans. R. Soc. B 355,

    769793.

    Samigullin, T.K., Yacentyuk, S.P., Degtyaryeva, G.V., Valiehoroman,

    K.M., Bobrova, V.K., Capesius, I., Martin, W.M., Troitsky, A.V.,

    Filin, V.R., Antonov, A.S., 2002. Paraphyly of bryophytes and closerelationship of hornworts and vascular plants inferredfrom analysis

    of chloroplast rDNA ITS (cpITS) sequences. Arctoa 11, 3143.

    Schuster, R.M., 1966. The Hepaticae and Anthocerotae of North

    America, vol. 1. Columbia University Press, New York, NY.

    Schuster, R.M., 2000. Austral Hepaticae, Part I. J. Cramer in der

    Gebrueder Borntraeger Verlagsbuchhandlung, Stuttgart.

    Sharp, P.A., 1981. Speculations on RNA splicing. Cell 23, 643646.

    Smith, D.K., Davison, P.G., 1993. Antheridia and sporophytes in

    Takakia ceratophylla (Mitt.) Grolle: evidence for reclassification

    among the mosses. J. Hattori Bot. Lab. 73, 263271.

    Steinhauser, S., Beckert, S., Capesius, I., Malek, O., Knoop, V., 1999.

    Plant mitochondrial RNA editing. J. Mol. Evol. 48, 303312.

    Sugiura, C., Kobayashi, Y., Aoki, S., Sugita, C., Sugita, M., 2003.

    Complete chloroplast DNA sequence of the moss Physcomitrella

    patens: evidence for the loss and relocation of rpoA from thechloroplast to the nucleus. Nucleic Acids Res. 31, 53245331.

    Swofford, D.L., 2003. PAUP*4.0b10: Phylogenetic analysis using

    parsimony. Sinauer, Sunderland, MA.

    Taylor, W.A., 1995. Spores in earliest land plants. Nature 373, 391

    392.

    Thomson, M.C., MacFarlane, J.L., Beagley, C.T., Wolstenholme,

    D.R., 1994. RNA editing ofmat-r transcripts in maize and soybean

    increases similarity of the encoded protein to fungal and bryophyte

    group II intron maturases: evidence that mat-r encodes a functional

    protein. Nucleic Acids Res. 22, 57455752.

    Thompson, J.D., Gibson, T.J., Plewniak, F., Jeanmougin, F., Higgins,

    D.G., 1997. The ClustalX windows interface: flexible strategies for

    multiple sequence alignment aided by quality analysis tools.

    Nucleic Acids Res. 24, 48764882.

    262 O. Dombrovska, Y.-L. Qiu / Molecular Phylogenetics and Evolution 32 (2004) 246263

  • 7/28/2019 1-s2.0-S1055790304000326-main

    18/18

    Turmel, M., Lemieux, C., Burger, G., Lang, B.F., Otis, C., Plante, I.,

    Gray, M.W., 1999. The complete mitochondrial DNA sequences of

    Nephroselmis olivacea and Pedinomonas minor: two radically differ-

    ent evolutionary patterns within green algae. Plant Cell 11, 1717

    1729.

    Turmel, M., Otis, C., Lemieux, C., 2002a. The complete mitochondrial

    DNA sequence ofMesostigma viride identifies this green alga as the

    earliest green plant divergence and a highly compact mitochondrial

    genomein theancestor of allgreen plants.Mol. Biol. Evol. 19,2438.

    Turmel, M., Otis, C., Lemieux, C., 2002b. The chloroplast and

    mitochondrial genome sequences of the charophyte Chaetosphae-

    ridium globosum: insights into the timing of the events that

    restructured organelle DNAs within the green algal linegae that

    led to land plants. Proc. Natl. Acad. Sci. USA 99, 1127511280.

    Turmel, M., Otis, C., Lemieux, C., 2003. The mitochondrial genome of

    Chara vulgaris: insights into the mitochondrial DNA architecture

    of the last common ancestor of green algae and land plants. Plant

    Cell 15, 18881903.

    Unseld, M., Marienfeld, J.R., Brand, P., Brennicke, A., 1997. The

    mitochondrial genome ofArabidopsis thaliana contains 57 genes in

    366,924 nucleotides. Nat. Genet. 15, 5761.

    Vangerow, S., Teekorn, T., Knoop, V., 1999. Phylogenetic information

    in the mitochondrial nad5 gene of pteridophytes: RNA editing and

    intron sequences. Plant Biol. 1, 235243.

    Wakasugi, T., Sugita, M., Tsudzuki, T., Sugiura, M., 1998. Updated

    gene map of tobacco chloroplast DNA. Plant Mol. Biol. Rep. 16,

    231241.

    Wahleithner, J.A., MacFarlane, J.L., Wolstenholme, D.R., 1990. A

    sequence encoding a maturase-related protein in a group II intron

    of a plant mitochondrial nad1 gene. Proc. Natl. Acad. Sci. USA 87,

    548552.

    Wellman, C.H., Osterloff, P.L., Mohiuddin, U., 2003. Fragments of

    the earliest land plants. Nature 425, 282285.

    Wissinger, B., Schuster, W., Brennicke, A., 1991. Trans-splicing in

    Oenothera mitochondria: nad1 mRNA are edited in exon and

    trans-splicing group II intron sequences. Cell 65, 473482.

    Wolf, P.G., Rowe, C.A., Sinclair, R.B., Hasebe, M., 2003. Complete

    nucleotide sequence of the chloroplast genome from a leptospo-

    rangiate fern, Adiantum capillus-veneris L. DNA Res. 10, 5965.

    Wolff, G., Plante, I., Lang, B.F., Kuck, U., Burger, G., 1994. Complete

    sequence of the mitochondrial DNA of the chlorophyte alga

    Prototheca wickerhamii: gene content and genome organization. J.

    Mol. Biol. 237, 7586.

    Won, H., Renner, S.S., 2003. Horizontal gene transfer from

    flowering plants to Gnetum. Proc. Natl. Acad. Sci. USA 100,

    1082410829.

    Zimmerly, S., Hausner, G., Wu, X.-C., 2001. Phylogenetic relation-

    ships among group II intron ORFs. Nucleic Acids Res. 29, 1238

    1250.

    Zuker, M., Mathews, D.H., Turner, D.H., 1999. Algorithms and

    thermodynamics for RNA secondary structure prediction: a

    practical guide. In: Barciszewski, J., Clark, B.F.C. (Eds.), RNA

    Biochemistry and Biotechnology. NATO ASI Series. Kluwer

    Academic Publishers, Dordrecht, pp. 1143.

    O. Dombrovska, Y.-L. Qiu / Molecular Phylogenetics and Evolution 32 (2004) 246263 263