Skip to main content

Chloroplast genomic comparison provides insights into the evolution of seagrasses

Abstract

Background

Seagrasses are a polyphyletic group of monocotyledonous angiosperms that have evolved to live entirely submerged in marine waters. Thus, these species are ideal for studying plant adaptation to marine environments. Herein, we sequenced the chloroplast (cp) genomes of two seagrass species (Zostera muelleri and Halophila ovalis) and performed a comparative analysis of them with 10 previously published seagrasses, resulting in various novel findings.

Results

The cp genomes of the seagrasses ranged in size from 143,877 bp (Zostera marina) to 178,261 bp (Thalassia hemprichii), and also varied in size among different families in the following order: Hydrocharitaceae > Cymodoceaceae > Ruppiaceae > Zosteraceae. The length differences between families were mainly related to the expansion and contraction of the IR region. In addition, we screened out 2,751 simple sequence repeats and 1,757 long repeat sequence types in the cp genome sequences of the 12 seagrass species, ultimately finding seven hot spots in coding regions. Interestingly, we found nine genes with positive selection sites, including two ATP subunit genes (atpA and atpF), three ribosome subunit genes (rps4, rps7, and rpl20), one photosystem subunit gene (psbH), and the ycf2, accD, and rbcL genes. These gene regions may have played critical roles in the adaptation of seagrasses to diverse environments. In addition, phylogenetic analysis strongly supported the division of the 12 seagrass species into four previously recognized major clades. Finally, the divergence time of the seagrasses inferred from the cp genome sequences was generally consistent with previous studies.

Conclusions

In this study, we compared chloroplast genomes from 12 seagrass species, covering the main phylogenetic clades. Our findings will provide valuable genetic data for research into the taxonomy, phylogeny, and species evolution of seagrasses.

Peer Review reports

Background

Along with coral reefs and mangroves, seagrass beds are one of the three main marine ecosystems, providing important habitats for marine life. They have significant ecological functions, such as stabilizing coastal sediments, providing nursery grounds for juvenile fish, and sequestering carbon [1,2,3,4]. The currently recognized seagrasses consist of approximately 74 species of six families and 13 genera in the order Alismatales, accounting for less than 1% of all flowering plant species [5,6,7]. Les et al. (1997) detected three separate origins of seagrasses using the rbcL gene, which was confirmed in subsequent studies [7,8,9]. Through independent evolutionary routes, seagrasses growing in intertidal and subtidal zones are characterized by similar environments, such as high salinity, low light, anaerobic soils, and extreme tides. Seagrasses have evolved shared traits such as salt tolerance, slender and soft leaves, carbon-concentrating mechanisms, and aerenchyma in the roots and rhizomes [10,11,12,13,14].

The unique evolutionary characteristics of seagrasses are similar to those of whales, which evolved from the land to aquatic environments [15]. The emergence of seagrasses is one of the most notable evolutionary transformations in the history of angiosperms. Recently, the broad application of technologies for genome sequencing has been demonstrated to be a valuable phylogenetic and evolutionary tool for revealing the genetic development and adaptive evolutionary mechanism of seagrasses to the marine environment [14, 16, 17]. For example, the first complete genome sequence and detailed genomic analysis of Zostera marina found that it had discarded several key innovative features of angiosperms, such as genes involved in stomatal and ethylene pathways, terpenoid synthesis, ultraviolet protection and far-infrared sensing, during its secondary entry into the sea, and also that it had recoded its cell wall components and expanded its sucrose synthesis and transport, ion transport and light-harvesting chlorophyll b-protein genes to adapt to the complex marine environment [17]. Meanwhile, Lee et al. (2016) found that genes related to hormone biosynthesis signal transduction and cell wall catabolism were also lost or modified in Zostera muelleri in the process of adapting to the marine environment [16]. Moreover, Lee et al. (2018) also sequenced the genomes of Halophila ovalis and explored the convergent evolutionary features with the above two seagrasses, finding that all three seagrasses lost genes related to ethylene and terpenoid biosynthesis while retaining those related to salt tolerance [14]. However, compared with higher terrestrial plants, genomics analysis in seagrasses is still in its infancy, and further research is urgently needed.

Plant chloroplasts are the main functional organs of plants, having been formed through endosymbiosis between early plants and cyanobacteria [18], and have multiple functions in plant cells, including playing a crucial role in photosynthesis and carbon fixation [19, 20]. Since the process of photosynthesis is influenced by three main factors (light irradiance and wavelength, carbon dioxide concentration, and temperature), it is reasonable to assume that the genetic basis of the cp genome changes in response to the conversion of living habitats. Chloroplasts are essential organelles in plant evolution due to their features. The cp genome is small and contains well-characterized features, including a relatively stable gene content and conserved structural features such as a large single copy (LSC) region, a small single copy (SSC) region, and a pair of inverted repeats (IRs) regions, together with the feature of a slow rate of nucleotide substitution, providing essential information to support comparative evolutionary research [19, 21, 22]. Adaptive evolution is defined as the improved adaptability of a species to changing environmental conditions. Given the conservation of the cp genome, observable alterations should be anticipated if adaptive evolution occurs in this molecule. Thus, a comparative analysis of the cp genomes of seagrasses should enhance our understanding of plant adaptation to the sea environment.

In the present study, we aimed to provide comprehensive insights into the evolution of cp genomes across prominent seagrass families. Here, we assembled and characterized the cp genome sequences of two seagrass species and compared them with 10 published seagrass cp genomes from four families, including Cymodoceaceae, Ruppiaceae, Zosteraceae, and Hydrocharitaceae. We also identified repeat sequences and positive selection, and reconstructed the phylogenetic relationships and molecular divergence dates for the major lineages within the seagrasses. These results will help develop genetic markers for seagrasses and provide an important theoretical basis for subsequent population genetic analysis and phylogenetic relationship analysis.

Results

Genomic features of seagrasses

This study analyzed 12 cp genomes covering the four families represented by seagrasses: Zosteraceae, Hydrocharitaceae, Ruppiaceae, and Cymodoceaceae. Among them, Z. muelleri and H. ovalis were assembled and annotated for the first time in this study (Fig. 1). All cp genomes had a typical four-part structure: a LSC, an SSC, and two IRs (Table 1). The length of the 12 cp genomes ranged from 143,877 bp to 178,261 bp and varied in size among the different families as follows: Hydrocharitaceae > Cymodoceaceae > Ruppiaceae > Zosteraceae. The IR was 24,399–44,815 bp in length, the LSC was 78,949–89,851 bp in length, and the SSC was 2,150–19,160 bp in length. The length differences between families were mainly related to the expansion and contraction of the IR region. The cp genomes of the seagrass species had 116–158 genes, including 78–122 protein-coding genes, eight ribosomal RNA genes (rRNA), and 30–42 transfer RNA genes (tRNA). The total GC content of the cp genomes was 35.5%–39.2%, and the GC content was more similar for species of the same family.

Fig. 1
figure 1

Chloroplast genome gene map of Halophila ovalis and Zostera muelleri. Genes on the inside of the outer circle are transcribed clockwise, while those outsides are transcribed counterclockwise

Table 1 Chloroplast genome features of 12 seagrass species

Repeat sequence analysis

In this study, a total of 2,751 simple sequence repeats (SSRs) were identified in the cp genome sequences of the 12 seagrass species (Fig. 2). Among them, the highest percentages of SSRs were mononucleotides (77.65%), followed by dinucleotides (12.40%), tetranucleotides (4.64%), trinucleotides (3.33%), pentanucleotides (1.22%), and hexanucleotides (0.76%) (Fig. 2E). Furthermore, we found that the Zosteraceae species had more SSRs, and A/T was the main mononucleotide SSR type in all species (Fig. 2B). In addition, we identified 1,757 long repeat sequence types in the cp genome sequences of the 12 seagrass species, including 1,182 tandem repeats, 279 palindromic repeats, and 296 dispersed repeats, respectively (Fig. 2C). The lengths of these long repeats were highly variable, with repeats of 1–20, 21–30, and 31–40 bp being more abundant among the three repeat types (Fig. 2D). Among them, Hydrocharitaceae species had more long repeat sequences (especially tandem repeats) and > 81 bp repeat types.

Fig. 2
figure 2

Analysis of the repeated sequences of the 12 chloroplast genomes of seagrasses compared in this study. A: The number of SSRs in six SSR types; B: summary of different SSR repeat unit types; C: the number of repeat elements in three long repeat types; D: summary of long repeat types by length; E: distribution of SSR types; and F: distribution of long repeat types

Comparative analysis

The overall sequence identity of the cp genomes of 12 seagrass species was visualized using the mVISTA program, using the E. acoroides annotated sequences as a reference. Figure 3 shows a genome-wide alignment with high sequence similarity (> 90% identity). The cp genomes of the same family showed higher similarity. The divergence level of the non-coding regions was higher than the coding regions. Additionally, the LSC and SSC regions showed a higher level of sequence divergence than the IR regions. In addition, 59 coding regions were extracted to calculate the nucleotide variability (Supplementary Fig. 1), and the loci with the largest variation were accD, clpP, infA, rpl22, rps15, rps18, and ycf1.

Fig. 3
figure 3

Visualization of the alignment of 12 chloroplast genome sequences of seagrasses. The chloroplast genome of Enhalus acoroides was used as the reference. The Y-axis depicts percent identity to the reference genome (50–100%) and the X-axis depicts sequence coordinates within the chloroplast genome

Phylogenetic relationships

To investigate the phylogenetic relationships of the seagrasses, we constructed a maximum likelihood (ML) tree using RAxML (Fig. 4). As indicated in the tree, the shared nucleotide-coding genes were divided into four major clades: Zosteraceae, Cymodoceaceae, Hydrocharitaceae, and Ruppiaceae, among which Cymodoceaceae and Ruppiaceae had a relatively close relationship. Moreover, Z. muelleri was more closely related to Z. japonica, followed by Z. nigricaulis and Z. marina within Zosteraceae. These results showed that phylogenetic proximity was associated with the traditional taxonomic group. The results of the MCMCtree analysis of species divergence times in seagrasses are also shown in Fig. 4. The results showed that the divergence time between S. polyrhiza and seagrasses was about 125.6 Mya (95% PHD = 117.7–133.9 Mya, calibration point = 128.0 Mya). The diversification of Hydrocharitaceae, Zosteraceae, Ruppiaceae, and Cymodoceaceae in Alismatales was about 65.54–101.62 Mya. Within Zosteraceae, the divergence between Phyllospadix and Zostera was approximately 38.29 Mya. The divergence time of Z. marina (mean age of 20.27 Mya) was much earlier than that of Z. nigricaulis (mean age = 7.97 Mya), and that of Z. muelleri and Z. japonica (mean age = 3.54 Mya).

Fig. 4
figure 4

Phylogenetic ML tree and time-calibrated phylogenetic tree of 59 single copy genes of 12 seagrasses

Adaptive evolution analysis

Using the M8 model and the Datamonkey web server SLAC, FEL, and MEME methods, a total of nine genes were detected with positive selection sites (Supplementary Table S1), including two ATP subunit genes (atpA and atpF), two ribosome small subunit genes (rps4 and rps7), one photosystem subunit gene (psbH), one ribosome large subunit gene (rpl20), and the ycf2, accD, and rbcL genes. Among them, the ycf2 gene harbored eight sites under positive selection, with two in accD, while the other seven genes each had only one positively selected site.

Discussion

Chloroplast sequence variation

The cp genomes of the 12 seagrasses in this study ranged in size from 143,877 to 178,261 bp, which is similar to most angiosperms [23]. However, the size of the cp genome within the seagrasses varied widely, with those of the species of Hydrocharitaceae being much larger than those of the other three families (Table 1). These differences in cp genome size may have been due to the expansion and contraction of the border positions between the IR and SSC regions [24, 25]. The GC content acts as a landmark for the physical location of functional elements in the genome [26] and is positively correlated with the rate of recombination and exon density [27]. Generally, cp genomes are characterized by a low GC content; however, in this study, Hydrocharitaceae had a higher GC content (38.4%–39.2%) than other seagrasses. Moreover, we used mVISTA to compare the whole cp of 12 species of seagrasses and used DnaSP to analyze the percentage of variable loci in 59 single-copy shared genes. Similar to previous results, the variation in noncoding regions was more significant than that of the coding regions [24, 28]. Genes with a relatively high mutation rate can be used as DNA barcodes to help distinguish between accessions within a given taxon and varieties in germplasm resources [29, 30]. We detected seven hot spots in coding regions, which can be used as candidate DNA barcodes for assessing the phylogenetic relationships and interspecific differences among seagrasses.

In addition, the 12 cp genomes of the seagrasses contained a high frequency of large repeats. Previous studies have suggested that larger and complex repeat sequences have played key roles in sequence rearrangements and cp genome evolution [31, 32]. In this study, we found that the Hydrocharitaceae genomes had the highest numbers of repeats and a significant correlation (R = 0.73, P = 0.0065) between the size of the cp genomes and the number of repeats (Supplementary Fig. 2). Furthermore, tandem repeats were the most abundant type of repeat in the Hydrocharitaceae cp genomes, which differs from the high content of dispersed repeats reported in other angiosperm lineages [33, 34]. From an evolutionary perspective, variations in repeated sequences among species are due to natural selection and adaptation by organisms to diverse environments [35]. Previous studies have shown that cp SSRs are dominated by A/T repeats, which contribute significantly to the AT richness of cp genomes [36]. In this study, SSRs in the seagrass cp genome also contained many AT units, and mononucleotide repeats accounted for 77.65% of the total SSRs. In addition, we found that dinucleotide repeats (AT/AT) were the most common of the different types in all the seagrasses, while pentanucleotide and hexanucleotide SSRs were rarely found. The copy number variation of SSRs in the cp genome was highly polymorphic, and these seagrass variants can be used as molecular genetic markers for future studies in population genetics, phylogeography, phylogenetics, and species identification [37,38,39].

Adaptive selection

In our study, we identified nine genes with positive selection sites. Among them, ATP synthase is essential for plant photosynthesis and is usually a product of two genetic systems in plants [40]. In this study, we found that two ATP subunit genes (atpA and atpF) exhibited site-specific selection. Additionally, three genes (rps4, rps7, and rpl20) encoding ribosomal subunits were identified under positive selection. One photosystem II subunit gene (psbH) was also under positive selection. ACCase is a key enzyme in fatty acid biosynthesis [41]. The β-CT subunit of ACCase is encoded by accD, which is the only component of plant lipid metabolism known to be encoded by the plastid genome [42, 43]. We identified positively selected sites (PSSs) in accD, which may have played key roles in seagrass fatty acid biosynthesis. Additionally, rbcL provides all the catalytically essential residues of RuBisCO, a critical enzyme for both the reductive and oxidative photosynthetic carbon cycles. A previous study showed that many aquatic plants had acquired carbon-concentrating mechanisms to overcome the potentially low, fluctuating supply of CO2 for underwater photosynthesis [44]. In particular, the rbcL gene evolved under positive selection in Potamogeton [45]. Although seagrasses can also use both CO2 and HCO3 (bicarbonate) for photosynthetic carbon reduction [46], they live mainly in seawater with low CO2 concentration, and so positive selection of this gene may be involved in increasing their CO2 utilization efficiency. We also found that ycf2 had eight sites under positive selection. This gene is the largest cp gene reported in angiosperms and is valuable for assessing sequence variations and evolutionary processes in plants [24, 47, 48]. Positive selection on the ycf2 gene has also been found in the adaptive evolution of other species [28, 49, 50]. In short, these positively selected genes may play a key role in the adaptation of seagrasses to the marine environment.

Phylogenetic relationships

Seagrasses are a paraphyletic group of marine angiosperms that evolved in parallel three to four times from land plants back to the sea [7, 17]. Therefore, seagrass is a generic term for a variety of aquatic angiosperms and therefore represents an ecological group rather than a taxonomic group. In this study, the phylogenetic trees of the seagrasses based on the whole-genome nucleotide coding sequences of the cp genomes were clearly divided into four major clades, representing Zosteraceae, Cymodoceaceae, Hydrocharitaceae, and Ruppiaceae, which is similar to previous analyses based on various molecular datasets [7, 24, 51,52,53]. Divergence time estimation dated the divergence of seagrasses and S. polyrhiza at around 125.6 Mya. This age is consistent with previous genome-based studies [17, 54]. The stem node ages of the seagrass lineages Zosteraceae and Cymodoceaceae/Ruppiaceae were at 84.22 Mya, which is slightly older than that in recent studies [52, 55]. The divergence time of E. acoroides/T. hemprichii and Halophila within Hydrocharitaceae was around 30.78 Mya. This age is slightly older than that reported by Li et al. (2012) at around 19.41 Mya [56], but is younger than that reported by Li et al. (2022) at around 34.64 Ma [52]. Among seagrasses, all species of Zosteraceae are seagrasses, whereas other families contain more than just seagrass species. Among Zosteraceae, the genus Phyllospadix has a clear taxonomic status owing to its morphological features and chromosome number that differs significantly from other species [57,58,59]. In this study, we also found that P. iwatensis was the first to diverge from the species of the Zosteraceae family in 38.29 Ma. This age is consistent with previous phylogenetic analyses and molecular clock estimates based on the cp rbcL and matK loci, which suggested that the family Zosteraceae emerged about 100 Ma, and the divergence of Zostera and Phyllospadix began around 36 Ma [60]. Fossil evidence indicates that seagrasses originated in the Late Cretaceous [61]. As higher angiosperms, seagrasses have an existing root system and can occupy a previously empty niche in shallow sedimentary shoreline marine systems. This may have been facilitated by the Cretaceous-Paleogene extinction event, which took place roughly 70–65 Mya and coincided with a lineage-specific whole-genome duplication in Zostera [17].

Conclusions

In this study, we sequenced the cp genomes of two seagrass species (Z. muelleri and H. ovalis) and revealed the cp genomic features with the available seagrass genomes obtained from the NCBI database. We screened out 2,751 SSRs and 1,757 long repeat sequence types in the cp genome sequences of the 12 seagrass species. We also identified nine positive selection genes and seven variable regions, which provide a reference for developing DNA markers and evaluating adaptive evolution in further studies of seagrass species. Phylogenetic and divergence time analysis based on the current data was generally consistent with previous studies. These findings will be valuable for further study of the cp genomes of seagrass species, and will provide valuable resources for studies of plant adaptation to marine environments.

Methods

Sampling, DNA extraction, and sequencing

In this study, two seagrasses, Z. muelleri and H. ovalis, were collected from their natural habitats, and the collection processes of both conformed to local and national regulations. The voucher specimens of Z. muelleri (voucher number: OUC-S120) and H. ovalis (voucher number: OUC-S121) were deposited in the Herbarium of Marine Ecology Laboratory, College of Marine Life Sciences, Ocean University of China (OUC). The samples were identified by Tang Xuexi, a Professor at OUC. Fresh leaves (100 mg) were preserved in silica gel immediately after sampling, after which TRIzol® reagent (Invitrogen, USA) was used in the laboratory to extract their total DNA following the manufacturer’s protocol. The integrity, quality, and concentration of the DNA were determined by 1% agarose gel electrophoresis and a NanoDrop Spectrophotometer 2000 (Thermo Fisher Scientific, Waltham, MA, USA). Illumina TruSeq™ Nano DNA Sample Prep Kits were used for Illumina sequencing library construction, and the DNA was sequenced using an Illumina HiSeq 4000 platform (150 bp*2). After Illumina sequencing, approximately 63,351.8 Mb and 97,155.8 Mb of raw data for Z. muelleri and H. ovalis were generated, and these raw reads were QC-filtered and trimmed using the Trimmomatic 0.39 software [62]. In total, 61,803.7 Mb and 91,116.5 Mb clean data were obtained. Afterwards, Z. marina was used as a reference sequence to assemble its genome using the NOVOPlasty v2.7.2 software [63], and the GapCloser software [64] was used to fill in the remaining local internal gaps and correct for single-base polymorphisms. Finally, the starting position and orientation of the chloroplast assembly sequence were determined using the reference genome, and the possible partition structure of the chloroplast (LSC/IR/SSC) was determined to obtain the final chloroplast genome sequence.

Genome annotation and comparative analysis

The assembled cp genomes were annotated using GeSeq software [65], with the following parameters: (1) protein search identity, 60; (2) rRNA, tRNA, DNA search identity, 35; and (3) third-party tRNA annotators, the tRNAscan-SE software [66]. Then, the annotation results were checked with BLAST and DOGMA [67]. The circular maps of the cp genomes of the two seagrass samples were presented using Organellar Genome DRAW software [68]. Subsequently, two complete cp genomes were deposited in GenBank with the following accession numbers: Z. muelleri (OP611572) and H. ovalis (OP611573). Alignments of the 12 complete cp seagrass genome sequences (Table 1) were visualized using mVISTA [69]. Finally, all shared coding regions of the extracted 12 seagrass cp genomes were aligned separately using MUSCLE [70], and the nucleotide variability of each selected region was evaluated separately using DNASP v5.10 [71].

Repeat sequence analysis

Repeat sequences within the seagrass cp genome, including forward (F), reverse (R), palindrome (P), and complement (C), were searched using the online software REPuter (https://bibiserv.cebitec.uni-bielefeld.de/reputer) [72], and identified with the following conditions: minimal size of 30 bp; 90% or greater sequence identity; and Hamming distance equal to 3. Tandem Repeats Finder v4.09 software (https://tandem.bu.edu/trf/trf.html) [73] was used to detect tandem repeats with > 6 bp repeat units. SSR markers were identified in the 12 seagrasses sequences using MISA (http://pgrc.ipk-gatersleben.de/misa/misa.html) [74] with a motif size of 1–6 nucleotides and thresholds of eight, five, four, three, three, and three repeat units for mono-, di-, tri-, tetra-, penta-, and hexanucleotide SSRs, respectively.

Phylogenetic relationships and divergence time estimation

Phylogenetic analysis of the 12 cp genomes of the seagrass species was performed with Spirodela polyrhiza as the outgroup. All shared nucleotide-coding sequences were concatenated into a super matrix and aligned with MUSCLE [70]. A phylogenetic tree was produced using the ML method based on the GTRGAMMA model with 1,000 bootstrap replicates using RAxML [75]. MCMCtree implemented in the PAML 4.7 package was used to estimate the speciation time [76]. The obtained ML tree was used as the input tree file for the analysis process. Three calibration points (E. acoroides vs. T. hemprichii: 8.59–34.64 MYA, Zostera vs. P. iwatensis: 15.29–42.00 MYA, and Zosteraceae vs. P. Hydrocharitaceae: 72.70–117.00 MYA) derived from the TimeTree database (http://www.timetree.org/) were applied to constrain the divergence times of the nodes.

Adaptive evolution analysis

The CODEML program in the PAML 4.7 package [76] was used to calculate the rate of nonsynonymous substitutions (dN) and synonymous substitutions (dS) of single-copy protein coding genes (PCGs). The dN/dS ratio, also known as the ω value, was used to measure the rate of gene evolution, where ω values greater than 1, equal to 1, and less than 1 represent positive selection, neutral selection and purifying selection, respectively. The site model allows for different ω values for different sites in the same sequence. The M7 (null hypothesis: 0 < ω < 1) and M8 (alternative hypothesis: ω > 1) models were used to detect PSSs on 12 seagrass species. In the likelihood ratio test (LRT) results of the two models, the alternative hypothesis M8 model was accepted if p < 0.05, and the null hypothesis M7 model was accepted if this was not true. With p < 0.05, the PSSs detected using the M8 model were considered as potential PSSs when the Bayes empirical Bayes (BEB) posterior probability was > 0.90 [77]. Meanwhile, PSSs were identified using the Data Monkey Web Server (http://www.datamonkey.org/) [78] based on three methods, namely the FEL, SLAC, and MEME models, with P < 0.1. In this study, positive selection sites detected using at least three or more of the methods above were considered to be positive selection sites.

Availability of data and materials

All newly sequenced chloroplast genomic sequences of the two seagrasses in this study can be downloaded from the National Center for Biotechnology Information (NCBI) under accession numbers OP611572 (https://www.ncbi.nlm.nih.gov/nuccore/OP611572) and OP611573 (https://www.ncbi.nlm.nih.gov/nuccore/OP611573). Information on the chloroplast genomic sequences of the other 10 seagrasses used for the combined analysis can also be downloaded from GenBank and found in Table 1.

Abbreviations

cp:

Chloroplast

LSC:

Large single copy

SSC:

Small single copy

IRs:

Inverted repeats

SSR:

Simple sequence repeat

ML:

Maximum likelihood

dN :

Nonsynonymous substitutions

dS :

Synonymous substitutions

LRT:

Likelihood ratio test

BEB:

Bayes empirical Bayes

rRNA:

Ribosomal RNA genes

tRNA:

Transfer RNA genes

PCGs:

Protein coding genes

PSSs:

Positively selected sites

References

  1. Suchanek TH, Williams SL, Ogden JC, Hubbard DK, Gill IP. Utilization of shallow-water seagrass detritus by Carribbean deep-sea macrofauna: δ13C evidence. Deep Sea Res Part A Oceanographic Res Papers. 1985;32(2):201–14.

    Article  Google Scholar 

  2. Hemminga MA, Duarte CM. Seagrass ecology. Cambridge: Cambridge University Press; 2000.

  3. Duarte CM, Middelburg JJ, Caraco N. Major role of marine vegetation on the oceanic carbon cycle. Biogeosciences. 2005;2(1):1–8.

    Article  CAS  Google Scholar 

  4. Orth RJ, Carruthers TJ, Dennison WC, Duarte CM, Fourqurean JW, Heck KL, Hughes AR, Kendrick GA, Kenworthy WJ, Olyarnik S. A global crisis for seagrass ecosystems. Bioscience. 2006;56(12):987–96.

    Article  Google Scholar 

  5. Short FT, Polidoro B, Livingstone SR, Carpenter KE, Bandeira S, Bujang JS, Calumpong HP, Carruthers TJ, Coles RG, Dennison WC. Extinction risk assessment of the world’s seagrass species. Biol Cons. 2011;144(7):1961–71.

    Article  Google Scholar 

  6. Huang X-P, Jiang Z-J, Fan H-Q, Chen X-Y, Zhou Y, Zhang J-P, Li W-T, Zhang P-D, Zheng F-Y, Han Q-Y. The nomenclature of the “algae” name of seagrasses in China. Oceanologia et Limnologia Sinica. 2016;47(1):290–4.

    Google Scholar 

  7. Les DH, Cleland MA, Waycott M. Phylogenetic studies in Alismatidae, II: evolution of marine angiosperms (seagrasses) and hydrophily. Systematic Botany. 1997;22:443–463.

  8. Janssen T, Bremer K. The age of major monocot groups inferred from 800+ rbcL sequences. Bot J Linn Soc. 2004;146(4):385–98.

    Article  Google Scholar 

  9. Larkum AW, Orth RJ, Duarte CM. Seagrasses: biology, ecologyand conservation. Phycologia. 2006;45(5):5.

    Google Scholar 

  10. Beer S, Bjork M, Hellblom F, Axelsson L. Inorganic carbon utilization in marine angiosperms (seagrasses). Funct Plant Biol. 2002;29(3):349–54.

    Article  CAS  PubMed  Google Scholar 

  11. Borum J, Pedersen O, Kotula L, Fraser MW, Statton J, Colmer TD, Kendrick GA. Photosynthetic response to globally increasing CO2 of co-occurring temperate seagrass species. Plant, Cell Environ. 2016;39(6):1240–50.

    Article  CAS  PubMed  Google Scholar 

  12. Ye C, Zhao K. Osmotically active compounds and their localization in the marine halophyte eelgrass. Biol Plant. 2003;46(1):137–40.

    Article  CAS  Google Scholar 

  13. Waycott M, Procaccini G, Les DH, Reusch T. Seagrass Evolution, Ecology and Conservation: A Genetic Perspective. Netherlands: Springer; 2007.

    Google Scholar 

  14. Lee H, Golicz AA, Bayer PE, Severn-Ellis AA, Chan CK-K, Batley J, Kendrick GA, Edwards D. Genomic comparison of two independent seagrass lineages reveals habitat-driven convergent evolution. J Exp Bot. 2018;69(15):3689–702.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Lambers H, Chapin FS, Pons TL. Plant physiological ecology. New York: Springer; 2008.

  16. Lee H, Golicz AA, Bayer PE, Jiao Y, Tang H, Paterson AH, Sablok G, Krishnaraj RR, Chan CK-K, Batley J. The genome of a southern hemisphere seagrass species (Zostera muelleri). Plant physiology. 2016;172(1):272–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Olsen JL, Rouzé P, Verhelst B, Lin Y-C, Bayer T, Collen J, Dattolo E, De Paoli E, Dittami S, Maumus F. The genome of the seagrass Zostera marina reveals angiosperm adaptation to the sea. Nature. 2016;530(7590):331–5.

    Article  CAS  PubMed  Google Scholar 

  18. Howe CJ, Barbrook AC, Koumandou VL, Nisbet RER, Symington HA, Wightman TF. Evolution of the chloroplast genome. Philos Trans R Soc Lond B Biol Sci. 2003;358(1429):99–107.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Wicke S, Schneeweiss GM, Depamphilis CW, Müller KF, Quandt D. The evolution of the plastid chromosome in land plants: gene content, gene order, gene function. Plant Mol Biol. 2011;76(3):273–97.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Daniell H, Lin C-S, Yu M, Chang W-J. Chloroplast genomes: diversity, evolution, and applications in genetic engineering. Genome Biol. 2016;17(1):1–29.

    Article  Google Scholar 

  21. Sugiura M. The chloroplast genome. Plant Mol Biol. 1992;19(1):149–68.

    Article  CAS  PubMed  Google Scholar 

  22. Korpelainen H. The evolutionary processes of mitochondrial and chloroplast genomes differ from those of nuclear genomes. Naturwissenschaften. 2004;91(11):505–18.

    Article  CAS  PubMed  Google Scholar 

  23. Zhang T, Fang Y, Wang X, Deng X, Zhang X, Hu S, Yu J. The complete chloroplast and mitochondrial genome sequences of Boea hygrometrica: insights into the evolution of plant organellar genomes. PLoS ONE. 2012;7(1):e30531.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Chen J, Zang Y, Shang S, Liang S, Zhu M, Wang Y, Tang X. Comparative chloroplast genomes of Zosteraceae species provide adaptive evolution insights into seagrass. Front Plant Sci. 1992;2021:12.

    Google Scholar 

  25. Wang W, Messing J. High-throughput sequencing of three Lemnoideae (duckweeds) chloroplast genomes from total DNA. PLoS ONE. 2011;6(9):e24670.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Zhang L, Kasif S, Cantor CR, Broude NE. GC/AT-content spikes as genomic punctuation marks. Proc Natl Acad Sci. 2004;101(48):16855–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Freudenberg J, Wang M, Yang Y, Li W. Partial correlation analysis indicates causal relationships between GC-content, exon density and recombination rate in the human genome. BMC Bioinformatics. 2009;10(1):1–10.

    Google Scholar 

  28. Fan W-B, Wu Y, Yang J, Shahzad K, Li Z-H. Comparative chloroplast genomics of dipsacales species: insights into sequence variation, adaptive evolution, and phylogenetic relationships. Front Plant Sci. 2018;9:689.

    Article  PubMed  PubMed Central  Google Scholar 

  29. Mehmood F, Shahzadi I, Waseem S, Mirza B, Ahmed I, Waheed MT. Chloroplast genome of Hibiscus rosa-sinensis (Malvaceae): comparative analyses and identification of mutational hotspots. Genomics. 2020;112(1):581–91.

    Article  PubMed  Google Scholar 

  30. Jiang L, Li M, Zhao F, Chu S, Zha L, Xu T, Peng H, Zhang W. Molecular identification and taxonomic implication of herbal species in genus Corydalis (Papaveraceae). Molecules. 2018;23(6):1393.

    Article  PubMed  PubMed Central  Google Scholar 

  31. Milligan BG, Hampton JN, Palmer JD. Dispersed repeats and structural reorganization in subclover chloroplast DNA. Mol Biol Evol. 1989;6(4):355–68.

    CAS  PubMed  Google Scholar 

  32. Cavalier-Smith T. Chloroplast evolution: secondary symbiogenesis and multiple losses. Curr Biol. 2002;12(2):R62–4.

    Article  CAS  PubMed  Google Scholar 

  33. Wu Y, Liu F, Yang D-G, Li W, Zhou X-J, Pei X-Y, Liu Y-G, He K-L, Zhang W-S, Ren Z-Y. Comparative chloroplast genomics of Gossypium species: insights into repeat sequence variations and phylogeny. Front Plant Sci. 2018;9:376.

    Article  PubMed  PubMed Central  Google Scholar 

  34. Haberle RC, Fourcade HM, Boore JL, Jansen RK. Extensive rearrangements in the chloroplast genome of Trachelium caeruleum are associated with repeats and tRNA genes. J Mol Evol. 2008;66(4):350–61.

    Article  CAS  PubMed  Google Scholar 

  35. Britten RJ, Kohne DE. Repeated Sequences in DNA: Hundreds of thousands of copies of DNA sequences have been incorporated into the genomes of higher organisms. Science (New York, NY). 1968;161(3841):529–40.

    Article  CAS  Google Scholar 

  36. Nie X, Lv S, Zhang Y, Du X, Wang L, Biradar SS, Tan X, Wan F, Weining S. Complete chloroplast genome sequence of a major invasive species, crofton weed (Ageratina adenophora). PLoS ONE. 2012;7(5):e36869.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Aradhya MK, Potter D, Gao F, Simon CJ. Molecular phylogeny of Juglans (Juglandaceae): a biogeographic perspective. Tree Genet Genomes. 2007;3(4):363–78.

    Article  Google Scholar 

  38. Wang S, Shi C, Gao L-Z. Plastid genome sequence of a wild woody oil species, Prinsepia utilis, provides insights into evolutionary and mutational patterns of Rosaceae chloroplast genomes. PLoS ONE. 2013;8(9):e73946.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Xue J, Wang S, Zhou SL. Polymorphic chloroplast microsatellite loci in Nelumbo (Nelumbonaceae). Am J Bot. 2012;99(6):e240–4.

    Article  PubMed  Google Scholar 

  40. Westhoff P, Alt J, Nelson N, Herrmann RG. Genes and transcripts for the ATP synthase CF 0 subunits I and II from spinach thylakoid membranes. Mol Gen Genet MGG. 1985;199(2):290–9.

    Article  CAS  Google Scholar 

  41. Rawsthorne S. Carbon flux and fatty acid synthesis in plants. Prog Lipid Res. 2002;41(2):182–96.

    Article  CAS  PubMed  Google Scholar 

  42. Ke J, Wen T-N, Nikolau BJ, Wurtele ES. Coordinate regulation of the nuclear and plastidic genes coding for the subunits of the heteromeric acetyl-coenzyme A carboxylase. Plant Physiol. 2000;122(4):1057–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Lee SS, Jeong WJ, Bae JM, Bang JW, Liu JR, Harn CH. Characterization of the plastid-encoded carboxyltransferase subunit (accD) gene of potato. Mol Cells. 2004;17(3):422–9.

    CAS  PubMed  Google Scholar 

  44. Maberly SC, Madsen TV. Freshwater angiosperm carbon concentrating mechanisms: processes and patterns. Funct Plant Biol. 2002;29(3):393–405.

    Article  CAS  PubMed  Google Scholar 

  45. Iida S, Miyagi A, Aoki S, Ito M, Kadono Y, Kosuge K. Molecular adaptation of rbcL in the heterophyllous aquatic plant Potamogeton. PLoS ONE. 2009;4(2):e4633.

    Article  PubMed  PubMed Central  Google Scholar 

  46. Beer S. Photosynthesis and photorespiration of marine angiosperms. Aquat Bot. 1989;34(1–3):153–66.

    Article  CAS  Google Scholar 

  47. Drescher A, Ruf S, Calsa T Jr, Carrer H, Bock R. The two largest chloroplast genome-encoded open reading frames of higher plants are essential genes. Plant J. 2000;22(2):97–104.

    Article  CAS  PubMed  Google Scholar 

  48. Huang JL, Sun GL, Zhang DM. Molecular evolution and phylogeny of the angiosperm ycf2 gene. J Syst Evol. 2010;48(4):240–8.

    Article  Google Scholar 

  49. Zhong Q, Yang S, Sun X, Wang L, Li Y. The complete chloroplast genome of the Jerusalem artichoke (Helianthus tuberosus L.) and an adaptive evolutionary analysis of the ycf2 gene. PeerJ. 2019;7:e7596.

    Article  PubMed  PubMed Central  Google Scholar 

  50. Wu Z, Liao R, Yang T, Dong X, Lan D, Qin R, Liu H. Analysis of six chloroplast genomes provides insight into the evolution of Chrysosplenium (Saxifragaceae). BMC Genomics. 2020;21(1):1–14.

    Article  Google Scholar 

  51. Ross TG, Barrett CF, Soto Gomez M, Lam VK, Henriquez CL, Les DH, Davis JI, Cuenca A, Petersen G, Seberg O. Plastid phylogenomics and molecular evolution of Alismatales. Cladistics. 2016;32(2):160–78.

    Article  PubMed  Google Scholar 

  52. Chen L-Y, Lu B, Morales-Briones DF, Moody ML, Liu F, Hu G-W, Huang C-H, Chen J-M, Wang Q-F. Phylogenomic analyses of Alismatales shed light into adaptations to aquatic environments. Mol Biol Evol. 2022;39(5):msac079.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Cuenca A, Petersen G, Seberg O, Davis JI, Stevenson DW. Are substitution rates and RNA editing correlated? BMC Evol Biol. 2010;10(1):1–15.

    Article  Google Scholar 

  54. An D, Zhou Y, Li C, Xiao Q, Wang T, Zhang Y, Wu Y, Li Y, Chao D-Y, Messing J. Plant evolution and environmental adaptation unveiled by long-read whole-genome sequencing of Spirodela. Proc Natl Acad Sci. 2019;116(38):18893–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Coyer J, Hoarau G, Kuo J, Tronholm A, Veldsink J, Olsen J. Phylogeny and temporal divergence of the seagrass family Zosteraceae using one nuclear and three chloroplast loci. Syst Biodivers. 2013;11(3):271–84.

    Article  Google Scholar 

  56. Chen L-Y, Chen J-M, Gituru RW, Wang Q-F. Generic phylogeny, historical biogeography and character evolution of the cosmopolitan aquatic plant family Hydrocharitaceae. BMC Evol Biol. 2012;12(1):1–12.

    Article  Google Scholar 

  57. Kuo J, Den Hartog C. Seagrass taxonomy and identification key. Glob Seagrass Res Methods. 2001;33:31–58.

    Article  Google Scholar 

  58. Kuo J. Chromosome numbers of the Australian Zosteraceae. Plant Syst Evol. 2001;226(3):155–63.

    Article  Google Scholar 

  59. Tanaka N, Kuo J, Omori Y, Nakaoka M, Aioi K. Phylogenetic relationships in the genera Zostera and Heterozostera (Zosteraceae) based on matK sequence data. J Plant Res. 2003;116(4):273–9.

    Article  PubMed  Google Scholar 

  60. Kato Y, Aioi K, Omori Y, Takahata N, Satta Y. Phylogenetic analyses of Zostera species based on rbcL and matK nucleotide sequences: implications for the origin and diversification of seagrasses in Japanese waters. Genes Genet Syst. 2003;78(5):329–42.

    Article  CAS  PubMed  Google Scholar 

  61. Larkum AW, Waycott M, Conran JG. Evolution and biogeography of seagrasses. In: Seagrasses of Australia. Cham: Springer; 2018. pp. 3–29.

  62. Bolger AM, Marc L, Bjoern U. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics (Oxford, England). 2014;15:2114–20.

    Google Scholar 

  63. Dierckxsens N, Mardulyn P, Smits G. NOVOPlasty: de novo assembly of organelle genomes from whole genome data. Nucleic Acids Res. 2017;45(4):e18–e18.

    PubMed  Google Scholar 

  64. Luo R, Liu B, Xie Y, Li Z, Huang W, Yuan J, He G, Chen Y, Pan Q, Liu Y. SOAPdenovo2: an empirically improved memory-efficient short-read de novo assembler. Gigascience. 2012;1(1):2047-2217X-2041–2018.

    Article  Google Scholar 

  65. Tillich M, Lehwark P, Pellizzer T, Ulbricht-Jones ES, Fischer A, Bock R, Greiner S. GeSeq–versatile and accurate annotation of organelle genomes. Nucleic Acids Res. 2017;45(W1):W6–11.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Lowe TM, Eddy SR. tRNAscan-SE: a program for improved detection of transfer RNA genes in genomic sequence. Nucleic Acids Res. 1997;25(5):955–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Wyman SK, Jansen RK, Boore JL. Automatic annotation of organellar genomes with DOGMA. Bioinformatics (Oxford, England). 2004;20(17):3252–5.

    CAS  PubMed  Google Scholar 

  68. Greiner S, Lehwark P, Bock R. OrganellarGenomeDRAW (OGDRAW) version 1.3. 1: expanded toolkit for the graphical visualization of organellar genomes. Nucleic Acids Res. 2019;47(W1):W59–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Frazer KA, Pachter L, Poliakov A, Rubin EM, Dubchak I. VISTA: computational tools for comparative genomics. Nucleic Acids Res. 2004;32(suppl_2):W273–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Edgar RC. MUSCLE: Multiple Sequence Alignment with Improved Accuracy and Speed. In: Computational Systems Bioinformatics Conference. 2004.

    Google Scholar 

  71. Librado P, Rozas J. DnaSP v5: a software for comprehensive analysis of DNA polymorphism data. Bioinformatics (Oxford, England). 2009;25(11):1451–2.

    CAS  PubMed  Google Scholar 

  72. Kurtz S, Choudhuri JV, Ohlebusch E, Schleiermacher C, Stoye J, Giegerich R. REPuter: the manifold applications of repeat analysis on a genomic scale. Nucleic Acids Res. 2001;29(22):4633–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Benson G. Tandem repeats finder: a program to analyze DNA sequences. Nucleic Acids Res. 1999;27(2):573–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Beier S, Thiel T, Münch T, Scholz U, Mascher M. MISA-web: a web server for microsatellite prediction. Bioinformatics (Oxford, England). 2017;33(16):2583–5.

    CAS  PubMed  Google Scholar 

  75. Stamatakis A. RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics (Oxford, England). 2014;30(9):1312–3.

    CAS  PubMed  Google Scholar 

  76. Yang Z. PAML 4: Phylogenetic Analysis by Maximum Likelihood. Mol Biol Evol. 2007;24(8):1586–91.

    Article  CAS  PubMed  Google Scholar 

  77. Ziheng Y, Wong WSW, Rasmus N. Bayes empirical bayes inference of amino acid sites under positive selection. Mol Biol Evol. 2005;22(4):1107–18.

    Article  Google Scholar 

  78. Pond SLK, Frost SDW. Datamonkey: rapid detection of selective pressure on individual sites of codon alignments. Bioinformatics. 2005;21(10):2531–3.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We are grateful to the editors and reviewers for their valuable suggestions, which helped us improve the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (42206117, 42176154, 42006144), the NSFC-Shandong Joint Fund (U1806213), and the National Key R&D Program of China (2019YFD0901204).

Author information

Authors and Affiliations

Authors

Contributions

JC, YZ and XXT collected and identified plant material and conducted the project. JC conducted the experiments, analyzed the data and drafted the manuscript. ZBY, LS and SX supervised the study. SS, YW and XXT revised the manuscript. All authors have read and approved the final draft of the manuscript.

Corresponding authors

Correspondence to Ying Wang or Xuexi Tang.

Ethics declarations

Ethics approval and consent to participate

We complied with all relevant institutional, national and international guidelines for the collection of Z. muelleri and H. ovalis with permissions from Ocean University of China.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interest.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1: Supplementary Figure 1.

Percentages of variable characters in protein-coding regions among the 12 chloroplast genomes of the seagrasses.

Additional file 2: Supplementary Figure 2.

Relationships between seagrass chloroplast genome sizes and the number of repeats.

Additional file 3: Table S1.

Positively selected sites of 59 single-copy genes shared by twelve seagrass species.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chen, J., Zang, Y., Shang, S. et al. Chloroplast genomic comparison provides insights into the evolution of seagrasses. BMC Plant Biol 23, 104 (2023). https://doi.org/10.1186/s12870-023-04119-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12870-023-04119-9

Keywords