Your new experience awaits. Try the new design now and help us make it even better

ORIGINAL RESEARCH article

Front. Plant Sci., 28 October 2025

Sec. Plant Systematics and Evolution

Volume 16 - 2025 | https://doi.org/10.3389/fpls.2025.1653128

Organelle genomes reveal adaptive evolution and phylogenetic position of the endangered Primula mallophylla

Wenqiao LiWenqiao Li1Le WangLe Wang1Youwei ZuoYouwei Zuo1Huan ZhangHuan Zhang1Xiao ZhangXiao Zhang1Lingxiang YangLingxiang Yang2Jiabin ZhangJiabin Zhang1Fengyuan ZhengFengyuan Zheng1Hongping Deng*Hongping Deng1*
  • 1School of Life Sciences, Southwest University, Chongqing, China
  • 2Administration Center of Chongqing Dabashan National Nature Reserve, Chongqing, China

Introduction: Species of Primula section Proliferae are predominantly distributed in alpine environments above 2,000 meters and represent an important group for studying environmental adaptation and phylogenetic evolution. Primula mallophylla, holds important ecological, ornamental, and conservation value. However, genomic resources for this species remain unavailable, hindering further research on its genetic evolution and conservation strategies.

Methods: In this study, we sequenced and assembled the complete mitochondrial and chloroplast genomes of P. mallophylla for the first time. Comprehensive analyses were conducted on their structural characteristics, repetitive sequences, RNA editing site prediction, codon usage bias, intracellular gene transfer, phylogenetic inference, and selective pressure.

Results: The results show that both the mitochondrial and chloroplast genomes exhibit typical master circular structures, with sizes of 340,219 bp and 150,733 bp, respectively. The mitochondrial genome has more abundant repetitive sequences and has undergone genomic rearrangements. There are only 9 MTPTs between mitochondria and chloroplasts, totaling 2,028 bp. In the regulation of mitochondrial genome expression, we predicted 475 RNA editing sites, with ccmB and mttB showing the highest potential editing frequencies. We found that P. mallophylla exhibits similar codon usage bias to most plants, and based on ENC-GC3s analysis, some genes appear to be under natural selection. Chloroplast genes rpl2, rpl22, rbcL, and ndhB exhibit branch-specific positive selection in Primula, reflecting adaptive evolution of photosynthesis and protein synthesis functions in high-altitude environments. Based on phylogenetic trees constructed from PCGs, Primulaceae are most closely related to Ebenaceae. Furthermore, phylogenetic analysis based on chloroplast genomes and PCGs of Primula species showed that sect. Proliferae is not monophyletic, with P. stenodonta being a sister species to P. mallophylla.

Discussion: These findings provide crucial genomic resources and insights into the adaptive evolution of P. mallophylla, while also clarifying phylogenetic relationships within sect. Proliferae and Primulaceae, thereby offering valuable guidance for conservation strategies and further evolutionary studies.

1 Introduction

Primula is the largest genus in the Primulaceae, comprising over 500 species primarily distributed in alpine regions and temperate zones of the Northern Hemisphere. In China, more than 300 species are found, most of which are endemic, with the Eastern Himalaya-Hengduan Mountains as the hotspot area (Richards, 2003; Hu and Kelso, 1996; Hu, 1998). This genus is of great ornamental value in horticulture and floriculture due to its long flowering period, colorful inflorescences, and distinctive shape, and is regarded as one of the greatest genera of garden plants in the world (Richards, 2002; Yan et al., 2010). In addition, the species of Primula have a long tradition of folk medicinal use, exhibiting significant therapeutic effects in treating various conditions, including rheumatic pain, headaches, insomnia, nervous tension, and ocular diseases (Budzianowski et al., 2005; Başbülbül et al., 2008; Majid et al., 2014; Khan et al., 2022).

Species of Primula are cross-pollinated plants and serve as an ideal model for evolution (Darwin, 1877; Mast et al., 2006; Gilmartin, 2015). Its systematic position and the phylogenetic relationships within the genus have long been focal points. Schott (1851) divided Primula into two subgenera for the first time, but this classification was only adopted by a few scholars. Pax (1889) divided Primula into 21 sections, but the boundaries between sections were not clear. Subsequently, Smith and Fletcher (1947) systematically classified the Primula into 31 sections. To better express the genetic relationship between sections, Wendelbo (1961) elevated some sections classified by Smith and Fletcher to the level of subgenus, and classified its 31 sections into 7 subgenera. Halda (1992) divided Primula into 8 subgenera, and took Omphalogramma as a subgenus within Primula. Then, Richards (2003) revised the Primula into a six-subgenera system, and divided it into 38 sections. However, Hu (1990, 1994, 1998) believed that there were cross and overlap among the sections of Primula, and their differentiation had not yet reached the subgenus level. Therefore, they divided the genus into 30 sections, including 24 sections in China. The above research has yielded substantial insights into the phylogenetic development of Primula taxa. Nevertheless, the interspecific hybridization, parallel evolution, gene introgression and other complex phenomena occur frequently in the long-term evolution of these species (Xie et al., 2017; Stubbs et al., 2024), which makes it difficult to clarify the phylogenetic relationship of the Primula, and the results of molecular phylogeny within the genus are in great conflict with the traditional classification system (Mast et al., 2001, Mast et al., 2006; Trift et al., 2002; Schmidt-Lebuhn et al., 2012). Numerous studies based on phylogenetic analyses of chloroplast genomes have demonstrated that some sections (such as: sect. Proliferae, sect. Crystallophlomis, sect. Monocarpicae, sect. Obconicolisteri) were not a monophyletic group (Li, 2022; Xie et al., 2023). Thus, additional genomic data are required to clarify the evolutionary history of the Primula further.

The organelle genomes of plants, including the mitochondrial genome (mtDNA) and chloroplast genome (cpDNA), are indispensable for the investigation of plant evolution, physiological regulation, and phylogeny (Maréchal and Brisson, 2010). In contrast to the nuclear genome, organelle genes have a low mutation rate and are predominantly maternally inherited, making them widely used in species identification, phylogenetic analysis, and studies of adaptive evolution (Wang et al., 2024a). The chloroplast genome exhibits a conserved structure, typically comprising a quadripartite structure with a large single-copy region (LSC), a small single-copy region (SSC), and two inverted repeat regions (IRs), and contains numerous functional genes that are involved in photosynthesis and key protein synthesis (Liu et al., 2024a). In contrast, the mitochondrial genome exhibits greater variability in both structure and size, including linear, circular, more complex branching or reticular structures, and is mainly involved in the respiratory chain, protein synthesis, and energy metabolism-related functions (Xia et al., 2023; Liu et al., 2024b; Yong et al., 2025). However, the abundance of numerous repeat sequences, insertion fragments, and exogenous DNA integration (including fragments from nuclear and chloroplast genomes) among its genes, especially post-transcriptional RNA editing events, may regulate gene expression and thus affect adaptation to environmental stresses. Thus, the mitochondrial genome has a unique role to play in the evolution of plant adaptations and the differentiation of species (Li et al., 2024c, Li et al., 2025). Nonetheless, systematic studies of organelle genomes in the genus Primula remain limited, with fewer than five mitochondrial genomes and fewer than 100 chloroplast genomes reported, which has constrained a deeper understanding of their phylogenetic relationships, species boundaries, and genetic structure.

Primula mallophylla Balf. F., belongs to the family Primulaceae, genus Primula, and section Proliferae. This plant vanished following its discovery in the late 19th century, only to be rediscovered by researchers after a span of 90 years, and is designated as Critically Endangered (CR) in the China Biodiversity Red List (Balfour, 1916; Wu et al., 2009; https://www.mee.gov.cn/xxgk2018/xxgk/xxgk01/202305/t20230522_1030745.html). Importantly, the vast majority of the plants in this section are distributed above 2km in altitude (Hu, 1990), which is a key taxon for studying environmental adaptations and evolutionary mechanisms. Therefore, the present study is the first to analyze the complete mitochondrial and chloroplast genomes of P. mallophylla, to comprehensively analyze its gene composition, structural features, repeat sequences, RNA editing, codon preferences, and selection pressure, exploring its evolutionary status and affinity in the Primula by combining phylogenetic analysis and collinear comparison. This study not only helps to clarify the taxonomic controversy and reveal the complex evolutionary process, but also offers a significant theoretical foundation for the conservation genetics of endangered plants.

2 Materials and methods

2.1 Plant materials, DNA extraction, and sequencing

Fresh young leaves of P. mallophylla were harvested from Chengkou County, Chongqing, China. Total DNA was extracted via the CTAB method (Doyle and Doyle, 1987), and its concentration and integrity were assessed using 0.7% agarose gel electrophoresis, NanoDrop One spectrophotometer (NanoDrop Technologies, Wilmington, DE, USA), and Qubit 3.0 fluorometer (Life Technologies, Carlsbad, CA, USA). A SMRTbell library with an insert size of 15-20 kb was then generated and sequenced on the PacBio Revio platform. Meanwhile, a paired-end library with an average insert size of 400 bp was generated and sequenced on the Illumina NovaSeq platform. Following quality control, adapter trimming, and the elimination of low-quality reads, 10 Gb HiFi long-reads and 4 Gb of Illumina short-reads were acquired.

2.2 Genome assembly and annotation

De novo assembly of the P. mallophylla mitochondrial genome was performed using Flye (v2.9) software with default parameters on HiFi long-reads (Kolmogorov et al., 2019). All assembled contigs were indexed using the makeblastdb program from BLAST (Altschul, 1997). Subsequently, the mitochondrial genome sequence of Primula sikkimensis (NC_082937.1-08293743.1) was used as a query to screen for contigs containing mitochondrial genome sequences via BLASTn (v2.13.0) (Chen et al., 2015), with parameter ‘-evalue 1e-10 -outfmt 6 -max_hsps 10 -word_size 7 -task blastn-short’. All long-reads were mapped to aligned contigs using Minimap2 (Li, 2018), and aligned reads were extracted and reassembled using Flye software. The assembly results were visualized using Bandage (v0.8.1) (Wick et al., 2015). The Protein-coding genes (PCGs) were annotated utilizing PMGA (Li et al., 2024a), referencing the mitochondrial genomes of P. sikkimensis, Arabidopsis thaliana (NC_037304), and Liriodendron tulipifera (NC_021152.1). The tRNA genes were annotated using tRNAscan-SE (v2.0.11) (Lowe and Eddy, 1997), whereas rRNA genes were annotated using BLASTn. All annotations underwent manual curation using Apollo (v1.11.8) (Lewis et al., 2002). The chloroplast genome was assembled via GetOrganelle (Jin et al., 2020), with the parameters ‘-R 15 -k 21,45,65,85,105 -F embplant_pt’. The assembly results were annotated with CPGAVAS2 (Shi et al., 2019), utilizing Primula stenodonta (NC_034677.1) as a reference. Annotation errors were manually modified using CPGView (Liu et al., 2023). Finally, both the mitochondrial and chloroplast genome maps were visualized using OGDRAW (Stephan et al., 2019).

2.3 Repeat sequence analysis

Repeat sequences of mitochondrial and chloroplast genomes, including simple sequence repeats (SSRs), tandem repeats, and dispersed repeats, were identified using the MISA webserver (v2.1) (Beier et al., 2017), the Tandem Repeats Finder (v4.09) (Benson, 1999), and the REPuter (Stefan et al., 2001), respectively. The outcomes were depicted with Microsoft Excel 2021 and the Circos package inside TBtools (Zhang et al., 2013; Chen et al., 2020).

2.4 RNA editing site prediction and codon usage bias analysis

RNA editing sites for mitochondrial genome PCGs were computationally predicted using Deepred-mt (Edera et al., 2021). The PCGs were extracted from mitochondrial and chloroplast genomes using Phylosuite (v1.1.16) (Zhang et al., 2020), and then relative synonymous codon usage (RSCU) values were calculated using MEGA (v7.0) (Kumar et al., 2016). The effective number of codons (ENC) values and GC content at the third synonymous codon position (GC3s) were calculated using CodonW (v1.4.4) (Grote et al., 2005). Results were visualized using ggplot2 (Wickham, 2011) in the R-package.

2.5 Intracellular gene transfer analysis

To detect potential sequence transfers between the mitochondrial and chloroplast genomes of P. mallophylla, pairwise sequence comparisons were performed using the BLASTn with an E-value threshold of 1e-5. Homologous fragments were then visualized using the Circos package inside TBtools.

2.6 Phylogenetic inference and divergence time estimation

A total of 16 species were selected for phylogenetic analyses based on mitochondrial and chloroplast genomes, of which 15 species belong to Ericales, with Hydrangea chinensis (Cornales) designated as the outgroup. The shared PCGs across mitochondrial and chloroplast genomes (15 and 59 genes, respectively) were extracted using PhyloSuite. After multiple sequence alignment with MAFFT (Katoh and Standley, 2013), correction using MACSE (Ranwez et al., 2018), and trimming with Gblocks (Talavera and Castresana, 2007), concatenated sequence matrices were generated for phylogenetic tree reconstruction. To evaluate phylogenetic signals under different evolutionary models, maximum likelihood (ML) and Bayesian inference (BI) analyses were conducted at both the nucleotide level and codon-partitioned level. ML analyses were performed using IQ-TREE (v1.6.12) (Nguyen et al., 2015), with the nucleotide model set as ‘-m MFP -B 1000 -alrt 1000’ and the codon-partitioned model as ‘-m MFP+MERGE -B 1000 -alrt 1000’. BI analyses were conducted in MrBayes v3.2.6 (Ronquist et al., 2012), with each dataset run for 2,000,000 generations, sampling every 1,000 generations, and discarding the initial 25% of samples as burn-in. Convergence was assessed using Tracer v1.7.2.

Additionally, phylogenetic analyses were performed based on complete chloroplast genomes of the 16 species to enhance resolution. To further resolve relationships within Primula, ML and BI trees were constructed based on 86 Primula species and two Androsace outgroup species using complete chloroplast genomes and 65 shared PCGs. To assess the consistency of phylogenetic signals, gene concordance factor (gCF) analyses were conducted on phylogenies inferred from chloroplast and mitochondrial PCGs using IQ-TREE under default settings, quantifying the concordance between individual gene trees and the overall species tree topology.

All published mitochondrial and chloroplast genomes of the species used in this analysis were retrieved from the NCBI database (accession numbers are listed in Supplementary Tables S1, S2). The best-fit nucleotide substitution models for each dataset were automatically selected by ModelFinder (Kalyaanamoorthy et al., 2017) implemented in IQ-TREE, and the specific PCGs datasets used for phylogenetic analyses are provided in Supplementary Table S3. Phylogenetic trees were visualized using iTOL v6 (Letunic and Bork, 2024).

Divergence times were estimated using the ML topology of the 16 species based on shared chloroplast PCGs. Three fossil calibration points were selected from Timetree5 (http://www.timetree.org/): P. mallophylla-Aegiceras corniculatum (34.6-54.5 Mya), Rhododendron delavayi-Camellia sinensis (82.8-106.0 Mya), and C. sinensis-H. chinensis (109.0–122.0 Mya). Divergence time estimation was conducted using the MCMCTree module of PAML v4.9j (Yang, 2007), with the trimmed concatenated nucleotide matrix as input. Analyses were run for 1,000,000 generations, sampling every 100 generations, with the initial 20% of samples discarded as burn-in. Two independent chains were run to verify consistency, and convergence was assessed using Tracer.

2.7 Selective pressure analysis

Based on the alignments of PCGs from the mitochondrial and chloroplast genomes of the 16 species, The KaKs_Calculator 2.0 (Wang et al., 2010) was used to calculate the nonsynonymous substitution rate (Ka), synonymous substitution rate (Ks), and the Ka/Ks ratio for each gene, providing a preliminary assessment of evolutionary rate variation among genes within the overall phylogenetic framework. R scripts were then used to generate density plots and boxplots.

Furthermore, based on the phylogenetic tree constructed from the mitochondrial and chloroplast PCGs of the 16 species, the codeml program in the PAML package was used to perform branch-site model analyses to detect potential signals of positive selection along specific lineages. The branch containing P. mallophylla was designated as the foreground branch, while all other branches were treated as background branches. Two models were fitted for comparison: Alternative model, which allows a proportion of sites on the foreground branch to be under positive selection (ω > 1), and the null model, which restricts all ω values to ≤ 1. The goodness of fit between the two models was compared using likelihood ratio tests (LRTs), and statistical significance was determined based on a chi-square distribution with one degree of freedom. Genes with p < 0.05 were considered to be under significant positive selection on the foreground branch.

In addition, to further explore the characteristics of positive selection within the genus Primula, the same analytical procedure was independently applied to the chloroplast PCGs dataset comprising 13 closely related Primula species and two outgroup species (Supplementary Tables S2, S3).

2.8 Genome structural comparison

The ten representative species of Ericales were selected for collinearity analysis of mitochondrial genomes. Conserved homologous sequences among these species were identified using BLASTn with parameters ‘-value 1e-5, - word_size 9, -gapopen 5, -gapextend 2, -reward 2, -penalty -3’. Based on the comparison results, the Multiple Synteny Plot was generated using MCscanX (Wang et al., 2012) for collinear blocks longer than 500 bp. Comparative analysis of the chloroplast genome structures of P. mallophylla and its four closely related species was conducted. The expansion and contraction of boundaries were examined and seen with IRscope (Amiryousefi et al., 2018), while sequence comparison was conducted using mVISTA (Frazer et al., 2004). To visually illustrate the rearrangements and structural differences of mitochondrial and chloroplast genomes, we performed whole-genome alignments using MUMmer v4 (Marçais et al., 2018) and generated gnuplot scripts with mummerplot, followed by the construction of dotplots.

3 Results

3.1 General features of organelle genomes

The high-quality mitochondrial and chloroplast genomes of P. mallophylla (Figures 1A, B) were sequenced, assembled, and annotated to identify variations and conservations (Figures 1C, D). The assembly results showed that each organelle genome of P. mallophylla was a closed-monocyclic structure (Supplementary Figure S1), with lengths of 340,219 bp and 150,733 bp, and GC contents of 45.57% and 37.09%, respectively. The chloroplast genome has a typical quadripartite structure, consisting of an 82,622 bp LSC region, a 17,743 bp SSC region, and two 25,184 bp IR regions.

Figure 1
A displays pink flowers of Primula mallophylla. B shows a lush, mountainous landscape. C and D feature circular diagrams of mitochondrial and chloroplast genomes of Primula mallophylla, annotated with gene locations and color-coded sections.

Figure 1. Morphology (A), habitat (B), and organelle genomes (C, D) of P. mallophylla. (A) shows a flowering individual, and both (A, B) were photographed in Houping Township, Chengkou County, Chongqing, China, by Prof. Hongping Deng. Genes in (C, D) are color-coded according to their functions, arrows indicate transcriptional direction, and the inner histogram represents GC content.

In the mitochondrial genome, we annotated a total of 62 genes, including 37 unique PCGs, 22 tRNA genes, and 3 rRNA genes (Supplementary Table S4). Among the 37 unique PCGs, 24 were considered core genes, comprising five ATP synthase genes, nine NADH dehydrogenase genes, four cytochrome c biogenesis genes, three cytochrome c oxidase genes, one protein transport subunit gene, one maturase gene, and one cytochrome b gene. In the chloroplast genome, a total of 175 genes were annotated, including 130 PCGs, 37 tRNA genes, and 8 rRNA genes.

3.2 Repeat sequences analysis

The mitochondrial genome and chloroplast genome of P. mallophylla differed significantly in the type and distribution of repeat sequences (Figure 2). The mitochondrial genome was more abundant in repeat sequences, with a total of 111 SSRs, 6 tandem repeats, and 196 dispersed repeats identified, which were widely distributed in various regions of the genome. Among the SSRs, monomeric repeats were dominant, particularly A/T, which accounted for 44.14%. The 6 tandem repeats were all located in the intergenic region, and the repeat units ranged from 11–57 bp. The dispersed repeats included only palindromic and forward repeats, accounting for 43.37% and 56.63%, respectively. However, most of the repeat units were between 30 bp and 100 bp in length, and only 13 repeat units were longer than 100 bp. The longest dispersed repeat was a 7,613 bp forward repeat, which could be a potential region for genomic rearrangement. In contrast, the chloroplast genome has relatively conserved repeat sequences. There were 36 SSRs, 15 tandem repeats, and 29 dispersed repeats in the chloroplast genome. The SSRs consisted of 29 monomeric repeats (all A/T), 5 dimeric, and 2 tetrameric, which were predominantly distributed in the LSC region. In contrast to the mitochondrial genome, the chloroplast genome contained more palindromic repeats than forward repeats, and the repeat units were all within 30–60 bp in length.

Figure 2
Diagram showing repeat sequences of mitochondrial and chloroplast genomes. Panels A and B display circular maps illustrating the spatial distribution of repeats within each genome. In the innermost circle, dispersed repeat pairs are connected by colored lines, with palindromic repeats in yellow and forward repeats in purple. Black line segments on the middle and outer circles mark the positions and lengths of tandem repeats and simple sequence repeats (SSRs). Panel C presents a bar chart comparing the number of different repeat types between the two organelle genomes.

Figure 2. Analysis of repeat sequences in two organelle genomes of P. mallophylla. (A, B) depict the spatial distribution of repeat sequences inside the mitochondrial and chloroplast genomes, respectively. In the innermost circle, pairs of dispersed repeats are connected by colored lines, with palindromic repeats shown in yellow and forward repeats shown in purple. The black line segments on the middle and outermost circles denote the locations and lengths of tandem repeats and SSRs, respectively. (C) presents a bar chart summarizing the counts of each type of repeat sequence in the organelle genomes.

3.3 Prediction of RNA editing events

We used 37 unique PCGs from the mitochondrial genome of P. mallophylla to predict RNA editing events, identifying a total of 475 predicted sites (Figure 3). Notably, the ccmB gene demonstrated the greatest frequency of RNA editing events, with 37 sites. In contrast, the rps14 and rpl2 genes exhibited less RNA editing activity, possessing only one editing site each, suggesting lower variability in these genes. Further analysis showed that 210 RNA editing sites (44.21%) changed from hydrophilic to hydrophobic amino acids, 155 sites (32.63%) from hydrophobic-hydrophobic, 67 sites (14.11%) from hydrophilic-hydrophilic, 36 sites (7.58%) from hydrophobic-hydrophilic, and 7 sites (1.47%) from hydrophilic-stop (Supplementary Table S5). In addition, 213 sites (44.84%) were changed to leucine, indicating leucine propensity.

Figure 3
Bar chart showing the number of RNA editing sites for various genes of mitochondrial genome. Peaks include atp9 with 36, cox2 with 30, and nad3 with 33. Genes like atp4, matR, and rpl16 have fewer sites, ranging from 1 to 6. The chart illustrates the variability in RNA editing across different genes.

Figure 3. Identification of RNA editing sites for PCGs in the mitochondrial genome of P. mallophylla.

3.4 Condon usage bias analysis

The codon usage frequencies in each organelle genome of P. mallophylla were approximately the same, but with some differences (Figures 4A, B). In both the mitochondrial and chloroplast genomes, the RSCU values for the majority of codons, excluding AUG (the start codon for methionine) and UGG (which encodes tryptophan), were either above or below 1. This suggests a generalized codon usage bias for the PCGs in each organelle genome. In the mitochondrial genome, 29 codons exhibited RSCU values over 1, with alanine showing the greatest preference for the GCU codon (RSCU = 1.6). In the chloroplast genome, 29 codons exhibited RSCU values over 1, with the UUA codon for leucine showing the strongest usage bias, having an RSCU value of 2.03 (Supplementary Table S6). In addition to this, of these codons, only 2 codons in the mitochondrial and 1 codon in the chloroplast genome ended with C/G endings, and the rest of the codons were A/U endings, showing a strong preference for A/U usage.

Figure 4
Bar charts and line graphs depict codon usage and gene expression data. Panels A and B show RSCU values for mitochondrial and chloroplast genes, respectively, with various codon types. Panels C and D display ENC versus GC3s for mitochondrial and chloroplast genes, indicating relationships between different gene categories. Color coding represents mitochondrial respiratory chain, chloroplast photosynthesis, transcription and translation, and other genes.

Figure 4. Codon usage characteristics of PCGs in the organelle genomes of P. mallophylla. (A, B) depict the distribution of RSCU values in the mitochondrial and chloroplast genomes, respectively. The horizontal axis represents 20 amino acids and stop codon (End), while the vertical axis indicates RSCU values for the different codons. (C, D) display ENC-GC3s plots for the mitochondrial and chloroplast genomes, respectively. The solid curve represents the expected ENC values under the assumption that codon usage bias is solely driven by GC3s composition. In both plots, green dots represent genes involved in the respiratory chain (C) or photosynthesis (D), pink dots denote genes related to transcription and translation, and orange dots correspond to other genes.

ENC-GC3s analysis can further help understand the underlying drivers behind codon usage (Li et al., 2023). In each organelle of P. mallophylla, most of the PCGs were distributed on or near the standard curve (Figures 4C, D). In mitochondria, rps1, rps13, and sdh3 were farthest from the standard curve, whereas in chloroplasts, atpF, psbA, and rps4 were farthest.

3.5 Homologous sequences analysis between two orangelle genomes

Horizontal gene transfer (HGT) is a common phenomenon in angiosperms. To investigate potential intracellular transfers, we examined the migration of chloroplast-derived sequences into the mitochondrial genome of P. mallophylla (Figure 5). A total of six mitochondrial plastid DNA (MTPTs) were identified, spanning 2,028 bp in total. These MTPTs contained six complete genes, including 1 PCGs (ycf15) and 5 tRNA-encoding genes (trnD-GUC, trnI-CAU, trnM-CAU, trnN-GUU, trnW-CCA) (Supplementary Table S7). Among these transfer segments, the longest homologous fragment detected was MTPT2, which was 728 bp in length and contained two genes, the complete gene ycf15 and part of the gene ycf2.

Figure 5
Circular diagram comparing mitochondrial DNA (mtDNA) and chloroplast DNA (cpDNA). The two arcs, one labeled mtDNA in blue and the other cpDNA in green, are connected by orange lines indicating syntenic regions. Gene names are marked around the arcs with positions in kilobases.

Figure 5. Intracellular genome transfer in P. mallophylla. The mitochondrial genome is illustrated in blue, the chloroplast genome in green, and the orange arcs within the circular ring represent homologous segments shared by both genomes. Complete genes are labeled in black on the outer circle, whereas partial genes are shown in gray.

3.6 Phylogenetic relationships and divergence time estimation

Phylogenetic trees were constructed based on the shared PCGs of mitochondrial and chloroplast genomes, including 15 species of Ericales and one outgroup. The results showed that (Figure 6; Supplementary Figures S2A, S2B), for both the DNA sequence and codon-partitioned datasets, the topologies of the ML and BI trees were fully congruent, indicating a high consistency of phylogenetic signal within each dataset. The results showed that both mitochondrial and chloroplast PCG-based phylogenetic trees consistently indicate a close relationship between Primulaceae and Ebenaceae. However, the phylogenetic trees based on mitochondrial and chloroplast PCGs showed some differences in certain clades. In the mitochondrial PCG tree, clade I (Primulaceae and Ebenaceae) was closer to clade II (Symplocaceae and Theaceae) with low support (Figure 6A), whereas in the chloroplast PCG tree, clade II was closer to clade III (Actinidiaceae and Ericaceae) (Figure 6B). Further analyses based on complete chloroplast genomes showed that the ML tree placed clade II closer to clade III (Supplementary Figure S2C), while the BI tree grouped all families except Primulaceae into one clade, with a low posterior probability of 0.7505 (Supplementary Figure S2D). To assess the congruence of phylogenetic signals, we conducted gCF analysis on trees constructed from mitochondrial and chloroplast genome PCGs. The results showed that the gCF values for clade I were 33.33% and 22.03% in the mitochondrial and chloroplast genomes, respectively. In the mitochondrial genome, the gCF value between clade I and clade II was 20%, whereas in the chloroplast genome, the gCF value between clade II and clade III was 16.95%. Overall, the phylogenetic signals of different organelle genomes showed some differences, but both supported the primary familial relationships within Ericales. Based on this 16-species phylogeny, the divergence time between P. mallophylla and P. sikkimensis was estimated at approximately 2.13 Mya, and the divergence between Primula and Aegiceras at approximately 42.99 Mya (Figure 6B).

Figure 6
Phylogenetic trees comparing Mitochondrion-PCGs and Chloroplast-PCGs, showcasing genetic relationships among various plant species. The trees display bootstrap/SH-aLRT support values, Bayesian posterior probabilities and species names with accession numbers, color-coded by family: Primulaceae, Ebenaceae, Symplocaceae, Theaceae, Actinidiaceae, Ericaceae, and Hydrangeaceae. Each tree includes a scale bar indicating branch length units.

Figure 6. Maximum likelihood (ML) and Bayesian inference (BI) phylogenetic trees constructed based on shared PCGs from the mitochondrial genome (A) and chloroplast genome (B) of 16 species. Branches show bootstrap/SH-aLRT support values above and Bayesian posterior probabilities below. In panel B, nodes indicate divergence times with 95% confidence intervals. NCBI accession numbers for each species are provided.

Further ML and BI trees were constructed for the complete chloroplast genomes (Figure 7A) and shared PCGs (Figure 7B) of 86 species in Primula to clarify the evolutionary relationships of this genus. The results showed that 88 species of Primula were clustered into three large clades, and the trees constructed based on chloroplast genomes and the PCGs were essentially consistent with each other. The sect. Proliferae was divided into 2 groups, of which P. mallophylla was clustered with P. stenodonta, P. bulleyana, P. beesiana, P. chungensis, and was closest related to P. stenodonta. In addition to the sect. Proliferae, the sect. Crystallophlomis and Petiolares were also not a monophyletic group.

Figure 7
Phylogenetic trees titled “Chloroplast (genomes)” and “Chloroplast (PCGs)” compare genetic relationships. Both trees show hierarchical clustering of species within sections, with numerical values indicating bootstrap/SH-aLRT support values, and Bayesian posterior probabilities. Color-coded sections include Crystallophomis, Ranunculoides, Cordifoliae, Proliferae, Ametytina, and Petiolares. Tree scales are provided for reference.

Figure 7. Maximum likelihood (ML) and Bayesian inference (BI) phylogenetic trees constructed from the chloroplast genomes (A) and shared PCGs (B) of 88 species. Branches show bootstrap/SH-aLRT support values above and Bayesian posterior probabilities below. NCBI accession numbers for each species are provided.

3.7 Selective pressure on PCGs in organelle genomes of Primula

To investigate the evolutionary patterns of PCGs in different organelle genomes, Ka/Ks analyses were conducted on mitochondrial and chloroplast PCGs based on a phylogeny of 16 species (15 Ericales species and one outgroup) (Supplementary Figure S3). Chloroplast genes exhibited higher Ks values than mitochondrial genes, indicating faster accumulation of neutral mutations. Most genes in both organelles had Ka/Ks < 1, suggesting strong purifying selection. In the mitochondrial genome, atp4, ccmB, and ccmFN showed evidence of positive selection, whereas no chloroplast genes were under significant positive selection. Branch-site analyses revealed that, when P. mallophylla was set as the foreground branch, no significantly selected genes were detected in either organelle. When Primula was the foreground branch, rpl22 and rbcL in the chloroplast genome were under significant positive selection (p < 0.05), indicating potential adaptive evolution (Supplementary Table S8).

Further analyses of 13 Primula species plus two outgroup showed that chloroplast PCGs were overall under strong purifying selection (Supplementary Figure S4), but in P. mallophylla, rpl2 and ndhB were significantly positively selected (p < 0.05), suggesting roles in species-specific adaptation (Supplementary Table S8).

3.8 Comparative analysis of organelle genome structure

Comparison of mitochondrial genome homologous fragments from 10 representative species of the Ericales showed that the length of their collinear blocks (>500bp) was short (Figure 8A). A total of 103 homologous collinear blocks in length were identified among the P. mallophylla, P. sikkimensis, and Aegiceras corniculatum, all belonging to the Primulaceae family. The longest collinear block up to 9,987bp between P. mallophylla and chromosome 2 of P. sikkimensis, which are sister species in the phylogenetic tree. This indicated that the mitochondrial genome of P. mallophylla underwent rearrangement with closely related species and was extremely non-conservative in the structure.

Figure 8
Diagram featuring three panels. Panel A shows a circular mitochondrial genome comparison across ten species, with connecting lines illustrating gene relationships. Panel B displays a chloroplast genome alignment among five species using color-coded blocks to indicate gene conservation and variations, labeled with gene names. Panel C illustrates linear chloroplast genome maps for five Primula species, highlighting inverted repeat regions and various gene segments with color differentiation.

Figure 8. Comparative analysis of the structural characteristics of organelle genomes. (A) Collinearity analysis of the mitochondrial genomes from 10 representative species of Ericales, with red arcs indicating inversion regions and gray areas representing homologous regions. (B) mVISTA analysis of the chloroplast genomes from five Primula species, where exons, untranslated regions (UTRs), conserved noncoding sequences (CNS), and mRNA are distinctly color-coded. Gray arrows indicate the direction of gene transcription, and the vertical axis denotes the proportion of sequence identity ranging from 50% to 100%. (C) Visualization of the chloroplast IR boundaries in five Primula species. JLB, JSB, JSA, and JLA represent the junctions between the LSC and IRb, SSC and IRb, SSC and IRa, and LSC and IRa, respectively.

Phylogenetic analysis of the chloroplast genome indicated that P. stenodonta, P. bulleyana, P. beesiana, and P. chungensis are most closely related to P. mallophylla. Consequently, we visualized and compared these five species' chloroplast genomes to reveal their genome structure variation characteristics more comprehensively (Figures 8B, C). The result showed that the coding regions exhibit greater conservation than non-coding regions. The majority of the observed changes were located in non-coding sequence regions (Figure 8B). Then, we compared the IR boundaries of five Primula species. The results indicated that IR regions exhibited greater conservation than the LSC and SSC regions, and all five species were found to have the same genes at the boundaries, but with slightly different sizes (Figure 8C). The mitochondrial and chloroplast genome dotplots of P. mallophylla and its closely related species are shown in Supplementary Figure S5.

4 Discussion

4.1 Genomic features of organelles of P. mallophylla

In this study, we successfully assembled high-quality mitochondrial and chloroplast genomes of the endangered species P. mallophylla. Due to the advantages of long segments and high accuracy of HiFi long-reads, it is able to span complete replicates for continuous and accurate assembly. Therefore, our assembled mitochondrial genomes are reliable and bring an extremely important reference to the study of Primula, which is currently very scarce. In comparison to the mitochondrial genomes of three Primula species constructed from second-generation data (Wei et al., 2022) and the unpublished mitochondrial genome of P. sikkimensis (OR402838.1- OR402843.1), the mitochondrial genomes of P. mallophylla were essentially the same as them in terms of length (340, 219 bp) and GC content (45.57%). Unlike the six chromosomes of P. sikkimensis, P. mallophylla exhibits the typical master circular molecules consistent with many higher plants. In addition, collinear analysis showed that P. mallophylla had more collinear blocks with P. sikkimensis, but most of them underwent rearrangements. Differences in mitochondrial genome structure between these two species may be a key driver of species differentiation and adaptation to different environmental stresses. In the common ancestor of angiosperms, there are typically 41 conserved and stable PCGs in mitochondrial genomes (Mower et al., 2012). Of these, 24 core PCGs are possessed in almost all species, leaving 17 variable genes that vary from species (Skippington et al., 2015). The mitochondrial genome of P. mallophylla has 37 PCGs, with 24 core genes. However, the ribosomal small subunit genes, rps2, rps7, rps11, and rps19 are missing among the variable genes. Except for rps19, P. mallophylla is in agreement with P. palinuri and P. smithiana (Wei et al., 2022). The rps2 gene has been reported to be lost from the mitochondrial genomes of most dicotyledons and transferred to the nucleus (Perrotta et al., 2002). However, it is still retained in a few dicotyledons, such as Nelumbo nucifera (Gui et al., 2016). The loss of the rps7, rps11, and rps19 is prevalent in angiosperms (Adams et al., 2002). Overall, the P. mallophylla mitochondrial genome is highly conserved during evolution.

The chloroplast genome of P. mallophylla exhibits a characteristic quadripartite structure, conserved in accordance with other Primula species (Ren et al., 2018; Xie et al., 2023). Because chloroplasts carry key functions of photosynthesis and basic metabolism in plants, they are driven by strong purifying selection so that they are usually highly conserved structurally (Jansen and Ruhlman, 2012). According to the phylogenetic results, P. mallophylla is closely related to P. stenodonta, P. bulleyana, P. beesiana, and P. chungensis. Comparative structural analyses indicated that the chloroplast genome structures of these five species were markedly similar and likely diverged within a relatively short time. In contrast, the chloroplast genomes of species in the other group of the sect. Poliferea are relatively more structurally diverse and may have undergone more evolutionary events, including genome rearrangements, expansions, or contractions (Xie et al., 2023).

In the organelle genomes of P. mallophylla, we detected a large number of repeat sequences, which are crucial for maintaining organelle genome structure and mediating genome reorganization. We found that the repeat sequences in mitochondria far exceeded those in chloroplasts. The mitochondrial genome has evolved to undergo more complex gene rearrangements, losses, and transfers, which together drive the high diversity and complexity of its structure (Wynn and Christensen, 2019). In contrast, the chloroplast genome is more conserved, with less accumulation of repeat sequences. In the mitochondrial genome, repeat sequences are mainly dominated by dispersed repeats, whereas in the chloroplast genome, they are dominated by SSRs, which is related to the differences in genome structure and function between the two.

In addition to the important exchange of genetic material that occurs between the organelle genomes and the nuclear genome, such exchanges are very common among organelles, especially chloroplast-to-mitochondrial sequence transfer (MTPT), which is crucial for plant growth and adaptive evolution (Thorsness and Weber, 1996; Cusimano and Wicke, 2016; Cheng et al., 2021). In P. mallophylla, we identified 9 MTPTs with a cumulative length of 2028 bp, constituting 0.60 % of the total length of the mitochondrial genome, which aligns with the structural characteristics of other Primula plants (Xie et al., 2023). Among these homologous fragments, the vast majority are complete or partial fragments of chloroplast trn genes (trnN-GUU, trnD-GUC, trnW-CCA, trnM-CAU, trnI-GAU, trnA-UGC, trnI-GAU). Research indicates that mitochondrial tRNAs are derived not only from ancestral mitochondria but also from chloroplasts through horizontal gene transfer (HGT) (Sprinzl and Vassilenko, 2005; Rice et al., 2013). Although the sequence lengths of the respective MTPTs in P. mallophylla are short, it is also well demonstrated that these chloroplast-derived tRNAs play an essential role in maintaining important mitochondrial functions (Kanno et al., 1997; Menand et al., 1998).

4.2 Expression patterns and adaptive evolution of organelle genomes in P. mallophylla

In plant organelle genomes, the process of gene expression is uniquely regulated, especially in terms of RNA editing at the post-transcriptional stage and codon usage preferences at the translational stage, which are important for the stability and adaptation of organelle function.

In plant mitochondria, post-transcriptional mRNAs often undergo C-U RNA editing, allowing the edited protein sequences to more closely resemble homologous proteins from other species, enhancing sequence conservation across species and further optimizing the efficiency of gene expression within mitochondria (Covello and Gray, 1989). In P. mallophylla, we predicted 475 editing sites for C-U. Among them, ccmB and mttB have the highest editing frequencies, which are also high in Primula (Wei et al., 2022), Dendrobium (Wang et al., 2024b), Solanum muricatum (Li et al., 2024b) and other species, and are usually the two most frequently edited genes, which may be related to the maintenance of mitochondrial respiratory chain policy operation and energy metabolism. Furthermore, we discovered that 44.21% of amino acids transitioned from hydrophilic to hydrophobic following RNA editing, enhancing the protein's folding and functionality (Yura and Go, 2008). It is noteworthy that the RNA editing results in this study were predicted using the Deepred-mt model. This approach efficiently identifies potential C-to-U editing sites in mitochondrial genes and provides valuable reference information. However, as model-based predictions are influenced by algorithmic and parameter settings, some false positives may exist. Therefore, the present results mainly represent the putative mitochondrial RNA editing characteristics of P. mallophylla.

The preference of amino acids for synonymous codon usage during the translation step is prevalent across various prokaryotic and eukaryotic organisms (Zhang et al., 2007). Codon use bias is shaped by mutational pressure, natural selection, and genetic drift, significantly contributing to the adaptive evolution of plants (Parvathy et al., 2022; Lyu et al., 2024). Eeach organelle genome of P. mallophylla has 29 codons with RSCU values over 1, with the predominant majority terminating in A/U, aligning with the preference shown in most plants, indicating a degree of consistency in codon usage among plant species (Liu et al., 2024a; Wang et al., 2024b). ENC-GC3s can further determine whether codon usage preferences are mainly influenced by mutational pressure or natural selection (Li et al., 2023; Wright, 1990). The ENC-GC3s theoretical curve represents the trend of the ENC with the GC content of GC3s driven by GC mutation preference only. If the measured values of ENC are significantly lower than the theoretical values, it suggests that natural selection may have a potential influence. The rps1, rps13, and sdh3 genes in mitochondria and the atpF, psbA, and rps4 genes in chloroplasts exhibit significant differences from theoretical values, which may be subject to strong selection due to adaptation to the environment.

In studies of adaptive evolution in organelle genomes, Ka and Ks are commonly used to evaluate gene mutation accumulation and the strength of natural selection. The Ka/Ks ratio can reveal whether a gene is subject to purifying selection, positive selection, or neutral evolution. Our analysis of the evolutionary rates and selective pressures of PCGs in the organelle genomes of Ericales species found that the chloroplast genome has higher Ks values than the mitochondrial genome, further confirming that plant mitochondrial genes evolve more slowly than chloroplast genes (Drouin et al., 2008). However, the overall Ka/Ks ratio of the chloroplast is lower than that of the mitochondrion, indicating that although chloroplast genes are more prone to mutations, their evolution is under stricter purifying selection and exhibits greater functional conservation. This difference may arise from distinctions between the two organelles in replication frequency, repair mechanisms, and the intensity of functional constraints (Christensen, 2013; Smith and Keeling, 2015). Overall, mitochondrial and chloroplast genes are under strong purifying selection due to their critical roles in core functions such as energy metabolism and photosynthesis. Further branch-site analyses were conducted to identify branch-specific positive selection. When the genus Primula was set as the foreground branch, rpl22 and rbcL in the chloroplast were detected as positively selected genes (p < 0.05). In an independent analysis of 13 Primula species plus two outgroup, chloroplast PCGs were generally under purifying selection, but rpl2 and ndhB were identified as positively selected in P. mallophylla (p < 0.05). These four genes are closely associated with plant adaptation to alpine stresses such as low temperature, high light, and UV-B radiation (Nyamgerel et al., 2024; Zhang et al., 2024; Lu et al., 2025). rbcL, encoding the large subunit of Rubisco, may help maintain photosynthetic efficiency under high-altitude light and temperature conditions; ndhB, involved in the chloroplast NADH dehydrogenase complex, may facilitate cyclic electron flow in photosystem I, mitigating oxidative damage under high light; rpl2 and rpl22, both ribosomal large subunit proteins, may reflect adaptive regulation of protein translation efficiency or stability in alpine environments.

4.3 Phylogenetic inference based on different organelle genomes

Mitochondrial and chloroplast genomes are usually matrilineally inherited and conserved, providing clear genetic signals that facilitate phylogenetic analysis. Due to the limited mitochondrial genomes in the Primulaceae, we reconstructed the phylogeny of related families of Ericales utilizing the conserved PCGs of the mitochondrial and chloroplast genomes. Phylogenetic analyses using different datasets and methods consistently support a stable placement of Primulaceae, generally showing a close relationship with Ebenaceae. Some topological differences were observed, with PCG-based trees clustering Primulaceae with Ebenaceae, while the BI tree based on complete chloroplast genomes grouped all other Ericales families together, consistent with chromosome-level genome analyses (Yang et al., 2022; Feng et al., 2021). These discrepancies likely reflect differences in evolutionary rates, gene rearrangements, and limited phylogenetic signal in individual PCGs, yet the overall results confirm the reliable phylogenetic position of Primulaceae.

The chloroplast genomes of 88 species were used to further explore the phylogenetic relationships of the Primula in depth. The ML and BI trees constructed on the basis of the whole chloroplast genome and conserved PCGs were basically consistent with each other, dividing the Primula into three major clades, which is in agreement with the results of a previous study (Xie et al., 2023). Focusing on the clades in which sect. Proliferae is located, revealing that sect. Proliferae, Crystallophlomis, and Petiolares are not monophyletic groups, which is inconsistent with traditional taxonomy (Hu, 1990). 13 species of the sect. Proliferae are clustered into two branches, one including P. mallophylla, P. stenodonta, P. bulleyana, P. beesiana, and P. chungensis, and the rest are in the other. The sect. Amethyatina and Petiolares nested in the Proliferae have been agreed upon in studies using the chloroplast genome (Xie et al., 2023). In morphology, the three sections also have similar characteristics, such as umbel, spherical capsule, bell calyx (Hu, 1990). However, in the nuclear gene phylogenetic trees or joint phylogenetic trees of nuclear and chloroplast genes, the species of the sect. Proliferae cluster into a monophyletic lineage (Liu, 2022; Yan et al., 2015), reflecting the conflict of phylogenetic signaling of nucleoplasmic genes. This phenomenon was also found in a comparison of whole genomes and chloroplast genomes of 80 individuals of Asian butternuts (Xu et al., 2021). The isolation-by-distance analysis and gene flow estimation indicated that, relative to seeds, pollen exhibited significantly higher gene flow levels. This may have impeded lineage differentiation and facilitated greater interaction of the nuclear genome compared to the chloroplast genome, resulting in discordant nucleoplasmic gene phylogenetic signals. With the interspecific hybridization, parallel evolution, gene introgression, and other complex phenomena occurring frequently, the phylogenetic relationship of Primula is becoming complex. Therefore, conducting genome-wide phylogenetic analyses of species is essential for acquiring a more thorough comprehension of their evolutionary history and elucidating the fundamental biological mechanisms by incorporating genomic variations.

5 Conclusions

In this study, we report high-quality mitochondrial genome and chloroplast genome structural features of P. mallophylla for the first time, including repeat sequences, RNA editing events, codon preference, and intracellular gene transfer. Combined analyses of Ka/Ks ratios and branch-site tests reveal the evolutionary rates and adaptive evolution features of P. mallophylla at the organelle genome level. The phylogenetic results showed that the P. stenodonta is sister to the P. mallophylla, and the sect. Proliferae in which it is found is not a monophyletic group. The collinear analysis with the relatives indicated that the mitochondrial genome of P. mallophylla experienced genomic rearrangement. However, the chloroplast genome is less different from its relatives in sequence comparison and boundary analysis, and the overall structure is more conservative. The results of this study not only provide molecular evidence for clarifying the taxonomic status and reconstructing the evolutionary background of P. mallophylla, but also provide important references for the conservation of germplasm resources and systematic evolution of endangered plants.

Data availability statement

The mitochondrial and chloroplast genome sequences of P. mallophylla generated in this study have been deposited in the NCBI GenBank database (https://www.ncbi.nlm.nih.gov/) under the accession numbers PV527762.1 and PQ31024.1, respectively. All raw data have been stored in the SRA database (BioProject: PRJNA1260928; BioSample: SAMN48407255; SRA: SRR33482105, SRR33482106).

Author contributions

WL: Methodology, Investigation, Writing – review & editing, Visualization, Formal Analysis, Writing – original draft, Software. LW: Writing – original draft, Methodology, Software, Visualization, Writing – review & editing. YZ: Writing – review & editing, Methodology, Supervision. HZ: Writing – review & editing, Data curation, Visualization, Validation, Software. XZ: Software, Data curation, Validation, Writing – review & editing. LY: Writing – review & editing, Resources, Validation. JZ: Writing – review & editing, Investigation, Validation. FZ: Validation, Investigation, Writing – review & editing. HD: Writing – review & editing, Methodology, Supervision, Project administration, Resources.

Funding

The author(s) declare financial support was received for the research and/or publication of this article. This work was supported by the National Key Protected Wildlife Protection Project -zlg2021-cq20211210, zlg2022-cq20220907.

Acknowledgments

We are grateful to the reviewers for their thorough reviews and suggestions that helped to improve this paper.

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.

Any alternative text (alt text) provided alongside figures in this article has been generated by Frontiers with the support of artificial intelligence and reasonable efforts have been made to ensure accuracy, including review by the authors wherever possible. If you identify any issues, please contact us.

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.2025.1653128/full#supplementary-material

References

Adams, K. L., Qiu, Y., Stoutemyer, M., and Palmer, J. D. (2002). Punctuated evolution of mitochondrial gene content: High and variable rates of mitochondrial gene loss and transfer to the nucleus during angiosperm evolution. Proc. Natl. Acad. Sci. U.S.A. 99, 9905–9912. doi: 10.1073/pnas.042694899

PubMed Abstract | Crossref Full Text | Google Scholar

Altschul, S. (1997). Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res. 25, 3389–3402. doi: 10.1093/nar/25.17.3389

PubMed Abstract | Crossref Full Text | Google Scholar

Amiryousefi, A., Hyvönen, J., and Poczai, P. (2018). IRscope: an online program to visualize the junction sites of chloroplast genomes. Bioinformatics 34, 3030–3031. doi: 10.1093/bioinformatics/bty220

PubMed Abstract | Crossref Full Text | Google Scholar

Balfour, I. B. (1916). New species of primula. Notes R. Bot. Gard. Edinburgh 9, 181. doi: 10.24823/nrbge.1916.2266

Crossref Full Text | Google Scholar

Başbülbül, G., Özmen, A., Biyik, H. H., and Şen, Ö. (2008). Antimitotic and antibacterial effects of the Primula veris L. flower extracts. Caryologia 61, 88–91. doi: 10.1080/00087114.2008.10589614

Crossref Full Text | Google Scholar

Beier, S., Thiel, T., Münch, T., Scholz, U., and 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

Budzianowski, J., Morozowska, M., and Wesołowska, M. (2005). Lipophilic flavones of Primula veris L. from field cultivation and in vitro cultures. Phytochemistry 66, 1033–1039. doi: 10.1016/j.phytochem.2005.03.024

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, Y., Ye, W., Zhang, Y., and Xu, Y. (2015). High speed BLASTN: an accelerated MegaBLAST search tool. Nucleic Acids Res. 43, 7762–7768. doi: 10.1093/nar/gkv784

PubMed Abstract | Crossref Full Text | Google Scholar

Cheng, L., Wang, W., Yao, Y., and Sun, Q. (2021). Mitochondrial RNase H1 activity regulates R-loop homeostasis to maintain genome integrity and enable early embryogenesis in Arabidopsis. PloS Biol. 19, e3001357. doi: 10.1371/journal.pbio.3001357

PubMed Abstract | Crossref Full Text | Google Scholar

Christensen, A. C. (2013). Plant mitochondrial genome evolution can be explained by DNA repair mechanisms. Genome Biol. Evol. 5, 1079–1086. doi: 10.1093/gbe/evt069

PubMed Abstract | Crossref Full Text | Google Scholar

Covello, P. S. and Gray, M. W. (1989). RNA editing in plant mitochondria. Nature 341, 662–666. doi: 10.1038/341662A0

PubMed Abstract | Crossref Full Text | Google Scholar

Cusimano, N. and Wicke, S. (2016). Massive intracellular gene transfer during plastid genome reduction in nongreen Orobanchaceae. New Phytol. 210, 680–693. doi: 10.1111/nph.13784

PubMed Abstract | Crossref Full Text | Google Scholar

Darwin, C. (1877). The different forms of flowers on plants of the same species. (London: John Murray).

Google Scholar

Doyle, J. J. and Doyle, J. L. (1987). A rapid DNA isolation procedure for small quantities of fresh leaf tissue. Phytochem. Bull. 19, 11–15.

Google Scholar

Drouin, G., Daoud, H., and 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

Edera, A. A., Small, I., Milone, D. H., and Sanchez-Puerta, M. V. (2021). Deepred-Mt: Deep representation learning for predicting C-to-U RNA editing in plant mitochondria. Comput. Biol. Med. 136, 104682. doi: 10.1016/j.compbiomed.2021.104682

PubMed Abstract | Crossref Full Text | Google Scholar

Feng, X., Li, G., Xu, S., Wu, W., Chen, Q., Shao, S., et al. (2021). Genomic insights into molecular adaptation to intertidal environments in the mangrove Aegiceras corniculatum. New Phytol. 231, 2346–2358. doi: 10.1111/nph.17551

PubMed Abstract | Crossref Full Text | Google Scholar

Frazer, K. A., Pachter, L., Poliakov, A., Rubin, E. M., and Dubchak, I. (2004). VISTA: Computational tools for comparative genomics. Nucleic Acids Res. 32, W273–W279. doi: 10.1093/nar/gkh458

PubMed Abstract | Crossref Full Text | Google Scholar

Gilmartin, P. M. (2015). On the origins of observations of heterostyly in Primula. New Phytol. 208, 39–51. doi: 10.1111/nph.13558

PubMed Abstract | Crossref Full Text | Google Scholar

Grote, A., Hiller, K., Scheer, M., Münch, R., Nörtemann, B., Hempel, D. C., et al. (2005). JCat: a novel tool to adapt codon usage of a target gene to its potential expression host. Nucleic Acids Res. 33, W526–W531. doi: 10.1093/nar/gki376

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

Halda, J. J. (1992). The genus primula in cultivation and in the wild (Prague: Tethys Books).

Google Scholar

Hu, C. M. (1990). Flora of China Vol. 59 (Beijing: Science Press).

Google Scholar

Hu, C. M. (1994). Geographical distribution of Primulaceae plants. J. Trop. Subtrop. Bot. 2, 1–14.

Google Scholar

Hu, C. M. (1998). Diversity and geographical distribution of Primula species in China. Plant Divers. Conserv. Taiwan Strait, 67–73.

Google Scholar

Hu, C. M. and Kelso, S. (1996). Flora of China Vol. 15 (Beijing: Science Press).

Google Scholar

Jansen, R. K. and Ruhlman, T. A. (2012). “Plastid genomes of seed plants,” in Genomics of chloroplasts and mitochondria. Eds. Bock, R. and Knoop, V. (Springer, Dordrecht).

Google Scholar

Jin, J., Yu, W., Yang, J., Song, Y., dePamphilis, C. W., Yi, T., 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

Kalyaanamoorthy, S., Minh, B. Q., Wong, T. K. F., von Haeseler, A., and Jermiin, L. S. (2017). ModelFinder: fast model selection for accurate phylogenetic estimates. Nat. Methods 14, 587–589. doi: 10.1038/NMETH.4285

PubMed Abstract | Crossref Full Text | Google Scholar

Kanno, A., Nakazono, M., Hirai, A., and Kameya, T. (1997). Maintenance of chloroplast-derived sequences in the mitochondrial DNA of Gramineae. Curr. Genet. 32, 413–419. doi: 10.1007/s002940050296

PubMed Abstract | Crossref Full Text | Google Scholar

Katoh, K. and 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

Khan, S., Shaheen, H., Mehmood, A., Nasar, S., and Khan, T. (2022). Ethnobotanical and antibacterial study of Primula plants traditionally used in the indigenous communities of Western Himalaya, Pakistan. Saudi J. Biol. Sci. 29, 3244–3254. doi: 10.1016/j.sjbs.2022.01.048

PubMed Abstract | Crossref Full Text | Google Scholar

Kolmogorov, M., Yuan, J., Lin, Y., and Pevzner, P. A. (2019). Assembly of long, error-prone reads using repeat graphs. Nat. Biotechnol. 37, 540–546. doi: 10.1038/s41587-019-0072-8

PubMed Abstract | Crossref Full Text | Google Scholar

Kumar, S., Stecher, G., and 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

Letunic, I. and Bork, P. (2024). Interactive Tree of Life (iTOL) v6: recent updates to the phylogenetic tree display and annotation tool. Nucleic Acids Res. 52, W78–W82. doi: 10.1093/nar/gkae268

PubMed Abstract | Crossref Full Text | Google Scholar

Lewis, S. E., Searle, S., Harris, N., Gibson, M., Iyer, V., Richter, J., et al. (2002). Apollo: a sequence annotation editor. Genome Biol. 3, 1–14. doi: 10.1186/gb-2002-3-12-research0082

PubMed Abstract | Crossref Full Text | 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, Q. (2022). The complete chloroplast genomes of Primula obconica provide insight that neither species nor natural section represent monophyletic taxa in Primula (Primulaceae). Genes 13, 567. doi: 10.3390/genes13040567

PubMed Abstract | Crossref Full Text | Google Scholar

Li, Z., Liu, J., Liang, M., Guo, Y., Chen, X., Wu, H., et al. (2024b). De novo assembly of the complete mitochondrial genome of pepino (Solanum muricatum) using PacBio HiFi sequencing: insights into structure, phylogenetic implications, and RNA editing. BMC Plant Biol. 24, 361. doi: 10.1186/s12870-024-04978-w

PubMed Abstract | Crossref Full Text | Google Scholar

Li, Y., Liu, Y., Zeng, X., Wu, P., Li, Q., Guo, S., et al. (2024c). Complete mitochondrial genome of Angelica dahurica and its implications on evolutionary analysis of complex mitochondrial genome architecture in Apiaceae. Front. Plant Sci. 15. doi: 10.3389/fpls.2024.1367299

PubMed Abstract | Crossref Full Text | Google Scholar

Li, J., Ni, Y., Lu, Q., Chen, H., and Liu, C. (2024a). PMGA: A plant mitochondrial genome annotator. Plant Commun. 6, 101191. doi: 10.1016/j.xplc.2024.101191

PubMed Abstract | Crossref Full Text | Google Scholar

Li, Z., Wang, Y., He, X., and Li, W. (2025). The Taihangia mitogenome provides new insights into its adaptation and organelle genome evolution in Rosaceae. Planta 261, 59. doi: 10.1007/s00425-025-04629-w

PubMed Abstract | Crossref Full Text | Google Scholar

Li, C., Zhou, L., Nie, J., Wu, S., Li, W., Liu, Y., et al. (2023). Codon usage bias and genetic diversity in chloroplast genomes of Elaeagnus species (Myrtiflorae: Elaeagnaceae). Physiol. Mol. Biol. Pla. 29, 239–251. doi: 10.1007/s12298-023-01289-6

PubMed Abstract | Crossref Full Text | Google Scholar

Liu, T. (2022). Phylogenetic genomics of the genus Primula based on transcriptomes (Guangzhou (Guangdong: South China Agricultural University).

Google Scholar

Liu, S., Ni, Y., Li, J., Zhang, X., Yang, H., Chen, H., et al. (2023). CPGView: A package for visualizing detailed chloroplast genome structures. Mol. Ecol. Resour. 23, 694–704. doi: 10.1111/1755-0998.13729

PubMed Abstract | Crossref Full Text | Google Scholar

Liu, D., Zhang, Z., Hao, Y., Li, M., Yu, H., Zhang, X., et al. (2024a). Decoding the complete organelle genomic architecture of Stewartia gemmata: an early-diverging species in Theaceae. BMC Genomics 25, 114. doi: 10.1186/s12864-024-10016-8

PubMed Abstract | Crossref Full Text | Google Scholar

Liu, G., Zuo, Y., Shan, Y., Yu, J., Li, J., Chen, Y., et al. (2024b). Structural analysis of the mitochondrial genome of Santalum album reveals a complex branched configuration. Genomics 116, 110935. doi: 10.1016/j.ygeno.2024.110935

PubMed Abstract | Crossref Full Text | Google Scholar

Lowe, T. M. and Eddy, S. R. (1997). tRNAscan-SE: a program for improved detection of transfer RNA genes in genomic sequence. Nucleic Acids Res. 25, 955–964. doi: 10.1093/nar/25.5.955

PubMed Abstract | Crossref Full Text | Google Scholar

Lu, T., Yin, Y., Luo, J., Chen, J., Wu, Y., Zhang, W., et al. (2025). Molecular phylogenetic relationships based on chloroplast genomes of Zingiberaceae species: Insights into evolution and adaptation to extreme environments. Front. Plant Sci. 16. doi: 10.3389/fpls.2025.1670568

PubMed Abstract | Crossref Full Text | Google Scholar

Lyu, Z., Yang, G., Zhou, X., Wang, S., Zhang, R., and Shen, S. (2024). Deciphering the complex organelle genomes of two Rhododendron species and insights into adaptive evolution patterns in high-altitude. BMC Plant Biol. 24, 1054. doi: 10.1186/s12870-024-05761-7

PubMed Abstract | Crossref Full Text | Google Scholar

Majid, A., Hassan, S., Hussain, W., Khan, A., Hassan, A., Khan, A., et al. (2014). In vitro approaches of Primula vulgaris leaves and roots extraction against human pathogenic bacterial strains. World Appl. Sci. J. 30, 575–580. doi: 10.5829/idosi.wasj.2014.30.05.82264

Crossref Full Text | Google Scholar

Marçais, G., Delcher, A. L., Phillippy, A. M., Coston, R., Salzberg, S. L., and Zimin, A. (2018). MUMmer4: A fast and versatile genome alignment system. Plos. Comput. Biol. 14, e1005944. doi: 10.1371/journal.pcbi.1005944

PubMed Abstract | Crossref Full Text | Google Scholar

Maréchal, A. and 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

Mast, A. R., Kelso, S., and Conti, E. (2006). Are any primroses (Primula) primitively monomorphic? New Phytol. 171, 605–616. doi: 10.1111/j.1469-8137.2006.01700.x

PubMed Abstract | Crossref Full Text | Google Scholar

Mast, A. R., Kelso, S., Richards, A. J., Lang, D. J., Feller, D. M., and Conti, E. (2001). Phylogenetic relationships in Primula L. and related genera (Primulaceae) based on noncoding chloroplast DNA. Int. J. Plant Sci. 162, 1381–1400. doi: 10.1086/323444

Crossref Full Text | Google Scholar

Menand, B., Maréchal-Drouard, L., Sakamoto, W., Dietrich, A., and Wintz, H. (1998). A single gene of chloroplast origin codes for mitochondrial and chloroplastic methionyl-tRNA synthetase in Arabidopsis thaliana. Proc. Natl. Acad. Sci. U.S.A. 95, 11014–11019. doi: 10.1073/pnas.95.18.11014

PubMed Abstract | Crossref Full Text | Google Scholar

Mower, J. P., Sloan, D. B., and Alverson, A. J. (2012). Plant mitochondrial genome diversity: the genomics revolution (Vienna: Springer).

Google Scholar

Nguyen, L. T., Schmidt, H. A., von Haeseler, A., and 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

Nyamgerel, N., Baasanmunkh, S., Oyuntsetseg, B., Tsegmed, Z., Bayarmaa, G. A., Lazkov, G., et al. (2024). Comparative plastome analysis and taxonomic classification of snow lotus species (Saussurea, Asteraceae) in Central Asia and Southern Siberia. Funct. Integr. Genomic. 24, 42. doi: 10.1007/s10142-024-01309-y

PubMed Abstract | Crossref Full Text | Google Scholar

Parvathy, S. T., Udayasuriyan, V., and Bhadana, V. (2022). Codon usage bias. Mol. Biol. Rep. 49, 539–565. doi: 10.1007/s11033-021-06749-4

PubMed Abstract | Crossref Full Text | Google Scholar

Pax, F. (1889). Monographische übersicht über die arten der gattung Primula. Bot. Jahrb. Syst. 10, 75–241.

Google Scholar

Perrotta, G., Grienenberger, J. M., and Gualberto, J. M. (2002). Plant mitochondrial rps2 genes code for proteins with a C-terminal extension that is processed. Plant Mol. Biol. 50, 523–533. doi: 10.1023/A:1019878212696

PubMed Abstract | Crossref Full Text | Google Scholar

Ranwez, V., Douzery, E. J. P., Cambon, C., Chantret, N., and Delsuc, F. (2018). MACSE v2: Toolkit for the alignment of coding sequences accounting for frameshifts and stop codons. Mol. Biol. Evol. 35, 2582–2584. doi: 10.1093/molbev/msy159

PubMed Abstract | Crossref Full Text | Google Scholar

Ren, T., Yang, Y., Zhou, T., and Liu, Z. (2018). Comparative plastid genomes of Primula species: Sequence divergence and phylogenetic relationships. Int. J. Mol. Sci. 19, 1050. doi: 10.3390/ijms19041050

PubMed Abstract | Crossref Full Text | Google Scholar

Rice, D. W., Alverson, A. J., Richardson, A. O., Young, G. J., Sanchez-Puerta, M. V., Munzinger, J., et al. (2013). Horizontal transfer of entire genomes via mitochondrial fusion in the angiosperm Amborella. Science 342, 1468–1473. doi: 10.1126/science.1246275

PubMed Abstract | Crossref Full Text | Google Scholar

Richards, A. J. (2002). Primula. 2nd ed (B. T. London: Batsford Ltd).

Google Scholar

Richards, A. J. (2003). Primula (Portland: Timber Press).

Google Scholar

Ronquist, F., Teslenko, M., van der Mark, P., Ayres, D. L., Darling, A., Hohna, S., et al. (2012). MrBayes 3.2: efficient Bayesian phylogenetic inference and model choice across a large model space. Syst. Biol. 61, 539–542. doi: 10.1093/sysbio/sys029

PubMed Abstract | Crossref Full Text | Google Scholar

Schmidt-Lebuhn, A. N., de Vos, J. M., Keller, B., and Conti, E. (2012). Phylogenetic analysis of Primula section Primula reveals rampant non-monophyly among morphologically distinct species. Mol. Phylogenet. Evol. 65, 23–34. doi: 10.1016/j.ympev.2012.05.015

PubMed Abstract | Crossref Full Text | Google Scholar

Schott, H. W. (1851). Die Sippen der österreichischen Primeln (Vienna: Gerold).

Google Scholar

Shi, L., Chen, H., Jiang, M., Wang, L., Wu, X., Huang, L., et al. (2019). CPGAVAS2, an integrated plastome sequence annotator and analyzer. Nucleic Acids Res. 47, W65–W73. doi: 10.1093/nar/gkz345

PubMed Abstract | Crossref Full Text | Google Scholar

Skippington, E., Barkman, T. J., Rice, D. W., and Palmer, J. D. (2015). Miniaturized mitogenome of the parasitic plant Viscum scurruloideum is extremely divergent and dynamic and has lost all nad genes. Proc. Natl. Acad. Sci. U.S.A. 112, E3515–E3524. doi: 10.1073/pnas.1504491112

PubMed Abstract | Crossref Full Text | Google Scholar

Smith, W. W. and Fletcher, H. R. (1947). The genus primula: sections obconica, sinenses, reinii, pinnatae, malacoides, bullatae, carolinella, grandis and denticulata. Trans. R. Soc Edinb. 61, 415–478. doi: 10.1017/S0080456800004828

Crossref Full Text | Google Scholar

Smith, D. R. and 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. and 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

Stefan, K., Choudhuri, J. V., Enno, O., Chris, S., Jens, S., and Robert, G. (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

Stephan, G., Pascal, L., and Ralph, B. (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

Stubbs, R. L., Theodoridis, S., Mora-Carrera, E., Keller, B., Potente, G., Yousefi, N., et al. (2024). The genomes of Darwin's primroses reveal chromosome-scale adaptive introgression and differential permeability of species boundaries. New Phytol. 241, 911–925. doi: 10.1111/nph.19361

PubMed Abstract | Crossref Full Text | Google Scholar

Talavera, G. and Castresana, J. (2007). Improvement of phylogenies after removing divergent and ambiguously aligned blocks from protein sequence alignments. Syst. Biol. 56, 564–577. doi: 10.1080/10635150701472164

PubMed Abstract | Crossref Full Text | Google Scholar

Thorsness, P. E. and Weber, E. R. (1996). Escape and migration of nucleic acids between chloroplasts, mitochondria, and the nucleus. Int. Rev. Cytol. 165, 207–234. doi: 10.1016/S0074-7696(08)62223-8

PubMed Abstract | Crossref Full Text | Google Scholar

Trift, I., Källersjö, M., and Anderberg, A. A. (2002). The monophyly of Primula (Primulaceae) evaluated by analysis of sequences from the chloroplast gene rbcL. Syst. Bot. 27, 396–407. doi: 10.1043/0363-6445-27.2.396

Crossref Full Text | Google Scholar

Wang, J., Kan, S., Liao, X., Zhou, J., Tembrock, L. R., Daniell, H., et al. (2024a). 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, L., Liu, X., Wang, Y., Ming, X., Qi, J., and Zhou, Y. (2024b). Comparative analysis of the mitochondrial genomes of four Dendrobium species (Orchidaceae) reveals heterogeneity in structure, synteny, intercellular gene transfer, and RNA editing. Front. Plant Sci. 15. doi: 10.3389/fpls.2024.1429545

PubMed Abstract | Crossref Full Text | Google Scholar

Wang, Y., Tang, H., DeBarry, J. D., Tan, X., Li, J., Wang, X., et al. (2012). MCScanX: a toolkit for detection and evolutionary analysis of gene synteny and collinearity. Nucleic Acids Res. 40, e49–e49. doi: 10.1093/nar/gkr1293

PubMed Abstract | Crossref Full Text | Google Scholar

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

PubMed Abstract | Crossref Full Text | Google Scholar

Wei, L., Liu, T., Hao, G., Ge, X., and Yan, H. (2022). Comparative analyses of three complete Primula mitogenomes with insights into mitogenome size variation in Ericales. BMC Genomics 23, 770. doi: 10.1186/s12864-022-08983-x

PubMed Abstract | Crossref Full Text | Google Scholar

Wendelbo, P. (1961). Studies in Primulaceae. II. An account of Primula subgenus Sphondylia (Syn. Sect. Floribundae) with a review of the subdivisions of the genus (Oslo: Norwegian Universities Press).

Google Scholar

Wick, R. R., Schultz, M. B., Zobel, J., and Holt, K. E. (2015). Bandage: interactive visualization of de novo genome assemblies. Bioinformatics 31, 3350–3352. doi: 10.1093/bioinformatics/btv383

PubMed Abstract | Crossref Full Text | Google Scholar

Wickham, H. (2011). ggplot2. WIREs Comput. Stat. 3, 180–185. doi: 10.1002/wics.147

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., Zhang, C., Qiao, Q., and Shen, M. (2009). Rediscovery of Primula mallophylla (Primulaceae), an endangered species endemic to Dabashan Mountain, China. Plant Divers. 31, 265–268. doi: 10.3724/SP.J.1143.2009.09003

Crossref Full Text | Google Scholar

Wynn, E. L. and Christensen, A. C. (2019). Repeats of unusual size in plant mitochondrial genomes: Identification, incidence and evolution. G3-Genes Genom Genet. 9, 549–559. doi: 10.1534/g3.118.200948

PubMed Abstract | Crossref Full Text | Google Scholar

Xia, C., Li, J., Zuo, Y., He, P., Zhang, H., Zhang, X., et al. (2023). Complete mitochondrial genome of Thuja sutchuenensis and its implications on evolutionary analysis of complex mitogenome architecture in Cupressaceae. BMC Plant Biol. 23, 84. doi: 10.1186/s12870-023-04054-9

PubMed Abstract | Crossref Full Text | Google Scholar

Xie, Y., Yang, G., Zhang, C., Zhang, X., and Jiang, X. (2023). Comparative analysis of chloroplast genomes of endangered heterostylous species Primula wilsonii and its closely related species. Ecol. Evol. 13, e9730. doi: 10.1002/ece3.9730

PubMed Abstract | Crossref Full Text | Google Scholar

Xie, Y., Zhu, X., Ma, Y., Zhao, J., Li, L., and Li, Q. (2017). Natural hybridization and reproductive isolation between two Primula species. J. Integr. Plant Biol. 59, 526–530. doi: 10.1111/jipb.12546

PubMed Abstract | Crossref Full Text | Google Scholar

Xu, L., Yu, R., Lin, X., Zhang, B., Li, N., Lin, K., et al. (2021). Different rates of pollen and seed gene flow cause branch-length and geographic cytonuclear discordance within Asian butternuts. New Phytol. 232, 388–403. doi: 10.1111/nph.17564

PubMed Abstract | Crossref Full Text | Google Scholar

Yan, H., He, C., Peng, C., Hu, C. M., and Hao, G. (2010). Circumscription of Primula subgenus Auganthus (Primulaceae) based on chloroplast DNA sequences. J. Syst. Evol. 48, 123–132. doi: 10.1111/j.1759-6831.2010.00068.x

Crossref Full Text | Google Scholar

Yan, H., Liu, Y., Xie, X., Zhang, C., Hu, C., Hao, G., et al. (2015). DNA barcoding evaluation and its taxonomic implications in the species-rich genus Primula L. @ in China. PloS One 10, e0122903. doi: 10.1371/journal.pone.0122903

PubMed Abstract | Crossref Full Text | Google Scholar

Yang, Z. (2007). PAML 4: phylogenetic analysis by maximum likelihood. Mol. Biol. Evol. 4, 1586–1591. doi: 10.1093/molbev/msm088

PubMed Abstract | Crossref Full Text | Google Scholar

Yang, Z., Liu, Z., Xu., H., Chen, Y., Du, P., Li, P., et al. (2022). The cchromosome-level genome of miracle druit (Synsepalum dulcificum) provides new insights into the evolution and function of miraculin. Front. Plant Sci. 12. doi: 10.3389/fpls.2021.804662

PubMed Abstract | Crossref Full Text | Google Scholar

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. doi: 10.3389/fpls.2025.1550467

PubMed Abstract | Crossref Full Text | Google Scholar

Yura, K. and 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, D., Gao, F., Jakovlić, I., Zou, H., Zhang, J., Li, W., et al. (2020). PhyloSuite: an integrated and scalable desktop platform for streamlined molecular sequence data management and evolutionary phylogenetics studies. Mol. Ecol. Resour. 20, 348–355. doi: 10.1111/1755-0998.13096

PubMed Abstract | Crossref Full Text | Google Scholar

Zhang, H., Meltzer, P., and Davis, S. (2013). RCircos: an R package for Circos 2D track plots. BMC Bioinf. 14, 1–5. doi: 10.1186/1471-2105-14-244

PubMed Abstract | Crossref Full Text | Google Scholar

Zhang, J., Zhou, D., Chen, W., Lin, P., Zhao, S., Wang, M., et al. (2024). Comparison of the chloroplast genomics of nine endangered Habenaria species and phylogenetic analysis. BMC Plant Biol. 24, 1046. doi: 10.1186/s12870-024-05766-2

PubMed Abstract | Crossref Full Text | Google Scholar

Zhang, W., Zhou, J., Li, Z., Wang, L., and Gu, X. (2007). Comparative analysis of codon usage patterns among mitochondrion, chloroplast and nuclear genes in Triticum aestivum L. J. Integr. Plant Biol. 49, 246–254. doi: 10.1111/j.1744-7909.2007.00404.x

Crossref Full Text | Google Scholar

Keywords: Primula mallophylla, mitochondrial genome, chloroplast genome, RNA editing event, adaptive evolution, phylogeny

Citation: Li W, Wang L, Zuo Y, Zhang H, Zhang X, Yang L, Zhang J, Zheng F and Deng H (2025) Organelle genomes reveal adaptive evolution and phylogenetic position of the endangered Primula mallophylla. Front. Plant Sci. 16:1653128. doi: 10.3389/fpls.2025.1653128

Received: 24 June 2025; Accepted: 13 October 2025;
Published: 28 October 2025.

Edited by:

Saraj Bahadur, Hainan University, China

Reviewed by:

Xiangdong Chen, Henan Institute of Science and Technology, China
Xun Wang, Shandong Agricultural University, China

Copyright © 2025 Li, Wang, Zuo, Zhang, Zhang, Yang, Zhang, Zheng and Deng. 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: Hongping Deng, ZGVuZ2hwQHN3dS5lZHUuY24=

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.