Vous êtes sur la page 1sur 5

A Tertiary Plastid Gains RNA Editing in Its New Host

Christopher J. Jackson,z Sebastian G. Gornik, and Ross F. Waller*


School of Botany, University of Melbourne, Victoria, Australia
z
Present address: Biology Department, University of New Brunswick, New Brunswick, Canada
*Corresponding author: E-mail: r.waller@unimelb.edu.au.
Associate editor: Hervé Philippe
New sequences are available in GenBank (accession nos. JX292160–JX292162, JX292164–JX292175, and JX945981–JX945986) or in
the supplementary data, Supplementary Material online.

Abstract
Dinoflagellates are known for their development of highly aberrant organelle genetic systems. Both their plastid and
mitochondrial genomes are extremely reduced in gene number and rearranged into numerous unconventional genomic
elements. Transcription processes are also elaborately modified including extensive RNA editing and trans-splicing. Some

Downloaded from http://mbe.oxfordjournals.org/ at Universidade Federal de Viçosa on July 17, 2013


dinoflagellates have replaced their original plastid through serial endosymbiotic events. Karlodinium veneficum is such an
example that now contains a haptophyte plastid. This tertiary plastid provides a case of a more conventional genetic
system introduced into a cellular environment with a known penchant for genetic oddities. Here, we show that
K. veneficum plastid transcripts undergo extensive substitutional editing. The substitution types are more diverse than
those seen in most other plastids but are similar to those of dinoflagellate organelles. There is no evidence for RNA editing
of plastid-encoded transcripts from extant haptophytes, suggesting that K. veneficum plastid editing developed after the
uptake of the tertiary endosymbiont.
Key words: RNA editing, plastid, endosymbiosis, dinoflagellate.

The development of endosymbiotic organelles requires two replaced their peridinin plastid by a subsequent, or serial,
cells to learn to live together. Plastids and mitochondria owe endosymbiosis (Archibald 2009). The sources of these new
their existence to bacteria that became resident inside a host plastids have been diverse, including from haptophytes, dia-
eukaryote. The development of these organelles required toms, cryptomonads (all resulting in tertiary plastids), and
metabolic integration of the two cells, as well as molecular green algae (serial secondary plastids) (Archibald 2009).
harmonization of their genetic functions. Typically, the endo- Dinoflagellates, therefore, present multiple opportunities for
symbiont surrenders much of its genetic content to the host examining the process of host–endosymbiont cell
cell nucleus but still maintains a reduced genome expressed integration.
semiautonomously using retained bacterial expression ma- The best studied of these recent endosymbioses in
Letter

chinery (Gould et al. 2008; Barbrook et al. 2010). All mito- dinoflagellates is the gain of a haptophyte to create the
chondria are derived from a single ancient endosymbiotic tertiary plastid found in three dinoflagellate genera,
event, whereas plastids have migrated horizontally into new Karlodinium, Karenia, and Takayama. The origin of this
hosts through several subsequent endosymbioses. These plas- new endosymbiont is clear as it contains the distinctive
tid transfers occur when a eukaryote, already bearing a pri- haptophyte accessory pigments (190 -hexanoyloxy-
0
mary plastid, becomes an endosymbiont itself in a new host. fucoxanthin and 19 -butanoyloxy-fucoxanthin) in place of
Plastids derived this way are known as secondary plastids or peridinin and phylogenies using plastid genes strongly
tertiary plastids if the endosymbiont already bore a secondary group it with haptophytes (Tengs et al. 2000; Ishida and
plastid (Archibald 2009). With each new endosymbiosis, a Green 2002; Patron et al. 2006). This endosymbiont is now
new round of molecular harmonization must commence be- well integrated into its dinoflagellate host, with the endosym-
tween the host cell and symbiont. biont nucleus eliminated and only the photosynthetic plastid
Dinoflagellate algae are unusual among eukaryotes in that retained, surrounded by 3–4 membranes (Dodge 1989). Most
this phylum has gained plastid endosymbionts multiple dif- endosymbiont genes are relocated to the host nucleus and,
ferent times. The ancestor of dinoflagellates and their sister therefore, are under the direct control of the new host (Ishida
phylum, Apicomplexa, most likely inherited a common plas- and Green 2002; Takishita et al. 2004; Nosenko et al. 2006;
tid originally derived from a red algal endosymbiont Patron et al. 2006; Shalchian-Tabrizi et al. 2006; Gabrielsen
(Archibald 2009; Janouškovec et al. 2010). In photosynthetic et al. 2011). In Karlodinium veneficum, only 70 protein
dinoflagellates (approximately half have now lost photosyn- genes are retained in the plastid, encoded on the plastid
thesis), this ancestral plastid contains the distinctive pigment genome inherited from the haptophyte (Gabrielsen et al.
peridinin. Four dinoflagellate groups have more recently 2011).

ß The Author 2012. Published by Oxford University Press on behalf of the Society for Molecular Biology and Evolution. All rights reserved. For permissions, please
e-mail: journals.permissions@oup.com

Mol. Biol. Evol. 30(4):788–792 doi:10.1093/molbev/mss270 Advance Access publication November 28, 2012 788
A Tertiary Plastid Gains RNA Editing in Its New Host . doi:10.1093/molbev/mss270 MBE
In this study, we present evidence that in K. veneficum, rpoB, and rpoC2 compared with E. huxleyi proteins affect
following endosymbiosis within the dinoflagellate host, plas- function in these cases). A second consequence of editing is
tid genetic processes have been re-engineered with the ac- a net increase in the A + T nucleotide content of plastid genes
quisition of an RNA editing system. From the recent 454 compared with transcripts, and this occurs predominantly at
assembly of the K. veneficum plastid genome, many gene codon position 1 then position 2 (table 2). The strong overall
sequences were noted as being unusual, including unex- A + T bias of this genome (72.9%, Gabrielsen et al. 2011) is
pected sequence divergence, the presence of canonical stop greatest at codon position 3 as this position least often con-
codons interrupting genes, and five genes with apparent strains the amino acid specified. Hence, the predominance of
frameshifts (Gabrielsen et al. 2011). To examine transcrip- A to G and U to C editing at positions 1 and 2 has allowed
tional processes in this endosymbiont, we generated gene further A + T drift to occur at these more limiting positions.
and transcript sequences corresponding to 14 plastid- Together these data indicate that RNA editing has allowed a
encoded genes (psaA, psbC, rbcL, atpG, atpI, rpl5, rps13, more extreme A + T rich genome to develop, while amelio-
rps18, rpl33, rpl36, petD, rpoB, rpoC2, and secY—totaling rating potential negative impacts to protein sequence and
7,373 nucleotides) including all five purported instances of function.
frameshifts, as well as cases of internal stop codons. We used a RNA editing in the K. veneficum plastid presents a dramatic

Downloaded from http://mbe.oxfordjournals.org/ at Universidade Federal de Viçosa on July 17, 2013


proof reading polymerase (Platinum Taq DNA Polymerase change from a nonediting to editing transcriptional system.
High Fidelity) and generated multiple independent amplicons So how and why has this system of RNA editing been
of each to verify sequencing fidelity (primer binding sites were acquired? To address how an editing system might have
excluded from sequence data). These data revealed extensive evolved requires knowledge of the molecular machinery
substitutional RNA editing occurring in the K. veneficum plas- used for editing. Throughout eukaryotes, RNA editing has
tid consisting exclusively of the four transitions, predomin- evolved many times, is most common in organelles, and is
antly A to G and U to C (table 1 and fig. 1). Editing affects on most often limited to one or two substitution types (Knoop
average 3% of nucleotides (up to 8.6% for some genes), is 2011; Gray 2012). However, the biochemistry of few substitu-
heavily biased for codon positions 1 and 2 (93%), and tional editing systems is well understood. The best known is
almost always results in nonsynonymous changes (95%). All from plant organelles, where a family of sequence-specific
cases of internal stop codons examined (in psaA, rps13, rpoB, RNA-binding proteins is thought to direct cyti-
and secY) were edited to restore open reading frames (e.g., dine deaminases that catalyze the base conversion C to U,
fig. 1). We found no evidence of insertional/deletional editing the major editing event in plants (Salone et al. 2007;
based on comparison of gene to transcript sequences. In four Chateigner-Boutin and Small 2010). Karlodinium veneficum
of the five genes reported by Gabrielsen et al. (2011) to con- plastid RNA editing consists of four substitution types, imply-
tain frameshifts (rpl5, rpoB, rpoC2, and secY), our sequences ing a more complex machinery than this. In other dinofla-
lacked the frameshifts, indicating 454 sequencing errors (that gellate organelles, even more extensive and diverse
often occurred in regions of homopolymers). In the fifth case substitutional RNA editing types occur, including transver-
(petD), transcript data suggest that this gene is simply shor- sions. Dinoflagellate mitochondrial editing is best docu-
tened by 18 nucleotides by a new but genuine stop codon. mented; it occurs widely in this phylum, 11 of a possible 12
Our data, therefore, suggest that no frameshifts are present in substitution types are known to occur and editing densities of
K. veneficum plastid genes. In free-living haptophytes, no re- up to 6% are seen (cox3, K. veneficum) (Lin et al. 2008; Waller
ports of RNA editing have been made, and we independently and Jackson 2009; Jackson et al. 2012). Fewer reports of editing
generated genomic DNA (gDNA) and cDNA sequences for in the peridinin plastid have been made, but from
three plastid genes (psaA, psbC, and rbcL) from the hapto- Heterocapsa triquetra and Lingulodinium polyedrum, up to
phyte Emiliania huxleyi and found no evidence of editing eight different substitution types have been reported
(70 edits occur in the corresponding K. veneficum genes). (Wang and Morse 2006; Dang and Green 2009). Although
This suggests that RNA editing in the K. veneficum plastid the mechanism of editing in dinoflagellate organelles is en-
was acquired after endosymbiosis in the dinoflagellate host. tirely uncharacterized, enzymatic base conversion is unable to
One effect of RNA editing in the K. veneficum plastid is that account for these more diverse types of transitions and trans-
transcript-encoded proteins share greater identity to homo- versions, and a nucleotide excision and replacement mech-
logs in related organisms than do the direct translations of the anism is envisaged for such editing systems.
gene sequences. We compared conceptual translations of It might be that the K. veneficum plastid has independently
K. veneficum gDNA and cDNA sequences to proteins from developed editing machinery de novo. However, the relative
haptophyte E. huxleyi (no other complete gene data set for a complexity of this editing system (four substitution types),
haptophyte or dinoflagellate peridinin plastid exists to enable and the presence of complex substitutional editing in both
broader comparison) and found an average increase of 2.0% mitochondria and peridinin plastids of dinoflagellates, pre-
identity after editing and up to 6.3% for one gene (rpl36) sents an alternative hypothesis of adoption of pre-existing
(table 1). Random nucleotide edits would most often editing machinery. In plants, common components of the
reduce protein identity, so the predominant increases seen "editosome" are used in both plastids and mitochondria,
indicate that editing events are under selection to maintain showing that a common machinery can drive RNA processing
protein similarity and function (it is not possible to know in both organelles (Takenaka et al. 2012). Further, in K. vene-
whether small decreases in identity seen in rps13, rps18, ficum, several plastid genes from the pre-existing peridinin

789
Jackson et al. . doi:10.1093/molbev/mss270 MBE
Table 1. RNA Editing Events in 14 Tertiary Plastid-Encoded Transcripts.
Gene Sequence RNA Edits Number gDNA Coding cDNA Coding Increase to
Identity to Identity to E. hux Protein
E. hux (%) E. hux (%) Identity After Editing (%)
A – G 26
a G – A 2
psaA (1,227 nt) C – U — 72.8 76.7 3.9
U – C 6
A – G 15
G – A 2
psbC (904 nt) C – U 1 79.0 82.0 3.0
U – C 7
A – G 4
G – A 1
rbcL (1,062 nt) C – U 2 83.3 83.6 0.3
U – C 4
A – G 7

Downloaded from http://mbe.oxfordjournals.org/ at Universidade Federal de Viçosa on July 17, 2013


G – A —
atpG (178 nt) C – U 1 47.5 49.2 1.7
U – C 7
A – G 16
G – A 1
atpI (307 nt) C – U — 61.4 67.3 5.9
U – C 4
A – G 1
G – A —
rpl5 (139 nt) C – U — 77.8 80.0 2.2
U – C —
A – G 9
a G – A 1
rps13 (351 nt) C – U 1 50.0 49.2 0.8
U – C 11
A – G 6
G – A —
rps18 (335 nt) C – U — 26.3 25.0 1.3
U – C 3
A – G 10
G – A 1
rpl33 (181 nt) C – U — 53.8 59.6 5.8
U – C 4
A – G 3
G – A 1
rpl36 (153 nt) C – U 1 62.5 68.8 6.3
U – C —
A – G 4
G – A 1
petD (174 nt) C – U — 36.0 36.0 0.0
U – C 1
A – G 10
a G – A 4
rpoB (580 nt) C – U — 31.3 30.8 0.5
U – C 6
A – G 2
G – A —
rpoC2 (1,268 nt) C – U — 16.2 15.8 0.4
U – C 2
A – G 18
G – A 1
secYa (514 nt) C – U 2 25.1 25.7 0.6
U – C 12
A – G 131 (59.3%)
G – A 15 (6.8%)
Total (7,373 nt) C – U 8 (3.6%) Average 51.6 Average 53.6 Average 2.0
U – C 67 (30.3%)
NOTE.—Protein identity to haptophyte Emiliania huxleyi (E. hux) proteins was determined by three-way alignments with K. veneficum gDNA and cDNA conceptual translations.
New sequences are available in GenBank (accessions JX292160–JX292162, JX292164–JX292175, and JX945981–JX945986) or in the supplementary data, Supplementary Material
online.
a
Genes where internal stop codons are removed by editing.
790
A Tertiary Plastid Gains RNA Editing in Its New Host . doi:10.1093/molbev/mss270 MBE
1 99
rps13 RNA

Protein M
I F R I I Y
A L G V VH
100 198
rps13 RNA

Protein K N V Y R I I
E D A H K T T
199 297

rps13 RNA

Protein N K E Y Y*
D R GH HQ
298 351

rps13 RNA =A
=G
=U
Protein * =C

Downloaded from http://mbe.oxfordjournals.org/ at Universidade Federal de Viçosa on July 17, 2013


FIG. 1. Representative K. veneficum plastid gene, rps13, showing position and types of transitional mRNA editing events. Note multiple edits can act on
a single codon, and an internal stop codon (asterisk) is removed. Pre- and post-RNA-edited nucleotides are indicated above and below, respectively.
Encoded amino acids changes are similarly indicated in protein schematic.

Table 2. Codon Positional A + T Content in 2,443 Plastid-Encoded editing system would have no immediate function. However,
Codons. once established, it might allow subsequent mutations to
Codon Position A + T% gDNA A + T% cDNA occur that are then reversed at the RNA level. In K. veneficum,
1st 56.5 53.0 one evolutionary consequence of adopting editing is an in-
2nd 64.1 61.8 crease in the A + T nucleotide bias of plastid genes compared
3rd 84.0 83.1 with the transcripts (and also compared with E. huxleyi where
All 68.2 66.0 the average A + T content of homologous sequences equals
61%). It is possible that there is some positive advantage for
allowing this increase in A + T content; however, such an
plastid have persisted in the host nucleus, and their proteins advantage is not obvious. If the K. veneficum plastid did use
are now targeted into the haptophyte tertiary plastid pre-existing editing machinery, then this would represent a
(Nosenko et al. 2006; Patron et al. 2006; Waller et al. 2006). good case of neutral evolution, cultivated by the novel genetic
Therefore, it is possible that either pre-existing plastid RNA environment of the dinoflagellate host. It is interesting to
editing machinery, or that for the existing mitochondrial note that the K. veneficum plastid genome has apparently
system (or even both), could have been recruited to this undergone further accelerated change after entering its dino-
new plastid and contributed to its development of editing. flagellate host, with substantial gene loss and genome re-
Confirmation of either hypothesis awaits knowledge of the arrangement, although not yet approaching the level of
RNA editing machinery in dinoflagellates. plDNA modification seen in peridinin plastids (Howe et al.
Why RNA editing has been recruited by this new dinofla- 2008; Gabrielsen et al. 2011). Although the mechanism caus-
gellate endosymbiont is a more challenging question and one ing these genomic changes is unclear, they provide further
that is equally relevant to RNA editing in all systems. It is evidence of the peculiar influence that the dinoflagellate en-
tempting to consider that editing evolved to correct muta- vironment can have on its new endosymbiotic partners.
tions after they occurred, and, indeed, editing appears to re- RNA editing in the related tertiary plastid of Karenia miki-
store K. veneficum plastid sequences to greater identity with motoi has been simultaneously reported (Dorrell and Howe
homologs (table 1), as is also the case in plant organelles 2012). Interestingly, both transition and transversion edits are
(Wakasugi et al. 2001). However, it is unlikely that evolution reported for K. mikimotoi, suggesting possible ongoing devel-
or recruitment of the necessary editosome complexity could opment of complexity of the plastid editing machinery in this
occur quickly enough to rescue spontaneous deleterious mu- lineage.
tations. Moreover, across eukaryotes, the presence of organ-
elle RNA editing systems negatively correlates with genome
Supplementary Material
mutation rates, suggesting that mutational pressure does not Supplementary data are available at Molecular Biology and
explain development of this form of genome complexity Evolution online (http://www.mbe.oxfordjournals.org/).
(Lynch et al. 2006). The theory of “constructive neutral evo-
lution” presents an alternative hypothesis, positing that de-
Acknowledgments
velopment of the machinery for editing precedes its This work was supported by an Australian Research Council
requirement (Lukes et al. 2011; Gray 2012). In such a scenario, Discovery grant (DP0663590) and by the University of
the gradual development or recruitment of a nondeleterious Melbourne Science Faculty Scholarship to C.J.J.

791
Jackson et al. . doi:10.1093/molbev/mss270 MBE
References and their integration into biological systems. New York: Wiley Press.
p. 280–309.
Archibald J. 2009. The puzzle of plastid evolution. Curr Biol. 19: Lukes J, Archibald JM, Keeling PJ, Doolittle WF, Gray MW. 2011. How a
R81–R88. neutral evolutionary ratchet can build cellular complexity. IUBMB
Barbrook AC, Howe CJ, Kurniawan DP, Tarr SJ. 2010. Organization and Life 63:528–537.
expression of organellar genomes. Phil Trans R Soc B. 365:785–797. Lynch M, Koskella B, Schaack S. 2006. Mutation pressure and the evo-
Chateigner-Boutin A-L, Small I. 2010. Plant RNA editing. RNA Biol. 7: lution of organelle genomic architecture. Science 311:1727–1730.
213–219. Nosenko T, Lidie KL, Van Dolah FM, Lindquist E, Cheng J-F, Bhattacharya
Dang Y, Green BR. 2009. Substitutional editing of Heterocapsa triquetra D. 2006. Chimeric plastid proteome in the Florida “red tide” dino-
chloroplast transcripts and a folding model for its divergent chloro- flagellate Karenia brevis. Mol Biol Evol. 23:2026–2038.
plast 16S rRNA. Gene 442:73–80. Patron NJ, Waller RF, Keeling PJ. 2006. A tertiary plastid uses genes from
Dodge JD. 1989. Phylogenetic relationships of dinoflagellates and their two endosymbionts. J Mol Biol. 357:1373–1382.
plastids. In: Green JC, Leadbeater BSC, Diver WL, editors. The chro- Salone V, Rüdinger M, Polsakiewicz M, Hoffmann B, Groth-Malonek M,
mophyte algae: problems and perspectives. Oxford: Clarenden Press.
Szurek B, Small I, Knoop V, Lurin C. 2007. A hypothesis on the
p. 207–227.
identification of the editing enzyme in plant organelles. FEBS Lett.
Dorrell G, Howe CJ. 2012. Functional remodeling of RNA processing in
581:4132–4138.
replacement chloroplasts by pathways retained from their prede-
Shalchian-Tabrizi K, Minge MA, Cavalier-Smith T, Nedreklepp JM,
cessors. Proc Natl Acad Sci U S A. 109:18879–18884.
Klaveness D, Jakobsen KS. 2006. Combined heat shock protein 90

Downloaded from http://mbe.oxfordjournals.org/ at Universidade Federal de Viçosa on July 17, 2013


Gabrielsen TM, Minge MA, Espelund M, et al. (11 co-authors). 2011.
Genome evolution of a tertiary dinoflagellate plastid. PLoS One 6: and ribosomal RNA sequence phylogeny supports multiple replace-
e19132. ments of dinoflagellate plastids. J Eukaryot Microbiol. 53:217–224.
Gould S, Waller R, McFadden G. 2008. Plastid evolution. Annu Rev Plant Takenaka M, Zehrmann A, Verbitskiy D, Kugelmann M, Härtel B,
Biol. 59:491–517. Brennicke A. 2012. Multiple organellar RNA editing factor
Gray MW. 2012. Evolutionary origin of RNA editing. Biochemistry 51: (MORF) family proteins are required for RNA editing in mitochon-
5235–5242. dria and plastids of plants. Proc Natl Acad Sci U S A. 109:5104–5109.
Howe CJ, Nisbet RER, Barbrook AC. 2008. The remarkable chloroplast Takishita K, Ishida K-I, Maruyama T. 2004. Phylogeny of nuclear-encoded
genome of dinoflagellates. J Exp Bot. 59:1035–1045. plastid-targeted GAPDH gene supports separate origins for the
Ishida K-I, Green BR. 2002. Second- and third-hand chloroplasts in dino- peridinin- and the fucoxanthin derivative-containing plastids of
flagellates: phylogeny of oxygen-evolving enhancer 1 (PsbO) protein dinoflagellates. Protist 155:447–458.
reveals replacement of a nuclear-encoded plastid gene by that of a Tengs T, Dahlberg OJ, Shalchian-Tabrizi K, Klaveness D, Rudi K, Delwiche
haptophyte tertiary endosymbiont. Proc Natl Acad Sci U S A. 99: CF, Jakobsen KS. 2000. Phylogenetic analyses indicate that the
9294–9299. 190 hexanoyloxy-fucoxanthin-containing dinoflagellates have tertiary
Jackson CJ, Gornik SG, Waller RF. 2012. The mitochondrial genome and plastids of haptophyte origin. Mol Biol Evol. 17:718–729.
transcriptome of the basal dinoflagellate Hematodinium sp.: charac- Wakasugi T, Tsudzuki T, Sugiura M. 2001. The genomics of land plant
ter evolution within the highly derived mitochondrial genomes of chloroplasts: gene content and alteration of genomic information by
dinoflagellates. Genome Biol Evol. 4:59–72. RNA editing. Photosynth Res. 70:107–118.
Janouškovec J, Horák A, Obornik M, Lukes J, Keeling PJ. 2010. A common Waller RF, Jackson CJ. 2009. Dinoflagellate mitochondrial genomes:
red algal origin of the apicomplexan, dinoflagellate, and heterokont stretching the rules of molecular biology. Bioessays 31:237–245.
plastids. Proc Natl Acad Sci U S A. 107:10949–10954. Waller RF, Slamovits CH, Keeling PJ. 2006. Lateral gene transfer of a
Knoop V. 2011. When you can’t trust the DNA: RNA editing changes multigene region from cyanobacteria to dinoflagellates resulting in
transcript sequences. Cell Mol Life Sci. 68:567–586. a novel plastid-targeted fusion protein. Mol Biol Evol. 23:1437–1443.
Lin S, Zhang H, Gray MW. 2008. RNA editing in dinoflagellates and its Wang Y, Morse D. 2006. Rampant polyuridylylation of plastid gene
implications for the evolutionary history of the editing machinery. transcripts in the dinoflagellate Lingulodinium. Nucleic Acids Res.
In: Smith HC, editor. RNA and DNA editing: molecular mechanisms 34:613–619.

792

Vous aimerez peut-être aussi