ArticlePDF Available

Prevention of early flowering by expression of FLOWERING LOCUS C requires methylation of histone H3 K36

Authors:

Abstract and Figures

Flowering represents a crucial transition from a vegetative to a reproductive phase of the plant life cycle. Despite extensive studies, the molecular mechanisms controlling flowering remain elusive. Although the enzymes involved are unknown, methylation of histone H3 K9 and K27 correlates with repression of FLOWERING LOCUS C (FLC), an essential transcriptional repressor involved in flowering time control in Arabidopsis thaliana; in contrast, methylation of H3K4 correlates with FLC activation. Here we show that loss-of-function of SET DOMAIN GROUP 8 (SDG 8), which encodes a homologue of the yeast SET2 histone methyltransferase, results in reduced dimethylation of histone H3K36, particularly in chromatin associated with the FLC promoter and the first intron, regions that contain essential cis-elements for transcription. sdg8 mutants display reduced FLC expression and flower early, establishing SDG8-mediated H3K36 methylation as a novel epigenetic memory code required for FLC expression in preventing early flowering. This is the first demonstrated role of H3K36 methylation in eukaryote development.
Content may be subject to copyright.
LETTERS
1256 NATURE CELL BIOLOGY VOLUME 7 | NUMBER 12 | DECEMBER 2005
Prevention of early flowering by expression of
FLOWERING LOCUS C requires methylation
of histone H3 K36
Zhong Zhao1, Yu Yu1, Denise Meyer1, Chengjun Wu1,2 and Wen-Hui Shen1,3
Flowering represents a crucial transition from a vegetative
to a reproductive phase of the plant life cycle. Despite
extensive studies, the molecular mechanisms controlling
flowering remain elusive. Although the enzymes involved are
unknown, methylation of histone H3 K9 and K27 correlates
with repression of FLOWERING LOCUS C (FLC), an essential
transcriptional repressor involved in flowering time control
in Arabidopsis thaliana; in contrast, methylation of H3K4
correlates with FLC activation1–4. Here we show that loss-of-
function of SET DOMAIN GROUP 8 (SDG 8), which encodes a
homologue of the yeast SET2 histone methyltransferase, results
in reduced dimethylation of histone H3K36, particularly in
chromatin associated with the FLC promoter and the first intron,
regions that contain essential cis-elements for transcription.
sdg8 mutants display reduced FLC expression and flower
early, establishing SDG8-mediated H3K36 methylation as a
novel epigenetic memory code required for FLC expression in
preventing early flowering. This is the first demonstrated role of
H3K36 methylation in eukaryote development.
In eukaryotes, genomic DNA is wrapped around histone octomers
(composed of two molecules each of histones H2A, H2B, H3 and H4),
resulting in highly packaged chromatin in the nucleus5. Current models
suggest that changes in chromatin structure positively or negatively effect
the accessibility of DNA to transcription factors, resulting in temporal
and spatial variation in gene expression. Histones, particularly their tails,
are subject to multiple types of covalent modification, including acetyla-
tion, methylation, phosphorylation and ubiquitination. Combinations
of such modifications have been proposed to serve as a ‘histone code’,
specifying a chromatin state that can determine the transcriptional activ-
ity of the embedded gene6.
To date, four lysines within histone H3 (K4, K9, K27 and K36) and
one lysine within histone H4 (K20) have been identified as being meth-
ylated by SET (Su(var)3-9, E(z) and Trithorax-conserved)-domain
1Institut de Biologie Moléculaire des Plantes (IBMP), Centre National de la Recherche Scientifique (CNRS), Université Louis Pasteur de Strasbourg (ULP), 12 rue
du Général Zimmer, 67084 Strasbourg Cédex, France. 2Present address: Biotechnology Research Institute of Yunnan Academy of Agriculture Sciences, Kunming
650223, Yunnan, People’s Republic of China.
3Correspondence should be addressed to W.-H. S. (e-mail: Wen-Hui.Shen@ibmp-ulp.u-strasbg.fr)
Received 27 May 2005; accepted 28 September 2005; published online: 20 November 2005; DOI: 10.1038/ncb1329
histone-lysine-methyltransferases (HKMTs), and one lysine within H3
(K79) as being methylated by a non-SET-domain HKMT7. Compared
to the well-established function of methylation at other lysines, little
is known about H3K36 methylation. H3K36 di- and tri-methylation is
enriched in actively transcribed genomic regions in yeast8,9 and meta-
zoa10, but the Saccharomyces cerevisiae set2 mutant lacking H3K36
methylation is viable and shows no marked growth defects11. The
mouse nuclear receptor binding protein, NSD1, which is involved in
acute myeloid leukaemia, was shown to methylate H3K36, as well as
H4K20, in vitro12. However, the in vivo function of H3K36 methylation
by NSD1 has not been established.
The Arabidopsis genome encodes 39 SET-domain proteins — named
SDG1 to SDG39 in the Plant Chromatin Database (www.chromdb.org).
SDG8, a 1759-amino-acid protein encoded by a gene (At1g77300) con-
taining 15 exons (Fig. 1a), shows the highest homology with SET2, the
sole H3K36 HKMT of S. cerevisiae11. The sequence homology between
the two proteins is limited to the region spanning the SET domain
and its surrounding cysteine-rich AWS (Associated With SET) and C
(Cysteine-rich) domains (see Supplementary Information, Fig. S1). The
plant protein also contains two motifs that are conserved within the
RPB1 subunits of RNA polymerase II, a CW (cysteine and tryptophan
conserved) domain that is also present in some other chromatin fac-
tors13, and a long amino-terminal extension (Fig. 1b).
To characterize the function of SDG8, we identified two mutants
(sdg8-1 and sdg8-2) from the Salk collection of Arabidopsis T-DNA inser-
tion lines14. Polymerase chain reaction (PCR) analysis confirmed that
sdg8-1 and sdg8-2 contain a T-DNA inserted in exon 7 and 2, respectively
(Fig. 1a). Both the sdg8-1 and sdg8-2 mutations are recessive and only
homozygous (hereafter called mutant) plants had a phenotype. At early
stages, the sdg8-1 and sdg8-2 mutants grew similarly to wild-type plants
(Fig. 1c). However, both sdg8-1 and sdg8-2 mutants flowered earlier,
slowing down rosette growth (Fig. 1d). Flowers of the mutant plants were
smaller (Fig. 1e) but contained morphologically normal flower organs
(four sepals, four petals, five–six stamens and two fused carpels per
ncb1329.indd 1256ncb1329.indd 1256 1/12/05 3:26:33 pm1/12/05 3:26:33 pm
Nature Publishing Group
© 2005
© 2005 Nature Publishing Group
NATURE CELL BIOLOGY VOLUME 7 | NUMBER 12 | DECEMBER 2005 1257
LETTERS
flower). Reverse transcriptase (RT)–PCR analysis revealed that SDG8
transcripts were not detectable in either the sdg8-1 or the sdg8-2 mutant
(Fig. 1f). More recently, a third mutant allele (sdg8-3) containing a T-
DNA inserted in intron 14 (Fig. 1a) was obtained from the Syngenta
collection, and this mutant showed a phenotype similar to sdg8-1 and
sdg8-2. Furthermore, a genomic fragment containing the SDG8 coding
region together with the upstream 3,023 bp and the downstream 308
bp non-coding regions could rescue both sdg8-1 and sdg8-2 mutants,
resulting in plants with a wild-type phenotype (Table 1). Together, these
data establish that SDG8 prevents early flowering.
Flowering represents the transition from vegetative to reproductive
development, a shift that is temporally controlled by developmental state
(for example, size and age) and regulated by environmental cues, such as
light and temperature1. Similarly to wild-type plants, all three (sdg8-1,
sdg8-2 and sdg8-3) mutants flowered earlier under inductive long-day
photoperiods than under short-day photoperiods (Table 1), indicating
that SDG8 is dispensable for the photoperiod response. Nevertheless,
when compared to wild-type plants, mutant plants flowered earlier
under all studied photoperiods. The strongest lag occurred under short-
day photoperiods, indicating a positive effect of light quantity on early
flowering. Exposure to a prolonged period of cold (vernalization) pro-
moted flowering of the wild-type plants but had no significant effect on
the sdg8-1 and sdg8-2 mutants (Table 1). Expression of SDG8 in wild-type
plants, however, was unchanged by vernalization (see Supplementary
Information, Fig. S2a). This latter result suggests that SDG8 does not
act as an initiator of the vernalization response pathway.
To investigate the molecular mechanism underlying early flowering of
the mutant plants, the expression of floral pathway genes was analysed.
Consistent with their normal photoperiod-responsive phenotype, the
mutants showed a normal expression of the CONSTANS ( CO) gene (data
not shown), a key regulator specific of the photoperiod pathway1. The
vernalization pathway and the autonomous (photoperiod-independent;
thought to be regulated in response to developmental cues) pathways
converge on FLC1, which encodes a MADS-box transcription factor. FLC
represses flowering in a dosage-dependent manner by negatively regulat-
ing the expression of several genes that promote flowering; for example,
FLOWERING LOCUS T (FT) and SUPPRESSOR OF OVEREXPRESSION
OF CO 1 (SOC1) (refs 15–18). In both sdg8-1 and sdg8-2 mutant plants,
FLC expression was markedly repressed whereas the downstream genes
FT and SOC1 were activated (Fig. 2a).
a
b
f
ce
d
sdg8-2
T-DNA
sdg8-1
T-DNA
sdg8-3
T-DNA
8651 bp
CW AWS SET C
1759 aa
RPB1
Col
Col
sdg8-1
Col sdg8-1 sdg8-2
Col
sdg8-1
sdg8-1
SDG8
ACTIN
Figure 1 Loss of SDG8 function causes early flowering. (a) Exon–intron
structure of the SDG8 gene. Boxes represent exons; lines represent introns;
triangles indicate T-DNA insertions. The T-DNA was located in the second
exon in sdg8-2, in the seventh exon in sdg8-1 and in the fourteenth intron
in sdg8-3. (b) Schematic diagram of the SDG8 protein. The CW, AWS, SET
and C boxes represent conserved domains. Two regions with homology to the
consensus sequence of the RPB1 motif are presented in sequence alignment.
(c) Wild-type Col and mutant sdg8-1 plants 29 days after germination (DAG).
(d) Wild-type Col and mutant sdg8-1 plants 46 DAG. (e) Wild-type Col and
mutant sdg8-1 flowers. (f) RT–PCR analysis of SDG8 expression in wild-type
Col and mutant sdg8-1 and sdg8-2 plants 20 DAG. ACTIN serves as an
internal control. Scale bars represent 10 mm in c and d, and 1 mm in e.
ncb1329.indd 1257ncb1329.indd 1257 1/12/05 3:26:38 pm1/12/05 3:26:38 pm
Nature Publishing Group
© 2005
© 2005 Nature Publishing Group
1258 NATURE CELL BIOLOGY VOLUME 7 | NUMBER 12 | DECEMBER 2005
LETTERS
The methylation level of different histone H3 lysine residues in mutant
and wild-type plants was analysed by western blotting with specific anti-
bodies2 to investigate whether SDG8 regulated FLC expression through
histone methylation. Although the levels of di-methylated H3K9 and
H3K27 were unchanged, both sdg8-1 and sdg8-2 mutants contained a
reduced level of di-methylated H3K36 (Fig. 2b), indicating that SDG8
encodes a major enzyme controlling H3K36 methylation in Arabidopsis.
A significant increase in di-methylated H3K4 was also observed in the
mutants (Fig. 2b), suggesting that decreased H3K36 methylation favours
H3K4 methylation. This finding is contrary to the previously described
cross-talk among histone modifications19.
We next examined histone H3 methylation at FLC by chromatin
immunoprecipitation (CHIP). Similar sets of PCR primers to those pre-
viously reported2 were used to cover the promoter (region A), the first
exon (region B), the first intron (regions C, D, E and F) and some down-
stream exons and introns (region G) of FLC (Fig. 3a). Consistent with a
previous report on non-vernalized plants2, FLC chromatin was found to
be rich in acetylated H3 and di-methylated H3K4, whereas di-methyl-
ated H3K9 and H3K27 were undetectable in wild-type plants (Fig. 3b).
Control experiments demonstrated that the actively expressed ACTIN20,21
was embedded in chromatin that was rich in acetylated H3 and di-meth-
ylated H3K4, whereas the silenced transposon Ta3 was primarily found in
chromatin rich in di-methylated H3K9 and H3K27. Furthermore our data
show that H3K36 di-methylation was strongly associated with FLC and
weakly with ACTIN, but was undetectable with Ta3 chromatin (Fig. 3b).
Compared with wild-type plants, sdg8-1 plants showed a marked reduc-
tion in H3K36 di-methylation in regions A, D, E and F, and a slightly
decreased level of di-methylated H3K36 in region B. There were slightly
increased levels of di-methylated H3K4 and actetylated H3 in region
B, but no changes in other regions or in H3K9 and H3K27 di-methyla-
tion in sdg8-1 plants (Fig. 3b). These results were reproducible in three
independent experiments. Although only a slight increase in di-methyl-
ated H3K4 was detected in region B of the FLC locus (Fig. 3b), a more
pronounced increase of di-methylated H3K4 was observed with whole
nuclear extracts of sdg8 mutants (Fig. 2b), indicating that sdg8 does not
uniformly affect H3K4 di-methylation over the entire genome.
The regions showing the most pronounced changes of H3K36 di-
methylation by sdg8-1 are located within the promoter and the first
intron of FLC, which correspond to regions with essential cis-acting
functions22. These regions have also been reported to be regulated
by H3K9 and H3K27 di-methylation2,4, H3K4 di- and tri-methyla-
tion2,3, and histone acetylation20,23. Interestingly, whereas the H3K36
di-methylation of FLC region B changed little in sdg8-1 mutants, this
region was highly responsive to H3K9 and H3K27 di-methylation in
vernalization2,4, and to H3K4 tri-methylation in early flowering elf7
and elf8 mutants3. In contrast, the regions D, E and F were markedly
reduced for H3K36 di-methylation in sdg8-1 but H3K9 di-methylation
was persistently undetectable in these regions upon vernalization2. Thus,
it seems t hat SDG8 targets H3K36 methylation in specific regions of FLC
chromatin that are distinct from the regions regulated by H3K4, H3K9
and H3K27 methylation in flowering time control.
Taken together, our genetic, physiological and molecular data indi-
cate that SDG8 methylates H3K36 of FLC chromatin, which positively
regulates FLC transcription to prevent early flowering. To test whether
ectopic ex pression of FLC could inhibit early flowering in an sdg8 mutant
background, sdg8-1 and sdg8-2 mutants, and wild-type plants were
transformed with a 35S:FLC construct that expresses FLC cDNA under
the control of the constitutive 35S promoter. 35S:FLC transformation
resulted in transgenic plants with different flowering times, including
early, unchanged and late, in either wild-type Col or mutants sdg8-1
and sdg8-2 backgrounds (see Supplementary Information, Table S1), as
previously reported for Ler and C24 ecotypes15,16. Specifically, 6 of 14
Col, 4 of 26 sdg8-1 and 6 of 42 sdg8-2 transgenic lines showed a very late
flowering phenotype (more than 18 rosette leaves under long-day pho-
toperiods). RT–PCR analysis demonstrated that the late-flowering phe-
notype correlates with high levels of FLC transcripts (see Supplementary
Information, Fig. S2b). Together, these data show that overexpression
of FLC alone is sufficient to cause a significant delay in flowering time
in sdg8 mutants and in wild-type plants. Nevertheless, 35S:FLC was
expressed at lower levels in the mutant than in the wild-type plants and
correlates with the reduced delay in flowering. This suggests that sdg8
also affects 35S:FLC expression.
Although the enzymes involved in H3K4, H3K9 and H3K27 methyla-
tion of FLC chromatin are not currently known, there is high sequence
homology of the ELF7 and ELF8 mutants to subunits of the yeast PAF1
transcriptional complex3. This suggests that, as in yeast, a SET1-like
HKMT may be recruited by the PAF1 complex during H3K4 methylation
of FLC chromatin. In yeast, the PAF1 complex also interacts with SET2,
consistent with an additional role in H3K36 methylation8,9. Functional
interaction between SDG8 and VERNALIZATION INDEPENDENCE
4 (VIP4; which encodes a homologue of LEO1, a subunit of the PAF1
complex8) was genetically tested. The vip4 mutant has been previously
shown to have reduced FLC expression and exhibit early flowering24,25.
The double mutant vip4–sdg8-1 flowered earlier than either the vip4 or
Table 1 Rosette leaf number at fl owering of wild-type (Col) and mutant plants in different growth conditions
SD–V MD–V LD–V SD + V MD + V LD + V
Col 63.0 ± 3.5 (12) 15.7 ± 0.3 (15) 11.1 ± 0.2 (18) 33.5 ± 0.6 (15) 12.8 ± 0.2 (15) 10.8 ± 0.3 (15)a
sdg8-1 22.5 ± 4.2 (12) 8.5 ± 0.2 (20) 8.0 ± 0.1 (21) 21.9 ± 1.0 (15) 8.8 ± 0.2 (15) 8.1 ± 0.2 (15)a
sdg8-2 23.4 ± 3.8 (12) 9.1 ± 0.2 (25) 7.9 ± 0.1 (21) 21.7 ± 0.7 (15) 9.2 ± 0.2 (15) 8.2 ± 0.2 (15)a
sdg8-3 21.4 ± 2.3 (20) 8.9 ± 0.5 (17) 8.1 ± 0.4 (20) n.t. n.t. n.t.
sdg8-1 pCSDG8b65.0 ± 4.4 (9) n.t. 11.0 ± 0.3 (11) n.t. n.t. n.t.
sdg8-2 pCSDG8b63.7 ± 3.0 (11) n.t. 11.3 ± 0.2 (27) n.t. n.t. n.t.
vip4–sdg8-1 18.3 ± 0.8 (15) 7.3 ± 0.4 (16) 6.0 ± 0.3 (15) n.t. n.t. n.t.
vip4 25.1 ± 1.2 (15) 12.1 ± 0.3 (24) 9.8 ± 0.4 (15) n.t. n.t. n.t.
Plants without (–) and with (+) vernalization (V) were tested in different photoperiods: long day (LD) was 16 h light and 8 h dark, medium day (MD) was 12 h light and 12 h dark, and short day
(SD) was 8 h light and 16 h dark. Values shown are mean number of rosette leaves ± standard deviation. Numbers in parentheses are the total numbers of plants evaluated. n.t., not tested. aThe
differences between wild-type and mutants are smallest at LD + V among all tested growth conditions; nevertheless these differences are highly signifi cant (P = 0.0001) according to a t-test.
bFrom 33 and 41 transformants, 11 and 13 independent transgenic lines (showing wild-type phenotype), were obtained for sdg8-1 pCSDG8 and sdg8-2 pCSDG8, respectively. The statistical
data shown were for progeny of a homozygous line of sdg8-1 pCSDG8 or sdg8-2 pCSDG8.
ncb1329.indd 1258ncb1329.indd 1258 1/12/05 3:26:40 pm1/12/05 3:26:40 pm
Nature Publishing Group
© 2005
© 2005 Nature Publishing Group
NATURE CELL BIOLOGY VOLUME 7 | NUMBER 12 | DECEMBER 2005 1259
LETTERS
sdg8-1 single mutants (Table 1), sug gesting that H3K4 and H3K36 meth-
ylation function synergistically to regulate FLC expression. Interestingly,
sdg8 mutants flowered earlier than the vip4 mutant (Table 1), indicating
that SDG8 does not function in an exclusively VIP4-dependent manner
and that additional VIP4-independent pathways exist in SDG8-mediated
flowering-time control.
RT–PCR analysis demonstrated that SDG8 is ubiquitously expressed
in different organs of Arabidopsis plants (data not shown). However, in
situ hybridization showed that SDG8 transcripts were more abundant
in young tissues but were distributed in a non-uniform manner in clus-
ters of cells (Fig. 4), implying that SDG8 expression is highly regulated.
The presence of abundant SDG8 transcripts in the inflorescence (ovules
as shown in Fig. 4c), where FLC expression is undetectable15, further
suggests that SDG8 may have additional roles that are independent of
FLC activation. Further exploration will define these additional roles of
SDG8 in plant growth and development and determine whether H3K36
methylation also has a critical function in other eukaryotes.
METHODS
Plant materials and growth conditions. All Arabidopsis alleles were derived
from the Columbia (Col) ecotype. sdg8-1, sdg8-2 and sdg8-3 alleles corre-
spond to Salk_065480, Salk_026442 and SAIL_880_B03, respectively, of the
T-DNA insertion strains from the Arabidopsis Biological Resource Center
(ABRC) and Nottingham Arabidopsis Stock Centre (NASC) (http://arabi-
dopsis.org). The vip4 mutant used in this study corresponds to Salk_122755,
which contains a T-DNA inserted in exon 4 of the VIP4 (At5g61150) gene.
Seeds were obtained from ABRC and NASC. The studied photoperiods were
8 h light and 16 h dark for short-day, 12 h light and 12 h dark for medium-
day, and 16 h light and 8 h dark for long-day. For vernalization tests, four-
day-old seedlings on MS media were maintained under short-day conditions
at 4 °C for 40 days then transferred to soil and grown under long-day at
22 °C. Flowering time was measured from a developmental perspective, as
the total number of rosette leaves15.
RT –PCR analysis. Total RNA was extracted with TRI Reagent (Molecular Research
Center Inc., Cincinnati, OH) from seedlings grown under medium-day conditions.
Contaminating DNA was eliminated by treatment with RNase-free RQ1-DNase
(Promega, Charbonnières, France). RT–PCR was performed as described17, using
gene specific primers (see Supplementary Information, Table S2).
Histone extraction and western-blot analysis. Histone-enriched protein extrac-
tion from 20-day-old seedlings and western blot analysis were performed as
described26. Upstate (Euromedex, Mundolshein, France) antibodies against his-
tone H3 dimethyl-Lys 36 (Catalogue no. 07-369), H3 dimethyl-Lys 4 (Catalogue
no. 07-030), H3 dimethyl-Lys 9 (Catalogue no. 07-212), H3 dimethyl-Lys 27
(Catalogue no. 07-322) and H3 (Catalogue no. 05-499) were used.
Chromatin immunoprecipitation. 20-day-old seedlings grown under medium-
day conditions were used for chromatin immunoprecipitation experiments as
previously des cribed21. Anti-acetyl-Histone H3-Lys 9 (Upstate Catalogue no. 06-
942) was used to detect acetylation. PCR detection of FLC regions2, ACTIN and
Ta3 (Ref. 21) was performed as described.
In sit u hybridization. Digoxigenin labelling of RNA probes, tissue preparation
and in situ hybridization were performed as described27. Tissue sections were
8 µm thick. A plasmid containing the 260 bp 3-untranslated region (UTR) of
SDG8 was used to prepare sense (negative control) and antisense probes.
Gene constructs and plant transformation. A 12 kb Nhe I restriction frag-
ment containing the SDG8 gene together with the upstream 3023 bp and
downstream 308 bp non-coding regions was cloned from the BAC clone
H3K36m >
H3K27m >
H3 >
H3K4m >
H3K9m >
FLC
FT
SOC1
A
CTIN
sdg8-1
sdg8-1
sdg8-2
Col
Col
sdg8-1
Col
sdg8-1
Col
sdg8-1
Col
DAG 4 8 12 16 20 24
ab
Figure 2 Effects of sdg8 on gene expression and histone H3 methylation.
(a) RT–PCR analysis of floral-pathway gene expression. Wild-type Col and
mutant sdg8-1 plants grown under 12-h light and 12-h dark photoperiods
were analysed for FLC, SOC1 and FT expression at 4, 8, 12, 16, 20 and 24
DAG. ACTIN served as an internal control. (b) Western blot analysis of lysine
methylation of histone H3. Histone-enriched protein extracts from 20-day-
old wild-type Col, and sdg8-1 and sdg8-2 mutant plants grown under a 12-h
light and 12-h dark photoperiod were probed with antibodies that specifically
recognize the indicated forms of histone H3.
a
b
PromoterExon 1 Exon 7
1 kb
Intron 1
AB C D E F G
Col
sdg8-1
Col
sdg8-1
Col
sdg8-1
Col
sdg8-1
Col
sdg8-1
Col
sdg8-1
Col
sdg8-1
Input H3K36mH3K4mH3K9mH3K27mH3K9a
FLC
A
B
C
D
E
F
G
ACTIN
Ta3
Figure 3 ChIP analysis of histone H3 modifications at the FLC locus.
(a) A schematic representation of FLC gene structure indicating the regions
examined by ChIP. Black boxes represent exons; lines represent promoter
and introns; bars labelled A–G represent regions amplified by PCR.
(b) ChIP analysis with antibodies against modified histone H3. Twenty-day-
old wild-type Col and mutant sdg8-1 plants grown under 12-h light and 12-h
dark photoperiods were analysed with antibodies that specifically recognize
modified histone H3, as indicated at the top of each panel. Input and no
antibody (–) are shown as positive and negative controls, respectively. FLC
regions and controls (ACTIN for an actively expressed gene and Ta3 for a
repressed transposon) were amplified by PCR.
ncb1329.indd 1259ncb1329.indd 1259 1/12/05 3:26:44 pm1/12/05 3:26:44 pm
Nature Publishing Group
© 2005
© 2005 Nature Publishing Group
1260 NATURE CELL BIOLOGY VOLUME 7 | NUMBER 12 | DECEMBER 2005
LETTERS
T14N5 (ABRC) into pCambia1300 (CAMBIATM, Canberra, NSW) resulting
in the binary vector pCSDG8. FLC cDNA containing the 88 bp 5-UTR, the
591 bp open reading frame (ORF) and the 238 bp 3-UTR was amplified by
RT–PCR and cloned behind the CaMV 35S promoter in pBinHyg-TX28, result-
ing in the binary vector p35S:FLC. The binary vectors were introduced into
Agrobacterium tumefaciens and the resulting strains were used to transform
Arabidopsis plants29.
Note: Supplementary Information is available on the Nature Cell Biology website.
ACKNOWLEDGEMENTS
The authors thank the Nottingham Arabidopsis Stock Center and the Arabidopsis
Biological Resource Center for seeds and BAC clone, and K. Richards and
H. Huang for helpful comments on the manuscript. Z.Z. is supported by a
postdoctoral fellowship from the French Ministère de la Recherche et des Nouvelles
Technologies, Y.Y. by a research training fellowship from the Association Franco-
Chinoise pour la Recherche Scientifique & Technique (AFCRST), and C.W. by a
four-month-visiting fellowship from the United Nations Educational, Scientific and
Cultural Organization (UNESCO). Research in the W.-H.S. laboratory is supported
by the Centre National de la Recherche Scientifique (CNRS).
COMPETING INTERESTS STATEMENT
The authors declare that they have no competing financial interests.
Published online at http://www.nature.com/naturecellbiology
Reprints and permissions information is available online at http://npg.nature.com/
reprintsandpermissions
1. Boss, P. K., Bastow, R. M., Mylne, J. S. & Dean, C. Multiple pathways in the decision to
flower: enabling, promoting, and resetting. Plant Cell 16, S18–S31 (2004).
2. Bastow, R. et al. Vernalization requires epigenetic silencing of FLC by histone methyla-
tion. Nature 427, 164–167 (2004).
3. He, Y., Doyle, M. R. & Amasino, R. M. PAF1-complex-mediated histone methylation of
FLOWERING LOCUS C chromatin is required for the vernalization-responsive, winter-
annual habit in Arabidopsis. Genes Dev. 18, 2774–2784 (2004).
4. Sung, S. & Amasino, R. M. Vernalization in Arabidopsis thaliana is mediated by the PHD
finger protein VIN3. Nature 427, 159–164 (2004).
5. Luger, K., et al. Crystal structure of the nucleosome core particle at 2.8 A resolution.
Nature 389, 251–260 (1997).
6. Strahl, B. D. & Allis, C. D. The language of covalent histone modifications. Nature 403,
41–45 (2000).
7. Sims III, R. J., Nishioka, K. & Reinberg, D. Histone lysine methylation: a signature for
chromatin function. Trends Genet. 19, 629–639 (2003).
8. Krogan, N. J. et al. Methylation of histone H3 by Set2 in Saccharomyces cerevisiae
is linked to transcriptional elongation by RNA polymerase II. Mol. Cell. Biol. 23,
4207–4218 (2003).
9. Xiao, T. et al. Phosphorylation of RNA polymerase II CTD regulates H3 methylation in
yeast. Genes Dev. 17, 654–663 (2003).
10. Bannister, A. J. et al. Spatial distribution of di- and tri-methyl lysine 36 of histone H3
at active genes. J. Biol. Chem. 280, 17732–17736.
11. Strahl, B. D. et al. Set2 is a nucleosomal histone H3-selective methyltransferase that
mediates transcriptional repression. Mol. Cell. Biol. 22, 1298–1306 (2002).
12. Rayasam, G. V. et al. NSD1 is essential for early post-implantation development and
has a catalytically active SET domain. EMBO J. 22, 3153–3163 (2003).
13. Perry, J. & Zhao, Y. The CW domain, a structural module shared amongst vertebrates, verte-
brate-infecting parasites and higher plants. Trends Biochem. Sci. 28, 576–580 (2003).
14. Alonso, J. M. et al. Genome-wide insertional mutagenesis of Arabidopsis thaliana.
Science 301, 653–657 (2003).
15. Michaels, S. D. & Amasino, R. M. FLOWERING LOCUS C encodes a novel MADS domain
protein that acts as a repressor of flowering. Plant Cell 11, 949–956 (1999).
16. Sheldon, C. C. et al. The FLF MADS box gene: a repressor of flowering in Arabidopsis
regulated by vernalization and methylation. Plant Cell 11, 445–458 (1999).
17. Lee et al. The AGAMOUS-LIKE 20 MADS domain protein integrates floral inductive
pathways in Arabidopsis. Genes Dev. 14, 2366–2376 (2000).
18. Samach, A. et al. Distinct roles of CONSTANS target genes in reproductive development
of Arabidopsis. Science 288, 1613–1616 (2000).
19. Fischle, W., Wang, Y. & Allis, C. D. Histone and chromatin cross-talk. Curr. Opin. Cell
Biol. 15, 172–183 (2003).
20. He, Y., Michaels, S. D. & Amasino, R. M. Regulation of flowering time by histone
acetylation in Arabidopsis. Science 302, 1751–1754 (2003).
21. Johnson, L. M., Cao, X. & Jacobsen, S. E. Interplay between two epigenetic marks: DNA
methylation and histone H3 lysine 9 methylation. Curr. Biol. 12, 1360–1367.
22. Sheldon, C. C., Conn, A. B., Dennis, E. S. & Peacock, W. J. Different regulatory regions
are required for the vernalization-induced repression of FLOWERING LOCUS C and for
the epigenetic maintenance of repression. Plant Cell 14, 2527–2537 (2002).
23. Ausin, I., et al. Regulation of flowering time by FVE, a retinoblastoma-associated
protein. Nature Genet. 36, 162–166 (2004).
24. Zhang, H. & van Nocker, S. The VERNALIZATION INDEPENDENCE 4 gene encodes a
novel regulator of FLOWERING LOCUS C. Plant J. 31, 663–673 (2002).
25. Oh, S., Zhang, H., Ludwig, P. & van Nocker, S. A mechanism related to the yeast
transcriptional regulator Paf1c is required for expression of the Arabidopsis FLC/MAF
MADS box gene family. Plant Cell 16, 2940–2953 (2004).
26. Yu, Y., Dong, A. & Shen, W.-H. Structure-function characterization of the tobacco SET-
domain protein NtSET1 reveals its role in histone methylation, chromatin binding and
segregation. Plant J. 40, 699–711 (2004).
27. Jackson, D. in Molecular Plant Pathology: A Practical Approach (eds Bowles, D. J., Gurr,
S. J. & McPherson, M.) 163–174 (Oxford University Press, Oxford, UK 1991).
28. Gatz, C. Novel inducible/repressible gene expression systems. Methods Cell Biol. 50,
411–424 (1995).
29. Clough, S. J. & Bent, A. F. Floral dip: a simplified method for Agrobacterium-mediated
transformation of Arabidopsis thaliana. Plant J. 16, 735–743 (1998).
a
b
c
SAM
L2
ads
abs
ms
ov ov
ov
ov
Placenta
ov
gp
ms
L3
L1
Figure 4 In situ hybridization analysis of SDG8 expression. (a) A longitudinal
section through a vegetative apex. The shoot apical meristem (SAM) and
leaves (L1 to L3) are indicated. (b) A longitudinal section through a floral
apex. The adaxial sepal (ads), the medial stamen (ms), the gynoecial
primordium (gp) and the abaxial sepal (abs) are indicated. (c) A longitudinal
section through part of a gynoecium. The ovule primodium (ov) and the
placenta are indicated. Hybridization signals are dark brown areas. A negative
control using a sense probe did not generate detectable hybridization signals
(data not shown). Wild-type Col plants grown under a 12-h light and 12-h
dark photoperiod were analysed. Scale bars represent 10 µm in all panels.
ncb1329.indd 1260ncb1329.indd 1260 1/12/05 3:26:48 pm1/12/05 3:26:48 pm
Nature Publishing Group
© 2005
© 2005 Nature Publishing Group
NATURE CELL BIOLOGY ADVANCE ONLINE PUBLICATION 1
ERRATUM
Owing to a technical error, the pages of this manuscript were origi-
nally mis-numbered by a 100 pages.  is has now been corrected on-
line.  e corrected online manuscript is numbered 100 pages higher
than the mis-numbered version.
© 2005 Nature Publishing Group
SUPPLEMENTARY INFORMATION
WWW.NATURE.COM/NATURECELLBIOLOGY 1
Figure S1 Alignment of SDG8 and SET2 amino acid sequences.
© 2005 Nature Publishing Group
© 2005 Nature Publishing Group
SUPPLEMENTARY INFORMATION
2 WWW.NATURE.COM/NATURECELLBIOLOGY
Figure S2 Effect of vernalization on SDG8 and FLC expression and correlation of late flowering with 35S:FLC overexpression.
© 2005 Nature Publishing Group
© 2005 Nature Publishing Group
SUPPLEMENTARY INFORMATION
WWW.NATURE.COM/NATURECELLBIOLOGY 3
Table S1 Flowering time of 35S:FLC transgenic T1 wild-type (Col) and mutant (sdg8-1 and sdg8-2) plants at long-day photoperiods
Phenotype No. of rosette leaves No. of plants Mean rosette leaf number per plant
Col control
93
10.1 ± 0.9 (10)
10 3
11 4
Col
35S:FLC
early 8 2 8.0 ± 0 (2)
unchanged
10 4
10.5 ± 0.8 (6)
11 1
12 1
very late
41 1
69.8 ± 17.2 (6)
57 1
76 1
80 2
85 1
sdg8-1 control
72
8.1 ± 0.7 (12)
87
93
sdg8-1
35S:FLC
early 51 5.8 ± 0.4 (6)
65
unchanged
74
7.8 ± 0.8 (9)
83
92
late 10 2
11 1
13 3 12.0 ± 1.5 (7)
14 1
23 1
very late 25 1 35.0 ± 12.1 (4)
39 1
53 1
sdg8-2 control
72
8.0 ± 0.6 (12)
88
92
sdg8-2
35S:FLC
early 51 5.9 ± 0.3 (8)
67
unchanged
77
8.1 ± 0.9 (17)
82
98
late
10 5
11.5 ± 1.4 (11)
11 1
12 3
13 2
14 2
18 1
26 1
very late 40 2 38.3 ± 13.9 (6)
44 1
62 1
© 2005 Nature Publishing Group
© 2005 Nature Publishing Group
SUPPLEMENTARY INFORMATION
4 WWW.NATURE.COM/NATURECELLBIOLOGY
Table S2 Gene-specifi c primers used in RT–PCR analysis
Gene Forward Primer Reverse Primer Corresponding region (nucleotide after stop codon = +1)
SDG8 AAACAGAGTGTTTCCTCCATGG TAAAGAAAGGAGAGGGATAGG –101 to +163
FLC AAGCTGAGATGGAGATGTC TCGATGCAATTCTCACACG –73 to +238
SOC1 CGAGCAAGAAAGACTCAAGTGTTTAAGG GAAGTGACTGAGAGAGAGAGAGTGAG –230 to +149
FT TACGAAAATCCAAGTCCCACTG AAACTCGCGAGTGTTGAAGTTC –206 to -87
ACTIN AAGTCATAACCATCGGAGCTG ACCAGATAAGACAAGACACAC –392 to +21
© 2005 Nature Publishing Group
© 2005 Nature Publishing Group

Supplementary resources (2)

... For example, loss of MET1 function in Arabidopsis leads to reduced methylation of the epigenetically imprinted AtFWA gene, resulting in a late-flowering phenotype (Kankel et al., 2003). Additionally, reduced methylation of the AtFLC promoter results in higher level of its expression, which leads to lower expression of AtFT and a delay in flowering time (Jean Finnegan et al., 2005;Sheldon et al., 2000;Zhao et al., 2005). Recently, Williams et al. (2022) found that drdd (dme; ros1; dml2; and dml3) quadruple mutants of Arabidopsis thaliana exhibited hypermethylated regions in the leaves and an early flowering phenotype due to 5′-hypermethylation and transcriptional down-regulation of AtFLC genes. ...
... Zm00001d023333 (Chr10:3606398-3621010, sdg127, SET domain gene127) encodes a histone-lysine N-methyltransferase ATXR7. Another two SET domain family genes, SET domain group 8 (SDG 8) in Arabidopsis thaliana (Zhao et al., 2005) and SDG712 in rice (Zhang et al., 2021) respectively. The above research findings suggest that Zm00001d023333 we identified in this study might affect PH and EH by delaying flowering time and lengthening vegetative growth. ...
Article
Full-text available
Plant height (PH) and ear height (EH) are important traits associated with biomass, lodging resistance, and grain yield in maize. There were strong effects of genotype x environment interaction (GEI) on plant height and ear height of maize. In this study, 203 maize inbred lines were grown at five locations across China’s Spring and Summer corn belts, and plant height (PH) and ear height (EH) phenotype data were collected and grouped using GGE biplot. Five locations fell into two distinct groups (or mega environments) that coincide with two corn ecological zones called Summer Corn Belt and Spring Corn Belt. In total, 73,174 SNPs collected using GBS sequencing platform were used as genotype data and a recently released multi-environment GWAS software package IIIVmrMLM was employed to identify QTNs and QTN x environment (corn belt) interaction (QEIs); 12 and 11 statistically significant QEIs for PH and EH were detected respectively and their phenotypic effects were further partitioned into Add*E and Dom*E components. There were 28 and 25 corn-belt-specific QTNs for PH and EH identified, respectively. The result shows that there are a large number of genetic loci underlying the PH and EH GEIs and IIIVmrMLM is a powerful tool in discovering QTNs that have significant QTN-by-Environment interaction. PH and EH candidate genes were annotated based on transcriptomic analysis and haplotype analysis. EH related-QEI S10_135 (Zm00001d025947, saur76, small auxin up RNA76) and PH related-QEI S4_4 (Zm00001d049692, mads32, encoding MADS-transcription factor 32), and corn-belt specific QTNs including S10_4 (Zm00001d023333, sdg127, set domain gene127) and S7_1 (Zm00001d018614, GLR3.4, and glutamate receptor 3.4 or Zm00001d018616, DDRGK domain-containing protein) were reported, and the relationship among GEIs, QEIs and phenotypic plasticity and their biological and breeding implications were discussed.
... Histone methylation played essential roles regulating of FLC expression. SET DOMAIN GROUP 8 (SDG 8) delayed flowering by promoting FLC expression and required the methylation of histone H3 at Lys36 (H3K36) (Zhao et al. 2005). ARABIDOPSIS TRITHORAX-RELATED 1 (ATX1) is dynamically activated FLC expression via histone H3 at Lys4 (H3K4) at FLC chromatin (Pien et al. 2008). ...
Article
Full-text available
Key message Two major QTLs for bolting time in radish were mapped to chromosome 02 and 07 in a 0.37 Mb and 0. 52 Mb interval, RsFLC1 and RsFLC2 is the critical genes. Abstract Radish (Raphanussativus L.) is an important vegetable crop of Cruciferae. The premature bolting and flowering reduces the yield and quality of the fleshy root of radish. However, the molecular mechanism underlying bolting and flowering in radish remains unknown. In YZH (early bolting) × XHT (late bolting) F2 population, a high-density genetic linkage map was constructed with genetic distance of 2497.74 cM and an average interval of 2.31 cM. A total of nine QTLs for bolting time and two QTLs for flowering time were detected. Three QTLs associated with bolting time in radish were identified by QTL-seq using radish GDE (early bolting) × GDL (late bolting) F2 population. Fine mapping narrowed down qBT2 and qBT7.2 to an 0.37 Mb and 0.52 Mb region on chromosome 02 and 07, respectively. RNA-seq and qRT-PCR analysis showed that RsFLC1 and RsFLC2 were the candidate gene for qBT7.2 and qBT2 locus, respectively. Subcellular localization exhibited that RsFLC1 and RsFLC2 were mainly expressed in the nucleus. A 1856-bp insertion in the first intron of RsFLC1 was responsible for bolting time. Overexpression of RsFLC2 in Arabidopsis was significantly delayed flowering. These findings will provide new insights into the exploring the molecular mechanism of late bolting and promote the marker-assisted selection for breeding late-bolting varieties in radish.
... All active ANMs catalyze the O-monomethylarginine (MMA). In Arabidopsis thaliana, several histone H3 lysine methyltransferases have been identi ed with important roles in gene silencing and growth regulation which con rm the results of this experiment (Jackson et al., 2002;Kim et al., 2005;Zhao et al., 2005). Recently, A. thaliana protein arginine methyltransferase 5 (AtANM5) was shown to symmetrically demethylate histone H4R3 Wang et al., 2007). ...
Preprint
Full-text available
The low performance of date palm cv. Barhi resulting from tissue culturing is one of the main challenges in the production of these plants. On the other hand, the final yield of the plant depends on various metabolic and biochemical factors which are caused by gene expression. The plant reacts to environmental factors to survive in different growth and environmental conditions through gene expression. This experiment was conducted to investigate the relationship between the expression of certain genes before and up to two weeks after pollination with the yield of off-shoot and tissue culture of Phoenix dactylifera L. cv. Barhi (10-year-old). Off-shoot and tissue-culture date palms were pollinated with Red Ghanami pollen, Green Ghanami pollen, and Green Ghanami + Red Ghanami pollen (50:50) based on a factorial design in randomized complete blocks with three replications. To this end, the relative expression levels of Histone acetyltransferase HAC1-like (LOC103717600) (HAC 1), Arginine N-methyltransferase 6.1 (LOC103716582) (ANM), TIME FOR COFFEE-like (LOC103716450) (TFC), Homeobox-leucine zipper protein HOX9-like (LOC103703962) (HOX 9), MADS-box transcription factor 2-like (LOC103702602) (GLO 1), and MADS-box transcription factor 16 (LOC103701267) (DEF 1) genes were examined using qRT-PCR method in 3 times: (1) Before pollination, (2) 1 week after pollination, and (3) 2 weeks after pollination. The fruit yield was measured at 180 days after pollination (the end of the experiment). The results showed that the tissue culture palms had lower yield but higher relative genes expression in all pollen treatments compared to off-shoot palms. Off-shoot date palm pollinated with Red Ghanami pollen (5.93 tons/ha) showed the highest yield while the lowest yield was recorded in tissue culture date palm pollinated with Green Ghanami pollen (2.09 tons/ha). The relative expression of the six studied genes significantly increased in all treatments two weeks after pollination. Two weeks after all investigated treatments, the GLO gene had the highest expression, while the HAC gene showed the lowest relative expression. The relationship between the yield and the genes expression in the three studied times showed that the HOX gene had no considerable effect on flowering and yield. The ANM and TFC genes expression before pollination exhibited a positive relationship, whereas the expression of HAC, DEF, and GLO genes had a negative relationship with 1 week after pollination. Two weeks after pollination, the expression of the ANM gene revealed a significant positive relationship with the final yield of the date palm. Moreover, the correlation results showed the prominent role of the genes in different stages of growth and yield of date palms by influencing the expression of each other. In general, it is possible to improve the yield of the tissue culture date palm by affecting the expression of the genes in specific stages based on their function.
... H3K36 methylation is predominantly carried out by the KMT3 subfamily, with SDG8/ASHH2/KMT3;1a serving as the representative gene responsible for modulating the owering process. In Arabidopsis, Brassica napus and other cruciferous plants, KMT3;1 represses owering by promoting FLC expression and downregulating the expression of FT (FLOWERING LOCUS T) and SOC1 (SUPPRESSOR OF OVEREXPRESSION OF CONSTANS 1) [31][32][33]. However, in rice, KMT3;1 promotes gene expression in the chromatin region of brassinolactin-related genes via H3K36me2/me3, accelerating owering [34,35]. ...
Preprint
Full-text available
Background SET DOMAIN GROUP (SDG) proteins act as histone lysine methyltransferases, which play pivotal roles in a variety of developmental processes, such as flowering determination, by modifying chromatin structure to regulate related gene transcription. Although there is extensive evidence that histone methylation plays a key role in plant growth and development, especially H3K36 methylation in Arabidopsis, little information is available for cotton. Results A total of 86 SDG genes were identified through genome-wide analysis of the Gossypium hirsutum genome. Subsequently, we thoroughly studied Arabidopsis, G. arboreum, and G. raimondii homologs and conducted a comprehensive analysis of their gene structure and conserved domain organization. A total of 12 conserved domains were detected in 86 GhSDGs and classified into seven main classes to clarify their evolutionary relationship, which was unevenly distributed across 25 chromosomes of G. hirsutum. Transcriptome data analysis of tissue and developmental stage expression showed that the GhSDG gene was expressed in different tissues, with high expression levels in the flowering stage and floral organs such as stamens and pistils, especially for KMT3 classes involved in activation of flowering, showing dramatically changed expression patterns. Subsequently, we performed reverse transcription-quantitative PCR for 8 KMT3 classes of GhSDGs in eight tissues (root, stem, leaf, sepal, petals, bud, stamens and pistils). In addition, two predicted GhSDG genes (GhKMT3;1a and GhKMT3;2a) were shown to regulate cotton development by virus-induced gene silencing. Budding and flowering were delayed in upland cotton plants with silenced GhKMT3;1a and GhKMT3;2a, and plant height was also lowered. We found that GhKMT3;1a and GhKMT3;2a-mediated H3K36 methylation regulates the expression of flowering-related genes and plays an important role in the flowering of upland cotton. Conclusion We comprehensively identified 86 GhSDG genes in upland cotton. Our data indicate that GhKMT3;1a and GhKMT3;2a-mediated H3K36 methylation regulate the expression of flowering-related genes and play an important role in the growth and development of cotton plants. These findings may also lay a foundation for breeding early-maturing cotton varieties in the future.
Article
Plants necessitate a refined coordination of growth and development to effectively respond to external triggers for survival and successful reproduction. This intricate harmonization of plant developmental processes and adaptability hinges on significant alterations within their epigenetic landscapes. In this review, we first delve into recent strides made in comprehending underpinning the dynamics of histones, driven by both internal and external cues. We encapsulate the prevailing working models through which cis/trans elements navigate the acquisition and removal of histone modifications, as well as the substitution of histone variants. As we look ahead, we anticipate that delving deeper into the dynamics of epigenetic regulation at the level of individual cells or specific cell types will significantly enrich our comprehension of how plant development unfolds under the influence of internal and external cues. Such exploration holds the potential to provide unprecedented resolution in understanding the orchestration of plant growth and development.
Preprint
As in origami, morphogenesis in living systems heavily relies on tissue curving and folding, through the interplay between biochemical and biomechanical cues. In contrast, certain organs maintain their flat posture over several days. Here we identified a pathway, which is required for the maintenance of organ flatness, taking the sepal, the outermost floral organ, in Arabidopsis as a model system. Through genetic, cellular and mechanical approaches, our results demonstrate that global gene expression regulator VERNALIZATION INDEPENDENCE 4 (VIP4) fine-tunes the mechanical properties of sepal cell walls and maintains balanced growth on both sides of the sepals, mainly by orchestrating the distribution pattern of AUXIN RESPONSE FACTOR 3 (ARF3). vip4 mutation results in softer cell walls and faster cell growth on the adaxial sepal side, which eventually cause sepals to bend outward. Downstream of VIP4, ARF3 works through modulating auxin signaling to down-regulate pectin methylesterase VANGUARD1, resulting in decreased cell wall stiffness. Our work unravels a 3-component module, which relates hormonal patterns to organ curvature, and actively maintains sepal flatness during its growth.
Article
Full-text available
Pineapple (Ananas comosus var. comosus) and ornamental bromeliads are commercially induced to flower by treatment with ethylene or its analogs. The apex is transformed from a vegetative to a floral meristem and shows morphological changes in 8 to 10 days, with flowers developing 8 to 10 weeks later. During eight sampling stages ranging from 6 h to 8 days after treatment, 7961 genes were found to exhibit differential expression (DE) after the application of ethylene. In the first 3 days after treatment, there was little change in ethylene synthesis or in the early stages of the ethylene response. Subsequently, three ethylene response transcription factors (ERTF) were up-regulated and the potential gene targets were predicted to be the positive flowering regulator CONSTANS-like 3 (CO), a WUSCHEL gene, two APETALA1/FRUITFULL (AP1/FUL) genes, an epidermal patterning gene, and a jasmonic acid synthesis gene. We confirm that pineapple has lost the flowering repressor FLOWERING LOCUS C. At the initial stages, the SUPPRESSOR OF OVEREXPRESSION OF CONSTANS 1 (SOC1) was not significantly involved in this transition. Another WUSCHEL gene and a PHD homeobox transcription factor, though not apparent direct targets of ERTF, were up-regulated within a day of treatment , their predicted targets being the up-regulated CO, auxin response factors, SQUAMOSA, and histone H3 genes with suppression of abscisic acid response genes. The FLOWERING LOCUS T (FT), TERMINAL FLOWER (TFL), AGAMOUS-like APETELAR (AP2), and SEPETALA (SEP) increased rapidly within 2 to 3 days after eth-ylene treatment. Two FT genes were up-regulated at the apex and not at the leaf bases after treatment, suggesting that transport did not occur. These results indicated that the ethylene response in pineapple and possibly most bromeliads act directly to promote the vegetative to flower transition via APETALA1/FRUITFULL (AP1/FUL) and its interaction with SPL, FT, TFL, SEP, and AP2. A model based on AP2/ERTF DE and predicted DE target genes was developed to give focus to future research. The identified candidate genes are potential targets for genetic manipulation to determine their molecular role in flower transition.
Article
The nuclear pore complex (NPC) has multiple functions beyond the nucleo-cytoplasmic transport of large molecules. Sub-nuclear compartmentalization of chromatin is critical for gene expression in animals and yeast. However, the mechanism by which the NPC regulates gene expression is poorly understood in plants. Here we report that the Y-complex (Nup107–160 complex, a subcomplex of the NPC) self-maintains its nucleoporin homeostasis and modulates FLOWERING LOCUS C (FLC) transcription via changing histone modifications at this locus. We show that Y-complex nucleoporins are intimately associated with FLC chromatin through their interactions with histone H2A at the nuclear membrane. Fluorescence in situ hybridization assays revealed that Nup96, a Y-complex nucleoporin, enhances FLC positioning at the nuclear periphery. Nup96 interacted with HISTONE DEACETYLASE 6 (HDA6), a key repressor of FLC expression via histone modification, at the nuclear membrane to attenuate HDA6-catalyzed deposition at the FLC locus and change histone modifications. Moreover, we demonstrate that Y-complex nucleoporins interact with RNA polymerase II to increase its occupancy at the FLC locus, facilitating transcription. Collectively, our findings identify an attractive mechanism for the Y-complex in regulating FLC expression via tethering the locus at the nuclear periphery and altering its histone modification.
Article
Full-text available
The very late-flowering behavior of Arabidopsiswinter-annual ecotypes is conferred mainly by two genes,FRIGIDA (FRI) and FLOWERING LOCUS C(FLC). A MADS-domain gene, AGAMOUS-LIKE 20(AGL20), was identified as a dominant FRI suppressor in activation tagging mutagenesis. Overexpression of AGL20suppresses not only the late flowering of plants that have functionalFRI and FLC alleles but also the delayed phase transitions during the vegetative stages of plant development. Interestingly, AGL20 expression is positively regulated not only by the redundant vernalization and autonomous pathways of flowering but also by the photoperiod pathway. Our results indicate that AGL20 is an important integrator of three pathways controlling flowering in Arabidopsis. Keywords • Flowering • MADS domain protein • AGL20 • phase transition • activation tagging
Article
Full-text available
The X-ray crystal structure of the nucleosome core particle of chromatin shows in atomic detail how the histone protein octamer is assembled and how 146 base pairs of DNA are organized into a superhelix around it. Both histone/histone and histone/DNA interactions depend on the histone fold domains and additional, well ordered structure elements extending from this motif. Histone amino-terminal tails pass over and between the gyres of the DNA superhelix to contact neighbouring particles. The lack of uniformity between multiple histone/DNA-binding sites causes the DNA to deviate from ideal superhelix geometry.
Article
Molecular plant pathology has directly benefited from advances in modern molecular techniques. These techniques have been applied both to pathogens and plants, enhancing our understanding of the organisms themselves and of the complex interactions which determine compatability between them and their host plants. This volume and its companion together provide the first comprehensive guide to the latest molecular techniques as well as the established approaches to the subject. Researchers in plant science, molecular biology, and related areas will find the two volumes of Molecular Plant Pathology to be an invaluable experimental resource for this exciting and fast-moving field, providing a wealth of easy to follow protocols supported by expert advice and guidance.
Article
A MADS box gene, FLF (for FLOWERING LOCUS F), isolated from a late-flowering, T-DNA–tagged Arabidopsis mutant, is a semidominant gene encoding a repressor of flowering. The FLF gene appears to integrate the vernalization-dependent and autonomous flowering pathways because its expression is regulated by genes in both pathways. The level of FLF mRNA is downregulated by vernalization and by a decrease in genomic DNA methylation, which is consistent with our previous suggestion that vernalization acts to induce flowering through changes in gene activity that are mediated through a reduction in DNA methylation. The flf-1 mutant requires a greater than normal amount of an exogenous gibberellin (GA3) to decrease flowering time compared with the wild type or with vernalization-responsive late-flowering mutants, suggesting that the FLF gene product may block the promotion of flowering by GAs. FLF maps to a region on chromosome 5 near the FLOWERING LOCUS C gene, which is a semidominant repressor of flowering in late-flowering ecotypes of Arabidopsis.
Article
This chapter discusses the novel inducible/repressible gene-expression systems. It describes a method that allows the regulated expression of a transgene upon a specific chemical stimulus. The control mechanism works at the level of the initiation of mRNA synthesis. Transcriptional regulation is generally mediated through the interaction of proteins with certain DNA sequences within a promoter. Plant promoters—like other eukaryotic promoters—contain two functionally distinct classes of elements that affect transcription. The TATA motif, which is typically located 30 nucleotides upstream of the transcription start site, makes up the first class. The second class of promoter elements is composed of small DNA sequences recognized specifically by DNA-binding proteins that synergistically regulate the activity of the transcription initiation complex. Proteins that activate transcription via these elements are modular, containing a DNA binding domain and an activation domain. An alternative approach is to place the general plant transcription machinery under the control of regulatory elements taken from other organisms, which respond to signals usually not encountered by plants.
Article
Background: The heterochromatin of many eukaryotes is marked by both DNA methylation and histone H3 lysine 9 (H3-K9) methylation, though the exact relationship between these epigenetic modifications is unknown. In Neurospora, H3-K9 methylation is required for the maintenance of all known DNA methylation. In Arabidopsis, H3-K9 methylation directs some of the CpNpG and asymmetric methylation. However, it is not known in any organism whether DNA methylation may also direct histone H3 methylation.Results: Using chromatin immunoprecipitation (ChIP) assays, we show that Arabidopsis heterochromatin is associated with H3-K9 methylation. This histone methylation is dependent on the KRYPTONITE and DDM1 genes (SU[VAR]3-9 and SWI2/SNF2 homologs, respectively). We also find that a decrease in DNA methylation does not directly cause a loss of H3-K9 methylation. Instead, a decrease in H3-K9 methylation is only seen at loci where transcription is derepressed.Conclusions: We conclude that DNA methylation does not control the methylation of histone H3-K9. We propose that loss of H3-K9 methylation is due to transcriptional reactivation, coupled with deposition of unmethylated nucleosomes. These findings are consistent with recent observations of DNA replication-independent deposition of histone H3.3 in Drosophila. Our results also suggest that, in Arabidopsis, DNA methylation is sufficient for gene silencing, but H3-K9 methylation is not.
Article
TheAgrobacteriumvacuum infiltration method has made it possible to transformArabidopsis thalianawithout plant tissue culture or regeneration. In the present study, this method was evaluated and a substantially modified transformation method was developed. The labor-intensive vacuum infiltration process was eliminated in favor of simple dipping of developing floral tissues into a solution containingAgrobacterium tumefaciens, 5% sucrose and 500 microliters per litre of surfactant Silwet L-77. Sucrose and surfactant were critical to the success of the floral dip method. Plants inoculated when numerous immature floral buds and few siliques were present produced transformed progeny at the highest rate. Plant tissue culture media, the hormone benzylamino purine and pH adjustment were unnecessary, andAgrobacteriumcould be applied to plants at a range of cell densities. Repeated application ofAgrobacteriumimproved transformation rates and overall yield of transformants approximately twofold. Covering plants for 1 day to retain humidity after inoculation also raised transformation rates twofold. Multiple ecotypes were transformable by this method. The modified method should facilitate high-throughput transformation ofArabidopsisfor efforts such as T-DNA gene tagging, positional cloning, or attempts at targeted gene replacement.
Article
The Agrobacterium vacuum infiltration method has made it possible to transform Arabidopsis thaliana without plant tissue culture or regeneration. In the present study, this method was evaluated and a substantially modified transformation method was developed. The labor-intensive vacuum infiltration process was eliminated in favor of simple dipping of developing floral tissues into a solution containing Agrobacterium tumefaciens, 5% sucrose and 500 microliters per litre of surfactant Silwet L-77. Sucrose and surfactant were critical to the success of the floral dip method. Plants inoculated when numerous immature floral buds and few siliques were present produced transformed progeny at the highest rate. Plant tissue culture media, the hormone benzylamino purine and pH adjustment were unnecessary, and Agrobacterium could be applied to plants at a range of cell densities. Repeated application of Agrobacterium improved transformation rates and overall yield of transformants approximately twofold. Covering plants for 1 day to retain humidity after inoculation also raised transformation rates twofold. Multiple ecotypes were transformable by this method. The modified method should facilitate high-throughput transformation of Arabidopsis for efforts such as T-DNA gene tagging, positional cloning, or attempts at targeted gene replacement.