Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Extensive chloroplast genome rearrangement amongst three closely related Halamphora spp. (Bacillariophyceae), and evidence for rapid evolution as compared to land plants

  • Sarah E. Hamsher ,

    Roles Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Project administration, Validation, Visualization, Writing – original draft, Writing – review & editing

    hamshers@gvsu.edu

    Affiliations Department of Biology, Grand Valley State University, Allendale, Michigan, United States of America, Annis Water Resources Institute, Grand Valley State University, Muskegon, Michigan, United States of America

  • Kyle G. Keepers,

    Roles Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Project administration, Software, Validation, Visualization, Writing – original draft, Writing – review & editing

    Affiliation Department of Ecology and Evolutionary Biology, University of Colorado, Boulder, Colorado, United States of America

  • Cloe S. Pogoda,

    Roles Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Project administration, Software, Validation, Visualization, Writing – original draft, Writing – review & editing

    Affiliation Department of Ecology and Evolutionary Biology, University of Colorado, Boulder, Colorado, United States of America

  • Joshua G. Stepanek,

    Roles Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Validation, Visualization, Writing – original draft, Writing – review & editing

    Affiliation Department of Biology, Colorado Mountain College, Edwards, Colorado, United States of America

  • Nolan C. Kane,

    Roles Conceptualization, Funding acquisition, Resources, Supervision, Validation, Writing – review & editing

    Affiliation Department of Ecology and Evolutionary Biology, University of Colorado, Boulder, Colorado, United States of America

  • J. Patrick Kociolek

    Roles Conceptualization, Funding acquisition, Resources, Supervision, Validation, Writing – review & editing

    Affiliations Department of Ecology and Evolutionary Biology, University of Colorado, Boulder, Colorado, United States of America, Museum of Natural History, University of Colorado, Boulder, Colorado, United States of America

Abstract

Diatoms are the most diverse lineage of algae, but the diversity of their chloroplast genomes, particularly within a genus, has not been well documented. Herein, we present three chloroplast genomes from the genus Halamphora (H. americana, H. calidilacuna, and H. coffeaeformis), the first pennate diatom genus to be represented by more than one species. Halamphora chloroplast genomes ranged in size from ~120 to 150 kb, representing a 24% size difference within the genus. Differences in genome size were due to changes in the length of the inverted repeat region, length of intergenic regions, and the variable presence of ORFs that appear to encode as-yet-undescribed proteins. All three species shared a set of 161 core features but differed in the presence of two genes, serC and tyrC of foreign and unknown origin, respectively. A comparison of these data to three previously published chloroplast genomes in the non-pennate genus Cyclotella (Thalassiosirales) revealed that Halamphora has undergone extensive chloroplast genome rearrangement compared to other genera, as well as containing variation within the genus. Finally, a comparison of Halamphora chloroplast genomes to those of land plants indicates diatom chloroplast genomes within this genus may be evolving at least ~4–7 times faster than those of land plants. Studies such as these provide deeper insights into diatom chloroplast evolution and important genetic resources for future analyses.

Introduction

Diatoms are single-celled eukaryotic algae with silica cell walls. They play an important role in global O2, CO2, and silica cycling [1,2] and have the most efficient Rubisco known [3]. Diatoms are the most diverse group of eukaryotic algae [4] with an estimated 100,000 species [5] and occupy many niches in marine and freshwater environments [6].

Halamphora (Cleve) Levkov [7] is a recently described genus composed of species formerly assigned to the genus Amphora Ehrenberg ex Kützing. Members of the genus occur across a wide ecological spectrum including fresh to hypersaline habitats [8] from the tropics to the polar regions [9,10], and are known to be prodigious oil producers [11,12]. Although diverse, recent taxonomic treatments [7,8] and broadly sampled molecular phylogenetic analyses [13,14] have combined to make Halamphora one of the most well studied groups of diatoms in terms of phylogenetic systematics.

Despite diatom diversity and importance as primary producers in most aquatic ecosystems, the chloroplast genomes of only 40 species have been analyzed to date and taxon sampling has primarily focused on the ‘polar’ centric diatoms (45% of published chloroplast genomes; [15,16,17]). Outside of this group, no two species from within the same genus have been examined prior to this study. As the cost of genomic sequencing has decreased [18], it has become more feasible to examine in greater detail the genetic makeup of biologically important, non-model organisms. This allows for salient comparisons and insight into evolutionary lifestyle mechanisms at the genetic level. Finer scale taxonomic sampling in non-model organisms, such as within genus-level comparisons, elucidates the time-scales of evolutionary processes that may occur at rates too rapid to yield meaningful comparisons at more sparse taxonomic sampling regimens.

Therefore, the purpose of this study is to explore the phylogenetic and genomic relationships between three closely related Halamphora species (H. americana Kociolek in Kociolek et al., H. calidilacuna Stepanek & Kociolek, and H. coffeaeformis (C.Agardh; Levkov)). We then compared overall genomic content between these newly sequenced and annotated genomes to currently published diatom plastid genomes. Strikingly, Halamphora demonstrates high levels of gene rearrangement in comparison to the genus Cyclotella. The levels of gene rearrangements in the genus Halamphora are comparable to the level observed in much older-diverging lineages of the major groups of land plants (dicots and monocots).

Materials and methods

Ethics statement

No permits were required for collection of benthic samples in Salt Alkaline Lake, ND or Blue Lake Warm Spring, UT. No specific permissions were required for these locations/activities. Both sites are publicly accessible and there are no regulations regarding the collection of algae from these sites. Field studies did not involve endangered or protected species.

Isolate collection and culturing

Environmental samples containing H. americana and H. coffeaeformis were collected from Salt Alkaline Lake, ND in 2011 and samples containing H. calidilacuna were collected from Blue Lake Warm Spring, UT in 2012 (Table 1). Conductivity and pH measurements were recorded from the sites at the time of collection using a YSI 556 multi-probe (YSI Incorporated, Yellow Springs, Ohio, USA). Individual cells were isolated into monoculture by micropipette serial dilution and grown in artificial brackish water medium created using the sea salts Instant Ocean (Spectrum Brands, Inc., Blacksburg, Virginia, USA). Conductivity of the medium was adjusted to 10 mS cm -1 to approximate the conductivity at the collection sites (Table 1) and added macro- and micronutrients were based on those of WC media [19] with the Na2SiO3 concentration increased to 56.85 mg l-1. Cultures were maintained at ~25°C, under fluorescent illumination with a 12:12 light cycle at an irradiance of ~50 μmol cm-2 S-1.

DNA extraction, library preparation, and sequencing

Cultures were harvested by centrifugation and DNA was extracted using the Qiagen DNeasy Plant Mini Kit (Qiagen, Crawley, UK) following manufacturer’s protocols.

Library preparation and sequencing followed Pogoda et al. [20]. Briefly, genomic libraries were prepared using Nextera XT DNA library prep kits (Illumina). The protocol calls for 1 ng total of input DNA and each gDNA sample was diluted to the appropriate concentration using a Qubit 3.0 fluorometer (ThermoFisher Scientific). Each sample was barcoded by the unique dual index adapters Nextera i5 and i7. Resulting libraries were cleaned using solid phase reversible immobilization (SPRI) to remove fragment sizes less than 300 base pairs via an epMotion 5075TMX automated liquid handling system. Sample quality control (QC) was conducted prior to normalizing the loading concentration of pooled samples to 1.8–2.1 pM with 1% PhiX control v3 added (Illumina). Samples were processed for paired end 151 base pair reads on the Illumina NextSeq sequencer at the University of Colorado’s BioFrontiers Institute Next-Generation Sequencing Facility in Boulder, Colorado. All wet lab work was performed in the Department of Ecology and Evolutionary Biology at the University of Colorado, Boulder.

Assembly and annotation

Raw de-multiplexed data were sub-sampled to approximately 2 GB per sample and trimmed using Trimmomatic-0.36 with the following parameters: ILLUMINACLIP:NexteraPE-PE.fa:2:20:10MINLEN:140 LEADING:20 TRAILING:20 [21]. Resulting fastq files were de novo assembled using SPAdes v3.9 with the following parameters: SPAdes-3.9.0-Linux/bin/spades.py—careful -k 35,55,85 [22]. Diatom chloroplast contigs were identified using command-line BLAST against Phaeodactylum tricornutum (accession EF067920) and confirmed using a web BLAST. Annotations were initiated in DOGMA [23] and completed in NCBI’s Sequin 15.10 (Bethesda, MD) using the protein-coding sequence of 23 diatom chloroplast genomes as references (S1 Table). Putative features unique to these genomes (or not found in the reference genomes) were identified using NCBI’s ORF Finder (Bethesda, MD). Annotated plastid genomes are available in GenBank using accession numbers MK045450 –MK045452.

Genome content was visualized using OGDraw v1.2 [24]. Total sequence length of protein coding genes (CDS), transfer RNAs (tRNAs), ribosomal RNAs (rRNAs), and non-coding DNA was calculated as a sum of the total content of each type of feature. Web BLAST was used to determine if unique ORFs identified using NCBI’s ORF Finder contained any sequence similarity to ORFs found in other species’ chloroplast genomes. A gene was considered present if there was an appropriate, full length BLAST hit to a reference diatom chloroplast. Otherwise, it was assumed absent. In addition, gene density (number of protein coding genes / genome size) was calculated [25].

Chloroplast phylogeny

Twenty molecular markers (psaA, psbC, petD, petG, atpA, atpG, rbcL, rbcS, rpoA, rpoB, rps14, rpl33, rnl, rns, ycf89, sufB, sufC, dnaK, dnaB, clpC) from 26 species were aligned (S1 File) using the MAFFT algorithm [26], manually edited as necessary, and partitioned by gene and codon position (except for ribosomal DNA regions). Maximum likelihood analysis with 50 independent tree searches and 1000 rapid bootstrap replicates was performed in RAxML [27] with the graphical user interface raxmlGUI ver. 1.2 [28] using the GTR+Γ+I model of evolution.

Pairwise genetic distances utilizing a Jukes-Cantor model of evolution were calculated for the twenty-marker concatenated alignment using R-Studio v1.1.456 using the ‘dist.dna’ function in the R package ape [29].

Synteny

Two separate alignments of biraphid pennate and thalassiosiroid diatoms and resulting locally collinear blocks (LCBs) were estimated with MAUVE 2.4.0 [30] after eliminating one copy of the inverted repeat (IR). Rearrangement distances between LCBs were measured using GRIMM 2.02 [31].

Results and discussion

General features

Halamphora chloroplast genomes varied in size from ~120–150 kb (Table 2), representing a size difference of ~20% within the genus. Despite this difference in size, they contained similar GC content (~30%; Table 2). The Halamphora genomes contain two canonical inverted repeats (IRs) separated by a small (SSC) and a large single-copy (LSC) region, a structure these genomes share with other diatoms (e.g., [16,32]), some red algae (e.g., [33]), glaucophytes (e.g. [34]), some green algae (e.g., [35]), and most land plants (e.g., [36]).

thumbnail
Table 2. Quantification of features in three Halamphora species, including % non-coding sequence, % GC, the number of ORFs (the number of ORFs shared by > 1 Halamphora genome), and gene density.

https://doi.org/10.1371/journal.pone.0217824.t002

The three Halamphora genomes shared a set of 130 protein-coding genes, three rDNAs, 27 tRNAs, one tmRNA (transfer-messenger RNA) and ffs (signal recognition particle RNA) (Fig 1 and S1S3 Figs) with a similar gene content to other pennate diatoms (S2 Table). With the exception of ORFs (see below), the chloroplast genomes of H. americana, H. calidilacuna, and H. coffeaeformis are nearly identical (99%) in gene content and differed only in the presence/absence of two genes– phosphoserine aminotransferase (serC) and cyclohexadienyl dehydrogenase (tyrC), both of which have been suggested to be of foreign or unknown origin, respectively [15].

thumbnail
Fig 1. Plastid genome map of Halamphora calidilacuna.

Genes on the outside are transcribed clockwise and those on the inside counterclockwise. The inner ring displays GC content in grey.

https://doi.org/10.1371/journal.pone.0217824.g001

Among the Halamphora genomes, serC was only present, and then only as a pseudogene, in H. calidilacuna and is presumed to be of plasmid origin. Plasmids have been found in only some pennate diatoms thus far [37,38]. serC has been found in five other diatom plastid genomes as a gene/pseudogene (S2 Table) and in plasmids of Cylindrotheca species [15,37,38]. An additional indication of the plasmid origin of this gene in H. calidilacuna is the presence of a partial (presumably non-functional) copy of ORF484 in the chloroplast genome, an ORF also found in the pCf2 plasmid of C. fusiformis [37,38].

The presence/absence of tyrC follows an interesting pattern, being present (and presumably functional) in H. calidilacuna, present and presumably non-functional pseudogene in H. americana, and absent in H. coffeaeformis. Although tyrC has also been found in other diatoms, green algae, and bacteria, its origin is less clear [15]. The presence as a gene/pseudogene in the two most closely related Halamphora taxa (Fig 2) and absence in the other taxon could indicate an acquisition via plasmid or from bacterial horizontal gene transfer (HGT) prior to the split of H. calidilacuna and H. americana, followed by a subsequent loss of function in H. americana. Alternatively, it is possible that all Halamphora species contain tyrC, but it is in transition to relocating to the nucleus.

thumbnail
Fig 2. Maximum likelihood phylogram inferred from twenty chloroplast encoded markers (see Materials and methods).

Node support is given as maximum likelihood bootstrap values (1000 bootstrap replicates). Asterisks indicate 100% support. Letters indicate morphological groups of diatoms as follows: (A) ‘radial’ centric; (B) ‘polar’ centric; (C) araphid; and (D) biraphid pennate diatoms.

https://doi.org/10.1371/journal.pone.0217824.g002

Differences in genome size are due to difference in length of the IRs and intergenic regions (Table 2). Gene density (calculated as the number of protein coding genes / genome size) was inversely proportional to genome size, with the greatest gene density in the smallest genome and vice versa (Table 2).

The largest Halamphora genome (H. calidilacuna; Table 2) also contains more open reading frames (ORFs) than the other congeners. Three ORFs are shared between at least two Halamphora genomes, but no ORF is common to all three genomes (S2 Table). Only two ORFs (ORF385, ORF484) were shared between Halamphora species and other diatom genomes. ORF385, which was free standing in the genome, was found in H. calidilacuna, Seminavis robusta, and Asterionellopsis glacialis. A partial (presumably non-functional) copy of ORF484, also free standing, was found in H. calidilacuna and within a plasmid of Cylindrotheca fusiformis [37,38]. Similar to the retrotransposons found in diatom mitochondrial genomes [20], two group II introns containing regions with homology to reverse transcriptases/maturases (ORF26 & ORF27) were found in H. calidilacuna and S. robusta. In H. calidilacuna, ORF26 was present in an intron of petD and ORF27 was in an intron of petB. These two ORFs show some similarity (~70% similarity at > 86% coverage) to group II introns found in Ulvophytes [39] and red algae [33].

The nucleotide identity between sister taxa H. calidilacuna and H. americana is 98%, and the identity between either of these taxa and H. coffeaeformis is 88%. In addition to overall sequence similarity, the utility of the rbcL (~1400 bp) and rbcL-3P (748 bp) as barcode markers [40] to distinguish between these closely related taxa was also evaluated. Both rbcL barcode markers (both phylogenetic and shorter rbcL-3P lengths; [40]) show the same pattern as the overall similarity; 1% divergence was observed between H. calidilacuna and H. americana and 5–6% (rbcL and rbcL-3P, respectively) divergence was observed between H. coffeaeformis and either of these taxa. Therefore, either of these barcoding markers could be used to distinguish between these closely related Halamphora species.

Seven pairs of overlapping genes were found in one or more Halamphora species, with some overlapping pairs also found in the Thalassiosirales. Overlapping pairs found in both Halamphora spp. and Thalassiosirales include: psbD-psbC by 53 bp; atpF-atpD by 4 bp; and rpl23-rpl4 by 11 bp in Halamphora spp. and 8–17 bp in the Thalassiosirales. Additional overlapping pairs include: sufB-sufC by 1 bp (H. americana, H. calidilacuna, and Thalassiosirales), ycf45-psaB by 4 bp (H. coffeaeformis and H. calidilacuna), within ORF25 by 15 bp (H. americana); and ORF385-serC by 2 bp (H. calidilacuna). Despite the widespread occurrence of overlapping genes, the origin, evolution and ramification of these overlaps remains unknown [41]. Some overlapping genes (e.g., psbD-psbC) are known to cause translational coupling, i.e, the translation of the psbC cistron depends on the translation of the psbD cistron [42]. Overlapping genes may also produce novel de novo proteins, a process common in viral genomes that can lead to changes in pathogenicity and possibly genome evolution [43]. Another type of alternative transcription (i.e., intron retention) is important to diatoms’ ability to adapt to changing nutrient conditions and not trivial in maintaining their physiology [44].

Phylogeny results

The resulting phylogeny (Fig 2) revealed ‘polar’ centric (B), araphid (C), and biraphid pennate (D) diatoms to be monophyletic groups and ‘radial’ centric (A) diatoms to be a paraphyletic group. The monophyly of araphid diatoms is most likely due to the limited taxon sampling in our tree, as this group is often paraphyletic in analyses that include a larger number of taxa in this group [16,45,46]. The monophyly of ‘polar’ centric diatoms may be due to taxon sampling as well [16,45,46], but some studies have shown this group to be monophyletic (e.g., [47,48]) and this is an area of ongoing research. Within the biraphid pennate diatoms, the relationships between the genus Halamphora and the remaining taxa largely agrees with the multi-gene phylogenies of Ruck & Theriot [49] and Stepanek & Kociolek [13]. However, a single gene (18S rDNA) phylogeny presented by Zgrundo et al. [50], which includes taxa from the genus Fistulifera H. Lange-Bertalot, recovered this genus within a clade consistently more closely related to the Halamphora than to the rest of the biraphid pennate diatoms [13,49,50]. The 20-gene phylogeny presented here continues to strongly support (BS 100) the monophyly of the genus Halamphora as well as the close sister relationship between H. americana and H. calidilacuna that have been recovered in several broadly-sampled multigene phylogenies of the group [13,14].

Synteny results

The MAUVE alignment of biraphid pennate diatoms resulted in 26 locally collinear blocks (LCBs) of sequence among the ten diatoms examined (Fig 3). This gene order comparison revealed inversions, translocations, and inversion/translocation combinations resulting in extensive plastid genome rearrangement across this group (Fig 3). Distances between gene orders of LCBs calculated using GRIMM revealed the gene order of Cylindrotheca closterium to be the most unique among these diatoms (S3 Table). Although gene order was more conserved within the clade containing Didymosphenia geminata, Phaeodactylum tricornutum, H. americana, H. calidilacuna, and H. coffeaeformis, no chloroplast genomes in these analyses had identical gene order (Fig 3 and S3 Table). Within Halamphora, gene order was not conserved and distances between gene orders of LCBs were ≥ 3 (S3 Table).

thumbnail
Fig 3. Gene order comparison of plastid genomes of eight biraphid pennate diatoms (three Halamphora spp. are from this study) with one copy of the inverted repeat removed prior to analysis.

Alignment and resulting locally collinear blocks (LCBs) were generated using MAUVE. Relationships between taxa displayed as a cladogram to the left of the diagram are based on Fig 2.

https://doi.org/10.1371/journal.pone.0217824.g003

Gene order of the thalassiosiroid diatoms [15,16] was examined as a comparison to the rearrangements observed within the biraphid pennate diatoms including Halamphora. The MAUVE alignment of thalassiosiroid diatoms resulted in 18 locally collinear blocks (LCBs) of sequence among the six diatoms examined (Fig 4). With the exception of T. oceanica, gene order is more conserved in the thalassiosiroid diatoms (Fig 4) and the resulting rearrangement distances are smaller (S4 Table and Fig 5), a pattern consistent with the findings of Ruck et al. [15] and Sabir et al. [16]. In particular, there was only one inversion between Cyclotella species (Fig 4), in contrast to the inversions, translocations, and inversion/translocation combinations within Halamphora spp. (Fig 3). A comparison of Jukes-Cantor genetic distance to rearrangement distance generated in GRIMM for both biraphid and thalassiosiroid diatoms (Fig 5) shows a weak but positive relationship between these two measures of evolution.

thumbnail
Fig 4. Gene order comparison of plastid genomes of six thalassiosiroid diatoms with one copy of the inverted repeat removed prior to analysis.

Alignment and resulting locally collinear blocks (LCBs) were generated using MAUVE. Relationships between taxa displayed as a cladogram to the left of the diagram are based on Fig 2.

https://doi.org/10.1371/journal.pone.0217824.g004

thumbnail
Fig 5. Plot describing the relationship between rearrangement distance and Jukes-Cantor genetic distance.

Blue circles represent biraphid diatoms. Red triangles represent thalassiosiroid diatoms. Regression coefficient for the biraphid diatoms is significant (p = 0.04, R2 = 0.136), whereas for the thalassiosiroid diatoms it is not (p = 0.34, R2 = 0.069).

https://doi.org/10.1371/journal.pone.0217824.g005

The number of inversions and translocations inferred within a genus of diatoms is surprisingly high, particularly among species of Halamphora, where congeners may differ by up to 7 detectable rearrangements (S3 Table). This represents a striking contrast to chloroplast genomes in most land plants, where gene order is typically highly conserved within genera, and even across diverse families, orders, and in some cases classes [51]. For example, the monocot Oryza sativa differs from the dicot Arabidopsis thaliana by only one inversion and a translocation, despite those lineages diverging an estimated 160 million years ago [52]. The level of sequence divergence (12%), differences in gene content (2 genes differ in presence/absence), and degree of genome feature rearrangement (~7 rearrangements) found in Halamphora is high, and in order to recapitulate this level of divergence in land plants, one must compare as disparate groups as angiosperms (i.e., A. thaliana: accession NC_000932) and Equisetales (i.e., Equisetum arvense: accession NC_014699) [52]. Based on estimates in a recent publication [45], Halamphora evolved ~75–100 MYA, compared to over 400 MYA between angiosperms and Equisetales [53]. The first occurrence of amphoroid diatoms in the fossil record comes from the Oamaru Formation [54], estimated to be of Eocene age (33–55 MYA). These data suggest that chloroplast genomes of this genus are evolving approximately four (based on the estimates of the first occurrence of the group) to seven (based on occurrence in the fossil record) times faster than those of land plants in multiple ways, including sequence divergence, gene content, and gene order.

Conclusions

Numerous studies have compared gene content and genome rearrangement among diatom chloroplast genomes (e.g., [15,16,17,32,55,56,57,58,59,60,61,62]), but this is the first to compare multiple genomes within a genus of biraphid pennate diatoms and several surprising patterns were identified. In regard to the phylogenetic position of Halamphora and the relationship between Halamphora species, our 20-gene phylogeny supported similar relationships to those revealed in other studies [13,14]. As with other diatoms, these chloroplast genomes are evolving relatively rapidly at the sequence level (12% divergence across the genus Halamphora). Halamphora plastid genomes also showed variation in gene content, with species incorporating two genes. Even more striking variation was observed in gene order, with multiple inversions, translocations, and inversion/translocation combinations found within this genus. Cyclotella, a thalassiosiroid genus, showed more conservation in gene order, with only one inversion. Overall, biraphid pennate diatoms appear to display more variation in gene order than the thalassiosiroid diatoms and significantly more variation than typical land plant chloroplasts (notable exceptions include the Geraniaceae family [63] and Amborella [64]). Although this pattern is conspicuous, only 0.04% of the estimated 100,000 diatom species’ chloroplast genomes have been examined and therefore, additional data and comparisons are necessary before generalizations should be made regarding overall diatom chloroplast genome evolution. Although these data are preliminary (comprising only a fraction of diatom diversity), they point to a comparable degree of variation within this one genus of diatoms to the divergence among distant divisions of vascular plants. This diversification within Halmaphora is accompanied by a substantially higher (4–7×) rate of evolution. These remarkable intrageneric and inter-kingdom comparisons require additional data to verify the results. However, if these data are supported by additional studies, they open the door to many questions about the rate and modes of molecular evolution of the chloroplast genome in this remarkable clade.

Supporting information

S1 Table. Diatom chloroplast genomes used as reference for annotating the Halamphora genomes.

https://doi.org/10.1371/journal.pone.0217824.s001

(DOCX)

S2 Table. Gene content comparison of three Halamphora spp. genomes (bold) with other published diatom plastid genomes.

X, present; P, pseudogenized copy of gene is present; -, absent.

https://doi.org/10.1371/journal.pone.0217824.s002

(XLSX)

S3 Table. Distances between gene orders of LCBs (generated in MAUVE) of biraphid pennate diatoms (taxa from this study are in bold) calculated using GRIMM.

Smaller values indicate more similar gene order.

https://doi.org/10.1371/journal.pone.0217824.s003

(DOCX)

S4 Table. Distances between gene orders of LCBs (generated in MAUVE) of thalassiosiroid diatoms calculated using GRIMM.

Smaller values indicate more similar gene order.

https://doi.org/10.1371/journal.pone.0217824.s004

(DOCX)

S1 Fig. Plastid genome map of Halamphora calidilacuna.

Genes on the outside are transcribed clockwise and those on the inside counterclockwise. The inner ring displays GC content in grey.

https://doi.org/10.1371/journal.pone.0217824.s005

(TIF)

S2 Fig. Plastid genome map of Halamphora americana.

Genes on the outside are transcribed clockwise and those on the inside counterclockwise. The inner ring displays GC content in grey.

https://doi.org/10.1371/journal.pone.0217824.s006

(TIF)

S3 Fig. Plastid genome map of Halamphora coffeaeformis.

Genes on the outside are transcribed clockwise and those on the inside counterclockwise. The inner ring displays GC content in grey.

https://doi.org/10.1371/journal.pone.0217824.s007

(TIF)

S1 File. Fasta alignment.

The fasta alignment of twenty molecular markers (psaA, psbC, petD, petG, atpA, atpG, rbcL, rbcS, rpoA, rpoB, rps14, rpl33, rnl, rns, ycf89, sufB, sufC, dnaK, dnaB, clpC) from 26 species used to generate the phylogeny presented in this study.

https://doi.org/10.1371/journal.pone.0217824.s008

(FASTA)

References

  1. 1. Nelson DM, Tréguer P, Brzezinski MA, Leynaert A, Quéguiner B. Production and dissolution of biogenic silica in the ocean: revised global estimates, comparison with regional data and relationship to biogenic sedimentation. Global Biogeochem Cy. 1995;9: 359–372.
  2. 2. Struyf E, Smis A, Van Damme S, Meire P, Conley DJ. The global biogeochemical silicon cycle. Silicon. 2009;1: 207–213.
  3. 3. Fabris M, Matthijs M, Rombauts S, Vyverman W, Goossens A, Baart GJE. The metabolic blueprint of Phaeodactylum tricornutum reveals a eukaryotic Entner-Doudoroff glycolytic pathway. Plant J. 2012;70: 1004–1014. pmid:22332784
  4. 4. Mann DG. The species concept in diatoms. Phycologia. 1999;38: 437–495.
  5. 5. Mann DG, Vanormelingen P. An inordinate fondness? The number, distributions, and origins of diatom species. J Eukaryot Microbiol. 2013;60: 414–420. pmid:23710621
  6. 6. Seckbach J, Kociolek P, editors. The Diatom World. Springer Netherlands; 2011. https://doi.org/10.1007/978-94-007-1327-7
  7. 7. Levkov Z. Amphora sensu lato. In: Lange-Bertalot H, editor. Diatoms of Europe: diatoms of the European inland waters and comparable habitats. Ruggell: ARG Gantner Verlag KG; 2009. p. 5–916.
  8. 8. Stepanek JG, Kociolek JP. Amphora and Halamphora from coastal waters and inland waters of the United States and Japan. Bibl Diatomologica. 2018;66: 1–260.
  9. 9. Cavalcante KP, Tremarin PI, Ludwig TAV. New records of amphoroid diatoms (Bacillariophyceae) from Cachoeira River, Northeast Brazil. Braz J Biol. 2014;74: 257–263. pmid:25055112
  10. 10. Van de Vijver B, Kopalova K, Zidarova R, Levkov Z. Revision of the genus Halamphora (Bacillariophyta) in the Antarctic region. Plant Ecol Evol. 2014;147: 347–391.
  11. 11. Sheehan J, Dunahay T, Benemann J, Roessler P. A look back at the US Department of Energy’s aquatic species program: biodiesel from algae. National Renewable Energy Laboratory, Golden, CO, 1998;TP-580-24190: 328 p.
  12. 12. Stepanek JG, Fields FJ, Kociolek JP. A comparison of lipid content metrics using six species from the genus Halamphora (Bacillariophyta). Biofuels. 2016;7: 521–528.
  13. 13. Stepanek JG, Kociolek JP. Molecular phylogeny of Amphora sensu lato (Bacillariophyta): an investigation into monophyly and classification of the amphoroid diatoms. Protist. 2014;165: 177–195. pmid:24646793
  14. 14. Stepanek JG, Kociolek JP. Molecular phylogeny of the diatom genera Amphora and Halamphora (Bacillariophyta) with a focus on morphological and ecological evolution. J Phycol. 2019;55: 442–456. pmid:30659609
  15. 15. Ruck EC, Nakov T, Jansen RK, Theriot EC, Alverson AJ. Serial gene losses and foreign DNA underlie size and sequence variation in the plastid genomes of diatoms. Genome Biol Evol. 2014;6: 644–654. pmid:24567305
  16. 16. Sabir JSM, Yu M, Ashworth MP, Baeshen NA, Baeshen MN, Bahieldin A, et al. Conserved gene order and expanded inverted repeats characterize plastid genomes of Thalassiosirales. PLOS One. 2014;9: e107854. pmid:25233465
  17. 17. Yu M, Ashworth MP, Hajrah NH, Khiyami MA, Sabir MJ, Alhebshi AM, et al. Evolution of the plastid genomes in diatoms. Adv Bot Res. 2018; in press.
  18. 18. Schuster SC. Next-generation sequencing transforms today's biology. Nature Methods. 2007;5: 16. pmid:18165802
  19. 19. Guillard RRL, Lorenzen CJ. Yellow-green algae with Chlorophyllide. J Phycol. 1972;8: 1014.
  20. 20. Pogoda CS, Keepers KG, Hamsher SE, Stepanek JG, Kane NC, Kociolek JP. Comparative analysis of the mitochondrial genomes of six newly sequenced diatoms reveals group II introns in the barcoding region of cox1. Mitochondr DNA Part A. 2019;30: 43–51. pmid:29527965
  21. 21. Bolger AM, Lohse M, Usadel B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics. 2014;30: 2114–2120. pmid:24695404
  22. 22. Bankevich A, Nurk S, Antipov D, Gurevich AA, Dvorkin M, Kulikov AS. SPAdes: a new genome assembly algorithm and its applications to single-cell sequencing. J Comput Biol. 2012;19: 455–477. pmid:22506599
  23. 23. Wyman SK, Jansen RK, Boore JL. Automatic annotation of organellar genomes with DOGMA. Bioinformatics. 2004;20: 3252–3255. pmid:15180927
  24. 24. Lohse M, Drechsel O, Kahlau S, Bock R. OrganellarGenomeDRAW—a suite of tools for generating physical maps of plastid and mitochondrial genomes and visualizing expression data sets. Nucleic Acids Res. 2013;41: W575–W581. pmid:23609545
  25. 25. Lee J, Yang EC, Graf L, Yang JH, Qiu H, Zelzion U, et al. Analysis of the draft genome of the red seaweed Gracilariopsis chorda provides insights into genome size evolution in Rhodophyta. Mol Biol Evol. 2018;35: 1869–1886. pmid:29688518
  26. 26. Katoh K, Standley DM. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol Biol Evol. 2013;30: 772–780. pmid:23329690
  27. 27. Stamatakis A. RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics. 2014;30: 1312–1313. pmid:24451623
  28. 28. Silvestro D, Michalak I. raxmlGUI: a graphical front-end for RAxML. Org Div Evol. 2012;12: 335–337.
  29. 29. Paradis E, Schliep K. ape 5.0: an environment for modern phylogenetics and evolutionary analysis in R. Bioinformatics. 2019;35: 526–528. pmid:30016406
  30. 30. Darling ACE, Mau B, Blattner FR, Perna NT. Mauve: multiple alignment of conserved genomic sequence with rearrangements. Genome Res. 2004;14: 13941403. pmid:15231754
  31. 31. Tesler G. GRIMM: genome rearrangements web server. Bioinformatics. 2002;18: 492–493. pmid:11934753
  32. 32. Imanian B, Pombert J-F, Keeling PJ. The complete plastid genomes of the two ‘dinotoms’ Durinskia baltica and Kryptoperidinium foliaceum. PLOS One. 2010;5: e10711. pmid:20502706
  33. 33. Muñoz-Gómez SA, Mejía-Franco FG, Durnin K, Colp M, Grisdale CJ, Archibald JM, et al. The new red algal sybphylum Proteorhodophytina comprises the largest and most divergent plastid genomes known. Curr Biol. 2017;27: 1677–1684. pmid:28528908
  34. 34. Reyes-Prieto A, Russell S, Figueroa-Martinez F, Jackson C. Comparative plastid genomics of Glaucophytes. Adv Bot Res. 2018;85: 95–127.
  35. 35. Letsch MR, Lewis LA. Chloroplast gene arrangement variation within a closely related group of green algae (Trebouxiophyceae, Chlorophyta). Mol Phylo Evol. 2012;64: 524–532. pmid:22659018
  36. 36. Shinozaki K, Ohme M, Tanaka M, Wakasugi T, Hayashida N, Matsubayashi T. The complete nucleotide sequence of the tobacco chloroplast genome: its gene organization and expression. EMBO J. 1986;5: 2043–2049. pmid:16453699
  37. 37. Hildebrand M, Corey DK, Ludwig JR, Kukel A, Feng T-Y, Volcani BE. Plasmids in diatom species. J Bacteriol. 1991;173: 5924–5927. 0021-9193/91/185924-04$02.00/0 pmid:1885558
  38. 38. Hildebrand M, Hasegawa P, Ord RW, Thorpe VS, Glass CA, Volcani BE. Nucleotide sequence of diatom plasmids: identification of open reading frames with similarity to site-specific recombinases. Plant Mol Biol. 1992;19: 759–770. pmid:1322740
  39. 39. Turmel M, Otis C, Lemieux C. Mitochondrion-to-chloroplast DNA transfers and intragenomic proliferation of chloroplast group II introns in Gloeotilopsis Green Algae (Ulotrichales, Ulvophyceae). Genome Biol Evol. 2017;8: 2789–2805. pmid:27503298
  40. 40. Hamsher SE, Evans KM, Mann DG, Poulíčková A, Saunders GW. Barcoding diatoms: exploring alternatives to COI-5P. Protist. 2011;162: 405–422. pmid:21239228
  41. 41. Huvet M, Stumpf MPH. Overlapping genes: a window on gene evolvability. BMC Genomics. 2014;15: 721. pmid:25159814
  42. 42. Adachi Y, Kuroda H, Yukawa Y, Sugiura M. Translation of partially overlapping psbD-psbC mRNAs in chloroplasts: the role of 5´-processing and translational coupling. Nucleic Acids Res. 2012;40: 3152–3158. pmid:22156163
  43. 43. Rancurel C, Khosravi M, Dunker AK, Romero PR, Karlin D. Overlapping genes produce proteins with unusual sequence properties and offer insight into de novo protein creation. J Virology. 2009;82: 10719–10736. pmid:19640978
  44. 44. Rastogi A, Maheswari U, Dorrell RG, Rocha Jimenez Vieira F, Maumus F, Kustka A, et al. Integrative analysis of large scale transcriptome data draws a comprehensive landscape of Phaeodactylum tricornutum genome evolutionary origin of diatoms. Sci Rep. 2018;8: 4834. pmid:29556065
  45. 45. Nakov T, Beaulieu JM, Alverson AJ. Accelerated diversification is related to life history and locomotion in a hyperdiverse lineage of microbial eukaryotes (Diatoms, Bacillariophyta). New Phytol. 2018;219: 462–473. pmid:29624698
  46. 46. Parks MB, Nakov T, Ruck EC, Wickett NJ, Alverson AJ. Phylogenomics reveals an extensive history of genome duplication in diatoms (Bacillariophyta). Am J Bot. 2018;105: 330–347. pmid:29665021
  47. 47. Ashworth MP, Nakov T, Theriot EC. Revisiting Ross and Sims (1971): toward a molecular phylogeny of the Biddulphiaceae and Eupodiscaceae (Bacillariophyceae). J Phycol. 2013;49: 1207–1222. pmid:27007638
  48. 48. Gargas CB, Theriot EC, Ashworth MP, Johansen JR. Phylogenetic analysis reveals that the ‘Radial Centric’ diatom Orthoseira Thwaites (Orthoseiraceae, Bacillariophyta) is a member of a ‘Multipolar’ diatom lineage. Protist 2018;169: 803–825. pmid:30448592
  49. 49. Ruck EC, Theriot EC. Origin and evolution of the canal raphe system in diatoms. Protist. 2011;162: 723–737. pmid:21440497
  50. 50. Zgrundo A, Lemke P, Pniewski F, Cox EJ, Latala A. Morphological and molecular phylogenetic studies on Fistulifera saprophila. Diatom Res. 2013;28: 431–443.
  51. 51. Jansen R. K., & Ruhlman T. A. (2012). Plastid genomes of seed plants. In Genomics of chloroplasts and mitochondria (pp. 103–126). Springer, Dordrecht
  52. 52. Xu J-H, Liu Q, Hu W, Wang T, Xue Q, Messing J. Dynamics of chloroplast genomes in green plants. Genomics. 2015;106: 221–231. pmid:26206079
  53. 53. Soltis PS, Soltis DE, Savolainen V, Crane PR, Barraclough TG. Rate heterogeneity among lineages of tracheophytes: integration of molecular and fossil data and evidence for molecular living fossils. P Natl Acad Sci USA. 2002;99: 4430–4435.
  54. 54. Schrader H-J. Die Pennaten Diatomeen aus dem Obereozän von Oamaru, Neuseeland. Beihefte zur Nova Hedwigia. 1969;28: 1–124.
  55. 55. Kowallik KV, Stoebe B, Schaffran I, Kroth-Pancic P, Freier U. The chloroplast genome of a chlorophyll a+c-containing alga, Odontella sinensis. Plant Mol Biol Rep. 1995;13: 336–342.
  56. 56. 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: 427–439. pmid:17252281
  57. 57. Lommer M, Roy A-S, Schilhabel M, Schreiber S, Rosenstiel P, LaRoche J. Recent transfer of an iron-regulated gene from the plastid to the nuclear genome in an oceanic diatom adapted to chronic iron limitation. BMC Genomics. 2010;11: 718. pmid:21171997
  58. 58. Ravin NV, Galachyants YP, Mardanov AV, Beletsky AV, Petrova DP, Sherbakova TA, et al. Complete sequence of the mitochondrial genome of a diatom alga Synedra acus and comparative analysis of diatom mitochondrial genomes. Curr Genet. 2010;56: 215–223. pmid:20309551
  59. 59. Tanaka T, Fukuda Y, Yoshino T, Maeda Y, Muto M, Matsumoto M, et al. High-throughput pyrosequencing of the chloroplast genome of a highly neutral-lipid-producing marine pennate diatom, Fistulifera sp. strain JPCC DA0580. Photosynth Res. 2011;109: 223–229. pmid:21290260
  60. 60. Brembu T, Winge P, Tooming-Klunderud A, Nederbragt AJ, Jakobsen KS, Bones AM. The chloroplast genome of the diatom Seminavis robusta: new features introduced through multiple mechanisms of horizontal gene transfer. Mar Genomics. 2014;16: 17–27. pmid:24365712
  61. 61. Galachyants YP, Zakharova YR, Petrova DP, Morozov AA, Sidorov IA, Marchenkov AM, et al. Sequencing of the complete genome of an araphid pennate diatom Synedra acus subsp. radians from Lake Baikal. Dokl Biochem Biophys. 2015;461: 84–88. pmid:25937221
  62. 62. Crowell RM, Nienow JA, Cahoon AB. The complete chloroplast and mitochondrial genomes of the diatom Nitzschia palea (Bacillariophyceae) demonstrate high sequence similarity to the endosymbiont organelles of the dinotom Durinskia baltica. J Phycol. 2018;55: 352–364. pmid:30536677
  63. 63. Guisinger MM, Kuehl JV, Boore JL, Jansen, RK. Extreme reconfiguration of plastid genomes in the angiosperm family Geraniaceae: rearrangements, repeats, and codon usage. Mol Biol Evol. 2010;28: 583–600.
  64. 64. Goremykin VV, Hirsch-Ernst KI, Wölfl S, Hellwig FH. (2003). Analysis of the Amborella trichopoda chloroplast genome sequence suggests that Amborella is not a basal angiosperm. Mol Biol Evol. 2003;20: 1499–1505. pmid:12832641