You are on page 1of 14

Research

Asexual reproduction in a close relative of Arabidopsis:


Blackwell Publishing Ltd

a genetic investigation of apomixis in Boechera


(Brassicaceae)
M. Eric Schranz1,*, Laksana Kantama2, Hans de Jong2 and Thomas Mitchell-Olds1
1
Department of Genetics and Evolution, Max Planck Institute for Chemical Ecology, Hans Knoll Strasse 6, D-07745 Jena, Germany; 2Laboratories of
Biochemistry and Genetics, Wageningen University, Arboretumlaan 4, NL-6703 BD Wageningen, the Netherlands; *Present address: Department of Biology,
Box 91000, Duke University, Durham, NC 27708, USA

Summary

Author for correspondence: • Understanding apomixis (asexual reproduction through seeds) is of great interest
M. Eric Schranz to both plant breeders and evolutionary biologists. The genus Boechera is an excellent
Tel: +1 (919) 613 8181 system for studying apomixis because of its close relationship to Arabidopsis, the
Fax: +1 (919) 613 8177
Email: eric.schranz@duke.edu
occurrence of apomixis at the diploid level, and its potentially simple inheritance by
transmission of a heterochromatic (Het) chromosome.
Received: 3 March 2006
• Diploid sexual Boechera stricta and diploid apomictic Boechera divaricarpa
Accepted: 24 March 2006
(carrying a Het chromosome) were crossed. Flow cytometry, karyotype analysis,
genomic in situ hybridization, pollen staining and seed-production measurements
were used to analyse the parents and resulting F1, F2 and selected F3 and test-cross
(TC) generations.
• The F1 plant was a low-fertility triploid that produced a swarm of aneuploid and
polyploid F2 progeny. Two of the F2 plants were fertile near-tetraploids, and analysis of
their F3 and TC progeny revealed that they were sexual and genomically stabilized.
• The apomictic phenotype was not transmitted by genetic crossing as a single
dominant locus on the Het chromosome, suggesting a complex genetic control of
apomixis that has implications for future genetic and evolutionary analyses in this group.

Key words: apomixis, Boechera (Brassicaceae), interspecific hybridization, polyploidy,


genome painting, heterochromatic chromosome.

New Phytologist (2006) 171: 425–438

© The Authors (2006). Journal compilation © New Phytologist (2006)


doi: 10.1111/j.1469-8137.2006.01765.x

directly derived from normal sexual pathways (Nogler, 1984a;


Introduction Holsinger, 2000; Koltunow & Grossniklaus, 2003). This
The study of apomixis (asexual reproduction through seeds) bypassing of normal sexuality may have resulted from gene-
is of great interest for both practical and theoretical reasons. expression changes during polyploidization and/or hybridization,
The harnessing of apomixis for agricultural purposes could be as most apomictic species are allopolyploids (Carman, 1997).
of great economic and humanitarian benefit, because it would A number of model species have been well developed for the
enable the propagation of hybrid genotypes indefinitely study of apomixis, each with their own advantages and disad-
(Spillane et al., 2004). For evolutionary biologists, apomictic vantages (Bicknell & Koltunow, 2004). Genetic studies with
systems allow for testing of hypotheses of the persistence of sexual these systems suggest that the trait is under the control of either
reproduction despite its twofold cost (Charlesworth, 1990; two or several genes (van Dijk et al., 1999; Noyes & Rieseberg,
West et al., 1999). Although biologists have long grappled with 2000; Albertini et al., 2001; Matzk et al., 2005), or by a single
apomictic mechanisms, their control is still relatively poorly dominant locus (Leblanc et al., 1995; Bicknell et al., 2000),
understood at the genetic and molecular levels (Bicknell & which can be located on supernumerary or hemizygous chromatin
Koltunow, 2004). It has been hypothesized that apomixis is (Ozias-Akins et al., 1998; Roche et al., 2001; Sharbel et al., 2004).

www.newphytologist.org 425
426 Research

In comparison with other apomictic complexes, apomictic system. Many of the major advances in our understanding of
reproduction in Boechera has several factors making it more the genetic control of apomixis have come from analysis of the
akin to normal sexual reproduction. These factors combine to progeny from such crosses (Ozias-Akins et al., 1998; van Dijk
make Boechera a compelling and tractable system for genetic et al., 1999; Noyes & Rieseberg, 2000). However, there are
studies of the control of apomixis. Perhaps the most promis- several important caveats for analysis of the inheritance of
ing and rare characteristic of Boechera is that apomixis can apomixis from such studies, including the problems of making
occur at the diploid level (Böcher, 1951, 1969; Sharbel & interploidy and/or interspecific crosses (discussed by Bicknell
Mitchell-Olds, 2001). Evidence of diploid apomictic Boechera & Koltunow, 2004; Matzk et al., 2005). In Boechera we can
lines has been gathered through cytological investigations partially mitigate these complications. The presence of diploid
(Böcher, 1951; Naumova et al., 2001); the occurrence of unre- apomicts means that we can avoid making interploidy crosses
duced gametes and heterozygous genotypes (DobeS et al., 2004b; (Schranz et al., 2005). Also, a robust phylogenetic hypothesis
Sharbel et al., 2004, 2005); and the persistence of heterozygosity exists for the relationship of sexual and apomictic lineages,
in progeny (Roy, 1995; Schranz et al., 2005). Additionally, these with direct evidence of past hybridization events (Sharbel &
diploid lines often contain heterochromatic (Het) and/or Mitchell-Olds, 2001; Koch et al., 2003; DobeS et al., 2004a,
supernumerary chromosomes that could be responsible for 2004b; Schranz et al., 2005). Specifically, the sexual species
the apomictic phenotype (Böcher, 1954; Sharbel et al., 2004; Boechera stricta is almost exclusively diploid, sexual, and forms
Kantama, 2005; Sharbel et al., 2005). Another major a well supported monophyletic clade (lineage II of DobeS
advantage is the close relationship of Boechera to the model et al., 2004b; Schranz et al., 2005). Boechera holboellii is
plant Arabidopsis thaliana (Koch et al., 2001). Boechera is diploid or triploid, often apomictic (Sharbel et al., 2004,
the only documented case of natural apomixis in the Brassica 2005), paraphyletic (I. A. Al-Shehbaz, pers. comm.), and is
family (Brassicaceae). In addition to A. thaliana having its scattered between two other lineages with other species of Boechera
complete genome sequenced, there is also an unrivalled (lineages I and III of DobeS et al., 2004b; Schranz et al., 2005).
understanding of normal sexual reproduction and a variety of The interspecific hybrid species Boechera divaricarpa has repeated
apomixis-like mutations that have been identified (Koltunow independent origins by hybridization of B. stricta and B. holboellii-
& Grossniklaus, 2003), which should provide an excellent like plants (Koch et al., 2003; DobeS et al., 2004a, 2004b). These
framework for comparison. B. divaricarpa lineages are diploid or triploid, and are often
Several aspects of the Boechera breeding system simplify genetic apomictic (Schranz et al., 2005). Thus apomictic lineages of
and molecular investigations, compared with other apomictic B. divaricarpa are known to have a mixed dosage of B. stricta
lineages. First, apomictic Boechera accessions have probably and B. holboellii-like chromosomes. By making crosses between
evolved from self-compatible and highly self-fertilizing sexual diploid B. stricta and B. divaricarpa, we are mirroring the evo-
types (Roy, 1995), whereas most other apomictic groups are lutionary steps that have probably happened within the group.
derived from self-incompatible and out-crossing taxa (Asker We reported previously on reciprocal crosses between a
& Jerling, 1992). Second, as in normal sexual taxa, the embryo sexual B. stricta line (SAD12 = ES6) and a number of diploid
is derived from the megaspore mother cell (MMC). The MMC apomictic B. divaricarpa lines, including individuals from the
generally enters meiosis I, but fails to complete the reductional Vipond Park site (VP9 = ES9) (Schranz et al., 2005). Analysis
phase (apomeiosis), and then undergoes normal meiosis of the reciprocal crosses of B. stricta SAD12 and B. divaricarpa
II to form a nonreduced restitution nucleus (Taraxacum-type VP9 found that all F1 progeny were diploid and apomictically
diplospory) (Böcher, 1951; Naumova et al., 2001; Taskin derived when VP9 was the maternal parent, whereas all F1
et al., 2004). In other forms of apomixis, the embryo either progeny were triploid and sexually derived when SAD12 was
develops from the nucellus (apospory) or else forgoes meiosis the maternal parent (Schranz et al., 2005).
(Antennaria-type diplospory) (Crane, 2001). Third, Boechera In this work, we continue our genetic analysis of these
apomicts are pseudogamous, meaning that fertilization of the genotypes. First, we provide additional information about the
central cell is still required for normal endosperm development diploid apomictic parent (VP9), including analysis of male
(Böcher, 1951; Naumova et al., 2001; Taskin et al., 2004). meiosis, pollen analysis showing nonreduced gamete production,
Pseudogamy is usually only found in conjunction with apospory cytology documenting the presence of a heterochromatic (Het)
(Richards, 1986; Asker & Jerling, 1992), and typically occurs chromosome, microsatellite analysis of progeny derived from
in lineages that descend from outbreeding taxa (Mogie, 1992). self-pollination demonstrating persistent heterozygosity, and
Finally, apomixis in Boechera is incomplete (facultative apomixis), genomic in situ hybridization (GISH) analysis showing its
with sexual reproduction still occurring at an often high frequency hybrid origin. From the cross we analysed the F1, F2 progeny,
(Böcher, 1951; Schranz et al., 2005). Both the formation of selected F3 and test-cross (TC) progeny using a combination
viable pollen and facultative apomixis allow for potential of flow cytometry, karyotype analysis, genome painting (GISH),
hybridization with sexual Boechera lineages. and seed and pollen grain analyses. In particular, we were interested
The ability to cross sexual and asexual lineages of Boechera is in testing the hypothesis that the Het chromosome would act
critical for the establishment of a genetically tractable research as a dominant locus and confer the apomictic phenotype

New Phytologist (2006) 171: 425–438 www.newphytologist.org © The Authors (2006). Journal compilation © New Phytologist (2006)
Research 427

on transmission. Considering Boechera’s great potential for a controlled growth room under long-day conditions (16 h light,
studies of apomixis, this study serves as a first genetic 8 h dark). Days to flowering were measured as the number of days
investigation into the control and transmission of apomixis in between seedling planting until the appearance of the first open
the group, and has important consequences for our understand- flower. The ES138 line required a 6-wk vernalization treatment
ing of the evolutionary history of apomixis and hybridization. at 4°C to induce flowering. While we were successful in using ES
138 for genetic crossing, the plant perished before we collected
measurements on the production of pollen or selfed seed.
Materials and Methods

Plant materials Chromosome preparation


Analyses of reciprocal crosses made between a common sexual Fast-growing root tips were collected between 09:00 and 10:00 h
diploid B. stricta (Graham) Al-Shehbaz tester (SAD12) and a and incubated in a 2-mM aqueous solution of 8-hydroxyquinoline
number of Boechera lines, including apomictic diploid B. at 15°C for 3 h, then fixed in ethanol/acetic acid (3 : 1) at
divaricarpa (A. Nelson) A. Löve & D. Löve lines from the 4°C. Meiotic preparations were made of spread meiotic cells
Vipond Park (VP9) site, have been described (Schranz et al., from inflorescences with anthers containing pollen mother
2005). In this study we have characterized the SAD12 × VP9 cells at meiosis. The antheres were collected and fixed in acetic
cross in greater detail (Table 1; Fig. 1). We analysed the acid : ethanol (1 : 3). To make the microscope slides, we rinsed
parental genotypes [SAD12 (ES 6) and VP9 (ES 9)]; two of the material three times in water and once in 10 mM citrate
the resulting F1 progeny (ES 116 and ES 117); 20 of their F2 buffer pH 4.5, then transferred the material to a pectolytic
progeny (including lines ES 136 and ES 137); 48 and 30 F3 enzyme mixture (0.3% pectolyase, 0.3% cellulase RS,
progeny from lines ES 136 and ES 137, respectively; and test 0.3% cytohelicase in 10 mM citrate buffer) for 2 h (meiotic
crosses involving ES 136 and ES 137 (Table 1; Fig. 1). material, 2.5 h) at 37°C. The root-tip or anther tissues were
For the test crosses, the ES 136 and ES 137 lines were used washed again, and with fine needles we dissected the root tip
as maternal parents, and a naturally occurring tetraploid B. stricta meristem in a small drop of 60% acetic acid on the slide, while
(ES 138) as the paternal pollen-donating parent. The ES 138 heating carefully on a hot plate at 45°C for 60–90 s. The
line of B. stricta was collected from the Quebec–Labrador material was spread on the microscopic slide by dropping approx.
border in Canada, on the north shore of the Strait of Belle 2 ml freshly prepared ethanol : acetic acid (3 : 1) onto the cells
Isle, by John E. Maunder and Nathalie Djan-Chékar of the in the acetic acid solution, followed by a brief dehydration in
Provincial Museum of Newfoundland (collection number NDC ethanol 98% and air-drying. We selected the best chromosome
99–348, Stn NDC Bor. 3; accession number NFM 3363). spread preparations under the phase-contrast microscope.
All chromosome preparations were counterstained with
20 µl 0.2 µg ml−1 4′,6-diamidine-2-phenylindole (DAPI)
Plant growth conditions
in Vectashield Mounting Medium (Vector Laboratories,
Seeds were germinated as described by Schranz et al. (2005) and Burlingame, CA, USA). Selected mitotic cell complements were
transplanted into 11 × 11 × 13-cm pots. Plants were grown in captured for quantification and karyotype analysis. Lengths of
chromosome arms, heterochromatic regions and secondary
constrictions (nucleolus organiser regions, NORs) were established
with the measuring tool of Adobe PHOTOSHOP, and statistical
analyses were performed in Microsoft EXCEL. For karyotype
analysis, we cut out individual chromosomes digitally, ordering
the short arm upwards and matching chromosomes where
possible on the basis of length, centromere position, hetero-
chromatin pattern and presence of satellites/NORs.

Genome painting
We isolated total genomic DNA from the diploid sexuals
B. holboellii (Hornemann) A. Löve & D. Löve, BH208; B. stricta,
BS2; and A. thaliana (accession Columbia) with the Nucleon
Phytopure extraction kit (Amersham Biosciences, Amersham,
UK). DNA was labelled with either digoxigenin-11-dUTP or
biotin-16-dUTP using the nick-translation kit of Roche (Roche
Fig. 1 Pedigree and development of the Boechera SAD 12 × VP9 Diagnostics GmbH, Roche Applied Science, Mannheim,
population. Germany). In the prehybridization step, we first dried the

© The Authors (2006). Journal compilation © New Phytologist (2006) www.newphytologist.org New Phytologist (2006) 171: 425–438
428 Research

Table 1 Information on plants from the Boechera SAD12 × VP9 population, including generation, line, chromosome number, genome size, days
to flowering, seed set and pollen viability

Chromosome Genome Days to Seeds Pollen


Generation Line number sizea floweringb per fruitc viability (%)d

Parental ES6 14 0.47 288 108.48 98.99


Parental ES9 14 0.50 122 77.56 72.82
F1 ES117 21 0.74 116 0.14 24.52
F1 ES116 21 0.75 101 nd
F2 ES116 #1 nd 0.67 75 nd
F2 ES116 #2 24 0.82 199 0.02
F2 ES117 #1 22 0.80 83 0.02
F2 ES117 #2 23 0.80 291 0.00 40.00
F2 ES117 #3 = ES136 27 0.96 102 10.66 75.88
F2 ES117 #4 24 0.86 129 0.16
F2 ES117 #5 nd 0.72 113 0.62 16.79
F2 ES117 #6 21 0.75 135 0.04 3.43
F2 ES117 #7 28 1.02 139 1.86
F2 ES117 #8 23 0.79 nd nd
F2 ES117 #9 mosaic, 24, 27 1.12 94 0.02
F2 ES117 #10 = ES137 27 0.98 91 9.53 72.81
F2 ES117 #11 nd 0.79 182 0.14
F2 ES117 #12 nd 0.88 206 nd
F2 ES117 #13 nd 0.67 99 0.00
F2 ES117 #14 22 0.76 105 0.04
F2 ES117 #15 21 0.74 101 0.00 1.87
F2 ES117 #16 21 0.73 200 2.66 33.21
F2 ES117 #17 27 0.99 81 0.06
F2 ES117 #19 nd 0.91 92 0.02
F3 ES 136 #3 nd 1.00 150 7.82
F3 ES 136 #9 nd 0.97 113 17.00 70.79
F3 ES 136 #10 nd 0.97 144 8.02 76.41
F3 ES 136 #16 nd 1.01 81 19.62 67.23
F3 ES 136 #24 nd 0.99 84 12.14 64.28
F3 ES 136 #30 nd 0.99 140 46.88 75.58
F3 ES 136 #34 nd 0.99 240 14.84 80.51
F3 ES 137 #9 nd 1.03 145 14.42 71.61
F3 ES 137 #10 nd 0.98 98 5.65 54.66
F3 ES 137 #14 nd 1.04 73 10.13 66.18
F3 ES 137 #15 nd 0.99 112 18.30 68.76
F3 ES 137 #16 nd 0.99 96 6.55 73.09
Tester ES 138 nd 1.01 No floweringe nd
Test cross ES 136 × ES 138 nd 0.95 No floweringf nd
Test cross ES 136 × ES 138 nd 0.97 No floweringf nd
Test cross ES 136 × ES 138 nd 0.99 No floweringf nd
Test cross ES 136 × ES 138 nd 1.01 No floweringf nd
Test cross ES 136 × ES 138 nd 1.03 No floweringf nd
Test cross ES 136 × ES 138 nd 0.99 No floweringf nd
Test cross ES 136 × ES 138 nd 0.92 No floweringf nd
Test cross ES 136 × ES 138 nd 1.04 No floweringf nd
Test cross ES 137 × ES 138 nd 0.96 No floweringf nd
Test cross ES 137 × ES 138 nd 0.97 No floweringf nd

a
Genome size presented as ratio compared with Brassica rapa standard.
b
Days to flowering of nonvernalized plants measured from transfer of seedlings to soil until appearance of first open flower.
c
Seed set per silique calculated by averaging number of seeds from five independent replicates of 10 siliques each.
d
Pollen viability was assessed for selected lines by staining with Alexander’s stain.
e
The tester line did not flower unless vernalized, so days to flowering, seed set and pollen viability were not assessed.
f
The test-cross lines did not flower following 300 d after transplant, so days to flowering, seed set and pollen viability were not assessed.

New Phytologist (2006) 171: 425–438 www.newphytologist.org © The Authors (2006). Journal compilation © New Phytologist (2006)
Research 429

microscopic preparations at 67°C for 30 min before incubation ICE14, H23) that were identified as being heterozygous
with 1 µg ml−1 RNAse-A in 2 × SSC at 37°C for 1 h, and two with two alleles each (bi-allelic) in the VP9 genotype (Schranz
wash steps in 2 × SSC for 5 min at 20°C. The preparations et al., 2005). Based on sequence similarity to Arabidopsis and
were rinsed in 10 mM HCl for 2 min and then treated with preliminary genetic mapping results (M.E.S. and T.M.O.,
100 µl pepsin (5 µg ml−1) in 10 mM HCl for 5 min at 37°C, unpublished data) these loci are unlinked. Ten self-derived
washed three times in 2 × SSC for 5 min, and fixed in 10% progeny from VP9, eight test-cross lines from ES136, and two
formaldehyde for 10 min followed by two washes in 2 × SSC test-cross lines from ES137 were analysed.
for 5 min (all steps at 20°C). The preparations were dehydrated
in ethanol series (70%, 90% and absolute ethanol for 3 min each)
Genome-size measurements
and air-dried. For single-colour genome painting, we used a
hybridization mixture of 50% formamide, 10% sodium dextrane Ploidy analyses were performed on a PARTEC (Münster,
sulphate, 2 × SSC, 0.25% SDS and 100 ng DNA probe (BH208 Germany) CCA-II flow cytometer using their CyStain UV
or BS2 genomic DNA) and 10 µg blocking DNA, in a total precise P nuclei extract and staining kit (PARTEC GmbH)
of 40 µl hybridization buffer. For the two-colour genome according to the manufacturer’s protocol. Sample leaf material
painting (GISH), the hybridization mixtures contained both was measured in combination with an internal size standard
BH208 and BS2 probes and blocking DNA (1 : 100 total (leaf material from Brassica rapa and/or Matthiola incana for
genomic DNA of A. thaliana). The mixture was denatured at larger genome individuals), and the ratio of the mean fluores-
100°C for 10 min and chilled immediately on ice for 10 min cence intensity values for the 2c peaks for both sample and
before use. For each preparation we added 40 µl of the standards were calculated (Schranz et al., 2005). When Matthiola
hybridization mix and covered it with a 24 × 50 mm cover was used as standard, the ratio was converted to the Brassica units
slide, followed by heating on an 80°C hot plate for 2.5 min before (by using the size ratio of Matthiolia compared with Brassica).
an overnight hybridization at 37°C in a humidified chamber. A minimum of 10 000 fluorescence counts was collected for
Post-hybridization washes involved three wash steps of 50% each sample run.
formamide in 2 × SSC (pH 7.0) at 42°C for 5 min, and two
steps in 2 × SSC for 5 min at 20°C. The hybridization signals
Seed collection and measurement
were detected with fluorescein isothiocyanate (FITC)-conjugated
anti-DIG antibodies and amplified with FITC-conjugated Seed collections were made from mature plants. Ten siliques
rabbit anti-sheep antibodies for the digoxigenin-labelled probe, from each of five mature reproductive axes per plant were
and with Avidin Texas-Red for the biotin-labelled probe that collected (for a total of 50 siliques). The number of seeds from
was amplified with biotinylated antiavidine and Avidin Texas- each group of 10 siliques was counted and the average number
Red. The preparations were counterstained in 100 µl of a 2 µg of seeds per fruit calculated (Table 1; see Fig. 6).
ml−1 DAPI solution in 100 mM citrate buffer pH 6.0 for 10
min in the dark, and finally mounted in Vectashield (Vector
Pollen grain staining and measurement
Laboratories) under a 24 × 50-mm cover slip. Chromosomes
were examined under a Zeiss Axioplan 2 photomicroscope Measurements of pollen counts and viability were estimated
equipped with epifluorescence illumination and filter sets using a modified Alexander’s stain (Alexander, 1980). Pollen
for DAPI, FITC and Texas Red fluorescence. The images were grains were counted by collecting two to five freshly opened
captured with a Photometrics Sensys 1305 × 1024 image array flowers from each plant. The Alexander stain that gave the
CCD camera and analysed using GENUS ver. 2.7 IMAGE ANALYSIS best results contained the minimum amount of lactic acid
WORKSTATION software (Applied Imaging International Ltd., (0.5 ml), suggested for thin-walled pollen (Alexander, 1980).
Newcastle upon Tyne, UK). DAPI images were sharpened with A dilution of 1 : 20 Alexander stain stock solution to 50%
a 7 × 7 Hi-Gauss high-pass spatial filter to accentuate minor glycerin was used. A total of 20 µl stain dilution was placed
details and heterochromatin differentiation of the chromosomes. on an object slide. An individual stamen from each flower
We used the levels and curves tools in Adobe PHOTOSHOP to was put on a microscopic slide containing a drop of the stain
improve DAPI heterochromatin differentiation banding, and dilution, and covered with a cover slip. We left the material
used the saturation tool to enhance colour saturation of the for at least 10 min before analysing the pollen grains. The
green and red fluorescence signals. numbers of viable (purple) and nonviable (blue/green) pollen
grains were recorded using an Axioskop 2 compound microscope
(Carl Zeiss GmbH, Jena, Germany) at ×20 magnification.
DNA extraction, microsatellite amplification and
Between 400 and 2000 pollen grains for each plant were counted,
analysis
from which the percentage of viable pollen was calculated (Table 1).
The isolation of DNA, microsatellite amplification and analysis The sizes of both viable and nonviable grains were also
were carried out as described by Schranz et al. (2005). Ampli- measured using the stained pollen grains. Photos of the pollen
fications were done with microsatellites (GC27, ICE3, Bf-3, grains were made with a Zeiss AxioCam camera and using the

© The Authors (2006). Journal compilation © New Phytologist (2006) www.newphytologist.org New Phytologist (2006) 171: 425–438
430 Research

program AXIOVISION ver. 3.0.6.1 (Carl Zeiss Vision GmbH, paired bivalents. Metaphase I cells diplayed seven regular
Hallbergmoos, Germany). The area of pollen grains was bivalents, showing that meiosis in this plant can be fully
estimated from the photos using the IMAGEJ ver. 1.33µ program synaptic. At this stage we could identify the Het chromosome
(Abramoff et al., 2004) from which the diameter of each in a few cells forming a heteromorphic bivalent with one of
grain was calculated. Between 20 and 44 individual photos, the other chromosomes (arrow, Fig. 3). We did not find any
representing between 94 and 825 individual pollen grains per anaphase I or anaphase II complement.
plant, were analysed (see Fig. 7).
Microsatellite analysis of self-progeny of VP9 and
Results test-cross lines
In our previous study (Schranz et al., 2005) we analysed micro-
Genome composition revealed by GISH
satellites that were bi-allelic in VP9. In a sample of 10 progeny
Genomic in situ hybridizations (GISH) were initially done with derived by self-pollination of VP9, heterozygosity was maintained
total genomic DNA from the diploid sexual B. stricta labelled (Fig. 4). The probability of obtaining all heterozygous genotypes
as a probe (red) and blocked with genomic DNA from B. holboellii by sexual reproduction with independent assortment is very
(Fig. 2a–d). Chromosome complements showed fluorescent unlikely (P < 0.001). This result strongly supports the conclusion
signals on the pericentromere regions of only the seven B. stricta- that VP9 is an apomictic lineage. The microsatellites used were
derived chromosomes in the hybrid B. divaricarpa VP9 genome also highly polymorphic between VP9 and both SAD12 and
(Fig. 2c), confirming its homoploid origin (hybridization with ES138. The test-cross lines, where ES136 and ES137 were
no increase in ploidy) between a B. stricta- and B. holboellii-like used as maternal plants and ES138 as a pollen donor, all
plant. Additionally, the hybrid B. divaricarpa VP9 genome was contained alleles derived from ES138. This result, and the
found to contain a highly heterochromatic chromosome (Het) dramatic change in phenotype (exemplified by the shift to very
(indicated by arrows in Fig. 2a,c) that may have properties late flowering time), both indicate that the near-tetraploid
similar to B-chromosomes and/or Het chromosomes identified lines ES136 and ES137 reproduce sexually.
in 15-chromosome Boechera apomicts (Sharbel et al., 2004;
Kantama, 2005). The fluorescent signal on the Het chromosome
Genome size, chromosome numbers and karyotypes
(Fig. 2c) suggests it is derived from the B. stricta ancestor of
this hybrid line. Genome painting of ES 117 F1 line chromo- Flow cytometry results of all lines were standardized to the
somes (Fig. 2d) further revealed a total of 14 B. stricta chromo- diploid B. rapa genome (Table 1), and chromosome counts
somes (seven derived from a reduced haploid B. stricta SAD12 were done for most of the parental, F1 and F2 plants (Table 1).
gamete; the other seven from the nonreduced B. divaricarpa Comparison of flow cytometry values with chromosome counts
VP9 gamete) and seven B. holboellii-like chromosomes (seven (assuming all chromosomes are of approximately the same
derived from the nonreduced B. divaricarpa VP9 gamete). size) showed very good correspondence (R 2 = 0.98) (Fig. 5a).
An additional two-colour genome painting was carried Chromosome counts and karyotype analysis (Fig. 5b) confirmed
out with the simultaneous hybridization of both B. stricta that the parental genotypes were diploid (2n = 2x = 14).
(green) and B. holboellii (red) probes, and blocking with total The F1 plants (ES 116 and ES 117) were also confirmed to be
genomic DNA of A. thaliana in order to further improve the triploids (2n = 3x = 21) (Fig. 5b), as hypothesized from earlier
discrimination of the parental species (Kantama, 2005). The flow cytometry results, probably caused by the union of a reduced
painting of one of the highly fertile F2 lines, ES 137, revealed and nonreduced gamete (Schranz et al., 2005)
14 B. stricta chromosomes (green), including the highly The most striking result was the swarm of chromosome
heterochromatic (Het) chromosome (arrow in Fig. 2e), and numbers of the F2 lines, ranging from triploid to tetraploid,
13 B. holboellii-like chromosomes (Fig. 2e). Thus the two with most being aneuploid (Table 1; Fig. 5a). While overall
likely aneuploid gametes derived from the ES 117 line (with genome-size measurements agreed with the chromosome
its 14 B. stricta and seven B. holboellii-like chromosomes) that counts, there was one notable exception. The F2 line ES 117
united to form the ES 137 plant were effectively able to restore #9 had the largest genome size measured by flow cytometry
the chromosome balance of one B. stricta to one B. holboellii-like (1.12), but gave a mosaic of aneuploid chromosome numbers
chromosome (Fig. 2e). The restored genomic ratio could explain (24 and 27). Two F2 lines, ES 136 and ES 137, were found to
the higher fertility of this line. However, we do not yet know be highly fertile (see below). Chromosome counts for both
if we have complete chromosome complements in these lines. these lines showed that they were nearly tetraploid (2n ≈ 4x ≈
27) (Table 1; Fig. 5b). Because of their fertility, we investi-
gated the genome size of their F3 progeny. All F3 progeny had
Analysis of pollen meiosis of VP9
genome sizes of the same magnitude (ranging from 0.94 to
Anther preparations of VP9 showed very few meiocytes. We 1.08) (Table 1; Fig. 5c), suggesting that the genomes of these
found pachytene and metaphase I complements with fully two lines have stabilized near tetraploidy.

New Phytologist (2006) 171: 425–438 www.newphytologist.org © The Authors (2006). Journal compilation © New Phytologist (2006)
Research 431

Fig. 2 Genome composition revealed by genomic in situ hybridization


(GISH) of selected Boechera SAD12 × VP9 population lines. DAPI
staining of chromosomes of: (a) paternal apomictic Boechera
divaricarpa genotype (VP9); (b) F1 line (ES 117). Fluorescence in
situ hybridizations done with total genomic DNA from the diploid
sexual Boechera stricta labelled as a probe (red) and blocked with
genomic DNA from Boechera holboellii staining the pericentromere
regions of: (c) seven B. stricta-derived chromosomes in the hybrid
B. divaricarpa (VP9) genome; (d) 14 B. stricta-like chromosomes in
the F1 line (ES 117) genome (arrows in Fig. 5(a,c) denote the Het
chromosome). Additional two-colour genome painting done by
the simultaneous hybridization of both B. stricta (green) and B.
holboellii (red) probes onto (e) one of the high-fertility F2 lines, ES
137, showing 14 B. stricta chromosomes (green) including a highly
heterochromatic chromosome (arrow) and 13 B. holboellii-like
chromosomes.

© The Authors (2006). Journal compilation © New Phytologist (2006) www.newphytologist.org New Phytologist (2006) 171: 425–438
432 Research

most lines of varying genome sizes were infertile, but the two
higher-fertility lines were nearly tetraploid. The F3 lines
derived from ES 136 and ES 137 all maintained high levels of
fertility, with an average of 15.11 seeds per fruit; however,
there was still great variation in seed production (5.65–48.88
seeds per silique; Table 1).

Pollen grain viability and size


The viability of pollen grains differed between the two parental
lines, with sexual B. stricta SAD12 producing almost all viable
pollen, and the apomictic B. divaricarpa VP9 producing 74.88%
viable pollen (Table 1). The triploid F1 plant (ES 117) had greatly
reduced pollen viability of 24.52% (Table 1). The swarm of
F2 plants displayed great variation in viability, from only
1.87% up to 75.88 and 72.81% for lines ES 136 and ES 137,
Fig. 3 DAPI-stained chromosome spread of anther cells of VP9.
Metaphase I complement with seven bivalents showing that meiosis
respectively (Table 1). The F3 progeny of ES 136 and ES 137
in this plant is fully synaptic. Arrow, Het chromosome in the maintained the higher-viability pollen, with an average of
heteromorphic bivalent. 69.92% (Table 1).
The analysis of pollen grain size of viable and nonviable pollen
revealed several striking results (Fig. 7). The sexual B. stricta
SAD12 produced almost exclusively reduced and viable
pollen grains with an average diameter of 16.53 µm (Fig. 7a).
The apomictic B. divaricarpa VP9 produced viable, nonreduced
pollen grains with an average diameter of 22.32 µm. Interest-
ingly, VP9 also produced some pollen grains of even larger
diameter (Fig. 7a). The F1 line ES 117 produced viable pollen
grains of varying size, from 17.66 to 25.82 µm (Fig. 7a), probably
Fig. 4 Microsatellite amplification of 10 progeny derived by self- reflecting a wide range of aneuploid gametes. However, no
pollination of VP9. At several unlinked microsatellite loci, including viable gametes of haploid size were detected. The nonviable
this locus (GC27), the initial VP9 plant was heterozygous. In a sample
of 10 progeny derived by self-pollination of VP9, heterozygosity was
pollen of ES 117 (Fig. 7b) was found to be smaller than that
maintained for all samples, suggesting that progeny were derived of the viable pollen. The size range of the nonviable pollen was
apomictically rather than sexually. from 12.51 to 22.31 µm, corresponding to gametes that would
be less than haploid to about diploid in size. The F2 line, ES
117 #15, was triploid and produced almost exclusively non-
The strong correlation between the flow cytometry results viable pollen (1.87% viability; Table 1). The size distribution
and chromosome number allowed us to estimate the genome of the nonviable grains for ES 117 #15 (Fig. 7b) ranged from
size of Boechera. A recent study (Johnston et al., 2005) estimated 11.92 to 19.90 µm, and thus were smaller than diploid in size.
the haploid genome of B. rapa (n = 10) at 1C = 529 Mb. Using Finally, one of the F3 lines derived from ES 137 (ES 137 #15)
this estimate, the haploid Boechera genome (n = 7) would be produced viable pollen grains from 18.89 to 25.09 µm, and
approx. 264 Mb. However, their estimate for the size of the nonviable pollen grains from 13.37 to 20.87 µm in diameter
A. thaliana genome is 157 Mb ( Johnston et al., 2005), which (Fig. 7a,b).
is larger than current estimates based on genome sequencing. The six lines discussed above (Fig. 7a,b) illustrate the
Using the latter value, we infer that the haploid Boechera variability in pollen viability and size from the various genera-
genome is roughly 1.7 times that of A. thaliana. tions. In addition, overall pollen grain viability and size for all
nonparental lines was also determined (Fig. 7c). The average
size of nonviable gametes was 15.88 µm, which is slightly less
Seed production of aneuploid and polyploid F2 lines
than that of haploid, and far less than the 21.91 µm of the
Analysis of seed production of F1 and F2 plants found that most diploid viable gametes (Fig. 7c). It should be noted that the
were highly infertile (Table 1; Fig. 6). However, two F2 lines, ratio of nonviable to viable pollen grains measured for size was
ES 136 and ES 137, had much greater seed set with c. 10 seeds lower than that obtained by simply counting nonviable and
per silique, compared with an average of only 0.36 seeds per viable pollen grains. This discrepancy was because nonviable
silique for the other lines (Table 1; Fig. 6). Correlation between pollen grains clumped together and thus it was difficult to
genome size and seed set (Fig. 6) was not significant, as take good photographs for size measurements.

New Phytologist (2006) 171: 425–438 www.newphytologist.org © The Authors (2006). Journal compilation © New Phytologist (2006)
Research 433

Fig. 5 Measures of genome size of lines in


the Boechera SAD12 × VP9 population.
(a) Correlation of genome size (measured by
flow cytometry and given as ratio to reference
Brassica rapa genome) and chromosome
numbers. (b) Karyotypes and chromosome
numbers of selected generations and lines.
(c) Distribution in genome size (measured by
flow cytometry and given as ratio to reference
B. rapa genome) of the F3 lines derived from
ES136 and ES137, demonstrating the genomic
stability of these lines near tetraploidy.

that fully pairs with one of the other chromosomes. Male meiosis
Discussion produced both reduced and unreduced gametes, suggesting
Here we present the first genetic analysis of the mechanisms first- or second-division restitution. Heterozygosity was also
and inheritance of apomictic reproduction in Boechera. First, maintained in self-pollinated progeny, thus supporting its
we characterized more fully the paternal diploid apomictic apomictic nature. We wondered if this Het chromosome carries
line VP9. Genome painting revealed its hybrid origin, whereas the genetic elements controlling apomixis in a manner similar
the analysis of male meiosis demonstrated that the chromosomes to several other systems where supernumerary or hemizygous
are fully synaptic, including a heterochromatic (Het) chromosome chromosomes carry an apomixis factor (Ozias-Akins et al.,

© The Authors (2006). Journal compilation © New Phytologist (2006) www.newphytologist.org New Phytologist (2006) 171: 425–438
434 Research

Fig. 6 Comparison of genome size (measured


by flow cytometry and given as ratio to
reference Brassica rapa genome) and seed
production of F2 lines of the Boechera
SAD12 × VP9 population. Most lines are
highly infertile except two tetraploids (ES 136
and ES 137) with greater seed production.

1998; Bicknell et al., 2000; Roche et al., 2001; Labombarda or supernumerary element (Nogler, 1984b; Sherwood et al.,
et al., 2002; Akiyama et al., 2005). Therefore we analysed the 1994; Bicknell et al., 2000; Roche et al., 2001). Our F1 plants
progeny derived from a cross between a diploid sexual and this were derived from the union of a reduced egg cell from the
diploid apomictic line. The resulting triploid F1 was largely diploid sexual maternal plant and a nonreduced pollen grain
sterile, producing limited numbers of viable pollen and seed. from the apomictic paternal plant, as shown by GISH analysis.
When the few seeds were grown, they produced plants that Hence the F1 progeny probably inherited the complete genome
were mostly aneuploid or near-tetraploid. Hence they were of the apomictic parent. However, there may have been cross-
not apomictically derived replicates of the three maternal over events during the failed meioses of the paternal VP9 plant
genomes, but rather the product of a reduced or unreduced that could have produced recombinant pollen grains. Future
meiosis. Two of the recovered lines were fertile, sexual and segregation analyses of molecular markers will be necessary to
near-tetraploid. Genome painting (GISH) analysis of one examine this possibility.
of these fertile lines found that it had a nearly balanced Our chromosome analyses showed that the diploid apomictic
chromosome complement of 14 B. stricta and 13 B. holboellii paternal plant had an aberrant chromosome resembling
chromosomes, suggesting that genome balance may play an the heterochromatic (Het) chromosome in other 14- and 15-
important role in establishing the sexual phenotype. However, chromosome Boechera apomicts (Kantama, 2005), but lacked
future work will be needed to establish if these correspond the extra chromosome found in the 15-chromosome apomicts
to complete chromosomal complements. Below we discuss (Sharbel et al., 2004, 2005). The offspring individuals in
in greater detail some of the implications of our results for the F2 progeny displayed varying aneuploid and polyploid
understanding the inheritance, control and evolution of chromosome numbers, suggesting that apomixis in Boechera
apomixis in Boechera, and the potential importance of the is not regulated by the simple inheritance of a single dominant
generation of sexual tetraploid lineages. locus or factor (e.g. the Het chromosome). If it were, then
the F1 line, which contains the Het chromosome, would be
apomictic with all or most derived F2 progeny being identical
Control of apomixis in Boechera
to the maternal plant.
Asexual seed formation is a complex trait resulting from modi- Various authors have pointed to polyploidy as being indis-
fication of the sexual life cycle (reviewed by Koltunow & pensible in the regulation and transmission of the apomixis trait
Grossniklaus, 2003; Bicknell & Koltunow, 2004). The normal (Quarin et al., 2001). But in Boechera, where naturally occurring
sexual pathway can be altered or deregulated by the inheritance triploids are indeed apomictic, they can be highly facultative
of a single dominant gene, hemizygous genomic region, and/ with many progeny derived sexually (Schranz et al., 2005). The

Fig. 7 Histograms of Boechera pollen grain size and viability. Alexander’s stain was used to differentiate viable (purple) and nonviable (blue/
green) pollen grains, and the diameter of each grain was measured by analysis of photos of pollen. Frequency based on total number of pollen
grains counted. Size distribution and selected pollen grain images of (a) viable; (b) nonviable pollen grains of six selected lines used to illustrate
the diversity of viability and size of the various lines. The lines used were SAD12 (diploid sexual); VP9 (diploid apomict); ES117 (triploid F1);
ES137 (near-tetraploid F2); ES117 #15 (aneuploid F2); ES137 #15 (near-tetraploid F3). (c) Size distribution of all viable and nonviable pollen grains
measured from aneuploid and polyploid lines (excluding the two diploid parental lines).

New Phytologist (2006) 171: 425–438 www.newphytologist.org © The Authors (2006). Journal compilation © New Phytologist (2006)
Research 435

© The Authors (2006). Journal compilation © New Phytologist (2006) www.newphytologist.org New Phytologist (2006) 171: 425–438
436 Research

results of our crossing have also shown that an increase in of the nonviable pollen grains might be caused simply by
ploidy from diploidy to triploidy is not necessarily associated chromosomal imbalances, particularly those generated from
with the transfer or expression of the apomictic phenotype. aneuploid F2 lines where the percentages of nonviable pollen
It has also been postulated that the hybrid constitution of grains were as high as 98%. However, both the apomictic
apomictic lineages, rather than polyploidy, will lead to expres- diploid parent and the recovered sexual tetraploids produced,
sion of the apomictic phenotype (Bicknell & Koltunow, 2004). on average, 25% nonviable gametes and 75% viable diploid-
Specifically, the hybridization-derived floral asynchrony hypo- sized pollen grains. This result would be consistent with the
thesis postulates that apomixis is caused by the differential recessive lethal gametophytic selection hypothesis of (Nogler,
expression and temporal regulation of genes in normal 1984b), in which an apomixis gene or factor is lethal when it
sexual reproductive pathways (Carman, 1997; Koltunow & occurs in a haploid pollen grain in the absence of a wild-type
Grossniklaus, 2003). In agreement with this hypothesis is our allele. Future studies examining the patterns of inheritance of
demonstration by GISH that the diploid apomictic parent molecular and/or cytological markers will be critical for resolving
(VP9) is of hybrid origin, with seven B. stricta and seven B. models for the control of apomixis in Boechera.
holboellii-like chromosomes. In addition, all diploid and triploid
lines that were apomictic and/or produced nonreduced
Escape from apomixis, low fertility and return to
gametes were also highly heterozygous, attesting to their likely
sexuality
hybrid origin (Schranz et al., 2005). Similarly, individuals
found to be heterozygous at microsatellite loci (DobeS et al., Traditionally, apomictic lineages were thought to have limited
2004b) and at a sequenced molecular marker (Sharbel et al., evolutionary potential and to be doomed to extinction (Darlington,
2004, 2005) tended to produce nonreduced gametes, and 1939; Stebbins, 1950). Later, researchers realized that the
were concluded to be apomictic. maintenance of male function and facultative apomixis provided
However, hybridization alone cannot explain apomixis in mechanisms for perpetuating apomictic lineages by transferring
Boechera. Our triploid F1, which contained a complete haploid apomixis genes via hybridization with sexual relatives (van
genome from the highly inbred maternal B. stricta plant and Dijk, 2003). The transfer of apomixis genes into new genetic
the potentially complete diploid B. divaricarpa genome from backgrounds would allow the purging or masking of deleterious
the apomictic paternal plant, is not apomictic. One possible mutations that may have accumulated in the apomict (Muller’s
explanation for this failure in transmission of the apomixis ratchet), and would generate new genetic diversity, allowing
phenotype could be segregation of the apomixis gene(s) caused escape from parasitism. Another possibility is that the hybridiza-
by recombination during the production of the 2n pollen tion of asexuals with sexuals allows genes from the apomictic
grains in the apomictic paternal plant. Chromosome analysis, parent to re-enter the sexual gene pool (Chapman et al., 2003),
however, showed that all or part of the heterochromatic undergo recombination, and potentially go on to form new
(Het) chromosome was transmitted to the F1 plant. Another apomictic lineages (de Wet, 1968).
explanation for the discrepancy may lie in the relative dosage We have detected such a shift in breeding system, from
of B. stricta and B. holboellii-like chromosomes. GISH analyses apomict diploid to low-fertility sexual triploid to fertile
of the diploid apomictic showed there is a 1 : 1 genome ratio, and sexual tetraploid, in our experimental cross of Boechera. In
but in the F1 triploid there is a ratio of 2 : 1 in favour of the many ways our cross follows the classic model of the formation
sexual B. stricta genome. The higher dose of B. stricta might allow of autotetraploids via a triploid bridge, and the inherent problems
for the correct, and disruptive, expression of sexuality in the of meiosis in triploids and the frequent endosperm develop-
triploid. Naturally occurring triploid apomictic lineages may mental problem of the triploid block (Ramsey & Schemske,
have two doses of B. holboellii-like chromosomes, or a dosage 1998; Husband, 2004; Henry et al., 2005).
closer to 1 : 1 if the triploid apomictic line is derived from the There are two routes from apomictic genotypes to the
union of two aneuploid gametes. However, one of the recovered formation of sexual tetraploids in Boechera. In this work we
F2 near-tetraploid lines has a nearly 1 : 1 genome ratio, yet is demonstrated that, by crossing the two diploids, we could
sexual. Sexuality at the tetraploid level may be feasible because generate sexual near-tetraploids in only two generations. In
of an interaction of dosage and polyploidization, or alternatively our earlier work, we frequently derived de novo tetraploids
because of the segregation of a gamete-lethality factor. directly from the crossing of a sexual diploid and an apomictic
The results of our pollen analysis may shed light on the triploid line. These sexual tetraploid lines can then undergo
possible segregation of a gamete-lethal factor or gene. In all all the classic advantages of sex, such as recombination and
samples, except the diploid sexual line, there a significant independent assortment. There could also be chances for
fraction of nonviable pollen grains were produced, which homologous pairing, translocations and epigenetic modifica-
were of approximately the same size as viable haploid pollen. tions often seen in new polyploid lines (Osborn et al., 2003),
In contrast, viable pollen grains in all samples, except for the all generating new phenotypic diversity. Also, it is important
sexual diploid, were of approximately diploid size. No viable to note that these tetraploid Boechera lines could potentially
haploid-sized pollen was observed for any of these lines. Some express apomixis, or some partial component of apomixis, at

New Phytologist (2006) 171: 425–438 www.newphytologist.org © The Authors (2006). Journal compilation © New Phytologist (2006)
Research 437

some frequency, caused by the penetration or expressiveness Albertini E, Porceddu A, Ferranti F, Reale L, Barcaccia G, Romano B,
of the traits. The partial expression of apomixis, especially Falcinelli M. 2001. Apospory and parthenogenesis may be uncoupled in
Poa pratensis: a cytological investigation. Sexual Plant Reproduction 14:
parthenogenesis, may be particularly important for the 213–217.
generation of new diploid apomictic lineages. The establishment Alexander MP. 1980. A versatile stain for pollen, fungi, yeast and bacteria.
of diploid hybrid apomictic haplotypes in Boechera could be Stain Technology 55: 13–18.
caused by homoploid hybridization, or alternatively by base Asker SE, Jerling L. 1992. Apomixis in Plants. Boca Raton, FL, USA: CRC
chromosome number reductions (Asker & Jerling, 1992). If Press.
Bicknell RA, Koltunow AM. 2004. Understanding apomixis: recent
reduced 2n egg cells from tetraploid Boechera lineages develop advances and remaining conundrums. Plant Cell 16: S228–S245.
parthenogenically, they could establish new diploid lineages Bicknell RA, Borst NK, Koltunow AM. 2000. Monogenic inheritance of
that could be either sexual or apomictic. Such cycles of ploidy apomixis in two Hieracium species with distinct developmental
have been described for other apomictic complexes, such as mechanisms. Heredity 84: 228–237.
Potentilla argentea and the Bothriochloa–Dichanthium complex, Böcher T. 1951. Cytological and embryological studies in the amphil–
apomictic Arabis holboellii complex. Biologiske Skrifter 6: 1–58.
and are known as the ‘diploid–tetraploid–dihaploid cycle’ (de Böcher T. 1954. Experimental taxonomical studies in the Arabis holboellii
Wet, 1968; Asker & Jerling, 1992). The independent assortment complex. Svensk Botanisk Tidskrift 48: 31–44.
and recombination of chromosomes derived from the parental Böcher T. 1969. Further studies in Arabis holboellii and allied species.
genotypes or species of the tetraploid means that there could Saertryk af Tidsskrift 64: 141–161.
be unusual constellations of B. stricta and B. holboellii chromo- Carman J. 1997. Asynchronous expression of duplicated genes in
angiosperms may cause apomixis, bispory, tetraspory, and polyembryony.
somes or chromosome regions in dihaploids produced Biological Journal of the Linnean Society 61: 51–94.
parthenogenically. Chapman H, Houliston GJ, Robson B, Iline I. 2003. A case of reversal:
Interestingly, few naturally occurring tetraploid Boechera the evolution and maintenance of sexuals from parthenogenetic clones in
lineages have been detected (reviewed by DobeS et al., 2006), Hieracium pilosella. International Journal of Plant Sciences 164: 719 –728.
but when they do occur, they are likely to be sexual (Böcher, 1969; Charlesworth B. 1990. Mutation–selection balance and the evolutionary
advantage of sex and recombination. Genetical Research 55: 199–221.
Johnson, 1970; Schranz et al., 2005). The low frequency of Crane CF. 2001. Classification of apomictic mechanisms. In: Savidan Y,
tetraploids could be explained by a potential propensity to Carman J, Dresselhaus T, eds. The Flowering of Apomixis: From
produce dihaploid offspring. Alternatively they may be selected Mechanisms to Genetic Engineering. Mexico: CIMMYT, IRD/European
against, possibly because of a fitness disadvantage. It is inter- Commission DG VI, 24–43.
esting to note that sexual tetraploids existing at low frequency Darlington C. 1939. The Evolution of Genetic Systems. Cambridge, UK:
Cambridge University Press.
in populations of dandelions have been postulated to be critical van Dijk PJ. 2003. Ecological and evolutionary opportunities of apomixis:
for the creation of new triploid apomictic cytotypes, by the insights from Taraxacum and Chondrilla. Philosophical Transactions of the
production of reduced 2n gametes that recombine with reduced Royal Society of London, Series B – Biological Sciences 358: 1113–1120.
n gametes from sexual diploids (Verduijn et al., 2004). Overall, van Dijk PJ, Tas ICQ, Falque M, Bakx-Schotman T. 1999. Crosses between
the historical occurrence of hybridization shifts in breeding sexual and apomictic dandelions (Taraxacum). II. The breakdown of
apomixis. Heredity 83: 715–721.
systems and alterations of ploidy have profound implications Dobes CH, Koch MA, Sharbel TF. 2006. Embryology, karyology, and
for our understanding of the inheritance, transmission and modes of reproduction in North American genus Boechera (Brassicaceae):
evolution of genetic and quantitative variation for Boechera. a compilation of seven decades of research. Annals of the Missouri Botanical
Garden. (In press.)
Dobes C, Mitchell-Olds T, Koch MA. 2004a. Intraspecific diversification in
Acknowledgements North American Boechera stricta (= Arabis drummondii ), Boechera ×
divaricarpa and Boechera holboellii (Brassicaceae) inferred from nuclear and
The authors thank Juliane Pfuetzenreuter, Silke Fuchs and Radim chloroplast molecular markers – an integrative approach. American Journal
Vasut for technical assistance. Thanks to John E. Maunder and of Botany 91: 2087–2101.
Nathalie Djan-Chékar of the Provincial Museum of New- Dobes CH, Mitchell-Olds T, Koch MA. 2004b. Extensive chloroplast
foundland and Labrador for providing seeds. The authors haplotype variation indicates Pleistocene hybridization, radiation of
North American Arabis drummondii Arabis drummondii, A. × divaricarpa,
also thank Peter van Dijk for helpful comments. Support for and A. holboellii (Brassicaceae). Molecular Ecology 13: 349–370.
this research was provided by the Max Planck Gesellschaft to Henry I, Dilkes B, Young K, Watson B, Wu H, Comai L. 2005. Aneuploidy
M.E.S. and by the Thai Government to L.K. and genetic variation in the Arabidopsis thaliana triploid response. Genetics
170: 1979–1988.
Holsinger KE. 2000. Reproductive systems and evolution in vascular plants.
References Proceedings of the National Academy of Sciences, USA 97: 7037–7042.
Husband BC. 2004. The role of triploid hybrids in the evolutionary
Abramoff MD, Magelhaes PJ, Ram SJ. 2004. Image processing with dynamics of mixed-ploidy populations. Biological Journal of the Linnean
IMAGEJ. Biophotonics International 11: 36 – 42. Society 82: 537–546.
Akiyama Y, Hanna W, Ozias-Akins P. 2005. High-resolution physical Johnston JS, Pepper AE, Hall AE, Chen ZJ, Hodnett G, Drabek J,
mapping reveals that the apospory-specific genomic region (ASGR) in Lopez R, Price HJ. 2005. Evolution of genome size in Brassicaceae.
Cenchrus ciliaris is located on a heterochromatic and hemizygous region Annals of Botany 95: 229–235.
of a single chromosome. Theoretical and Applied Genetics 111: 1042– Johnson TF. 1970. Investigations in the floral biology of the Arabis holboellii
1051. complex. MSc thesis. Seattle, WA, USA: University of Washington.

© The Authors (2006). Journal compilation © New Phytologist (2006) www.newphytologist.org New Phytologist (2006) 171: 425–438
438 Research

Kantama L. 2005. Chromosome studies and genetic analyses of natural have no allelic form in sexual genotypes. Proceedings of the National
and synthetic apomictic model species. PhD thesis. Wageningen, Academy of Sciences, USA 95: 5127–5132.
the Netherlands: Wageningen University. Quarin CL, Espinoza F, Martinez EJ, Pessino SC, Bovo OA. 2001. A rise
Koch M, Haubold B, Mitchell-Olds T. 2001. Molecular systematics of of ploidy level induces the expression of apomixis in Paspalum notatum.
the Brassicaceae: evidence from coding plastidic matK and nuclear Sexual Plant Reproduction 13: 243–249.
Chs sequences. American Journal of Botany 88: 534 – 544. Ramsey J, Schemske DW. 1998. Pathways, mechanisms, and rates of
Koch MA, Dobes C, Mitchell-Olds T. 2003. Multiple hybrid formation in polyploid formation in flowering plants. Annual Review of Ecology and
natural populations: concerted evolution of the internal transcribed spacer Systematics 29: 467–501.
of nuclear ribosomal DNA (ITS) in north American Arabis divaricarpa Richards A. 1986. Plant Breeding Systems. London: George Allen & Unwin.
(Brassicaceae). Molecular Biology and Evolution 20: 338 – 350. Roche D, Hanna WW, Ozias-Akins P. 2001. Is supernumerary chromatin
Koltunow AM, Grossniklaus U. 2003. Apomixis: a developmental involved in gametophytic apomixis of polyploid plants? Sexual Plant
perspective. Annual Review of Plant Biology 54: 547– 574. Reproduction 13: 343–349.
Labombarda P, Busti A, Caceres ME, Pupilli F, Arcioni S. 2002. An Roy BA. 1995. The breeding systems of six species of Arabis (Brassicaceae).
AFLP marker tightly linked to apomixis reveals hemizygosity in a portion American Journal of Botany 82: 869–877.
of the apomixis-controlling locus in Paspalum simplex. Genome 45: Schranz ME, Dobes C, Koch MA, Mitchell-Olds T. 2005. Sexual
513–519. reproduction, hybridization, apomixis, and polyploidization in the genus
Leblanc O, Grimanelli D, Gonzalez-d, e-Leon D, Savidan Y. 1995. Boechera (Brassicaceae). American Journal of Botany 92: 1797–1810.
Detection of the apomictic mode of reproduction in maize–Tripsacum Sharbel TF, Mitchell-Olds T. 2001. Recurrent polyploid origins and
hybrids using maize RFLP markers. Theoretical and Applied Genetics 90: chloroplast phylogeography in the Arabis holboellii complex (Brassicaceae).
1198–1203. Heredity 87: 59–68.
Matzk F, Prodanovic S, Baumlein H, Schubert I. 2005. The inheritance of Sharbel TF, Voigt ML, Mitchell-Olds T, Kantama L, de Jong H. 2004. Is
apomixis in Poa pratensis confirms a five locus model with differences in the aneuploid chromosome in an apomictic Boechera holboellii a genuine
gene expressivity and penetrance. Plant Cell 17: 13 –24. B chromosome? Cytogenetic and Genome Research 106: 173–183.
Mogie M. 1992. The Evolution of Asexual Reproduction in Plants. New York, Sharbel TF, Mitchell-Olds T, Dobes C, Kantama L, de Jong H. 2005.
NY, USA: Chapman & Hall. Biogeographic distribution of polyploidy and B chromosomes in the
Naumova TN, van der Laak J, Osadtchiy J, Matzk F, Kravtchenko A, apomictic Boechera holboellii complex. Cytogenetic and Genome Research
Bergervoet J, Ramulu KS, Boutilier K. 2001. Reproductive development 109: 283–292.
in apomictic populations of Arabis holboellii (Brassicaceae). Sexual Plant Sherwood RT, Berg CC, Young BA. 1994. Inheritance of apospory in
Reproduction 14: 195–200. buffelgrass. Crop Science 34: 1490–1494.
Nogler GA. 1984a. Gametophytic apomixis. In: Johri B, ed. Embryology Spillane C, Curtis MD, Grossniklaus U. 2004. Apomixis technology
of Angiosperms. Berlin: Springer, 475 – 518. development – virgin births in farmers’ fields? Nature Biotechnology 22:
Nogler GA. 1984b. Genetics of apospory in apomictic Ranunculus 687–691.
auricomus. 5. Conclusion. Botanica Helvetica 94: 411– 422. Stebbins G. 1950. Variation and Evolution in Higher Plants. New York:
Noyes RD, Rieseberg LH. 2000. Two independent loci control Columbia University Press.
agamospermy (apomixis) in the triploid flowering plant Erigeron annuus. Taskin KM, Turgut K, Scott RJ. 2004. Apomictic development in Arabis
Genetics 155: 379–390. gunnisoniana. Israel Journal of Plant Sciences 52: 155–160.
Osborn TC, Pires JC, Birchler JA, Auger DL, Chen ZJ, Lee HS, Comai L, Verduijn MH, Van Dijk PJ, Van Damme JMM. 2004. The role of
Madlung A, Doerge RW, Colot V, Martienssen RA. 2003. tetraploids in the sexual–asexual cycle in dandelions (Taraxacum). Heredity
Understanding mechanisms of novel gene expression in polyploids. Trends 93: 390–398.
in Genetics 19: 141–147. West S, Lively C, Read A. 1999. A pluralist approach to sex and
Ozias-Akins P, Roche D, Hanna WW. 1998. Tight clustering and recombination. Journal of Evolutionary Biology 12: 1003–1012.
hemizygosity of apomixis-linked molecular markers in Pennisetum de Wet J. 1968. Diploid–tetraploid–haploid cycles and the origin of
squamulatum genetic control of apospory by a divergent locus that may variability in Dichanthium agamospecies. Evolution 22: 394–397.

New Phytologist (2006) 171: 425–438 www.newphytologist.org © The Authors (2006). Journal compilation © New Phytologist (2006)

You might also like