Skip to main content

Chloroplast genome sequencing analysis of Heterosigma akashiwo CCMP452 (West Atlantic) and NIES293 (West Pacific) strains

Abstract

Background

Heterokont algae form a monophyletic group within the stramenopile branch of the tree of life. These organisms display wide morphological diversity, ranging from minute unicells to massive, bladed forms. Surprisingly, chloroplast genome sequences are available only for diatoms, representing two (Coscinodiscophyceae and Bacillariophyceae) of approximately 18 classes of algae that comprise this taxonomic cluster.

A universal challenge to chloroplast genome sequencing studies is the retrieval of highly purified DNA in quantities sufficient for analytical processing. To circumvent this problem, we have developed a simplified method for sequencing chloroplast genomes, using fosmids selected from a total cellular DNA library. The technique has been used to sequence chloroplast DNA of two Heterosigma akashiwo strains. This raphidophyte has served as a model system for studies of stramenopile chloroplast biogenesis and evolution.

Results

H. akashiwo strain CCMP452 (West Atlantic) chloroplast DNA is 160,149 bp in size with a 21,822-bp inverted repeat, whereas NIES293 (West Pacific) chloroplast DNA is 159,370 bp in size and has an inverted repeat of 21,665 bp. The fosmid cloning technique reveals that both strains contain an isomeric chloroplast DNA population resulting from an inversion of their single copy domains. Both strains contain multiple small inverted and tandem repeats, non-randomly distributed within the genomes. Although both CCMP452 and NIES293 chloroplast DNAs contains 197 genes, multiple nucleotide polymorphisms are present in both coding and intergenic regions. Several protein-coding genes contain large, in-frame inserts relative to orthologous genes in other plastids. These inserts are maintained in mRNA products. Two genes of interest in H. akashiwo, not previously reported in any chloroplast genome, include tyr C, a tyrosine recombinase, which we hypothesize may be a result of a lateral gene transfer event, and an unidentified 456 amino acid protein, which we hypothesize serves as a G-protein-coupled receptor. The H. akashiwo chloroplast genomes share little synteny with other algal chloroplast genomes sequenced to date.

Conclusion

The fosmid cloning technique eliminates chloroplast isolation, does not require chloroplast DNA purification, and reduces sequencing processing time. Application of this method has provided new insights into chloroplast genome architecture, gene content and evolution within the stramenopile cluster.

Background

Stramenopiles represent an enormous eukaryotic assemblage of 500,000 to one million species which includes both algae and colorless protists [1, 2]. Algal representatives within this major branch in the tree of life are exceptionally diverse. They include recently discovered minute, picoplanktonic unicells (Pinguiophyceae), as well as colonial forms (Synurophyceae), the silicious diatoms (Coscinodiscophyceae, Bacillariophyceae and Fragilariophyceae), and the large pseudoparenchymatous kelps (Phaeophyceae), which may attain lengths of at least 150 feet. These autotrophic eukaryotes serve as primary producers that fix at least 40% of the total carbon processed on earth and significantly impact global sulfur and nitrogen cycles [3–7]. Although some stramenopiles adversely affect aquaculture endeavors and ecosystem health through formation of toxic blooms [8–10], others form dense underwater forests which serve as habitat for myriad vertebrate and invertebrate species. Stramenopiles are not only used extensively in industry, in aquaculture and as a human food source, but they also provide research opportunities for novel pharmaceutical discovery and nanotechnological development [11].

Autotrophic stramenopiles evolved approximately 100 million years ago [12–16]. Their chloroplasts (secondary endosymbionts) significantly differ from those of green algae, land plants or rhodophytes (primary endosymbionts), in morphology, pigment composition, storage materials and chromosome gene content [17]. For this reason, one cannot assume identical chloroplast function among representatives of these disparate taxa. Presently, over 100 chloroplast genomes have been sequenced, predominantly from terrestrial plants. In contrast, few molecular data exist describing the underlying genetic profiles of chloroplast DNA (cpDNA) among the approximately 18 classes of autotrophic stramenopiles. At this writing, the only stramenopile chloroplast genomes that have been published, are those of the diatoms Odontella sinensis, Thalassiosira pseudonana (both in the class Coscinodiscophyceae) and Phaeodactylum tricornutum (Bacillariophyceae) [18–20]. One factor that has hindered progress in stramenopile chloroplast genome sequencing is difficulty in obtaining purified cpDNA. Typically, this process is accomplished by physically isolating chloroplasts before DNA extraction, or by separating cpDNA from mitochondrial and nuclear DNA in cesium chloride gradients. The first approach is extremely difficult in this group of organisms, particularly those of picoplanktonic size, and the second is labor intensive, requiring sufficient biomass for DNA isolation, and repeated series of multi-day centrifugation spins [21].

In this study we sequenced the chloroplast genome of two Heterosigma akashiwo (Raphidophyceae) strains originating from West Atlantic (CCMP452) and West Pacific (NIES293) coastal waters. We initiated our study of H. akashiwo cpDNA using a standard shotgun sequencing method with highly purified cpDNA retrieved from over 80 liters of cell culture. Alternatively, to bypass the tedious process of cpDNA purification, we used a simplified whole genome fosmid cloning approach to determine cpDNA sequences. For each strain, we constructed a fosmid library using whole cellular DNA (nuclear, mitochondrial and chloroplast) from approximately 2 liters of culture. Chloroplast clones were selected from the total genomic DNA preparations using bioinformatic analysis of fosmid end-sequences, obtained via high throughput sequencing. Sequencing fosmid subclones independently aided in final finishing of the genomes, as has been discussed previously [22, 23].

Heterosigma akashiwo is a small (12 μm), naturally wall-less unicell that forms toxic brown tides in temperate and subtropical regions world-wide [24–26]. As a coastal-dwelling organism, H. akashiwo also contributes significantly to primary productivity within these critically important ecosystems [27]. Significant research on its morphology [28], physiology [29–31], molecular biology [32–34], toxicology [35, 36], and biochemistry [37–39] define H. akashiwo as one of the most broadly studied non-diatomaceous stramenopiles. Much of this attention has been focused on events associated with chloroplast biology. For example, both photoperiod and light intensity determine the number of chloroplasts per cell (13 to 40) and the phase, amplitude and period of their synchronized division [40, 41]. A chloroplast run-on transcription system (the only one developed for stramenopiles) not only shows that chloroplast RNA abundance is regulated predominantly at the transcriptional level, but that transcriptional response is also modified by the physiological challenges imposed on the cell [42, 43]. An average H. akashiwo cell contains about 600 copies of its chloroplast genome [40]. Electron microscope studies [21], combined with restriction enzyme digestion [44], reassociation kinetic analysis [45], and physical mapping [46, 47] reveal that the approximately 154 kb H. akashiwo chloroplast genome is a circular molecule which contains a large, inverted repeat (IR). Demonstration of a chloroplast-encoded rubisco small subunit [46, 48] and documentation of the presence of bacterial-like two-component signal transduction arrays [49, 50] gave early evidence that the chloroplast genome of H. akashiwo may be functionally distinct from those of green algae and land plants.

The existence of an extensive database augments H. akashiwo's potential as a model system for studies in stramenopile chloroplast evolution and biogenesis. It has been suggested that H. akashiwo strain CCMP452 serve as the reference genotype for this organism [51]. New data reported here show that the chloroplast genome sequence of H. akashiwo: (a) displays marginal synteny with other chloroplast genomes including those of the diatoms; (b) contains six genes encoding proteins of unknown function; (c) lacks introns; and (d) has genes that appear to have been obtained via lateral transfer.

Results and Discussion

Sequencing strategy: conventional vs. fosmid approach

We compared two methods to obtain sequencing templates for these two strains, a standard CsCl cpDNA preparation, and total genomic DNA cloning into fosmid vectors. Using the standard approach, CsCl-purified H. akashiwo CCMP452 cpDNA was cloned into pUC18 plasmids and sequenced by the conventional shotgun cloning described in the Materials and Methods. A total of 1152 clones were sequenced in both forward and reverse direction, providing greater than 8× coverage, given an average read length of 550 base pairs (bp) and an estimated genome size of 150,000 bp. Purification of cpDNA sequencing template by this commonly used method was extremely labor intensive. It required the generation of large quantities of cells followed by the recovery of highly purified cpDNA using CsCl gradients. To avoid these technical challenges, we adapted a large-insert (fosmid) cloning method for total genomic DNA to cpDNA sequencing (Fig. 1). This fosmid cloning method requires minimal biological material and avoids the isolation of pure cpDNA. Our conventionally sequenced H. akashiwo CCMP452 chloroplast genome served as a reference for this endeavor. Briefly, total genomic DNA (nuclear, mitochondrial and chloroplast) was used to construct a large insert fosmid library. Using high-throughput fosmid DNA isolation and end-sequencing methods, these fosmids were then end-sequenced from their vector/insert junctions to determine clones of chloroplast origin.

Figure 1
figure 1

Fosmid cloning technique. High molecular weight, total DNA is subject to pulse-field electrophoresis to recover sheared DNA of 45 to 50 kb. This DNA is used to generate a fosmid library which is selectively screened for cpDNA-containing clones, which are then sequenced, annotated and assembled.

Chloroplast fosmid identity was determined two ways. The sequenced fosmid ends were compared to: (1) the draft sequence generated by the shotgun method and (2) a customized blast database consisting only of published chloroplast genome sequences. Earlier reports used hybridization to macroarrays comprised of chloroplast-genomic probes to screen for cpDNA-containing clones [22, 23]. In contrast, our end-sequence based approach does not rely on a priori knowledge of the cpDNA sequence. Hybridization screening could produce a high number of false positives given the homology of chloroplast gene sequences to bacterial and nuclear gene sequences, or missed clones given the divergence of stramenopile genes at the DNA sequence level. In addition, our method is easily updated and made more powerful as newly sequenced chloroplast genomes are added to the reference database. For additional genomes of autotrophic stramenopile taxa sequenced entirely from fosmids (Aureoumbra lagunensis, Pinguiococcus pyrenoidosus), we have found that relatively little finishing is required to obtain the complete genome once chloroplast genome fosmids are sequenced (unpublished, Cattolico et al.). Of 1,920 fosmids generated from H. akashiwo CCMP452 total DNA, twenty gave clear chloroplast signatures when compared to the draft conventionally sequenced genome. All twenty of these fosmids were also identified using the genome-independent bioinformatic approach, demonstrating that this method is feasible for de novo sequencing. Eight fosmids were fully sequenced to assemble the H. akashiwo CCMP452 chloroplast genome (Fig. 2A [GenBank Accession: EU168191]).

Figure 2
figure 2

H. akashiwo CCMP452 (A) and NIES293 (B) genome maps. Outer rim: genes on plus and minus strand, color coded according to function (see legend); Second ring: small inverted (red) and tandem (blue) repeats; Third ring: sequence comparison to the other H. akashiwo genome, including SNPs (blue), small insertions (green), deletions (red) and regions of extremely poor alignment (orange); Fourth ring: Location and size of fosmid clones color coded according to their orientation: supports depicted isoform (green), supports alternate isoform (pink), uninformative (black); Fifth ring: location of inverted repeats, large and small single copy domains. Red bar depicts location of 8 kb region inverted in CCMP452 relative to NIES293; inner circle: GC content.

Because the fosmid cloning technique for generating template DNA proved to be rapid, efficient and cost effective, it was also chosen to sequence the cpDNA of H. akashiwo NIES293, West Pacific strain. A total of 3,072 fosmids were end-sequenced using high-throughput methods to identify fosmids of chloroplast origin for sequencing. 2,304 additional clones were screened by Real Time PCR once the partial genome sequence had been obtained. Primers were designed from the draft genome sequence to search for clones that spanned gaps. In total twenty three fosmids were identified as chloroplast-derived and ten of these fosmids were fully sequenced to assemble the H. akashiwo NIES293 chloroplast genome (Fig. 2B [GenBank accession: EU168190]).

As noted above, although our ongoing studies show that entire stramenopile chloroplast genomes are clonable into fosmids, the fosmid coverage for both H. akashiwo CCMP452 and NIES293 cpDNA was not complete. Fosmids generated from some cpDNA domains were abundant, whereas others were minimal. As shown in Fig. 2, great difficulty in fosmid recovery was experienced for an identical region in both H. akashiwo strains. The reasons for extremely low coverage in this particular cpDNA region are not known. One might suggest that the genes encoded in this region (e.g., those necessary for ATP synthesis, cytochrome function, and DNA replication) influence the survival of bacterial host cells during fosmid library construction. Alternatively, insert packaging could be impeded by the presence of structural anomalies, such as branched replication or recombination intermediates, within a localized region of the cpDNA.

PCR was used to span those areas of the genome that were not found in clone libraries. For example, a gap of approximately 10 kb existed in NIES293 for which no fosmid clone was retrieved. To close this gap, a series of PCR primers was designed to create 1200 bp products, offset by an average of 350 bp per product. Primers were designed using the completed CCMP452 cpDNA sequence as reference. The sequenced PCR products were assembled, and confirmed to overlap with the fosmid sequences flanking the gaps. Similarly, a 0.1 kb gap in CCMP452 lacking shotgun clones was spanned by sequencing a single PCR product.

Global genome structure

The H. akashiwo CCMP452 chloroplast genome is 160,149 bp in size (Table 1). This chromosome contains a 21,822 bp IR which divides the molecule into large single copy (LSC: 77,470 bp) and small single copy (SSC: 39,035 bp) domains (Fig. 2A). The 159,370 bp H. akashiwo NIES293 chloroplast genome is shorter in the IR (21,665 bp) as well as the LSC (77,206 bp) and SSC (38,834 bp) domains (Fig. 2B). Notably, the H. akashiwo NIES293 SSC domain contains an ~8.0 kb inversion when compared to that of H. akashiwo CCMP452 (Fig. 2). An overall GC content of 30.5% is seen for CCMP452 while a GC content of 30.4% occurs in NIES293 cpDNA (Table 1, Fig. 2).

Table 1 Overview of H. akashiwo strains CCMP 452 and NIES 293 chloroplast genomes

The genomes of both H. akashiwo strains exist in two isomeric configurations. Both sequencing fosmids that span the repeats, and long PCR confirmed this observation. For H. akashiwo CCMP452, three fosmids (FA2278; FA2279; FA4020) which spanned the entire repeat, including some part of both single copy domains, were chosen for shotgun sequence analysis. Two of these fosmids (FA2279; FA4020) assembled into isomeric form A (Fig. 2A) while the third showed the alternate isomer, form B. Similarly, for H. akashiwo NIES293, three sequenced fosmids spanned the IRs, one belonging to isomeric form A (FA3944) and two to the alternate form B (FA4254, FA8926) (Fig. 2B). To further confirm the presence of two isomeric forms in H. akashiwo CCMP452, primers designed to the ends of each single copy region (Fig. 2A) were used in multiple combinations in long PCR to probe for the presence of both potential configurations. The isomers found in these chloroplast genomes may have been formed by a recombination event within the IR which resulted in the inversion of the single copy domains relative to one another (Fig. 3).

Figure 3
figure 3

Isomeric cpDNA populations. Single copy regions are flipped resulting from a recombination event. Arrows show positions of sequence in large and small single copy regions.

The observation that cpDNAs exist as a heterogeneous population is not new. In 1983, Palmer hypothesized that a recombination event within the IR of Phaseolus vulgaris generated an equimolar population of isomeric cpDNA molecules which differed only by the orientation of their single copy regions [52]. The subsequent demonstration of "polarity reversal" of the single copy region resulting in the generation of isomeric cpDNAs in angiosperms [53], in a chlorophytic alga [54], in the stramenopiles Vaucheria bursa [55], Cyclotella meneghiniana [56], and H. akashiwo (this work), argues for the widespread occurrence of this process across divergent taxa. Our fosmid cloning approach eliminates the laborious process of using extensive restriction analysis of cpDNA to document the flipping of single copy domains. By judiciously choosing fosmids (40 to 45 kb), one can easily document cpDNA isomerization. An additional advantage of the fosmid technique is that the investigator can readily distinguish the identity of IR number one from IR number two. In conventional shotgun sequencing strategies, assignment of a sequence to a specific repeat domain is frequently challenging [22], especially if the IR is large, as is often found in terrestrial plants. When assembling the genome from shotgun data, the large IR elements collapse and final finishing typically requires in-silico duplication of the IR to complete the genome sequence. This approach may lead to errors, especially if the repeats are not identical as seen in the cryptophyte Guillardia theta [57].

It is well established that repeat size can both expand and contract [52, 53]. The ~22 kb H. akashiwo IR is similar in size to that found in T. pseudonana [~18 kb], C. meneghiniana [~17 kb], and Skeletonema costatum [~20 kb]) but significantly larger than the 6 kb (sufficient in size solely to encode the ribosomal operon) repeat domain seen in the genomes of rhodophytes and most algae that contain chloroplasts of secondary endosymbiotic origin [55, 56, 58, 59]. Many stramenopile chloroplast genomes appear to maintain an IR (e.g., Dictyota dichotoma, O. sinensis, P. tricornutum, Pylaiella littoralis, V. bursa) [60]. New sequencing data suggest that other stramenopile chloroplast genomes may lack this architectural feature altogether (e.g. A. lagunensis; unpublished data). Although data are sparse, haptophyte [61] and cryptophyte [57] chloroplasts also appear to maintain a small IR. Rhodophyte chloroplast genomes [58, 62, 63] display an inverted or direct repeat (e.g., Cyanidium caldarium, Cyanidioschyzon merolae, Galderia sulphuraria, Gracilaria tenuistipitata) or may lack a repeat entirely (e.g., Chondrus crispus, Griffithsia pacifica, Porphyra yezoensis).

Gene Content

The H. akashiwo CCMP452 and NIES293 genomes are co-linear with respect to gene content, with exception of ten genes (see below) which are located within the ~8.0 kb inversion inside the small single copy region (Fig. 2). An overall protein coding content of 68.5% is seen for CCMP452 and 69.0 % occurs in NIES293 cpDNA (Table 1, Fig. 2).

RNA genes include the ribosomal RNA operons, one copy in each IR, one tmRNA, one threonine pseudo-transfer RNA (anticodon UGU), and 34 tRNA genes whose anticodons encompass 20 different amino acids. Seven of these tRNA genes are located in each IR, resulting in a total of 27 distinct tRNA genes. Three tRNA genes have anticodons for methionine, although previous studies suggest one of these tRNAs may be subsequently modified to a tRNA isoleucine [64]. Also present is the widely conserved tRNA glutamine (UUC), which contributes to translation and also plays an integral role in the biosynthetic pathway of δ-aminolevulinic acid, the precursor for generating the tetrapyrole-containing pigments, heme, chlorophyll and bilin in bacteria and algae as well as in terrestrial plants [65–67]. Many codons found in the genes of the H. akashiwo genomes have no corresponding anticodon in the tRNAs that are encoded in the cpDNA. Although tRNAs are imported into the mitochondrion [68], presently there is no evidence that they are similarly imported into the chloroplast. Comparing the codon usage of the predicted ORFs to the anticodons of the resident tRNA complement, one might suggest that 50% of the tRNAs use a wobble base at the third codon position. This codon-anticodon discrepancy is also present in other chloroplast genomes of secondary endosymbiotic origin.

Both H. akashiwo chloroplast genomes contain genes encoding 156 predicted proteins, including a core set of 45 genes which are conserved in all chloroplast genomes sequenced to date. An additional 48 genes are conserved in chloroplast genomes of rhodophytes and in algae with chloroplast genomes of secondary endosymbiotic origin [61]. Of the 156 genes for predicted proteins, approximately one-third encode products used in photosynthesis or energy generation. All the ATP synthase genes (atp A, D, G, H, I) are found with the exception of atp C; all the genes of the electron transfer chain (pet A, B, D, F, G, J, L, M, N) as well as genes important in Calvin cycle function (Form II rubisco large and small subunits rbc L and rbc S, the putative rubisco expression protein cfx Q [cbb X], and rubisco transcriptional regulator ycf 30 [rbc R]) are also present. The genomes also contain 19 conserved hypothetical genes common to other chloroplast genomes (ycf s) and six open reading frames with no sequence homology to genes in other chloroplast genomes.

The chloroplast genomes of H. akashiwo and the diatoms T. pseudonana, O. sinensis, and P. tricornutum have diverged in gene content. The three diatom genomes are extremely similar in gene content; there are only 3 genes (acp P, syf B, tsf) encoded by at least one but not all 3 of these algae. In contrast, although both diatoms and H. akashiwo share an identical set of 125 protein-coding genes (both identified and ycf's), H. akashiwo also maintains genes found in rhodophytic cpDNA (e.g., acs F, ftr B, ilv B, ilv H, pet J, rps 1, trg 1, tsg 1, as well as ycf 17, ycf 34, ycf 36, ycf 54, ycf 65). Conversely, the three diatoms contain seven genes not present in H. akashiwo (the rps 6, sec G, ycf 42, ycf 88, ycf 89, and ycf 90 protein-coding genes as well as ffs, the 4.5S RNA signal recognition particle component).

Novel genes

We have now entered an era in which the comparative genomics of autotrophic eukaryotes can be studied. By cataloguing genes from broadly sampled taxa, we increase both our understanding of chloroplast evolution and gain insight into biochemical mechanisms that drive chloroplast homeostasis. However, this task is not easily accomplished, for chloroplast genomes probably represent a chimeric assemblage of genes which originate from both ancestral symbiont and lateral gene transfer events. For example, the H. akashiwo chloroplast genome retains the genes trg1 and tsg1, encoding a functional two-component His-to-Asp signal transduction circuit [49]. Similar circuits are found in all cyanobacterial cells, the putative ancestral source of chloroplast genomes. The sensor kinase/response regulator protein pair is responsible for converting physiological information from the environment to a program that regulates gene transcription. Although genes for one or both of these proteins are found in most genomes of rhodophytic lineage, no His-to-Asp pair is encoded in the three diatom cpDNAs which have been sequenced. Thus by analyzing these proteins, we document the retention of ancestral proteins (evolutionary footprints?), and describe a mechanism of gene regulation which is confined to a specific taxonomic cluster (see [49] for discussion). Expanding this approach, we have determined a possible function for two additional genes present in H. akashiwo which have not been found in any other chloroplast genome.

tyrC

Both H. akashiwo chloroplast genomes contain a gene that encodes a putative site-specific tyrosine recombinase, which we have named tyr C (tyrosine recombinase/chloroplast). The translated H. akashiwo TyrC protein is 318 and 298 amino acids in length in strains NIES293 and CCMP452 respectively (Fig. 4). In strain NIES293 residues 129 and 130 are lacking. A significant change in the CCMP452 tyr C gene is effected by the inversion that occurs in the SSC region of this genome (Fig. 2). This flip relocates 69 bp of the tyr C 3' terminus to a new location which is ~8.0 kb downstream. The predicted amino acids encoded by the displaced region in CCMP452 retain 100% sequence identity to those present in the intact NIES293 protein.

Figure 4
figure 4

Comparison of H. akashiwo CCMP452, H. akashiwo NIES293, Chaetosphaeridium globosum and Prototheca wickerhamii recombinases. Gray shading indicates residues completely conserved among the four proteins. Stars indicated conserved residues important in catalytic function. Overline and underline are box I and box II respectively.

Proteins with the greatest similarity to the putative H. akashiwo recombinase are found in the mitochondrial genomes of Prototheca wickerhamii, a chlorophyte closely related to Chlorella vulgaris, and in the charophyte Chaetosphaeridium globosum (Fig. 4). In addition to these algal mitochondrial tyrosine recombinases, H. akashiwo TyrC has amino acid sequence similarity to the recombinases found in Lactobacillus leichmannii, Picrophilus torridus and Methanococcus maripaludis. Furthermore, the H. akashiwo tyr C genes have a 25% GC content in the third codon position, markedly higher than the 14% average for genes on the H. akashiwo cpDNA, suggesting that this gene may be the product of a lateral gene transfer event.

Because there is such a limited sequence similarity among known integrases the identification of these proteins often relies upon the identification of essential catalytic residues [69]. The putative H. akashiwo TyrC protein contains numerous motifs defined for the integrase family of recombinases [70]. This protein retains the critically important catalytic residues (CCMP452 numbering): Arg 143 (with a conserved glutamate located three amino acids downstream), His 248, Arg 251 and Tyr 283 (Fig. 4). These residues have been shown to lie close to the active site when the protein is folded. Mutation of any one of these amino acids reduces or eliminates recombinase activity [69, 71, 72]. All bacterial sequences with similarity to H. akashiwo TyrC noted above also retain the Arg-His-Arg amino acid triad as well as the Tyr nucleophile component. Additionally, H. akashiwo TyrC displays the highly conserved domains designated Box I and II by Nunes-Duby and colleagues [73] in their comparative analysis of 105 site-specific recombinases.

Though the tyr C gene is expressed in both H. akashiwo strains (Deodato and Cattolico, unpublished), presently, we can only speculate on the function of its translated protein product. In bacteria, site-specific recombination often utilizes the tyrosine recombinase pair XerC and XerD, which may be evolutionary derivatives of a single ancestral protein [73, 74]. Conventionally, the XerC/D protein pair breaks and rejoins DNA strands at short, conserved, 28 base-pair domains (dif sites) through the formation of Holliday junction intermediates [75–77]. This docking domain usually consists of two 11-base-pair "arms" with a 6-nucleotide central region (Table 2). Four types of putative dif recognition domains are present in the H. akashiwo chloroplast genomes (Table 2). Whether these nucleotide domains truly serve as points for intramolecular recombination, or sites where multimeric [21] H. akashiwo cpDNA molecules are converted to monomers, warrants further experimentation.

Table 2 Comparison of putative dif sites in H. akashiwo chloroplast genomes with those of selected bacteria and viruses

Trans-membrane protein

An extremely large protein comprised of 456 amino acids is encoded in the IR of both strains (Heak452_Cp006/Heak452_Cp062; Heak293_Cp006/Heak293_Cp062). Expression of this large gene has been verified by quantitative RT-PCR in both strains (Deodato and Cattolico, unpublished). A variety of sequence analysis techniques have been used to gain some insight into the nature of this unique chloroplast gene. Standard BLAST queries against all routinely available databases reveal no significant known homologs. Searches with PSI-BLAST [78] indicate that the most closely related proteins in standard databases are a series of putative G protein-coupled receptors (GPCR) in C. elegans. Other significant partial hits (i.e., alignment of fragments of 60–120 residues with ~30% sequence identity and 40–60% identity plus conservative substitution with minimal to modest gapping) include FMLP receptors (human and mouse), LSH receptor (human and pig), melanocortin-3 receptor (rat), and metabotropic glutamate receptor 5 (rat). Hydrophobicity analyses and membrane topology prediction suggest that the undescribed H. akashiwo protein sequence possesses seven probable transmembrane segments; the length and hydrophobic residue repeat patterns in the putative transmembrane segments are consistent with an alpha-helical structural motif. The qualitative features of the transmembrane helix prediction profiles are more similar to the profiles observed in other G protein-coupled receptors from the rhodopsin/beta-adrenergic class (6 clear transmembrane segments, and a seventh segment which is at the threshold margin for transmembrane assignment) than they are to bacterial halorhodopsin proteins, which have seven strong transmembrane segments [79–81].

Attempts to align the undescribed H. akashiwo protein sequence with a collection of sequences from the rhodopsin/beta-adrenergic (Group A) receptor family were largely unsuccessful. We were unable to generate an alignment although the H. akashiwo protein sequence displays 12–18% amino acid sequence identity with various members of a compiled GPCR data set, comparable to the sequence identity observed for bovine rhodopsin with many adrenergic receptors. The H. akashiwo protein sequence does exhibit some key signature features of G protein-coupled receptors, such as an NRF motif at the carboxy terminal end of the third putative transmembrane segment, which is an observed variant of the well-characterized DRY motif in the GPCR superfamily. In contrast the H. akashiwo protein sequence does not possess the highly conserved disulfide bond observed in the extracellular loops of many GPCRs. The H. akashiwo protein does possess a number of glycosylation, myristoylation, and phosphorylation sites in combinations and locations similar those observed for G-protein-coupled-receptor sequences.

On the basis of these analyses, the H. akashiwo protein sequence appears to be an integral membrane protein with seven probable transmembrane segments. It exhibits sequence characteristics that suggest it may be a G protein-coupled receptor, related most closely to the rhodopsin/beta-adrenergic receptor family, although we have not been able to generate convincing pairwise or multiple sequence alignments with other members of the GPCR superfamily. If the H. akashiwo protein sequence is indeed the first member of the GPCR superfamily in the chloroplast of an alga, it is obviously strongly diverged from the GPCRs seen in animals. However, because this protein looks far more like a G protein-coupled receptor than it does anything else currently present in sequence databases, more detailed biochemical characterization of the H. akashiwo protein sequence is warranted.

Gene arrangement

Four protein-coding genes use GTG starts (rbc S, psb F, PRSP-3 [ycf 65], rps 3). There is no consistency within stramenopiles or rhodophytes for chloroplast genes that initiate with a non-ATG start. Two sets of overlapping genes are common to both genomes: psb C and psb D (32 codons), and Heak452Cp_021/gro EL (3 codons). Additionally, in CCMP452, the Heak452_Cp014 (orf97)/chl I genes overlap by 7 codons. However, a one base-pair insertion in NIES293 results in a frame shift that causes orf97 and chl I genes to be contiguous. Sequence alignment of NIES293 orf97 and the functional CCMP452 96-amino acid sequence shows that the amino termini of these polypeptides are virtually identical (98% homology among the first 65 amino acids). Given that CCMP452 orf97 is differentially expressed over the cell cycle [34], it will be of interest to determine whether the altered NIES293 protein retains its functionality.

Unlike terrestrial plant and green algal chloroplast genomes, but similar to rhodophytic chloroplast genomes and other chloroplast genomes of secondary endosymbiotic origin, no introns have been detected in H. akashiwo chloroplast-encoded genes. However, a conserved putative intein [82] in dna B is maintained, and numerous other genes encode proteins that contain in-frame amino acid deletions or insertions when compared to homologues in other algal chloroplast genomes. Proteins having the largest inserts include ClpC (multiple: 90, 43, 41 amino acids) and RpoA (79 amino acids). Among the 16 protein-coding genes modified by inserts, it appears that some common functional identities occur. These include five members of the ATP complex, AtpA (2 amino acids), D (4, 5, 12, and 2 amino acids), G (2 amino acids), B (1 amino acid) and E (1 amino acid) as well as five ribosomal proteins, RpL4 (14 amino acids), RpL18 (20 amino acids), Rps5 (2 amino acids), Rps9 (5, 2, and 3 amino acids), and Rps10 (11 amino acids). Proteins that have significant, extended carboxy termini include Rps10 (31 amino acids), Ycf16 (32 amino acids), and ClpC (46 amino acids). Comparison of genomic sequences to cDNAs generated for clp C, rpo A, rpl 18, rps 5, and rps 10 shows that the inserts are retained in mature mRNA. Whether they are removed after translation remains unknown.

Globally, H. akashiwo cpDNA in either isomeric form shows little synteny with published cpDNAs (Fig. 5), though sub-domains of conservation in gene placement are evident. As in other chloroplast genomes of the rhodophytic or secondary endosymbiotic lineage, the ribosomal protein genes occur in clusters. The largest of these conserved arrays is the "ribosomal protein block" which includes 26 ribosomal genes as well as tuf A, rpo A and sec Y [83]. Dna K is almost universally found 3' to this ribosomal protein-coding domain. This gene cluster may represent an evolutionarily conserved, prokaryotic-like transcriptional operon in which large numbers of ribosomal protein genes are co-transcribed [84]. Indeed, northern analysis using probes spanning the entire "ribosomal protein block" of G. theta cpDNA revealed the production of an mRNA transcript of approximately 16 kb. Smaller mRNAs in this northern analysis, likely a product of primary transcript processing, were also detected [85].

Figure 5
figure 5

Synteny among stramenopile and red-lineage chloroplast genomes. H. akashiwo vs (A) Odontella sinensis and (B) Porphyra purpurea; Thalassiosira pseudonana vs (C) Odontella sinensis and (D) Porphyra purpurea.

Numerous smaller, intact motifs seen in all rhodophytic and secondary endosymbiotic chloroplasts examined to date are maintained in H. akashiwo cpDNA. Among the conserved gene clusters are the atp B/atp E and atp I/atp H/atp G/atp F/atp D/atp A complexes, the ribosomal genes rpl 11/rpl 1/rpl 12; rpl 27/rpl 21, the photosynthetic genes psa A/psa B, psb D/psb C, psb B/psb T/psb N/psb H as well as the Calvin cycle rbc L/rbc S genes (often in association with cfx Q) (Fig. 2). Conservation in gene order is maintained in the placement of the H. akashiwo initiator methionine tRNA. As in rhodophytes and algae having chloroplasts of secondary endosymbiotic origin, this tRNA is embedded between psa D and ycf 36. Interestingly, rps 14, which is adjacent to initiator methionine tRNA in most green algae and land plants, lies immediately upstream of the psa D gene in the H. akashiwo chloroplast genomes. In the rhodophytic lineage the rpo C2C1 B1/rps 20/gln B/rpl 33/rps 18 polymerase cluster appears to have undergone dissolution through a series of independent events. Two genes (rps 20 and gln B) in the cluster appear to have been targeted for removal or transfer to the nucleus. The intact cluster is present in Porphyra purpurea and P. yezoensis. Cluster integrity is maintained in H. akashiwo, O. sinensis, P. tricornutum, G. theta and G. tenuistipitata, although gln B is lost. In C. caldarium rps 20 rather than gln B has been eliminated. A. lagunensis lacks both rps 20 and gln B, as does the haptophyte Emiliania huxleyi, which also splits rpo C2C1 B1 and rpl 33/rps 18 into distantly-located clusters.

Analysis of cluster integrity has been a valuable tool in the assessment of phylogenetic identity and evolutionary processes (e.g. [86, 87]). The data presented here give evidence that both gene cluster maintenance and dissolution have occurred in the H. akashiwo chloroplast genomes. Unfortunately, comparative analysis of gene flux solely within the stramenopiles is hampered by the paucity of available data, since H. akashiwo is the only non-diatom genome published from this group. However, the small data set available already suggests that the stramenopiles will present a significant challenge, especially in deciphering the dynamics of gene cluster flux and variations in gene co-linearity patterns within this taxon.

Indels and SNPs

Though the genomes of H. akashiwo CCMP452 and NIES293 are largely co-linear and have identical gene content, there are 150 single nucleotide polymorphisms (SNPs) between them. Within the 35 protein-coding genes containing SNPs, both synonymous (30) and/or non-synonymous (36) changes are noted (Table 3). These changes occur in informational (e.g., rpo B, rps 14) as well as operational (e.g., fts H, sec Y) genes. Also seen are small, variable regions containing deletions and insertions of one to six nucleotides. These small variable regions are clustered into "hot spots" which appear throughout the genome (Fig. 2). Additionally, six large, variable regions, which are predominantly located in the SSC region, represent the major cpDNA sequences between the two H. akashiwo strains.

Table 3 Presence of Single Nucleotide Polymorphisms in protein coding genes between H. akashiwo CCMP452 and NIES293

The extent to which cpDNA sequence varies among H. akashiwo ecotypes is not known. Unicellular algae, such as H. akashiwo, often exist in high-density populations that are generated via rapid cell division. If DNA replication serves as a mutational driver, then chloroplast genetic profiles might be expected to shift during the biogenesis of an algal bloom [88]. When examining genetic difference between strains, analyzing incomplete genomes or standard nuclear markers may be misleading. For example, analysis of chloroplast rbc L/S as well as nuclear 18S and ITS rDNA (markers that have proven to be reliable in other taxa) suggested that approximately 40 H. akashiwo strains of different geographic origin were of identical genotype (Ki and Han, 2007; Connell, 2000). This conclusion led the authors to propose that geographic distribution of H. akashiwo is due to a global dispersal mechanism. By sequencing whole genomes, the presence of appreciable genetic differences in cpDNA between strains was made clear, and suggests a diverged ancestry for CCMP452 and NIES293. Continued sequence analysis of additional strains may show an even greater variation among H. akashiwo populations. For example, six variants of the cfx Q gene (1 to 2 nucleotide changes) are seen when 24 H. akashiwo strains are analyzed (Lee, Hoyt, Lakeman and Cattolico, unpub.). In-silico modeling suggests that the non-synonymous changes observed in the sequence of cfxQ, may impact protein function [89].

Repeats

Analysis of the H. akashiwo chloroplast genome reveals the presence of numerous AT-rich repeats (Table 4). CCMP452 has 40 inverted and 25 tandem repeats that represent 2.62% of the total genome, whereas NIES293 cpDNA has 36 inverted and 23 tandem repeats encompassing 2.38% of this genome. Both strains retain many identical repeat structures. Substitution, loss or gain of nucleotides within a repeat motif is not confined to one H. akashiwo strain. Essentially all major changes in these repeat elements occur within intergenic regions.

Table 4 Occurrence of tandem and inverted repeats in chloroplast genomes of rhodophytes and algae with chloroplasts of secondary endosymbiotic origin

Inverted repeats found in H. akashiwo cpDNA are comprised of a stem structure which ranges from 17 to 87 bp in length (average 36.9 +/- 15.6 bp). The loop domain of these inverted repeat arrays is very small, averaging only 5.49 +/- 3.5 bp. Thus the average inverted repeat structure is approximately 42 bp in size. Tandem repeats have a period of 18.1 +/- 5.9 bp (CCMP452) or 19.9 +/- 4.0 bp (NIES293) with a copy number ranging from 1.9 to 7.5. Thus, the average tandem repeat element is 37.5 +/- 5.0 bp in size. Whether the repeat size maintenance of approximately 40 bp for both inverted and tandem repeats has functional significance is not known.

Notably, many repeats (including both tandem and inverted types) are localized within the spacer region that lies between the 3' terminus of two genes that are transcribed toward one another on opposite DNA strands. These "shared repeats" are located at seventeen identical sites within H. akashiwo CCMP452 and NIES293 cpDNA including between psb A /psb C, psa C/ccs A, psa L/pet A, psa I/clp C, ycf 54/psb Y and ycf 30/pet J. CCMP452 has three additional sites. The observation of repeat sharing between two genes is similar to that seen in bacterial genomes where inverted repeats with stem lengths longer than eight nucleotide pairs are found most frequently in "short non-coding regions bounded by two 3' ends of convergent genes" [90]. Additionally, both H. akashiwo genomes have repeats, at 15 identical sites, that lie in the spacer region between genes that are transcribed on the same DNA strand. In some cases, inverted repeats overlap with the genes themselves. The largest examples include overlaps at the 3' end of psb I (20 bp), psa I (36 bp), pet D (39 bp), and dna K (24 bp). Repeats are also found internal to genes. CCMP452 orf 97 (Heak452_Cp014), which overlaps chl I, contains a perfect 24 base pair tandem repeat. This repeat is located 61 bases 5' to the ATG start of chl I [34]. A tandem repeat is also found within the 3' terminus of rpo B (CCMP452, 26 bp; NIES293, 36 bp).

Dispersed repeats occur in both H. akashiwo CCMP452 and NIES293 chloroplast genomes, but they are of low similarity and number (less than 100 total dispersed repeats greater than 90% similarity). The largest and most similar of these are conserved between the two H. akashiwo genomes. These elements are likely to have limited influence on recombination, unlike those observed for Chlamydomonas reinhardtii [52].

Though repeats are found in rhodophytic chloroplast genomes and other chloroplast genomes of secondary endosymbiotic origin, they are often present at a much lower frequency than that seen in H. akashiwo (Table 4). The glaucophyte Cyanophora paradoxa and the thermo-tolerant unicell, C. merolae, appear to be exceptions to this observation. The former retains high numbers of both tandem and inverted motifs while the latter appears to have retained almost exclusively tandem arrays.

It was of interest to determine whether a repeat structure is associated with a specific gene and whether that association is maintained among chloroplast genomes that maintain regional, but little global (Fig. 5), gene co-linearity. Notably, genes encoding cytochromes appear to be targeted for repeat embellishment. In H. akashiwo an inverted repeat is found within the 3' spacer of all pet genes (except pet L) and the gene css A, which encodes a cytochrome assembly protein [91]. This pattern of inverted repeat localization for the cytochrome complex is partially maintained in all the taxa examined in Table 5. Also striking is the uniformity of repeat placement among many taxa in the 3' spacer adjacent to rbc S, rps 10, and atp A genes. For example in the glaucophyte C. paradoxa not only is an inverted repeat associated with the 3' termini of pet A, B/D, F, G, L, rbc S, and atp A, but a 3' inverted repeat remains associated with rps 10 even though the "ribosomal protein block" is significantly disrupted in this chloroplast genome. Maintenance of repeat association with a specific gene is particularly notable in a genome such as P. purpurea, which has many coding genes (253) and few repeats (11). In this red algal chloroplast genome, the probability of finding an inverted repeat in the 3' spacer of any one gene is approximately 4.3%. Selective placement of specific repeats may extend beyond the rhodophytes and algae with chloroplast genomes of secondary endosymbiotic origin. For example, although rbc S is nuclear-localized in terrestrial plants and green algae, in those cases, the remaining chloroplast-encoded rbc L gene is usually followed by a repeat element in its 3' intergenic region.

Table 5 Conservation of gene-associated repeats

The highly conserved association of a secondary element with a specific gene in one taxon may offer clues for its function in others. For example, both strains of H. akashiwo retain a tandem repeat (77 bp) and an inverted repeat (212 bp) in the spacer 5' to rpl 3, which is the first gene in the putative ribosomal operon. Like bacteria [84], chloroplasts [85] transcribe the approximately 30 genes within this motif as a single transcript. Disruption of the E. coli inverted repeat structure that lies 50 bp upstream of the rpl 3 gene eliminates the transcription of this operon [92]. Well-documented information is available concerning the impact on terrestrial plant and green algal chloroplast mRNA function by the presence of inverted repeats within both the 5' and 3'UTR of a gene [93–95]. There is no doubt that intergenic regions contain significant information critical to organelle function. As more chloroplast genome sequences become available, we may find it just as instructive to compare and catalogue these domains, as it is to compare "coding" domains.

Conclusion

The fosmid-cloning-based chloroplast genome sequencing approach described here allows chloroplast genomic analysis for algal species that would be refractory to conventional organellar DNA isolation and analysis. In this study, we have presented new information on the chloroplast genome architecture and function in the stramenopile class raphidophyceae. Our ongoing studies target additional underrepresented stramenopile taxa for chloroplast genome analysis. The generated data will help resolve evolutionary patterns and provide insight into the mechanisms of chloroplast genome function within this marginally analyzed taxon.

Methods

Algal growth and strains

H. akashiwo (Hada) Hada ex Hara et Chihara strains CCMP452 and NIES293 were used in this study. CCMP452 was isolated from Long Island sound in 1952 and is commercially available from the Provasoli-Guillard National Center for Culture of Marine Phytoplankton; NIES293, isolated from Onagawa Bay, Japan in 1984, is from the collection of the National Institute for Environmental Studies in Japan. Vegetative cultures were axenically maintained on an artificial sea-water (O-3 medium) as previously described [50, 96]. One-liter cultures were grown in 2.8 liter Fernbach flasks with continuous rotary shaking at 60 rpm under 60 μmol Q m-2s-1 cool white light on a 12 hr light: 12 hr dark (diel) photoperiod. Cells were counted using a Coulter Counter (model ZBI, Coulter Electronics Inc., Hialeah, Fla.) equipped with a 100 × 120 μm aperture. All cultures were tested for fungal and bacterial contamination by inoculating 1 ml of H. akashiwo culture into 5.0 ml of a medium containing 2.0 g of nutrient broth (Difco laboratories, MI) and 1.25 g yeast extract in 0.25 liter of O-3 algal growth medium.

Chloroplast DNA purification

cpDNA from H. akashiwo CCMP452 was purified using a modified Hoescht dye/CsCl technique [97–99]. Pellets of approximately 6 × 108 late logarithmically growing cells (roughly 2 L of culture per pellet), were resuspended in 20 ml of 50 mM Tris- 50 mM EDTA buffer, pH 8.0 (TE buffer) at 5°C, after which 1 ml SDS (20% SDS in TE buffer) was added. After gentle mixing, 0.25 ml of Hoescht dye (10 mg/ml dH2O) was added, the mixture was placed on ice for 5 min, then 20 g of solid CsCl was added. When the salt dissolved, the refractive index was adjusted to 1.398. The solution was centrifuged using a Beckman Ti70.1 fixed angle rotor at 45,000 rpm for 20 hrs at 20°C. This centrifugation step separates the nuclear (highest density), mitochondrial (middle density) and chloroplast (lowest density) DNAs according to their different %G+C content. cpDNA, visualized by UV light, was recovered by puncturing the centrifuge tube wall using a 20-gauge needle. cpDNA fractions were pooled into a 5.0 ml tube, the refractive index readjusted to 1.3080 and the solution centrifuged for 20 hrs at 45,000 rpm and 20°C in a vertical Beckman Vti65.2 rotor. This last step was repeated until a single, clean cpDNA band was recovered. Hoescht dye was removed by adding to the DNA/CsCl solution an equal volume of isopropanol that was extracted with NaCl- saturated TE buffer. The isopropanol wash was repeated 10 times. To remove salts, the cpDNA solution was dialyzed (22 mm snake skin dialysis tubing, Pierce, Rockford, Il) overnight with stirring at 4°C against 2 liters of TE buffer. To concentrate the DNA solution, 100% butanol was added (0.5 ml butanol:1 ml DNA solution), the alcohol discarded, and the process repeated until the final DNA solution was reduced to approximately 0.5 ml. cpDNA was precipitated by the addition of 50 μl of 3 M sodium acetate (in H2O, pH 6.0) and 1 ml of 95% ethanol. The purified cpDNA was stored at -20°C until use. Approximately 80 liters of culture were harvested to retrieve sufficient cpDNA (10 μg) for the conventional shotgun sequencing protocol (about 15 cpDNA purification runs).

Total genomic DNA purification

Total high molecular weight DNA was extracted for long PCR and for fosmid library construction using Qiagen Genomic-Tip kits (100 G or 500 G) according to manufacturer's directions (Qiagen, Valencia, CA, USA). Briefly, H. akashiwo cells, grown to a density of 1.3 × 105 cells/ml, were harvested by centrifugation at 1,000 × g for 5 min. Cells were resuspended at a density of 8.7 × 105 cells/ml in 20 ml of cold lysis buffer (20 mM EDTA, 10 mM TrisCl, pH 8, 1% Triton X, 500 mM Guanidine-HCl, and 200 mM NaCl). The lysed cell suspension was incubated at 37°C for 1 hour with gentle agitation. The DNA was further treated with RNAse (20 μ g/ml) for 30 minutes at 37°C followed by Proteinase K (0.8 mg/ml) treatment for 2 h at 50°C with gentle agitation. To remove cell debris, the lysed cell suspension was pelleted by centrifugation at 9,750 × g for 20 min and the cleared lysate was removed. Three ml of the lysate were added to each Qiagen Genomic tip, previously equilibrated with QBT (750 mM NaCl, 50 mM MOPS, pH 7.0, 15% isopropanol, 0.15% Triton). The columns were washed twice with 10 ml of buffer QC (1.0 M NaCl, 50 mM MOPS, pH 7.0, 15% isopropanol). DNA was eluted from the genomic tip with buffer QF (1.25 M NaCl, 50 mM Tris-Cl, pH 8.5, 15% isopropanol) and precipitated by the addition of 0.7 volume of room-temperature isoproponal. The DNA was pelleted by centrifugation at 9,750 × g for 20 minutes. This pellet was then washed with 4 ml of cold 70% ethanol, and centrifuged at 9,750 × g for 10 minutes, before the supernatant was removed and the pellet air-dried. The pellet was resuspended in a total of 1 ml of Tris-Cl, pH 8.5. A single round of total DNA purification from 2 L of culture produced sufficient DNA (50 μg) to make a fosmid library.

Shotgun library preparation, DNA sequencing and genome assembly

DNA (CsCl-purified, or cosmid or fosmid clones) was sheared to 3–5 kb fragments using a Hydra-Shear (GeneMachines Inc. USA), and transformed into a blunt-ended pUC18 library, using 100 μg/mL carbenicillin and X-Gal/IPTG on for selection on solid agar bioassay plates (Nunc #240845). White colonies were picked using a Q-pix automated colony picker (Genetix Ltd. UK) and inoculated into 384-well freezing plates (Genetix cat# X7001) using UWGC freezing medium (10 g/L tryptone, 5 g/L yeast extract, 10 g/L NaCl, 6.3 g/L K2HPO4, 1.8 g/L KH2PO4, 0.5 g/L sodium citrate, 0.9 g/L (NH4)2SO4, 4.4% glycerol, 100 μg/mL Carbenicillin). Templates were amplified using TempliPhi (Amersham/GE USA), and sequenced according to standard protocols using the Big Dye Terminator reagent BDT v3.1 (0.25 μL per reaction). Sequencing reactions were analyzed using ABI 3730 automated sequencers (Applied Biosystems USA). Sequencing reads were processed using the phred/Phrap/consed package of base-calling, sequence assembly, and finishing/editing software [100–103].

Long PCR

To determine the orientation of the LSC relative to the SSC, four primers were designed based on H. akashiwo CCMP452 cpDNA sequence obtained from shot-gun cloning. The primers were designed to the unique regions of the chloroplast genome and were used to amplify cpDNA from the SSC region through the IR to the LSC region. The primer set one ORAC 210 (5' cgatcgttaactagtggtacttgctgtc 3') and ORAC 214 (5' caatcagtggaacacaagcagtgaag 3') generates a ~28 kb fragment while primer set two, ORAC 212 (5' ccacgtttctatacgacagatttcgag 3') and ORAC 216 (5'catatgcatcagaaacccaaatacctg 3'), produces a ~29 kb product. These primers were also used in two alternate combinations: set three (ORAC 212; ORAC 214) and set four (ORAC 216 and ORAC 210) were expected to generate ~29 kb and ~26 kb PCR products respectively only if a second isomeric form of cpDNA was present.

The long PCR reactions were performed using the LA Taq™ PCR system from Takara Mirus Bio inc. (Madison, WI) in a 50 μ l reaction following the manufacturer's recommendations. The PCR reaction contained 1 X LA PCR™ buffer II (Mg2+ plus), 400 μ M of each dNTP, 200 nM each of the downstream and upstream primers, 2 U of Takara LA Taq™ and 280 ng of high molecular weight DNA. A negative control was performed for each primer set by excluding the DNA from the PCR reaction. The PCR reactions were mixed by pipetting, briefly centrifuged, then placed in the thermal cycler (Eppendorf Mastercycle Gradient) for an initial denaturation step at 94°C for 3 min followed by 29 cycles of 94°C for 30 sec, and 68°C for 20 min. After the 30th cycle, a final extension was performed at 68°C for 10 min. The size of the PCR products was estimated using Roche DNA molecular weight marker XV (Roche Applied Science, Indianapolis, In) on a 0.5% TAE gel (4.84 g/L Tris-Base, 1.1% glacial acetic acid, 1 mM EDTA, pH 8.5 plus 5 g/L electrophoresis-grade agarose) run at 10 volts for 60 h. The PCR products were cloned into Expand Vector III vector using the Expand Cloning Kit from Roche according to the manufacturer's instructions. The presence of inserts was confirmed using the restriction enzyme Not1 (Roche). The four unique cosmid clones were shotgun sequenced to confirm the orientation of the SSC and LSC regions relative to the IR.

Fosmid library construction, and end-sequencing

Large-insert fosmid clones were prepared from high molecular weight DNA as previously described [104]. Briefly, sheared (45 kb) total cellular DNA was size-selected by agarose gel-electrophoresis using a DRIII CHEF gel apparatus (Bio-Rad, Hercules, CA), followed by end-repair and packaging into the PCC1Fos Vector, using the Epicentre CopyControl Fosmid Library Production Kit (Cat CCFOS110, Epicentre Biotechnologies, Madison, WI). Clones were plated after chloramphenicol selection, and picked using the Q-pix automated colony picker (Genetix Ltd. UK) and inoculated into 384-well freezing plates using UWGC freezing medium (defined above, under Shotgun library preparation, but with 12 ug/mL chloramphenicol as the antobiotic). Fosmid DNA was recovered using a standard alkaline-lysis protocol, and sequenced according to ABI manufacturer's directions, in an 8 μL reaction using 0.5 μL BDT version 3.1, 5 pmol of vector end-sequencing primers, and 100 ng DNA per reaction. Cycle sequencing was carried out in standard thermocycling conditions (3 min denature at 94°C, followed by 99 cycles of the following regime: 94°C 30 sec, 50°C 20 sec, 60°C 4 min), and analyzed on an ABI 3730 automated sequencer (ABI Biosystems, USA). Vector sequences were removed and sequences were further trimmed from both ends until a window of 12 bp with 90% of positions having a Phred score of Q20 or greater was reached. Sequences were compared using BLASTX to the GenBank non-redundant database and to a custom database consisting of published chloroplast genomes. Fosmids in which both end sequences had high quality matches (E value < 10-4) to a chloroplast gene as judged by both BLAST analyses were identified as chloroplast derived. All fosmid end sequences are available on our web site [105]. In addition to end-sequencing, six 384-well freezer plates of fosmids from the NIES293 library were screened using Real-Time PCR (RT-PCR) and assayed on an ABI 7900HT Sequence Detection System. PCR reactions were prepared using ABI Sybr Green PCR Master Mix (ABI Cat #4334973), and primer pairs designed to regions of the draft NIES293 genome (as well as the completed CCMP452 genome, since it was available). Primer pairs were standard oligonucleotide primers, designed to produce a 150 bp product. Reactions were inoculated using a 384-pin plastic plate replicator (ISC bio express cat# g32404) directly from the 384-well fosmid glycerol stock (see above). Positive clones were end-sequenced to confirm their identity, and sequenced by shotgun methods (see above).

Annotation

Open reading frames were initially predicted using Glimmer 2.0 [106] and then refined manually. The comparative RNA Database [107] was used to refine the locations of the ribosomal RNAs. Genes for tRNAs and tmRNAs were identified using tRNASCAN-SE [108]. SRPscan [109] was used to search for signal recognition particle RNAs. An initiator methionine tRNA was differentiated from the two elongator methionine tRNAs by identifying the conserved, characteristic nucleotide sequence of its anticodon loop (ttgggctcataacccga) using a chloroplast-specific tRNA data-base [110]. Predicted gene functions were assigned using a BLASTP search of the GenBank Non-Redundant database [78]. Conserved protein motifs were identified using the PFAM [111] database. BLASTP searches were used to identify orthologous genes (reciprocal best BLAST hits) in other chloroplast genomes. Tandem repeats were found with Tandem Repeat Finder [112] using default settings. Inverted repeats were found with E-inverted from the EMBOSS package [113] using the default settings and the additional constraint that repeats had to be more than 80% similar and the length of the loop shorter than the stem. Dispersed repeats were found using the cross-match function within Consed with the following parameters: minmatch = 12, minscore = 20, % similarity = 90%. A more stringent % similarity was used to filter out spurious repeats identified as extensions of more exact repeats. Additional dispersed repeats were found using pipMaker [114], using the default parameters and comparing each genome to itself. For analysis of the putative G-protein coupled receptor protein trans-membrane segment prediction was performed using the HMMTOP [115], TopPredII [116] and TMpred [117] programs. Global synteny analysis and SNP identification was performed using MUMMER [118]. Artemis and the Artemis Comparison Tool were used to visualize the comparative genome architecture and localization of SNPs [119, 120]. Circular genome maps were created with CGview [121]. All genome data used in this manuscript may be accessed through our publicly available website [105].

Abbreviations

cpDNA:

Chloroplast DNA

IR:

Inverted Repeat

LSC:

Large Single Copy

SSC:

Small Single Copy

PCR:

Polymerase Chain Reaction

bp:

base pairs

References

  1. Daugbjerg N, Andersen RA: Phylogenetic analyses of the rbc L sequences from haptophytes and heterokontalgae suggest their chloroplasts are unrelated. Mol Biol Evol. 1997, 14: 1242-1251.

    PubMed  Google Scholar 

  2. Medlin LK, Kaczmarska I: Evolution of the diatoms V: morphological and cytological support for the major clades and a taxonomic revision. Phycologia. 2004, 43: 245-270.

    Google Scholar 

  3. Bolin B, Degens ET, Duvigneaurd P, Kempe S: The global biogeochemical carbon cycle. The Global Carbon Cycle. Edited by: Bolin B, Degens ET, Kempe S, Ketner P. 1977, New York: J. Wiley & Sons, 1-53.

    Google Scholar 

  4. Jickells TD: Emissions from the oceans to the atmosphere, deposition from the atmosphere to the oceans and the interactions between them. Challenges of a changing earth Proceedings of the global change open science conference. Edited by: Steffen W, Jager J, Carson DJ, Bradshaw C. 2002, Amsterdam, The Netherlands: Springer-Verlag

    Google Scholar 

  5. Li WK: Primary production of prochlorophytes, cyanobacteria, and eukaryotic phytoplankton: Measurements from flow cytometric sorting. Limnol Oceanogr. 1994, 39: 169-175.

    Google Scholar 

  6. Nelson DM, Treguer P, Brzezinski MA, Leynaert A, Queguiner B: Production and dissolution of biogenic silica in the ocean – revised global estimates, comparison with regional data and relationship to biogenic sedimentation. Global Biogeochemical Cycles. 1995, 9: 359-372. 10.1029/95GB01070.

    Google Scholar 

  7. Simó R: Production of atmospheric sulfur by oceanic plankton: biogeochemical, ecological and evolutionary links. Trends Ecol Evol. 2001, 16: 287-294. 10.1016/S0169-5347(01)02152-8.

    PubMed  Google Scholar 

  8. Buskey EJ, Wysor B, Hyatt C: The role of hypersalinity in the persistence of the Texas 'brown tide' in the Laguna Madre. J Plankton Res. 1998, 20 (8): 1553-1565. 10.1093/plankt/20.8.1553.

    Google Scholar 

  9. Haigh R, Taylor FJ: Distribution of potentially harmful phytoplankton species of the northern Strait of Georgia, British Columbia (Canada). Can J Fish Aquatic Sci. 1990, 47: 2339-2350.

    Google Scholar 

  10. Liu H, Laws EA, Villareal TA, Buskey EJ: Nutrient-limited growth of Aureoumbra lagunensis (Pelagophyceae) with implications for its capability to outgrow other phytoplankton species in phosphate-limited environments. Journal of Phycology. 2001, 37 (4): 500-10.1046/j.1529-8817.2001.037004500.x.

    Google Scholar 

  11. Gordon R, Parkinson J: Potential roles for diatomists in nanotechnology. J Nanosci Nanotechnol. 2005, 5: 35-40. 10.1166/jnn.2005.002.

    PubMed  Google Scholar 

  12. Falkowski PG, Katz ME, Knoll AH, Quigg A, Raven JA, Schofield O, Taylor FJR: The evolution of modern eukaryotic phytoplankton. Science. 2004, 305: 354-360. 10.1126/science.1095964.

    PubMed  Google Scholar 

  13. Gibbs SP: The chloroplasts of some algal groups may have evolved from endosymbiotic eucaryotic algae. Ann N Y Acad Sci. 1981, 361: 193-208. 10.1111/j.1749-6632.1981.tb46519.x.

    PubMed  Google Scholar 

  14. Gibbs SP: The chloroplast endoplasmic reticulum: Structure, function, and evolutionary significance. Int Rev Cytol. 1981, 72: 19-99.

    Google Scholar 

  15. Keeling PJ: Diversity and evolutionary history of plastids and their hosts. American Journal of Botany. 2004, 91: 1481-1493. 10.3732/ajb.91.10.1481.

    PubMed  Google Scholar 

  16. Stiller JW, Reel DC: A single origin of plastids revisited: convergent evolution in organellar genome content. J Phycol. 2003, 39: 95-105. 10.1046/j.1529-8817.2003.02070.x.

    Google Scholar 

  17. Bjornland T, Liaaen-Jensen S: Distribution patterns of carotenoids in relation to chromophyte phylogeny and systematics. The Chromophyte Algae: Problems and Perspectives. Edited by: Green JC, Leadbeater BSC, Diver WL. 1989, Oxford: Clarendon Press, 38: 37-60.

    Google Scholar 

  18. Kowallik KV, Stroebe B, Schaffran I, Freier U: The chloroplast genome of a chlorophyll a+c-containing alga, Odontella sinensis. Plant Molecular Biology Reporter. 1995, 13: 336-342. 10.1007/BF02669188.

    Google Scholar 

  19. Armbrust EV, Berges JA, Bowler C, Green BR, Martinez D, Putnam NH, Zhou S, Allen AE, Apt KE, Bechner M, Brzezinski MA, Chaal BK, Chiovitti A, Davis AK, Demarest MS, Detter JC, Glavina T, Goodstein D, Hadi MZ, Hellsten U, Hildebrand M, Jenkins BD, Jurka J, Kapitonov VV, Kroger N, Lau WWY, Lane TW, Larimer FW, Lippmeier JC, Lucas S, 15 others: The Genome of the Diatom Thalassiosira Pseudonana : Ecology, Evolution, and Metabolism. Science. 2004, 306 (5693): 79-86. 10.1126/science.1101156.

    PubMed  Google Scholar 

  20. Oudot-Le Secq MP, Grimwood J, Shapiro H, Armbrust EV, Bowler C, Green BR: Chloroplast genomes of the diatoms Phaeodactylum tricornutum and Thalassiosira pseudonana: comparison with other plastid genomes of the red lineage. Mol Genet Genomics. 2007, 277 (4): 427-439. 10.1007/s00438-006-0199-4.

    PubMed  Google Scholar 

  21. Aldrich JK, Cattolico RA: Isolation and characterization of chloroplast DNA from the marine chromophyte Olisthodiscus luteus : Electron microscopic visualization of isomeric molecular forms. Plant Physiology. 1981, 68: 641-647.

    PubMed  PubMed Central  Google Scholar 

  22. Jansen RK, Raubeson LA, Boore JL, DePamphilis CW, Chumley TW, Haberle RC, Wyman SK, Alverson AJ, Peery R, Herman SJ, Fourcade HM, Kuehl JV, McNeal JR, Leebens-Mack J, Cui L: Methods for obtaining and analyzing whole chloroplast genome sequences. Methods in Enzymology. 2005, 395: 348-384.

    PubMed  Google Scholar 

  23. McNeal JR, Leebens-Mack JH, Arumuganathan K, Kuehl JV, Boore JL, DePamphilis CW: Using partial genomic fosmid libraries for sequencing complete organellar genomes. Biotechniques. 2006, 41 (1): 69-73.

    PubMed  Google Scholar 

  24. Bearon RN, Grunbaum D, Cattolico RA: Relating cell-level swimming behaviors to vertical population distributions in Heterosigma akashiwo (Raphidophyceae), a harmful alga. Limnology and Oceanography. 2004, 49 (2): 607-613.

    Google Scholar 

  25. Bowers HA, Tomas C, Tengs T, Kempton JW, Lewitus AJ, Oldach DW: Raphidophyceae [Chadefaud ex Silva] systematics and rapid identification: sequence analyses and real-time PCR assays. Journal of Phycology. 2006, 42: 1333-1348. 10.1111/j.1529-8817.2006.00285.x.

    PubMed  PubMed Central  Google Scholar 

  26. Cattolico RA, Boothroyd JC, Gibbs SP: Synchronous growth and plastid replication in the naturally wall-less alga Olisthodiscus luteus. Plant Physiol. 1976, 57: 497-503.

    PubMed  PubMed Central  Google Scholar 

  27. Han MS, Furuya K: Size and species-specific primary productivity and community structure of phytoplankton in Tokyo Bay. Journal of Plankton Research. 2000, 22: 1221-1235. 10.1093/plankt/22.7.1221.

    Google Scholar 

  28. Hashimoto H: Electron-opaque annular structure girdling the constricting isthmus of the dividing chloroplasts of Heterosigma akashiwo (Raphidophyceae, Chromophyta). Protoplasma. 1997, 197 (3–4): 210-216. 10.1007/BF01288030.

    Google Scholar 

  29. Bearon RN, Grunbaum D, Cattolico RA: Effects of salinity structure on swimming behavior and harmful algal bloom formation in Heterosigma akashiwo, a toxic raphidophyte. Mar Ecol Prog Ser. 2006, 306: 153-163. 10.3354/meps306153.

    Google Scholar 

  30. Ersland DR, Cattolico RA: Nuclear deoxyribonucleic acid characterization of the marine chromophyte Olisthodiscus luteus. Biochemistry. 1981, 20 (24): 6886-6893. 10.1021/bi00527a023.

    PubMed  Google Scholar 

  31. Miyagi N, Satoh S, Fujii T: A nitrate-inducible plasma membrane protein of a marine alga, Heterosigma akashiwo. Plant & Cell Physiology. 1992, 33 (7): 971-976.

    Google Scholar 

  32. Chaal BK, Ishida K, Green BR: A thylakoidal processing peptidase from the heterokont alga Heterosigma akashiwo. Plant Molecular Biology. 2003, 52: 463-472. 10.1023/A:1023900100803.

    PubMed  Google Scholar 

  33. Ishida KI, Cavalier-Smith T, Green B: Endomembrane structure and the chloroplast protein targeting pathway in Heterosigma akashiwo (Raphidophyceae, Chromista). J Phycol. 2000, 36: 1135-1144. 10.1046/j.1529-8817.2000.00071.x.

    Google Scholar 

  34. Valentin K, Fischer S, Cattolico RA: The chloroplast bchI gene encodes a subunit of magnesium chelatase in the marine heterokont alga Heterosigma carterae. Eur J Phycol. 1998, 33: 113-120. 10.1080/09670269810001736603.

    Google Scholar 

  35. Ono K, Khan S, Onoue Y: Effects of temperature and light intensity on the growth and toxicity of Heterosigma akashiwo (Raphidophyceae). Aquac Res. 2000, 31: 427-433. 10.1046/j.1365-2109.2000.00463.x.

    Google Scholar 

  36. Twiner MJ, Trick CG: Possible physiological mechanisms for production of hydrogen peroxide by the ichthyotoxic flagellate Heterosigma akashiwo. J Plankton Res. 2000, 22 (10): 1961-1975. 10.1093/plankt/22.10.1961.

    Google Scholar 

  37. Hariharan T, Johnson PJ, Cattolico RA: Purification and characterization of phosphoribulokinase from the marine chromophytic alga Heterosigma carterae. Plant Physiol. 1998, 117 (1): 321-329. 10.1104/pp.117.1.321.

    PubMed  PubMed Central  Google Scholar 

  38. Okamoto T, Kim D, Oda T, Matsuoka K, Ishimatsu A, Muramatsu T: Concanavalin A-induced discharge of glycocalyx of raphidophycean flagellates, Chattonella marina and Heterosigma akashiwo. Biosci Biotechnol Biochem. 2000, 64 (8): 1767-1770. 10.1271/bbb.64.1767.

    PubMed  Google Scholar 

  39. Shono M, Hara Y, Wada M, Fujii T: A sodium pump in the plasma membrane of the marine alga Heterosigma akashiwo. Plant & Cell Physiology. 1996, 37 (3): 385-388.

    Google Scholar 

  40. Cattolico RA: Variation in plastid number: effect on chloroplast and nuclear DNA complement in the unicellular alga Olisthodiscus luteus. Plant Physiology. 1978, 62: 558-562.

    PubMed  PubMed Central  Google Scholar 

  41. Satoh E, Watanabe MM, Fujii T: Photoperiodic regulation of cell division and chloroplast replication in Heterosigma akashiwo. Plant Cell Physiol. 1987, 28 (6): 1093-1099.

    Google Scholar 

  42. Doran E, Cattolico RA: Photoregulation of chloroplast gene transcription in the chromophytic alga Heterosigma carterae. Plant Physiol. 1997, 115: 773-781.

    PubMed  PubMed Central  Google Scholar 

  43. Reynolds AE, McConaughy BL, Cattolico RA: Chloroplast genes of the marine alga Heterosigma carterae are transcriptionally regulated during a light/dark cycle. Mol Mar Biol Biotech. 1993, 2: 121-128.

    Google Scholar 

  44. Aldrich JK, Gelvin S, Cattolico RA: Extranuclear DNA of a marine chromophytic alga: restriction enzyme analysis. Plant Physiology. 1982, 69: 1189-1195.

    PubMed  PubMed Central  Google Scholar 

  45. Ersland DR, Aldrich JK, Cattolico RA: Kinetic complexity, homogeneity and copy number of chloroplast DNA from the marine alga Olisthodiscus luteus. Plant Physiology. 1981, 68: 1468-1473.

    PubMed  PubMed Central  Google Scholar 

  46. Reith ME, Cattolico RA: The inverted repeat of Olisthodiscus luteus ctDNA contains the genes for both subunits of RuBPcase and the 32,000 QB protein: phylogenetic implication. Proc Natl Acad Sci USA. 1986, 83: 8599-8603. 10.1073/pnas.83.22.8599.

    PubMed  PubMed Central  Google Scholar 

  47. Shivji MS, Li N, Cattolico RA: Structure and organization of rhodophyte and chromophyte plastid genomes: implications for the ancestry of plastids. Mol Gen Genet. 1992, 232 (1): 65-73. 10.1007/BF00299138.

    PubMed  Google Scholar 

  48. Boczar B, Delaney T, Cattolico RA: Gene for the ribulose-1,5-biphosphate carboxylase small subunit protein of the marine chromophyte Olisthodiscus luteus is similar to that of a chemoautotrophic bacterium. Proc Natl Acad Sci USA. 1989, 86: 4996-4999. 10.1073/pnas.86.13.4996.

    PubMed  PubMed Central  Google Scholar 

  49. Duplessis MR, Karol KG, Adman ET, Choi LYS, Jacobs MA, Cattolico RA: Chloroplast His-to-Asp signal transduction: a potential mechanism for plastid gene regulation in Heterosigma akashiwo (Raphidophyceae). BMC Evolutionary Biology. 2007, 7: 70-10.1186/1471-2148-7-70.

    PubMed  PubMed Central  Google Scholar 

  50. Jacobs MA, Connell L, Cattolico RA: A conserved His-Asp signal response regulator-like gene in Heterosigma akashiwo chloroplasts. Plant Mol Biol. 1999, 41 (5): 645-655. 10.1023/A:1006394925182.

    PubMed  Google Scholar 

  51. Ki J-S, Han M-S: Nuclear rDNA and chloroplast rbcL, rbcS and IGS sequence data, and their implications from the Japanese, Korean, and North American harmful algae, Heterosigma akashiwo (Raphidophyceae). Environmental Research. 2007, 103: 299-304. 10.1016/j.envres.2006.08.014.

    PubMed  Google Scholar 

  52. Palmer JD: Chloroplast DNA exists in two orientations. Nature. 1983, 301: 92-93. 10.1038/301092a0.

    Google Scholar 

  53. Palmer JD: Comparative organization of chloroplast genomes. Ann Rev Genet. 1985, 19: 325-354. 10.1146/annurev.ge.19.120185.001545.

    PubMed  Google Scholar 

  54. Aldrich J, Cherney B, Merlin E, Williams C, Mets L: Recombination within the inverted repeat sequences of the Chlamydomonas reinhardii chloroplast genome produces two orientation isomers. Current Genetics. 1985, 9: 233-238. 10.1007/BF00420317.

    PubMed  Google Scholar 

  55. von Berg K-HL, Kowallik KV: Structural organization of the chloroplast genome of the chromophytic alga Vaucheria bursata. Plant Mol Biol. 1992, 18: 83-95. 10.1007/BF00018459.

    Google Scholar 

  56. Bourne CM, Palmer JD, Stoermer EF: Organization of the chloroplast genome of the freshwater centric diatom Cyclotella meneghiniana. J Phycol. 1992, 28: 347-355. 10.1111/j.0022-3646.1992.00347.x.

    Google Scholar 

  57. Douglas SE, Penny SL: The plastid genome of the cryptophyte alga, Guillardia theta : complete sequence and conserved synteny groups confirm its common ancestry with red algae. J Mol Evol. 1999, 48 (2): 236-244. 10.1007/PL00006462.

    PubMed  Google Scholar 

  58. Hagopian JC, Reis M, Kitajima JP, Bhattacharya D, Oliveira MCd: Comparative analysis of the complete plastid genome sequence of the red alga Gracilaria tenuistipitata var. liui provides insights into the evolution of rhodoplasts and their relationship to other plastids. J Mol Evol. 2004, 59: 464-477. 10.1007/s00239-004-2638-3.

    PubMed  Google Scholar 

  59. Stabile JE, Gallagher JC, Wurtzel ET: Colinearity of chloroplast genomes in divergent ecotypes of the marine diatom Skeletonema costatum (Bacillariophyta). J Phycol. 1995, 31: 795-800. 10.1111/j.0022-3646.1995.00795.x.

    Google Scholar 

  60. Reith M: Molecular biology of rhodophyte and chromophyte plastids. Annual Review of Plant Physiology and Plant Molecular Biology. 1995, 46: 549-575. 10.1146/annurev.pp.46.060195.003001.

    Google Scholar 

  61. Sanchez-Puerta MV, Bachvaroff TR, Delwiche CF: The complete plastid genome sequence of the haptophyte Emiliania huxleyi : a comparison to other plastid genomes. DNA Research. 2005, 12: 151-156. 10.1093/dnares/12.2.151.

    PubMed  Google Scholar 

  62. Li N, Cattolico RA: Chloroplast genome characterization in the red alga Griffithsia pacifica. Molec Gen Genet. 1987, 209: 343-351. 10.1007/BF00329664.

    PubMed  Google Scholar 

  63. Reith ME, Munholland J: A high-resolution gene map of the chloroplast genome of the red alga Porphyra purpurea. Plant Cell. 1993, 5: 465-475. 10.1105/tpc.5.4.465.

    PubMed  PubMed Central  Google Scholar 

  64. Kashdan MA, Dudock BS: The gene for spinach chloroplast Isoleucine tRNA has a Methionine anticodon. Journal of Biological Chemistry. 1982, 257 (19): 11191-11194.

    PubMed  Google Scholar 

  65. Avissar YJ, Beale SI: Biosynthesis of tetrapyrrole pigment precursors. Plant Physiol. 1988, 88: 879-886.

    PubMed  PubMed Central  Google Scholar 

  66. Beale SI: Biosynthesis of the tetrapyrrole pigment precursor, δ-aminolevulinic acid, from glutamate. Plant Physiol. 1990, 93: 1273-1279.

    PubMed  PubMed Central  Google Scholar 

  67. Nuiza L, Beale S: Physical and kinetic interaction between glutamyl-tRNA reductase and glutamate-1-semialdehyde aminotransferase of Chlamydomonas reinhardtii. J Biol Chem. 2005, 280: 24301-24307. 10.1074/jbc.M502483200.

    Google Scholar 

  68. Kumar R, Small I, Marechal-Drouad L, Akama K: Striking differences in mitochondrial RNA import between different plants. Mol Gen Genet. 1996, 252: 404-411.

    PubMed  Google Scholar 

  69. Esposito D, Scocca JJ: The integrase family of tyrosine recombinases: evolution of a conserved active site domain. Nucleic Acids Research. 1997, 25 (18): 3605-3614. 10.1093/nar/25.18.3605.

    PubMed  PubMed Central  Google Scholar 

  70. Abremski KE, Hoess RH: Evidence for a second conserved arginine residue in the integrase family of recombination proteins. Protein Engineering. 1992, 5 (1): 87-91. 10.1093/protein/5.1.87.

    PubMed  Google Scholar 

  71. Argos P, Landy A, Abremski K, Egan JB, Haggard-Ljungquist E, Hoess RH, Kahn ML, Kalionis B, Narayana SV, Piersonr LS: The integrase family of site-specific recombinases: regional similarities and global diversity. EMBO J. 1986, 5 (2): 433-440.

    PubMed  PubMed Central  Google Scholar 

  72. Sciochetti SA, Piggot PJ: A tale of two genomes: resolution of dimeric chromosomes in Escherichia coli and Bacillus subtilis. Res Microbiol. 2000, 151: 503-511. 10.1016/S0923-2508(00)00220-5.

    PubMed  Google Scholar 

  73. Nunes-Düby SE, Kwon HJ, Tirumalai RS, Ellenberger T, Landy A: Similarities and differences among 105 members of the Int family of site-specific recombinases. Nucleic Acids Res. 1998, 26 (2): 391-406. 10.1093/nar/26.2.391.

    PubMed  PubMed Central  Google Scholar 

  74. Sciochetti SA, Piggot PJ, Blakely GW: Identification and characterization of the dif site from Bacillus subtilis. J Bacteriol. 2001, 183 (3): 1058-1068. 10.1128/JB.183.3.1058-1068.2001.

    PubMed  PubMed Central  Google Scholar 

  75. Barre F-X, Søballe B, Michel B, Aroyo M, Robertson M, Sherratt D: Circles: the replication-recombination-chromosome segregation connection. PNAS. 2001, 98 (15): 8189-8195. 10.1073/pnas.111008998.

    PubMed  PubMed Central  Google Scholar 

  76. Blakely GW, Sherratt DJ: Interactions of the site-specific recombinases XerC and XerD with the recombination site dif. Nucleic Acids Research. 1994, 22 (25): 5613-5620. 10.1093/nar/22.25.5613.

    PubMed  PubMed Central  Google Scholar 

  77. Lesterlin C, Barre F, Cornet F: Genetic recombination and the cell cycle: What we have learned from chromosome dimers. Mol Microbiol. 2004, 54 (5): 1151-1160. 10.1111/j.1365-2958.2004.04356.x.

    PubMed  Google Scholar 

  78. Altschul SF, Madden TL, SchÄffer AA, Zhang J, Zhang Z, Miller W, Lipman DJ: Gapped BLAST and PSI-BLAST: A new generation of protein database search programs. Nucleic Acids Res. 1997, 25 (17): 3389-3402. 10.1093/nar/25.17.3389.

    PubMed  PubMed Central  Google Scholar 

  79. Kolbe M, Besir H, Essen L-O, Oesterhelt D: Structure of the light-driven chloride pump halorhodopsin at 1.8 Ã… resolution. Science. 2000, 288: 1390-1396. 10.1126/science.288.5470.1390.

    PubMed  Google Scholar 

  80. Leucke H, Schobert B, Lanyi JK, Spudich EN, Spudich JL: Crystal structure of sensory rhodopsin II at 2.4 angstroms: insights into color tuning and transducer interaction. Science. 2001, 293: 1499-1503. 10.1126/science.1062977.

    Google Scholar 

  81. Pebay-Peyroula E, Rummel G, Rosenbusch JP, Landau EM: X-ray structure of bacteriorhodopsin at 2.5 angstroms from microcrystals grown in lipidic cubic phases. Science. 1997, 277: 1676-1681. 10.1126/science.277.5332.1676.

    PubMed  Google Scholar 

  82. Pietrokovski S: A new intein in cyanobacteria and its significance for the spread of inteins. TIG. 1996, 12 (8): 287-288.

    PubMed  Google Scholar 

  83. Stoebe B, Kowallik KV: Gene-cluster analysis in chloroplast genomics. Trends in Genetics. 1999, 15 (9): 344-347. 10.1016/S0168-9525(99)01815-6.

    PubMed  Google Scholar 

  84. Li X, Lindahl L, Sha Y, Zengel JM: Analysis of the Bacillus subtilis S10 ribosomal protein gene cluster identifies two promoters that may be responsible for transcription of the entire 15-kilobase S10-spc-α cluster. J Bacteriol. 1997, 179 (22): 7046-7054.

    PubMed  PubMed Central  Google Scholar 

  85. Wang SL, Liu XQ, Douglas SE: The large ribosomal protein gene cluster of a cryptomonad plastid: gene organization, sequence and evolutionary implications. Biochem Mol Biol Int. 1997, 41 (5): 1035-1044. 10.1080/15216549700202101.

    PubMed  Google Scholar 

  86. Pombert J-F, Lemieux C, Turmel M: The complete chloroplast DNA sequence of the green alga Oltmannsiellopsis viridis reveals distinctive quadripartite architecture in the chloroplast genome of early diverging ulvophytes. BMC Biol. 2006, 4: 15-10.1186/1741-7007-4-3.

    Google Scholar 

  87. Turmel M, Otis C, Lemieux C: The complete chloroplast DNA sequences of the charophycean green algae Staurastrum and Zygnema reveal that the chloroplast genome underwent extensive changes during the evolution of the Zygnematales. BMC Biol. 2005, 3: 22-10.1186/1741-7007-3-22.

    PubMed  PubMed Central  Google Scholar 

  88. Lakeman MB, Cattolico RA: Cryptic diversity in phytoplankton cultures is revealed using a simple plating technique. Journal of Phycology. 2007, 43: 662-774. 10.1111/j.1529-8817.2007.00364.x.

    Google Scholar 

  89. Lee KH, Cattolico RA: Putative RuBisCo activase CfxQ in the toxic alga Heterosigma akashiwo. J Phycol PSA Abstracts. 2006, 42: 1-48.

    Google Scholar 

  90. Lillo F, Basile S, Mantegna RN: Comparative genomics study of inverted repeats in bacteria. Bioinformatics. 2002, 18: 971-979. 10.1093/bioinformatics/18.7.971.

    PubMed  Google Scholar 

  91. Xie Z, Merchant S: The plastid-encoded ccsA gene is required for heme attachment to chloroplast c-type cytochromes. J Biol Chem. 1996, 271 (9): 4632-4639. 10.1074/jbc.271.9.4632.

    PubMed  Google Scholar 

  92. Shen P, Zengel JM, Lindahl L: Secondary structure of the leader transcript from the Escherichia coli S10 ribosomal protein operon. Nucleic Acids Research. 1988, 16 (18): 8905-8924. 10.1093/nar/16.18.8905.

    PubMed  PubMed Central  Google Scholar 

  93. Anthonisen IL, Salvador ML, Klein U: Specific sequence elements in the 5' untranslated regions of rbc L and atp B gene mRNAs stabilize transcripts in the chloroplast of Chlamydomonas reinhardtii. RNA. 2001, 7: 1024-1033. 10.1017/S1355838201001479.

    PubMed  PubMed Central  Google Scholar 

  94. Bollenbach TJ, Stern DB: Secondary structures common to chloroplast mRNA 3' UTRs direct cleavage by CSP41, an endoribonuclease belonging to the short chain dehydrogenase/reductase superfamily. J Biol Chem. 2003, 278 (28): 25832-25838. 10.1074/jbc.M303559200.

    PubMed  Google Scholar 

  95. Hayes R, Kudla J, Schuster G, Gabay L, Maliga P, Gruissem W: Chloroplast mRNA 3'-end processing by a high molecular weight protein complex is regulated by nuclear encoded RNA binding proteins. EMBO J. 1996, 15 (5): 1132-1141.

    PubMed  PubMed Central  Google Scholar 

  96. McIntosh L, Cattolico RA: Preservation of algal and higher plant ribosomal RNA integrity during extraction and electrophoretic quantitation. Anal Biochem. 1978, 91 (2): 600-612. 10.1016/0003-2697(78)90546-8.

    PubMed  Google Scholar 

  97. Aldrich JK, Cattolico RA: Isolation and Characterization of chloroplast DNA from the marine chromophyte Olithodiscus luteus : Electron microscopic visualization of isomeric molecular forms. Plant Physiol. 1981, 68: 641-647.

    PubMed  PubMed Central  Google Scholar 

  98. Delaney T, Cattolico RA: Chloroplast ribosomal DNA organization in the chromopytic alga Olisthodiscus luteus. Curr Genet. 1989, 15: 221-229. 10.1007/BF00435509.

    PubMed  Google Scholar 

  99. Li N, Cattolico RA: Identification of chloroplast DNA heterogeneity using field inversion gel electrophoresis. Curr Genet. 1991, 20: 157-159. 10.1007/BF00312779.

    Google Scholar 

  100. Ewing B, Green P: Base-calling of automated sequencer traces using Phred: Error probabilities. 1998, Cold Spring Harbor Laboratory Press, 8: 186-194.

    Google Scholar 

  101. Gordon D, Abajian C, Green P: A graphical tool for sequence finishing. 1998, Cold Spring Harbor Laboratory Press, 8: 195-202.

    Google Scholar 

  102. Nickerson DA, Tobe TO, Taylor SL: Polyphred: automating the detection and genotyping of single nucleotide substitutions using fluorescence-based resquencing. Nucleic Acids Res. 1997, 25 (14): 2745-2741. 10.1093/nar/25.14.2745.

    PubMed  PubMed Central  Google Scholar 

  103. Rieder MJ, Taylor SL, Tobe V, Nickerson DA: Automating the identification of DNA variations using quality-based fluorescence re-sequencing: Analysis of the human mitochondrial genome. 1998, Oxford University Press, 26: 967-973.

    Google Scholar 

  104. Raymond CK, Subramanian S, Paddock M, Qiu R, Deodato C, Palmieri A, Chang J, Radke T, Haugen E, Kas A, Waring D, Bovee D, Stacy R, Kaul R, Olson MV: Targeted, haplotype-resolved resequencing of long segments of the human genome. Genomics. 2005, 86 (6): 759-766. 10.1016/j.ygeno.2005.08.013.

    PubMed  Google Scholar 

  105. The Stramenopile Chloroplast Genome Database. [http://chloroplast.ocean.washington.edu]

  106. Delcher A, Harmon D, Kasif S, White O, Salzberg S: Improved microbial gene identification with GLIMMER. Nucleic Acids Research. 1999, 27: 4636-4641. 10.1093/nar/27.23.4636.

    PubMed  PubMed Central  Google Scholar 

  107. Cannone JJ, Subramanian S, Schnare MN, Collett JR, D'Souza LM, Du Y, Feng B, Lin N, Madabusi LV, Müller KM, Pande N, Shang Z, Yu N, Gutell RR: The comparative RNA web (CRW) site: An online database of comparative sequence and structure information for ribosomal, intron, and other RNAs. Biomed Central Bioinformatics. 2002, 3: 2-

    PubMed  PubMed Central  Google Scholar 

  108. Lowe TM, Eddy SR: tRNAscan-SE: A program for improved detection of transfer RNA genes in genomic sequence. Nucleic Acids Research. 1997, 25: 955-964. 10.1093/nar/25.5.955.

    PubMed  PubMed Central  Google Scholar 

  109. Regalia M, Rosenblad MA, Samuelsson T: Prediction of signal recognition particle RNA genes. Nucleic Acids Research. 2002, 30 (15): 3368-3377. 10.1093/nar/gkf468.

    PubMed  PubMed Central  Google Scholar 

  110. Kurihara K, Kunisawa T: A gene order database of plastid genomes. Data Science Journal. 2004, 3: 60-79. 10.2481/dsj.3.60.

    Google Scholar 

  111. Finn RD, Mistry J, Schuster-Böckler B, Griffiths-Jones S, Hollich V, Lassmann T, Moxon S, Marshall M, Khanna A, Durbin R, Eddy SR, Sonnhammer EL, Bateman A: Pfam: clans, web tools and services. Nucleic Acids Res. 2006, 34 (Database issue): D247-D251. 10.1093/nar/gkj149.

    PubMed  PubMed Central  Google Scholar 

  112. Benson G: Tandem repeats finder: a program to analyze DNA sequences. Nucleic Acids Res. 1999, 27 (2): 573-580. 10.1093/nar/27.2.573.

    PubMed  PubMed Central  Google Scholar 

  113. Rice P, Longden I, Bleasby A: The European Molecular Biology Open Software Suite. Trends in Genetics. 2000, 16 (6): 276-277. 10.1016/S0168-9525(00)02024-2.

    PubMed  Google Scholar 

  114. Schwartz S, Zhang Z, Frazer K, Smit A, Riener C, Bouck J, Gibbs R, Hardison R, Miller W: PipMaker – A web server for aligning two genomic DNA sequences. Genome Research. 2000, 10 (4): 577-586. 10.1101/gr.10.4.577.

    PubMed  PubMed Central  Google Scholar 

  115. Tusnady GE, Simon I: The HMMTOP transmembrane topology prediction server. Bioinformatics. 2001, 17: 849-850. 10.1093/bioinformatics/17.9.849.

    PubMed  Google Scholar 

  116. Claros MG, Heijne Gv: TopPred II: an improved software for membrane protein structure predictions. CABIOS. 1994, 10: 685-686.

    PubMed  Google Scholar 

  117. Hofmann K, Stoffel W: TMbase-a database of membrane spanning protein segments. Biol Chem Hoppe-Seyler. 1993, 166-374

  118. Kurtz S, Phillippy A, Delcher AL, Smoot M, Shumway M, Antonescu C, Salzberg SL: Versatile and open software for comparing large genomes. Genome Biology. 2004, 5: R12-10.1186/gb-2004-5-2-r12.

    PubMed  PubMed Central  Google Scholar 

  119. Carver TJ, Rutherford KM, Berriman M, Rajandream MA, Barrell BG, Parkhill J: ACT: The Artemis Comparison Tool. Bioinformatics. 2005, 21: 3422-3423. 10.1093/bioinformatics/bti553.

    PubMed  Google Scholar 

  120. Rutherford K, Parkhill J, Crook J, Horsnell T, Rice P, Rajandream MA, Barrell B: Artemis: sequence visualization and annotation. Bioinformatics. 2000, 16: 944-945. 10.1093/bioinformatics/16.10.944.

    PubMed  Google Scholar 

  121. Stothard P, Wishart DS: Circular genome visualization and exploration using CGView. Bioinformatics. 2005, 21: 537-539. 10.1093/bioinformatics/bti054.

    PubMed  Google Scholar 

Download references

Acknowledgements

We wish to thank Maynard Olson for helping RAC and MJ initiate this project; Quin Tu who generated all purified cpDNA for the conventional cloning of Heterosigma akashiwo CCMP452, Jean M. Velluppillai who initially identified and analyzed the putative G-protein coupled receptor; Kathy Charing for support in repeat analysis; Chloe Deodato, Andrea Kunkle and Mary Nicholson for help in strain verification. William Hatheway and Molly Brown provided editorial and library assistance. This study was supported by Washington Sea Grants NA16RG1044 AM09 and NA040AR170032 to RAC and NSF 0523756 to GR and RAC. JMV and Han Ong were supported by a NIH Vision Research pre-doctoral training grant T32 EY07031 and an NIH/NHGRI Genome Training Grant T32 HG00035 post-doctoral award respectively.

RAC dedicates this manuscript to Sarah Gibbs who recognized Heterosigma akashiwo (aka Olisthodiscus luteus) as a unique system to study chloroplast biogenesis, provided many stimulating discussions, and served as a dynamic role model.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Rose Ann Cattolico.

Additional information

Authors' contributions

RAC conceived the study, performed the analysis of TyrC, determined repeat placement in cpDNAs and wrote a major portion of the manuscript. MAJ developed the application of fosmid cloning technology to chloroplast sequencing, refined fosmid end-sequencing protocols, designed custom PCR for genome finishing and fosmid screening, and contributed to manuscript writing. JC produced both fosmid and shotgun libraries, and ran DNA quality analyses. MD isolated cpDNA used in the conventional cloning of cpDNA, did the initial annotation of the Heterosigma CCMP452 genome, as well as verified the presence of isomeric cpDNAs using long PCR. TL performed analysis on the putative G-protein coupled receptor. JM was responsible for genome analysis software development. HCO conducted the sequence alignment of proteins containing large inserts and showed that these inserts were contained in the mature RNAs. ES developed Sybr screening method for chloroplast fosmid retrieval, did fosmid end-sequencing, and DNA preparations. YZ was responsible for genome sequence finishing, and quality check on completed sequences. GR developed the bioinformatic screen of fosmid end-sequences, completed the final annotation of both genomes, performed the comparative genomic analyses (SNPs and genome synteny) and contributed to manuscript writing.

Authors’ original submitted files for images

Rights and permissions

This article is published under license to BioMed Central Ltd. This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article

Cattolico, R.A., Jacobs, M.A., Zhou, Y. et al. Chloroplast genome sequencing analysis of Heterosigma akashiwo CCMP452 (West Atlantic) and NIES293 (West Pacific) strains. BMC Genomics 9, 211 (2008). https://doi.org/10.1186/1471-2164-9-211

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1471-2164-9-211

Keywords