Botanical Garden of the University of Osnabrück, Osnabrück, Germany
Department of Biology, University of Leicester, Leicester, England
![]() |
Abstract |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
![]() |
Introduction |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
Retroelements and their derivatives are ubiquitous and abundant components of plant genomes (Flavell, Smith, and Kumar 1992
; Voytas et al. 1992
; Hirochika and Hirochika 1993
; Matsuoka and Tsunewaki 1996, 1999), often representing 50% of all the DNA (Pearce et al. 1996
; SanMiguel et al. 1996
). Based on their structure, the retrotransposons are divided into two groups: those that are flanked by long terminal repeats (LTRs), and non-LTR retrotransposons, or long interspersed nuclear elements (LINEs; see reviews by Kumar and Bennetzen 1999
; Schmidt 1999
). LTR retrotransposons are further divided, most importantly into the two groups Ty1 or copia, and Ty3 or gypsy. The major structural difference between copia and gypsy groups is in the order of the reverse transcriptase (RT) and integrase domains in their pol genes. Gypsy group elements have similarities to retroviruses (see reviews in Bennetzen 1996, 2000
; Kumar and Bennetzen 1999
). The RT genes have conserved amino acid domains, some of which are characteristic of each retroelement group (Xiong and Eickbush 1990
; Eickbush 1994
). In plants, degenerate oligonucleotide primers have been designed to amplify these domains by PCR and used for detection and assessment of their distribution and evolution. The detailed characterization of different plant taxa with respect to the content, variability, and physical distribution of retrotransposons makes a major contribution to our understanding of host genome organization and evolution.
Copia group sequences have been found in diverse species, including single-cell algae, bryophytes, gymnosperms, and angiosperms (Voytas and Ausubel 1988
; Grandbastien et al. 1998
; Flavell, Smith, and Kumar 1992
; Hirochika, Fukuchi, and Kikuchi 1992
; Voytas et al. 1992
; Kamm et al. 1996
; Heslop-Harrison et al. 1997
). Gypsy-like elements have been reported from major taxonomic groups of plants (pine [IFG7; Kossack and Kinlaw 1999
], lily [del1; Smyth et al. 1989
], maize [magellan; Purugganan and Wessler 1994
], tomato [Su and Brown 1997
], pineapple [Tomson, Thomas, and Dietzgen 1998
], rice [Kumekawa et al. 1999
], several angiosperms and gymnosperms [e.g., Brandes et al. 1997
; Suoniemi, Tanskanen, and Schulman 1998
]). It is likely that the detection methods using heterologous primers in polymerase chain reactions (PCRs) are less efficient due to relatively higher sequence heterogeneity among gypsy-like elements (Su and Brown 1997
).
The ubiquity of plant retrotransposons, their extant sequence heterogeneity, and the function including reverse transcription suggest that their major functions were present in the first eukaryotes (see Heslop-Harrison 2000)
, although retrotransposons may have originated after the creation of the first eukaryotes and reached their current wide dispersal by a combination of vertical and horizontal transmission (Kumar and Bennetzen 1999
). An important question in retroelement research centers on the contribution of vertical or horizontal transmission to retroelements' sequence evolution and species dispersion.
In this study, we isolated, cloned, and sequenced part of the RT gene of gypsy-like and copia-like retrotransposons from different species, focusing on the major taxonomic groups of the gymnosperms and comparing them with published sequences. We aimed to reveal the lineages of gypsy- and copia-like retrotransposons in gymnosperms and the relationships between taxonomic groups. We also aimed to characterize the content and distribution of retroelement-related DNA sequences to identify the locations of such sequences within and between chromosomes and examine their chromosomal conservation in the Pinaceae family.
![]() |
Materials and Methods |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
|
PCR Assay for Reverse Transcriptases
Degenerate oligonucleotides for gypsy-like retrotransposons were newly designed by inspection of conserved amino acid sequences in the RT domains of different published gypsy-like retrotransposons (see Results): CyRT1 = MRNATGTGYGTNGAYTAYMG, encoding the peptide RMCVDYR, and GyRT4 = RCAYTTNSWNARYTTNGCR, encoding YAKLSKC, where R = A + G, Y = C + T, M = A + C, S = G + C, W = A + T, and N = A + G + C + T. PCR was carried out in 50 µl including 100200 ng genomic DNA, 50 pmol of each primer, 2 U Taq DNA polymerase, 5 µl buffer (Life Technologies), and 3.5 µl 50 mM MgCl for amplification of copia-like retrotransposons following Flavell et al. (1992)
. PCR products were gel-purified and cloned in pGEM T-Easy vector (Promega).
DNA Sequencing
Clones were amplified by PCR with M13 primers, and 40 ng of the product was used in a 10-µl cycle sequencing reaction with the ABI BigDye Terminator Kit on an ABI 377 DNA sequencer (ABI, Foster City, Calif.). Most clones were sequenced in both strands.
Sequence Analysis
For both the copia and the gypsy data sets, initial sequence alignments and neighbor-joining trees were constructed with CLUSTAL X (Thompson et al. 1997
) and improved manually. Generalized parsimony analyses
were performed with PAUP*, version 3.1, with the branch-and-bound search option, MULPARS, ACCTRAN, TBR branch swapping, and gaps treated as missing. For bootstrap support (Felsenstein 1985
), the same settings as in the initial tree searches were used.
DNA Labeling and Membrane Hybridization
The nonradioactive chemiluminescence method Alk-Phos Direct (Amersham) was used for DNA labeling, hybridization, and detection. Southern blots were prepared using standard protocols (Sambrook, Fritsch, and Maniatis 1989
). Five micrograms, or in some cases 10 µg, of digested genomic DNA (HindIII, HaeIII) from different conifer species was used for each row. Digested fragments were separated on 1.2% agarose gels, blotted, and probed with different clones of Picea abies, Pinus pinaster, and Ginkgo biloba.
In Situ Hybridization
Methods for chromosome preparation and in situ hybridization essentially followed Schwarzacher and Heslop-Harrison (2000)
. Briefly, seedling root tips were placed in ice water overnight, followed by a 3-h pretreatment with 0.05% colchicine prior to fixation in alcohol : acetic acid (3:1). Roots were partially digested with enzymes, and cells were spread on glass slides. Clones were labeled by PCR with biotin or digoxigenin. The clone pTa71, containing rDNA from Triticum aestivum (Gerlach and Bedbrook 1979
), was used for detection of 45S rDNA sites, and clone pTa794, containing a 410-bp BamH1 fragment of the 5S rDNA from of Triticum aestivum (Gerlach and Dyer 1980
), was used to detect 5S rDNA sites. The hybridization mixture (40% formamide, 2 x SSC, 10% dextran sulfate, 1 µg salmon sperm DNA, 0.15% SDS, and about 100 ng of two different labeled probes) was denatured and applied to the slides, and probes and slides were denatured together at 85°C for 5 min. After overnight hybridization at 37°C, slides were washed, with the most stringent wash being at 42°C in 20% formamide, 0.1 x SSC. The hybridization sites were detected using antidigoxigenin conjugated to FITC (Roche) and Cy3 conjugated to streptavidin (Sigma). After detection, the slides were washed, counterstained with DAPI (4',6-diamidino-2-phenylindole), and photographed with appropriate filters on a Nikon epifluorescence microscope. Negatives were scanned and printed from Adobe PhotoShop 5.0 using only contrast and brightness functions affecting the whole image equally.
![]() |
Results |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
Consistent with the spacing of the amino acid sequence domains, the degenerate oligonucleotide primers for gypsy and copia RT domain fragments amplified PCR products of 420 bp and
260 bp, respectively, from the gymnosperm species studied here (table 1 ; 79 clones of gypsy-like and 86 copia-like retrotransposons were isolated from 21 species of gymnosperm). With the copia primers, about 95% of the 260-bp PCR fragments were identified as copia-like retrotransposons by sequence analysis, with occasional products of
420 bp and
600 bp that were gypsy-like and LINE-like, respectively. After cloning of the 420-bp PCR product from the gypsy primers, only half of the clones were homologous to the RT domain of gypsy-like elements. Products of other sizes were also analyzed, and some were retroelement-related: the gypsy primers also amplified
260-bp copia-like products and
600-bp LINE-related fragments. PCR reactions were carried out more than once, and there was evidence that each reaction amplified different subsets of the target sequences: for example, in the amplification of copia-like elements from P. abies, clones within each of the three reactions (pPaty1-pPaty11, pPaty12-pPaty18, and pPaty19-pPaty29) were more similar than those between the reactions, suggesting that the degenerate primers sampled only part of the diversity of the RT genes in each amplification procedure.
Phylogenetic Analysis of Sequences
Gypsy-like Retrotransposons
Many gypsy-like sequences fell into two clades (fig. 1
; clades labeled A and C/B). Pairwise comparisons of the 76 gypsy-like retrotransposon fragments showed nucleotide homologies of 37.7% (Ps11ty/Ar6li) to 99.4% (Psbgy4/Tbgy4) between species, and a similar range of 41.5% (Pagy9/Pagy52) to 99.0% (Pagy14/Pagy16) within species (see EMBL database accessions). More detailed analysis of gymnosperm gypsy-like elements (420 bp) used maximum-parsimony (MP) analysis with sequences from table 1
and the database (accession numbers shown on the tree) and Ty3 as the outgroup, as shown in figure 1
. MP phylogeny provided strong bootstrap support for a monophyletic origin of plant gypsy-like elements but showed high diversity within all species. The tree showed a clade of 32 sequences (fig. 1
, branch B) with 92% homology isolated from 10 divergent gymnosperm species (G. biloba, Araucaria araucana, Taxus baccata, Podocarpus totara and Podocarpus nivalis, Pinus species, and P. abies). The second well-supported clade (branch A) also represented the gymnosperm species, including G. biloba, Taxodium distichum, and most studied species from the family Pinaceae. A larger number of well-supported small clades were not resolved by a strict-consensus tree (fig. 1
), while other major phylogenetic divisions were at best weak.
|
Copia-like Retrotransposons
For the gymnosperm sequences analyzed, within-species similarity ranged from 40.1% (Paty14/Paty23) to 98.1% (Paty24/Paty26) in 24 sequences of copia-like elements from P. abies. Between-species homology ranged from 28.2% (Pitaty2/Pipty1) to 85.8% (Ppty18/Gity4), thus suggesting slightly higher similarity within than between species, in agreement with the suggestion and results obtained by Flavell et al. (1992)
for angiosperms.
The phylogenetic picture emerging from the MP analysis of copia-like element sequences (260 bp, giving lower character numbers and hence lower bootstrap support values than the gypsy tree) is slightly different from that for gypsy elements (fig. 2 ). A monophyletic origin of plant sequences is not supported by bootstrap support values of over 50% with respect to either Ty1 from Saccharomyces or copia from Drosophila melanogaster. However, more clades of two to four sequences from single or related taxa were well supported for the copia-like sequences than for the gypsy-like sequences. As with the gypsy tree, similar sequences were frequently found in divergent taxa: Cajanus cajan (Angiospermae, Dicotyledonae) and Gnetum montanum (Gnetophyta); Dicranum scoparium (Bryophyta) and A. thaliana (Angiospermae; Dicotyledonae); Equisetum arvense (Sphenopsida) and L. esculentum (Angiospermae; Dicotyledonae); and Pinus coulteri (Coniferopsida) and Secale cereale (Angiospermae; Liliopsida) all showed more than 60% bootstrap support in the MP tree. Two conifer-specific branches existed in the MP copia tree. The first clade had 100% bootstrap support (branch A), with sequences only from different species of the genus Pinus, and the second clade had bootstrap support of 65% (branch B) and was divided into two sister subgroups: the first included 31 sequences from 12 Pinaceae species and P. nivalis (Podocarpaceae), with strong support (bootstrap support 100%), and the second included four sequences from another family of conifers (Cryptomeria and Sequoiadendron, Taxodiaceae). As with the gypsy sequences, A. thaliana copia sequences were found throughout the tree. The monophyly of the third supported main gymnosperm branch (branch C) in the MP tree (bootstrap support 85%) was destroyed by inclusion of a sequence from Equisetum scirpoides (Sphenopsida). Branch C was divided into the Equisetum branch and two others, one including sequences from only Pinus, and the second including sequences from G. biloba, P. abies, P. pinaster, and Larix decidua.
|
Six gypsy clones, representative of their diversity in P. abies (Pagy5, Pagy7, Pagy9, Pagy14, Pagy16, and Pa15ty), and one clone (Gigy6) from G. biloba were used for Southern hybridization to DNA digests from Pinus, Picea, and other representative gymnosperms (fig. 3 ).
|
Three copia clones, representative of their diversity in P. abies, were used as probes: Paty5 and Paty14 belong to one clade, and Paty11 belongs to another (fig. 2 ). Paty11 showed a different hybridization pattern from Paty5 and Paty14. All three probes were present in Picea, Pinus, and Abies DNA but showed differences in genomic organization (pattern of bands) and abundance between the genera (fig. 4 ). It is notable that probe Paty11 was abundant in G. biloba but weak in P. nivalis and A. araucana.
|
In Situ Hybridization
rDNA probes were used to assist with chromosome identification by localizing the blocks of repetitive sequences and to provide a comparison with the dispersed signals seen with retroelement probes. On the P. abies (2n = 24) chromosomes, we detected 12 major and 2 minor sites of 45S rDNA, all at intercalary positions (fig. 5A
), and two major intercalary pairs of 5S rDNA sites, with additional pairs of terminal sites (fig. 5I
).
|
The copia probes Paty5, Paty11, and Paty14 (fig. 5G and J ) also showed a dispersed distribution over the chromosomes, but Paty14 showed multiple, more intense, bands in the intercalary regions of many chromosome arms. Regions of weaker hybridization at rDNA sites were also revealed, particularly with Paty11 (fig. 5G ).
After hybridization with a 390-bp probe of 18S rDNA genes from P. pinaster, chromosomes of P. pinaster (2n = 24) showed five chromosome pairs with strong intercalary hybridization signals and one with a weak intercalary hybridization signal, and additional clear strong signals at the centromeric regions of some chromosome pairs (fig. 5L
). With 45S rDNA probe (pTa71), hybridization to the centromeric regions was very weak (data not shown). Hizume, Ishida, and Murata (1992)
and Lubaretz et al. (1996)
also reported weak hybridization signals with the rDNA probe to the centromeric regions of Pinus thunbergii and Pinus sylvestris, respectively. The gypsy-like element Ppgy1, which is isolated in the trees (e.g., fig. 1B
), is collocalized with the 18S rDNA probe, with additional signals on the other parts of the chromosomes, including centromeric regions (fig. 5M and N
). The other gypsy-like retroelements (Ppgy3 and Ppgy5) are dispersed on all chromosomes (data not shown).
![]() |
Discussion |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
Although the sequencing of retroelements from different species is far from complete and is largely based on PCR amplification using at least partially selective primers, there are enough data that useful phylogenetic inferences can be made. For examination of the relationships between all groups of retroelements, an unrooted neighbor-joining tree was used with fragments of the RT gene for 26 gypsy-like and badnaviral retroelements, 28 copia-like elements, and three LINEs (fig. 6
). The sequences included retroelement fragments from gymnosperms (selected to sample the diversity present based on the trees; figs. 1 and 2
) and published retroelement sequences, including gypsy and copia from D. melanogaster and Ty1 and Ty3 from S. cerevisiae. The alignment of individual RT domains, each typically 260 bp long, spanned only 278 bp, showing relatively high similarity. The tree supported the monophyletic origin of the copia and LINE clades. At the base of the gypsy clade, there was a grouping including banana streak badnavirus (BSV), two retroelements from legumes (CyclopsPisum sativum; broad bean elementVicia faba), gypsy from D. melanogaster, and Ty3. Most of the plant gypsy elements also showed a monophyletic origins, and the Ty3 retrotransposon from S. cerevisiae is the next relative to the plant gypsy lineage. The broad bean and pea Cyclops retroelements (also placed in an anomalous position by Miller et al. [1999
] and Chavanne et al. ]1998
[), along with the BSV sequence (Harper et al. 1999), are an exception, representing another lineage of retroelements.
|
Among the plant branches, there were several supported groupings representing only gymnosperms or conifers. The hybridization of genomic DNA from three species to clones representing many gymnosperm retrotransposons supported the suggestion that all species have similar diversities of retrotransposons but major differences in copy number.
The evidence from the trees allows us to explore support for three suggestions, not mutually exclusive, about the evolution of the RT domain of retroelements in gymnosperms and plants. First, there may have been an explosive radiation of retroelements of each group in the common ancestor of all plants, with most of these families having been maintained with only limited further divergence in sequence in the subsequent 350 Myr. Second, multiple horizontal transfers of the retroelements may have occurred between physically and phylogenetically distant populations or taxa. Finally, the constraints on retroelement evolution may be such that retroelements have reached near-identical ranges of sequence diversity in widespread modern plant taxa (convergent evolution) which are different from the sequence diversity ranges in the animal and fungal kingdoms. The contribution of the three models can also be examined within the gymnosperm-specific elements: were the retrotransposon families present in all gymnosperms before their divergence, becoming differentially amplified in different genomes? Or are the families showing horizontal transmission between species?
The monophyletic grouping of all types of retroelements in figure 6 and the monophyletic grouping of some families of gymnosperm-specific clones in figures 1 and 2
, taken with the hybridization results, give some evidence that common ancestry and rapid radiation are the major factors in the current diversity of retroelements. Perhaps the stress events during early evolution of the kingdoms led to widespread activation of retroelements, known to be a stress response (Grandbastien 1998
). As the conserved regions are over 3040 amino acids long and many elements are nonfunctional due to the presence of stop codons and frameshifts, it is unlikely that similarities are due to convergence of function and then of sequence. Such a phenomenon may explain the similarities between two sequences, but it cannot explain the similarities between several sequences belonging to different superfamilies and classes of elements (Capy et al. 1997
). If the model of early generation of the different RT domains of the retroelement families is correct, then differential amplification (and perhaps loss) of different families, without lineage-specific diversification, must have occurred. However, given that most domains are apparently nonfunctional, it is unclear why there should be this lack of diversification of the sequences (fig. 2
, e.g., branch B).
There are widespread viruses and biting insects that could be responsible for the horizontal transfer of retroelements between plants. There is good evidence for horizontal transfer of gypsy in Drosophila (Robertson and Lampe 1995
; Jordan, Matyunina, and McDonald 1999
; Terzian et al. 2000
), and the model of widespread horizontal transmission of elements between plants (Hirochika and Hirochika 1993
), which would lead to the homogeneous population structure seen in large plant groupings, has been considered. But the clear monophyletic structure of the trees suggests that cross-kingdom transfer has not occurred: perhaps barriers such as methylation, codon usage, specificity of polymerases and other enzymes, or integration site specificity act to prevent such transfers.
Determination of the phylogeny of any group of retroelements comparison with the phylogeny of their hosts would suggest the validity of either model: if they were identical, then the explosive evolution model would be most likely (Eickbush 1994
). However, it would be impossible to obtain unequivocal phylogenies because of the difficulty in obtaining and distinguishing members of a retroelement lineage and in determining an independent plant phylogeny. Nevertheless, there is a possibility that Ginkgo is more closely related to the family Pinaceae than is normally recognized.
The data presented here add to our knowledge of genome structure of the Pinaceae, allowing construction of molecular karyotypes of Pinus and Picea and integrating information about major classes of repetitive DNA sequences with the morphology of chromosomes. The rDNA sequences show chromosome-specific distribution patterns and hence allow identification of individual chromosomes, as was found in Pinus elliottii (Doudrick et al. 1995
). The information about chromosome identification will be useful for the integration of genetic and physical maps and for comparative analysis of conserved synteny between the various conifer species as detailed maps of agronomic traits become available. The sequence data from the complete A. thaliana genome, with retroelement sequences distributed widely in our trees, shows the importance of nonselective genomic sequencing and how it can be used in understanding genome evolution in diverse species. The addition of information about biodiversity present within the retroelements, as one of the most abundant genomic components, is a valuable addition to data about gymnosperm genomes, in which molecular diversity at the DNA level seems more limited than in angiosperms, and adds another tool to those available to reconstruct the evolutionary history of the gymnosperms and the plant kingdom.
![]() |
Supplementary Material |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
|
![]() |
Acknowledgements |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
![]() |
Footnotes |
---|
1 Abbreviations: LTR, long terminal repeat; MP, maximum parsimony; RT, reverse transcriptase.
2 Keywords: Picea abies,
Pinus
pine
spruce
retroelements
biodiversity
evolution
genome organization
gymnosperms
phylogenetics
3 Address for correspondence and reprints: Pat Heslop-Harrison, Department of Biology, University of Leicester, Leicester LE1 7RH, United Kingdom. E-mail: phh4{at}le.ac.uk
![]() |
References |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
Bennetzen J. L., 1996 The contributions of retroelements to plant genome organisation, function and evolution Trends Microbiol 4:347-353[ISI][Medline]
. 2000 Transposable element contributions to plant gene and genome evolution Plant Mol. Biol 42:251-269[ISI][Medline]
Brandes A., J. S. Heslop-Harrison, A. Kamm, R. L. Doudrick, T. Schmidt, 1997 Comparative analysis of the chromosomal and genomic organization of Ty1-copia-like retrotransposons in pteridophytes, gymnosperms and angiosperm Plant Mol. Biol 33:11-21[ISI][Medline]
Brown G. R., V. Amarasinghe, G. Kiss, J. E. Carlson, 1993 Preliminary karyotype and chromosomal location of rDNA sites in white spruce using fluorescence in situ hybridization Genome 36:310-316.[ISI]
Brown G. R., J. E. Carlson, 1997 Molecular cytogenetics of the genes encoding 18s5. 8s26s rRNA and 5s rRNA in two species of spruce Picea. Theor. Appl. Genet 95:1-9[ISI]
Capy P., T. Langin, D. Higuet, P. Maurer, C. Bazin, 1997 Do the integrases of LTR-retrotransposons and class II element transposases have a common ancestor? Genetica 100:63-72[ISI][Medline]
Chavanne F., D. X. Zhang, M. F. Liaud, R. Cerff, 1998 Structure and evolution of Cyclops: a novel giant retrotransposon of the Ty3/Gypsy family highly amplified in pea and other legume species Plant Mol. Evol 37:363-375
Doudrick R. L., J. S. Heslop-Harrison, C. D. Nelson, T. Schmidt, W. L. Nanse, T. Schwarzacher, 1995 Karyotyping slash pine Pinus elliottii var. elliottii using patterns of fluorescence in situ hybridization and fluorochrome banding J. Hered 86:289-296[ISI]
Eickbush T., 1994 Origin and evolutionary relationships of retroelements Pp. 121157 in S. Morse, ed. The evolutionary biology of viruses. Raven Press, New York
Felsenstein J., 1985 Confidence limits on phylogenies: an approach using the bootstrap Evolution 39:783-791[ISI]
Flavell A. J., 1992 Ty1-copia group retrotransposons and the evolution of retroelements in eukaryotes Genetica 86:203-214[ISI][Medline]
Flavell A. J., E. Dunbar, R. Anderson, S. R. Pearce, R. Hartley, A. Kumar, 1992 Ty1-copia group retrotransposons are ubiquitous and heterogeneous in higher plants Nucleic Acids Res 20:3639-3644[Abstract]
Flavell A. J., D. B. Smith, A. Kumar, 1992 Extreme heterogeneity of Ty1-copia group retrotransposons in plants Mol. Gen. Genet 231:233-242[ISI][Medline]
Gerlach W. L., J. R. Bedbrook, 1979 Cloning and characterization of ribosomal RNA genes from wheat and barley Nucleic Acids Res 7:1869-1885[Abstract]
Gerlach W. L., T. A. Dyer, 1980 Sequence organization of the repeated units in the nucleus of wheat which contain 5s-rRNA genes Nucleic Acids Res 8:4851-4865[Abstract]
Grandbastien M. A., 1998 Activation of plant retrotransposons under stress conditions Trends Plant Sci 3:181-187[ISI]
Harper G., J. O. Osuji, J. S. Heslop-Harrison, R. Hull, 1999 Integration of banana streak badnavirus into the Musa genome: molecular and cytogenetic evidence Virology 255:207-213[ISI][Medline]
Heslop-Harrison J. S., 2000 RNA, genes, genomes and chromosomes: repetitive DNA sequences in plants Chromosomes Today 13:4556 + plate VI
Heslop-Harrison J. S., A. Brandes, S. Taketa, et al. (15 co-authors) 1997 The chromosomal distribution of Ty1-copia group retrotransposable elements in higher plants and their implications for genome evolution Genetica 100:197-204[ISI][Medline]
Hirochika H., A. Fukuchi, F. Kikuchi, 1992 Retrotransposon families in rice Mol. Gen. Genet 233:209-216[ISI][Medline]
Hirochika H., R. Hirochika, 1993 Ty1-copia group retrotransposons as ubiquitous components of plant genomes Jpn. J. Genet 68:35-46[Medline]
Hizume M., F. Ishida, M. Murata, 1992 Multiple location of the rRNA genes in chromosomes of pines, Pinus densiflora and P. thunbergii. Jpn. J. Genet 67:389-396
Hizume M., H. H. Tominaga, K. Kondo, Z. Gu, Z. Yue, 1993 Fluorescent chromosome banding in six taxa of Eurasian Larix. Pinaceae. Khromosomo 69:2342-2354
Jordan I. K., L. V. Matyunina, J. F. McDonald, 1999 Evidence for the recent horizontal transfer of long terminal repeat retrotransposon Proc. Natl. Acad. Sci. USA 26:12621-12625
Kamm A., R. L. Doudrick, J. S. Heslop-Harrison, T. Schmidt, 1996 The genomic and physical organization of Ty1-copia-like sequences as a component of large genomes in Pinus elliottii var. elliotii and other gymnosperms Proc. Natl. Acad. Sci. USA 93:2708-2713
Khoshoo T. N., 1959 Polyploidy in gymnosperms Evolution 13:24-39[ISI]
. 1961 Chromosome numbers in gymnosperms Silvae Genet 1:1-9
Kossack D. S., C. S. Kinlaw, 1999 IFG, a gypsy-like retrotransposon in Pinus, Pinaceae, has an extensive history in pines Plant Mol. Biol 39:417-426[ISI][Medline]
Kumar A., J. L. Bennetzen, 1999 Plant retrotransposons Annu. Rev. Genet 33:479-532[ISI][Medline]
Kumekawa N., H. Ohtsubo, T. Horiuchi, E. Ohtsubo, 1999 Identification and characterization of novel retrotransposons of the gypsy type in rice Mol. Gen. Genet 260:593-602[ISI][Medline]
Le Q. H., S. Wright, Z. Yu, T. Bureau, 2000 Transposon diversity in Arabidopsis thaliana Proc. Natl. Acad. Sci. USA 97:7376-7381
Lubaretz O., J. Fuchs, R. Ahne, A. Meister, I. Schubert, 1996 Karyotyping of three Pinaceae species via fluorescent in situ hybridization and computer-aided chromosome analysis Theor. Appl. Genet 92:411-416[ISI]
Matsuoka Y., K. Tsunewaki, 1996 Wheat retrotransposon family identified by reverse transcriptase domain analysis Mol. Biol. Evol 13:1384-1392
. 1999 Evolutionary dynamics of Ty1-copia group retrotransposons in grass shown by reverse transcriptase domain analysis Mol. Biol. Evol 16:208-217[Abstract]
Miller K., C. Lynch, J. Martin, E. Herniou, M. Tristem, 1999 Identification of multiple gypsy LTR-retrotransposon lineages in vertebrate genomes J. Mol. Evol 49:358-366[ISI][Medline]
Murray B. G., 1998 Nuclear DNA amount in gymnosperms Ann. Bot 82: (Suppl. A) 3-15
Pearce S. R., G. Harrison, D. Li, J. S. Heslop-Harrison, A. Kumar, A. J. Flavell, 1996 The Ty1-copia group retrotransposons in Vicia species: copy number, sequence heterogeneity and chromosomal localisation Mol. Gen. Genet 25:305-315
Presting G. G., L. Malysheva, J. Fuchs, I. Schubert, 1998 A Ty3/gypsy retrotransposon-like sequence localizes to the centromeric regions of cereal chromosomes Plant J 16:721-728[ISI][Medline]
Purugganan M. D., S. R. Wessler, 1994 Molecular evolution of magellan, a maize Ty3/gypsy-like retrotransposon Proc. Natl. Acad. Sci. USA 91:11674-11678
Robertson H. M., D. J. Lampe, 1995 Recent horizontal transfer of a mariner transposable element among and between Diptera and Neuroptera. Mol. Biol. Evol 12:850-862[Abstract]
Sambrook J., E. F. Fritsch, T. Maniatis, 1989 Molecular cloning: a laboratory manual Cold Spring Harbor Laboratory Press, Cold Spring Harbor, N.Y
SanMiguel P., B. S. Gaut, A. Tikhonov, Y. Nakajiama, J. Bennetzen, 1998 The paleontology of intergene retrotransposons of maize Nat. Gen 20:43-45[ISI][Medline]
SanMiguel P., A. Thikhonov, Y. K. Jin, et al. (11 co-authors) 1996 Nested retrotransposons in the intergenic regions of the maize genome Science 274:765-768
Schmidt A., R. L. Doudrick, J. S. Heslop-Harrison, T. Schmidt, 2000 The contribution of short repeats of low sequence complexity to large conifer genomes Theor. Appl. Genet 101:7-14[ISI]
Schmidt T., 1999 LINEs, SINEs and repetitive DNA: non-LTR retrotransposons in plant genomes Plant Mol. Biol 4:903-910
Schwarzacher T., J. S. Heslop-Harrison, 2000 Practical in situ hybridization Bios, Oxford, England
Scotti I., F. Magni, R. Fink, W. Powell, G. Binelli, P. E. Hedley, 2000 Microsatellite repeats are not randomly distributed within Norway spruce (Picea abies K.) expressed sequences Genome 43:41-46[ISI][Medline]
Sederoff R. R., A. M. Stomp, 1993 DNA transfer to conifers Pp. 241254 in M. R. Ahuja and W. J. Libby, eds. Clonal forestry. I. Genetics and biotechnology. Springer Verlag, Berlin/Heidelberg
Smyth D. R., P. Kalitsis, J. L. Joseph, J. W. Sentry, 1989 Plant retrotransposon from Lilium henryi is related to Ty3 of yeast and the gypsy group of Drosophila. Proc. Natl. Acad. Sci. USA 86:515-519[ISI]
Su P. Y., T. A. Brown, 1997 Ty3/gypsy-like retrotransposon sequences in tomato Plasmid 38:148-157[ISI][Medline]
Suoniemi A., J. Tanskanen, A. H. Schulman, 1998 Gypsy-like retrotransposons are widespread in the plant kingdom Plant J 13:699-705[ISI][Medline]
Swofford D. L., 1991 PAUP: phylogenetic analysis using parsimony Version 3.1.1. Smithsonian Institution, Washington, D.C
Terzian C., C. Ferraz, J. Demaille, A. Bucheton, 2000 Evolution of the gypsy endogenous retrovirus in the Drosophila melanogaster subgroup Mol. Biol. Evol 17:908-914
Thompson J. D., Y. J. Gibson, F. Plewniak, F. Jeanmougin, D. G. Higgins, 1997 The CLUSTAL-X windows interface: flexible strategies for multiple sequence alignment aided by quality analysis tools Nucleic Acids Res 25:4876-4882
Tomson K. G., J. E. Thomas, R. G. Dietzgen, 1998 Retrotransposon-like sequence integrated into the genome of pineapple, Ananas comosus. Plant Mol. Biol 38:461-465[ISI][Medline]
Vernhettes S., M. A. Grandbastien, J. M. Casacuberta, 1998 The evolutionary analysis of the Tnt1 retrotransposon in Nicotiana species reveals the high variability of its regulatory sequences Mol. Biol. Evol 15:827-836[Abstract]
Voytas D. F., F. M. Ausubel, 1988 A copia-like transposable element family in Arabidopsis thaliana. Nature 336:242-244[ISI][Medline]
Voytas D. F., M. P. Cummings A. Konieczny, F. M. Ausubel, S. Rodermel, 1992 Copia-like retrotransposons are ubiquitous among plants Proc. Natl. Acad. Sci. USA 89:7124-7128[Abstract]
Xiong Y., T. H. Eickbush, 1990 Origin and evolution of retroelements based on their reverse transcriptase sequences EMBO J 9:3353-3362[Abstract]