Skip to main content

ORIGINAL RESEARCH article

Front. Plant Sci. , 17 February 2025

Sec. Plant Systematics and Evolution

Volume 16 - 2025 | https://doi.org/10.3389/fpls.2025.1550467

The organelle genomes of the endangered seagrass Zostera caespitosa reveal sequence divergences, massive gene transfer, and uncommon RNA editing types

Yushun YongYushun Yong1Yulian WangYulian Wang2Dawei WangDawei Wang2Xingfang Yuan*Xingfang Yuan2*Quansheng Zhang*Quansheng Zhang1*
  • 1Ocean School, Yantai University, Yantai, China
  • 2No. 6 Geological Team, Shandong Provincial Bureau of Geology and Mineral Resources, Weihai, China

Introduction: Zostera caespitosa, a rare submerged angiosperm, is considered endemic to the northwestern Pacific.

Methods: This study assembled and compared the mitochondrial (mt) and chloroplast (cp) genomes of Z. caespitosa to understand the organelle evolutionary patterns.

Results and discussion: The cp genome (143,972 bp) was the second smallest within the seagrasses, whereas the mt genomes (192,246 bp) of Z. caespitosa and other seagrasses were smaller compared to those of other monocotyledons. The protein-coding genes (PCGs) in the organelle genome exhibit a strong A/U bias at codon endings, a selection-driven codon bias. The rates of nonsynonymous (Ka) and synonymous (Ks) substitutions in the mt genes of Zostera were two times higher than those in the cp genes. Additionally, 50 mitochondrial plastid DNA (MTPT) segments, totaling 44,662 bp, were identified, constituting 23.23% of the mt genome, which is significantly higher than those in most land plants. Phylogenetic analysis of 13 seagrass core cp-PCGs supported previous studies showing two genera in family Zosteraceae: Phyllospadix and Zostera, the latter comprising Zostera and Zosterella as subgenera. RNA editing was remarkably abundant in the 167 mt-PCGs and 172 in cp-PCGs, particularly in the cp genome. There are 11 different RNA editing types in the cp and 3 in the mt, most of which are C to U. Unexpectedly rare editing events, such as A to C, A to U, U to A, G to C, and U to G, have also been found in the cp.

1 Introduction

Zostera caespitosa is a submerged perennial herb belonging to the class Monocotyledoneae, order Alismatales, family Zosteraceae, and genus Zostera. The distribution of Z. caespitosa is local, with only a few populations found along the east coast of Korea, the southern coast of Japan, and the coastal areas of Liaoning and Shandong in China (Short et al., 2007; Xu et al., 2021). Z. caespitosa is among the typical representative species in seagrass beds, playing a vital role in erosion protection, bacterial suppression, nutrient cycling, and significantly regulating carbon sequestration (Fourqurean et al., 2012; Lamb et al., 2017; Unsworth et al., 2022). However, the increasing pressures from human activities, such as ocean warming, coastal modification, and water quality degradation, are causing global seagrass loss (Orth et al., 2006; Waycott et al., 2009). This marine plant species was listed as “endangered” on the IUCN and China Biodiversity - Higher Plants Volume in 2021 and 2022 Red Lists, respectively (Short et al., 2011; Jiang et al., 2024). Previous studies on Z. caespitosa have mainly focused on its distribution and biological traits (Short et al., 2007; Xu et al., 2021; Jiang et al., 2024; Im et al., 2024). Nevertheless, it lacks basic genetic resources, limiting further research and conservation efforts.

Mitochondrial (mt) and chloroplast (cp) are endosymbiotic organelles with independent genetic material separate from those in the cell nucleus (Birky, 2001). These organelles share common traits, including replication methods, mutation patterns, and inheritance mechanisms and are significantly different in land plants (Sloan et al., 2012; Maréchal and Brisson, 2010; Birky, 1995). Plant mt genomes are typically larger and more complex, with a wide range of genome sizes (ca. 100–10,000 kb), varied structures, low density of genes, and numerous repetitive sequences (Jansen and Ruhlman, 2012), making their conformation challenging (Gualberto et al., 2014). In contrast, cp genomes of angiosperms exhibit a relatively simple structure, typically consisting of approximately 120–130 genes and a smaller size (~107–218 kb) (Daniell et al., 2016). These genomes comprise a conserved four-part structure with large single copy (LSC), small single copy (SSC), and inverted repeats (IRs) region (Henry, 2005; Jansen and Ruhlman, 2012). Thus, they are ideal systems for investigating the phylogenetic relationships among different plant species (Mower and Vickrey, 2018; Nizam et al., 2022). In addition, the RNA editing process was extremely abundant, diverse, and remarkably complex in organellar RNAs (Kaur, 2020; Takenaka et al., 2013). To date, no research has been conducted on identifying RNA editing events through RNA-Seq read mapping in Z. caespitosa.

Recent developments in sequencing technologies have significantly increased the number of plant cp and mt genomes. Currently, the NCBI database contains approximately 13,000 cp genomes and 673 mt genomes, yet only 285 species have both genomes assembled (Wang et al., 2024). Although various evolutionary patterns have been proposed for these genomes, knowledge gaps persist, primarily due to unequal sampling. Seagrasses, comprising approximately 74 species, represent a crucial transformative event in higher plant evolution (Olsen et al., 2016). Unfortunately, complete organelle genomes have been published for only five seagrass species: Z. marina, Z. japonica, Phyllospadix iwatensis, Ruppia sinensis, and Cymodocea nodosa (Petersen et al., 2017; Chen et al., 2021, 2022, 2023; Ma et al., 2024). In the Yellow–Bohai Sea, a temperate seagrass habitat in the North Pacific, five species from three genera (Zostera, Phyllospadix, and Ruppia) have currently been identified, namely, Z. marina, Z. japonica, Z. caespitosa, P. iwatensis, and R. sinensis (Xu et al., 2021). Recent research has extensively used mt and cp genomes to explore molecular evolution (Li et al., 2023; Huang et al., 2020). As a major member of the Yellow–Bohai Seas seagrass species, the differences between the mt and cp genomes of Z. caespitosa remain uncharacterized, limiting insights into its origin and adaptive evolution.

Thus, this study sequenced, assembled, and annotated the organelle genome of Z. caespitosa. Our objectives were to (1) describe the features of the organelle genomes; (2) analyze evolutionary differences in sequences (repeated elements, codon bias, mutation rate, and phylogenetic relationships); (3) identify horizontal gene transfer (HGT) events; and (4) explore the characteristics of RNA editing events.

2 Results

2.1 Composition of the Z. caespitosa organelle genome

The complete circular mt genome of Z. caespitosa was 192,246 bp long with 45.63% in GC content (Figure 1A). Furthermore, the mt genome has 86.02%, 13.98%, 0.75%, and 2.71% intergenic regions, protein-coding genes (PCGs), tRNA, and rRNA, respectively (Supplementary Table S1). The mt genome encodes 50 genes, namely, 27 PCGs, 20 tRNA, and 3 rRNA (rrn5, rrn18, and rrn26) (Table 1). The PCGs comprised several genes, including five ATP synthase (atp1, atp4, atp6, atp8, and atp9), four cytochrome c biogenesis (ccmB, ccmC, ccmFc, and ccmFn), three cytochrome c oxidase (cox1, cox2, and cox3), and one ubiquinol cytochrome c reductase (cob). Additionally, PCGs have one maturase (matR), one membrane transport protein (mttB), nine NADH dehydrogenase (nad1, nad2, nad3, nad4, nad4L, nad5, nad6, nad7, and nad9), one ribosomal protein (SSU) (rps7), and two copies of atp6 and nad4L genes. Gene nad1, nad2, nad5, and nad7 possessed four introns, nad4 possessed three introns, and cox2 and ccmFc possessed one intron.

Figure 1
www.frontiersin.org

Figure 1. The organelle genomes map of Z. caespitosa. (A ,B) represent the mt and cpgenomes, respectively. The genes located inside the circles are transcribed in a clockwise direction, while those outside the circle are transcribed counterclockwise. Different colored genes represent different functions.

Table 1
www.frontiersin.org

Table 1. List of genes encoded by the mt genome of Z. caespitosa.

The complete cp genome of Z. caespitosa was 143,972 bp, and its typical structure includes an SSC of 12,405 bp, an LSC of 83,313 bp, and two IRs region, each 24,127 bp (Figure 1B and Supplementary Table S1). Further, a total of 131 genes were annotated in the Z. caespitosa cp genome, corresponding to 85 PCGs (counting genes in the inverted repeats twice), 38 rRNAs, and 8 tRNAs (Table 2). Seventeen genes had introns in the cp genome with two introns for three genes (clpP1, rps12, and ycf3) and one intron for the other 14 (ndhA, ndhB, petB, petD, atpF, rpl16, rpl2, rpoC1, trnA-UGC, trnG-UUC, trnI-GAU, trnK-UUU, trnL-UAA, and trnA-UAC).

Table 2
www.frontiersin.org

Table 2. List of genes encoded by the cp genome of Z. caespitosa.

2.2 Repeat sequence analysis

A total of 21 and 103 microsatellites [simple repeat sequences (SSRs)] were identified in the mt and cp genomes, respectively, which were 1 to 6 bp DNA fragments (Figure 2 and Supplementary Table S2). Motifs 1, 2, 3, 4, 5, and 6 were 1, 7, 2, 10, 0, and 1 mt repeats, respectively. In contrast, their cp repeats measured 71, 21, 0, 9, 0, and 2 bp, respectively (Figure 2C). Tetramer SSRs were the most dominant in the mt genome (47.6%), whereas monomer SSRs dominated the cp genome (68.9%). Furthermore, the mt and cp genomes had 110 and 39 tandem repeats, respectively, besides dispersed repeats that were also prevalent throughout these genomes (Figure 2 and Supplementary Table S3). The mt and cp genomes had 5,028 and 37 pairs of dispersed repeats, respectively, depicting significantly higher dispersed repeats in the mt genome (Figure 2 and Supplementary Tables S4, S5). There were 0 complementary, 2,847 forward, 2,165 palindromic, and 16 reverse repeats in the mt genome, while there were 0 complementary, 15 forward, 19 palindromic repeats, and 3 reverse repeats in the cp genome (Figure 2D). The organellar genomes lacked complementary repeats but had the abundant forward and palindromic repeats.

Figure 2
www.frontiersin.org

Figure 2. Repeat sequences within mt and cp genomes. Colored lines in the inner circle represent dispersed repeats: pink for forward, purple for palindromic, and orange for reverse repeats. The black line in the first circle marks tandem repeats, while the second circle highlights SSRs. (A, B) show the physical distribution of repeats in mt and cp genomes, while panels (C, D) illustrate the different types of dispersed repeats and SSRs.

2.3 Codon bias analysis

The different organelles displayed a degree of variation in the codon usage frequency. The Relative Synonymous Codon Usage (RSCU) values in the mt genome were 0.49 (GCG in alanine) to 1.5 (CAA in glutamine) and 0.28 (CUG in leucine) to 2.26 (UUA in leucine) in the cp genome (Figures 3A, B and Supplementary Table S6). Only UGG (tryptophan) and AUG (start codon) had RSCU values 1 in the mt and cp genome; the rest were >1 or <1, indicating a general preference for codons in PCGs. The RSCU values of 30 codons in the mt and 29 codons in the cp genomes exceeded 1, indicating that these codons showed biased usage. The Z. caespitosa organellar genomes have a strong A/U preference in codon usage.

Figure 3
www.frontiersin.org

Figure 3. Summary of codon usage in PCGs of the Z. caespitosa organelle genome: (A, B) show RSCU distribution for mt and cp genomes, with horizontal axes representing 21 amino acids and vertical axes showing RSCU values. (C, D) plot ENC against GC3s for mt and cp genes, respectively. The solid line represents the expected trend based on GC3s composition. Blue dots indicate respiratory chain or photosynthesis genes, brown dots mark translation-related/transcription genes, and pink dots represent other genes.

Effective codon numbers versus the third position GC content (ENC-GC3s) plots are commonly used to assess factors influencing codon usage patterns. The mt and cp genomes exhibited similar bias patterns, with most PCG genes deviating from the standard curve and a minority positioned on or close to it (Figures 3C, D and Supplementary Table S7). Respiration caused the remarkably pronounced deviations in mt, while photosynthesis primarily caused those in cp. The mt PCG gene atp9 has ENC values of 33.25 (less than 35). In the cp genome, PCG genes with <35 ENC included psbI (33.96), petN (28.65), psaI (34.04), psaJ (33.05), rpl36 (23.23), and rpl32 (28.76).

2.4 Mutation rate analysis of Zostera species

To resolve the variations in evolutionary rates across the organelle genomes of Zostera, we examined the rates of evolution for the PCGs common to all Zostera species. The rates of nonsynonymous (Ka) and synonymous (Ks) substitutions of cp genome is roughly 1/2 of that of the mt genome, indicating that the cp genome probably suffered strong purifying selection and is more conserved at the level of PCGs (Figure 4A). The Ka/Ks value also reveals that the document rate overlaps between loci across the two organelles. Comparing the Ka/Ks of each gene from both organelles corroborated that mt genes evolve faster than cp (Figures 4B,C). Most genes from all major organellar classes are under strong purifying selection. ATP synthase genes had particularly the lowest Ka/Ks ratios in the cp genome, while they had the highest Ka/Ks ratios in the mt genome. In contrast to photosystem genes, ribosomal genes possibly suffered weaker purifying selection, with significantly greater Ka/Ks values from mt than cp genes.

Figure 4
www.frontiersin.org

Figure 4. (A) The density distribution of Ka/Ks. (B, C) Boxplots show the median Ka/Ks for classes of mt and cp genes across the Zostera species.

2.5 Phylogenetic analysis

To investigate the phylogenetic relationship of Zosterceae, we analyzed the shared PCGs of the cp genome of 13 seagrass species. Spirodela pholyhiza was selected as the outgroup. Because of the high congruence between the ML and BI analyses of cp genomes, only the ML topology is presented here. Phylogenomic analysis using common 57 PCGs genes revealed four primary clades: Hydrocharitaceae, Cymodoceaceae, Ruppiaceae, and Zosteraceae, of which Cymodoceaceae and Ruppiaceae had a relatively close relationship (Figure 5). Furthermore, Zosteraceae were divided into three major clades: Phyllospadix, subgenus Zostera, and subgenus Zosterella. Within Zosteraceae, Z. caespitosa was more closely related to Z. marina, followed by Z. nigricaulis, Z. japonica, Z. muelleri, and P. iwatensis.

Figure 5
www.frontiersin.org

Figure 5. The ML tree of 13 seagrasses based on all shared cp-PCGs.

2.6 Identification of homologous sequences

Homologous sequences between the mt and cp genomes of Z. caespitosa were searched to further investigate gene residues from cp genome in the mt genome. Sequence similarity analysis identified 50 mitochondrial-plastid DNA transfers (MTPTs) in Z. caespitosa (Figure 6). The MTPTs were 44,662 bp long, accounting for 23.23% of the mt genome length and 31.02% of the total cp genome (Supplementary Table S8). Eleven fragments were longer than 1,000 bp, with the longest being MTPT1 (2,961 bp), which consists of the rrn4.5, rrn5, and rrn23 genes. Most MTPTs in the mt genome of Z. caespitosa are derived from the tRNA gene and rRNA gene in the cp genome. Two proteins (rbcL and rpl23), two rRNA genes (rrn4.5 and rrn5), and six tRNA genes (trnC-GCA, trnY-UAC, trnT-ACC, trnW-CCA, trnL-UUG, and trnR-CGU) were fully intact, and others were partial sequences.

Figure 6
www.frontiersin.org

Figure 6. Homologous sequences between organelle genomes are shown as blue (mt) and green (cp) arcs. The inner arcs represent MTPT fragments, while outer gray and black labels indicate partial and intact genes, respectively.

2.7 RNA editing events

To identify RNA editing events of Z. caespitosa, we mapped RNA-seq data to the mt and cp genomes. A total of 167 and 172 RNA editing sites were identified in the mt and cp PCGs, respectively (Figure 7A and Supplementary Tables S9, S10). Four mt (nad7, cox1, nad4, and cob) and cp genes (matK, rpoC2, accD, and ndhD) had many editing sites. In addition, the fewest edits were observed in most cp genes, with only one RNA editing site. Furthermore, cp PCGs had 11 different types of RNA editing sites, including 33% of C–U, 21% of U–C, 13% of G–A, 12% of A–G, 7% of C–A, 4% of A–C, 3% of G–C, 3% of G–U, 2% of A–U, 2% of U–A, and 1% of U–G (Figure 7B). Most mt PCGs had this C-to-U editing type. Over 62% (106) cp and 34% (57) mt editing sites exhibited near 100% editing efficiency, indicating a more efficient editing process than initially expected (Figure 7C). Most importantly, 92.21% and 55.81% in mt and cp occurred above the first two bases of the codon, changing the corresponding amino acids (Supplementary Tables S9, S10). We further examined the number of the amino acid changes and found that 95.2% (159) in mt and 57.6% (99) in cp caused amino acid alterations. The most frequent amino acid changes in the Z. caespitosa organelle genome were from Pro to Leu and Ser to Leu (Figures 7D, E).

Figure 7
www.frontiersin.org

Figure 7. The features of RNA editing sites identified in PCGs of the organelle genome include (A) the number of RNA editing sites in each PCG; (B) types of RNA editing; (C) efficiency of RNA editing; and (D, E) the number of amino acid changes in mt and cp, respectively.

3 Discussion

3.1 The structure and size of the organelle genomes in Z. caespitosa

Studying the genomes of plant organelles enhances our comprehension of their functions, inheritance patterns, and replication mechanisms, while also offering valuable insights into their evolution and adaptation (Wang et al., 2024). This study successfully de novo assembled the complete organelle genomes of Z. caespitosa. Consistent with the common structure of organelle genomes in seagrass (Petersen et al., 2017; Chen et al., 2021), both genomes are circular, sized at 192,246 bp and 143,972 bp, respectively. However, a recent study found varying degrees of genome size, completeness, and fragmentation in the seagrass mt genomes, which are attributed to recombination (Ma et al., 2024). Compared to the cp genome, the larger mt genome is due to a high abundance of repetitive sequences, particularly dispersed repeats (the mt genome contains 5,028 dispersed repeats, and cp has 37) (Supplementary Tables S4, S5). Similar disparities in repeat content have been observed in Dystaenia takeshimana (Park and Park, 2023). The length of the Z. caespitosa mt genome is well among the ranges reported for seagrasses (Chen et al., 2022). Notably, seagrasses’ mt genomes were smaller compared to those of other monocotyledons (Supplementary Table S11). Several factors, including horizontal gene transfer (HGT), the number of repeated sequences, and the gain or loss of large intrageneric fragments, influence variations in plant mt genome sizes (Wu et al., 2022).

Furthermore, the cp genome of Z. caespitosa contains a distinctive quadripartite organization, with two IRs, an LSC, and an SSC, similar to the structural characteristics of cp genomes of other angiosperms (Daniell et al., 2016). This genome is the second smallest within the Zosteraceae family, with cp genomes ranging from 143,877 bp in Z. marina to 178,261 bp in Thalassia hemprichii (Chen et al., 2023). This difference may be caused by very short ycf1 genes in the IR regions in Z. caespitosa (Olsen et al., 2016). The lengths of the cp genomes in the same family were similar (Supplementary Table S11), suggesting that the cp genome size variability is linked to phylogenetic relationships and evolutionary history (Zheng et al., 2017). Additionally, the expansion and contraction of the SC border and IR regions are major factors contributing to size variations in cp genomes, influencing their evolutionary rate (Dodsworth et al., 2015).

3.2 Codon usage pattern analysis in Z. caespitosa

Codon usage bias in genes, a notable evolutionary feature, occurs in various prokaryotic and eukaryotic organisms (Wang et al., 2023). These biases result from a complex interaction of natural selection, genetic drift, and mutational forces across long evolutionary periods (Chen et al., 2004; Zhang et al., 2022). According to angiosperm organelle genomes, this research showed that most codons ending with A/U have >1 RSCU values, possibly because of composition bias toward a high A/T ratio (Duan et al., 2021). This pattern is likely due to the structural stability of polyA and polyT, compared to polyC and polyG (Gragg et al., 2002).

The ENC-GC3 analysis suggests that mutations predominantly affect genes closely aligned with or overlapping the standard curve in the ENC-GC3 plot, probably altering their codon bias. Conversely, genes deviating from this curve are likely influenced by natural selection (Li et al., 2023; Wright, 1990). This study demonstrates that the majority of photosynthesis-associated genes in the cp genome and respiration-related genes in the mt genome of Z. caespitosa depart from the standard curve, indicating a greater impact of natural selection. This deviation pattern is consistent across species within the tea plant family (Li et al., 2023). Additionally, this study reveals that mt genome genes like atp9, with an ENC value of 33.25, and cp genome genes including psbI (33.96), petN (28.65), psaI (34.04), psaJ (33.05), rpl36 (23.23), and rpl32 (28.76), all with ENC values below 35, exhibit mutation-influenced codon usage biases. In contrast, genes with ENC values above 35 are predominantly shaped by natural or artificial selection (Wright, 1990).

3.3 Mutation rate analysis of Zostera species

The Ka/Ks is employed to evaluate the evolutionary rate of nucleotides and selective pressure, serving as an important indicator in species evolution (Li et al., 1985). In this study, the Ka/Ks of the cp genome is roughly 1/2 of that of the mt genome. The result was consistent with a previous study on the red alga Porphyra and phytoplankton Phaeocystis (Haptophyta), which have secondary cp; their mt mutation rates were higher than their cp rates (Smith et al., 2012, 2014). However, this trend contrasts most land plants and green algae, where the Ka/Ks of cp were three times higher than that of mt (Clegg et al., 1994; Smith, 2015; Smith and Keeling, 2015). The mutation rate variation across organisms is unclear but likely reflects the differences in the efficiency of DNA replication machinery between organelles and across taxonomic lineages (Drouin et al., 2008; Zhu et al., 2014; Smith, 2015; Gualberto and Newton, 2017). Additionally, the endosymbiotic history, specifically the number of endosymbiosis events, may be important, but their role in influencing these rates is poorly understood (Smith and Keeling, 2015).

3.4 Phylogenetic relationships in seagrass

Phylogenetic relationships within the family Zosteraceae, particularly regarding genus Zostera species, have been subjects of intense debate (Tanaka et al., 2003; Coyer et al., 2013). These studies used core barcodes to distinguish Zostera species within the Zosteraceae, which have potential limitations. In response, researchers have increasingly turned to cp genomes as a super barcode, which has proven successful in identifying numerous species and individuals across diverse studies (Kane et al., 2012; Chen et al., 2019). This study constructed phylogenetic trees based on common PCGs from 13 seagrass cp genomes. Results showed monophyly of subgenus Zostera and subgenus Zosterella. Interestingly, one good phylogenomic bootstrap branch indicated that Z. caespitosa typically emerged as a sister to Z. marina, demonstrating that Z. caespitosa was more closely related to Z. marina than other Zosteraceae species. Additionally, the results support the infrageneric classifications proposed by Chen et al., which divide seagrasses into four major clades: Zosteraceae, Ruppiaceae, Cymodoceaceae, and Hydrocharitaceae, with Cymodoceaceae and Ruppiaceae showing a relatively close relationship (Les and Tippery, 2013; Chen et al., 2022).

3.5 Horizontal gene transfers between organellar genomes in Z. caespitosa

Horizontal gene transfer between mt and cp genomes has been a significant phenomenon in the long-term evolution of angiosperms (Sloan and Wu, 2014; Gui et al., 2016). This study identified 50 MTPTs between the cp and mt genomes of Z. caespitosa, with a total length of 44,662 bp, accounting for 23.23% of the mt genome. Generally, the ratio of MTPTs in mt genomes varies from 0.56% in Marchantia polymorpha to 10.85% in Phoenix dactylifera (Zhao et al., 2018). The percentage in the mt genome of Z. caespitosa far exceeds the highest previously reported rates of approximately 10%–12% in Boea and Cucurbita (Zhang et al., 2012; Alverson et al., 2010). In the Z. caespitosa and Z. marina mt genomes, the frequent occurrence of MTPTs is linked to the recent integration of several large sequence transfers (Petersen et al., 2017). Among these MTPTs, MTPT1 is the longest. Previous studies have also documented the presence of large fragments of transferred cpDNA in mt genomes (Zhu et al., 2023). These large fragments likely play a significant role in angiosperms evolution by enhancing genetic diversity. Additionally, the origin of tRNAs in angiosperms mt is twofold: some are derived from the mt ancestor, while others are obtained from the cp via HGT (Sprinzl and Vassilenko, 2005). In addition to gene transfer between mt and cp genomes, HGT most frequently occurs from organelles to the nuclear genome, resulting in the production of nuclear plastid DNA (NUPTs) and nuclear mitochondrial DNA (NUMTs). Recent studies on Thalassia testudinum have found large, uninterrupted nuclear mitochondrial DNA sequences (NUMTs), indicating a recent mitochondrial DNA transfer. In contrast, Z. marina shows fewer such sequences, suggesting less frequent mitochondrial DNA integration. These large, uninterrupted NUMTs in T. testudinum may result from genome instability caused by TE expansion (Ma et al., 2024).

3.6 RNA editing events

Recent studies have demonstrated that RNA editing is prevalent across plant organelle genes and is crucial in energy metabolism and regulating genetics (Li et al., 2019) In this study, the RNA-seq data identified 167 RNA editing sites across mt-PCGs and 172 RNA across cp-PCGs. The number of editing sites is higher than those reported in most angiosperm cp genomes (20–60) (Knoop, 2023). Mt genes such as nad7, cox1, nad4, cob, and cp genes, including matK, rpoC2, accD, and ndhD had notably high editing frequencies. This study further investigated the impact of editing on protein structure, confirming that no new transmembrane domains were created (Supplementary Figures S1, S2). Additionally, editing nine differential sites (accD-157, accD-624, accD-1220, rpoC2-2255, rpoC2-2752, rpoC2-2822, mat-748, matK-178, and matK-799) changed the secondary structure composition surrounding the editing sites in the cp (Supplementary Figure S3). Editing 12 different sites (nad7-927, cox1-254, cox1-352, cox1-868, cox1-1037, cox1-1279, nad4-29, nad4-197, nad4-362, nad4-599, cob-400, and cob-794) altered the secondary structure composition surrounding the editing sites in the mt. This editing mechanism effectively repairs potential defects by disrupting and folding alpha-helices, enhancing structural protein stability (Yuar and Go, 2008).

Similarly, previous studies revealed C-to-U as the predominant type of editing in mt-PCGs (98.8%) and pt-PCGs (40%) of Z. caespitosa (Gray and Covello, 1993). Additionally, Z. caespitosa had non-canonical edits such as U to C, G to U, C to A, A to G, and G to A, typically found only in ancestral land plants (Chateigner-Boutin and Small, 2010; Uthaipaisanwong et al., 2012) Z. caespitosa also had unexpectedly rare editing events like A to C, A to U, U to A, G to C, and U to G, not previously reported in other species. These RNA editing events potentially alter protein structure or interaction by nonsynonymously replacing conserved amino acids, most notably changing Pro to Leu and Ser to Leu. These substitution shifts the physicochemical properties from hydrophilic to hydrophobic, increasing the hydrophobicity of interface residues, crucial for protein–protein interactions and enzyme efficiency (Jobson and Qiu, 2008; Giegé and Brennicke, 1999).

4 Conclusions

This study reported the complete organelle genomes of Z. caespitosa, enabling their comprehensive comparison. The PCGs in the organelle genome exhibits a strong A/U bias at codon endings, a selection-driven codon bias. The mt genome includes abundant cpDNA transferred fragments, much higher than in most angiosperms. The Ka/Ks of the Zostera mt genome are twofold higher than those in the cp genome. Remarkably, numerous RNA editing events were identified in cp-PCGs, including five rare types of RNA editing.

5 Materials and methods

5.1 Sample collection and organelle genome sequencing

Fresh Z. caespitosa leaves were collected from Yantai, China (121.4534 E, 37.5193 N). Prof. Quansheng Zhang verified the voucher specimens stored in the Ocean School of Yantai University. Leaves’ DNA was extracted using the DNA extraction kit (Invitrogen, CA, USA). DNA library was constructed using the Illumina TruSeq Library Preparation Kit (Illumina, CA, USA), in accordance with the manufacturer’s guidelines. These libraries were sequenced using the Illumina NovaSeq 6000 (Biozeron, Shanghai, China), producing raw data consisting of 150-bp paired-end reads. Constructing the long fragment library, its quality was assessed with Qubit, and then it was sequenced on the ONT platform (Biozeron, Shanghai, China).

5.2 Assembly of Z. caespitosa mt genome

The Illumina sequencing raw data were filtered with Trimmomatic v0.39 before the mt genome assembly (Bolger et al., 2014). All ONT-generated long reads were aligned to the Z. marina (NC_035345.1) mt genome, with Minimap2 v2.10-r761 using its default settings, processed into a Pairwise Mapping Format (PAF) file (Li, 2018). Homologous reads with a mapping quality >20 were treated as potential mt sequences. Subsequently, de novo assembly was performed on the homologous ONT long reads and the clean paired-end reads using Canu v2.2 and GetOrganelle v1.7.5 with default settings (Koren et al., 2017; Jin et al., 2020). BLASTN was used to align the draft contigs from Canu and GetOrganelle with the Z. marina mt CDSs to detect the candidate mt genome contigs. Using overlapping markers as a guide, the selected contigs were manually joined to generate the complete mt genome sequences. The circularity of the assembly was confirmed using the “check_circularity.pl” script, part of the sprai package (http://zombie.cb.k.u-tokyo.ac.jp/sprai/).

The alignment of long ONT reads and short Illumina to the mt genomes was performed using BWA, removing multi-mapped reads, unmapped reads, and PCR duplicates. Coverage information for the mt genomes was obtained by sorting the Binary Alignment/Map (BAM) files, and the accuracy was manually checked using the Integrative Genomics Viewer (IGV) with the BAM files as references.

5.3 Assembly of Z. caespitosa cp genome

Quality control was performed on the raw sequencing reads using FastQC, followed by trimming of redundant reads or low-quality reads (Q < 20) using Trimmomatic v0.39 (Bolger et al., 2014). The trimmed reads were assembled using GetOrganelle (v1.7.5) with default parameters (Jin et al., 2020). The Z. marina (NC_036014.1) cp genome sequence served as the reference for re-assembly, generating multiple potential cp genome assemblies. The final cp genome assembly’s accuracy, especially regarding the IR order and continuity, was checked and manually corrected as needed, using BLASTN (E-value cutoff of 10{sp}−5{/sp}) against reference cp genomes. The “check_circularity.pl” script from the sprai package was used to evaluate the circularity of the cp assembly. The boundaries of the SSC, LSC, IR, and regions were determined using BLASTN self-alignment.

5.4 Mt and cp genome of Z. caespitosa annotation

The mt and cp genes were annotated using GeSeq (Tillich et al., 2017) and BLASTN, with Z. marina (NC_035345.1) and Z. marina (NC_036014.1) as references, respectively. Exon/intron and start/stop codons boundaries in PCGs were manually corrected with Snap Gene Viewer, using reference gene models for guidance. The Organellar Genome DRAW (OGDRAW) software was used to visualize the mt and cp genomes graphically (Greiner et al., 2019). These mt and cp genome sequences are available in GenBank with the accession numbers PP566026 and PP566025, respectively.

5.5 Repeat sequence analysis

Here, SSRs were identified using the MISA tool (Beier et al., 2017), with the minimum repeat numbers set to 10, 5, 4, 3, 3, and 3 for mono-, di-, tri-, tetra-, penta-, and hexa-nucleotides, respectively. Dispersed repetitive, including forward (F), reverse (R), complementary (C), and palindromic (P) repeats, were identified using the REPuter software (Kurtz et al., 2001). The analysis was conducted with a 30-bp minimum repeat length, a 90-bp maximum repeat size, and a Hamming distance threshold of three. The detection of tandem repeats was performed using Tandem Repeats Finder (Benson, 1999).

5.6 Codon usage bias analysis

Phylosuite (v1.1.16) was used to extract the PCGs from mt and cp genomes. The codon preference of PCGs in these genomes was assessed using MEGA (v7.0), and RSCU values were calculated (Kumar et al., 2016). CodonW (v1.4.4) was employed to calculate the ENC-GC3 values, and the data were visualized with ggplot2 in R package.

5.7 Selective pressure estimation and phylogenetic analysis

Using MAFFT version 7.427 (Katoh and Standley, 2013), the core PCG sequences were aligned, and the Ka/Ks ratios for Z. caespitosa, Z. marina, and Z. japonica were calculated via the MLWL-based Ka/Ks Calculator (Wang et al., 2010). Python and R software were used to visualize the plots. The core CDSs of 13 seagrass cp genomes were aligned using MAFFT, and the alignments were used to construct Bayesian trees with MrBayes 3.2. The most suitable nucleotide substitution model was chosen through ModelFinder (Nguyen et al., 2015). Further, ML analyses were performed with IQ-TREE version 1.5.5 (Nguyen et al., 2015) utilizing ultrafast bootstrap (1,000 replicates) and a partition model. Finally, the trees were visualized using normal tree tools.

5.8 Gene transfer between mt and cp genomes

The mt and cp genome sequences for Z. caespitosa were aligned using the BLASTn tool, with the default parameters. Subsequently, the distribution of MTPTs was displayed using TBtools, which integrates the Circos package (Chen et al., 2020). Additionally, GeSeq was used to annotate the MTPTs, and duplicates, such as the two MTPTs in the cp genome’s IRs, were excluded.

5.9 Identification of RNA editing events

RNA editing events in the PCGs of the Z. caespitosa organelle genomes were identified using RNA-seq and genomic sequencing data. Quality control of the raw RNA reads was conducted via FastQC using default parameters. The RNA-seq reads were then mapped to the cp and mt genomes of Z. caespitosa using Tophat (Trapnell et al., 2009). Picard (https://github.com/broadinstitute/picard) was employed to remove duplicate records from the resulting BAM files. Next, single-nucleotide polymorphisms (SNPs) were identified using GATK (McKenna et al., 2010) and Samtools (Li et al., 2009) for genotyping analysis. The IGV was used to examine the mapped reads. The efficiency of RNA editing was quantified as the ratio of edited reads to total mapped reads (Li et al., 2009).

Data availability statement

The raw data were deposited at the NCBI SRA database https://www.ncbi.nlm.nih.gov/sra; with the accession number: SRR31595597–SRR31595599.

Author contributions

YY: Writing – review & editing, Formal analysis, Methodology, Software, Writing – original draft. YW: Writing – original draft, Validation. DW: Validation, Writing – original draft. XY: Conceptualization, Funding acquisition, Writing – review & editing. QZ: Conceptualization, Funding acquisition, Writing – review & editing, Supervision.

Funding

The author(s) declare financial support was received for the research, authorship, and/or publication of this article. This work was supported by the National Natural Science Foundation of China (Grant No. 42476112) and the Shandong Provincial Bureau of Geology and Mineral Resources project (Grant No. HJ202510).

Conflict of interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Generative AI statement

The author(s) declare that no Generative AI was used in the creation of this manuscript.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

Supplementary material

The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/fpls.2025.1550467/full#supplementary-material

References

Alverson, A. J., Wei, X., Rice, D. W., Stern, D. B., Barry, K., Palmer, J. D. (2010). Insights into the evolution of mitochondrial genome size from complete sequences of Citrullus lanatus and Cucurbita pepo (Cucurbitaceae). Mol. Biol. Evol. 27, 1436–1448. doi: 10.1093/molbev/msq029

PubMed Abstract | Crossref Full Text | Google Scholar

Beier, S., Thiel, T., Münch, T., Scholz, U., Mascher, M. (2017). MISA-web: A web server for microsatellite prediction. Bioinformatics 33, 2583–2585. doi: 10.1093/bioinformatics/btx198

PubMed Abstract | Crossref Full Text | Google Scholar

Benson, G. (1999). Tandem repeats finder: A program to analyze DNA sequences. Nucleic Acids Res. 27, 573–580. doi: 10.1093/nar/27.2.573

PubMed Abstract | Crossref Full Text | Google Scholar

Birky, C. W., Jr. (1995). Uniparental inheritance of mitochondrial and chloroplast genes: Mechanisms and evolution. Proc. Natl. Acad. Sci. U. S. A. 92, 11331–11338. doi: 10.1073/pnas.92.25.11331

PubMed Abstract | Crossref Full Text | Google Scholar

Birky, C. W., Jr. (2001). The inheritance of genes in mitochondria and chloroplasts: Laws, mechanisms, and models. Annu. Rev. Genet. 35, 125–148. doi: 10.1146/annurev.genet.35.102401.090231

PubMed Abstract | Crossref Full Text | Google Scholar

Bolger, A. M., Lohse, M., Usadel, B. (2014). Trimmomatic: A flexible trimmer for Illumina sequence data. Bioinformatics 30, 2114–2120. doi: 10.1093/bioinformatics/btu170

PubMed Abstract | Crossref Full Text | Google Scholar

Chateigner-Boutin, A. L., Small, I. (2010). Plant RNA editing. RNA Biol. 7, 213–219. doi: 10.4161/rna.7.2.11343

PubMed Abstract | Crossref Full Text | Google Scholar

Chen, C., Chen, H., Zhang, Y., Thomas, H. R., Frank, M. H., He, Y., et al. (2020). TBtools: An integrative toolkit developed for interactive analyses of big biological data. Mol. Plant 13, 1194–1202. doi: 10.1016/j.molp.2020.06.009

PubMed Abstract | Crossref Full Text | Google Scholar

Chen, J., Zang, Y., Liang, S., Xue, S., Shang, S., Zhu, M., et al. (2022). Comparative analysis of mitochondrial genomes reveals marine adaptation in seagrasses. BMC Genomics 23, 800. doi: 10.1186/s12864-022-09046-x

PubMed Abstract | Crossref Full Text | Google Scholar

Chen, J., Zang, Y., Shang, S., Liang, S., Zhu, M., Wang, Y., et al. (2021). Comparative chloroplast genomes of Zosteraceae species provide adaptive evolution insights into seagrass. Front. Plant Sci. 12. doi: 10.3389/fpls.2021.741152

PubMed Abstract | Crossref Full Text | Google Scholar

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

PubMed Abstract | Crossref Full Text | Google Scholar

Chen, S. L., Lee, W., Hottes, A. K., Shapiro, L., McAdams, H. H. (2004). Codon usage between genomes is constrained by genome-wide mutational processes. Proc. Natl. Acad. Sci. U. S. A. 101, 3480–3485. doi: 10.1073/pnas.0307827100

PubMed Abstract | Crossref Full Text | Google Scholar

Chen, X., Cui, Y., Nie, L., Hu, H., Xu, Z., Sun, W., et al. (2019). Identification and phylogenetic analysis of the complete chloroplast genomes of three Ephedra herbs containing ephedrine. Biomed. Res. Int. 2019, 5921725. doi: 10.1155/2019/5921725

PubMed Abstract | Crossref Full Text | Google Scholar

Clegg, M. T., Gaut, B. S., Learn, G. H., Morton, B. R. (1994). Rates and patterns of chloroplast DNA evolution. Proc. Natl. Acad. Sci. U. S. A. 91, 6795–6801. doi: 10.1073/pnas.91.15.6795

PubMed Abstract | Crossref Full Text | Google Scholar

Coyer, J. A., Hoarau, G., Kuo, J., Tronholm, A., Veldsink, J., Olsen, J. L. (2013). Phylogeny and temporal divergence of the seagrass family Zosteraceae using one nuclear and three chloroplast loci. Syst. Biodivers. 11, 271–284. doi: 10.1080/14772000.2013.821187

Crossref Full Text | Google Scholar

Daniell, H., Lin, C. S., Yu, M., Chang, W. J. (2016). Chloroplast genomes: Diversity, evolution, and applications in genetic engineering. Genome Biol. 17, 134. doi: 10.1186/s13059-016-1004-2

PubMed Abstract | Crossref Full Text | Google Scholar

Dodsworth, S., Leitch, A. R., Leitch, I. J. (2015). Genome size diversity in angiosperms and its influence on gene space. Curr. Opin. Genet. Dev. 35, 73–78. doi: 10.1016/j.gde.2015.10.006

PubMed Abstract | Crossref Full Text | Google Scholar

Drouin, G., Daoud, H., Xia, J. (2008). Relative rates of synonymous substitutions in the mitochondrial, chloroplast, and nuclear genomes of seed plants. Mol. Phylogenet. Evol. 49, 827–831. doi: 10.1016/j.ympev.2008.09.009

PubMed Abstract | Crossref Full Text | Google Scholar

Duan, H., Zhang, Q., Wang, C., Li, F., Tian, F., Lu, Y., et al. (2021). Analysis of codon usage patterns of the chloroplast genome in Delphinium grandiflorum L. reveals a preference for AT-ending codons as a result of major selection constraints. PeerJ 9, e10787. doi: 10.7717/peerj.10787

PubMed Abstract | Crossref Full Text | Google Scholar

Fourqurean, J. W., Duarte, C. M., Kennedy, H., Marbà, N., Holmer, M., Mateo, M. A., et al. (2012). Seagrass ecosystems as a globally significant carbon stock. Nat. Geosci. 5, 505–509. doi: 10.1038/ngeo1477

Crossref Full Text | Google Scholar

Giegé, P., Brennicke, A. (1999). RNA editing in Arabidopsis mitochondria effects 441 C to U changes in ORFs. Proc. Natl. Acad. Sci. U. S. A. 96, 15324–15329. doi: 10.1073/pnas.96.26.15324

PubMed Abstract | Crossref Full Text | Google Scholar

Gragg, H., Harfe, B., Jinks-Robertson, S. (2002). Base composition of mononucleotide runs affects DNA polymerase slippage and removal of frameshift intermediates by mismatch repair in Saccharomyces cerevisiae. Mol. Cell Biol. 22, 8756–8762. doi: 10.1128/MCB.22.24.8756-8762.2002

PubMed Abstract | Crossref Full Text | Google Scholar

Gray, M. W., Covello, P. S. (1993). RNA editing in plant mitochondria and chloroplasts. FASEB J. 7, 64–71. doi: 10.1096/fasebj.7.1.8422976

PubMed Abstract | Crossref Full Text | Google Scholar

Greiner, S., Lehwark, P., Bock, R. (2019). OrganellarGenomeDRAW (OGDRAW) version 1.3.1: Expanded toolkit for the graphical visualization of organellar genomes. Nucleic Acids Res. 47, W59–W64. doi: 10.1093/nar/gkz238

PubMed Abstract | Crossref Full Text | Google Scholar

Gualberto, J. M., Mileshina, D., Wallet, C., Niazi, A. K., Weber-Lotfi, F., Dietrich, A. (2014). The plant mitochondrial genome: Dynamics and maintenance. Biochimie 100, 107–120. doi: 10.1016/j.biochi.2013.09.016

PubMed Abstract | Crossref Full Text | Google Scholar

Gualberto, J. M., Newton, K. J. (2017). Plant mitochondrial genomes: Dynamics and mechanisms of mutation. Annu. Rev. Plant Biol. 68, 225–252. doi: 10.1146/annurev-arplant-043015-112232

PubMed Abstract | Crossref Full Text | Google Scholar

Gui, S., Wu, Z., Zhang, H., Zheng, Y., Zhu, Z., Liang, D., et al. (2016). The mitochondrial genome map of Nelumbo nucifera reveals ancient evolutionary features. Sci. Rep. 6, 30158. doi: 10.1038/srep30158

PubMed Abstract | Crossref Full Text | Google Scholar

Henry, R. J. (2005). Plant diversity and evolution: Genotypic and phenotypic variation in higher plants (Wallingford, UK: CABI Publishing), 332.

Google Scholar

Huang, J., Yu, Y., Liu, Y.-M., Xie, D.-F., He, X.-J., Zhou, S.-D. (2020). Comparative chloroplast genomics of Fritillaria (Liliaceae), inferences for phylogenetic relationships between Fritillaria and Lilium and plastome evolution. Plants 9, 133. doi: 10.3390/plants9020133

PubMed Abstract | Crossref Full Text | Google Scholar

Im, D., Kim, S. H., Kim, H., Suonan, Z., Zhang, F., Song, H.-J., et al. (2024). Tuft dynamics and the reproductive phenology of Zostera caespitosa on the southern coast of Korea. J. Mar. Sci. Eng. 12, 1738. doi: 10.3390/jmse12101738

Crossref Full Text | Google Scholar

Jansen, R. K., Ruhlman, T. A. (2012). “Plastid genomes of seed plants,” in Genomics of Chloroplasts and Mitochondria. Eds. Bock, R., Knoop, V. (Springer, Dordrecht, The Netherlands), 103–126.

Google Scholar

Jiang, R., Gao, X., Liu, J., Liu, Z., Qiu, G., Shao, L., et al. (2024). Effect of temperature on photosynthetic physiology and the kinetics of nutrient absorption in Zostera caespitosa. Environ. Exp. Bot. 219, 105617. doi: 10.1016/j.envexpbot.2023.105617

Crossref Full Text | Google Scholar

Jin, J. J., Yu, W. B., Yang, J. B., Song, Y., dePamphilis, C. W., Yi, T. S., et al. (2020). GetOrganelle: A fast and versatile toolkit for accurate de novo assembly of organelle genomes. Genome Biol. 21, 241. doi: 10.1186/s13059-020-02154-5

PubMed Abstract | Crossref Full Text | Google Scholar

Jobson, R. W., Qiu, Y. L. (2008). Did RNA editing in plant organellar genomes originate under natural selection or through genetic drift? Biol. Direct. 3, 43. doi: 10.1186/1745-6150-3-43

PubMed Abstract | Crossref Full Text | Google Scholar

Kane, N., Sveinsson, S., Dempewolf, H., Yang, J. Y., Zhang, D., Engels, J. M. M., et al. (2012). Ultra-barcoding in cacao (Theobroma spp.; Malvaceae) using whole chloroplast genomes and nuclear ribosomal DNA. Am. J. Bot. 99, 320–329. doi: 10.3732/ajb.1100570

PubMed Abstract | Crossref Full Text | Google Scholar

Katoh, K., Standley, D. M. (2013). MAFFT multiple sequence alignment software version 7: Improvements in performance and usability. Mol. Biol. Evol. 30, 772–780. doi: 10.1093/molbev/mst010

PubMed Abstract | Crossref Full Text | Google Scholar

Kaur, B. (2020). Gene fragmentation and RNA editing without borders: Eccentric mitochondrial genomes of diplonemids. Nucleic Acids Res. 48, 2694–2708. doi: 10.1093/nar/gkz1215

PubMed Abstract | Crossref Full Text | Google Scholar

Knoop, V. (2023). C-to-U and U-to-C: RNA editing in plant organelles and beyond. J. Exp. Bot. 74, 2273–2294. doi: 10.1093/jxb/erac488

PubMed Abstract | Crossref Full Text | Google Scholar

Koren, S., Walenz, B. P., Berlin, K., Miller, J. R., Bergman, N. H., Phillippy, A. M. (2017). Canu: Scalable and accurate long-read assembly via adaptive k-mer weighting and repeat separation. Genome Res. 27, 722–736. doi: 10.1101/gr.215087.116

PubMed Abstract | Crossref Full Text | Google Scholar

Kumar, S., Stecher, G., Tamura, K. (2016). MEGA7: Molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol. Biol. Evol. 33, 1870–1874. doi: 10.1093/molbev/msw054

PubMed Abstract | Crossref Full Text | Google Scholar

Kurtz, S., Choudhuri, J. V., Ohlebusch, E., Schleiermacher, C., Stoye, J., Giegerich, R. (2001). REPuter: The manifold applications of repeat analysis on a genomic scale. Nucleic Acids Res. 29, 4633–4642. doi: 10.1093/nar/29.22.4633

PubMed Abstract | Crossref Full Text | Google Scholar

Lamb, J. B., van de Water, J. A. J. M., Bourne, D. G., Altier, C., Hein, M. Y., Fiorenza, E. A., et al. (2017). Seagrass ecosystems reduce exposure to bacterial pathogens of humans, fishes, and invertebrates. Science 355, 731–733. doi: 10.1126/science.aal1956

PubMed Abstract | Crossref Full Text | Google Scholar

Les, D. H., Tippery, N. P. (2013). “In time and with water … The systematics of Alismatid monocotyledons,” in Early Events in Monocot Evolution. Systematics Association Special Volume Series. Eds. Wilkin, P., Mayo, S. J. (Cambridge University Press, Cambridge, UK), 118–164.

Google Scholar

Li, H. (2018). Minimap2: Pairwise alignment for nucleotide sequences. Bioinformatics 34, 3094–3100. doi: 10.1093/bioinformatics/bty191

PubMed Abstract | Crossref Full Text | Google Scholar

Li, H., Handsaker, B., Wysoker, A., Fennell, T., Ruan, J., Homer, N., et al. (2009). The sequence alignment/map format and SAMtools. Bioinformatics 25, 2078–2079. doi: 10.1093/bioinformatics/btp352

PubMed Abstract | Crossref Full Text | Google Scholar

Li, J., Tang, H., Luo, H., Tang, J., Zhong, N., Xiao, L. (2023). Complete mitochondrial genome assembly and comparison of Camellia sinensis var. Assamica cv. Duntsa. Front. Plant Sci. 14. doi: 10.3389/fpls.2023.1117002

PubMed Abstract | Crossref Full Text | Google Scholar

Li, W. H., Wu, C. I., Luo, C. C. (1985). A new method for estimating synonymous and nonsynonymous rates of nucleotide substitution considering the relative likelihood of nucleotide and codon changes. Mol. Biol. Evol. 2, 150–174. doi: 10.1093/oxfordjournals.molbev.a040343

PubMed Abstract | Crossref Full Text | Google Scholar

Li, M., Xia, L., Zhang, Y., Niu, G., Li, M., Wang, P., et al. (2019). Plant editosome database: A curated database of RNA editosome in plants. Nucleic Acids Res. 47, D170–D174. doi: 10.1093/nar/gky1026

PubMed Abstract | Crossref Full Text | Google Scholar

Ma, X., Vanneste, S., Chang, J., Ambrosino, L., Barry, K., Bayer, T., et al. (2024). Seagrass genomes reveal ancient polyploidy and adaptations to the marine environment. Nat. Plants. 10, 240–255. doi: 10.1038/s41477-023-01608-5

PubMed Abstract | Crossref Full Text | Google Scholar

Maréchal, A., Brisson, N. (2010). Recombination and the maintenance of plant organelle genome stability. New Phytol. 186, 299–317. doi: 10.1111/j.1469-8137.2010.03195.x

PubMed Abstract | Crossref Full Text | Google Scholar

McKenna, A., Hanna, M., Banks, E., Sivachenko, A., Cibulskis, K., Kernytsky, A., et al. (2010). The Genome Analysis Toolkit: A MapReduce framework for analyzing next-generation DNA sequencing data. Genome Res. 20, 1297–1303. doi: 10.1101/gr.107524.110

PubMed Abstract | Crossref Full Text | Google Scholar

Mower, J. P., Vickrey, T. L. (2018). Structural diversity among plastid genomes of land plants. Adv. Bot. Res. 85, 263–292. doi: 10.1016/bs.abr.2017.11.013

Crossref Full Text | Google Scholar

Nguyen, L., Schmidt, H. A., von Haeseler, A., Minh, B. Q. (2015). IQ-TREE: A fast and effective stochastic algorithm for estimating maximum-likelihood phylogenies. Mol. Biol. Evol. 32, 268–274. doi: 10.1093/molbev/msu300

PubMed Abstract | Crossref Full Text | Google Scholar

Nizam, A., Meera, S. P., Kumar, A. (2022). Genetic and molecular mechanisms underlying mangrove adaptations to intertidal environments. iScience 25, 103547. doi: 10.1016/j.isci.2021.103547

PubMed Abstract | Crossref Full Text | Google Scholar

Olsen, J. L., Rouzé, P., Verhelst, B., Lin, Y. C., Bayer, T., Collen, J., et al. (2016). The genome of the seagrass Zostera marina reveals angiosperm adaptation to the sea. Nature 530, 331–335. doi: 10.1038/nature16548

PubMed Abstract | Crossref Full Text | Google Scholar

Orth, R. J., Carruthers, T. J., Dennison, W. C., Duarte, C. M., Fourqurean, J. W., Heck, K. L., et al. (2006). A global crisis for seagrass ecosystems. Bioscience 56, 987–996. doi: 10.1641/0006-3568(2006)56[987:AGCFSE]2.0.CO;2

Crossref Full Text | Google Scholar

Park, S., Park, S. J. (2023). Intrageneric structural variation in organelle genomes from the genus Dystaenia (Apiaceae): Genome rearrangement and mitochondrion-to-plastid DNA transfer. Front. Plant Sci. 14. doi: 10.3389/fpls.2023.1283292

PubMed Abstract | Crossref Full Text | Google Scholar

Petersen, G., Cuenca, A., Zervas, A., Ross, G. T., Graham, S. W., Barrett, C. F., et al. (2017). Mitochondrial genome evolution in Alismatales: Size reduction and extensive loss of ribosomal protein genes. PloS One 12, e0177606. doi: 10.1371/journal.pone.0177606

PubMed Abstract | Crossref Full Text | Google Scholar

Short, F., Carruthers, T., Dennison, W., Waycott, M. (2007). Global seagrass distribution and diversity: A bioregional model. J. Exp. Mar. Biol. Ecol. 350, 3–20. doi: 10.1016/j.jembe.2007.06.012

Crossref Full Text | Google Scholar

Short, F. T., Polidoro, B., Livingstone, S. R., Carpenter, K. E., Bandeira, S., Bujang, J. S., et al. (2011). Extinction risk assessment of the world’s seagrass species. Biol. Conserv. 144, 1961–1971. doi: 10.1016/j.biocon.2011.04.010

Crossref Full Text | Google Scholar

Sloan, D. B., Alverson, A. J., Wu, M., Palmer, J. D., Taylor, D. R. (2012). Recent acceleration of plastid sequence and structural evolution coincides with extreme mitochondrial divergence in the angiosperm genus Silene. Genome Biol. Evol. 4, 294–306. doi: 10.1093/gbe/evs006

PubMed Abstract | Crossref Full Text | Google Scholar

Sloan, D. B., Wu, Z. (2014). History of plastid DNA insertions reveals weak deletion and AT mutation biases in angiosperm mitochondrial genomes. Gen. Biol. Evol. 6, 3210–3221. doi: 10.1093/gbe/evu253

PubMed Abstract | Crossref Full Text | Google Scholar

Smith, D. R. (2015). Mutation rates in plastid genomes: They are lower than you might think. Genome Biol. Evol. 7, 1227–1234. doi: 10.1093/gbe/evv069

PubMed Abstract | Crossref Full Text | Google Scholar

Smith, D. R., Arrigo, K. R., Alderkamp, A. C., Allen, A. E. (2014). Massive difference in synonymous substitution rates among mitochondrial, plastid, and nuclear genes of Phaeocystis algae. Mol. Phylogenet. Evol. 71, 36–40. doi: 10.1016/j.ympev.2013.10.018

PubMed Abstract | Crossref Full Text | Google Scholar

Smith, D. R., Hua, J., Lee, R. W., Keeling, P. J. (2012). Relative rates of evolution among the three genetic compartments of the red alga Porphyra differ from those of green plants and do not correlate with genome architecture. Mol. Phylogenet. Evol. 65, 339–344. doi: 10.1016/j.ympev.2012.06.017

PubMed Abstract | Crossref Full Text | Google Scholar

Smith, D. R., Keeling, P. J. (2015). Mitochondrial and plastid genome architecture: Reoccurring themes, but significant differences at the extremes. Proc. Natl. Acad. Sci. U. S. A. 112, 10177–10184. doi: 10.1073/pnas.1422049112

PubMed Abstract | Crossref Full Text | Google Scholar

Sprinzl, M., Vassilenko, K. S. (2005). Compilation of tRNA sequences and sequences of tRNA genes. Nucleic Acids Res. 33, D139–D140. doi: 10.1093/nar/gki012

PubMed Abstract | Crossref Full Text | Google Scholar

Takenaka, M., Zehrmann, A., Verbitskiy, D., Härtel, B., Brennicke, A. (2013). RNA editing in plants and its evolution. Annu. Rev. Genet. 47, 335–352. doi: 10.1146/annurev-genet-111212-133519

PubMed Abstract | Crossref Full Text | Google Scholar

Tanaka, N., Kuo, J., Omori, Y., Nakaoka, M., Aioi, K. (2003). Phylogenetic relationships in the genera Zostera and Heterozostera (Zosteraceae) based on matK sequence data. J. Plant Res. 116, 273–279. doi: 10.1007/s10265-003-0090-x

PubMed Abstract | Crossref Full Text | Google Scholar

Tillich, M., Lehwark, P., Pellizzer, T., Ulbricht-Jones, E. S., Fischer, A., Bock, R., et al. (2017). GeSeq: Versatile and accurate annotation of organelle genomes. Nucleic Acids Res. 45, W6–W11. doi: 10.1093/nar/gkx391

PubMed Abstract | Crossref Full Text | Google Scholar

Trapnell, C., Pachter, L., Salzberg, S. L. (2009). TopHat: Discovering splice junctions with RNA-Seq. Bioinformatics 25, 1105–1111. doi: 10.1093/bioinformatics/btp120

PubMed Abstract | Crossref Full Text | Google Scholar

Unsworth, R. K. F., Cullen-Unsworth, L. C., Jones, B. L., Lilley, R. J., James, R. K., Jones, R., et al. (2022). The planetary role of seagrass conservation. Science 377, 609–613. doi: 10.1126/science.abq6923

PubMed Abstract | Crossref Full Text | Google Scholar

Uthaipaisanwong, P., Chanprasert, J., Shearman, J. R., Sangsrakru, D., Yoocha, T., Jomchai, N., et al. (2012). Characterization of the chloroplast genome sequence of oil palm (Elaeis guineensis Jacq.). Gene 500, 172–180. doi: 10.1016/j.gene.2012.03.061

PubMed Abstract | Crossref Full Text | Google Scholar

Wang, Y., Jiang, D., Guo, K., Zhao, L., Meng, F., Xiao, J., et al. (2023). Comparative analysis of codon usage patterns in chloroplast genomes of ten Epimedium species. BMC Genom. Data. 24, 3. doi: 10.1186/s12863-023-01104-x

PubMed Abstract | Crossref Full Text | Google Scholar

Wang, J., Kan, S., Liao, X., Zhou, J., Tembrock, L. R., Daniell, H., et al. (2024). Plant organellar genomes: Much done, much more to do. Trends Plant Sci. 29, 754–769. doi: 10.1016/j.tplants.2023.12.014

PubMed Abstract | Crossref Full Text | Google Scholar

Wang, D., Zhang, Y., Zhang, Z., Zhu, J., Yu, J. (2010). KaKs_Calculator 2.0: A toolkit incorporating gamma-series methods and sliding window strategies. Genom. Proteom. Bioinform. 8, 77–80. doi: 10.1016/S1672-0229(10)60008-3

PubMed Abstract | Crossref Full Text | Google Scholar

Waycott, M., Duarte, C. M., Carruthers, T. J., Orth, R. J., Dennison, W. C., Olyarnik, S., et al. (2009). Accelerating loss of seagrasses across the globe threatens coastal ecosystems. Proc. Natl. Acad. Sci. U. S. A. 106, 12377–12381. doi: 10.1073/pnas.0905620106

PubMed Abstract | Crossref Full Text | Google Scholar

Wright, F. (1990). The “effective number of codons” used in a gene. Gene 87, 23–29. doi: 10.1016/0378-1119(90)90491-9

PubMed Abstract | Crossref Full Text | Google Scholar

Wu, Z. Q., Liao, X. Z., Zhang, X. N., Tembrock, L. R., Broz, A. (2022). Genomic architectural variation of plant mitochondria—A review of multichromosomal structuring. J. Syst. Evol. 60, 160–168. doi: 10.1111/jse.12655

Crossref Full Text | Google Scholar

Xu, S., Qiao, Y., Xu, S., Yue, S., Zhang, Y., Liu, M., et al. (2021). Diversity, distribution, and conservation of seagrass in coastal waters of the Liaodong Peninsula, North Yellow Sea, Northern China: Implications for seagrass conservation. Mar. pollut. Bull. 167, 112261. doi: 10.1016/j.marpolbul.2021.112261

PubMed Abstract | Crossref Full Text | Google Scholar

Yuar, K., Go, M. (2008). Correlation between amino acid residues converted by RNA editing and functional residues in protein three-dimensional structures in plant organelles. BMC Plant Biol. 8, 79. doi: 10.1186/1471-2229-8-79

PubMed Abstract | Crossref Full Text | Google Scholar

Zhang, T., Fang, Y., Wang, X., Deng, X., Zhang, X., Hu, S., et al. (2012). The complete chloroplast and mitochondrial genome sequences of Boea hygrometrica: Insights into the evolution of plant organellar genomes. PloS One 7, e30531. doi: 10.1371/journal.pone.0030531

PubMed Abstract | Crossref Full Text | Google Scholar

Zhang, Y., Shen, Z., Meng, X., Zhang, L., Liu, Z., Liu, M., et al. (2022). Codon usage patterns across seven Rosales species. BMC Plant Biol. 22, 65. doi: 10.1186/s12870-022-03450-x

PubMed Abstract | Crossref Full Text | Google Scholar

Zhao, N., Wang, Y., Hua, J. (2018). The roles of mitochondrion in intergenomic gene transfer in plants: A source and a pool. Int. J. Mol. Sci. 19, 547. doi: 10.3390/ijms19020547

PubMed Abstract | Crossref Full Text | Google Scholar

Zheng, X.-M., Wang, J.-R., Feng, L., Liu, S., Pang, H.-B., Qi, L., et al. (2017). Inferring the evolutionary mechanism of the chloroplast genome size by comparing whole-chloroplast genome sequences in seed plants. Sci. Rep. 7, 1555. doi: 10.1038/s41598-017-01518-5

PubMed Abstract | Crossref Full Text | Google Scholar

Zhu, A., Guo, W., Jain, K., Mower, J. P. (2014). Unprecedented heterogeneity in the synonymous substitution rate within a plant genome. Mol. Biol. Evol. 31, 1228–1236. doi: 10.1093/molbev/msu079

PubMed Abstract | Crossref Full Text | Google Scholar

Zhu, H., Shan, Y., Li, J., Zhang, X., Yu, J., Wang, H. (2023). Assembly and comparative analysis of the complete mitochondrial genome of Viburnum chinshanense. BMC Plant Biol. 23, 487. doi: 10.1186/s12870-023-04493-4

PubMed Abstract | Crossref Full Text | Google Scholar

Keywords: Zostera caespitosa, mitochondrial genome, chloroplast genome, sequence evolution, gene transfer, RNA editing

Citation: Yong Y, Wang Y, Wang D, Yuan X and Zhang Q (2025) The organelle genomes of the endangered seagrass Zostera caespitosa reveal sequence divergences, massive gene transfer, and uncommon RNA editing types. Front. Plant Sci. 16:1550467. doi: 10.3389/fpls.2025.1550467

Received: 23 December 2024; Accepted: 20 January 2025;
Published: 17 February 2025.

Edited by:

Michael L. Moody, The University of Texas at El Paso, United States

Reviewed by:

Jun Chen, Ocean University of China, China
Xiao Ma, Ghent University, Belgium

Copyright © 2025 Yong, Wang, Wang, Yuan and Zhang. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Quansheng Zhang, eXR1cXN6QGhvdG1haWwuY29t; Xingfang Yuan, MTM5NjU5MDI3NEBxcS5jb20=

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.

Research integrity at Frontiers

Man ultramarathon runner in the mountains he trains at sunset

94% of researchers rate our articles as excellent or good

Learn more about the work of our research integrity team to safeguard the quality of each article we publish.


Find out more