ArticlePDF Available

Direct measurement of the transfer rate of chloroplast DNA into the nucleus

Authors:

Abstract and Figures

Gene transfer from the chloroplast to the nucleus has occurred over evolutionary time. Functional gene establishment in the nucleus is rare, but DNA transfer without functionality is presumably more frequent. Here, we measured directly the transfer rate of chloroplast DNA (cpDNA) into the nucleus of tobacco plants (Nicotiana tabacum). To visualize this process, a nucleus-specific neomycin phosphotransferase gene (neoSTLS2) was integrated into the chloroplast genome, and the transfer of cpDNA to the nucleus was detected by screening for kanamycin-resistant seedlings in progeny. A screen for kanamycin-resistant seedlings was conducted with about 250,000 progeny produced by fertilization of wild-type females with pollen from plants containing cp-neoSTLS2. Sixteen plants of independent origin were identified and their progenies showed stable inheritance of neoSTLS2, characteristic of nuclear genes. Thus, we provide a quantitative estimate of one transposition event in about 16,000 pollen grains for the frequency of transfer of cpDNA to the nucleus. In addition to its evident role in organellar evolution, transposition of cpDNA to the nucleus in tobacco occurs at a rate that must have significant consequences for existing nuclear genes.
Content may be subject to copyright.
17. Sebastiani, F., Barberio, C., Casalone, E., Cavalieri, D. & Polsinelli, M. Crosses between Saccharomyces
cerevisiae and Saccharomyces bayanus generate fertile hybrids. Res. Microbiol. 153, 53–58 (2002).
18. Tamai, Y., Momma, T., Yoshimoto,H. & Kaneko, Y. Co-existence of two types of chromosome in the
bottom-fermenting yeast, Saccharomyces pastorianus. Yea st 14, 923–937 (1998).
19. Kielland-Brandt, M. C., Nilson-Tillgren, T., Peterson, J. G. L., Holemberg, S. & Gjermansen, C. in
Yeast Genetics (eds Spencer, J. F. T., Spencer, D. M. & Smith, A. R. W.) 421–438 (Springer, New York,
1983).
20. Wolfe, K. H. & Shields, D. C. Molecularevi dencefor an ancient duplication of the entire yeast genome.
Nature 387, 708–713 (1997).
21. Llorente, B. et al. Genomic exploration of the hemiascomycetous yeasts: 20. Evolution of gene
redundancy compared to Saccharomyces cerevisiae.FEBS Lett. 487, 122–133 (2000).
22. Greig, D., Borts, R. H., Louis, E. J. & Travisano, M. Epistasis and hybrid sterility in Saccharomyces.
Proc. R. Soc. Lond. B. 269, 1167–1671 (2002).
23. Hunter, N., Chambers, S. R., Louis, E. J. & Borts, R. H. The mismatch repair system contributes to
meiotic sterility in an interspecific yeast. EMBO J. 15, 1726–1733 (1996).
24. Adams, A., Gottschling, D. E., Kaiser, C. A. & Stearns, T. in Methods in Yeast Genetics (ed. Dickerson,
M. M.) 145–175 (Cold Spring Harbor Laboratory Press, New York, 1997).
25. Gietz, R. D., Schiestl, R. H., Willems, A. R. & Woods, R. A. Studies on the transformation of intact
yeast cells by the LiAc/SS-DNA/PEG procedure. Yea st 11, 355–360 (1995).
26. Gardner, D. C. J., Heale, S. M., Stateva, L. I. & Oliver, S. G. Treatment of yeast cells with wall lytic
enzymes is not required to prepare chromosomes for pulsed-field gel analysis. Ye as t 9, 1053–1055
(1993).
Supplementary Information accompanies the paper on Nature’s website
(çhttp://www.nature.com/nature).
Acknowledgements This work was supported by grants from the Biotechnology and Biological
Sciences Research Council (to E.J.L., I.N.R. and S.G.O.) and the Wellcome Trust (to S.G.O.).
We thank S. James and L. Lockhart for their help in some early analyses, and B. Dujon for
discussions.
Competing interests statement The authors declare that they have no competing financial
interests.
Correspondence and requests for materials should be addressed to S.G.O.
(e-mail: steve.oliver@man.ac.uk).
..............................................................
Direct measurement of the
transfer rate of chloroplast
DNA into the nucleus
Chun Y. Huang*, Michael A. Ayliffe& Jeremy N. Timmis*
*Department of Molecular Biosciences, The University of Adelaide,
South Australia, 5005, Australia
CSIRO Plant Industry, GPO Box 1600, Australian Capital Territory 2601,
Australia
.............................................................................................................................................................................
Gene transfer from the chloroplast to the nucleus has occurred
over evolutionary time
1
. Functional gene establishment in the
nucleus is rare, but DNA transfer without functionality is pre-
sumably more frequent. Here, we measured directly the transfer
rate of chloroplast DNA (cpDNA) into the nucleus of tobacco
plants (Nicotiana tabacum). To visualize this process, a nucleus-
specific neomycin phosphotransferase gene (neoSTLS2) was inte-
grated into the chloroplast genome, and the transfer of cpDNA to
the nucleus was detected by screening for kanamycin-resistant
seedlings in progeny. A screen for kanamycin-resistant seedlings
was conducted with about 250,000 progeny produced by fertili-
zation of wild-type females with pollen from plants containing
cp-neoSTLS2. Sixteen plants of independent origin were identi-
fied and their progenies showed stable inheritance of neoSTLS2,
characteristic of nuclear genes. Thus, we provide a quantitative
estimate of one transposition event in about 16,000 pollen grains
for the frequency of transfer of cpDNA to the nucleus. In addition
to its evident role in organellar evolution, transposition of
cpDNA to the nucleus in tobacco occurs at a rate that must
have significant consequences for existing nuclear genes.
The genomes of cytoplasmic organelles have been depleted
during endosymbiotic evolution, as many genes were either lost
or transferred to the nucleus. The plastome of higher plants has
been reduced to approximately 130 genes from an ancestral cyano-
bacterial genome that probably contained over 3,000 genes
2
.It
has been estimated that 1,700 protein-coding nuclear genes of
Arabidopsis thaliana were acquired from cyanobacteria
1
. Even
greater genetic erosion has occurred during mitochondrial genome
evolution
3
. The rps10 gene, located in the mitochondrial genome in
most angiosperms, has been transferred independently to the
nucleus hundreds of times during endosymbiotic evolution
4
. Simi-
lar multiple independent gene transfers from the chloroplast to the
nuclear genome have also been reported
5
.
Long tracts of extant organelle DNA are also found in the
eukaryotic chromosomes of plants. For example, about 620 kilo-
bases (kb) of mitochondrial DNA (mtDNA) are found on chromo-
some 2 of Arabidopsis
6
, and there are about 33 kb of cpDNA on
chromosome 10L of rice
7
. Indeed, eukaryotes generally contain
DNA sequences in their nuclear genome that show close similarity
Figure 1 Introduction of aadA and neoSTLS2 genes into the tobacco plastid genome by
homologous recombination. a, Construction and integration of pPRV111A::neoSTLS2.
Restriction sites are: A, ApaI; B, BamHI; E, EcoRI; H, HindIII. b, Southern analysis of two
transplastomic lines. Total DNA (5
m
g) was digested with BamHI and hybridized with the
probes indicated (see a). c, Northern analysis of transplastomic lines and their progeny.
Total leaf RNA (3
m
g) was hybridized with the neo probe (see a). 18S RNA stained with
ethidium bromide provides a loading control. Molecular sizes (in kilobases (kb)) are
indicated on the left. n þve, nuclear neoSTLS2 control line.
letters to nature
NATURE| VOL 422 | 6 MARCH 2003 | www.nature.com/nature72 © 2003 Nature Publishing Group
to organellar DNA
8–11
, suggesting that transfer of chloroplast and
mtDNA are ongoing, possibly frequent processes.
Transfer of DNA from yeast mitochondria to the nucleus, in the
form of a non-integrating recombinant plasmid, occurs at a rate of
one in 2.5 £10
4
cells per generation
12
, but nothing is known about
the frequency of transposition of cpDNA to the nucleus. To quantify
this process, we inserted a nuclear-specific neoSTLS2 gene (ref. 13)
together with a plastid-selectable aminoglycoside 3 0-adeyltransfer-
ase marker gene, (aadA)
14
, into the plastome of tobacco (Fig. 1a).
The aadA gene confers spectinomycin resistance, enabling selection
of transplastomic cells. The neoSTLS2 gene features a constitutive
plant viral promoter, CaMV-35S
15
, and a nuclear intron. It was
designed to be functional only when transposed to the nucleus
and to confer kanamycin resistance in young seedlings. An intron,
from the potato nuclear ST-LS1 gene
16
, was integrated within the
reading frame to prevent synthesis of a functional protein in the
chloroplast
13
.
Two independent transplastomic (tp) lines (tp7 and tp17),
containing both neoSTLS2 and aadA, were obtained by particle
bombardment of tobacco leaves with pPRV111A::neoSTLS2
(Fig. 1a). DNA blot analysis confirmed the homoplasmic presence
of the transgenes in tp7 and tp17 lines (Fig. 1b). Homoplasmy is not
essential for our transposition experiments, but it may be critical for
genetic stability of the transplastomic lines in the absence of
spectinomycin selection.
Crosses were made to determine the inheritance and function-
ality of the selectable markers in the two transplastomic lines.
Seeds set from tp7 or tp17 as the female parent were able to
germinate on medium containing 500
m
gml
21
spectinomycin,
and develop uniformly into green, healthy seedlings similar to a
positive control, SSuH2 (ref. 17), containing a plastid aadA gene
(Fig. 2a). In contrast, seedlings of wild-type tobacco and seedlings
from wild-type females crossed with tp7 or tp17 as male parents,
bleached rapidly on spectinomycin medium (Fig. 2a and Table 1).
This reciprocal difference in the inheritance of aadA is characteristic
of maternal inheritance of plastid genes in tobacco.
Notably, when the seeds set from tp7 or tp17 female parents were
grown on medium containing 150
m
gml
21
kanamycin, all seedlings
showed partial resistance to the antibiotic. They were uniformly
paler green and smaller in size than positive control seedlings from a
nuclear neo tobacco line, NS23 (ref. 18; Fig. 2b and Table 1). Wild-
type seedlings and seedlings from wild-type females crossed with
tp7 or tp17 male parents were all fully sensitive to kanamycin
Figure 2 Seedling tests for resistance to spectinomycin and kanamycin. a,b, Seedlings
from six progenies grown on medium containing 500
m
gml
21
spectinomycin (a)or
150
m
gml
21
kanamycin (b) with positive (SSuH2 or NS23) and wild-type controls. c, One
kanamycin-resistant plant among the progeny of self-pollinated tp7 plants on kanamycin
(150
m
gml
21
) and geneticin (20
m
gml
21
) medium. d, Segregation of kanamycin
resistance in progeny of wild type £tp-kr1 plants on 150
m
gml
21
kanamycin medium.
e, A kanamycin-resistant plant obtained in the progeny of wild type £tp7 cross.
f, Progenies of four kanamycin-resistant plants with controls on spectinomycin medium.
Table 1 Progenies analysed for resistance to spectinomycin and kanamycin
Cross
Spectinomycin
(500
m
gml
21
)
Kanamycin
(150
m
gml
21
)
kr:ks ratio
rsr s (P)
.............................................................................................................................................................................
Section I
tp7 self 168 0 0 827†
tp7 £WT 129 0 0 685†
WT £tp7 0 164 7 92,900
tp17 self 90 0 0 1,022†
tp17 £WT 127 0 0 532†
WT £tp17 0 90 11 165,000
Section II
WT £tp-kr1 0 142 166 149 1:1 (NS)
kr1 self ND ND 140 34 3:1 (NS)
Section IIIA
kr3 self 0 80 137 42 3:1 (NS)
kr6 self 0 38 107 35 3:1 (NS)
kr7 self 0 49 64 12 3:1 (NS)
kr8 self 0 82 165 49 3:1 (NS)
kr9 self 0 39 51 13 3:1 (NS)
kr10 self 0 35 62 17 3:1 (NS)
WT £kr12 0 79 129 120 1:1 (NS)
WT £kr13 0 100 133 133 1:1 (NS)
kr14 self 0 41 64 18 3:1 (NS)
kr16 self 0 48 126 60 3:1 (NS)
kr17 self 0 64 110 44 3:1 (NS)
kr18 self 0 40 52 15 3:1 (NS)
Section IIIB
kr4 self 0 76 136 330 3:1 (*)
WT £kr4 0 84 80 72 1:1 (NS)
kr5 self ND ND 39 144 3:1 (*)
WT £kr5 0 73 72 76 1:1 (NS)
kr11 self 0 59 12 105 3:1 (*)
kr19 self 0 54 36 59 3:1 (*)
Section IIIC
kr2 self 0 81 0 185 3:1 (*)
WT £kr2 ND ND 0 224 1:1 (*)
kr15 self 0 39 0 162 3:1 (*)
WT £kr15 ND ND 0 311 1:1 (*)
.............................................................................................................................................................................
Progenies are grouped in sections according to the pattern of inheritance. Section III is subdivided to
reflect genotypic and phenotypic differences. The parental origins of the 18 kanamycin-resistant
plants in section III are not specified, as tp7 and tp17 were indistinguishable in every respect (Figs
1c, 2a, b and 3c, d). ND, not determined; NS, not significant; kr, kanamycin resistant; ks, kanamycin
sensitive; r, resistant; s, sensitive; WT, wild type.
*P.0.01 (
x
2
).
†Partial resistance to 150
m
gml
21
kanamycin.
letters to nature
NATURE| VOL 422 | 6 MARCH 2003 | www.nature.com/nature 73
© 2003 Nature Publishing Group
(Fig. 2b). These results indicate that the partial kanamycin-resistant
phenotype is strictly maternally inherited, as expected for
cp-neoSTLS2 (refs 14, 19), and demonstrate that there is no
functional neoSTLS2 gene in the nucleus of tp7 or tp17 lines,
which could have been introduced concomitantly during chloro-
plast transformation.
Two neo-hybridizing transcripts (3.3 and 2.2 kb) were identified
in tp7 and tp17 RNA that were larger than correctly processed neo
transcripts (1.2 kb) in RNA of a nuclear neoSTLS2 control line,
derived from pCMneoSTLS2 (ref. 13; Fig. 1c). However, a low level
of the intron-spliced messenger RNA was detected in these two
transplastomic lines by the more sensitive polymerase chain reac-
tion with reverse transcription (RT–PCR) (data not shown). There-
fore cp-neoSTLS2 in the chloroplast gives rise to relatively abundant
transcripts that, although inefficiently processed, probably account
for the partial kanamycin resistance observed in the transplastomic
lines and their progenies (Fig. 2b). To suppress this partial resist-
ance, several analogues of kanamycin were tested. Supplementation
of kanamycin medium with a low concentration of geneticin was
found to be effective (Fig. 2c).
One kanamycin-resistant seedling was identified in an initial
screen of 9,300 seeds from the progeny of self-pollinated tp7 on
kanamycin and geneticin medium (Fig. 2c). This plant was desig-
nated tp-kr1 to reflect that it probably contained both chloroplast
and nuclear copies of neoSTLS2. Abundant, correctly processed neo
transcripts (1.2 kb) were identified in tp-kr1 RNA in addition to the
3.3- and 2.2-kb plastid transcripts (Fig. 1c). These 1.2-kb transcripts
remained in kanamycin-resistant progeny after wild-type females
were crossed with tp-kr1, whereas the 3.3- and 2.2-kb chloroplast
transcripts were absent owing to uniparental inheritance (data not
shown).
To gain further evidence that the transferred cpDNA had inte-
grated into the nucleus of tp-kr1, segregation analysis of kanamycin
and spectinomycin resistance was applied to the progeny of wild
type £tp-kr1 crosses. A 1:1 segregation ratio (kanamycin resistant/
kanamycin sensitive) was observed (Fig. 2d and Table 1), demon-
strating that at least one expressed copy of neoSTLS2 had integrated
into the nuclear genome of tp-kr1 at a single locus. Seedlings from
this cross were uniformly sensitive to spectinomycin (Table 1),
confirming non-transmission of the paternal transplastome. The
kanamycin-resistant progeny of the wild type £tp-kr1 cross were
designated kr1 to reflect the replacement of the transplastome with
the wild type plastome. Self-fertilization of kr1 yielded a 3:1
segregation ratio (kanamycin resistant/kanamycin sensitive) in the
progeny, demonstrating stable nuclear transmission of the trans-
posed neoSTLS2 gene (Table 1).A single fragment hybridized to neo
and aadA probes in Southern hybridizations of DraI-digested kr1
DNA, consistent with a single cpDNA insert in nuclear DNA (data
not shown). These data provide genetic and molecular evidence of a
nuclear integrant containing neoSTLS2 in tp-kr1. The unequivocal
verification of nuclear integration provided by sequence analysis of
a transplastomic/nuclear junction in kr1 is described later (see
Fig. 3e).
More kanamycin-resistant individuals arising from independent
transposition events were sought to quantify the frequency of
cpDNA transfer by large-scale screening of seeds derived from
Figure 3 Analysis of kanamycin-resistant plants and three transplastomic/nuclear
junctions. a,NcoI (N) and XbaI (X) transplastome fragments (double-headed arrows).
b,NcoI-digested DNA from 13 kanamycin-resistant plants hybridized with neo.cd,XbaI-
digested DNA from 17 kanamycin-resistant plants hybridized with neo (c)oraadA (d).
Lanes contained 5
m
g of DNA except for tp7 and tp17 (0.005
m
g). eg, Unique
transplastomic/nuclear junctions (circle enclosing a cross), PCR primers (arrows),
chloroplast sequences (filled boxes), unknown sequences (open boxes) and nuclear
sequences with database homology (light grey boxes). Templates for PCR from the
progeny of wild type £tp-kr1 (e), self-pollinated kr17 (f) or kr18 (g) are shown. 2ve, no
DNA; þve, iPCR product. Gels were stained with ethidium bromide. Fragment sizes are
indicated in kilobases. kr, kanamycin resistant; ks, kanamycin sensitive.
letters to nature
NATURE| VOL 422 | 6 MARCH 2003 | www.nature.com/nature74 © 2003 Nature Publishing Group
wild type £tp7 and wild type £tp17 backcrosses. Eighteen kana-
mycin-resistant plants (kr2–19) were identified among approxi-
mately 250,000 seedlings (Fig. 2e and Table 1). Backcross and self-
pollinated progenies of all 18 kanamycin-resistant plants were
uniformly sensitive to spectinomycin (Fig. 2f and Table 1), con-
firming the absence of pollen transmission of transplastomes from
their paternal transplastomic lines. Kanamycin resistance segregated
in the progenies of 16 kanamycin-resistant plants, 12 of which
showed mendelian inheritance: either 3:1 (kanamycin resistant/
kanamycin sensitive) after self-pollination or 1:1 after crossing to
wild type (Table 1). Therefore at least one expressed copy of
neoSTLS2 is integrated into the nuclear genome at a single locus
in these plants. The other four progenies of self-pollinated plants
(kr4, kr5, kr11 and kr19) had segregation ratios that deviated
significantly from 3:1. However, normal 1:1 ratios were restored
in wild type £kr4 and wild type £kr5, the two backcross progenies
tested (Table 1). Southern analysis identified multiple copies of
neoSTLS2 in these four plants, indicating that complex transgene
interactions
20
may be responsible for the abnormal ratios observed
in the progeny of self-pollinated plants (Fig. 3c).
Progenies of two kanamycin-resistant plants (kr2 and kr15)
derived from self-pollination or backcrossing to wild-type females,
showed uniform sensitivity to kanamycin (Table 1). Southern
analysis confirmed that the neoSTLS2 gene was present in both
parent plants (Fig. 3b) but absent from any offspring (data not
shown), indicating that the nuclear neoSTLS2 gene failed to trans-
mit. These data suggest that deletion may follow insertion at some
chromosomal sites.
Total DNA samples from kr1 and 12 kanamycin-resistant plants
from the large-scale screen were investigated to establish the
independence of the transposition events. NcoI digestion and
hybridization with the neo probe revealed two to four hybridizing
fragments in each sample (Fig. 3a, b). All 13 kanamycin-resistant
plants contained different sized fragments compared with the
control nuclear plant. As neoSTLS2 contains an NcoI site in its
coding region, two neo-hybridizing fragments (3.4 and 5.9 kb) are
expected in the transplastome (Fig. 3a). Ten out of the thirteen plant
DNAs contained these two fragments, as did a low loading of DNA
from tp17 (Fig. 3b). Therefore at least 9.3 kb of DNA, including
aadA, neoSTLS2 and adjacent cpDNA, was transferred from the
transplastome and integrated into their nuclear genomes. These
aadA/neoSTLS2 integrants are flanked by at least 6.5 kb of cpDNA,
which is larger than the plastid targeting sequence (3.0 kb) in the
pPRV111A::neoSTLS2 transformation vector (Fig. 1a). This rules
out the possibility that the neo-hybridizing fragments in most of the
kanamycin-resistant plants could originate from any source other
than the experimental transplastome.
To resolve further the independence of integration events, DNA
samples from all kanamycin-resistant plants, except for kr2 and
kr15, which did not transmit kanamycin resistance, were digested
with XbaI for Southern analysis. The neo probe hybridized to one or
more fragments, which differed in size and/or number, in each
sample (Fig. 3c). Hybridization of the same membrane with the
aadA probe similarly revealed fragments of different size and/or
number in each sample (Fig. 3d). The patterns of hybridizing
fragments for each DNA sample did not resemble that from any
other kanamycin-resistant plant or of the parental transplastomic
lines (Fig. 3c, d), indicating that each plant resulted from an
independent integration event. No hybridization was observed in
pooled DNA from 50 kanamycin-sensitive seedlings derived from
wild type £tp7 or wild type £tp17 crosses (Fig. 3c, d), reconfirm-
ing that no neo or aadA sequences are present in the nuclear genome
of the two transplastomic parental lines.
Junctions between three of the transposed chloroplast sequences
and nuclear DNA were obtained (see Supplementary Information)
by inverse polymerase chain reactions (iPCR). Analysis of a junction
sequence from kr1 (1,315 bp) identified 1,128 bp of non-plastid
DNA adjoining the transposed cpDNA. A sequence located at the 50
end of this non-plastid DNA showed similarity to two nuclear
encoded genes: 67 bp to a Nicotiana plumbaginifolia manganese
superoxide dismutase gene (89%) and 66 bp to tobacco TA-29/
TSJT1 (86%) (EMBL accession numbers Z67979 and X52283,
respectively). No other significant homology was identified within
this sequence, although a 300-bp open reading frame was present at
the 30end. To exclude the possibility that this iPCR product was
artefactual, the junction sequence was amplified by PCR from
genomic DNA. A 0.72-kb junction fragment was amplified from
the parental tp7-kr1 DNA, DNAs from five kanamycin-resistant
progeny of wild type £tp7-kr1 crosses and from the iPCR product
(Fig. 3e). No PCR product was amplified from 5 kanamycin-
sensitive progeny (Fig. 3e) or from the other 16 kanamycin-resistant
plants derived from independent transpositions (data not shown).
These results, together with the genetic and sequence data,
unequivocally demonstrate that the transposed cpDNA in tp-kr1
is integrated into nuclear DNA.
Junction sequences from kr17 (1,451 bp) and kr18 (585bp) DNA
were shown to be unique to the nuclear genome of the plants from
which they were derived and their kanamycin-resistant progeny
(Fig. 3f, g). Sequence analysis of the non-plastid DNA adjacent to
the kr17 and kr18 junctions showed no homology to sequences
deposited in databases except for three short tracts (20–60 bp) that
were similar to chloroplast DNA or mtDNA (Fig. 3f, g). As pre-
existing cpDNA and mtDNA may be very common in the tobacco
genome
11
, it is not surprising to find other organellar DNA in the
vicinity of an experimental integrant. The three junction sequences
indicate that the transposed plastid fragments have integrated into
different nuclear chromosomal locations.
These data demonstrate that DNA is transferred from the
chloroplast and integrated into the nucleus at a frequency of one
in approximately 16,000 tobacco pollen grains (16 independent,
heritable nuclear insertions in 250,000 seedlings). This is a mini-
mum estimate of the proportion of pollen grains that contain newly
transposed cpDNA because our experimental approach would have
identified only those integrants that were sufficiently large to
contain an expressed neoSTLS2 gene. In addition, as the integrants
that we identified are smaller than the entire plastome, we expect
similar transposition events involving other parts of the chloroplast
genome that cannot be monitored by the neoSTLS2 reporter gene.
Transfer of organellar DNA to the nucleus has been a major
driving force in the evolution of eukaryotic cells. Transfer of cpDNA
provides primary sequences for the evolution of new nuclear genes
1
and its frequency is such that integration could have a potentially
wide impact on the function of native nuclear genes and genome
organization. Despite this frequent transfer process the tobacco
nuclear genome does not seem to be continuously expanding, so we
assume that an equilibrium exists between integration and deletion.
It is noteworthy that two kanamycin-resistant plants did not
transmit the transposed plastid sequence to their progeny.
Transfer of cpDNA is the first evolutionary event necessary for
functional establishment of a plastid gene in the nucleus. As the
experimental cp-neoSTLS2 gene contains a constitutive nuclear
promoter, it allowed us to measure the primary rate of transposition
to the nucleus, which does not reflect the problems faced by native
chloroplast genes. The latter need to obtain appropriate nuclear
sequences for transcription, translation and protein transport back
into the chloroplast. A rate several orders of magnitude lower is
expected for such complex DNA arrangements
21
.
The chloroplast is an increasingly attractive location for high level
expression of prokaryote and eukaryote genes
22,23
, and there is the
advantage of natural biological containment afforded by maternal
inheritance that prevents pollen-mediated transmission of the
transplastome in many species
19,24
. Our findings identify a potential
pathway for transfer of plastid transgene DNA in tobacco. However,
genes tailored for expression in the chloroplast are unlikely to
letters to nature
NATURE| VOL 422 | 6 MARCH 2003 | www.nature.com/nature 75
© 2003 Nature Publishing Group
function in the nucleus as is demonstrated by the absence of
spectinomycin resistance, despite the presence of a nuclear aadA
gene in all of the kanamycin-resistant plants selected. A
Methods
pPRV111A::neoSTLS2 construction and transplastomic plants
pPRV111A::neoSTLS2 was produced by ligating the HindIII fragment containing
neoSTLS2 from plasmid pCMneoSTLS2 (ref. 13) into pPRV111A
14
. This transformation
vector was introduced into the plastids of Nicotiana tabacum L. Petit Havana (N,N) by
biolistic bombardment, and homoplasmic plants were regenerated as described
previously
14
. The homoplasmic plants were transferred to soil and grown in a controlled
environment chambe r with a 14 h light/10 h dark and 25 8C day/20 8C night growth
regime. The photon flux density was approximately 300
m
mol m
22
s
21
at the plant surface.
The nuclear neoSTLS2 control line was generated by biolistic bombardment with plasmid
pCMneoSTLS2 (ref. 13).
Seedling tests for resistance to kanamycin and spectinomycin
Surface-sterilized seeds were plated on 150-mm plates containing 50 ml of 0.5 MS salt
medium
25
and either spectinomycin dihydrochloride (500
m
gml
21
) or kanamycin
sulphate (150
m
gml
21
). For screening seeds of self-pollinated tp7, the latter medium was
supplemented with geneticin disulphate (20
m
gml
21
) when seedlings were two weeks old.
The plates were placed at 25 8C with continuous fluorescent light. To determine
nondestructively the kanamycin-resistance phenotype for verification of transplastomic/
nuclear junctions, leaf pieces were taken from individual seedlings grown in 0.5 MS
medium without kanamycin, and then cultured in MS medium containing 150
m
gml
21
kanamycin and the required plant hormones
25
. Kanamycin resistance was judged by callus
growth.
Molecular analysis
DNA and RNA blot analyses were carried out as described
11,26
. Junction sequences were
obtained using iPCR as described (see http://arabi4.agr.hokudai.ac.jp/ArabiE/protocols/
general/general.html). Key differences in the restriction fragment sizes between nuclear
and transplastomic DNA were used to design cpDNA primers adjacent to the site of
integration. These primers were used in conjunction with a second primer specific for
either aadA or neoSTLS2. Primers used in iPCR were 5 0-GAAGTTTCCAAAAGGTCGTT-
30with 5 0-CTCGCCATCTATTTTCATTG-30for kr1, 5 0-CCAGATTCCAAATGAACAAA-
30(f2) with 5 0-CAATAGCCCTCTGGTCTTCT-30(r3) for kr17, and f2 with r3 for kr18.
PCR amplification
11
of the junction sequences was undertaken with primers 5 0-
GCACTGTGTCATTCAATACT-30(f1) and 5 0-CCAATTGTGACATCCCTTCT-30(r1) for
kr1, f2 and 50-GGTTTTCCAAAGGGGTTTT-30(r2) for kr17, and 5 0-GCCGTCATCAC
TAACCATT-3 0(f3) and r3 for kr18.
Received 26 November 2002; accepted 13 January 2003; doi:10.1038/nature01435.
Published online 5 February 2003.
1. Martin, W. et al. Evolutionary analysis of Arabidopsis, cyanobacterial, and chloroplast genomes reveals
plastid phylogeny and thousands of cyanobacterial genes in the nucleus. Proc. Natl Acad. Sci. USA 99,
12246–12251 (2002).
2. Kaneko, T. et al. Sequence analysis of the genome of the unicellular cyanobacterium Synechocystis sp.
strain PCC6803. II. Sequence determination of the entire genome and assignment of potential
protein-coding regions. DNA Res. 3, 109–136 (1996).
3. Gray, M. W. The endosymbiont hypothesis revisited. Int. Rev. Cytol. 141, 233–357 (1992).
4. Adams, K. L., Daley, D. O., Qiu, Y. L., Whelan, J. & Palmer, J.D. Repeated, recent and diverse transfers
of a mitochondrial gene to the nucleus in flowering plants. Nature 408, 354–357 (2000).
5. Millen, R. S. et al. Many parallel losses of infA from chloroplast DNA during angiosperm evolution
with multiple independent transfers to the nucleus. Plant Cell 13, 645–658 (2001).
6. Stupar, R. M. et al. Complex mtDNA constitutes an approximate 620-kb inserti on on Arabidopsis
thaliana chromosome 2: implication of potential sequencing errors caused by large-unit repeats. Proc.
Natl Acad. Sci. USA 98, 5099–5103 (2001).
7. Yuan, Q. et al. Genome sequencing of 239-kb region of rice chromosome 10L reveals a high
frequency of gene duplication and a large chloroplast DNA insertion. Mol. Genet. Gen. 267, 713–720
(2002).
8. Timmis, J. N. & Scott, N. S. Spinach nuclear and chloroplast DNAs have homologous sequences.
Nature 205, 65–67 (1983).
9. Mourier, T., Hansen, A. J., Willerslev, E. & Arctander, P. The human genome project reveals a
continuous transfer of large mitochondrial fragments to the nucleus. Mol. Biol. Evol. 18, 1833–1837
(2001).
10. Ricchetti, M., Fairhead, C. & Dujon, B. Mitochondrial DNA repairs double-strand breaks in yeast
chromosomes. Nature 402, 96–100 (1998).
11. Ayliffe, M. A. & Timmis, J. N. Tobacco nuclear DNA contains long tracts of homology to chloroplast
DNA. Theor. Appl. Genet. 85, 229–238 (1992).
12. Thorsness, P. E. & Fox, T. D. Escape of DNA from the mitochondria to the nucleus in the yeast,
Saccharomyces cerevisiae.Nature 346, 376–379 (1990).
13. Maas, C. et al. Expression of intron modified NPTII genes in monocotyledonous and dicotyledonous
plant cells. Mol. Breed. 3, 15–28 (1997).
14. Zoubenko, O. V., Allison, L. A., Svab, Z. & Maliga, P. Efficient targeting of foreign genes into the
tobacco plastid genome. Nucleic Acids Res. 22, 3819–3824 (1994).
15. Benfey, P. N. & Chua, N. H. The cauliflower mosaic virus 35S promoter: combinatorial regulation of
transcription in plants. Science 250, 959–966 (1990).
16. Eckes, P., Rosahl, S., Schell, J. & Willmitzer, L. Isolation and characterization of a light-inducible,
organ-specific gene from potato and analysis of its expression after tagging and transfer into tobacco
and potato shoots. Mol. Gen. Genet. 205, 14–22 (1986).
17. Whitney, S. M. & Andrews, T. J. The gene for the ribulose-1, 5-bisphosphate carboxylase/oxygenase
(Rubisco) small subunit relocated to the plastid genome of tobacco directs the synthesis of small
subunits that assemble into Rubisco. Plant Cell 13, 193–205 (2001).
18. Horser, C., Abbott, D., Wesley, V., Smith, N. & Waterhouse, P. Gene silencing
principles and
application. Genet. Eng. Principles Meth. 24, 239–256 (2002).
19. Daniell, H., Datta, R., Varma, S., Gray, S. & Lee, S. B. Containment of herbicide resistance through
genetic engineering of the chloroplast genome. Nature Biotechnol. 16, 345–348 (1998).
20. Fagard, M. & Vaucheret, H. (Trans)gene silencing in plants: how many mechanisms? Annu. Rev. Plant
Physiol. Plant Mol. Biol. 51, 167–194 (2000).
21. Babiychuk, E., Fuangthong, M., van Montagu, M., Inze
´, D. & Kushnir, S. Efficient gene tagging in
Arabidopsis thaliana using a gene trap approach. Proc. Natl Acad. Sci. USA 94, 12722–12727
(1997).
22. Maliga, P. Engineering the plastid genome of higher plants. Curr. Opin. Plant Biol. 5, 164–172 (2002).
23. Daniell, H., Khan, M. S. & Allison, L. Milestones in chloroplast genetic engineering: an
environmentally friendly era in biotechnology. Trends Plant Sci. 7, 84–91 (2002).
24. Scott, S. E. & Wilkinson, M. J. Low probability of chloroplast movement from oilseed rape (Brassica
napus)intowildBrassica rapa.Nature Biotechnol. 17, 390–392 (1999).
25. Murashige, T. & Skoog, F. A revised medium for rapid growth and bioassays with tobacco tissue
cultures. Physiol. Plant 15, 473–497 (1962).
26. Huang, C. Y., Barker, S. J., Langridge, P., Smith, F. W. & Graham, R. D. Zinc deficiency up-regulates
expression of high-affinity phosphate transporter genes in both phosphate-sufficient and -deficient
barley roots. Plant Physiol. 124, 415–422 (2000).
Supplementary Information accompanies the paper on Nature’s website
(çhttp://www.nature.com/nature).
Acknowledgements We thank the Australian Research Council for financial support; P. Maliga
for pPRV111A; S. Whitney and J.Andrews for facilities and advice on chloroplast transformation,
for SSuH2 seeds and for comments on the manuscript; T. J. Higgins, P. Whitfeld, W. Martin and
R. Lockington for discussion; C. Maas for pCMneoSTLS2; N. Smith for the NS23 seeds; and S. Elts
for technical assistance.
Competing interests statement The authors declare that they have no competing financial
interests.
Correspondence and requests for materials should be addressed to J.N.T.
(e-mail: jeremy.timmis@adelaide.edu.au). Nucleotide sequences of three transplastomic/nuclear
junctions are deposited in EMBL under accession numbers AJ495859 for kr1, AJ517467 for kr17
and AJ517468 for kr18.
..............................................................
Optimal transsaccadic integration
explains distorted spatial perception
Matthias Niemeier*†‡, J. Douglas Crawford†‡ & Douglas B. Tweed*†‡
*Departments of Physiology and Medicine, University of Toronto, 1 King’s College
Circle, Toronto M5S 1A8, Canada
Canadian Institutes of Health Research, Group for Action and Perception
Centre for Vision Research, York University, 4700 Keele Street, Toronto M3J 1P3,
Canada
.............................................................................................................................................................................
We scan our surroundings with quick eye movements called
saccades, and from the resulting sequence of images we build a
unified percept by a process known as transsaccadic integration.
This integration is often said to be flawed, because around the
time of saccades, our perception is distorted
1–6
and we show
saccadic suppression of displacement (SSD): we fail to notice if
objects change location during the eye movement
7,8
. Here we
show that transsaccadic integration works by optimal inference.
We simulated a visuomotor system with realistic saccades, retinal
acuity, motion detectors and eye-position sense, and pro-
grammed it to make optimal use of these imperfect data when
interpreting scenes. This optimized model showed human-like
SSD and distortions of spatial perception. It made new predic-
tions, including tight correlations between perception and motor
action (for example, more SSD in people with less-precise eye
control) and a graded contraction of perceived jumps; we verified
these predictions experimentally. Our results suggest that the
brain constructs its evolving picture of the world by optimally
integrating each new piece of sensory or motor information.
letters to nature
NATURE| VOL 422 | 6 MARCH 2003 | www.nature.com/nature76 © 2003 Nature Publishing Group
... More than 11 Kb DNA insertions were also detected in Arabidopsis [18]. Huang et al. used a neomycin phosphotransferase to demonstrate that a gene was transferred from the chloroplast to the nucleus [19]. However, the mechanism by which plastid DNA sequences are transferred has yet to be clarified, and little is known about how the nuclear genome is integrated. ...
... During the course of evolution, there have been frequent exchanges of genes among the different genetic systems. DNA movement occurs when genes are transferred from organelles to the nucleus [19,56]. In this study, we identified 163-194 orthologous genes of chloroplasts and transcriptomes of five species of Deutzia spp. ...
Article
Full-text available
In this study, the chloroplast genomes and transcriptomes of five Deutzia genus species were sequenced, characterized, combined, and analyzed. A phylogenetic tree was constructed, including 32 other chloroplast genome sequences of Hydrangeoideae species. The results showed that the five Deutzia chloroplast genomes were typical circular genomes 156,860-157,025 bp in length, with 37.58-37.6% GC content. Repeat analysis showed that the Deutzia species had 41-45 scattered repeats and 199-201 simple sequence repeats. Comparative genomic and pi analyses indicated that the genomes are conservative and that the gene structures are stable. According to the phylogenetic tree, Deutzia species appear to be closely related to Kirengeshoma palmata and Philadelphus. By combining chloroplast genomic and transcriptomic analyses, 29-31 RNA editing events and 163-194 orthologous genes were identified. The ndh, rpo, rps, and atp genes had the most editing sites, and all RNA editing events were of the C-to-U type. Most of the orthologous genes were annotated to the chloroplast, mitochondria, and nucleus, with functions including energy production and conversion, translation, and protein transport. Genes related to the biosynthesis of monoterpenoids and flavonoids were also identified from the transcriptome of Deutzia spp. Our results will contribute to further studies of the genomic information and potential uses of the Deutzia spp.
... For example, in yeast (Saccharomyces cerevisiae), gene transfer from mitochondria to the nucleus occurs at a high frequency, but the opposite process of gene transfer from the nucleus to mitochondria is much less frequent [8]. In tobacco (Nicotiana tabacum), gene transfer from chloroplasts to the nucleus occurs at a high frequency during pollen development and in somatic leaf tissue [9,10], and mild heat treatment mimicking environmental stress increases this frequency, as well as that of gene transfer from mitochondria to the nucleus, probably because heat treatment disrupts these organelles [11]. Thus, there is ongoing and fairly frequent gene ux between organelles and the nucleus; however, it remains unknown whether gene transfer also occurs from chloroplasts to mitochondria. ...
... Although the sulfadiazine selection may increase the frequency, this frequency should not re ect naturally occurring gene transfer from chloroplasts to mitochondria, since such transferred genes should be randomly integrated into the chloroplast genome without any homologous sequences, and it would be technically di cult to detect such random integration. Moreover, the frequency of such transfer could vary among cell types, considering that gene transfer from chloroplasts to the nucleus is more frequent in reproductive cells than in somatic cells [9,10,32], and gene transfer needs to occur in reproductive cells to be inherited in the next generation. In summary, our results indicate that there is ongoing gene transfer from chloroplasts to mitochondria. ...
Preprint
Full-text available
Angiosperm mitochondrial genomes have highly complex and diverse structures that are partly due to frequent insertions of nuclear and chloroplast DNA (cpDNA) into mitochondrial DNA (mtDNA). This suggests the existence of mechanisms for gene transfer from chloroplasts to mitochondria, but these have yet to be discovered. In this study, we aimed to detect chloroplast-to-mitochondrion gene transfer by analyzing the translocation of a marker gene, sul , encoding a bacterial dihydropteroate synthase that confers sulfonamide resistance in tobacco ( Nicotiana tabacum ), to mtDNA. First, we created tobacco chloroplast transformants in which sul , surrounded on both sides by ~ 1 kb of mitochondrial homologous sequences that enable targeted integration into mtDNA, was introduced into the chloroplast genome. Heat shock enhanced sul expression in the transformants, suggesting that chloroplast degradation stimulates gene transfer from chloroplasts to mitochondria. Shoot regeneration using the heat-shocked chloroplast transformants under sulfadiazine selection resulted in several transformants showing moderately resistant to sulfadiazine. Deep sequencing analysis of the target mitochondrial locus detected sul in the SR plants with an integration efficiency of 0.0007–0.0036%, and we validated the results by ruling out two types of artifactual outcomes, PCR jumping and sul integration into nuclear mitochondrial DNA (NuMT). From these results, we propose that gene transfer from chloroplasts to mitochondria is ongoing in tobacco.
... The presence of a nuclear and two organellar (plastid and mitochondrial) genomes results in protein synthesis occurring in three separate compartments. Although the bacterial progenitors of plastids and mitochondria harbored all genetic components required for translation, their genomes have since been extensively reduced, and numerous proteins involved in organellar translation are now encoded in the nucleus and imported into the organelles (Huang et al. 2003;Timmis et al. 2004;Giannakis et al. 2022). Transfer RNAs (tRNAs) are some of the last remaining translational components encoded in organellar genomes. ...
Article
Full-text available
The number of tRNAs encoded in plant mitochondrial genomes varies considerably. Ongoing loss of bacterial-like mitochondrial tRNA genes in many lineages necessitates the import of nuclear-encoded counterparts that share little sequence similarity. Because tRNAs are involved in highly specific molecular interactions, this replacement process raises questions about the identity and trafficking of enzymes necessary for the maturation and function of newly imported tRNAs. In particular, the aminoacyl-tRNA synthetases (aaRSs) that charge tRNAs are usually divided into distinct classes that specialize on either organellar (mitochondrial and plastid) or nuclear-encoded (cytosolic) tRNAs. Here, we investigate the evolution of aaRS subcellular localization in a plant lineage (Sileneae) that has experienced extensive and rapid mitochondrial tRNA loss. By analyzing full-length mRNA transcripts (PacBio Iso-Seq), we found predicted retargeting of many ancestrally cytosolic aaRSs to the mitochondrion and confirmed these results with colocalization microscopy assays. However, we also found cases where aaRS localization does not appear to change despite functional tRNA replacement, suggesting evolution of novel interactions and charging relationships. Therefore, the history of repeated tRNA replacement in Sileneae mitochondria reveals that differing constraints on tRNA/aaRS interactions may determine which of these alternative coevolutionary paths is used to maintain organellar translation in plant cells.
... The reduction of the size and gene content by transfer the plastid and the nuclear genome interdependent. Experiments assessing DNA transfer from plastids to nuclei estimate the rate of transfer to be as high as 1 in 16,000 in pollen grains (Huang et al., 2003) to as low as 1 in 5 million in vegetative cells (Stegemann et al., 2003), suggesting this transfer may have occurred rapidly after endosymbiosis. ...
Article
Full-text available
The spectacular diversity of plastids in non-green organs such as flowers, fruits, roots, tubers, and senescing leaves represents a Universe of metabolic processes in higher plants that remain to be completely characterized. The endosymbiosis of the plastid and the subsequent export of the ancestral cyanobacterial genome to the nuclear genome, and adaptation of the plants to all types of environments has resulted in the emergence of diverse and a highly orchestrated metabolism across the plant kingdom that is entirely reliant on a complex protein import and translocation system. The TOC and TIC translocons, critical for importing nuclear-encoded proteins into the plastid stroma, remain poorly resolved, especially in the case of TIC. From the stroma, three core pathways (cpTat, cpSec, and cpSRP) may localize imported proteins to the thylakoid. Non-canonical routes only utilizing TOC also exist for the insertion of many inner and outer membrane proteins, or in the case of some modified proteins, a vesicular import route. Understanding this complex protein import system is further compounded by the highly heterogeneous nature of transit peptides, and the varying transit peptide specificity of plastids depending on species and the developmental and trophic stage of the plant organs. Computational tools provide an increasingly sophisticated means of predicting protein import into highly diverse non-green plastids across higher plants, which need to be validated using proteomics and metabolic approaches. The myriad plastid functions enable higher plants to interact and respond to all kinds of environments. Unraveling the diversity of non-green plastid functions across the higher plants has the potential to provide knowledge that will help in developing climate resilient crops.
... Horizontally transferred DNA is highly unstable in the genome and is often deleted within one plant generation (Huang et al., 2003). The instability of these endogenized elements in the host genome is thus inconsistent with the persistence of EVEs in the plant genome for thousands of years after integration. ...
Thesis
Full-text available
Florendoviruses are novel endogenous viruses in the family Caulimoviridae, which are widely distributed in the genomes of angiosperms. It is assumed that endogenous florendovirus elements (EFEs) have become fixed in plant genomes because they contribute beneficial functions to normal plant physiology. Caulimovirid promoters, such as the cauliflower mosaic virus 35S promoter, are typically very strong promoters. Thus, the insertion of a florendovirus promoter upstream of an endogenous plant gene could dramatically change the level of its activity. In the course of this study, florendovirus promoters from Lotus japonicus and Glycine max were identified and determined to be active in a transient transformation assay. By site-directed mutagenesis, predictions of the positions of the TATA-box elements in the promoters were also confirmed. Reversetranscription polymerase chain reaction assays demonstrated that sequences downstream of the florendovirus promoters were transcribed in their host plant, indicating the involvement of EFEs in the gene regulatory networks of the plant. These findings will lay the foundation for future studies on the functions and contributions of these viruses to normal plant physiology.
... Organellar sequences continue to be identified in sequenced eukaryotic nuclear genomes, and there is growing evidence that mitochondrial and plastid DNA continues to insert into nuclear sequence and may play a role in human diseases, including cancer (Wei et al., 2022). Two independent groups observed de novo integrations of plastid DNA (NUPTs) in the tobacco nuclear genome (Huang et al., 2003;Stegemann et al., 2003). NUMTs have been observed and their frequency increased by mutations in Saccharomyces cerevisiae Fox, 1990, 1993). ...
Article
Full-text available
Genomic assemblies of the unicellular green alga, Chlamydomonas reinhardtii provided important resources for researchers. However, assembly errors, large gaps, and unplaced scaffolds as well as strain-specific variants currently impede many types of analyses. By combining PacBio HiFi and Oxford Nanopore long-read technologies, we generated a de novo genome assembly for strain CC-5816, derived from crosses of strains CC-125 and CC-124. Multiple methods of evaluating genome completeness and base-pair error rate suggest the final telomere-to-telomere assembly is highly accurate. The CC-5816 assembly enabled previously difficult analyses that include characterization of the 17 centromeres, rDNA arrays on 3 chromosomes, and 56 insertions of organellar DNA into the nuclear genome. Using Nanopore sequencing, we identified sites of cytosine (CpG) methylation, which are enriched at centromeres. We analyzed CRISPR/Cas9 insertional mutants in the PF23 gene. Two of the three alleles produced progeny that displayed patterns of meiotic inviability that suggested the presence of a chromosomal aberration. Mapping Nanopore reads from pf23-2 and pf23-3 onto the CC-5816 genome showed that these two strains each carry a translocation that was initiated at the PF23 gene locus on Chromosome 11 and joined with Chromosomes 5 or 3, respectively. The translocations were verified by demonstrating linkage between loci on the two translocated chromosomes in meiotic progeny. The three pf23 alleles display the expected short cilia phenotype and pf23-2 lacks the PF23 protein by immunoblotting. Our CC-5816 genome assembly will undoubtedly provide an important tool for the Chlamydomonas research community.
Article
Horizontal gene transfer (HGT), or lateral gene transfer, is the non-sexual movement of genetic information between genomes. It has played a pronounced part in bacterial and archaeal evolution, but its role in eukaryotes is less clear. Behaviours unique to eukaryotic cells - phagocytosis and endosymbiosis - have been proposed to increase the frequency of HGT, but nuclear genomes encode fewer HGTs than bacteria and archaea. Here, I review the existing theory in the context of the growing body of data on HGT in eukaryotes, which suggests that any increased chance of acquiring new genes through phagocytosis and endosymbiosis is offset by a reduced need for these genes in eukaryotes, because selection in most eukaryotes operates on variation not readily generated by HGT.
Article
Full-text available
Plastid transformation technology has been widely used to express traits of potential commercial importance, though the technology has been limited to traits that function while sequestered in the organelle. Prior research indicates that plastid contents can escape from the organelle, suggesting a possible mechanism for engineering plastid transgenes to function in other cellular locations. To test this hypothesis, we created tobacco (Nicotiana tabacum cv. Petit Havana) plastid transformants that express a fragment of the nuclear-encoded Phytoene desaturase (PDS) gene capable of catalyzing post-transcriptional gene silencing if RNA escapes into the cytoplasm . We found multiple lines of direct evidence that plastid-encoded PDS transgenes affect nuclear PDS gene silencing: knockdown of the nuclear-encoded PDS mRNA and/or its apparent translational inhibition, biogenesis of 21-nucleotide (nt) phased small interfering RNAs (phasiRNAs), and pigment- deficient plants. Furthermore, plastid-expressed double-stranded RNA (dsRNA) with no cognate nuclear-encoded pairing partner also produced abundant 21-nt phasiRNAs in the cytoplasm, demonstrating that a nuclear-encoded template is not required for siRNA biogenesis. Our results indicate that RNA escape from plastids to the cytoplasm occurs generally, with functional consequences that include entry into the gene silencing pathway. Furthermore, we uncover a method to produce plastid-encoded traits with functions outside of the organelle and open additional fields of study in plastid development, compartmentalization, and small RNA biogenesis.
Article
Full-text available
Background Species of Broussonetia (family Moraceae) are commonly used to make textiles and high-grade paper. The distribution of Broussonetia papyrifera L. is considered to be related to the spread and location of humans. The complete chloroplast (cp) genomes of B. papyrifera , Broussonetia kazinoki Sieb., and Broussonetia kaempferi Sieb. were analyzed to better understand the status and evolutionary biology of the genus Broussonetia . Methods The cp genomes were assembled and characterized using SOAPdenovo2 and DOGMA. Phylogenetic and molecular dating analysis were performed using the concatenated nucleotide sequences of 35 species in the Moraceae family and were based on 66 protein-coding genes (PCGs). An analysis of the sequence divergence (pi) of each PCG among the 35 cp genomes was conducted using DnaSP v6. Codon usage indices were calculated using the CodonW program. Results All three cp genomes had the typical land plant quadripartite structure, ranging in size from 160,239 bp to 160,841 bp. The ribosomal protein L22 gene ( RPL22 ) was either incomplete or missing in all three Broussonetia species. Phylogenetic analysis revealed two clades. Clade 1 included Morus and Artocarpus , whereas clade 2 included the other seven genera. Malaisia scandens Lour. was clustered within the genus Broussonetia . The differentiation of Broussonetia was estimated to have taken place 26 million years ago. The PCGs’ pi values ranged from 0.0005 to 0.0419, indicating small differences within the Moraceae family. The distribution of most of the genes in the effective number of codons plot (ENc-plot) fell on or near the trend line; the slopes of the trend line of neutrality plots were within the range of 0.0363–0.171. These results will facilitate the identification, taxonomy, and utilization of the Broussonetia species and further the evolutionary studies of the Moraceae family.
Article
Full-text available
Long tracts of DNA with high sequence homology to chloroplast DNA were isolated from nuclear genomic libraries of Nicotiana tabacum. One lambda EMBL4 clone was characterised in detail and assigned to nuclear DNA. The majority of the 15.5-kb sequence is greater than 99% homologous with its chloroplast DNA counterpart, but a single base deletion causes premature termination of the reading frame of the psaA gene. One region of the clone contains a concentration of deleted regions, and these were used to identify and quantify the sequence in native nuclear DNA by polymerase chain reaction (PCR) methods. An estimated 15 copies of this specific region are present in a 1c tobacco nucleus.
Article
Full-text available
Relatively few chloroplast (ct) or mitochondrial (mt) proteins are encoded and synthesized within the organelles, the majority being made on cytoplasmic ribosomes from messenger RNAs transcribed from nuclear genes1,2. How this interaction of genetic systems evolved is poorly understood. Whether the chloroplast or mitochondrial specific genes now residing in the nucleus arose de novo, or by genetic exchange between the nucleus and a symbiotic autonomous prokaryote, is an intriguing unanswered question, although that such genetic exchange is possible is strongly suggested by recent reports of homology between maize mitochondrial and chloroplast genomes3 and the presence of mitochondrial sequences in yeast nuclei4. We now present evidence that the spinach nucleus contains integrated sequences that are homologous to ctDNA sequences and that these sequences are incorporated at specific sites within the genome. The experiments also show a region of homology between ctDNA and mtDNA.
Article
Full-text available
A central component of the endosymbiotic theory for the bacterial origin of the mitochondrion is that many of its genes were transferred to the nucleus. Most of this transfer occurred early in mitochondrial evolution; functional transfer of mitochondrial genes has ceased in animals. Although mitochondrial gene transfer continues to occur in plants, no comprehensive study of the frequency and timing of transfers during plant evolution has been conducted. Here we report frequent loss (26 times) and transfer to the nucleus of the mitochondrial gene rps10 among 277 diverse angiosperms. Characterization of nuclear rps10 genes from 16 out of 26 loss lineages implies that many independent, RNA-mediated rps10 transfers occurred during recent angiosperm evolution; each of the genes may represent a separate functional gene transfer. Thus, rps10 has been transferred to the nucleus at a surprisingly high rate during angiosperm evolution. The structures of several nuclear rps10 genes reveal diverse mechanisms by which transferred genes become activated, including parasitism of pre-existing nuclear genes for mitochondrial or cytoplasmic proteins, and activation without gain of a mitochondrial targeting sequence.
Article
Phosphate (P) is taken up by plants through high-affinity P transporter proteins embedded in the plasma membrane of certain cell types in plant roots. Expression of the genes that encode these transporters responds to the P status of the plants, and their transcription is normally tightly controlled. However, this tight control of P uptake is lost under Zn deficiency, leading to very high accumulation of P in plants. We examined the effect of plant Zn status on the expression of the genes encoding the HVPT1 and HVPT2 high-affinity P transporters in barley (Hordeum vulgareL. cv Weeah) roots. The results show that the expression of these genes is intimately linked to the Zn status of the plants. Zn deficiency induced the expression of genes encoding these P transporters in plants grown in either P-sufficient or -deficient conditions. Moreover, the role of Zn in the regulation of these genes is specific in that it cannot be replaced by manganese (a divalent cation similar to Zn). It appears that Zn plays a specific role in the signal transduction pathway responsible for the regulation of genes encoding high-affinity P transporters in plant roots. The significance of Zn involvement in the regulation of genes involved in P uptake is discussed.
Article
We used DNA sequencing and gel blot surveys to assess the integrity of the chloroplast gene infA, which codes for translation initiation factor 1, in >300 diverse angiosperms. Whereas most angiosperms appear to contain an intact chloroplast infA gene, the gene has repeatedly become defunct in ∼24 separate lineages of angiosperms, including almost all rosid species. In four species in which chloroplast infA is defunct, transferred and expressed copies of the gene were found in the nucleus, complete with putative chloroplast transit peptide sequences. The transit peptide sequences of the nuclear infA genes from soybean and Arabidopsis were shown to be functional by their ability to target green fluorescent protein to chloroplasts in vivo. Phylogenetic analysis of infA sequences and assessment of transit peptide homology indicate that the four nuclear infA genes are probably derived from four independent gene transfers from chloroplast to nuclear DNA during angiosperm evolution. Considering this and the many separate losses of infA from chloroplast DNA, the gene has probably been transferred many more times, making infA by far the most mobile chloroplast gene known in plants.
Article
The sequence determination of the entire genome of the Synechocystis sp. strain PCC6803 was com- pleted. The total length of the genome finally confirmed was 3,573,470 bp, including the previously reported sequence of 1,003,450 bp from map position 64% to 92% of the genome. The entire sequence was assem- bled from the sequences of the physical map-based contigs of cosmid clones and of A clones and long PCR products which were used for gap-filling. The accuracy of the sequence was guaranteed by analysis of both strands of DNA through the entire genome. The authenticity of the assembled sequence was supported by restriction analysis of long PCR products, which were directly amplified from the genomic DNA using the assembled sequence data. To predict the potential protein-coding regions, analysis of open reading frames (ORFs), analysis by the GeneMark program and similarity search to databases were performed. As a result, a total of 3,168 potential protein genes were assigned on the genome, in which 145 (4.6%) were identical to reported genes and 1,257 (39.6%) and 340 (10.8%) showed similarity to reported and hypothet- ical genes, respectively. The remaining 1,426 (45.0%) had no apparent similarity to any genes in databases. Among the potential protein genes assigned, 128 were related to the genes participating in photosynthetic reactions. The sum of the sequences coding for potential protein genes occupies 87% of the genome length. By adding rRNA an,d tRNA genes, therefore, the genome has a very compact arrangement of protein- and RNA-coding regions. A notable feature on the gene organization of the genome was that 99 ORFs, which showed similarity to transposase genes and could be classified into 6 groups, were found spread all over the genome, and at least 26 of them appeared to remain intact. The result implies that rearrangement of the genome occurred frequently during and after establishment of this species.
Article
Intron sequences from monocotyledonous and dicotyledonous origin were used to abolish marker gene expression in prokaryotes (Escherichia coli and Agrobacterium tumefaciens) but permit expression in selected eukaryotic systems using the eukaryotic specific splicing mechanism. A 1014 bp maize Shrunken-1 (Sh 1) intron 1 flanked by exon1 and exon2 sequences was cloned into the N-terminal of the NPT II-coding region. Transient gene expression analysis revealed that the modified neomycin phosphotransferase II (NPT II) gene, driven by the cauliflower mosaic virus (CaMV) 35S promoter, is expressed in barley protoplasts, but poorly expressed in tobacco protoplasts. In dicotyledonous cells AU-rich sequences are known to be important for efficient splicing and therefore an attempt was made to improve expression of the NPT II gene, containing the Sh 1 intron 1, in tobacco by increasing the AU content from 57% to 69%. Reverse transcriptase PCR analysis of RNA from transiently expressed NPT II transcripts from tobacco protoplasts revealed that despite the increase in AU-content, NPT II was still poorly expressed. Cryptic splice sites were identified as one possible cause for missplicing of the Sh1 intron 1 in dicots and poor levels of expression. Alternatively, cloning of the 198 bp intron 2 of the potato STLS 1 gene (81% AU) into the N-terminal part of the NPT II-coding region resulted in proper expression of NPT II in tobacco as well as in barley protoplasts and abolished marker gene expression in prokaryotes. The successful insertion of an intron into a selectable marker gene which completely abolishes gene expression in prokaryotes, without affecting expression of chimeric genes in monocotyledonous and dicotyledonous plant cells provides a suitable system to reduce the number of false-positives in transgenic plant production.
Article
The isolation and analysis of several cDNAs and of one genomic clone encoding ST-LS1, a single-copy gene fromSolanum tuberosum with leaf/stem-specific, light-inducible expression, is described. The structure of the gene was determined by sequencing several overlapping partial cDNA clones as well as the genomic clone and by determining the transcription start site by RNA protection experiments. A “tagged” derivative of the gene, obtained by exon modification, was reintroduced into potato and into tobacco shoots usingAgrobacterium/Ti-plasmid vector systems. The modified gene was expressed in both tobacco and potato shoots giving rise to an RNA of approximately 1,200 nucleotides which exceeded the length of the RNA made from the endogenous gene by the expected size of the “tag” (470 nucleotides). The level of expression of the modified gene varied substantially between independent transformants. A high proportion of the transformants (20%–40%) synthesized as much RNA from the added gene as from the resident gene. Expression of the transferred gene was light induced. Qualitatively and quantitatively the expression of the introduced gene was similar in a homologous (potato) and in a heterologous, but related, cellular background (tobacco).
Article
This chapter highlights endosymbiont hypothesis. All contemporary genomes (including those of plastids and mitochondria) ultimately derive from a single genome—the genome of a single, presumably cellular, entity which was the ancestor of all surviving forms of live. The construction and interpretation of phylogenetic trees based on small subunit (SSU, or 16S-like) and large subunit (LSU, or 23S-like) rRNA sequences have proven especially informative in instances where morphological diversity tends to confound traditional methods of phylogenetic analysis. In addition to ribosomal RNA (rRNA) sequence, there are a number of traits that characterize the archaebacteria as a distinct group of organisms, separate from all other prokaryotes. These include (1) cell walls that, when present, lack peptidoglycan and muramic acid; (2) the presence of lipids containing phytanyl side groups in ether linkage; and (3) the presence of RNA polymerases that are distinct from those of eubacteria in subunit composition, response to RNA synthesis inhibitors or stimulators, immunological reactivity, and gene sequence. Two major divisions of archaebacteria include (1) sulfur-dependent, extreme thermophiles; and (2) methane-producers (methanogens) and their relatives, including the extreme halophiles.