Download Suppression of RICE TELOMERE BINDING

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Mitosis wikipedia , lookup

List of types of proteins wikipedia , lookup

JADE1 wikipedia , lookup

Transcript
The Plant Cell, Vol. 19: 1770–1781, June 2007, www.plantcell.org ª 2007 American Society of Plant Biologists
Suppression of RICE TELOMERE BINDING PROTEIN1 Results
in Severe and Gradual Developmental Defects Accompanied
by Genome Instability in Rice
Jong-Pil Hong,a Mi Young Byun,a Dal-Hoe Koo,b Kyungsook An,c Jae-Wook Bang,b In Kwon Chung,a
Gynheung An,c and Woo Taek Kima,1
a Department
of Biology, College of Science, Yonsei University, Seoul 120-749, Korea
of Bioscience and Biotechnology, Chungnam National University, Daejeon 305-764, Korea
c National Research Laboratory of Plant Functional Genomics, Department of Life Science,
Pohang University of Science and Technology, Pohang 790-784, Korea
b School
Although several potential telomere binding proteins have been identified in higher plants, their in vivo functions are still
unknown at the plant level. Both knockout and antisense mutants of RICE TELOMERE BINDING PROTEIN1 (RTBP1)
exhibited markedly longer telomeres relative to those of the wild type, indicating that the amount of functional RTBP1 is
inversely correlated with telomere length. rtbp1 plants displayed progressive and severe developmental abnormalities in
both germination and postgermination growth of vegetative organs over four generations (G1 to G4). Reproductive organ
formation, including panicles, stamens, and spikelets, was also gradually and severely impaired in G1 to G4 mutants. Up to
11.4, 17.2, and 26.7% of anaphases in G2, G3, and G4 mutant pollen mother cells, respectively, exhibited one or more
chromosomal fusions, and this progressively increasing aberrant morphology was correlated with an increased frequency
of anaphase bridges containing telomeric repeat DNA. Furthermore, 35S:anti-RTBP1 plants expressing lower levels of
RTBP1 mRNA exhibited developmental phenotypes intermediate between the wild type and mutants in all aspects
examined, including telomere length, vegetative and reproductive growth, and degree of genomic anomaly. These results
suggest that RTBP1 plays dual roles in rice (Oryza sativa), as both a negative regulator of telomere length and one of
positive and functional components for proper architecture of telomeres.
INTRODUCTION
Telomeres, which form the extreme ends of linear eukaryotic
chromosomes, consist of tandem repeats of short, G-rich sequence elements. They are responsible for preserving chromosome integrity and for protection against end-to-end fusion and
recombination with other chromosomes and for exonucleolytic
degradation (Blackburn, 1991; Greider, 1996). Telomeres are
packaged into specialized nucleoprotein complexes, and nonhistone telomere binding proteins appear to be essential components for telomere structure and functions (Collins, 2000;
Blackburn, 2001; Shore, 2001). Telomere binding proteins can be
divided into two distinct classes according to their substrate
specificities. Members of the first group, which bind to the singlestranded 39 extension at telomere termini, are necessary for chromosome capping and telomerase regulation. The second class
comprises double-stranded telomeric repeat binding proteins.
Double-stranded telomere binding proteins, including yeast Rap1
and Taz1 and human TRF1/Pin2 and TRF2, have been exten1 To
whom correspondence should be addressed. E-mail wtkim@
yonsei.ac.kr; fax 82-2-312-5657.
The author responsible for distribution of materials integral to the
findings presented in this article in accordance with the policy described
in the Instructions for Authors (www.plantcell.org) is: Woo Taek Kim
([email protected]).
www.plantcell.org/cgi/doi/10.1105/tpc.107.051953
sively characterized (Shore, 1994; Chong et al., 1995; Bilaud et al.,
1997; Cooper et al., 1997). These proteins are involved in the
control of telomere length and stability in vivo. A negative feedback function that regulates telomere length has been identified
for Rap1 and TRF1/Pin2, which inhibit the action of telomerase at
the ends of telomeres (van Steensel and de Lange, 1997; Wotton
and Shore, 1997; Smogorzewska et al., 2000). Recent studies
have revealed that human telomere complex is composed of six
proteins (de Lange, 2005). Three subunits, TRF1, TRF2, and POT1,
directly interact with TTAGGG telomere repeats, while three
additional subunits, TIN2, TPP1, and Rap1, are associated with a
complex. This human telomere-protein complex, referred to as
shelterin, exhibits DNA remodeling activity and acts to change
the structure of the telomeric DNA, thereby protecting chromosome ends (de Lange, 2005).
Although the plant telomere repeat sequence TTTAGGG is
highly conserved with that of human, its structural compositions
and biological roles have remained much less well understood in
higher plants. Recently, cDNAs encoding potential telomeric
DNA binding proteins have been identified from several plant
species, including rice (Oryza sativa), Arabidopsis thaliana, and
tobacco (Nicotiana tabacum) (Yu et al., 2000; Chen et al., 2001;
Hwang et al., 2001; Yang et al., 2003; Karamysheva et al., 2004;
Kuchar and Fajkus, 2004; Schrumpfova et al., 2004; Hwang and
Cho, 2007). Like human TRF1/Pin2 and TRF2, the predicted proteins possess a single DNA binding Myb-like domain, displaying
Rice Telomere Binding Protein RTBP1
sequence identity to double-stranded telomere binding proteins from yeast and human. In Arabidopsis, there are at least 12
TRF-like (TRFL) genes that fall into two different gene families
(Karamysheva et al., 2004). Bacterially expressed proteins from
TRFL family 1 formed homo- and heterodimers and specifically
interacted with double-stranded plant telomeric repeats in vitro.
These results suggest that plant telomeres are also comprised of
telomere-specific protein components that contribute to their
proper structure and functions. Possible in vivo functions of
telomere binding protein Ng-TRF1 were shown in tobacco BY-2
suspension cells. Both 35S:Ng-TRF1 and 35S:anti-Ng-TRF1 transgenic BY-2 cells exhibited significant changes in telomere length
and cell viability, indicating that Ng-TRF1 is involved in the
control of telomere length and stability in tobacco cultured cells
(Yang et al., 2004). However, in vivo functions of plant telomere
binding proteins have yet to be determined at the plant level.
We aim to elucidate the physiological roles of telomere binding
proteins with respect to telomere structure and functions in higher
plants. In this report, we have used rice as a molecular genetic
and cytological model system and obtained plants containing
a T-DNA copy integrated into the RICE TELOMERE BINDING
PROTEIN1 (RTBP1) gene and transgenic plants carrying 35S:
RTBP1 and 35S:anti-RTBP1 constructs. Pulse-field gel electrophoresis showed that both knockout and antisense lines exhibited markedly longer telomeres compared with those of the
wild-type plants. Homozygous rtbp1 lines displayed progressive
and severe developmental abnormalities in both vegetative and
reproductive organs accompanied by genome instability during
four consecutive generations (G1 to G4). In G2 rtbp1 mutants,
abnormal chromosome bridges were detected in 11.4% of anaphases examined, while the anaphase bridges increased to 17.2
and 26.7% in G3 and G4 mutants, respectively. These results
could lead to a better understanding of RTBP1 function not only
at the cellular level but also in the whole plant and suggest that
RTBP1 participates in the control of telomere length and telomere stability in rice plants.
RESULTS
Isolation of a T-DNA Insertion Mutant of RTBP1 and
Construction of 35S:RTBP1 and 35S:anti-RTBP1
Transgenic Rice Plants
In the past few years, there has been a marked increase of
interest in structure and functions of plant telomeres. Most of the
work has dealt with the identification of the proteins that interact
with telomere sequence. Consequently, a number of proteins
that bind in vitro to oligonucleotides containing telomeric TTTAGGG
repeats have been isolated from several plant species. Until now,
however, only a few of these proteins have been shown to reflect
a preference for a structural feature of plant telomeres in vivo.
RTBP1 was previously identified as a double-stranded telomere
binding protein in rice (Yu et al., 2000). However, its in vivo
function was not known. It was therefore pertinent to establish if
the presence or absence of RTBP1 affects the architecture of
plant telomeres. To define the cellular functions of RTBP1 in rice,
we employed a reverse genetic approach. The gene-specific primer
M1 along with a T-DNA–specific primer LBa-1 were used to
1771
screen DNA pools from a collection of 20,500 T-DNA–transformed
rice mutant lines (Jeon et al., 2000). A 2.2-kb PCR product was
amplified using these primers. After PCR screening of successively smaller mutant pools, we were able to isolate a single rice
line that contained a T-DNA insertion in the RTPB1 gene and
referred to the mutant as rtbp1. The insertion was mapped to the
fifth exon in RTBP1 located on chromosome 2 (line 2D-00626;
Figure 1A). Plants homozygous for the T-DNA insertion were
identified by multiplex PCR with primers M2, M3, and RBa-1
(Figure 1B). T-DNA disruption of RTBP1 was further verified by
RT-PCR, demonstrating that the rice mutant seedlings contained
a negligible amount of both full-length and partial RTBP1 mRNAs
(Figures 1C and 1D). This indicates that rtbp1 is null for the RTBP1
gene. Genomic DNA gel blot analysis using a b-glucuronidase
(GUS) cDNA probe confirmed that the rtbp1 mutant plants
contained a single copy of T-DNA integrated into the RTBP1
gene (Figure 1E). We also established transgenic rice that overexpressed or suppressed RTBP1 by introducing a cauliflower mosaic
virus 35S promoter-pRTBP1 construct in the sense (35S:RTBP1)
or antisense (35S:anti-RTBP1) orientation. Elevated expression
of RTBP1 was observed in 35S:RTBP1 transgenic lines, while a
markedly lower level of mRNA was detected in the 35S:antiRTBP1 plants (Figure 1C).
Knockout Mutation and Suppression of RTBP1 Resulted in
Increased Telomere Length in Rice Plants
To address whether the altered expression of RTBP1 affects
telomere metabolism in rice, we measured the length of telomeres in wild-type, rtbp1, 35S:RTBP1, and 35S:anti-RTBP1 plants.
Total genomic DNA was isolated from each mutant or transgenic
line, digested with the restriction enzyme TaqI, and resolved by
pulse-field gel electrophoresis. DNA on the gel was then blotted
and hybridized with the telomere repeat probe (TTTAGGG)70
(Figure 2) Samples from callus, leaf, and root of wild-type plants
migrated as a typical telomeric smear, with fragments ranging
from ;5 to 10 kb. On the other hand, the rtbp1 plant showed
markedly longer telomeres, whose lengths ranged between 10
and 30 kb in both heterozygous and homozygous G1 mutant
populations (Figure 2A). These long telomeres were further maintained throughout the G2 to G4 plants, reaching a new stable set
point (Figure 2B). Overexpression of antisense RTBP1 mRNA
caused a significant enhancement of telomere elongation in T2
progeny, resulting in telomeres 8 to 25 kb long (Figure 2C). We
interpret these results as evidence that there is an inverse
correlation between the amount of functional RTBP1 and the
length of telomeres in rice plants. On the other hand, telomeric
DNA tracts were maintained at a constant size in T2 35S:RTBP1
transgenic plants (data not shown). This raises the possibility that
the amount of RTBP1 might be saturated in the wild-type plants;
hence, overexpression of RTBP1 did not exert any effects on the
telomere length. With this in mind, we speculate that RTBP1 is
involved in negative regulation of telomere length in rice plants.
Phenotypic Analysis of rtbp1
To explore the mechanistic outcome of decrease in the level of
RTBP1 (Figure 1) and changes in telomere length (Figure 2), we
1772
The Plant Cell
Figure 1. Molecular Characterization of the T-DNA Insertion into the Rice RTBP1 Gene.
(A) Schematic representation of the genomic organization of RTBP1 with the position of the integrated T-DNA. Solid bars indicate exons. Solid lines
represent introns. T-DNA insertion site and orientation of left and right borders are indicated. Gene-specific (M1, M2, M3, M4, M5, and M6) and T-DNA–
specific (LBa-1 and RBa-1) primers used in the genotyping and RT-PCR are shown with arrows.
(B) Genotyping of the rtbp1 mutant plants. A set of gene-specific and T-DNA–specific primers used for genomic PCR are indicated. W, wild type; T,
T-DNA insertion.
(C) RT-PCR analysis of RTBP1 and actin mRNAs in the wild type, homozygous G1 rtbp1 mutant line, and 35S:RTBP1 (#1 and #15), and 35S:anti-RTBP1
(#A and #G) T2 transgenic rice plants using M1 and M4 primers. The actin gene was used as a loading control.
(D) RT-PCR analysis of RTBP1 and actin mRNAs in the wild-type plant and homozygous G1 rtbp1 mutant line. In this experiment, two different primer
sets (M1/M5 and M6/M4) were used for RTBP1, and no RT-PCR products were detected, indicating that rtbp1 was null for the RTBP1 gene. The actin
gene was used as a loading control. M, molecular weight marker.
(E) Genomic DNA gel blot analysis in rice genomic DNA. Total leaf genomic DNAs were isolated from wild-type plants and heterozygous and
homozygous G1 mutant lines. Rice genomic DNAs (10 mg per lane) were digested with various restriction enzymes as indicated, blotted onto nylon
membranes, and hybridized with the 32P-labeled cDNA probe corresponding to the GUS gene. The blots were visualized by autoradiography. M,
molecular weight marker; He, heterozygous G1 rtbp1 mutant; Ho; homozygous G1 rtbp1 mutant.
Rice Telomere Binding Protein RTBP1
1773
Figure 2. Telomere Length in the Wild Type, rtbp1 Mutant Line, and 35S:RTBP1 and 35S:anti-RTBP1 Transgenic Rice Plants.
(A) Telomere length in callus, leaf, and root of wild-type plants and G1 heterozygous and G1 homozygous rtbp1 mutant lines.
(B) Marked increase in telomere length in rtbp1 mutant plants over four successive generations (G1 to G4).
(C) Enhanced telomere elongation in T2 35S:anti-RTBP1 transgenic plants (#A and #G). Pulse-field gel electrophoresis was performed more than five
times independently using randomly chosen individual rice plants.
examined the rtbp1 plants for phenotypic differences from the
wild-type plants. We observed that rtbp1 displayed abnormal
growth of vegetative organs. As presented in Figure 3A, rtbp1 exhibited a gradual retardation in growth of roots and shoots over
four successive generations (G1 to G4). The aberrant growth of
rtbp1 was generally apparent in early stages of development,
resulting in severely shortened seedlings with less pronounced
tap roots and increased numbers of lateral roots under the light
growth conditions through G3 to G4 generations. The germination efficiency of the rtbp1 line, scored as radicle emergence
from the seed coat, was also reduced progressively to 90, 55, 40,
and 35% in G1, G2, G3, and G4, respectively (Figure 3B). Mutant
seeds, which failed to germinate, were analyzed by microscopy.
The results indicate that the G3 rtbp1 mutant seeds may not be
dormant because radicle and coleoptile tissues appeared to be
formed 1 d after imbibition (Figure 3C, b). At 2 d after imbibition,
however, mutant radicle and coleoptile were unable to overcome
the constraints of the surrounding tissues, resulting in the reduction of germination rate (Figure 3C, d). In addition, the root
apical meristem and root cap were not normally developed in
these G3 mutant seeds. The germinating G3 and G4 mutant
seeds were able to develop intact seedlings, but their growth
was significantly retarded compared with that of the wild type.
Consequently, the height and number of leaves of mature
4-month-old G4 homozygous rtbp1 progeny grown in the greenhouse conditions displayed only ;50% to those of wild-type rice
(Figure 3D, left panel). Thus, both the germination and postgermination growth of the rtbp1 mutant line was markedly impaired compared with the wild-type plants.
We next analyzed the development of reproductive organs.
Under our experimental conditions, the wild-type rice contains
10 6 1 panicles per plant. However, the average number of panicle in rtbp1 decreased gradually through the G1 to G4 generations, and the G4 mutant progeny contained only 4 6 1 panicles
per plant (Figure 4A). We also compared the formation of floral
organs in the panicles between the wild-type and rtbp1 line. The
number of flowers per panicle was significantly reduced in the
rtbp1 mutant relative to that of the wild type. Normal rice has 55 6
15 flowers per panicle on average, whereas the flower number
per panicle was reduced markedly to 30 6 8 and 24 6 10 in the
G3 and G4 mutant populations, respectively (Figure 4B). Interestingly, the morphology of rtbp1 flowers appeared to be different
1774
The Plant Cell
Figure 3. Developmental Defects of Vegetative Organs in the rtbp1 Mutant Line and 35S:anti-RTBP1 Transgenic T2 Rice Plants.
(A) Progressive and severe retardation of vegetative growth of the rtbp1 mutant during four successive generations. Wild type, rtbp1 (G1 to G4 generations),
and transgenic T2 35S:anti-RTBP1 (#A and #G) and T2 35S:RTBP1 (#1 and #15) seedlings were germinated and grown in MS agar for 5 d under the light
growth conditions (left panel), and the growth patterns of roots and shoots were monitored (right panel). The values are means 6 SD (n ¼ 50).
Rice Telomere Binding Protein RTBP1
to that of wild-type flowers. The normal flower is typified by one
pair of glumes, one lemma, one palea, and six stamens (Jeon
et al., 2000). The flowers of the rtbp1 mutant, however, showed
varying numbers of stamens; in the G3 mutant population, 6.8%
of rtbp1 flowers contained a reduced number of stamens, while
9.5% of G4 mutant flowers had abnormal stamens (Figure 4C,
Table 1). In some of the G3 and G4 flowers, there is only one
stamen (Figure 4C, d), and sometimes the reduced number of
stamen resulted in the combined anthers in the one filament
(Figure 4C, c and e). In addition, an abnormal structure of spikelets was often detected in the G4 mutant due to the abnormal
growth and structure of palea and lemma (Figure 4D, b, c and d),
while we did not observe such abnormality in the wild-type
flowers. The morphology of glumes, however, was similar to that
of the wild type (Figure 4D). These results indicate that the development of reproductive organs occurs abnormally in the rtbp1
mutant. Overall, our phenotypic analysis implies that suppression of the RTBP1 gene yields severe defects in both vegetative
and reproductive organs of rice plants.
As a next experiment, we investigated the phenotypic differences
of 35S:anti-RTBP1 transgenic plants. Because 35S:anti-RTBP1
lines contained significantly lower levels of RTBP1 transcript relative to the wild type (Figure 1C), we expected that the antisense
plants would also display growth defects. Figures 3A, 3B, and 3D
(right panel) depict the visible alterations in the 35S:anti-RTBP1
transgenic plants during the vegetative stages. As was the case
for the rtbp1 line, the growth and germination ratio of these antisense lines was significantly disturbed in the T2 generation. In
addition, the T2 35S:anti-RTBP1 plants displayed disordered development of reproductive organs, including abnormal number
and structure of panicles, flowers, and stamens (Figures 4A and
4B, Table 1). These defects, however, were generally less pronounced than those seen for the G2 rtbp1 mutant plant; hence,
35S:anti-RTBP1 transgenic lines exhibited a phenotype intermediate between that of wild-type and rtbp1 mutant plants. Thus, it
is consistent with the notion that the proper level of RTBP1 is
required for the normal growth and development of rice plants.
Finally, we examined the phenotype of 35S:RTBP1 transgenic
plants. In contrast with what we observed in 35S:anti-RTBP1, the
RTBP1-overexpressing lines were indistinguishable from wildtype plants in both vegetative and reproductive stages in terms of
development and morphology (Figures 3A, 3B, 4A, 4B, and 4D).
Thus, this is in line with the result that the 35S:RTBP1 plants
possess telomeres of normal length. It is still conceivable that
rice cells might require a much more increased level of RTBP1 for
the alteration of their phenotype, but our experimental conditions
were not optimal to detect such changes. This led us to hypothesize that the basal level of RTBP1 might be already saturated or
1775
high enough to play a functional role in the wild-type plants; hence,
its overexpression may result in no visible effects on the 35S:RTBP1
T2 lines. Taken together, our mutant and transgenic analyses support the notion that RTBP1 plays a critical role in the development of
both vegetative and reproductive organs in rice plants.
Cytogenetic Analysis
Riha et al. (2001) isolated the telomerase-deficient Arabidopsis
At-TERT/ mutants and analyzed their phenotypes. Although
Arabidopsis can survive for up to 10 generations without telomerase, the last five generations show severe and progressive
developmental abnormalities in both vegetative and reproductive
organs, which are associated with massive genome instability.
Detailed cytogenetic studies indicated that genomic instability in
G8 telomerase mutants was caused by multiple rounds of the
breakage-fusion-bridge cycle due to dysfunctional telomeres
(Siroky et al., 2003). In addition, perturbation of At-Pot1 or AtPot2 gene expression, which encode Pot (protection of telomeres)-like proteins in Arabidopsis, induced serious growth
aberrations and sterility with bridged chromosomes and aneuploidy (Shakirov et al., 2005). These results indicate that the developmental anomalies correlated with telomere dysfunction are
intimately tied with meiosis disorders in Arabidopsis. Thus, we
considered the possibility that disturbed phenotypes of the rtbp1
rice mutant line are also accompanied by cytogenetic damage
due to the structural dysfunction of telomeres resulting from the
absence of telomere binding protein RTBP1.
To investigate this possibility, we performed cytogenetic analysis. Meiotic progression in wild-type and rtbp1 pollen mother
cells (PMCs) was monitored via 49,6-diamidino-2-phenylindole
(DAPI) staining of chromosomes. Figures 5A and 5B). show typical meiotic events in wild-type PMCs. All 12 chromosomes from
wild-type cells were clearly separated throughout meiosis, resulting in the complete migration of chromosomes to the opposite pole of the cells in metaphase and anaphase, consistent with
their normal morphology. On the other hand, meiotic anomalies
were often detected in rtbp1 mutant cells. As shown in Table 1
and Figures 5C and 5D, we observed the presence of anaphase
bridges in rtbp1 PMCs, which were not observed in wild-type
cells. Among 70 anaphases examined from G2 mutant cells, eight
anaphases (11.4%) displayed bridged chromosomes, while 17.2%
(10/58) of G3 anaphases exhibited abnormal chromosome fusions (Figure 5C). Finally, in G4 mutant cells, the number of bridges
further rose, and up to 26.7% (16/60) of anaphases contained
one or more chromosome fusions (Figure 5D). Chromosome
bridges were also detected in the mutant metaphase, although
Figure 3. (continued).
(B) Germination ratio of the wild type, rtbp1, and transgenic sense and antisense lines. The germination efficiency was scored as radicle emergence
from the seed coat. The values are means 6 SD (n ¼ 50).
(C) Microscopy analysis of wild-type (a) and G3 rtbp1 mutant (b) seeds at 1 d after imbibition and of wild-type (c) and G3 mutant (d) roots at 2 d after
imbibition. sc, scutellum; cp, coleoptile; ep, epiblast; s, shoot apex; r, radicle; rm, root meristem; rc, root cap. Bars ¼ 500 mm in (a) and (b) and 200 mm in
(c) and (d).
(D) Gross morphology of 4-month-old wild-type, three different generations (G2 to G4) of the rtbp1 mutant, and transgenic T2 35S:anti-RTBP1 (#A and
#G) and T2 35S:RTBP1 (#1 and #15) plants grown under greenhouse conditions.
Figure 4. Developmental Defects of Floral Organs in the rtbp1 Mutant Line and 35S:anti-RTBP1 Transgenic T2 Rice Plants.
(A) Comparison of the wild type and rtbp1 (G2 to G4 generations) after bolting (left panel) and the number of panicle per plant in the wild type, rtbp1
(G2 to G4 generations), T2 35S:anti-RTBP1 (#A and #G), and T2 35S:RTBP1 (#1 and #15) rice (right panel). The values are means 6 SD (n ¼ 30).
(B) Panicle structure (left panel) and the number of flowers per panicle (right panel) in the wild type, rtbp1 (G2 to G4 generations), T2 35S:anti-RTBP1 (#A
and #G), and T2 35S:RTBP1 (#1 and #15) rice. The values are means 6 SD (n ¼ 70).
(C) Close-up view of the stamens of wild-type (a), G3 rtbp1 (b and c), and G4 rtbp1 (d and e) flowers. Multiple anthers in one filament are indicated by
asterisks. c, carpel; s, stamen. Bars ¼ 0.5 cm.
(D) A spikelet from the wild type (a) and the G4 rtbp1 line (b, c, and d). g, glume; l, lemma; p, palea. Bars ¼ 0.5 cm.
Rice Telomere Binding Protein RTBP1
Table 1. Frequency of Abnormal Stamen and Anaphase Bridges
Genotype
Abnormal Stamen
Anaphase Bridge
Wild type
rtbp1 G2
rtbp1 G3
rtbp1 G4
35S:anti-RTBP1 #A
35S:anti-RTBP1 #G
0/60
2/62
4/59
6/63
1/58
3/72
0/80 (0%)
8/70 (11.4%)
10/58 (17.2%)
16/60 (26.7%)
2/73 (2.7%)
3/68 (4.4%)
(0%)
(3.2%)
(6.8%)
(9.5%)
(1.7%)
(4.2%)
the frequency was lower (Figure 5E). Thus, increased frequency
of anaphase bridges during four consecutive generations (G1 to
G4) of rtbp1 PMCs is coupled with the gradually increased severity
of the morphology of the mutant plants over these generations.
Using the fluorescence in situ hybridization (FISH) technique,
Vannier et al. (2006) recently observed anaphase bridges containing telomeric repeats in Arabidopsis tert and atrad50/tert mutant cells, which indicates that the abnormal chromosome-end
fusions include telomeric repeats at the fusion points. To address
whether the anaphase bridges detected in rtbp1 PMCs were the
result of telomere dysfunction, we next performed a FISH experiment using two different kinds of probes: a telomeric DNA repeat
probe containing (TTTAGGG)70 (the same probe as that used
for pulse-field gel electrophoresis) and rice centromere-specific
165-bp CentO satellite DNA (Zhang et al., 2005). As shown in Figure 5H, we detected hybridization signal of the telomeric probe
(green) in anaphase bridge of G4 rtbp1 PMCs, while we were not
able to detect such hybridization of CentO probe (red), suggesting that the abnormal anaphase bridges in rtbp1 PMCs contain
telomeric DNAs. Overall, these results support the notion that
deficiency of RTBP1 resulted in telomere dysfunction, which in
turn caused impaired growth of rtbp1 plants. In addition, chromosome bridges were also observed in the T2 generation of
35S:anti-RTBP1 lines (2.7 to 4.4%) (Figure 4F, Table 1). As was
the case for development of vegetative and reproductive organs,
the frequency of anaphase bridges in antisense lines was intermediate between the wild-type and G2 rtbp1 plants. These cytogenetic results argue that RTBP1 is crucially involved in the
positive control of telomere stability in rice plants.
DISCUSSION
Telomeres have highly conserved architecture composed of
tandemly arrayed G-rich sequences of DNA that are synthesized and maintained by the action of telomerase at the extremity of eukaryotic linear chromosomes. Telomeres of higher
plants are made up of typical TTTAGGG arrays, although a few
plants exceptionally contain TTAGGG vertebrate-type repeats
(Weiss-Schneeweiss et al., 2004) or no detectable telomeric repeats (Pich et al., 1996). In humans, telomere structure and telomerase activity are closely coupled with the program of cellular
proliferation, differentiation, senescence, and tumor growth. Despite limited insight into the plant telomeres compared with mammalian systems, plant telomeres have recently attracted much
interest, since it is becoming increasingly apparent that the function and architectural features of telomeres as well as the pat-
1777
terns of telomerase regulation are largely conserved in higher
plants and animals (reviewed in Riha and Shippen, 2003a; McKnight
and Shippen, 2004; Gallego and White, 2005; Lamb et al., 2007;
Zellinger and Riha, 2007). As in human and other animal species,
the stable maintenance of telomeres is essential for growth and
development in Arabidopsis, and a number of nuclear proteins,
such as RAD50, KU70/80, MRE11, Pot1/2, ATM, and ATR, have
been shown to be involved in telomere regulation (Gallego and
White, 2001; Bundock et al., 2002; Bundock and Hooykaas, 2002;
Riha and Shippen, 2003b; Gallego et al., 2003; Puizina et al.,
2004; Shakirov et al., 2005; Vespa et al., 2005; Vannier et al.,
2006). Disruption of the genes encoding these telomere-related
proteins causes strikingly different developmental phenotypes
accompanied by an abrupt onset of genome instability. Thus,
protection and homeostasis of telomeres are crucial for the
normal life cycle of Arabidopsis.
In addition to the above-mentioned telomere-related proteins,
genes for putative double-stranded telomeric repeat binding factors have been isolated in several plant species. The majority of
the work, however, focused on the in vitro characterization of the
encoded proteins, which results in a lack of information about in
vivo functions of the proteins (Yu et al., 2000; Chen et al., 2001;
Hwang et al., 2001; Yang et al., 2003; Karamysheva et al., 2004;
Kuchar and Fajkus, 2004; Schrumpfova et al., 2004). Recently, it
was reported that telomere length and cell viability were markedly
changed when expression of double-stranded telomere binding
factor Ng-TRF1 was altered in transgenic tobacco BY-2 suspension culture cells, providing the first evidence for in vivo role
of telomere binding protein in plants at the cellular level (Yang
et al., 2004). However, functions of telomere binding proteins
have yet to be determined at the whole-plant level. Rice is regarded as a model monocot cereal because of its economical
value, small genome size, and ease of transformation (Jeon et al.,
2000). Recent genomic analysis revealed that rice telomeric regions
are significantly heterogeneous in both sequence and structure,
suggesting the existence of dynamic mechanisms of controlling
telomere length (Mizuno et al., 2006). In addition, most telomere
studies in plants have been done with the dicot Arabidopsis. These
factors prompted us to investigate the possible function of RTBP1
in rice.
In this study, we examined the effects of inactivating, silencing,
or overexpressing RTBP1 on rice telomere, morphology, and development. We detected typical telomeric smears consisting of
fragments 5 to 10 kb long in wild-type rice plants (Figure 2A), which
agrees with the results determined by both fiber-FISH and terminal restriction fragment assay (Mizuno et al., 2006). However,
we found that the rtbp1 mutant lines had markedly longer telomeres (10 to 30 kb), whereas the 35S:anti-RTBP1 constructs
contained 8- to 25-kb telomeric tracts. Thus, the activity of the
RTBP1 gene seems to be inversely related to telomere length,
suggesting that RTBP1 is involved in a negative control of telomere length homeostasis in rice plants. This raises the tantalizing
possibility that RTBP1 inhibits the action of telomerase at the extreme ends of telomeres in vivo, as is the case for human TRF1/
PIN2 (van Steensel and de Lange, 1997; Wotton and Shore,
1997; Smogorzewska et al., 2000). It is worth noting that markedly lengthened telomeric DNA tracts were also found in heterozygous G1 progeny (Figure 2A). This suggests that control of
1778
The Plant Cell
Figure 5. Cytogenetic Analysis of Wild-Type, rtbp1 Mutant, and 35S:anti-RTBP1 T2 Transgenic PMCs.
(A) and (B) DAPI-stained metaphase and anaphase chromosomes of wild-type PMCs. Bars ¼ 10 mM.
(C) and (D) A high incidence of anaphase chromosome fusions was detected in G3 (17.2%) and G4 (26.7%) rtbp1 mutants. Examples of anaphase
chromosomes with two bridges (C) and one bridge (D) are presented. Bars ¼ 10 mM.
(E) and (F) Abnormal chromosomal fusions were also detected during metaphase in rtbp1 mutant PMCs (E) and anaphase in 35S:anti-RTBP1 T2
transgenic PMCs (F). Bars ¼ 10 mM.
(G) and (H) FISH analysis of anaphase chromosomes in the wild-type (left panel) and G4 rtbp1 (right panel) PMCs with the telomeric repeat (TTTAGGG)70
(green) and rice centromere-specific CentO satellite DNA (red) fluorescent probes. Telomeric repeat signal detected in the anaphase bridge is indicated
by an arrow. No CentO signal was detected in the anaphase bridge. Bars ¼ 5 mM.
telomere length was sensitive to cellular amount of RTBP1. In
addition, although the severity of the phenotype of rtbp1 increased gradually during each of four successive generations,
long telomeres in G1 plants were maintained at almost constant
size in subsequent G2 to G4 generations of rtbp1 mutant (Figure
2B). Thus, an as yet unidentified RTBP1-independent factor(s)
may also be in place that controls telomere homeostasis in rice
plants. Isolation and characterization of telomere-related mutants
other than rtbp1 will be necessary to examine this possibility.
Another striking phenotype of rtbp1 is severely disturbed development of vegetative and reproductive organs. Closer inspection revealed that both germination and postgermination growth
of shoots and roots as well as generation of floral tissues, including stamens, spikelets, lemma, palea, and anthers, of homozygous rtbp1 lines were gradually and severely impaired over
four successive generations (Figures 3 and 4). Furthermore, cytogenetic analysis showed that up to 26.7% of anaphases of the
rtbp1 PMCs suffered one or more chromosomal fusions, which
might cause a significant degree of genome instability (Figure 5,
Table 1). With the aid of FISH analysis, we could detect the
telomeric repeat sequence in the abnormal anaphase bridges of
mutant PMCs (Figure 5H). Accumulating evidence in mammalian
systems indicates that bridged anaphase chromosomes and
unequal segregation of chromosomes are typical results from
Rice Telomere Binding Protein RTBP1
the fusion of unprotected chromosome termini due to structural
defects of telomeres (Hande et al., 1999; Artandi et al., 2000).
Arabidopsis mutants in which telomere-related genes are disrupted suffer genomic instability due to anaphase chromosomal
fusions, as do the later generation progeny of telomerase-deficient
mutants (Riha et al., 2001; Riha and Shippen, 2003b; Siroky et al.,
2003; Puizina et al., 2004; Shakirov et al., 2005; Vespa et al.,
2005; Vannier et al., 2006). In this regard, it should be noted that
increased frequency of anaphase bridges in rtbp1 PMCs occurs
in parallel with progressively increased aberrant morphology of
the mutant rice plants (Table 1). In addition, although telomere
length is reset to a new, longer size after just one generation of
RTBP1 loss and the length of telomeres does not change in subsequent generations (Figure 2), the rate of chromosome fusion
continues to increase. Thus, abnormal phenotypes of rtbp1 are
not due to the increased telomere length per se but because of
genome instability caused by uncapped telomeres. Based on
these results, we propose that RTBP1 may play dual functions as
a negative regulator of telomere length and as one of the required
functional components for the proper architecture of telomeres
in rice plants.
Because our extensive screening of the T-DNA–transformed
20,500 rice mutant collection yielded only one mutant line for
RTBP1, we established transgenic 35S:anti-RTBP1 rice plants for
comparison (Figure 1). It is interesting to note that the 35S:antiRTBP1 lines exhibited a phenotype intermediate between the
wild-type and rtbp1 knockout plants in all of the criteria examined, including telomere length (Figure 2), both vegetative and
reproductive growth (Figures 3 and 4), and degree of genomic
anomaly (Figure 5, Table 1). Because the 35S:anti-RTBP1 progeny contained lower but detectable levels of RTBP1 mRNA
(Figure 1), these results, along with those obtained from rtbp1,
provide evidence that the amount of functional RTBP1 is critically
associated with the program of normal life cycle of rice plants.
Human telomeres are protected by sheltrin, a complex consisting of six telomere-associated proteins (de Lange, 2005). Sheltrin
has a DNA remodeling activity and provides for the dynamic
structural flexibility of the chromosome ends, which is necessary
for the protection of telomeres. Electron microscopy examination of mammalian telomeric DNA revealed large duplex loops
with tails, which were termed t-loops (Griffith et al., 1999). This structure was proposed to be a solution to the problem of telomere
protection. Subsequent studies showed that human TRF2 is involved in the formation of t-loops (Zhu et al., 2000; Stansel et al.,
2001). Telomeric DNA of higher plants is also arranged into t-loops.
Recent electron microscopy studies showed frequent t-loops
with highly varying sizes, ranging from 2 to 75 kb, in Pisum sativum
telomeres and suggest that t-loops will be found in plant and animal species that contain detectable telomeric repeat sequences
(Cesare et al., 2003). In the case of the P. sativum t-loops, the
largest loops detected were ;85 kb. It seems likely that there is
some active mechanism to generate such large structures. Thus,
it would be of interest to investigate whether RTBP1 participates
in t-loop formation of rice telomeres.
It was reported that, although telomere tracts are fairly uniformly distributed in 14 wild accessions of Arabidopsis, the
Wassilewskija ecotype exhibited a bimodal size distribution, with
different individuals harboring telomeres with 2 to 5 kb or 4 to 9
1779
kb, indicating that telomere length is not modulated by a single
genetic factor (Shakirov and Shippen, 2004). In addition, analysis
of the subtelomeric region of 35 wild-type accessions revealed
higher structural variation relative to centromere region, which
implies the variety of genomic events that drive the fluidity of chromosome termini (Kuo et al., 2006). Maillet et al. (2006) analyzed
telomere lengths in 16 different interecotype crosses between
Arabidopsis plants with different telomere sizes and observed
that the interecotype hybrid plants presented a new telomerelength set point, intermediate between that of the two parent plants.
These recent studies support the notion that plant telomeres are
structurally and functionally dynamic. Therefore, further functional
characterizations of RTBP1 and identification of the proteins it
interacts with will be critical to a more complete understanding of
such dynamic structures and functions of telomeres in rice, a
model monocot cereal plant.
METHODS
Plant Growth
Dry rice (Oryza sativa var japonica cv Dongjin) seeds were soaked once
with 70% ethanol and then rinsed extensively with sterilized water. Seeds
of the wild type, rtbp1 mutants, 35S:RTBP1, and 35S:anti-RTBP1 were
germinated on Murashige and Skoog medium containing MS basal salt
(Wako Pure Chemical), 3% sucrose, 0.2% phytogel, and 0.55 mM myoinositol. The seedlings were grown for 1 week at 258C under continuous
light and then transplanted to soil in the greenhouse and raised to maturity.
Total RNA Isolation and RT-PCR
Total RNAs of wild-type, rtbp1, 35S:RTBP1, and 35S:anti-RTBP1 rice
leaves were obtained by a method described previously (Hong et al.,
2005). The first-strand cDNA, synthesized from 10 mg of total RNA, was
PCR amplified with RTBP1-specific primers using high-fidelity Ex-Tag
polymerase (Takara) as described (Yang et al., 2002).
Genomic DNA Isolation and Pulse-Field Gel Electrophoresis
Total genomic DNA was isolated from mature rice leaves as indicated
previously (Hong et al., 2005) and digested with TaqI restriction enzyme.
The DNA fragments (10 mg per lane) were separated by pulse-field gel
electrophoresis using a CHEF-DRIII system (Bio-Rad) according to Yang
et al. (2004). The gel was blotted onto a nylon membrane filter (Bio-Rad)
and then hybridized to a 32P-labeled (TTTAGGG)70 fragment under high
stringency conditions (Yang et al., 2004).
PCR Screening for the rtbp1 Mutant Line
The primer M1 (59-ATGGTGTTGCAGAAGAGGTTGGAC-39) was used in
combination with a T-DNA left border primer LBa-1 (59-TCCGAAACTATCAGTGTCTAGCTAGA-39) in a PCR screen on pooled rice DNA (Jeon
et al., 2000) to identify rice lines containing a T-DNA copy integrated into
the RTBP1 gene. Using primers M1 and LBa-1, a 2.2-kb PCR fragment
was specifically amplified from the largest DNA pool and detected in consecutive smaller DNA pools until a single positive rice plant was identified.
Plants homozygous for the T-DNA insertion were identified by multiplex PCR with primers M2 (59-AAAACCCTTAGATGACCACATACA-39),
M3 (59-AGTTAACCATCAAATCTTTCAACA-39), and RBa-1 (59-GGGTTGGGGTTTCTACAGGACGTAAC-39) (Figure 2B).
1780
The Plant Cell
Genomic DNA Gel Blot Analysis
Rice leaf genomic DNA (10 mg per lane) was digested with appropriate
enzymes, separated by electrophoresis on a 0.7% agarose gel, and blotted
onto a nylon membrane filter (Amersham). The filter was hybridized with
32P-labeled GUS cDNA probe under high stringency conditions as described by Yang et al. (2003). The GUS probe was prepared from 1.8-kb
BamH1-EcoRI fragment (Jeon et al., 2000). The blot was visualized by autoradiography at 808C using Kodak XAR-5 film and an intensifying screen.
Rural Development Administration of the Republic of Korea) to W.T.K.
and from the Crop Functional Genomics Program (CG1111) to G.A.
M.Y.B. was a recipient of a BK21 graduate student scholarship.
Received March 30, 2007; revised May 19, 2007; accepted June 1, 2007;
published June 22, 2007.
REFERENCES
Generation of the 35S:RTBP1 and 35S:anti-RTBP1 Constructs and
Transformation of Rice
The full-length pRTBP1 cDNA was inserted into the corresponding sites
of the binary vector pCambia in the sense or antisense orientation. The
fusion gene construct was transferred to Agrobacterium tumefaciens strain
AGL1 by electroporation as described by Cho et al. (2006). Rice transformation was accomplished by Agrobacterium-mediated cocultivation
methods as previously described (Jeon et al., 2000). All transgenic rice
plants were generated on a medium containing 40 mg L1 hygromycin B.
The regenerated plants were grown in a greenhouse of typically 308C during
the day and 208C at night. The light/dark cycle in the greenhouse was 14/10 h.
Cytogenetic Studies
Anaphase and metaphase spreads were obtained from PMCs of the wild
type and the rtbp1 mutant and stained with DAPI as described (MartinezZapater and Salinas, 1998) with some modifications. Briefly, anthers were
fixed in 1:3 (v/v) acetic acid:ethanol for 2 h and soaked in distilled water,
followed by digestion with 2% cellulose, 1.5% macerozyme, 0.3% pectolyase, and 1 mM EDTA, pH 4.2, in distilled water for 50 to 60 min. After
being squashed on slide glass, samples were stained with DAPI and observed by fluorescence microscopy (Axio Imager; Zeiss).
FISH
The telomeric probe (TTTAGGG)70 and rice centromere-specific 165-bp
CentO satellite DNA were labeled with biotin-16-dUTP or digoxigenin-11dUTP using a nick translation system (Invitrogen). CentO DNA was obtained by PCR using genomic DNA as a template (forward primer,
59-CACGTGGGTGCGATGTTTTTG-39; reverse primer, 59-TGCCAATATTTGCATTAATTG-39). FISH analysis was performed with slides obtained as
described above for DAPI staining according to Koo et al. (2004).
Chromosomal DNA on the slides was denatured with 70% formamide
at 708C for 2.5 min, followed by dehydration in a 70, 85, 95, and 100%
ethanol series at 208C for 3 min each. The probe mixture was applied to
the denatured chromosomal DNA and covered with a glass cover slip.
Slides were placed in a humid chamber at 378C for 18 h. Probes were
detected with avidin-FITC or anti-digoxigenin-Cy3 (Roche). The chromosomes were counterstained with DAPI.
Accession Numbers
The GenBank accession numbers for the Arabidopsis TERT, tobacco NgTRF1, and RTBP1 are NM121691, AF543195, and AF242298, respectively.
ACKNOWLEDGMENTS
We thank June M. Kwak (University of Maryland, College Park) and the
members of W.T.K.’s laboratory for their help and discussion. This work
was supported by grants from the Plant Diversity Research Center (21st
Century Frontier Research Program funded by the Ministry of Science
and Technology of Korea) and the BioGreen 21 Program (funded by the
Artandi, S.E., Chang, S., Lee, S.L., Alson, S., Gottlieb, G.J., Chin, L., and
Depinho, R.A. (2000). Telomere dysfunction promotes non-reciprocal
translocations and epithelial cancers in mice. Nature 406: 641–645.
Bilaud, T., Brun, C., Ancelin, K., Koering, C.E., Laroche, T., and
Gilson, E. (1997). Telomeric localization of TRF2, a novel human
telobox protein. Nat. Genet. 17: 236–239.
Blackburn, E.H. (1991). Structure and function of telomeres. Nature
350: 569–573.
Blackburn, E.H. (2001). Switching and signaling at the telomere. Cell
106: 661–673.
Bundock, P., and Hooykaas, P. (2002). Severe developmental defects,
hypersensitivity to DNA-damaging agents, and lengthened telomeres
in Arabidopsis MRE11 mutants. Plant Cell 14: 2451–2462.
Bundock, P., van Attikum, H., and Hooykaas, P. (2002). Increased
telomere length and hypersensitivity to DNA damaging agents in an
Arabidopsis KU70 mutant. Nucleic Acids Res. 30: 3395–3400.
Cesare, A.J., Quinney, N., Willcox, S., Subramanian, D., and Griffith,
J.D. (2003). Telomere looping in P. sativum (common garden pea).
Plant J. 36: 271–279.
Chen, C.M., Wang, C.T., and Ho, C.H. (2001). A plant gene encoding a
myb-like protein that binds telomeric GGTTTAG repeats in vitro. J.
Biol. Chem. 276: 16511–16519.
Cho, S.K., Chung, H.S., Ryu, M.Y., Park, M.J., Lee, M.M., Bahk, Y.-Y.,
Kim, J., Pai, H.S., and Kim, W.T. (2006). Heterologous expression and
molecular and cellular characterization of CaPUB1 encoding a hot pepper U-box E3 ubiquitin ligase homolog. Plant Physiol. 142: 1664–1682.
Chong, L., van Steensel, B., Broccoli, D., Erdjument-Bromage, H.,
Hanish, J., Tempst, P., and de Lange, T. (1995). A human telomeric
protein. Science 270: 1663–1667.
Collins, K. (2000). Mammalian telomeres and telomerase. Curr. Opin.
Cell Biol. 12: 378–383.
Cooper, J.P., Nimmo, E.R., Allshire, R.C., and Cech, T.R. (1997).
Regulation of telomere length and function by a Myb-domain protein
in fission yeast. Nature 385: 744–747.
de Lange, T. (2005). Shelterin: The protein complex that shapes and
safeguards human telomeres. Genes Dev. 19: 2100–2110.
Gallego, M.E., Jalut, N., and White, C.I. (2003). Telomerase dependence of telomere lengthening in Ku80 mutant Arabidopsis. Plant Cell
15: 782–789.
Gallego, M.E., and White, C.I. (2001). RAD50 function is essential for
telomere maintenance in Arabidopsis. Proc. Natl. Acad. Sci. USA 98:
1711–1716.
Gallego, M.E., and White, C.I. (2005). DNA repair and recombination
functions in Arabidopsis telomere maintenance. Chromosome Res.
13: 481–491.
Greider, C.W. (1996). Telomere length regulation. Annu. Rev. Biochem.
65: 337–365.
Griffith, J.D., Comeau, L., Rosenfield, S., Stansel, R.M., Bianchi, A.,
Moss, H., and de Lange, T. (1999). Mammalian telomeres end in a
large duplex loop. Cell 97: 503–514.
Hande, M.P., Samper, E., Lansdorp, P., and Blasco, M.A. (1999).
Telomere length dynamics and chromosomal instability in cells derived from telomerase null mice. J. Cell Biol. 144: 589–601.
Rice Telomere Binding Protein RTBP1
Hong, J.-P., Kim, S.M., Ryu, M.Y., Choe, S., Park, P.B., An, G., and
Kim, W.T. (2005). Structure and expression of OsMRE11 in rice. J.
Plant Biol. 48: 229–236.
Hwang, M.G., and Cho, M.H. (2007). Arabidopsis thaliana telomeric
DNA-binding protein 1 is requited for telomere length homeostasis
and its Myb-extension domain stabilizes plant telomeric DNA binding.
Nucleic Acids Res. 35: 1333–1342.
Hwang, M.G., Chung, I.K., Kang, B.G., and Cho, M.H. (2001). Sequencespecific binding property of Arabidopsis thaliana telomeric DNA binding protein 1 (AtTBP1). FEBS Lett. 503: 35–40.
Jeon, J.S., et al. (2000). T-DNA insertional mutagenesis for functional
genomics in rice. Plant J. 22: 561–570.
Karamysheva, Z.N., Surovtseva, Y.V., Vespa, L., Shakirov, E.V., and
Shippen, D.E. (2004). A C-terminal Myb-extension domain defines a
novel family of double-strand telomeric DNA binding proteins in
Arabidopsis. J. Biol. Chem. 279: 47799–47807.
Koo, D.-H., Plaha, P., Lim, Y.P., Hur, Y., and Bang, J.-W. (2004). A
high-resolution karyotype of Brassica rapa ssp. pekinensis revealed
by pachytene analysis and multicolor fluorescence in situ hybridization. Theor. Appl. Genet. 109: 1346–1352.
Kuchar, M., and Fajkus, J. (2004). Interactions of putative telomerebinding proteins in Arabidopsis thaliana: Identification of functional
TRF2 homolog in plants. FEBS Lett. 578: 311–315.
Kuo, H.-F., Olsen, K.M., and Richards, E.J. (2006). Natural variation in
a subtelomeric region of Arabidopsis: Implications for the genomic
dynamics of a chromosome end. Genetics 173: 401–417.
Lamb, J.C., Yu, W., Han, F., and Birchler, J.A. (2007). Plant chromosomes from end to end: Telomeres, heterochromatin and centromeres. Curr. Opin. Plant Biol. 10: 116–122.
Maillet, G., White, C.I., and Gallego, M.E. (2006). Telomere-length
regulation in inter-ecotype crosses of Arabidopsis. Plant Mol. Biol. 62:
859–866.
Martinez-Zapater, J.M., and Salinas, J. (1998). Arabidopsis Protocol.
(Totowa, NJ: Human Press).
McKnight, T.D., and Shippen, D.E. (2004). Plant telomere biology.
Plant Cell 16: 794–803.
Mizuno, H., Wu, J., Kanamori, H., Fujisawa, M., Namiki, N., Saji, S.,
Katagiri, S., Katayose, Y., Sasaki, T., and Matsumoto, T. (2006).
Sequencing and characterization of telomere and subtelomere regions on
rice chromosomes 1S, 2S, 2L, 6L, 7S, 7L and 8S. Plant J. 46: 206–217.
Pich, U., Fuchs, J., and Schubert, I. (1996). How do Alliaceae stabilize
their chromosome ends in the absence of TTTAGGG sequences?
Chromosome Res. 4: 207–213.
Puizina, J., Siroky, J., Mokros, P., Schweizer, D., and Riha, K. (2004).
Mre11 deficiency in Arabidopsis is associated with chromosomal instability in somatic cells and Spo11-dependent genome fragmentation
during meiosis. Plant Cell 16: 1968–1978.
Riha, K., McKnight, T.D., Griffing, L.R., and Shippen, D.E. (2001).
Living with genome instability: Plant response to telomere dysfunction. Science 291: 1797–1800.
Riha, K., and Shippen, D.E. (2003a). Telomere structure, function and
maintenance in Arabidopsis. Chromosome Res. 11: 263–275.
Riha, K., and Shippen, D.E. (2003b). Ku is required for telomeric C-rich
strand maintenance but not for end-to-end chromosome fusions in
Arabidopsis. Proc. Natl. Acad. Sci. USA 100: 611–615.
Schrumpfova, P., Kuchar, M., Mikova, G., Skrisovska, L., Kubicarova,
T., and Fajkus, J. (2004). Characterization of two Arabidopsis thaliana
myb-like proteins showing affinity to telomeric DNA sequence. Genome
47: 316–324.
Shakirov, E., and Shippen, D.E. (2004). Length regulation and dynamics of
individual telomere tracts in wild-type Arabidopsis. Plant Cell 16: 1959–1967.
1781
Shakirov, E.V., Surovtseva, Y.V., Osbun, N., and Shippen, D.E.
(2005). The Arabidopsis Pot1 and Pot2 proteins function in telomere
length homeostasis and chromosome end protection. Mol. Cell. Biol.
25: 7725–7733.
Shore, D. (1994). Rap1: A protean regulator in yeast. Trends Genet. 10:
408–412.
Shore, D. (2001). Telomeric chromatin: Replicating and wrapping up
chromosome ends. Curr. Opin. Genet. Dev. 11: 189–198.
Siroky, J., Zluvova, J., Riha, K., Shippen, D.E., and Vyskot, B. (2003).
Rearrangement of ribosomal DNA clusters in lare generation telomerasedeficient Arabidopsis. Chromosoma 112: 116–123.
Smogorzewska, A., van Steensel, B., Bianchi, A., Oelmann, S.,
Schaefer, M.R., Schnapp, G., and de Lange, T. (2000). Control of
human telomere length by TRF1 and TRF2. Mol. Cell. Biol. 20: 1659–
1668.
Stansel, R.M., de Lange, T., and Griffith, J.D. (2001). T-loop assembly
in vitro involves binding of TRF2 near the 39 telomeric overhang.
EMBO J. 20: 5532–5540.
Vannier, J.-B., Depeiges, A., White, C., and Gallego, M.E. (2006).
Two roles for Rad50 in telomere maintenance. EMBO J. 25: 4577–
4585.
van Steensel, B., and de Lange, T. (1997). Control of telomere length
by the human telomeric protein TRF1. Nature 385: 740–743.
Vespa, L., Couvillion, M., Spangler, E., and Shippen, D.E. (2005). ATM
and ATR make distinct contributions to chromosome end protection
and the maintenance of telomeric DNA in Arabidopsis. Genes Dev. 19:
2111–2115.
Weiss-Schneeweiss, H., Riha, K., Jang, C.G., Puizina, J., Scherthan,
H., and Schweizer, D. (2004). Chromosome termini of the monocot
plant Othocallis siberica are maintained by telomerase, which specifically synthesises vertebrate-type telomere sequences. Plant J. 37:
484–493.
Wotton, D., and Shore, D. (1997). A novel Rap1p-interacting factor,
Rif2p, cooperates with Rif1p to regulate telomere length in Saccharomyces cerevisiae. Genes Dev. 11: 748–760.
Yang, S.W., Jin, E.S., Chung, I.K., and Kim, W.T. (2002). Cell cycledependent regulation of telomerase activity by auxin, ABA and protein
phosphorylation in tobacco BY-2 suspension culture cells. Plant J. 29:
617–626.
Yang, S.W., Kim, D.H., Lee, J.J., Chun, Y.J., Lee, J.-H., Kim, Y.J.,
Chung, I.K., and Kim, W.T. (2003). Expression of the telomeric repeat
binding factor gene NgTRF1 is closely coordinated with the cell division program in tobacco BY-2 suspension culture cells. J. Biol. Chem.
278: 21395–21407.
Yang, S.W., Kim, S.K., and Kim, W.T. (2004). Perturbation of NgTRF1
expression induces apoptosis-like cell death in tobacco BY-2 cells
and implicates NgTRF1 in the control of telomere length and stability.
Plant Cell 16: 3370–3385.
Yu, E.Y., Kim, S.E., Kim, J.H., Ko, J.H., Cho, M.H., and Chung, I.K.
(2000). Sequence-specific DNA recognition by the Myb-like domain of plant telomeric protein RTBP1. J. Biol. Chem. 275: 24208–
24214.
Zellinger, B., and Riha, K. (2007). Composition of plant telomeres.
Biochim. Biophys. Acta 1769: 399–409.
Zhang, W., Yi, C., Bao, W., Liu, B., Cui, J., Yu, H., Cao, X., Gu, M., Liu,
M., and Cheng, Z. (2005). The transcribed 165-bp CentO satellite is
the major functional centromeric element in the wild rice species
Oryza punctata. Plant Physiol. 139: 306–315.
Zhu, X.-D., Kuster, B., Mann, M., Petrini, J.H.J., and de Lange, T.
(2000). Cell-cycle-regulated association of RAD50/MRE11/NBS1 with
TRF2 and human telomeres. Nat. Genet. 25: 347–352.
Suppression of RICE TELOMERE BINDING PROTEIN1 Results in Severe and Gradual
Developmental Defects Accompanied by Genome Instability in Rice
Jong-Pil Hong, Mi Young Byun, Dal-Hoe Koo, Kyungsook An, Jae-Wook Bang, In Kwon Chung,
Gynheung An and Woo Taek Kim
Plant Cell 2007;19;1770-1781; originally published online June 22, 2007;
DOI 10.1105/tpc.107.051953
This information is current as of June 14, 2017
References
This article cites 56 articles, 26 of which can be accessed free at:
/content/19/6/1770.full.html#ref-list-1
Permissions
https://www.copyright.com/ccc/openurl.do?sid=pd_hw1532298X&issn=1532298X&WT.mc_id=pd_hw1532298X
eTOCs
Sign up for eTOCs at:
http://www.plantcell.org/cgi/alerts/ctmain
CiteTrack Alerts
Sign up for CiteTrack Alerts at:
http://www.plantcell.org/cgi/alerts/ctmain
Subscription Information
Subscription Information for The Plant Cell and Plant Physiology is available at:
http://www.aspb.org/publications/subscriptions.cfm
© American Society of Plant Biologists
ADVANCING THE SCIENCE OF PLANT BIOLOGY