ArticlePDF Available

Chloroplast DNA Replication : Mechanism, Enzymes and Replication Origins

Authors:

Abstract

Chloroplasts contain circular DNA molecules which are found in low copy number in proplastids but are amplified to very high copy number in actively dividing leaf cells. A double displacement loop (D-loop) mechanism for chloroplast DNA (ctDNA) replication has been proposed, and pairs of replication origins which fit this model have been identified in some species. It appears that ctDNA replication is under the control of at least some nuclear gene products, as genes for DNA polymerase, topoisomerases, DNA primase and other accessory replication proteins have not been reported in the sequenced chloroplast genomes, and ctDNA replication remains active in the absence of active chloroplast transcription or translation. Only a few chloroplast replication proteins have been isolated, and to date most have not been characterized in detail. The mechanism by which ctDNA copy number is regulated during plant development is not known. In this review we summarize the current understanding of ctDNA replication.
eXtra Botany
COMMENTARY
Mechanisms for maintenance,
replication, and repair of the
chloroplast genome in plants
Brent L. Nielsen*, John D. Cupp and
Jeffrey Brammer
Department of Microbiology and Molecular Biology, Brigham
Young University, Provo, Utah 84602, USA
* E-mail: brentnielsen@byu.edu
Journal of Experimental Botany, Vol. 61, No. 10, pp. 2535–2537,
2010
doi:10.1093/jxb/erq163
Photosynthesis is a complex process that occurs in chlor-
oplasts of higher plants, and requires a large number of
proteins to assemble the photosynthetic machinery. Many
chloroplast-localized proteins are nuclear-encoded and must
be imported into the chloroplasts from the cytoplasm. A
considerable number of genes for photosynthesis and other
chloroplast functions, including transcription and trans-
lation, are encoded in the chloroplast genome (ctDNA),
which ranges in size from about 130–160 kbp in most higher
plants. CtDNA replication is not linked with the plant
cell cycle and the chloroplast genome can be amplified to
a very high copy number per cell in rapidly dividing leaf
tissue. Later in leaf development and plant growth, the
ctDNA levels reduce to very low levels (Oldenburg and
Bendich, 2004b). The controls that regulate ctDNA replica-
tion initiation, replication, and copy number are not
understood. From earlier publications on a number of plant
species it appears that ctDNA may replicate by more than
one mechanism, including a recombination-dependent
replication mechanism (Rowan et al., 2010, this issue;
Oldenburg and Bendich, 2004b;Marechal and Brisson,
2010), a double D-loop mechanism (Chiu and Sears, 1992;
Kunnimalaiyaan and Nielsen, 1997a,b), and rolling circle
replication (Kolodner and Tewari, 1975).
In this issue, Rowan et al. (2010) report on the role of
chloroplast-targeted RecA (cpRecA) in the maintenance
of ctDNA in Arabidopsis. Previously published reports
provide evidence that some ctDNA molecules may be
recombination intermediates as shown by the presence of
branched DNA molecules in some DNA preparations
(Oldenburg and Bendich, 2004a,b;Scharff and Koop,
2007). As summarized in a review by Marechal and Brisson
(2010), recombination has been shown to be involved in the
repair of double-strand breaks and point mutations in
ctDNA. It has been known for some time that a plant
homologue of bacterial RecA is localized in chloroplasts
(Cerutti et al., 1992), but, to date, little is known about the
role of DNA recombination in the maintenance of ctDNA.
Rowan et al. (2010) show clear evidence that cpRecA is
involved in the maintenance of the chloroplast genome copy
number in plants, as T-DNA insertions (from the Agro-
bacterium Ti plasmid) in the nuclear gene encoding this
protein led to a reduction in ctDNA copy number in the
mutant plants relative to wild-type plants and to a change
in the structure of the ctDNA. The levels of detectable
single-stranded DNA increased in the mutants, which is
compatible with the decreased amount of cpRecA which
would normally coat the single-stranded DNA regions and
thus block its detection. After a few generations the mutants
began to show significant signs of distress and reduced
chloroplast function, including variegation and necrosis.
These findings represent a significant advance in our
understanding of the mechanisms involved in the mainte-
nance of ctDNA integrity. The authors suggest that the role
of cpRecA is primarily in DNA repair, as supported by the
analysis of wild-type plants that have been treated with
ciprofloxacin, which induces double-strand DNA breaks. In
these plants, altered ctDNA structures were observed as
in the cpRecA plants. Similar experiments with insertions
in the DRT 100 homologue, which has only very weak
homology to bacterial RecA but can partially complement
E. coli recA mutants showed no effect, suggesting that DRT
100 may not be directly involved in the repair of ctDNA.
The role of cpRecA in DNA repair is clearly supported
by these experiments; it is also possible that cpRecA may
be involved in recombination-mediated replication of the
chloroplast genome.
CpRecA and DRT 100 are not the only RecA homo-
logues localized to chloroplasts. A dual-targeted (to both
chloroplasts and mitochondria) RecA (distinguished from
the others as RecA2) has been identified in the Arabidopsis
nuclear genome (Christensen et al., 2005). T-DNA inser-
tions in this gene lead to non-viable plants (BL Nielsen,
JD Cupp, unpublished observations; Shedge et al., 2007),
suggesting that RecA2 may be essential for ctDNA and/or
mtDNA maintenance and plant development. However, at
this point in time there are no data to determine whether
the lethal phenotype is due to the disruption of chloroplast
or mitochondrial DNA maintenance mechanisms, or both.
The RecA2 gene was not included in the current study by
Rowan et al. (2010, this issue) but its role in ctDNA
replication should be evaluated. The observation that
T-DNA insertions in cpRecA were not lethal may be due
to functional (at least partial) complementation by RecA2.
ªThe Author [2010]. Published by Oxford University Press [on behalf of the Society for Experimental Biology]. All rights reserved.
For Permissions, please e-mail: journals.permissions@oxfordjournals.org
at Serials Dept., Harold B. Lee Library, Brigham Young University on June 29, 2010 http://jxb.oxfordjournals.orgDownloaded from
But because RecA2 insertions appear to be lethal, RecA2
may play some specific and essential role in ctDNA and/or
mtDNA maintenance. It will be important to obtain and
analyse mutants of RecA2, which may need to be generated
by other approaches such as inducible RNAi, microRNA or
some other technique. If such mutants can be obtained,
then similar approaches to those used by Rowan et al.
should be used to analyse the role of RecA2 in maintaining
ctDNA copy number.
It is possible that early ctDNA replication during
germination and seedling development may be initiated
by one mechanism, such as the double D-loop mechanism,
and high level amplification of ctDNA may occur by
rolling circle replication. Alternatively, perhaps rolling circle
replication occurs initially, maybe in combination with
recombination-dependent replication, and the double D-loop
mechanism is then used as a way of maintaining ctDNA
later in plant development. Rolling circle replication is used
by many bacteriophages to produce large numbers of new
DNA molecules for progeny phage very quickly. Some
bacteriophages, including lambda phage, initiate replication
bidirectionally from a specific origin, similar to bacterial and
eukaryotic chromosomal DNA replication, but after com-
pletion of a unit circle shift to rolling circle DNA repli-
cation. Other phages such as T4 and T7 also have multiple
replication mechanisms, including replication from a specific
origin for both, and a recombination-dependent mechanism
for T4 (Mosig, 1998) and replication as concatamers for
T7 (as described in Oldenberg and Bendich, 2004a;Scharff
and Koop, 2006). Thus there is ample evidence from
other organisms and bacteriophages for multiple replication
mechanisms for individual genomes. This possibility should
be seriously considered for ctDNA, as many aspects of
chloroplast genome structure, transcription and replication
share similarities with bacteriophage mechanisms.
A double D-loop mechanism for ctDNA replication was
reported more than 30 years ago for Pisum sativum
(Kolodner and Tewari, 1975), which lacks the large inverted
repeat common to most higher plant chloroplast genomes.
In P. sativum, the two origins map within the rRNA spacer
region and just downstream of the 5S rRNA gene, about
6 kbp apart. The rRNA operon is present in the large
inverted repeat, so in species that have the inverted repeat
the two replication origins are duplicated. For example,
in tobacco ctDNA there are two identical pairs (one pair
in each of the large inverted repeats) of replication
origins implicated by both in vivo and in vitro analysis
(Kunnimalaiyaan and Nielsen, 1997a,b). However, there is
growing evidence that these are not the only replication
origins in ctDNA.
Support for the involvement of more than one mecha-
nism for ctDNA replication and/or additional replication
origins can be inferred from the results of the Koop
laboratory on insertions in the oriA and oriB replication
origins in tobacco ctDNA (Scharff and Koop, 2006,2007).
Targeted inactivation of either or both of these origins was
not lethal, although some deletions resulted in reduced
growth rate of the plants and reduced ctDNA copy number,
particularly later in leaf development (Scharff and Koop,
2007). This may suggest that the double D-loop mechanism
involving these origins is involved in ctDNA replication
during the transition from rapidly dividing cells to maturing
cells. Scharff and Koop (2006) reported the presence of a
significant amount of linear ctDNA molecules with defined
ends in tobacco, a substantial portion of which mapped to
previously reported ctDNA replication origins, and some of
which mapped to novel specific locations. Similar results
were reported earlier for maize ctDNA by Oldenberg and
Bendich (2004a). The Bendich laboratory reported that the
majority of ctDNA in maize is linear, and that the structure
and copy number of ctDNA molecules change during
development (Oldenburg and Bendich, 2004b). The earlier
published work of Kolodner and Tewari (1975) suggested
that a rolling circle replication initiation site may be present
at a different location from the D-loop origins in the
chloroplast genome. This rolling circle replication site
would be in the single-copy region in species with the
inverted repeat.
Another factor to consider in the control of ctDNA
maintenance is the presence of two nuclear-encoded DNA
polymerases that are both dual targeted to chloroplasts and
mitochondria (Christensen et al., 2005; Carrie et al., 2009).
These plant DNA polymerases share significant homology
with bacterial DNA polymerase I rather than with the
animal mitochondrial DNA polymerase c(Ono et al.,
2007). In Arabidopsis and tobacco the coding regions
for these genes are very highly conserved, suggesting that
they may be functionally redundant. However, the up-
stream promoter regions share no homology (BL Nielsen,
J Brammer, unpublished results), raising the possibility that
the two genes are differentially regulated and may be
expressed at different times and have different roles in
organelle DNA replication and maintenance. Indeed, our
preliminary data suggest that the two enzymes are not
expressed at equal levels or at the same time during plant
development (J Brammer, BL Nielsen, unpublished results).
One or both of these DNA polymerases would be essential
for any of the above-mentioned replication mechanisms.
Future work should examine the involvement of these two
DNA polymerases with cpRecA in the maintenance of
ctDNA.
An origin-binding protein or specificity factor, similar to
dnaA for the bacterial chromosome or rep proteins involved
in plasmid DNA replication initiation, or an enzyme that
nicks the DNA to initiate rolling circle replication, is also
likely to be required for ctDNA replication. While an
origin-binding activity has been characterized for ctDNA
replication in Chlamydomonas (Wu et al., 1989), which
shares some similarity with ctDNA replication in higher
plants, no such protein has been identified in higher plants.
Given the wide range in ctDNA levels in different tissues
during plant development, it seems clear that one or more
protein(s) involved in controlling initiation of ctDNA
replication must be present in plants.
From this new report and previous work from a number
of laboratories, there is strong support for the presence of
2536 |Commentary
at Serials Dept., Harold B. Lee Library, Brigham Young University on June 29, 2010 http://jxb.oxfordjournals.orgDownloaded from
multiple replication origins and/or replication mechanisms,
suggesting that the maintenance of ctDNA is more complex
than in bacteria, the endosymbiotic ancestor (Scharff and
Koop, 2007). There is growing evidence that more than one
mechanism is involved in replication of the chloroplast
genome, and each may function at different times during
chloroplast development and on the different forms of
ctDNA (linear and circular DNA). The majority of ctDNA
in most plant tissues is linear, with a varying proportion
of branched and/or circular molecules, providing support
for the premise that a recombination-mediated replication
mechanism may be involved in ctDNA replication, compat-
ible with the results reported by Rowan et al. (2010, this
issue). It is exciting that some new progress is being made
on ctDNA repair and replication, but there is still much to
be learned about the process of ctDNA maintenance in
plants during the various stages of plant development.
References
Carrie C, Kuhn K, Murcha MW, Duncan O, Small ID, O’Toole N,
Whelan J. 2009. Approaches to defining dual-targeted proteins in
Arabidopsis. The Plant Journal 57, 1128–1139.
Cerutti H, Osman M, Grandoni P, Jagendorf AT. 1992. A homolog
of Escherichia coli RecA protein in plastids of higher plants. Proceedings
of the National Academy of Sciences, USA 89, 8068–8072.
Chiu WL, Sears BB. 1992. Electron microscopic localization of
replication origins in Oenothera chloroplast DNA. Molecular and
General Genetics 232, 33–39.
Christensen AC, Lyznik A, Mohammed S, Elowsky CG, Elo A,
Yule R, Mackenzie S. 2005. Dual-domain, dual-targeting organellar
protein presequences in Arabidopsis can use non-AUG start codons.
The Plant Cell 17, 2805–2816.
Kolodner RD, Tewari KK. 1975. Chloroplast DNA from higher plants
replicates by both the Cairns and rolling circle mechanism. Nature
256, 708–711.
Kunnimalaiyaan M, Nielsen BL. 1997a. Fine mapping of replication
origins (oriA and oriB) in Nicotiana tabacum chloroplast DNA. Nucleic
Acids Research 25, 3681–3686.
Kunnimalaiyaan M, Nielsen BL. 1997b. Chloroplast DNA
replication: mechanism, enzymes and replication origins. Journal of
Plant Biochemistry and Biotechnology 6, 1–7.
Marechal A, Brisson N. 2010. Recombination and the
maintenance of plant organelle genome stability. New Phytologist
186, 299–317.
Mosig G. 1998. Recombination and recombination-dependent DNA
replication in bacteriophage T4. Annual Review of Genetics 32,
379–413.
Oldenburg DJ, Bendich AJ. 2004a. Most chloroplast DNA of maize
seedlings in linear molecules with defined ends and branched forms.
Journal of Molecular Biology 335, 953–970.
Oldenburg DJ, Bendich AJ. 2004b. Changes in the structure of
DNA molecules and theamount of DNA per plastid during
chloroplast development in maize. Journal of Molecular Biology 344,
1311–1330.
Ono N, Sakai A, Takechi K, Takio S, Takusagawa M, Takano H.
2007. NtPolI-like1 and NtPolI-like2, bacterial DNA polymerase I
homologs isolated from BY-2 cultured tobacco cells, encode DNA
polymerases engaged in DNA replication in both plastids and
mitochondria. Plant and Cell Physiology 48, 1679–1692.
Rowan BA, Oldenburg DJ, Bendich AJ. 2010. RecA maintains the
integrity of chloroplast DNA molecules in Arabidopsis.Journal of
Experimental Botany 61, 2575–2588.
Scharff LB, Koop HU. 2006. Linear molecules of tobacco ptDNA
end at known replication origins and additional loci. Plant Molecular
Biology 62, 611–621.
Scharff LB, Koop HU. 2007. Targeted inactivation of the tobacco
plastome origins of replication A and B. The Plant Journal 50,
782–794.
Shedge V, Arrieta-Montiel M, Christensen AC, Mackenzie SA.
2007. Plant mitochondrial recombination surveillance requires unusual
RecA and MutS homologs. The Plant Cell 19, 1251–1264.
Wu M, Nie ZQ, Yang J. 1989. The 18-kD protein that binds to
the chloroplast DNA replication origin is an iron-sulfur protein
related to a subunit of NADH dehydrogenase. The Plant Cell 1,
551–557.
Commentary |2537
at Serials Dept., Harold B. Lee Library, Brigham Young University on June 29, 2010 http://jxb.oxfordjournals.orgDownloaded from
... Chloroplasts utilize a double displacement loop strategy to initiate DNA replication [60]. The two displacement loops begin on opposite strands and begin replicating unidirectionally towards each other until they join to create a bidirectional replication bubble [61,62]. At this point, the displacement loops fuse, forming a Cairns or theta structure and DNA replication continues bidirectionally until two daughter molecules are created. ...
... At this point, the displacement loops fuse, forming a Cairns or theta structure and DNA replication continues bidirectionally until two daughter molecules are created. Rolling circle and recombination-dependent replication have also been proposed for cpDNA [24,61,62]. MOC1 has been identified as a Holliday junction (recombination intermediate) resolvase that mediates chloroplast nucleoid segregation [63]. ...
... also been proposed for cpDNA [24,61,62]. MOC1 has been identified as a Holliday junction (recombination intermediate) resolvase that mediates chloroplast nucleoid segregation [63]. ...
Article
Full-text available
Mitochondria and chloroplasts perform essential functions in respiration, ATP production, and photosynthesis, and both organelles contain genomes that encode only some of the proteins that are required for these functions. The proteins and mechanisms for organelle DNA replication are very similar to bacterial or phage systems. The minimal replisome may consist of DNA polymerase, a primase/helicase, and a single-stranded DNA binding protein (SSB), similar to that found in bacteriophage T7. In Arabidopsis, there are two genes for organellar DNA polymerases and multiple potential genes for SSB, but there is only one known primase/helicase protein to date. Genome copy number varies widely between type and age of plant tissues. Replication mechanisms are only poorly understood at present, and may involve multiple processes, including recombination-dependent replication (RDR) in plant mitochondria and perhaps also in chloroplasts. There are still important questions remaining as to how the genomes are maintained in new organelles, and how genome copy number is determined. This review summarizes our current understanding of these processes.
... Regions of the plastid genome that best supported DNA synthesis in vitro were designated as replication origins (oris), and led to the assignment of two major oris (known as oriA and oriB) in Oenothera, tobacco, and pea (Heinhorst and Cannon, 1993;Kunnimalaiyaan and Nielsen, 1997). Sequences similar to those of oriA and oriB have been identified in the plastid genomes of many plants (Oldenburg and Bendich, 2004a;Shaver et al., 2008;Krishnan and Rao, 2009). ...
... Three types of replication mechanism have been proposed for ptDNA: theta replication, rolling circle replication (RCR), and recombination-dependent replication (RDR; Kunnimalaiyaan and Nielsen, 1997;Marechal and Brisson, 2010). Although circular ptDNA molecules were reported for chloroplasts from entire light-grown shoots of several plants (Kolodner and Tewari, 1972b;Lamppa and Bendich, 1979b;Bendich and Smith, 1990;Lilly et al., 2001), support for the theta and RCR mechanisms would seem to require the presence of circular ptDNA molecules in meristematic tissues. ...
Article
Full-text available
The DNA molecules in plastids and mitochondria of plants have been studied for over 40 years. Here, we review the data on the circular or linear form, replication, repair, and persistence of the organellar DNA (orgDNA) in plants. The bacterial origin of orgDNA appears to have profoundly influenced ideas about the properties of chromosomal DNA molecules in these organelles to the point of dismissing data inconsistent with ideas from the 1970s. When found at all, circular genome-sized molecules comprise a few percent of orgDNA. In cells active in orgDNA replication, most orgDNA is found as linear and branched-linear forms larger than the size of the genome, likely a consequence of a virus-like DNA replication mechanism. In contrast to the stable chromosomal DNA molecules in bacteria and the plant nucleus, the molecular integrity of orgDNA declines during leaf development at a rate that varies among plant species. This decline is attributed to degradation of damaged-but-not-repaired molecules, with a proposed repair cost-saving benefit most evident in grasses. All orgDNA maintenance activities are proposed to occur on the nucleoid tethered to organellar membranes by developmentally-regulated proteins.
... Arabidopsis chloroplast genome also consists of 29 genes, including 4 genes for chloroplast RNA polymerase subunits and 25 genes for components of the ribosome (Sato et al., 1999). Chloroplasts have been shown to utilize a double displacement loop strategy to initiate their DNA replication (Kunnimalaiyaan and Nielsen, 1997). In addition, the rolling circle and recombinant-dependent replication (RDR) process have also been proposed for cp-DNA replication. ...
Article
Full-text available
Besides the nuclear genome, plants possess two small extra chromosomal genomes in mitochondria and chloroplast, respectively, which contribute a small fraction of the organelles’ proteome. Both mitochondrial and chloroplast DNA have originated endosymbiotically and most of their prokaryotic genes were either lost or transferred to the nuclear genome through endosymbiotic gene transfer during the course of evolution. Due to their immobile nature, plant nuclear and organellar genomes face continuous threat from diverse exogenous agents as well as some reactive by-products or intermediates released from various endogenous metabolic pathways. These factors eventually affect the overall plant growth and development and finally productivity. The detailed mechanism of DNA damage response and repair following accumulation of various forms of DNA lesions, including single and double-strand breaks (SSBs and DSBs) have been well documented for the nuclear genome and now it has been extended to the organelles also. Recently, it has been shown that both mitochondria and chloroplast possess a counterpart of most of the nuclear DNA damage repair pathways and share remarkable similarities with different damage repair proteins present in the nucleus. Among various repair pathways, homologous recombination (HR) is crucial for the repair as well as the evolution of organellar genomes. Along with the repair pathways, various other factors, such as the MSH1 and WHIRLY family proteins, WHY1, WHY2, and WHY3 are also known to be involved in maintaining low mutation rates and structural integrity of mitochondrial and chloroplast genome. SOG1, the central regulator in DNA damage response in plants, has also been found to mediate endoreduplication and cell-cycle progression through chloroplast to nucleus retrograde signaling in response to chloroplast genome instability. Various proteins associated with the maintenance of genome stability are targeted to both nuclear and organellar compartments, establishing communication between organelles as well as organelles and nucleus. Therefore, understanding the mechanism of DNA damage repair and inter compartmental crosstalk mechanism in various sub-cellular organelles following induction of DNA damage and identification of key components of such signaling cascades may eventually be translated into strategies for crop improvement under abiotic and genotoxic stress conditions. This review mainly highlights the current understanding as well as the importance of different aspects of organelle genome maintenance mechanisms in higher plants.
... El número de plastomas por plastidio, y el número de plastidios por célula vegetal son muy variables. En células embriogénicas hay entre 10-30 copias del plastoma, mientras que en cloroplastos desarrollados hay 200-300, y entre 20-60 plastos por célula, lo que hace que haya entre 10.000 y 12.000 copias de ADN por célula (Kunnimalaiyaan y Nielsen, 1997). El elevado número de copias del plastoma varía entre especies, así, en remolacha puede haber 1900 copias por célula y en trigo puede llegar a 50.000, lo que supone que el contenido de ADN plastidial sea del 10-20% del contenido total de ADN de una célula vegetal (Bendich, 1987). ...
Thesis
Maize is one of the most important crops for the nutritional intake of humans, being such via direct utilization as food or through livestock breeding. The constant growth of human population, the increasing meat consumption in emerging economy countries, the use of grain for biofuel production or the climate change are demanding an increase in the productivity of this crop. Plant biotechnology offers diverse procedures to achieve this objective, being plastid transformation one of them. This technique brings important advantages to plant transformation such as high expression level of the gene of interest or prevent the transgene transmission through pollen to other crops due to the maternal inheritance of plastids. Achieving plastid transformation of maize would allow the transfer of those advantages to this important crop. We have developed diverse in vitro culture procedures and worked with different maize genotypes to select the best candidates to be plastidially transformed. The best protocol tested has been the regeneration through somatic embryogenesis induced from immature embryos. We also obtained a green calli culture procedure using the peptide hormone -PSK, designed a second round of regeneration on selective media using mature somatic embryos and a protocol for culturing callus cell aggregates on liquid media as an alternative to cell suspensions. The three selection schemes tested were antibiotic based: the gene aadA gene and streptomycin, the nptII gene and kanamycin and the Aph(4) gene and hygromycin selection. Those genes were cloned in four different plastid transformation vectors which introduced the transgenes into the following regions of the inverted repeat regions of the plastome: 16SRNAr-trnV-ORF85/ORF58 or 16SRNAr-trnI/trnA-23SRNAr. A total of 54 transformation experiments were made with 12 different types of materials and 1738 maize regenerants were obtained, 7 of which were transplastomic regenerants with a high level of heteroplasmy. Six of them were obtained in streptomycin selection and 1 with hygromycin. It was also designed an alternative selection scheme which was based on the complementation of 3 nuclear photosynthetic albino mutants: csr1-1, crs1-2 y crs2-2. We obtained embryogenic calli from those mutants cultured in light or dark conditions from immature embryos. The mutated gene was introduced in a plastid transformation vector which introduces the gene into the 16SRNAr-trnI/trnA-23SRNAr recombination zone. No regenerants were obtained from 6 transformation experiments.
... The ptDNA origins of replication (oris) have been mapped for several plant species and two closely spaced oris are often found. For maize ptDNA three origins had been identified: oriA and oriB by sequence similarity to Oenothera ptDNA; and the third, designated here as oriC, by similarity to Chlamydomonas reinhardtii (Heinhorst and Cannon 1993;Kunnimalaiyaan and Nielsen 1997;Oldenburg and Bendich 2004a). The size of the plastid genome in maize, Zea mays L., is 140,384 bp (Maier et al. 1995), although the extracted ptDNA is found in various structural forms, including branched-linear multigenomic molecules, unit-genomesized linear isomers and head-to-tail concatemers, and less-than-genome-sized fragments (Oldenburg and Bendich 2004a). ...
Article
Full-text available
The structure of a chromosomal DNA molecule may influence the way in which it is replicated and inherited. For decades plastid DNA (ptDNA) was believed to be circular, with breakage invoked to explain linear forms found upon extraction from the cell. Recent evidence indicates that ptDNA in vivo consists of linear molecules with discrete termini, although these ends were not characterized. We report the sequences of two terminal regions, End1 and End2, for maize (Zea mays L.) ptDNA. We describe structural features of these terminal regions and similarities found in other plant ptDNAs. The terminal sequences are within inverted repeat regions (leading to four genomic isomers) and adjacent to origins of replication. Conceptually, stem-loop structures may be formed following melting of the double-stranded DNA ends. Exonuclease digestion indicates that the ends in maize are unobstructed, but tobacco (Nicotiana tabacum L.) ends may have a 5′-protein. If the terminal structure of ptDNA molecules influences the retention of ptDNA, the unprotected molecular ends in mature leaves of maize may be more susceptible to degradation in vivo than the protected ends in tobacco. The terminal sequences and cumulative GC skew profiles are nearly identical for maize, wheat (Triticum aestivum L.) and rice (Oryza sativa L.), with less similarity among other plants. The linear structure is now confirmed for maize ptDNA and inferred for other plants and suggests a virus-like recombination-dependent replication mechanism for ptDNA. Plastid transformation vectors containing the terminal sequences may increase the chances of success in generating transplastomic cereals.
... The plastome most probably exists as circular chromosomes (Krause, 2008;Wakasugi et al., 2001) and is replicated on the whole (Heinhorst und Cannon, 1993;Kunnimalaiyaan und Nielsen, 1997). It is thus to be expected that copy numbers of different genes do not differ from each other, because they should be replicated equally. ...
Thesis
Zusätzlich zu der eubakteriellen RNA-Polymerase (RNAP) der Plastiden sind im Zellkern von Arabidopsis thaliana drei weitere, phagentypische RNAP kodiert, die jeweils aus nur einer Einheit aufgebaut sind. Die Enzyme RpoTp und RpoTm werden in die Plastiden, bzw. die Mitochondrien transportiert, während RpoTmp in beiden Organellen zu finden ist. Um die Lichtabhängigkeit der RpoT-Gene zu untersuchen, wurde die lichtinduzierte Akkumulation ihrer Transkripte in 7-Tage alten Keimlingen, sowie 3- bzw. 9-Wochen alten Rosettenblättern mittels quantitativer real-time PCR ermittelt. Die entwicklungsabhängige Regulation der RpoT-Transkript-Akkumulation wurde außerdem während der Blattentwicklung analysiert. Zusätzlich wurde der Einfluss des circadianen Rhythmus untersucht. Es stellte sich heraus, dass die Transkriptakkumulation aller drei RpoT-Gene stark lichtinduziert war und nur marginalen circadianen Schwankungen unterlag. In weiteren Versuchen mit verschiedenen Lichtrezeptor-Mutanten und unterschiedlichen Lichtqualitäten wurde der Einfluss multipler Rezeptoren auf den Prozess der Lichtinduktion gezeigt. In den Zellen höherer Pflanzen finden sich drei Genome. Die Biogenese von Chloroplasten und Mitochondrien, sowie lebenswichtige Prozesse, wie Atmung und Photosynthese setzen oftmals die Aktivität von Genen auf mindestens zwei dieser Genome voraus. Eine intrazelluläre Kommunikation zwischen den verschiedenen Genomen ist daher unumgänglich für einen funktionierenden Stoffwechsel der Pflanze. In dieser Arbeit wurde herausgestellt, dass die Zahl mitochondrialer Genkopien in photosynthetisch inaktiven Arabidopsis-Keimlingen drastisch erhöht ist. Bei der Untersuchung des DNA-Gehaltes in Proben, die Altersstufen von 2-Tage alten Keimblättern bis hin zu 37-Tage alten, seneszenten Rosettenblättern umfassten, fand sich ein deutlicher Anstieg der Kopienzahlen in älteren Rosettenblättern. Außerdem unterschieden sich die Kopienzahlen der untersuchten Gene zum Teil erheblich voneinander.
Article
Full-text available
Chloroplast genetic engineering offers a number of unique advantages, including high-level trans-gene expression, multigene engineering in a single transformation event, transgene containment via maternal inheritance, lack of gene silencing, position and pleiotropic effects and undesirable foreign DNA. More than 40 transgenes have been stably integrated and expressed via the tobacco chloro-plast genome to confer desired agronomic traits or express high levels of vaccine antigens and biopharmaceuticals. Despite such significant progress, this technology has not been extended to other important plant species. For example, Arabidopsis may be an ideal model system for chloro-plast functional genomics. The employment of chloroplast transformation technology in Arabidopsis has been hampered by the lack of an efficient and reproducible protocol that provides fertile chloro-plast transgenic plants. Transformation of the Arabidopsis chloroplast genome was achieved via organogenesis but the efficiency was at least a 100-fold lower than in tobacco and had the drawback of polyploidy in the leaf tissue that resulted in sterile transgenic plants. This problem can be overcome by adapting procedures that are now available to regenerate plants from both diploid and tetraploid explants via callus. In addition, it is feasible to regenerate Arabidopsis via somatic embryogenesis. Recent breakthroughs in highly efficient plastid transformation of recalcitrant crops such as cotton and soybean have opened the possibility of engineering Arabidopsis plastid genome via somatic embryogenesis. Therefore, protocols of recent improvements in tissue culture, DNA delivery, and the novel vector designs are provided here in order to achieve highly efficient plastid transformation in Arabidopsis.
Article
Full-text available
Chloroplast genome copy number is very high in leaf tissue, with upwards of 10,000 or more copies of the chloroplast DNA (ctDNA) per leaf cell. This is often promoted as a major advantage for engineering the plastid genome, as it provides high gene copy number and thus is expected to result in high expression of foreign proteins from integrated genes. However, it is also known that ctDNA copy number and ctDNA integrity decrease as cells age. Quantitative PCR (qPCR) allows measurement of organelle DNA levels relative to a nuclear gene target. We have used this approach to determine changes in copy number of ctDNA relative to the nuclear genome at different ages of Arabidopsis plant growth and in organellar DNA polymerase mutants. The mutant plant lines have T-DNA insertions in genes encoding the two organelle localized DNA polymerases (PolIA and PolIB). Each of these mutant lines exhibits some delay in plant growth and development as compared to wild-type plants, with the PolIB plants having a more pronounced delay. Both mutant lines develop to maturity and produce viable seeds. Mutants for both proteins were observed to have a reduction in ctDNA and mtDNA copy number relative to wild type plants at all time points as measured by qPCR. Both DNA polymerase mutants had a fairly similar decrease in ctDNA copy number, while the PolIB mutant had a greater effect of reduction in mtDNA levels. However, despite similar decreases in genome copy number, RT-PCR analysis of PolIA mutants show that PolIB expression remains unchanged, suggesting that PolIA may not be essential to plant survival. Furthermore, genotypic analysis of plants from heterozygous parents display a strong pressure to maintain two functioning copies of PolIB. These results indicate that the two DNA polymerases are both important in ctDNA replication, and they are not fully redundant to each other, suggesting each has a specific function in plant organelles.
Chapter
We review current progress in our understanding of chloroplast genomes (plastomes) of liverworts, mosses, hornworts, lycophytes and monilophytes. We briefly cover some of the methods used to obtain complete nucleotide sequences of plastomes and we summarize the published sequences from the plant groups above. We explore some of the evolutionary changes that have occurred in terms of gene content, introns and position of the inverted repeat boundaries. We also discuss RNA editing, which is especially high in plastome genes of some non-seed land plants. We finish with a phylogenetic analysis of available plastome genes and we suggest some possible directions for future research.
Chapter
Most living plant cells contain plastids, which harbour their own DNA: the plastome. Plastomes are significantly less diverse than nuclear genomes, but this lower diversity has the advantage that comparisons can be made across all clades of green plants. Large blocks of synteny can be found among plastomes from different algal lineages including land plants. In this chapter, I review the nature of plastome diversity, starting with some history of research and what is known about packaging, replication, and inheritance of plastid DNA. I review the composition of plastomes, the current status of available data (and future needs), and then examine plastome origins, structural evolution, patterns of nucleotide substitution, gene expression (but not translation or post-translational stages), and plastome biotechnology. The goal is a broad overview, but this is at the expense of important detail to which I defer to cited literature.
Article
Full-text available
A DNA-relaxing enzyme was purified 5 000-fold to homogeneity from isolated chloroplasts of Pisum sativum. The enzyme consists of a single polypeptide of 112 kDa. The enzyme was able to relax negatively supercoiled DNA in the absence of ATP. It is resistant to nalidixic acid and novobiocin, and causes a unit change in the linkage number of supercoiled DNA. The enzyme shows optimum activity at 37°C with 50 mM KCl and 10 mM MgCl2. From these properties, the enzyme can be classified as a prokaryotic type I topoisomerase. Using a partiall purified pea chloroplast DNA polymerase fraction devoid of topoisomerase I activity for in vitro replication on clones containing the pea chloroplast DNA origins of replication, a 2–6-fold stimulation of replication activity was obtained when the purified topoisomerase I was added to the reaction at 70–100 mM KCl. However, when the same reaction was carried out at 125 mM KCl, which does not affect DNA polymerase activity on calf thymus DNA but is completely inhibitory for topoisomerase I activity, a 4-fold drop in activity resulted. Novobiocin, an inhibitor of topoisomerase II, was not found to inhibit the in vitro replication of chloroplast DNA.
Article
Full-text available
Three distinct chloroplast (cp) DNA fragments from Petunia hybrida, which promote autonomous replication in yeast, were mapped on the chloroplast genome. Sequence analysis revealed that these fragments (called ARS A, B and C) have a high AT content, numerous short direct and inverted repeats and at least one yeast ARS consensus sequence 5A/TTTTATPuTTTA/T, essential for yeast ARS activity. ARS A and B also showed the presence of (semi-)conserved sequences, present in all Chlamydomanas reinhardii cpDNA regions that promote autonomous replication in yeast (ARS sequences) or in C. reinhardii (ARC sequences). A 431 bp BamHI/EcoRI fragment, close to one of the inverted repeats and adjacent to the ARS B subfragment contains an AT-rich stretch of about 100 nucleotides that show extensive homology with an Euglena gracilis cpDNA fragment which is part of the replication origin region. This conserved region contains direct and inverted repeats, stem-and-loop structures can be folded and it contains an ARS consensus sequence. In the near vicinity a GC-rich block is present. All these features make this cpDNA region the best candidate for being the origin of replication of P. hybrida cpDNA.
Chapter
Plants are unique among eukaryotes in possessing two DNA-containing organelles—the plastid and the mitochondrion. Moreover, the green alga Chlamydomonas reinhardtii has recently been shown to contain a third extranuclear genome—that of the basal body (Hall et al., 1989). Nothing is known about the origin, phylogenetic distribution and evolution of basal body DNA, and therefore this genome will not be considered in this chapter. In contrast, we now possess a rather detailed picture of the tempo and mode of evolution of chloroplast DNA (cpDNA) and mitochondrial DNA (mtDNA) in land plants. Review of this topic will form the heart of this chapter, as presented in Sects. III–V. Data for both genomes will be presented in an integrated format in order to highlight the striking contrasts in their evolution in land plants. The much more limited evolutionary data base available for algal organelle genomes will be discussed in Sect. VI. All plastid and mitochondrial genomes are of endosymbiotic, bacterial origin. However, as discussed in the next section, considerable uncertainty remains as to the precise number and nature of endosymbioses that have taken place.
Chapter
Plastids are organelles found in all eukaryotic plant cells and not in the cells of animals or fungi. They are bound by an envelope consisting of two membranes and contain a variety of internal structures. Plastids cannot arise de novo; they are formed by the division of existing plastids. Plastid division by constriction-fission is common to all plant groups. In the majority of higher plants, a range of plastid types develops from smaller undifferentiated proplastids found in the meristematic tissue of the shoot and root and in embryonic tissue. The chapter provides the evidence of control associated with cell division and also the evidence of control during cell differentiation in the leaf. It also discusses the effects of external factors—such as light, temperature, ionizing radiation, chemical inhibitors, and nutrition—on plastid division. The relationships of chloroplast number per cell in discs cultured at different light intensities suggest that there is an energy requirement for at least chloroplast division. Chloroplast development requires the translation of plastid DNA, and functional chloroplast protein complexes are formed only after nuclear-coded gene products have combined with those of the plastid.
Article
We have examined the effects of novobiocin and template topology on the transcription of two chloroplast genes encoding the large subunit of ribulose 1,5-bisphosphate carboxylase (rbcL) and the beta subunit of the chloroplast ATPase (atpB), in an in vitro transcription system. The template topology was monitored by agarose gel electrophoresis while the in vitro transcripts were determined by 5' S1 nuclease analysis under identical conditions. We discovered that our chloroplast transcription extracts contain a DNA gyrase activity and a chromatographically separable topoisomerase I activity. Incubation of a supercoiled template with the extracts under the same conditions in which transcription assays were carried out leads to a decrease in the supercoiled from and concomitant appearance of distinct topoisomers. More extensive relaxation of the supercoiled template occurs when nucleotide triphosphates are omitted from the reaction mixture or when a low concentration (25 μg/ml) of novobiocin is added. Higher concentrations (≥ 250 μg/ml) of the drug, however, also inhibit the topoisomerase I activity. The transcription of the atpB gene is inhibited by lower concentrations of novobiocin as compared to the rbcL gene in the same reaction mixture. Relaxed, closed circular template and linearized DNA are not substrates for chloroplast transcription extracts, although they are transcribed accurately by the E. coli RNA polymerase under our conditions. Control of in vitro transcription of the two chloroplast genes by template topology can also be demonstrated by modulating the relative activity for the topoisomerases in the transcription extract. Our results suggest that changes in template topology may be a mechanism by which chloroplast genes are differentially regulated and the chloroplast DNA gyrase and topoisomerase I are key enzymes for this mode of regulation in vivo.
Article
The relationship between nuclear and plastid DNA synthesis in cultured tobacco cells was measured by following3H-thymidine incorporation into total cellular DNA in the absence or presence of specific inhibitors. Plastid DNA synthesis was determined by hybridization of total radiolabeled cellular DNA to cloned chloroplast DNA. Cycloheximide, an inhibitor of nuclear encoded cytoplasmic protein synthesis, caused a rapid and severe inhibition of nuclear DNA synthesis and a delayed inhibition of plastid DNA synthesis. By contrast, chloramphenicol which only inhibits plastid and mitochondrial protein production, shows little inhibition of either nuclear or plastid DNA synthesis even after 24 h of exposure to the cells. The inhibition of nuclear DNA synthesis by aphidicolin, which specifically blocks the nuclear DNA polymeraseα, has no significant effect on plastid DNA formation. Conversely, the restraint of plastid DNA synthesis exerted by low levels of ethidium bromide has no effect on nuclear DNA synthesis. These results show that the synthesis of plastid and nuclear DNA are not coupled to one another. However, both genomes require the formation of cytoplasmic proteins for their replication, though our data suggest that different proteins regulate the biosynthesis of nuclear and plastid DNA.
Article
Chloroplasts contain their own autonomously replicating DNA genome. The majority of proteins present in the chloroplasts are encoded by nuclear DNA, but the rest are encoded by chloroplast DNA and synthesized by the chloroplast transcription–translation machinery1–4. Although the nucleotide sequences of many chloroplast genes from various plant species have been determined, the entire gene organization of the chloroplast genome has not yet been elucidated for any species of plants. To improve our understanding of the chloroplast gene system, we have determined the complete sequence of the chloroplast DNA from a liverwort, Marchantia polymorpha, and deduced the gene organization. As reported here the liverwort chloroplast DNA contains 121,024 base pairs (bp), consisting of a set of large inverted repeats (IRA and IRB, each of 10,058 bp) separated by a small single-copy region (SSC, 19,813 bp) and a large single-copy region (LSC, 81,095 bp). We detected 128 possible genes throughout the liverwort chloroplast genome, including coding sequences for four kinds of ribosomal RNAs, 32 species of transfer RNAs and 55 identified open reading frames (ORFs) for proteins, which are separated by short A+T-rich spacers (Fig. 1). Twenty genes (8 encoding tRNAs, 12 encoding proteins) contain introns in their coding sequences. These introns can be classified as belonging to either group I or group II, as described for mitochondria5. Interestingly, seven of the identified ORFs show high homology to unidentified reading frames (URFs) found in human mitochondria6,7.
Article
Chloroplast DNA (ctDNA) generally occurs as circular molecules with molecular weights (MWs) in the range 70–130 × 106 depending on the species1,2. In Euglena gracilis, ctDNA (44 µm, 92 × 106 MW3) replicates through Cairns-type intermediates4 having structural aspects suggesting bidirectional replication5. Pea and corn ctDNA were shown to contain two displacement loops (D-loops) located 7,100 base pairs (bp) apart6. The displacing strands of the two D-loops are located on opposite strands of the parental DNA; they expand towards each other and form a Cairns replicative intermediate when the two strands elongate past each other. Rolling circle intermediates7, apparently resulting from the continuation of Cairns rounds of replication, have also been observed8. The origins of replication were not located on physical maps for any of these studies. We present here the results of an electron microscopic (EM) study indicating that replication of ctDNA in Euglena is initiated near the 5′ end of the supplementary 16S ribosomal RNA gene9.
Article
A 5.3 kb chloroplast restriction fragment of Chlamydomonas reinhardii containing an origin of DNA replication and a sequence capable of promoting autonomous replication in C. reinhardii (ARC sequence) also carries an ARS sequence (autonomous replication in yeast). The ARC and ARS elements have been physically mapped and shown to be distinct from the origin of DNA replication. Similarly, restriction fragments containing the origin of chloroplast DNA replication from Euglena gracilis are unable to promote autonomous replication in yeast.