ArticlePDF Available

Wheat Hybridization and Polyploidization Results in Deregulation of Small RNAs

Authors:

Abstract and Figures

Speciation via interspecific or intergeneric hybridization and polyploidization triggers genomic responses involving genetic and epigenetic alterations. Such modifications may be induced by small RNAs, which affect key cellular processes, including gene expression, chromatin structure, cytosine methylation and transposable element (TE) activity. To date, the role of small RNAs in the context of wide hybridization and polyploidization has received little attention. In this work, we performed high-throughput sequencing of small RNAs of parental, intergeneric hybrid, and allopolyploid plants that mimic the genomic changes occurring during bread wheat speciation. We found that the percentage of small RNAs corresponding to miRNAs increased with ploidy level, while the percentage of siRNAs corresponding to TEs decreased. The abundance of most miRNA species was similar to midparent values in the hybrid, with some deviations, as seen in overrepresentation of miR168, in the allopolyploid. In contrast, the number of siRNAs corresponding to TEs strongly decreased upon allopolyploidization, but not upon hybridization. The reduction in corresponding siRNAs, together with decreased CpG methylation, as shown here for the Veju element, represent hallmarks of TE activation. TE-siRNA downregulation in the allopolyploid may contribute to genome destabilization at the initial stages of speciation. This phenomenon is reminiscent of hybrid dysgenesis in Drosophila.
Content may be subject to copyright.
Copyright © 2011 by the Genetics Society of America
DOI: 10.1534/genetics.111.128348
Wheat Hybridization and Polyploidization Results in Deregulation
of Small RNAs
Michal Kenan-Eichler,* Dena Leshkowitz,
Lior Tal,* Elad Noor,* Cathy Melamed-Bessudo,*
Moshe Feldman* and Avraham A. Levy*
,1
*
Department of Plant Sciences, Weizmann Institute of Science, Rehovot, 76100 Israel and
Bioinformatics Unit,
Weizmann Institute of Science, Rehovot, 76100 Israel
Manuscript received March 3, 2011
Accepted for publication April 4, 2011
ABSTRACT
Speciation via interspecic or intergeneric hybridization and polyploidization triggers genomic responses
involving genetic and epigenetic alterations. Such modications may be induced by small RNAs, which
affect key cellular processes, including gene expression, chromatin structure, cytosine methylation and
transposable element (TE) activity. To date, the role of small RNAs in the context of wide hybridization and
polyploidization has received little attention. In this work, we performed high-throughput sequencing of
small RNAs of parental, intergeneric hybrid, and allopolyploid plants that mimic the genomic changes
occurring during bread wheat speciation. We found that the percentage of small RNAs corresponding to
miRNAs increased with ploidy level, while the percentage of siRNAs corresponding to TEs decreased. The
abundance of most miRNA species was similar to midparent values in the hybrid, with some deviations, as seen
in overrepresentation of miR168, in the allopolyploid. In contrast, the number of siRNAs corresponding to TEs
strongly decreased upon allopolyploidization, but not upon hybridization. The reduction in corresponding
siRNAs, together with decreased CpG methylation, as shown here for the Veju element, represent hallmarks of
TE activation. TE-siRNA downregulation in the allopolyploid may contribute to genome destabilization at the
initial stages of speciation. This phenomenon is reminiscent of hybrid dysgenesis in Drosophila.
INTERSPECIFIC or intergeneric hybridization and
whole-genome doubling provide a mechanism for
overnightspeciation (Rieseberg and Willis 2007;
Leitch and Leitch 2008; Soltis and Soltis 2009;
Wood et al. 2009). Wheat, Spartina, Senecio, and Trag-
opogon are examples of recent speciation via wide hy-
bridization and polyploidization. It has been argued
that in nature, there are plenty of opportunities for in-
terspecic hybridization and genome doubling, with an
estimated 10% hybridization among animals and 25%
among plants (Mallet 2007). Moreover, unreduced
gametes, which are occasionally produced in most plant
species (Bretagnolle and Thompson 1995), were
shown to occur at a frequency of up to 50% in some
combinations of intergeneric hybrids of wheat (Kihara
and Lilienfeld 1949). Therefore, one might expect
new hybrid and polyploid species to be formed more
frequently than documented (Van DePeer et al. 2009).
This apparent paradox may be resolved by works dem-
onstrating the genome-wide genetic and epigenetic im-
pact of genome merging or doubling, consequently
challenging maintenance of genome functionality and
stability. While some intergenomic combinations show
hybrid incompatibility (Bomblies and Weigel 2007;
Ishikawa and Kinoshita 2009; Walia et al. 2009) or
hybrid dysgenesis (Malone and Hannon 2009), or fea-
ture reduced tness, others show improved vigor (Chen
2010). The mechanisms dictating nascent hybrid and
polyploid survival are therefore of particular interest
when attempting to understand the opportunities and
bottlenecks of speciation.
A genetic model for such hybrid incompatibility was
proposed by Dobzhansky (1936) and thoroughly re-
viewed by Landry et al. (2007), whereby divergent loci
or alleles demonstrate incompatibility when merged in-
to the same nucleus. More recently, Tirosh et al. (2009)
performed a genome-wide analysis of gene expression
in yeast interspecic hybrids and analyzed interactions
between cis- and trans-acting regulatory factors and their
relation to gene expression rewiring in hybrid genomes.
Distinct expression patterns in the hybrid were shown
to account for both hybrid failure (Landry et al. 2007)
and hybrid vigor (Niet al. 2009; Birchler et al. 2010).
Whole-genome doubling of interspecic or interge-
neric hybrids gives rise to allopolyploid species, which
present the same type of novel intergenomic interac-
tions as those found in interspecic hybrids while also
providing fertility rescue. Plant allopolyploidy also offers
Supporting information is available online at http://www.genetics.org/
cgi/content/full/genetics.111.128348/DC1.
Sequence data from this article have been deposited with the GenBank
Data Libraries under accession no. GSE29243.
1
Corresponding author: Weizmann Institute of Science, Plant Sciences
Department, Rehovot, Israel, 76100. E-mail: avi.levy@weizmann.ac.il
Genetics 188: 263272 ( June 2011)
an additional level of opportunities for the newly
formed species, such as mutation buffering, sub- and
neofunctionalization of duplicated genes, xed hetero-
zygosity, and a broader range of dosage responses (Comai
2005; Doyle et al. 2008; Chen 2010). Most angiosperms
undergo one or more polyploidization events in their
lineage (Van DePeer et al. 2009) and 24% of all spe-
ciation events seem to occur via polyploidization (Otto
and Whitton 2000). Despite these potential advantages,
polyploidization events can lead to genomic shockdue
to gene redundancy, imbalanced and antagonistic gene
expression, orchestration of DNA replication among the
multiple and sometimes different genomes, and pairing
between multiple homologous and homeologous chro-
mosomes. Studies in several species have shown a broad
range of adverse genetic and epigenetic responses, which
occur soon after hybridization and polyploidization, in-
cluding DNA deletions, rearrangements or cytosine meth-
ylation, gene silencing, activation of transposons, and
modication of parental imprinting (Comai 2005;
Doyle et al. 2008; Chen 2010). Small RNAs have been
associated with all of these events (Matzke and Birchler
2005) and therefore, changes in small RNA species in
hybrids and polyploids might provide mechanistic in-
sight into the control of genetic and epigenetic changes
that occur in response to polyploidization.
Endogenous plant small RNAs can be divided into
several classes (Ghildiyal and Zamore 2009), where two
of the most prominent classes include the 21-nucleotide
(nt)-long species, mostly corresponding to microRNAs
(miRNA), and the small interfering RNAs (siRNAs),
typically 24 nt in length. Several dsRNA-mediated path-
ways operating at the level of the nuclear genome have
been described in plants and include RNA-directed DNA
methylation (Matzke and Birchler 2005) and small
RNA-mediated DNA demethylation (Penterman et al.
2007). Small RNAs have been shown to be involved in
a broad range of functions including heterochromatin
formation and silencing (Lippman and Martienssen
2004). siRNAs seem to function as guardians against
transposable elements (TEs) during plant development
(Levy and Walbot 1990; Hsieh et al. 2009; Gehrin
et al. 2009; Mosher et al. 1990; Slotkin et al. 2009).
However, their role in transposon regulation in the con-
text of hybridization or polyploidization has received far
less attention.
In this work, we performed a high-throughput screen
of small RNAs in parental, hybrid, and allopolyploid
plants mimicking the events that shaped the genome
during the speciation of bread wheat. We show that
miRNAs and repeat-derived siRNAs respond differently
to changes in ploidy level. In addition, the siRNA pools
corresponding to transposons were signicantly reduced
upon allopolyploidization. In parallel, we show that the
reduction or total disappearance of siRNAs correlated
with decreased CpG methylation of target transposons.
This deregulation of siRNAs, and the associated reduc-
tion in transposon methylation, may contribute to genome
instability and may hinder speciation via hybridization
and polyploidization.
MATERIALS AND METHODS
Plant material: Seeds were germinated on wet 3MM papers
in petri dishes. All seeds were imbibed for 24 hr at room
temperature under regular light. Seeds of Aegilops tauschii
then underwent 36 weeks of vernalization, at 4in the dark.
After germination, plantlets were placed in 5-liter pots and
grown in a greenhouse. All plants were grown during the
winter, at 20and under short-daylight conditions.
Pollen from the diploid wheat Aegilops tauschii, cultivar
TQ113 (genome DD, kindly provided by J. Dvorˇak), was used
to manually pollinate stigmas of the tetraploid wheat Triticum
turgidum ssp. durum, cultivar Langdon (TTR16, genome BBAA),
which served as the female parent, to generate BAD F1 hybrids.
Pollination was performed early in the morning, using freshly
collected pollen, 23 days after emasculation of the female-
parent owers. The hybrid plants were mostly sterile. F1 spikes
were bagged to prevent outcrossing. Occasionally, seeds were
obtained from the spikes of F1 plants. Typically, spikes from
F1 plants spontaneously gave zero to one seed, corresponding
to an average frequency of 1%, consistent with reports of
frequent occurrence of unreduced gametes, in similar interge-
neric hybrids of wheat (Kihara and Lilienfeld 1949). Allopoly-
ploid plants derived from these seeds were then characterized
(see RESULTS for details).
Preparation of metaphase chromosomes from wheat root
tips: To count chromosomes in F1 hybrids and in newly
synthesized amphiploids, root tips of germinated seeds were
cut, subjected to 24 hr cold treatment in double-distilled water
to arrest mitosis, and then transferred to a 2% acetocarmine
staining solution for 23 days. Before squashing, root tips were
briey heated in a solution containing 4 drops of 2% aceto-
carmine and 1 drop of 1 NHCl. Squashing was performed on
a preparative slide, in one drop of 2% acetocarmine.
RNA extraction: Total RNA was extracted from 2 g of whole
2-month-old plant tiller tissue, using 25 ml TRI reagent
(Molecular Research Center, Cincinnati, OH), according to
the manufacturers protocol, with the exception of isopropa-
nol precipitations and ethanol washes, which were performed
overnight, at 220. The plant material used in this study con-
sisted of six TTR16 plants, six TQ113 plants, three F1 hybrid
plants, and six newly synthesized S1 amphiploid plants, whose
polyploid nature had been previously conrmed by karyotype
analysis. Tissues from all genotypes were at the vegetative phase.
Small RNA library preparation: RNA quality of all samples
was veried using a bio-analyzer. Cloning of small RNAs was
performed by Illumina (San Diego, CA), using the DGE small-
RNA sample preparation kit protocol v1.0 (Illumina, Hayward,
CA). In brief, small RNAs (1832 nt long) were size selected,
puried, and ligated with 39and 59adaptors. Four cDNA librar-
ies of size-selected RNA from the four pools of plants were
prepared by reverse-transcription PCR, followed by PCR.
Bioinformatic analysis of small RNA high-throughput se-
quencing data: Eighteen million reads were obtained from
high-throughput sequencing of small RNAs from the four
libraries. Within each library, the unique sequences were re-
ported as tags containing sequence information and frequency.
There were 9.8 million tags in total for all four libraries. A script
was developed to trim the 39adaptor: in the rst step, the rst 7
bases from the adaptor and any downstream sequence were
removed. Then the process was repeated allowing one mis-
match in the adaptor sequence. Sequences that contained
264 M. Kenan-Eichler et al.
more than 3 Ns were also eliminated. Tag frequencies were
recalculated after these steps and resulted in 8.9 million tags.
Following quality control, tags with a frequency of at least
30 reads were assembled into short contigs using the Staden
sequence analysis programs(Staden 1996) normal shotgun
assemblymodule with a minimal initial match of 20 bases,
allowing 5% maximal mismatch, and then allowing additional
assembly with 17 bases with no mismatch. Eventually, 21,787
tags were assembled into 6892 contigs. The number of reads
and tags in each library is summarized in supporting informa-
tion,Table S1. The Staden program produces a le, which
reveals the relationship between the contigs and the sequen-
ces and thus helps determine the contigs composition and
library frequencies. Contigs were annotated using Eland ver-
sion 0.3 and BLAST 2.2.17 (Altschul et al. 1990). Searches
were performed against the following databases: NCBI nr/nt
( July 2008), the rice genome (ftp://ftp.plantbiology.msu.
edu/pub/data/Eukaryotic_Projects/o_sativa/annotation_dbs/
pseudomolecules/version_5.0 and version_6.0), wheat repeats
(http://www.tigr.org/tdb/e2k1/tae1/wheat_downloads.shtml
wheat repeat database), T. aestivum release 2 (ftp://ftp.tigr.
org/pub/data/plantta/Triticum_aestivum), T. turgidum release
1 (ftp://ftp.tigr.org /pub/data/plantta/Triticum_turgidum),
and mature miRNA (http://microrna.sanger.ac.uk/sequences/).
Specically, for miRNA detection, Eland_21 (seed of 21 ba-
ses) was run against plant miRNAs using a script that slides
awindowof21basesalongthecontigsequence.OnlymiR-
NAs with no mismatches were reported. Using the output of
BLASTn table against the TIGR wheat repeat database, the
contig hits were divided into several subcategories: retroele-
ments, DNA transposons, repetitive element, and unclassi-
ed repeats.
Contig frequency was calculated by summing the frequency
of each tag comprising the contig. Interlibrary normalization
was performed by dividing the contig frequency count per
library, by a factor that was calculated by dividing by the total
number of reads from tags with more than 30 reads. Therefore,
the calculation represents the frequency per million reads. A
rigorous signicance test was performed according to Audic
and Claverie (1997). For miRNA counting, all contigs with
the same seed sequence were summed.
Northern blot analyses: Northern blot analyses were per-
formed as previously described (Brown and Mackey 2001).
Briey, 15 mg of total RNA was loaded on formamide-agarose
gels and later transferred to a Hybond N
1
nylon membrane
(Amersham) via the standard wet transfer method. The Wis2-
1A LTR probe was prepared using the forward primer 59
TGTTGGAAATATGCCCTAGAG 39and the reverse primer
59GCACAATCTCATGGTCTAAGG 39; the Veju LTR probe
was prepared using the forward primer 59TAGAATAAATC
CGAGGCATACC 39and the reverse primer 59TTAGGTTA
CAGTTGGACTTGG 39, for amplication from the T. aestivum
var. Chinese Spring genome.
Probes were labeled with [a-
32
P]dCTP (Amersham) using
Klenow DNA polymerase (Fermentas). The membranes were
blotted overnight with 20 pmol of the labeled probe and then
exposed overnight to a phosphorimager screen, and images
were visualized using the Image Gauge program. Membranes
were stripped for reuse, in 0.1% SDS and veried to give no
signal.
Bisulte sequencing: Hybrid and polyploid genomic DNA
was extracted from 100 mg of leaf tissue, using the GeneElute
Plant genomic DNA miniprep kit (Sigma). Two micrograms of
genomic DNA suspended in 0.5 ml double-distilled water was
sheared on ice by ultrasound (Microson sonicator, Misonix)
set at 3 W output power, four pulses of 10 sec each with 50-sec
intervals between pulses. Sheared DNA was concentrated to
anal volume of 20 ml, using a speed vacuum set at 60for
2 hr. The genomic DNA was then subjected to bisulte conver-
sion and purication using the EpiTect Bisulte kit (Qiagen),
according to the manufacturers instructions. Bisulte ions
convert nonmethylated cytosine residues to uracil residues.
Fifty nanograms of puried DNA were then subjected to
PCR, using degenerate primers (Metabion), designed to am-
plify the Veju retrotransposon LTR (Veju degenerate forward
primer, 59AGTGAATGTYAAGTTGTTGGTG 39;Veju degener-
ate reverse primer, 59TCRAACAACCTARCTCATRATAC 39).
The PCR products were then separated on a 2% agarose gel
and puried using a PCR purication kit (RBC). Cleaned PCR
products were ligated to a pGem-T easy vector (Promega),
which were then used to transform Escherichia coli (top 10
strain); transformed bacteria were then plated on selective
plates. Plasmid DNA was extracted from 25 bacterial colonies
for each plant and then sequenced. Sequences were analyzed
using the Sequencher 4.9 program, and methylation patterns
were analyzed using the Kismeth program http://katahdin.
mssm.edu/kismeth/revpage.pl (Gruntman et al. 2008).
RESULTS
Phenotypic analysis of hybrid and polyploid wheat:
To gain insight into the molecular mechanisms leading
to the genetic and epigenetic changes occurring upon
hybrid and polyploid formation, the small RNA proles
of parental wheat lines were compared to those of the
derived hybrid and the rst allopolyploid generation.
More specically, the four wheat species analyzed
included the parental tetraploid T. turgidum ssp. durum
(genome BBAA) and the diploid Ae. tauschii (genome
DD), their synthetic triploid hybrid (genome BAD),
and their derived hexaploid (genome BBAADD). This
synthetic hexaploid is analogous in genome structure to
bread wheat T. aestivum ssp. aestivum. Parental lines were
self-pollinated for several generations, with spike bag-
ging to prevent cross-pollination, and were thus consid-
ered inbred. Hybrid seeds were obtained by crossing
T. turgidum and Ae. tauschii. The hybrid F1 seeds were
obtained from the female tetraploid plants and were
found to be very small and shriveled compared to their
parents (Figure 1). Upon germination, F1 seeds gave rise
to F1 plants bearing necrotic leaves (Figure 1), a typical
hybrid incompatibility phenotype (Bomblies and Weigel
2007). Some F1 plants died shortly after germination.
Nevertheless, those that survived were vigorous and fea-
tured spikes larger than their parents (Figure 1), but
most were sterile. Occasionally, seeds were spontaneously
obtained from the spikes of F1 plants (See materials
and methods) and were termed S1.While these
seeds were slightly shriveled, they were larger in size
than the seeds of either parent (Figure 1). Plants ger-
minating from S1 seeds has a duplicated genome as
conrmed by karyotype analysis (Figure S1). The result-
ing S1 plants were fertile and bore spikes that yielded S2
seeds, which were less shriveled than, but similar in size
to, S1 seeds. While S1 plant leaves were highly necrotic,
the plants were vigorous and featured spikes larger than
those in the hybrid or in the parents and leaves similar
in size to those of the tetraploid parent (Figure 1).
Deregulation of Small RNAs 265
High-throughput sequencing of small RNAs: Total
RNA was isolated from tillers of 2-month-old wheat
plants, which primarily included somatic tissues, namely
leaves and stems and some meristematic tissues. A small
RNA library was prepared for each of the analyzed
wheat types and was composed of a pool of RNA col-
lected from three to six plants of each genotype. Eigh-
teen million reads were obtained from small RNA
sequencing of the four libraries (Table S1). The unique
sequences within each library were reported as tags
(containing sequence information and frequency).
These tags underwent strict quality checks (Figure S2
and materials and methods), and only tags that had
30 reads or more were included in the analysis, due to
lack of a fully sequenced reference genome. As the
data from a number of small RNA sequencing experi-
ments proved highly reproducible (with a correlation of
Spearmansr¼0.950.98) for tags with $30 reads
(Fahlgren et al. 2009), this number was set as the thresh-
old for data analysis. Of the 18 million reads obtained, 6
million high-quality reads, corresponding to 21,787
unique sequence tags, were analyzed. The small RNA
sizes ranged between 18 and 35 nt and included the
two prominent classes of 21 nt- and 24 nt-long small RNAs
(Figure 2). The 21-nt class corresponds mainly to miRNA,
while the 24-nt class corresponds most likely to siRNAs
(Ghildiyal and Zamore 2009). The 24-nt-long small
RNAs were most abundant within the two parental and
and the hybrid libraries, whereas, the 21-nt-long small
RNAs were most prevalent in the allopolyploid library.
The various small RNAs obtained from tags with $30
reads were categorized according to sequence similarity
(Figure S3). miRNAs were one of the most abundant
small RNA species (2144%, depending on the geno-
type). Small RNAs that matched repeats, which were
largely TEs, represented 12% of the total reads for
all libraries, aside from the allopolyploid library, where
a signicant decrease (P(x
2
),0.0001) to only 6% was
observed. Surprisingly, tRNAs comprised 9% of total
hits in the parental and hybrid libraries and rose to 20%
(P(x
2
),0.0001) in the allopolyploid library (Figure S3).
Small RNA contigs matching tRNA genes ranged in size
between 18 and 36 bases. However, 55% of the small
RNA tags corresponding to tRNAs showed distinct peaks
at 19, 20, and 21 nt, suggesting that they were not deg-
radation products of larger tRNA molecules. No excep-
tional RNA degradation was observed in any of the
samples, hence degradation could not explain the higher
tRNA frequency observed in allopolyploid samples.
Abundance of miRNAs in parent, hybrid and allo-
polyploid libraries: Twenty-one known plant miRNA
sequences were identied with high certainty, i.e., with
$30 reads each, and were identical to known miRNAs
from other plant species (Table S2). For example,
miR168, the most abundant miRNA in rice (Luet al.
2008), was the most abundant in the present data. In-
terestingly, the amount of miRNA, relative to total small
RNA, increased with increasing levels of ploidy, being
the lowest (21%) in the diploid Ae. tauschii DD genome,
33% in the triploid hybrid genome (BAD), 38% in the
tetraploid T. turgidum BBAA genome, and 44% in the
synthetic hexaploid BBAADD genome (Figure 3).
Figure 1.Parents, hybrids, and allopolyploid plant mate-
rial. Triticum turgidum durum (female parent, genome BBAA)
and Aegilops tauschii (male parent, genome DD) leaves, spikes,
and seeds are shown at the top left and top right, respectively.
Plant crossing yielded small and shriveled hybrid (F1) seeds
(genome BAD, in the center), from which hybrid F1 plants
were grown. F1 plants featured necrotic leaves and were mostly
sterile, occasionally and spontaneously giving rise to allopoly-
ploid (S1) seeds (middle), with a doubled chromosome num-
ber. S1 seeds were larger than the seeds of both parents and
were somewhat shriveled. S1 plants (bottom), derived from S1
seeds, bore leaves as broad as the maternal plant, spikes larger
than those of both parents, and yielded S2 seeds (bottom left),
which were as large as S1 seeds, and a little less shriveled. Scale
bar, 0.5 cm.
266 M. Kenan-Eichler et al.
To assess the changes in the abundance and prole
of miRNAs expressed as a result of hybridization and
polyploidization, we compared the normalized number
of hits (in reads per million) found in the hybrid and
the polyploid libraries to the average number of hits in
the two parental libraries, namely the midparent value
(MPV). Under the assumption of additive expression,
the number of hits in F1 or S1 should be similar to the
midparent value: log
2
(F1/MPV) or log
2
(S1/MPV) ¼0.
Positive log
2
values indicate that the number of small
RNAs for a given tag is higher in the hybrid or the
polyploid than in the parent average; negative values
indicate the opposite. When applying a cutoff value in
which jlog
2
j,0.5 is considered similar to MPV, most
miRNAs were expressed to similar degrees as the MPV
(Figure 4). However, miR390a was underrepresented
(log
2
(F1/MPV) ¼24.8), and miR160f was overrepre-
sented (log
2
(F1/MPV) ¼4.86) when comparing hybrid
miRNA expression proles to those of the parent librar-
ies. Interestingly, both miR390 and miR160 play key
roles in auxin regulation, via regulation of TAS3, a
trans-acting siRNA involved in auxin signaling (Fahlgren
et al. 2006) by the former and via targeting of ARF10
(Liu et al. 2007) by the latter. In the allopolyploid
plants, several miRNA expression patterns signicantly
deviated from those observed in the parental lines.
miR157a, miR160f, miR159a, miR396d, and miR168
were signicantly overrepresented relative to MPV, while
miR156b, miR165a, and miR1135 were signicantly un-
derrepresented. miR168 had the highest number of
reads in almost all libraries, amounting to 25% of the
polyploid library reads and 119,710 and 79,244 reads in
T. turgidum and Ae. tauschii libraries, respectively (Table
S2). It was overrepresented by 1.27-fold in the hybrid
and by 2.46-fold in the synthetic polyploid, relative to
the midparent value (MPV ¼99,477) and is suggested
to have genome-wide effects on small RNA species (see
discussion). miR156 was the second-highest tag ex-
pressed in the model, while miR1135 was only moderately
expressed (Table S2). miR156a was overrepresented by
2.4-fold in the polyploid relative to the MPV. miR156
and its related sequence is of particular interest, as its
overexpression has been suggested to prolong the vegeta-
tive phase and to delay owering in Arabidopsis (Schwarz
et al. 2008) and may explain the heterotic effects seen here
in the allopolyploid plants.
siRNAs that correspond to repeats: Small 24-nt-long
RNAs were abundant in all the libraries, where a large
percentage of these siRNAs corresponded to repeats:
1113 Staden contigs out of a total of 6892 correspond
to transposons and retrotransposons. An additional
130 contigs corresponded to ribosomal RNA genes and
telomeric and centromeric repeats. siRNAs matching
known genes often corresponded to repeats as well.
Here, we focused on repeats, using the complete wheat
repeat database (http://wheat.pw.usda.gov/ITMI/Repeats/
atle.total) as reference.
Transposons, which make up approximately 80% of
the wheat genome (Charles et al. 2008), comprise a
major class of repeats. Recently, small RNAs have been
mapped to the repeat sequences of the Triticeae repeats
database (Cantu et al. 2010). This work demonstrates
the changes in TE-related small RNAs following hybrid-
ization and polyploidization and conrms that TEs are
targets of small RNAs, particularly at their termini. In
contrast to miRNAs, the amount of siRNAs correspond-
ing to transposons decreased with increased ploidy (Fig-
ure 3). In the hybrid, 42% of siRNAs corresponding to
transposons were represented similarly to the MPV (de-
ned here as jlog
2
j,0.5), while 39% were underrepre-
sented and 19% were overrepresented. In the allopolyploid,
a massive reduction and signicant shift (P(x
2
),0.0001)
in the relative abundance of siRNAs corresponding to
transposons was observed, with 85% of the hits below
the MPV (Figure 5). This underrepresentation affected
Figure 2.Length distribution of small RNAs. Length dis-
tribution of small RNAs in the four libraries. Small RNAs
ranged in size from 18 to 35 nucleotides, with two prominent
peaks at 21 and 24 nucleotides. Note that the x-axis is not
linear at the end.
Figure 3.Changes in small RNA expression levels as
a function of plant ploidy. Percentage of siRNAs correspond-
ing to repeats, and of miRNA reads, as a function of genome
ploidy (where X is a haploid genome) and genomic compo-
sition. The percentage of reads was determined from the total
number of small RNA reads in each library, thus enabling
a comparison between the different ploidy groups.
Deregulation of Small RNAs 267
both DNA elements and retroelements (Table 1). Small
RNA matching to the sequenced and annotated T. aestivum
BAC genome region (GenBank accession no. CT009735),
using IGB (Integrative Genome Browser), conrmed the
massive reduction of small RNAs corresponding to trans-
posons, in the polyploid compared to the parental lines
(Figure 6). Note that in the data used by Cantu et al.
(2010), the number of reads corresponding to Veju in
natural hexaploid wheat was very low (1/400,000), sim-
ilar to the data shown here for the synthetic hexaploid,
but contrasting with the present data collected from
natural tetraploid and diploid wheat.
We then examined whether the reduction in TE-
related siRNAs correlated with hallmarks of transposons
activation. Table 1 summarizes several examples of ret-
rotransposons and DNA transposons for which small
RNAs were underrepresented in the polyploid com-
pared to the parental lines. We then tested whether
siRNA underrepresentation correlated with transcrip-
tional activation of the copia-like retrotransposon Wis2-
1A and of Veju,aterminal-repeat retrotransposon in
miniature(TRIM) element (Sanmiguel et al. 2002).
In both cases, the allopolyploid expressed higher tran-
script levels than the hybrid, further substantiating the
observed reduction in corresponding siRNAs. However,
the tetraploid parent also expressed high transcript lev-
els, thus, no straightforward correlation can be drawn
between small RNA and corresponding mRNA levels,
detected by Northern hybridization (Figure S4).
Cytosine methylation represents an additional hall-
mark of transposon activity and has been reported to be
reduced in active and hypermethylated in silent trans-
posons (Chandler and Walbot 1986). As a case study,
we performed bisulte sequencing of the LTR of a Veju
element, using the available sequence from a transposon-
rich region in the T. aestivum genome (GenBank acces-
sion no. CT009735, coordinates 38,98939,363). Seven
different small RNAs, corresponding to hundreds of
reads, were mapped along the LTR sequence (Figure
7, middle) and shown to derive from both strands. All
seven were almost fully suppressed in the allopolyploid
line (Figure 7, top). Sequencing of 25 different Veju-
LTR clones per plant type, following bisulte conver-
sion, enabled us to determine the average methylation
of Veju elements similar in sequence to the T. aestivum
LTR (GenBank accession no. CT009735; see meth-
ods). A decrease in CG methylation (Figure 7, bottom)
but not in CHG or CHH methylation (data not shown),
was observed in the allopolyploid line and correlated
with decreased abundance of Veju LTR small RNAs in
Figure 4.miRNA abundance in the
hybrid and polyploid lines. miRNA
abundance in the hybrid (F1, gray)
and the polyploid (S1, black) relative
to the midparent value (MPV) is shown
as a logarithmic cumulative distribution
function (n¼21). Positive values indi-
cate higher abundance in the hybrid or
polyploid compared to the MPV, nega-
tive values indicate the opposite and
values close to zero indicate similar
abundance between the analyzed and
parental lines.
Figure 5.Abundance of small RNAs correspond-
ing to transposons. The abundance in the hybrid (F1,
gray) and the polyploid (S1, black) relative to the
midparent value is shown as a logarithmic cumulative
distribution function (n¼1113). The box on the top
left summarizes the abundance relative to the midpar-
ent value. In the hybrid, 42% of the small RNAs cor-
responding to transposons had an abundance similar
to the midparent (jlog
2
j,0.5), 19% were overrepre-
sented (log
2
.0.5), and 39% were underrepresented
(log
2
,20.5). In the polyploid, 12% of the small RNAs
corresponding to transposons had an abundance simi-
lar to the midparent value, 3% were overrepresented,
and 85% were underrepresented, compared to
the midparent value. This underrepresentation
was highly signicant P(x
2
),0.0001.
268 M. Kenan-Eichler et al.
the polyploid. These results are consistent with and fur-
ther elucidate data reported by Kraitshtein et al.
(2010) with regard to AFLP-based Veju methylation.
DISCUSSION
Small RNAs are involved in a number of key cellular
processes, and perturbations in their steady-state ex-
pression levels can lead to genome-wide changes in
gene expression or in chromatin and genome structure.
Modied small RNA expression levels have been widely
reported in the context of developmental regulation of
plants, but to a lesser extent in the context of speciation
via interspecic hybridization and allopolyploidization
(Haet al. 2009). Haet al. (2009) describe an allotetra-
ploid hybrid, analogous to Arabidopsis suecica, formed
between a synthetic Arabidopsis autotetraploid line
and the natural Arabidopsis tetraploid, A. arenosa, and
report signicant deviations in hybrid miRNA expres-
sion from MPV. In this study, most hybrid miRNA ex-
pression proles did not deviate from their MPVs
(Figure 4). However, a number of deviations were found
in the allopolyploid (Table S2). In addition, some classes,
such as miR156 and miR166, included several sequence
variants, each demonstrating differential expression pat-
terns. However, in the absence of the wheat genome
sequence, it is difcult to associate changes in miRNA
levels to the expression of target genes or to plant
phenotypes.
Another difference between the present model and
that described in the Arabidopsis allopolyploid analysis
(Haet al. 2009) lies in the nature of the genetic material
analyzed, where the Arabidopsis work involved plants
of the same ploidy level (parents and hybrid were all
tetraploids), while this study encompassed lines of
different ploidy levels (parents [2·and 4·], hybrid
[3·], and a derived allohexaploid [6·] analogous to
natural bread wheat). The wheat material offered
unique and novel insight into an unexpected response
to changes in ploidy levels. Remarkably, and unexpect-
edly, the relative amount of small RNAs corresponding
to miRNAs increased with ploidy (Figure 3). Inversely,
the relative amount of 24-nt small RNAs corresponding
to transposons decreased with increased ploidy (Figure
3). Similarly, in the work of Cantu et al. (2010), the
count per basepair of small RNAs corresponding to
transposons was lower in hexaploid than in tetraploid
wheat varieties (Cantu et al. 2010). This ploidy depen-
dence seems to be insensitive to genomic composition,
but sensitive to dosage. Indeed, while the diploid, tetra-
ploid, and hexaploid wheat had different genomic
TABLE 1
Number of reads (normalized to reads per million) of siRNAs
corresponding to TEs, in parents, hybrid, and rst-generation
allopolyploid (S1)
TEs
a
durum
BBAA
tauschii
DD F1 BAD
S1
BBAADD
Wis2-1A 756 528 475 0
Latidu 1267 486 747 94
Veju 326 211 400 6
WHAM 83 2646 420 69
Caspar 701 315 472 8
Thalos 532 656 753 5
a
Caspar and Thalos are DNA transposons, and other TEs are
retroelements.
Figure 6.Viewing of small RNAs on a BAC sequence. A view of small RNAs mapped to the T. aestivum BAC CT009735, using
the IGB tool from Affymetrix, with coordinates indicating the nucleotide number (nt). Horizontal gray bars correspond to trans-
posable elements, and vertical black lines indicate the abundance of corresponding small RNAs in each library.
Deregulation of Small RNAs 269
composition, the triploid (hybrid) and allohexaploid,
which both featured the same three wheat genomes
(A, B, and D) but at varying doses (3·vs. 6·), expressed
divergent small RNAs proles.
Mechanistically, a correlation can be drawn between
specic miRNA overrepresentation and repressed siR-
NAs formation. More specically, miR168 overrepresen-
tation in polyploids may account for ARGONAUTE 1 (AGO1)
suppression (Mallory and Vaucheret 2009), which
in turn reduces the production of siRNAs (Vaucheret
2008). Although, the regulation of AGO1 is quite com-
plex (Mallory and Vaucheret 2009), it is conceivable
that affecting regulatory loops responsible for AGO1 ac-
tivity, e.g., via alterations in miR168, may bear genome-
wide effects on siRNAs. In fact, the novel intergenomic
interactions and novel dosage effects demonstrated in
the allopolyploid lines may account for the disruption
of the activity of several genes related to small RNA
expression and stability. Previously, we have shown that
novel interactions between cis- and trans-acting regula-
tory elements can lead to overexpression or, alterna-
tively, to suppression of 10% of the genes of an
interspecic hybrid (Tirosh et al. 2009). Deregulation
of central genes involved in the gene silencing machin-
ery could account for some of the alterations in small
RNA proles reported here, which in turn can affect TE
activities, as discussed below.
siRNAs play a critical role in heterochromatin mainte-
nance and in transposable element silencing (Zaratiegui
et al. 2007). The strong decrease in the percentage of
siRNAs corresponding to TEs reported here could thus
lead to transposon activation. Transposons, notoriously
responsive to genomic shocks(Mcclintock 1984),
may be responsive to hybridization- and polyploidization-
induced shocks,linked with downregulation of small
RNAs. Indeed, it has been reported that silent trans-
posons and retrotransposons may become transcrip-
tionally and sometime transpositionally active upon
hybrid and polyploid formation (Kashkush et al. 2003;
Madlung et al. 2005; Chen and Ni2006; Kraitshtein
et al. 2010). The correlation between small RNAs cytosine
methylation and transposon silencing (Matzke et al.
2007) may further explain transposon activation follow-
ing small RNAs-related demethylation, as reported here
for the Veju TE.
A fascinating aspect of TE regulation involves their
capacity to transit between inactive and active states.
Remarkably, a hypomethylated TE can become reme-
thylated within one generation, as shown here for a
specic element (Figure 7) and as shown on a genome-
wide scale for the Veju element (Kraitshtein et al.
2010). An attractive model for such kinetics is demon-
strated via the disappearance of small RNAs, as seen
here in S1 for several elements, including Wis2-1A (Table
1), which then leads to transcription of otherwise silent
TEs. As a result, high levels of aberrant and dsRNAs
may be generated, as many TEs are nested within each
other in opposite orientations (Sanmiguel et al. 1996).
In the subsequent generation, small RNAs produced
by such transcriptional activation, might reverse the
element to its original methylated and silenced state.
This proposed mechanism may account for the
reported maintenance silencing of repeats by small
RNAs produced by RNA polymerases IV and V
(Pikaard et al. 2008; Matzke et al. 2009). In other
words, TE activation may catalyze TE silencing via up-
regulation of small RNA production.
However, the association between small RNAs and TE
transcription still remains elusive. In an earlier work
(Kashkush et al. 2003), we analyzed an allopolyploid
resulting from a cross between two diploid parents and
measured S1 transcriptional activation of transposons
that were silent in the parents. The present model, us-
ing different parental species, is a bit more complex
Figure 7.Veju-related small RNAs and CpG methylation.
Top: The abundance of small RNAs that correspond to the
Veju retrotransposon LTR found on BAC CT009735 for the
midparent value, the hybrid (F1), and the polyploid (S1).
gDNA: The coordinates of small RNAs on the Veju LTR geno-
mic DNA. Bottom: The percentage of CpG methylation was
determined by bisulte sequencing for the Veju LTR region. A
decrease in CpG methylation in the rst generation of the
polyploid (S1), compared to Ae.s tauschii (genome DD), was
observed. In the second polyploid generation (S2), CpG meth-
ylation increased. A natural hexaploid (Genome BBAADD,
variety Chinese Spring) is shown for comparison.
270 M. Kenan-Eichler et al.
(Figure S4), where S1 transposon transcript levels were
higher than those in F1, substantiating an earlier report
of transcriptional activation in the S1 generation
(Kashkush et al. 2003). However, high levels of both
small RNAs and TE transcripts were found in the tetra-
ploid parent (Figure S4), which may be due to transcrip-
tion of a divergent subfamily of transposons, not targeted
by the small RNAs, or to alteration in small RNA mobil-
ity, processing, or other unknown phenomenon.
Another intriguing question remaining relates to the
changes measured in allopolyploid small RNA species,
and to a lesser extent in the hybrid. This observation
may be linked to polyploidy per se. Alternatively, this may
be a result of the kinetics of small RNA alterations,
which begin in the hybrid and peak in the S1 genera-
tion. Another interesting possibility is that global deme-
thylation and activation of TEs may occur through
disruption of developmentally regulated processes in
the germline. Such modications may occur during
meiosis, or gametogenesis, or during the development
of an embryo derived from deregulatedgametes. In
such cases, massive changes would be found in the S1
only, and not in the F1.
In summary, the data reported here, together with
previous ndings, suggest that deregulation of small
RNAs may stimulate TE activation in interspecic
hybrids and allopolyploids. This phenomenon is remi-
niscent of hybrid dysgenesis in Drosophila (Malone
and Hannon 2009), a mechanism of incompatibil-
ity that can hinder speciation via hybridization and
polyploidization. The correlation between the different
hallmarks of transposon activation (small RNAs, hypome-
thylation, transcription, and transposition) is complex,
as they undergo rapid changes between generations. In
addition, it is important to note that this study was per-
formed on plants that survived hybridization and allo-
polyploidization. The events of seedling death and seed
abortion, which were observed in the hybrid and allo-
polyploid lines, may have occurred as a result of severe
transposon activation and were not considered in our
analysis.
We thank Naomi Avivi-Ragolski and other members of the Levy
laboratory for their help and discussions; Idan Efroni for help in data
analysis; Yehudit Posen for editing the manuscript; and members of
the bioinformatics unit, especially Jaime Prilusky for writing the quality-
control script and Ester Feldmesser for help with statistical analysis. This
work was funded by a grant from the Israeli Science Foundation, no.
616/09, to A.A.L.
LITERATURE CITED
Altschul,S.F.,W.Gish,W.Miller,E.W.Myers and D. J. Lipman,
1990 Basic local alignment search tool. J. Mol. Biol. 215: 403410.
Audic, S., and J. M. Claverie, 1997 The signicance of digital gene
expression proles. Genome Res. 7: 986995.
Birchler, J. A., H. Yao,S.Chudalayandi,D.Vaiman and R. A.
Veitia, 2010 Heterosis. Plant Cell 22: 21052112.
Bomblies, K., and D. Weigel, 2007 Hybrid necrosis: autoimmunity
as a potential gene-ow barrier in plant species. Nat. Rev. Genet.
8: 382393.
Bretagnolle, F., and J. D. Thompson, 1995 Tansley Review No. 78:
gametes with the somatic chromosome number: mechanisms of
their formation and role in the evolution of autopolyploid plants.
New Phytol. 129: 122.
Brown, T., and K. Mackey, 2001 Analysis of RNA by northern blot
hybridization. Curr. Protoc. Hum. Genet. 3(Appendix): 3K.
Cantu, D., L. S. Vanzetti,A.Sumner,M.Dubcovsky,M.Matvienko
et al., 2010 Small RNAs, DNA methylation and transposable ele-
ments in wheat. BMC Genomics 11: 408.
Chandler, V. L., and V. Walbot, 1986 DNA modication of a maize
transposable element correlates with loss of activity. Proc. Natl.
Acad. Sci. USA 83: 17671771.
Charles, M., H. Belcram,J.Just,C.Huneau,A.Viollet et al.,
2008 Dynamics and differential proliferation of transposable
elements during the evolution of the B and A genomes of wheat.
Genetics 180: 10711086.
Chen, Z. J., 2010 Molecular mechanisms of polyploidy and hybrid
vigor. Trends Plant Sci. 15: 5771.
Chen, Z. J., and Z. F. Ni, 2006 Mechanisms of genomic rearrange-
ments and gene expression changes in plant polyploids. Bioessays
28: 240252.
Comai, L., 2005 The advantages and disadvantages of being poly-
ploid. Nat. Rev. Genet. 6: 836846.
Dobzhansky, T., 1936 Studies on hybrid sterility. II. Localization of ster-
ility factors in Drosophila pseudoobscura hybrids. Genetics 21: 113135.
Doyle, J. J., L. E. Flagel,A.H.Paterson,R.A.Rapp,D.E.Soltis
et al., 2008 Evolutionary genetics of genome merger and dou-
bling in plants. Annu. Rev. Genet. 42: 443461.
Fahlgren, N., T. A. Montgomery,M.D.Howell,E.Allen,S.K.
Dvorak et al., 2006 Regulation of AUXIN RESPONSE FACTOR3
by TAS3 ta-siRNA affects developmental timing and patterning in
Arabidopsis. Curr. Biol. 16: 939944.
Fahlgren, N., C. M. Sullivan,K.D.Kasschau,E.J.Chapman,J.S.
Cumbie et al., 2009 Computational and analytical framework for
small RNA proling by high-throughput sequencing. RNA 15:
9921002.
Gehring, M., K. L. Bubb and S. Henikoff, 2009 Extensive deme-
thylation of repetitive elements during seed development under-
lies gene imprinting. Science 324: 14471451.
Ghildiyal, M., and P. D. Zamore, 2009 Small silencing RNAs: an
expanding universe. Nat. Rev. Genet. 10: 94108.
Gruntman, E., Y. Qi,R.K.Slotkin,T.Roeder,R.A.Martienssen
et al., 2008 Kismeth: analyzer of plant methylation states
through bisulte sequencing. BMC Bioinformatics 9: 371.
Ha, M., J. Lu,L.Tian,V.Ramachandran,K.D.Kasschau et al.,
2009 Small RNAs serve as a genetic buffer against genomic
shock in Arabidopsis interspecic hybrids and allopolyploids.
Proc. Natl. Acad. Sci. USA 106: 1783517840.
Hsieh, T. F., C. A. Ibarra,P.Silva,A.Zemach,L.Eshed-Williams
et al., 2009 Genome-wide demethylation of Arabidopsis endo-
sperm. Science 324: 14511454.
Ishikawa, R., and T. Kinoshita, 2009 Epigenetic programming: the
challenge to species hybridization. Mol. Plant 2: 589599.
Kashkush, K., M. Feldman and A. A. Levy, 2003 Transcriptional
activation of retrotransposons alters the expression of adjacent
genes in wheat. Nat. Genet. 33: 102106.
Kihara, H., and F. Lilienfeld, 1949 A new synthesized 6x-wheat,
pp. 307319 in Proceedings of Eighth International Congress of Genet-
ics, edited by G. B. A. R. Larsson. Hereditas Supplement, Stock-
holm, Sweden.
Kraitshtein, Z., B. Yaakov,V.Khasdan and K. Kashkush,
2010 Genetic and epigenetic dynamics of a retrotransposon
after allopolyploidization of wheat. Genetics 186: 801812.
Landry, C. R., D. L. Hartl and J. M. Ranz, 2007 Genome clashes in
hybrids: insights from gene expression. Heredity 99: 483493.
Leitch, A. R., and I. J. Leitch, 2008 Genomic plasticity and the
diversity of polyploid plants. Science 320: 481483.
Levy, A. A., and V. Walbot, 1990 Regulation of the timing of trans-
posable element excision during maize development. Science
248: 15341537.
Lippman, Z., and R. Martienssen, 2004 The role of RNA interfer-
ence in heterochromatic silencing. Nature 431: 364370.
Deregulation of Small RNAs 271
Liu,P.P.,T.A.Montgomery,N.Fahlgren,K.D.Kasschau,H.
Nonogaki et al., 2007 Repression of AUXIN RESPONSE
FACTOR10 by microRNA160 is critical for seed germination
and post-germination stages. Plant J. 52: 133146.
Lu, C., D. H. Jeong,K.Kulkarni,M.Pillay,K.Nobuta et al.,
2008 Genome-wide analysis for discovery of rice microRNAs re-
veals natural antisense microRNAs (nat-miRNAs). Proc. Natl.
Acad. Sci. USA 105: 49514956.
Madlung, A., A. P. Tyagi,B.Watson,H.M.Jiang,T.Kagochi et al.,
2005 Genomic changes in synthetic Arabidopsis polyploids.
Plant J. 41: 221230.
Mallet, J., 2007 Hybrid speciation. Nature 446: 279283.
Mallory, A. C., and H. Vaucheret, 2009 ARGONAUTE 1 homeo-
stasis invokes the coordinate action of the microRNA and siRNA
pathways. EMBO Rep. 10: 521526.
Malone, C. D., and G. J. Hannon, 2009 Small RNAs as guardians of
the genome. Cell 136: 656668.
Matzke, M. A., and J. A. Birchler, 2005 RNAi-mediated pathways
in the nucleus. Nat. Rev. Genet. 6: 2435.
Matzke, M., T. Kanno,B.Huettel,L.Daxinger and A. J. Matzke,
2007 Targets of RNA-directed DNA methylation. Curr. Opin.
Plant Biol. 10: 512519.
Matzke, M., T. Kanno,L.Daxinger,B.Huettel and A. J. Matzke,
2009 RNA-mediated chromatin-based silencing in plants. Curr.
Opin. Cell Biol. 21: 367376.
McClintock, B., 1984 The signicance of responses of the genome
to challenge. Science 226: 792801.
Mosher,R.A.,C.W.Melnyk,K.A.Kelly,R.M.Dunn,D.J.Studholme
et al., 2009 Uniparental expression of PolIV-dependent siRNAs
in developing endosperm of Arabidopsis. Nature 460: 283286.
Ni, Z., E. D. Kim,M.Ha,E.Lackey,J.Liu et al., 2009 Altered
circadian rhythms regulate growth vigour in hybrids and allopo-
lyploids. Nature 457: 327331.
Otto, S. P., and J. Whitton, 2000 Polyploid incidence and evolu-
tion. Annu. Rev. Genet. 34: 401437.
Penterman, J., D. Zilberman,J.H.Huh,T.Ballinger,S.Henikoff
et al., 2007 DNA demethylation in the Arabidopsis genome.
Proc. Natl. Acad. Sci. USA 104: 67526757.
Pikaard,C.S.,J.R.Haag,T.Ream and A. T. Wierzbicki, 2008 Roles of
RNA polymerase IV in gene silencing. Trends Plant Sci. 13: 390397.
Rieseberg, L. H., and J. H. Willis, 2007 Plant speciation. Science
317: 910914.
SanMiguel,P.,A.Tikhonov,Y.K.Jin,N.Motchoulskaia,D.
Zakharov et al., 1996 Nested retrotransposons in the inter-
genic regions of the maize genome. Science 274: 765768.
SanMiguel, P. J., W. Ramakrishna,J.L.Bennetzen,C.S.Busso and
J. Dubcovsky, 2002 Transposable elements, genes and recom-
bination in a 215-kb contig from wheat chromosome 5A(m).
Funct. Integr. Genomics 2: 7080.
Schwarz,S.,A.V.Grande,N.Bujdoso,H.Saedler and P.
Huijser, 2008 The microRNA regulated SBP-box genes
SPL9 and SPL15 control shoot maturation in Arabidopsis. Plant
Mol. Biol. 67: 183195.
Slotkin, R. K., M. Vaughn,F.Borges,M.Tanurdzic,J.D.Becker
et al., 2009 Epigenetic reprogramming and small RNA silencing
of transposable elements in pollen. Cell 136: 461472.
Soltis, P. S., and D. E. Soltis, 2009 The role of hybridization in
plant speciation. Annu. Rev. Plant Biol. 60: 561588.
Staden, R., 1996 The Staden sequence analysis package. Mol. Bio-
technol. 5: 233241.
Tirosh, I., S. Reikhav,A.A.Levy and N. Barkai, 2009 A yeast
hybrid provides insight into the evolution of gene expression
regulation. Science 324: 659662.
Van de Peer, Y., S. Maere and A. Meyer, 2009 The evolutionary
signicance of ancient genome duplications. Nat. Rev. Genet. 10:
725732.
Vaucheret, H., 2008 Plant ARGONAUTES. Trends Plant Sci. 13:
350358.
Walia, H., C. Wilson,A.M.Ismail,T.J.Close and X. Cui,
2009 Comparing genomic expression patterns across plant spe-
cies reveals highly diverged transcriptional dynamics in response
to salt stress. BMC Genomics 10: 398.
Wood, T. E., N. Takebayashi,M.S.Barker,I.Mayrose,P.B.
Greenspoon et al., 2009 The frequency of polyploid speciation
in vascular plants. Proc. Natl. Acad. Sci. USA 106: 1387513879.
Zaratiegui, M., D. V. Irvine and R. A. Martienssen, 2007 Non-
coding RNAs and gene silencing. Cell 128: 763776.
Communicating editor: J. Schimenti
272 M. Kenan-Eichler et al.
GENETICS
Supporting Information
http://www.genetics.org/cgi/content/full/genetics.111.128348/DC1
Wheat Hybridization and Polyploidization Results in Deregulation
of Small RNAs
Michal Kenan-Eichler, Dena Leshkowitz, Lior Tal, Elad Noor, Cathy Melamed-Bessudo,
Moshe Feldman and Avraham A. Levy
Copyright ©2011 by the Genetics Society of America
DOI: 10.1534/genetics.111.128348
M. Kenan-Eichler et al. 2 SI









































FIGURE S1.—Metaphase chromosomes from a root tip of an S1 allopolyploid plant, spontaneously derived from
a TTR16XTQ113 F1 hybrid. The expected 42 chromosomes are shown.
M. Kenan-Eichler et al. 3 SI
:2(.131%!$2
:()'(04!+)381%!$2
4!+)38#(%#*
1),,)-'.&;/1),%1
+),)-!3)-'1%!$26)3(,.1%3(!-;2
22%,"+)-'!++1%!$26)3()$%-3)#!+2%04%-#%2)-3.4-)04%3!'2
+),)-!3)-'!++3!'26)3(+%223(!-1%!$2
).)-&.1,!3)#!-!+82)2
.-3)'#.-2314#3).-42)-'3!$%-
),)+!1)382%!1#(.&2%+%#3%$-4#+%.3)$%$!3!"!2%242)-'!(!2(!+'.1)3(,
+!23-1-3$!3!"!2%
.1,!+)9!3).-.&3(%+)"1!18#.4-323.1%!$2/%1,)++).-
!1%-35!1!-'$.-3%31!/+.)$
!1%-35!1$)/+.)$
8"1)$ 31)/+.)$
.+8/+.)$ (%7!/+.)$
:!--.3!3%$#.-3)'2:#.-3)'26)3(-.()32
FIGURE S2.—Flow chart of small RNA high-throughput sequencing data analysis.
M. Kenan-Eichler et al. 4 SI











"$ "#







   
 $"#
"$

"$

 
%"

 %! 
FIGURE S3.—Abundance of small RNA classes in the four analyzed libraries, namely, the tetraploid parent
(genome BBAA), the diploid parent (genome DD), their hybrid and derived allopolyploid (first generation= S1),
and the calculated mid-parent value (MPV). The total number of reads in each library is shown at the top of each
column.
M. Kenan-Eichler et al. 5 SI







 






 
 
FIGURE S4.—Northern blot analyses of Wis2-1A (left) and Veju (right) retrotransposon LTR regions in the
tetraploid parent (genome BBAA), the diploid parent (genome DD), their hybrid and derived allopolyploid.
Methylene blue staining of the membranes shows equal loading of RNA samples. Probes were derived from a
consensus sequence from wheat Wis2-1A LTRs (a 354bp fragment) and from Veju1_TM_LTR entry at TREP (a
339 bp fragment).
M. Kenan-Eichler et al. 6 SI
TABLE S1
Number of small RNA reads and tags in the various libraries: TTR16 –tetraploid parent
(Genome BBAA), TQ113 – diploid parent (Genome DD), F1 – triploid hybrid (genome BAD),
S1 – first generation of synthetic hexaploid (Genome BBAADD).
Libraries: TTR16 TQ113 Hybrid Polyploid Total
Total reads 5,633,938 3,064,285 6,421,516 3,458,759 18,578,498
Total tags 3,726,879 1,879,921 3,068,481 1,092,783 -
Tags after QCa 991,936 2,829,480 3,423,567 1,627,189 -
Tags with >30
reads
7967 4292 6756 2772 -
Reads from
tags with >30
reads
1,576,108 1,202,651 1,689,159 1,561,663 6,029,581
aQC: Quality Control process including 1. Trimming of 3’ PCR primer; 2. Elimination of reads with
>3Ns; 3. Assembly of reads with identical sequences into unique tags.
M. Kenan-Eichler et al. 7 SI
TABLE S2
Micro RNAs in high-throughput small RNA sequencing data
Librarires
miRNA name miRNA target TTR16 TQ113 Hybrid Polyploid
ath-miR156a SBP protein family 80003 36175 53344 140378
bna-miR156a SBP protein family 125499 69857 101207 58
ath-miR157a SBP protein family 0 0 0 115
osa-miR159a GAMYB transcription factor 854 969 1142 5102
osa-miR160f auxin reponse factor 0 0 21 56
ath-miR164a NAC1 transcription factor 1432 595 1634 2302
ath-miR165a classIII hd-leu zipper 43 0 40 0
ath-miR166a class III hd-leu zipper 17463 7565 18397 8228
osa-miR166g class III hd-leu zipper 1746 552 1378 366
osa-miR166k class III hd-leu zipper 5983 3095 5893 1777
ath-miR167d auxin reponse factor 8 19877 7066 12782 9790
osa-miR168a Argonaute 1 regulation 119710 79244 126428 245051
ath-miR169b CCAAT-binding tf 18 66 74 91
osa-miR171b scarecrow-like protein 36 0 34 98
ath-miR172a Apetala 2 5633 3540 4350 10126
ath-miR390a not found 0 32 0 18
ath-miR393a transport inhibitor protein 1 923 907 1022 1349
ath-miR396b not found 136 136 197 258
osa-miR396d not found 2259 1711 3771 12121
osa-miR528 not found 2020 2311 1003 2925
tae-miR1135 early light-inducible protein 312 0 186 26
Total counts 383956 213833 332913 440247
The number of reads are normalized to reads per million for comparison between the libraries. TTR16 –tetraploid
parent (Genome BBAA), TQ113 – diploid parent (Genome DD), F1 – triploid hybrid (genome BAD), S1 – first
generation of synthetic hexaploid (Genome BBAADD).
... tritici) is suppressed in the hexaploid state by Med15, a component of the Mediator complex encoded by the D sub-genome (Hiebert et al., 2020). These findings are also aligned with the observed alternations in gene expression during allopolyploidization (Kenan-Eichler et al., 2011;Li et al., 2014;Vasudevan et al., 2023;M. Yu et al., 2017). ...
Article
Full-text available
Bread wheat (Triticum aestivum L.) is a globally important food crop, which was domesticated about 8–10,000 years ago. Bread wheat is an allopolyploid, and it evolved from two hybridization events of three species. To widen the genetic base in breeding, bread wheat has been re‐synthesized by crossing durum wheat (Triticum turgidum ssp. durum) and goat grass (Aegilops tauschii Coss), leading to so‐called synthetic hexaploid wheat (SHW). We applied the quantitative genetics tools of “hybrid prediction”—originally developed for the prediction of wheat hybrids generated from different heterotic groups — to a situation of allopolyploidization. Our use‐case predicts the phenotypes of SHW for three quantitatively inherited global wheat diseases, namely tan spot (TS), septoria nodorum blotch (SNB), and spot blotch (SB). Our results revealed prediction abilities comparable to studies in ‘traditional’ elite or hybrid wheat. Prediction abilities were highest using a marker model and performing random cross‐validation, predicting the performance of untested SHW (0.483 for SB to 0.730 for TS). When testing parents not necessarily used in SHW, combination prediction abilities were slightly lower (0.378 for SB to 0.718 for TS), yet still promising. Despite the limited phenotypic data, our results provide a general example for predictive models targeting an allopolyploidization event and a method that can guide the use of genetic resources available in gene banks.
... However, five generations later, these allotetraploids display a relative abundance comparable to their parental diploids, suggesting a gradual return of the sRNA profile to its original state (Palacios et al. 2019). In that work, the first generation revealed that 21-nt sRNAs were over-represented, whereas the 24-nt sRNAs were underrepresented, which is consistent with previous reports (Kenan-Eichler et al. 2011;Barber et al. 2012). Nevertheless, an important difference between both polyploids relies on the formation of the successive generations. ...
Article
Full-text available
Key message The shock produced by the allopolyploidization process on a potato interspecific diploid hybrid displays a non-random remobilization of the small RNAs profile on a variety of genomic features. Abstract Allopolyploidy, a complex process involving interspecific hybridization and whole genome duplication, significantly impacts plant evolution, leading to the emergence of novel phenotypes. Polyploids often present phenotypic nuances that enhance adaptability, enabling them to compete better and occasionally to colonize new habitats. Whole-genome duplication represents a genomic “shock” that can trigger genetic and epigenetic changes that yield novel expression patterns. In this work, we investigate the polyploidization effect on a diploid interspecific hybrid obtained through the cross between the cultivated potato Solanum tuberosum and the wild potato Solanum kurtzianum, by assessing the small RNAs (sRNAs) profile of the parental diploid hybrid and its derived allopolyploid. Small RNAs are key components of the epigenetic mechanisms involved in silencing by RNA-directed DNA Methylation (RdDM). A sRNA sequencing (sRNA-Seq) analysis was performed to individually profile the 21 to 22 nucleotide (21 to 22-nt) and 24-nt sRNA size classes due to their unique mechanism of biogenesis and mode of function. The composition and distribution of different genomic features and differentially accumulated (DA) sRNAs were evaluated throughout the potato genome. We selected a subset of genes associated with DA sRNAs for messenger RNA (mRNA) expression analysis to assess potential impacts on the transcriptome. Interestingly, we noted that 24-nt DA sRNAs that exclusively mapped to exons were correlated with differentially expressed mRNAs between genotypes, while this behavior was not observed when 24-nt DA sRNAs were mapped to intronic regions. These findings collectively emphasize the nonstochastic nature of sRNA remobilization in response to the genomic shock induced by allopolyploidization.
... The intact or partial genome of an alien species can be combined with the wheat genome as amphiploids by hybridizing wheat to the alien species and subsequent chromosome doubling, which is the case for triticale, the generation of hexaploid synthetic wheat, and many other amphiploids ( Figure 2) [27]. Generally, amphiploids combine the characteristics of both wheat and alien species, but that combination is not equivalent to a simple one-plus-one chromosomal addition due to intergenomic interactions, polyploidization-related genome modifications, or epigenetic changes [86][87][88][89][90]. For example, an artificial wheat-alien species amphiploid can be developed into a new species, such as triticale (×Triticosecale), that combines and presents elements of both wheat and rye (Secale cereale) genomes. ...
Article
Full-text available
Wheat, including durum and common wheat, respectively, is an allopolyploid with two or three homoeologous subgenomes originating from diploid wild ancestral species. The wheat genome’s polyploid origin consisting of just three diploid ancestors has constrained its genetic variation, which has bottlenecked improvement. However, wheat has a large number of relatives, including cultivated crop species (e.g., barley and rye), wild grass species, and ancestral species. Moreover, each ancestor and relative has many other related subspecies that have evolved to inhabit specific geographic areas. Cumulatively, they represent an invaluable source of genetic diversity and variation available to enrich and diversify the wheat genome. The ancestral species share one or more homologous genomes with wheat, which can be utilized in breeding efforts through typical meiotic homologous recombination. Additionally, genome introgressions of distant relatives can be moved into wheat using chromosome engineering-based approaches that feature induced meiotic homoeologous recombination. Recent advances in genomics have dramatically improved the efficacy and throughput of chromosome engineering for alien introgressions, which has served to boost the genetic potential of the wheat genome in breeding efforts. Here, we report research strategies and progress made using alien introgressions toward the enrichment and diversification of the wheat genome in the genomics era.
... Kenan-Eichler et al. (2011) [119] studied the presence of small RNAs after A. tauschii × T. turgidum hybridization and allopolyploidization, in special, small RNAs corresponding to TEs. They found that the percentage of siRNAs corresponding to TEs strongly decreased upon allopolyploidization, but not upon hybridization. ...
Article
Full-text available
Transposable elements (TEs) are major components of plant genomes with the ability to change their position in the genome or to create new copies of themselves in other positions in the genome. These can cause gene disruption and large-scale genomic alterations, including inversions, deletions, and duplications. Host organisms have evolved a set of mechanisms to suppress TE activity and counter the threat that they pose to genome integrity. These includes the epigenetic silencing of TEs mediated by a process of RNA-directed DNA methylation (RdDM). In most cases, the silencing machinery is very efficient for the vast majority of TEs. However, there are specific circumstances in which TEs can evade such silencing mechanisms, for example, a variety of biotic and abiotic stresses or in vitro culture. Hybridization is also proposed as an inductor of TE proliferation. In fact, the discoverer of the transposons, Barbara McClintock, first hypothesized that interspecific hybridization provides a “genomic shock” that inhibits the TE control mechanisms leading to the mobilization of TEs. However, the studies carried out on this topic have yielded diverse results, showing in some cases a total absence of mobilization or being limited to only some TE families. Here, we review the current knowledge about the impact of interspecific hybridization on TEs in plants and the possible implications of changes in the epigenetic mechanisms.
... Several reports supported the hypothesis that these multi-functional sRNAs might also be involved in heterosis manifestation, since their accumulation in hybrids of model species was significantly lower than that in parents, correlating to decreased methylation patterns (Groszmann et al., 2011). This correlation was confirmed in many independent studies carried out on wheat, maize and rice, in which sRNA accumulation analysis showed significant variation in sRNA amounts between hybrids and their respective parental lines Chodavarapu et al., 2012;Kenan-Eichler et al., 2011). However, the specific methylated regions and the mechanisms influencing hybrid performance remain unclear and have become a stimulating topic to be explored in horticulture research. ...
Article
Full-text available
Heterosis in plants has been among the challenging topics for plant scientists worldwide. The production of F1 hybrid varieties of seed-propagated horticultural species is one of the most successful applications of plant breeding techniques. The exploitation of the heterosis phenomenon promotes homogeneity and maximizes crop yields and is a way for breeders to legally control and protect their commercial products. In the past heterosis has been largely studied and explored in cereal crop systems, considering maize as a model for understanding the genetic bases of this phenomenon. To date, crossbreeding in horticultural vegetables has also rapidly progressed. F1 hybrid varieties are available for many horticultural crops, including both allogamous and autogamous species. Several genetic and nongenetic mechanisms have been applied to facilitate the large-scale production of F1 hybrid seeds in vegetable crops to prevent undesirable selfing. Although the development and commercialization of F1 hybrids is currently common in agriculture, this phenomenon is still being investigated at different levels. With the rapid accumulation of knowledge on plant genome structures and gene activities and the advancement of new genomics platforms and methodologies, significant progress has been achieved in recent years in the study of the genetic and molecular bases of heterosis. This paper provides a brief overview of current theoretical advances and practical predictions of the molecular mechanisms underlying heterosis in plants. The aim is to carefully summarize the fundamental mechanisms of heterosis in plants, focusing on horticultural plant breeding, to improve the existing knowledge in this research area. We describe the quantitative genetic model of phenotypic variation and combine evolutionary, phenotypic and molecular genetic views to explain the origin and manifestation of heterosis and its significance for breeding F1 hybrid varieties in horticultural crops. The principles of genomic prediction and its applications in genomic selection are then covered.
... Recent application of next generation sequencing-based approaches revealed that rye Bs contain more than 4000 putative genic sequences (Martis et al. 2012 (Carchilan et al. 2009) and, from these studies, regulatory interactions between A-and B-located coding sequences have been proposed (Banaei-Moghaddam et al. 2015). It is likely that Bs influence A-localized sequences through epigenetic mechanisms, such as homology-dependent RNA interference pathways (Slotkin and Martienssen 2007), as has been proposed for the modulation of gene-activity in newly formed hybrids and allopolyploids (Comai 2005;Kenan-Eichler et al. 2011). Bs may also exert control over A chromosomes via the spatial organization of As in interphase nuclei, and it has been suggested that spatial positioning of genes and chromosomes can influence gene expression (Misteli 2007). ...
Chapter
Full-text available
This chapter describes supernumerary or accessory chromosomes (B-chromosomes) in several grasses focusing on those in species of the sub-tribe Triticineae of the tribe Triticeae. It refers to their origin, molecular characterization, preferential transmission (accumulation mechanism), effect on morphology, fitness, and chromosomal pairing in species and hybrids, and their transcriptional activity.
Article
Full-text available
Trans-chromosomal interactions resulting in changes in DNA methylation during hybridization have been observed in several plant species. However, little is known about the causes or consequences of these interactions. Here, we compared DNA methylomes of F1 hybrids that are mutant for a small RNA biogenesis gene, Mop1 (mediator of paramutation1) with that of their parents, wild-type siblings, and backcrossed progeny in maize (Zea mays). Our data show that hybridization triggers global changes in both trans-chromosomal methylation (TCM) and trans-chromosomal demethylation (TCdM), most of which involved changes in CHH methylation. In more than 60% of these TCM differentially methylated regions (DMRs) in which small RNAs are available, no significant changes in the quantity of small RNAs were observed. Methylation at the CHH TCM DMRs was largely lost in the mop1 mutant, although the effects of this mutant varied depending on the location of these DMRs. Interestingly, an increase in CHH at TCM DMRs was associated with enhanced expression of a subset of highly expressed genes and suppressed expression of a small number of lowly expressed genes. Examination of the methylation levels in backcrossed plants demonstrates that both TCM and TCdM can be maintained in the subsequent generation, but that TCdM is more stable than TCM. Surprisingly, although increased CHH methylation in most TCM DMRs in F1 plants required Mop1, initiation of a new epigenetic state of these DMRs did not require a functional copy of this gene, suggesting that initiation of these changes is independent of RNA-directed DNA methylation.
Chapter
Full-text available
This chapter describes characteristic features of the chromosomes and genomes of Triticeae species. Centromeres contain typical CENH3 nucleosomes, but these are associated with repeats that are larger than in other plant species. The sub-telomeric ends are rich in transposable elements and contain diverse repeats and recombination hotspots. The nucleolar organizer regions contain hundreds or thousands of ribosomal genes, rDNA repeats, arranged in tandem arrays that form a constriction known as the nucleolar organizer (NOR). We describe their mapping as well as the phenomenon known as Nucleolar dominance. Genome sizes in the Triticeae are large, with 1C values ranging in diploids from 4.0–9.4 pg, compared to related grasses such as rice (1C = 0.5 pg). These size differences are mostly due to a large amount of repetitive DNA, in particular of transposable elements, with retroelements as the most prominent repeats. In hexaploid bread wheat, genome size reaches 1C = 16 pg, with ~ 108,000 high-confidence protein-coding genes, and a high number of pseudogenes and RNA genes. The wheat transcriptome shows complex expression patterns for homoeologous loci. We discuss gene organization in islands as well as the high synteny between the different species and the role of introgression in shaping genomes.
Chapter
Full-text available
The chapter deals with the various steps, periods, and processes that led to the domestication of the wheat as well as with the archaeological sites where domestication took place. Additionally, the chapter describes the ecogeographical characteristics of the area of wheat domestication, the selection of non-brittle rachis, large grain size, rapid and synchronous germination, free-threshing grains, and yield. The genetic basis of non-brittle rachis and free-threshing grains are delt with in details. The formation of hexaploid wheat, T. aestivum , and the spread of its free-threshing form to almost all parts of the globe to become the main cultivated wheat, are reviewed. The production of synthetic Triticum aestivum , and Triticale are also referred to in this chapter.
Chapter
Full-text available
The chapter describes the mode, time, and place of origin of the allopolyploids of the genus Triticum . In addition, genetic and epigenetic changes due to allopolyploidization that brought about to cytological diploidization (exclusive homologous chromosome pairing), are discussed within the chapter. The suppression of pairing between homoeologous chromosomes in hexaploid wheat Triticum aestivum by the Ph1 gene, its discovery, the induction of mutations in this gene, its isolation, the theories concerning its mode of its action, and its origin, are discussed in the chapter. Other pairing genes (suppressors and promoters) that exist in T. aestivum , and in its relatives are reviewed. The chapter also describes processes leading to genetic diploidization and subgenomic asymmetry in the control of various traits in allopolyploid wheats. Several aspects of evolution during the life of the allopolyploids are discussed too.
Article
Full-text available
Allopolyploidy, or the combination of two or more distinct genomes in one nucleus, is usually accompanied by radical genomic changes involving transposable elements (TEs). The dynamics of TEs after an allopolyploidization event are poorly understood. In this study, we analyzed the methylation state and genetic rearrangements of a high copied, newly amplified terminal-repeat retrotransposon in miniature (TRIM) family in wheat termed Veju. We found that Veju insertion sites underwent massive methylation changes in the first four generations of a newly formed wheat allohexaploid. Hypomethylation or hypermethylation occurred in ∼43% of the tested insertion sites; while hypomethylation was significantly predominant in the first three generations of the newly formed allohexaploid, hypermethylation became predominant in the subsequent generation. In addition, we determined that the methylation state of Veju long terminal repeats (LTRs) might be correlated with the deletion and/or insertion of the TE. While most of the methylation changes and deletions of Veju occurred in the first generation of the newly formed allohexaploid, most Veju insertions were seen in the second generation. Finally, using quantitative PCR, we quantitatively assessed the genome composition of Veju in the newly formed allohexaploid and found that up to 50% of Veju LTRs were deleted in the first generation. Retrotransposition bursts in subsequent generations, however, led to increases in Veju elements. In light of these findings, the underlying mechanisms of TRIM rearrangements are discussed.
Article
Full-text available
Heterosis refers to the phenomenon that progeny of diverse varieties of a species or crosses between species exhibit greater biomass, speed of development, and fertility than both parents. Various models have been posited to explain heterosis, including dominance, overdominance, and pseudo-overdominance. In this Perspective, we consider that it might be useful to the field to abandon these terms that by their nature constrain data interpretation and instead attempt a progression to a quantitative genetic framework involving interactions in hierarchical networks. While we do not provide a comprehensive model to explain the phenomenology of heterosis, we provide the details of what needs to be explained and a direction of pursuit that we feel should be fruitful.
Article
Full-text available
More than 80% of the wheat genome is composed of transposable elements (TEs). Since active TEs can move to different locations and potentially impose a significant mutational load, their expression is suppressed in the genome via small non-coding RNAs (sRNAs). sRNAs guide silencing of TEs at the transcriptional (mainly 24-nt sRNAs) and post-transcriptional (mainly 21-nt sRNAs) levels. In this study, we report the distribution of these two types of sRNAs among the different classes of wheat TEs, the regions targeted within the TEs, and their impact on the methylation patterns of the targeted regions. We constructed an sRNA library from hexaploid wheat and developed a database that included our library and three other publicly available sRNA libraries from wheat. For five completely-sequenced wheat BAC contigs, most perfectly matching sRNAs represented TE sequences, suggesting that a large fraction of the wheat sRNAs originated from TEs. An analysis of all wheat TEs present in the Triticeae Repeat Sequence database showed that sRNA abundance was correlated with the estimated number of TEs within each class. Most of the sRNAs perfectly matching miniature inverted repeat transposable elements (MITEs) belonged to the 21-nt class and were mainly targeted to the terminal inverted repeats (TIRs). In contrast, most of the sRNAs matching class I and class II TEs belonged to the 24-nt class and were mainly targeted to the long terminal repeats (LTRs) in the class I TEs and to the terminal repeats in CACTA transposons. An analysis of the mutation frequency in potentially methylated sites revealed a three-fold increase in TE mutation frequency relative to intron and untranslated genic regions. This increase is consistent with wheat TEs being preferentially methylated, likely by sRNA targeting. Our study examines the wheat epigenome in relation to known TEs. sRNA-directed transcriptional and post-transcriptional silencing plays important roles in the short-term suppression of TEs in the wheat genome, whereas DNA methylation and increased mutation rates may provide a long-term mechanism to inactivate TEs.
Article
Soon after its discovery 75 years ago, heterochromatin, a dense chromosomal material, was found to silence genes. But its importance in regulating gene expression was controversial. Long thought to be inert, heterochromatin is now known to give rise to small RNAs, which, by means of RNA interference, direct the modification of proteins and DNA in heterochromatic repeats and transposable elements. Heterochromatin has thus emerged as a key factor in epigenetic regulation of gene expression, chromosome behaviour and evolution.
Article
Plant non-specific lipid transfer proteins(nsLTPs) are a group of abundantly expressed small basic proteins,which can reversibly bind and transport hydrophobic molecules in vitro.Nine types of nsLTP have been identified from a variety of plants.All the type of nsLTPs possesses the conserved eight cysteine residue motif with a three-dimensional structure of an internal hydrophobic cavity and the lipid binding site.Based on the growing knowledge about the structure,gene expression,regulation and in vitro activity,nsLTPs are considered to play a role in key processes of plant physiology,including wax synthesis and transport,abiotic stress resistance,disease resistance,and plant reproduction.This review aims at presenting comprehensive information of key topics,including basic features,classification,gene cloning,expression profiles,and functional studies of nsLTP.Finally the perspectives were included on the future study of nsLTP family.
Article
The production of 2n gametes in plants, i.e. gametes with a somatic chromosome number, is considered to be the dominant process involved in the origin of polyploid plants. In this review, we provide a synthesis of current knowledge concerning the production of 2n gametes. Firstly, we describe the different methods used to detect and quantify the production of 2n gametes in plants, which include morphological and flow cytometry screening of the occurrence of 2n pollen, the analysis of crosses among diploid and tetraploid parents and the instigation of micro‐and mega‐sporogenesis. Secondly, the high level of inter‐ and infra‐specific variation in 2n gametes production is described. Thirdly, the various cytological anomalies responsible for the production of 2n gametes are reviewed, with particular reference to the relative genetic consequences of the first and second restitution divisions that give rise to 2n gametes. Fourthly, the significance of 2n gametes in crop plant improvement is discussed, in relation to somatic chromosome doubling to obtain new polyploid varieties. In particular, we compare the genetic and yield consequences of methods based on unilateral and bilateral sexual polyploidization. Finally, we outline how knowledge of the variety of mechanisms involved in 2n gamete production have increased our understanding of the evolutionary significance of polyploidy and the population biology of polyploid plants. Contents Summary 1 I. Introduction 2 II. Methods used to detect the presence and frequency of 2n gametes 3 III. Frequency of 2n gamete production 5 IV. Mechanisms of formation and the influence of external factors 6 V. The genetic consequences of First Division Restitution (FDR) and Second Division Restitution (SDR) 12 VI. 2n gametes and the unilateral and bilateral sexual polyploidization of crop plants 13 VII. The evolutionary significance of 2n gamete production 15 Acknowledgements 18 References 18
Article
Ecological factors, hybrid sterility and differences in ploidy levels are well known for contributing to gene-flow barriers in plants. Another common postzygotic incompatibility, hybrid necrosis, has received comparatively little attention in the evolutionary genetics literature. Hybrid necrosis is associated with a suite of phenotypic characteristics that are similar to those elicited in response to various environmental stresses, including pathogen attack. The genetic architecture is generally simple, and complies with the Bateson-Dobzhansky-Muller model for hybrid incompatibility between species. We survey the extensive literature on this topic and present the hypothesis that hybrid necrosis can result from autoimmunity, perhaps as a pleiotropic effect of evolution of genes that are involved in pathogen response.
Article
Hybrids such as maize (Zea mays) or domestic dog (Canis lupus familiaris) grow bigger and stronger than their parents. This is also true for allopolyploids such as wheat (Triticum spp.) or frog (i.e. Xenopus and Silurana) that contain two or more sets of chromosomes from different species. The phenomenon, known as hybrid vigor or heterosis, was systematically characterized by Charles Darwin (1876). The rediscovery of heterosis in maize a century ago has revolutionized plant and animal breeding and production. Although genetic models for heterosis have been rigorously tested, the molecular bases remain elusive. Recent studies have determined the roles of nonadditive gene expression, small RNAs, and epigenetic regulation, including circadian-mediated metabolic pathways, in hybrid vigor, which could lead to better use and exploitation of the increased biomass and yield in hybrids and allopolyploids for food, feed, and biofuels.