IMR Press / FBL / Volume 28 / Issue 10 / DOI: 10.31083/j.fbl2810233
Open Access Original Research
Assembly and Characterization of the Mitochondrial Genome of Fallopia aubertii (L. Henry) Holub
Show Less
1 Qinghai Key Laboratory of Phytochemistry on Qinghai-Tibet Plateau, Qinghai Minzu University, 810007 Xining, Qinghai, China
2 College of Pharmacy, Qinghai Minzu University, 810007 Xining, Qinghai, China
3 Country Qinghai Provincial Key Laboratory of High-value Utilization of Characteristic Economic Plants, Qinghai Minzu University, 810007 Xining, Qinghai, China
*Correspondence: qhlych@126.com (Yong-Chang Lu); wang_jiul@163.com (Jiu-Li Wang)
Front. Biosci. (Landmark Ed) 2023, 28(10), 233; https://doi.org/10.31083/j.fbl2810233
Submitted: 12 June 2023 | Revised: 29 August 2023 | Accepted: 8 September 2023 | Published: 28 September 2023
Copyright: © 2023 The Author(s). Published by IMR Press.
This is an open access article under the CC BY 4.0 license.
Abstract

Background: Fallopia aubertii (L. Henry) Holub is a perennial semi-shrub with both ornamental and medicinal value. The mitochondrial genomes of plants contain valuable genetic traits that can be utilized for the exploitation of genetic resources. The parsing of F. aubertii mitochondrial genome can provide insight into the role of mitochondria in plant growth and development, metabolism regulation, evolution, and response to environmental stress. Methods: In this study, we sequenced the mitochondrial genome of F. aubertii using the Illumina NovaSeq 6000 platform and Nanopore platform. We conducted a comprehensive analysis of the mitochondrial genome of F. aubertii, which involved examining various aspects such as gene composition, repetitive sequences, RNA editing sites, phylogeny, and organelle genome homology. To achieve this, we employed several bioinformatics methods including sequence alignment analysis, repetitive sequence analysis, phylogeny analysis, and more. Results: The mitochondrial genome of F. aubertii has 64 genes, including 34 protein-coding genes (PCGs), three rRNAs, and 27 tRNAs. There were 77 short tandem repeat sequences detected in the mitochondrial genome, five tandem repeat sequences identified by Tandem Repeats Finder (TRF), and 50 scattered repeat sequences observed, including 22 forward repeat sequences and 28 palindrome repeat sequences. A total of 367 RNA coding sites were predicted in PCGs, with the highest number (33) found within ccmB. Ka/Ks values estimated for mitochondrial genes of F. aubertii and three closely related species representing Caryophyllales were less than 1 for most of the genes. The maximum likelihood evolutionary tree showed that F. aubertii and Nepenthes ×ventrata are most closely related. Conclusions: In this study, we obtained basic information on the mitochondrial genome of F. aubertii and this study investigated repeat sequences and homologous segments, predicted RNA editing sites, and utilized the Ka/Ks ratio to estimate the selection pressure on mitochondrial genes of F. aubertii. We also discussed the systematic evolutionary position of F. aubertii based on mitochondrial genome sequences. Our study revealed variations in the sequence and structure of mitochondrial genomes in Caryophyllales. These findings are of great significance for identifying and improving valuable plant traits and serve as a reference for future molecular studies of F. aubertii.

Keywords
Caryophyllales
organellar genomes
comparative genomics
evolution
phylogeny
1. Introduction

The mitochondrial genome, which is circular or linear DNA [1, 2], usually contains dozens of genes that encode proteins and RNA molecules involved in the regulation of mitochondrial function and morphology. The primary function of mitochondria is the production of adenosine triphosphate (ATP), which provides energy to cells [3]. Mitochondrial DNA plays a crucial role in metabolic processes like respiration and photosynthesis, which have a direct or indirect impact on the growth and development of organisms [4]. However, mutations, deletions, or insertions in mitochondrial genes can alter this effect [5, 6, 7]. The parsing of higher plant mitochondrial genes can provide insight into the role of mitochondria in plant growth and development, metabolism regulation, and response to environmental stress [8, 9]. Similarly, the mitochondrial genome contains valuable information that can be utilized for the development of molecular markers, genetic engineering, and elucidation of the phylogenetic and evolutionary connections among plant species [10, 11].

Fallopia aubertii (L. Henry) Holub is a perennial semi-shrub or deciduous vine plant that is widely distributed in Asia, Europe, and North America [12, 13, 14]. It is a heliophilous plant with strong adaptability but is intolerant of shade or water-logging [15], and is commonly found on slopes, riverbanks, forest edges, and beaches. Regarding population ecology, F. aubertii is a relatively dispersed species that is malleable and adaptable to various environments. In recent years, researchers have studied the growth and development, genetic variation, chemical constituents, and pharmacodynamic effects of F. aubertii [14, 16, 17]. Our research team has isolated multiple compounds from F. aubertii, studying its anti-gout efficacy, and sequenced and annotated its chloroplast genome [18, 19, 20].

In this study, we utilize high-throughput sequencing technology and bioinformatics methods to analyze the mitochondrial genome features of F. aubertii. In addition, we reconstruct the phylogenetic relationships among F. aubertii and its closest relatives by using the mitochondrial genome sequence. The current study provides a scientific foundation and a point of reference for the exploitation and development of the species’ resources, as well as the preservation of its biological diversity.

2. Materials and Methods
2.1 Sample Collection and DNA Sequencing

The sample of F. aubertii was collected from the campus of Qinghai Minzu University in Qinghai, China (36.59 N, 101.82 E) and identified by Professor Yong-chang Lu. The fresh young leaves were dried in silica gel and stored at –20 °C. Genomic DNA was extracted from the whole sample of F. aubertii using the modified CTAB method and was subsequently evaluated for purity and integrity using agarose gel electrophoresis and a NanoDrop 2000c spectrophotometer (NanoDrop Technologies, Wilmington, DE, USA). The classified DNA samples were sent to Shanghai Origingene Biotechnology Pharmaceutical Technology Co., Ltd., where the genome DNA was sequenced using the Illumina NovaSeq 6000 platform (BIOZERON Co., Ltd., Shanghai, China) and Nanopore platform (Oxford Nanopore Technologies, Oxford, UK).

2.2 Assembly and Annotation of Mitochondrial Genome

We used GetOrganelle v1.7.1 (Max Planck Institute of Molecular Plant Physiology, Munich, Germany) [21] to perform de novo assembly of the mitochondrial genome. The mitochondrial genome of Bougainvillea spectabilis (GenBank Accession Number: MW167296), a closely related species, was used as the reference sequence. To select the potential mitochondrial reads, we used BLAST searches against the mitogenome of B. spectabilis and the GetOrganelle results from a pool of Illumina reads. The mitochondrial Illumina reads were assembled into 50 mitogenome contigs using SPAdes v3.14.1 (Algorithmic Biology Lab, St. Petersburg Academic University of the Russian Academy of Sciences, St Petersburg, Russia) [22]. Nanopore reads were aligned against the GetOrganelle and SPAdes assembled scaffolds using BWA v0.7.1 (Wellcome Trust Sanger Institute, Wellcome Trust Genome Campus, Cambridge, UK) [23]. The aligned Nanopore reads were extracted to perform de novo assembly of the mitochondrial genome using Canu v2.0, and then used Pilon v1.23 (https://github.com/broadinstitute/pilon/; Broad Institute of MIT and Harvard, Cambridge, MA, USA) [24] for error correction. The genes were predicted based on the method of homology alignment prediction. The coding genes, tRNAs, rRNAs, and possible pseudogenes were annotated by Blast+ 2.7.1 and tRNAscan-SE v2.0.7 (Todd Lowe Lab, Dept. of Biomolecular Engineering, School of Engineering University of California, Santa Cruz, CA, USA) [25]. tRNAscan-SE was also used to predict the secondary structure of tRNAs in F. aubertii. The mitochondrial genome map was drawn using OGDRAW (https://chlorobox.mpimp-golm.mpg.de/OGDraw.html) [26].

2.3 Analysis of Mitochondrial Genome Repeat Sequences

Tandem repeat sequences in the genome were identified by Tandem Repeats Finder (TRF) [27] (https://tandem.bu.edu/trf/trf.html) with the following parameters: match +2, mismatch –7, indel –7. We reported only the repeats with scores greater than 50. Identification of short tandem repeat (STR) or simple sequence repeats (SSR) was performed using MISA v2.1 MISA (Leibniz Institute of Plant Genetics and Crop Plant Research, Stadt Seeland, Germany) [28] with the following parameters:1–10, 2–5, 3–4, 4–3, 5–3, 6–3. The REPuter online version [29] (https://bibiserv.cebitec.uni-bielefeld.de/reputer/) was used to search scattered repeats in the mitochondrial genome with the following parameters: hamming distance-3, maximum computed repeats-50, and minimal repeat size-8.

2.4 Analysis of RNA Editing

To predict the RNA-editing sites within the mitochondrial genes of F. aubertii the online PmtREP (http://www.genepioneer.com/) was used, with a threshold value of 0.2 for the mitochondrial sequences.

2.5 Analysis of GC Content and GC Skew

We applied the cloud platform (http://www.genepioneer.com/) to analyze the GC content of the F. aubertii mitochondrial coding sequence (CDS). To visualize the GC skew, we uploaded the CDS of F. aubertii to the online Proksee [30] tool and selected the “GC Content” and “GC Skew” tabs.

2.6 Selection Pressure Analysis

The ratio of nonsynonymous substitutions (Ka) to synonymous substitutions (Ks) of protein-coding genes (PCGs) between F. aubertii and three closely related species (Nepenthes ×ventrata (MH798871.1), Agrostemma githago (MW553037.1), and Tetragonia tetragonoides (MW971440.1)) was calculated. The homologous protein-coding sequences were obtained by comparing protein-coding sequences among mitochondrial genomes of F. aubertii and its closely related species to find the best match using BLAST [31]. Mafft v7.427 (Research Institute for Microbial Diseases, Suita, Osaka, Japan) [32] (https://mafft.cbrc.jp/alignment/server/) was used for homologous protein sequence alignment, whereas a Perl script was utilized to map the aligned protein sequences to the coding sequence to obtain an aligned coding sequence. Finally, Ka/Ks Calculator 2.0 [33] 96–97 with MLWL model was applied to calculate Ka and Ks values.

2.7 Phylogenetic Analysis

A phylogenetic tree was constructed using the mitochondrial genome sequences of F. aubertii, Malania oleifera (Olacaceae, Santalales), and 15 other species from the order Caryophyllales, with M. oleifera used as an outgroup. We extracted the shared coding sequences from the 17 genomes mentioned above and used BLAST to identify homologous sequences. The resulting sequences were aligned using Mafft, and then concatenated. A trimming percentage of 0.7 was set using the trim function to remove any poorly aligned bases. The maximum likelihood phylogenetic tree was constructed, and the substitution model GTR+F+R3 was selected for the analysis using MEGA 7 [34]. The phylogenetic tree’s step values were determined for each branch by performing 1000 spontaneous replicate analyses.

2.8 Analysis of Homologous Fragments of Chloroplast-Mitochondrial Genomes

A BLAST search was conducted between the mitochondrial and chloroplast genomes of F. aubertii to identify homologous regions. The resulting homologous segments were then compared and analyzed for sequence similarity, number of protein-coding genes, length, and composition of intergenic regions.

3. Results
3.1 Basic Features of the Mitochondrial Genome

The second-generation sequencing platform yielded a data set of 4645.7 Mb of raw reads and 4432 Mb of clean reads, whereas the third-generation sequencing platform generated 21,993 reads with a total size of 67 Mb.

The genome (NCBI accession number: MW664926.1) obtained in this study is a circular DNA molecule with a length of 350,156 bp (Fig. 1), and a GC content of 44.73%. The base composition of the mitochondrial genome is asymmetric, and the coding and non-coding regions’ base composition, size, and proportion are shown in Table 1. The GC content in tRNA and rRNA is relatively high, accounting for 0.59% and 1.47% of the mitochondrial genome sequence, respectively.

Fig. 1.

A circular map of the F. aubertii mt genome. The counterclockwise genes are located on the outer side of the loop, whereas the clockwise genes are located on the inner side of the circle. The diagram shows the names of the contained genes. Colors are used to represent the different functional groups of genes. The grey circles inside represent GC content. The asterisks (*) represent intron-containing genes.

Table 1.Characteristics of F. aubertii mt genome, CDS and Non-coding region.
Feature A (%) T (%) G (%) C (%) GC (%) Size (bp) Proportion in Genome (%)
Genome 27.69 27.58 22.2 22.53 44.73 350,156 100
CDS 26.64 31.39 21.39 20.58 41.97 30,276 8.65
Cis-spliced intron 25.64 23.14 26.94 24.27 51.22 22,868 6.53
tRNA 23.05 26.34 28.28 22.32 50.61 2065 0.59
rRNA 26.18 22.41 28.82 22.59 51.41 5153 1.47
Non-coding region 27.85 27.81 22.08 22.26 44.34 289,955 82.81

Note: CDS: coding sequence.

The F. aubertii mitochondrial genome contains 64 genes, including 34 PCGs, three rRNAs, and 27 tRNAs (Table 2). Out of the 34 protein-coding genes, only five (ccmFC, rps3, nad4, nad5, and nad7) contain introns and were categorized as cis-splice genes (Fig. 2). We examined 27 tRNA genes and determined that they can transport 17 standard amino acids. The length of tRNA ranged from 71 bp to 88 bp, with a length of 1816 bp in total. Among these genes, multicopy genes were found for trnE-UUC, trnH-GUG, and trnM-CAU. Additionally, we predicted the secondary structures of tRNAs (Fig. 3). All genes except for tRNA-Leu, tRNA-Ser, and tRNA-Tyr were predicted to have typical cloverleaf structures, and the base mismatches were mostly G-U. The mitochondrial genome contains three different tRNAs, each with different anticodons that specifically transport serine (Ser). These results are of great significance for further study of the function and stability of tRNA.

Table 2.List of genes annotated in F. aubertii mt genome.
Group of genes Gene name
ATP synthase atp1, atp4, atp6, atp8, atp9
Cytochrome c biogenesis ccmB, ccmC, ccmFC*, ccmFN
Ubiquinol cytochrome c reductase cob
Cytochrome c oxidase cox1, cox2, cox3
Maturases matR
Transport membrane protein mttB
NADH dehydrogenase nad1****, nad2****, nad3, nad4***, nad4L, nad5****, nad6, nad7****, nad9
Ribosomal proteins (LSU) rpl16, rpl5
Ribosomal proteins (SSU) rps1, rps12, rps13, rps14, rps3*, rps4, rps7
Succinate dehydrogenase sdh4
Ribosomal RNAs rrn18, rrn26, rrn5
Transfer RNAs trnC-GCA, trnD-GUC, trnE-CUC, trnE-UUC (2), trnF-GAA, trnG-GCC, trnH-GUG (2), trnI-GAU, trnK-UUU, trnL-CAA, trnL-UAA, trnM-CAU (4), trnN-GUU, trnP-UGG, trnQ-UUG, trnR-ACG, trnS-GCU, trnS-GGA, trnS-UGA, trnV-GAC, trnW-CCA, trnY-GUA

Notes: *: intron number. The number of asterisks (*) refers to the number of introns. The number in brackets next to the gene name indicates the number of copies of multi-copy genes.

Fig. 2.

Schematic map of the cis-splicing genes. Black blocks represent exons, and white blocks represent introns. The numbers beneath the protein-coding genes indicate the locations of introns and exons. One intron is present in both ccmFC and rps3. Three introns are in nad4 and four in nad5 and nad7.

Fig. 3.

tRNA secondary structure prediction. Twenty-four tRNA secondary structures were identified in the mitochondrial genome of F. aubertii. The top left of the structure is the amino acid abbreviation letter. Base mismatches exist, usually U-C base mismatches. The mismatched bases are marked in red boxes.

3.2 Mitochondrial Repetitive Sequences Analysis

Repetitive sequences can serve as genome-specific genetic markers for phylogenetic relationships among species. STRs show a high degree of polymorphism and are commonly used as molecular markers for genetic diversity studies and germplasm characteristics, identification, and selection. In the analysis of STR in the mitochondrial genome of F. aubertii, 77 STR loci were detected. There are 28 single nucleotide motifs, 16 dinucleotide motifs, seven trinucleotide motifs, 22 tetranucleotide motifs, and four pentanucleotide motifs. No hexanucleotide motifs were detected (Table 3). Additionally, five long tandem repeat sequences were detected by TRF software in the mitochondrial genome of F. aubertii (Table 4).

Table 3.Number of short tandem repeats in F. aubertii mt genome.
Motif Number Proportion (%)
Mononucleotide 28 36.4
Dinucleotide 16 20.8
Trinucleotide 7 9.1
Tetranucleotide 22 28.5
Pentanucleotide 4 5.2
Total 77 100
Table 4.Long tandem repeats of F. aubertii mt genome.
Indices Period size of the repeat Copy number Consensus size Percent of matches Percent of indels A C G T Entropy
48,505–48,551 20 2.2 22 81 11 29 25 21 23 1.99
93,390–93,432 18 2.4 18 81 18 34 25 20 18 1.96
119,178–119,223 24 1.9 24 90 0 28 21 21 28 1.99
276,633–276,698 32 2.1 32 94 0 27 9 22 40 1.84
289,717–289,760 21 2.1 21 91 0 36 15 20 27 1.93

Scattered repetitive sequences are crucial in the study of gene mutation, genome origin and evolution, and species formation [35]. In this study, a total of 50 scattered repeats were identified in the mitochondrial genome of F. aubertii, consisting of 22 forward repeats and 28 palindromic repeats. No reverse repeats, or complementary repeats were detected. The longest identified repeat sequence was 5272 bp, whereas most of the repeats were 50–200 bp in length (Fig. 4).

Fig. 4.

Scattered repeats of F. aubertii mt genome. Different types of repeats are represented by different colors, blue represents forward repeats and orange represents palindromic repeats. The height of the colored bar represents the number of sequences. No scattered repeats in the length range of 2000–2999 bp were found.

3.3 Prediction of RNA Editing Sites

The prediction of RNA editing sites in the mitochondrial genome coding genes of F. aubertii resulted in a total of 367 sites in 30 genes (Fig. 5). Among them, ccmB and nad4 contained the majority of RNA editing sites, with ccmB containing most of them (33 sites), accounting for 8.99% of the total RNA editing sites. The atp8 had the fewest RNA editing sites, accounting for only 0.27% of the total number of sites. The results showed that 7.9% of hydrophobic amino acids were converted to hydrophilic amino acids, 49.86% of hydrophilic amino acids were converted into hydrophobic amino acids, and 42.24% of hydrophobic amino acids remained unchanged (Table 5). RNA editing sites were C to T conversion, and the first base of the codon accounted for 32.0% of the editing sites, whereas the second base accounted for 68.0%. Some RNA editing sites may result in the formation of stop codons, but we did not identify such phenomena in the mitochondrial genome of F. aubertii. Furthermore, the proportion of amino acid conversion to leucine after RNA editing was the largest, accounting for 46.59%.

Fig. 5.

Distribution of RNA editing sites in F. aubertii mt protein-coding genes (PCGs). RNA editing sites in 30 coding genes in F. aubertii were predicted. There are no RNA editing sites in cox1, ccmFC, ccmFN, or atp9. The horizontal axis is the protein-coding gene, and the number on the top of the color bar refers to the number of predicted RNA editing sites.

Table 5.Basic statistics on RNA editing sites predicted in F. aubertii mt genes.
Type RNA-editing Number Percentage (%)
Hydrophobic CCA (P) CTA (L) 31 28.07
CCC (P) CTC (L) 3
CCG (P) CTG (L) 26
CCT (P) CTT (L) 18
CCC (P) TTC (F) 6
CCT (P) TTT (F) 2
GCT (A) GTT (V) 1
GCG (A) GTG (V) 2
GCA (A) GTA (V) 2
CTT (L) TTT (F) 9
CTC (L) TTC (F) 3
Hydrophilic CAT (H) TAT (Y) 15 14.17
CAC (H) TAC (Y) 6
CGT (R) TGT (C) 23
CGC (R) TGC (C) 8
Hydrophobic-hydrophilic CCT (P) TCT (S) 14 7.90
CCG (P) TCG (S) 6
CCC (P) TCC (S) 4
CCA (P) TCA (S) 5
Hydrophilic-hydrophobic CGG (R) TGG (W) 19 49.86
TCT (S) TTT (F) 28
TCC (S) TTC (F) 25
TCG (S) TTG (L) 37
TCA (S) TTA (L) 56
ACT (T) ATT (I) 4
ACC (T) ATC (I) 6
ACG (T) ATG (M) 8
3.4 Mitochondrial Genome Size and GC Content of F. aubertii Compared with Other Species

The CDS of F. aubertii were analyzed to determine the GC content of each gene and calculate the average GC content (Table 6). The nad5 gene was the longest, possibly due to its introns. The GC content varied across PCGs, with matR gene having the highest GC content (51.5%) and the GC content of the nad4L being the lowest (34.65%). The AT skew average value was negative, and 13 genes had positive AT skew. Visual analysis of GC skew showed that 22 genes with positive GC skew had higher G-base content than C-base content, and 12 genes with negative GC skew and high C-base content (Fig. 6). Caryophyllales plants, including F. aubertii, had a mitochondrial genome size range of 247 Kb–509 Kb, with an average GC content of about 44% (Table 7).

Fig. 6.

GC skew of F. aubertii mt CDS. The outermost circle represents the coding area of a gene, the innermost circle represents the variation in GC content across the genome, and the innermost circle represents the GC skew, where green represents skew+ and purple represents skew-.

Table 6.GC content of F. aubertii mt PCGs.
Genes Length GC1 GC2 GC3 GC all AT skew GC skew
cox1 1611 46 41.15 36.87 41.34 –0.17 0.02
nad4L 303 37.62 36.63 29.7 34.65 –0.24 0.1
nad9 573 49.74 40.31 31.41 40.49 0.01 0.07
nad2 1467 40.9 42.74 32.72 38.79 –0.21 –0.05
ccmFC 1344 48.21 43.53 42.63 44.79 –0.07 0.01
ccmFN 1740 49.66 46.55 40.52 45.57 –0.16 –0.01
cox2 783 51.72 36.4 30.65 39.59 –0.08 0.03
sdh4 420 42.14 34.29 42.14 39.52 –0.09 –0.02
cox3 798 50.75 41.73 33.08 41.85 –0.26 0.07
atp8 483 35.4 37.27 41.61 38.1 –0.02 0.03
nad1 978 53.99 41.1 28.83 41.31 –0.17 0.03
rpl16 480 49.38 53.12 31.87 44.79 0.09 0.18
rps3 1686 41.46 40.57 40.93 40.98 0.12 0.11
rps12 378 53.97 48.41 30.95 44.44 0.21 0.11
nad3 357 42.02 42.86 35.29 40.06 –0.3 –0.01
ccmB 621 44.44 44.44 33.82 40.9 –0.25 –0.17
rps13 351 49.57 39.32 26.5 38.46 0.15 0.14
nad4 1488 44.76 40.73 35.08 40.19 –0.19 –0.03
rps7 456 48.03 38.16 36.18 40.79 0.12 0.2
mttB 786 42.37 40.46 42.75 41.86 –0.2 –0.04
cob 1182 49.49 41.12 30.96 40.52 –0.22 0.03
rps14 303 42.57 46.53 33.66 40.92 –0.02 0.08
rpl5 561 50.27 37.97 40.64 42.96 0.1 0.02
nad5 2013 43.37 43.07 33.68 40.04 –0.23 –0.01
atp6 801 44.19 35.21 30.34 36.58 –0.18 –0.06
rps1 612 44.61 39.71 42.65 42.32 0.06 0.14
atp1 1524 57.28 41.54 33.66 44.16 0.06 0.07
atp9 225 48 42.67 37.33 42.67 –0.22 0.08
matR 1965 52.67 44.43 57.4 51.5 0.13 –0.03
atp4 597 47.24 41.71 36.18 41.71 0.02 0.03
ccmC 753 46.22 44.62 34.26 41.7 –0.21 –0.1
rps4 828 41.3 38.77 34.78 38.29 0.14 –0.05
nad6 624 41.83 37.02 35.58 38.14 –0.15 0.08
nad7 1185 55.7 43.8 32.66 44.05 0.02 0.05
All 30,276 47.24 41.78 36.9 41.97 –0.08 0.02
Table 7.Mitochondrial genome and GC content in F. aubertii and 15 species of Caryophyllales.
Species Accession number Category Genome size GC
Fallopia aubertii MW664926.1 Polygonaceae 350,156 44.73
Nepenthes ×ventrata MH798871.1 Nepenthaceae 520,764 44.17
Alternanthera philoxeroides MN166292.1 Amaranthaceae 283,258 44.07
Spinacia oleracea NC_035618.1 Amaranthaceae 329,613 43.41
Beta macrocarpa NC_015994.1 Amaranthaceae 385,220 43.89
Beta vulgaris NC_015099.1 Amaranthaceae 364,950 43.91
Chenopodium quinoa MK182703.1 Amaranthaceae 315,003 43.83
Suaeda glauca MW561632.1 Amaranthaceae 474,330 44.07
Bougainvillea spectabilis MW167296.1 Nyctaginaceae 343,746 44.06
Mirabilis jalapa NC_056991.1 Nyctaginaceae 267,334 44.75
Mirabilis himalaica NC_048974.1 Nyctaginaceae 346,363 44.65
Agrostemma githago MW553037.1 Caryophyllaceae 262,903 44.66
Silene latifolia HM562727.1 Caryophyllaceae 253,413 42.56
Pereskia aculeata ON496936.1 Cactaceae 515,187 44.05
Sesuvium portulacastrum MN683736.1 Aizoaceae 392,221 44.16
Tetragonia tetragonoides MW971440.1 Aizoaceae 347,227 43.84
3.5 Ka, Ks Analysis

Comparing the Ka and Ks in homologous genes during evolution is essential for the study of gene functions and evolutionary relationships across species, as well as for exploring issues such as adaptation and genetic diversity of plant communities. The Ka/Ks ratios of 34 PCGs in the mitochondrial genomes of F. aubertii, N. ventrata, A. githago, and T. tetragonoides were analyzed (Fig. 7). Among these 34 genes, the rsp16 gene had a zero Ka/Ks value. Overall, four of the 34 genes had Ka/Ks values greater than one, indicating that these four genes were under positive selective pressure during evolution, which included nad4. rps13, rps1 and ccmFN. The remaining genes had values of Ka/Ks less than 1, which implies purifying selection and relative conservation.

Fig. 7.

Ka/Ks values of the PCGs. The mitochondrial protein-encoding genes of F. aubertii were compared with N. ventrata (MH798871.1), A. githago (MW553037.1), and T. tetragonoides (MW971440.1), respectively, for Ka/Ks analysis. The three closely related species contain protein-coding genes different from those in F. aubertii, so the Ka/Ks values of some genes in the figure are 0.

3.6 Phylogenetic Analysis

To better understand the phylogenetic position of F. aubertii, 16 plant mitogenomes from the NCBI database were downloaded (Table 7). The phylogenetic tree showed that most branch nodes had high support values above 99%, and species from the same family were grouped together, indicating high result reliability (Fig. 8). The phylogenetic tree also strongly demonstrated that F. aubertii and the N. ventrata clustered into one clade with a 92% bootstrap value (Fig. 7). Caryophyllales plants were further divided into two subgroups [36, 37]. The F. aubertii and the N. ventrata were non-core taxa of Caryophyllales, and other species belonged to the core group of Caryophyllales.

Fig. 8.

A phylogenetic tree of 17 species. Reconstruction of phylogenetic relationships was based on the shared CDS using the maximum likelihood model with M. oleifera (Olacaceae) as an outgroup. F. aubertii is indicated using the star icon. The plants grouped together are marked according to the family.

3.7 Homologous Fragment Analysis of F. aubertii Chloroplast and Mitochondrial Genomes

By analyzing the homology of chloroplast and mitochondrial genomes, we can deeply explore their evolutionary history and their roles in evolution. Fifty-eight homologous fragments were observed in the organellar genome of F. aubertii, with a total length of 47,757 bp, accounting for 29.2% and 13.6% of the chloroplast and mitochondrial genomes, respectively. There were 14 homologous genes in the mitochondrial genome, including trnW-CCA, trnV-GAC, trnS-GGA, trnR-ACG, trnN-GUU, trnM-CAU, trnL-CAA, trnI-GAU, trnH-GUG, trnD-GUC, trnA-UGC, rrn26, rrn18, and ccmC(Table 8).

Table 8.Homologous regions of organellar genomes of F. aubertii.
Sequence Chloroplast genome Mitochondrial genome Identity
Gene Sequence position (bp) Gene Sequence position (bp)
1 trnR-ACG, rrn5, rrn4.5, rrn23, trnA-UGC, trnI-GAU, rrn16, trnV-GAC, rps7 138,146–149,356 trnR-ACG, trnA-UGC, trnI-GAU, trnV-GAC 275,594–286,804 100
2 rps7, trnV-GAC, rrn16, trnI-GAU, trnA-UGC, rrn23, rrn4.5, rrn5, trnR-ACG 100,317–111,527 trnR-ACG, trnA-UGC, trnI-GAU, trnV-GAC 286,804–275,594 100
3 ycf2, trnL-CAA, ndhB 90,089–98,017 trnL-CAA 286,794–294,720 99
4 trnL-CAA, ycf2, ndhB 151,656–159,584 trnL-CAA 294,720–286,794 99
5 rps19, rpl2, rpl23, trnM-CAU 87,280–89,471 trnM-CAU, ccmC 303,457–305,648 100
6 trnI-CAU, rpl23, rpl2 160,202–162,393 trnM-CAU, ccmC 305,648–303,457 100
7 psbA 206–937 / 15,690–14,968 94
8 trnH-GUG 1–217 trnH-GUG 303,456–303,240 100
9 rbcL 58,741–58,955 / 241,189–240,975 97
10 trnW-CCA 68,795–68,976 trnW-CCA 44,932–44,752 90
11 rrn23 139,746–139,869 / 109,167–109,290 95
12 rrn23 109,804–109,927 / 109,290–109,167 95
13 trnD-GUC 31,665–31,802 trnD-GUC 171,543–171,407 92
14 trnS-GGA 47,519–47,611 trnS-GGA 261,692–261,784 96
15 trnH-GUG 2–78 trnH-GUG 257,385–257,461 100
16 rrn16 104,286–104,394 rrn18 13,207–13,315 92
17 rrn16 145,279–145,387 rrn18 13,315–13,207 92
18 petG 68,653–68,737 / 45,085–45,001 95
19 rrn23 139,686–139,933 rrn26 92,350–92,597 81
20 rrn23 109,740–109,987 rrn27 92,597–92,350 81
21 trnN-GUU 111,672–111,750 trnN-GUU 57,454–57,532 96
22 trnN-GUU 137,923–138,001 trnN-GUU 57,532–574,54 96
23 / 205–263 / 303,197–303,255 100
24 trnM-CAU 54,431–54,505 trnM-CAU 200,075–200,149 94
25 rrn16 103,420–103,478 / 2162–2220 100
26 rrn16 146,195–146,253 / 2220–2162 100
27 rrn16 103,730–103,815 rrn18 12,655–12,740 91
28 rrn16 145,858–145,943 rrn18 12,740–12,655 91
29 rrn23 140,317–140,410 rrn26 92,965–93,058 87
30 rrn23 109,263–109,356 rrn26 93,058–92,965 87
31 psbD, psbC 35,865–35,965 / 256,992–257,092 86
32 trnP-UGG 69,203–69,251 / 44,659–44,611 97
33 ycf3 46,462–46,537 / 260,822–260,897 88
34 ndhA 128,336–128,377 / 285,575–285,534 97
35 rrn16 104,701–104,818 rrn18 13,950–14,066 83
36 rrn16 144,855–144,972 rrn18 14,066–13,950 83
37 rrn16 104,580–104,649 rrn18 13,830–13,899 88
38 rrn16 145,024–145,093 rrn18 13,899–13,830 88
39 rrn23 140,568–140,612 / 93,214–93,258 95
40 rrn23 109,061–109,105 / 93,258–93,214 95
41 rpl2 87,768–87,803 / 101,974–102,009 100
42 rpl2 161,870–161,905 / 102,009–101,974 100
43 rrn16 103,910–103,981 rrn18 12,837–12,908 87
44 rrn16 145,692–145,763 rrn18 12,908–12,837 87
45 ycf3 45,725–45,763 / 285,573–285,535 97
46 / 46,265–46,315 / 260,632–260,682 92
47 rrn16 104,451–104,525 rrn18 13,371–13,445 86
48 rrn16 145,148–145,222 rrn18 13,445–13,371 86
49 psbA 248–285 / 303,209–303,272 97
50 / 68,036–68,103 / 19,391–19,458 86
51 rrn23 141,817–141,862 rrn26 94,788–94,833 91
52 rrn23 107,811–107,856 rrn26 94,833–94,788 91
53 rrn16 103,467–103,523 rrn18 12,351–12,407 87
54 rrn16 146,150–146,206 rrn18 12,407–12,351 87
55 trnS-GCU 8666–8693 trnS-GGA 261,781–261,754 100
56 rpoC1 22,600–22,627 / 110,671–110,698 100
57 rrn16 104,889–104,920 rrn18 14,184–14,215 96
58 rrn16 144,753–144,784 rrn18 14,215–14,184 96

Note: “/” indicates no gene at the locus.

4. Discussion

Analyzing mitochondrial genes can lead to the optimization of plant growth characteristics at the genetic level, thus increasing yield and enhancing plant adaptation to the environment, which is essential for agricultural production and food safety [38, 39, 40, 41]. The study of genome size and GC content is crucial for understanding plant evolution and adaptability, as they greatly influence traits such as morphology, physiology, and biochemistry of plants [42]. The size of angiosperm mitochondrial genomes is usually 200–800 Kb [43], and the entire length of F. aubertii mitochondrial genome is 350,156 bp with a GC content of 44.73%, which is similar to other Caryophyllales plants [44, 45]. Research on genome size and GC content can provide important insights into plants’ genetic characteristics and biological functions [46] and form the scientific basis for applications such as plant breeding and biotechnology [47, 48].

Repeat sequences are essential for gene expression and regulation [49]. They are the main driving force for gene diversification and evolution [50]—the present study identified STRs, long tandem repeats, and scattered repeats. A total of 77 STR loci, five long tandem repeats, and 50 scattered repeat sequences were detected by analyzing the F. aubertii mitochondrial genome repeats. Repetitive sequences play an important role in the processes of insertion, deletion, and other rearrangements that may occur within the mitochondrial genome [51]. Mitochondrial repetitive sequences have been widely used in plant genetic improvement and molecular marker-assisted breeding [52, 53]. Analysis of the mitochondrial repeat sequences of F. aubertii has helped us to gain insights into the organelle sequences, genetic diversity, and other relevant features of the plant, which are essential for species conservation and germplasm identification [54].

RNA editing is a type of genetic variation event in RNA post-transcriptional modification, which is crucial for maintaining the stability of gene expression [55, 56]. Compared with DNA editing, RNA editing can respond more quickly to external stimuli or environmental changes, thereby adapting to the needs of survivors [57]. The diversity and complexity of RNA function, as well as its impact on gene expression and regulation, can be better understood by exploring the location and type of RNA editing, leading to further study into RNA function and regulatory mechanisms [58]. In this study, the RNA editing sites in the mitochondrial genome of F. aubertii were mainly C to T conversions, with the second base of the codon being the most frequently changed. The proportion of amino acid conversion to leucine was the largest, accounting for 46.59%. Leucine is the main amino acid produced after editing, similar to RNA editing results in higher angiosperms [59].

To understand the various selection pressures that took place during the evolution of the gene and to determine whether a gene has undergone rapid evolution, Ka/Ks analysis can be used [60]. In addition to predicting protein structure and function, Ka/Ks analysis can be used to research the gene families’ functional diversity and evolutionary pathways [61]. When the value of Ka/Ks >1, it indicates a positive selection pressure during the gene’s evolution. Conversely, Ka/Ks <1 denotes the negative selection. Most of the genes in F. aubertii are conserved. The four genes rps1, nad4, rps13, and ccmFN all have Ka/Ks >1, which is significant for understanding the positive selection pressure these four genes experienced during evolution. We referred to the literature on the mitochondrial genome of order Caryophyllales [62] and found that the phenomenon of four genes with Ka/Ks >1 is unique to F. aubertii. Nevertheless, there are less studies concerning the selective pressure on the mitochondrial genome of Caryophyllales and our findings need to be confirmed by further studies. The positive selection of the above genes would contribute to a faster rate of evolution in organisms and would also increase the adaptive divergence of mitochondrial genes [63, 64].

In this study, a maximum likelihood phylogenetic tree was constructed based on the shared mitochondrial CDSs. The F. aubertii and the N. ventrata formed a sister clade indicating a closer relationship between them, and agreed with previous research [62, 65]. In the evolution of the chloroplast genome system in Caryophyllales, Polygonaceae plants are also closely related to Nepenthaceae plants [66, 67, 68]. Our study provides a scientific basis for further studies on the phylogeny of the Caryophyllales.

There is a horizontal transfer of genes between organelle genomes [69]. We made a comparison of the mitochondrial and chloroplast genomes of F. aubertii. Some genes were found to be similar, and these genes were highly homologous. Between the chloroplast and mitochondrial genomes of F. aubertii, there were 58 homologous segments and 14 homologous genes, including 11 tRNAs, two ribosomal protein genes, and a gene for cytochrome c biogenesis (ccmC). In angiosperms, tRNA genes are frequently transferred from the chloroplast genome to the mitochondrial genome [70, 71]. The homology of chloroplast and mitochondrial genomes is an important research direction in biology and evolution. By studying the functional activity of migrating genes, we have gained a greater understanding of species evolution [72].

5. Conclusions

In this study, we provided a comprehensive analysis of the mitochondrial genome of F. aubertii. The genome was found to have a circular structure with a length of 350,156 bp and a GC content of 44.73%. The genome contains 64 genes, including 34 protein-coding genes, three rRNAs, and 27 tRNAs. Analyses of repeat sequences revealed the presence of short tandem repeats, long tandem repeats, and scattered repeats in the mitochondrial genome of F. aubertii. RNA editing sites were predicted, and the majority of editing events involved the conversion of C to T, with leucine being the most frequently converted amino acid. Analysis of the Ka/Ks ratios indicated that most mitochondrial genes in F. aubertii have undergone negative selection, whereas four genes (rps1, nad4, rps13, and ccmFN) showed evidence of positive selection. Phylogenetic analysis revealed that F. aubertii and Nepenthes × ventrata are closely related, supporting previous studies on the evolutionary relationships within the order Caryophyllales. Additionally, a comparison of the mitochondrial and chloroplast genomes showed the presence of homologous regions and genes, suggesting potential gene transfer events between these organellar genomes.

This study provides valuable information on the mitochondrial genome of F. aubertii and serves as a reference for future molecular studies and the exploitation of genetic resources in this species. The findings contribute to the identification and improvement of valuable plant traits and highlight the significance of mitochondrial genomes in plant evolution and adaptation.

Availability of Data and Materials

The datasets used and/or analyzed during the current study can be obtained from the corresponding authors upon reasonable request. The raw data for the mitochondrial genomes sequencing is available on the website https://www.ncbi.nlm.nih.gov/sra/ under the reference number SRX10859887. Additionally, the mitochondrial genome itself is available on the website https://www.ncbi.nlm.nih.gov/genbank/ under the reference number MW167296.

Author Contributions

XZ performed the research, analyzed the data, and wrote the manuscript; JLW and YCL conceptualized and designed the study and reviewed the manuscript. All authors contributed to editorial changes in the manuscript. All authors read and approved the final manuscript. All authors have participated sufficiently in the work and agreed to be accountable for all aspects of the work.

Ethics Approval and Consent to Participate

Not applicable.

Acknowledgment

Not applicable.

Funding

This research was funded by Natural Science Foundation of Qinghai Province Science and Technology Department (Grant No. 2019-ZJ-915) and Foundation of Qinghai Minzu University (Grant No. 2022-JYQN-002, 2021XJG17).

Conflict of Interest

The authors declare no conflict of interest.

References
[1]
Yurina NP, Odintsova MS. Mitochondrial Genome Structure of Photosynthetic Eukaryotes. Biochemistry. Biokhimiia. 2016; 81: 101–113.
[2]
Lohse M, Drechsel O, Kahlau S, Bock R. OrganellarGenomeDRAW–a suite of tools for generating physical maps of plastid and mitochondrial genomes and visualizing expression data sets. Nucleic Acids Research. 2013; 41: W575–W581.
[3]
He L, Maheshwari A. Mitochondria in Early Life. Current Pediatric Reviews. 2023; 19: 395–416.
[4]
Knoop V. The mitochondrial DNA of land plants: peculiarities in phylogenetic perspective. Current Genetics. 2004; 46: 123–139.
[5]
Wallace DC. Mitochondrial diseases in man and mouse. Science (New York, N.Y.). 1999; 283: 1482–1488.
[6]
Xia M, Zhang Y, Jin K, Lu Z, Zeng Z, Xiong W. Communication between mitochondria and other organelles: a brand-new perspective on mitochondria in cancer. Cell & Bioscience. 2019; 9: 27.
[7]
Bohovych I, Khalimonchuk O. Sending Out an SOS: Mitochondria as a Signaling Hub. Frontiers in Cell and Developmental Biology. 2016; 4: 109.
[8]
Ng S, De Clercq I, Van Aken O, Law SR, Ivanova A, Willems P, et al. Anterograde and retrograde regulation of nuclear genes encoding mitochondrial proteins during growth, development, and stress. Molecular Plant. 2014; 7: 1075–1093.
[9]
Liberatore KL, Dukowic-Schulze S, Miller ME, Chen C, Kianian SF. The role of mitochondria in plant development and stress tolerance. Free Radical Biology & Medicine. 2016; 100: 238–256.
[10]
Rogalski M, do Nascimento Vieira L, Fraga HP, Guerra MP. Plastid genomics in horticultural species: importance and applications for plant population genetics, evolution, and biotechnology. Frontiers in Plant Science. 2015; 6: 586.
[11]
Wu CC, Chang SH, Tung CW, Ho CK, Gogorcena Y, Chu FH. Identification of hybridization and introgression between Cinnamomum kanehirae Hayata and C. camphora (L.) Presl using genotyping-by-sequencing. Scientific Reports. 2020; 10: 15995.
[12]
Jakubska-Busse A, Dziadas M, Gruss I, Kobyłka MJ. Floral Volatile Organic Compounds and a List of Pollinators of Fallopia baldschuanica (Polygonaceae). Insects. 2022; 13: 904.
[13]
Schuster TM, Reveal JL, Bayly MJ, Kron KA. An updated molecular phylogeny of Polygonoideae (Polygonaceae): Relationships of Oxygonum, Pteroxygonum, and Rumex, and a new circumscription of Koenigia. Taxon. 2015; 64: 1188–1208.
[14]
Wang JF, Ma GX, Hu JZ. Phytochemical and chemotaxonomic study on Fallopia aubertii (L. Henry) Holub. Biochemical Systematics and Ecology. 2019; 85: 50–53.
[15]
Schuster TM, Wilson KL, Kron KA. Phylogenetic relationships of Muehlenbeckia, Fallopia, and Reynoutria (Polygonaceae) investigated with chloroplast and nuclear sequence data. International Journal of Plant Sciences. 2011; 172: 1053–1066.
[16]
Olaru OT, Venables L, VAN DE Venter M, Nitulescu GM, Margina D, Spandidos DA, et al. Anticancer potential of selected Fallopia Adans species. Oncology Letters. 2015; 10: 1323–1332.
[17]
Zhao LQ, Zhang XP, Zhang CF. Separation and ldentification of Chemical Constituents in Fallopia auberti Stems and Leaves. Food Science. 2012; 33: 1–5. (In Chinese)
[18]
Ji D, Wang Q, Wang H, Ma Q, Wang M, Lu Y. Preparative separation of gallic acid from Fallopia aubertii using middle-pressure chromatogram isolated gel coupled with reversed-phase chromatography with hydrophilic groups. RSC Advances. 2021; 11: 27276–27282.
[19]
Lu Y, Wang W, Yu Y, Ma Q, Wang H. The chloroplast genome of Fallopia aubertii (Polygonaceae) from Xining, China. Mitochondrial DNA. Part B, Resources. 2021; 6: 3458–3460.
[20]
Ma Q, Zhao PC, Lu YC. Study on Inhibition of Xanthine Oxidase Activty in Different Extract Parts of Fallopia auberti L. Henry Holub. Chinese Journal of Ethnomedicine and Ethnopharmacy. 2020; 29: 16–20. (In Chinese)
[21]
Jin JJ, Yu WB, Yang JB, Song Y, dePamphilis CW, Yi TS, et al. GetOrganelle: a fast and versatile toolkit for accurate de novo assembly of organelle genomes. Genome Biology. 2020; 21: 241.
[22]
Bankevich A, Nurk S, Antipov D, Gurevich AA, Dvorkin M, Kulikov AS, et al. SPAdes: a new genome assembly algorithm and its applications to single-cell sequencing. Journal of Computational Biology: a Journal of Computational Molecular Cell Biology. 2012; 19: 455–477.
[23]
Li H, Durbin R. Fast and accurate short read alignment with Burrows-Wheeler transform. Bioinformatics (Oxford, England). 2009; 25: 1754–1760.
[24]
Walker BJ, Abeel T, Shea T, Priest M, Abouelliel A, Sakthikumar S, et al. Pilon: an integrated tool for comprehensive microbial variant detection and genome assembly improvement. PloS One. 2014; 9: e112963.
[25]
Chan PP, Lin BY, Mak AJ, Lowe TM. tRNAscan-SE 2.0: improved detection and functional classification of transfer RNA genes. Nucleic Acids Research. 2021; 49: 9077–9096.
[26]
Greiner S, Lehwark P, Bock R. OrganellarGenomeDRAW (OGDRAW) version 1.3.1: expanded toolkit for the graphical visualization of organellar genomes. Nucleic Acids Research. 2019; 47: W59–W64.
[27]
Benson G. Tandem repeats finder: a program to analyze DNA sequences. Nucleic Acids Research. 1999; 27: 573–580.
[28]
Beier S, Thiel T, Münch T, Scholz U, Mascher M. MISA-web: a web server for microsatellite prediction. Bioinformatics (Oxford, England). 2017; 33: 2583–2585.
[29]
Kurtz S, Choudhuri JV, Ohlebusch E, Schleiermacher C, Stoye J, Giegerich R. REPuter: the manifold applications of repeat analysis on a genomic scale. Nucleic Acids Research. 2001; 29: 4633–4642.
[30]
Grant JR, Enns E, Marinier E, Mandal A, Herman EK, Chen CY, et al. Proksee: in-depth characterization and visualization of bacterial genomes. Nucleic Acids Research. 2023; 51: W484–W492.
[31]
Hu G, Kurgan L. Sequence Similarity Searching. Current Protocols in Protein Science. 2019; 95: e71.
[32]
Katoh K, Rozewicki J, Yamada KD. MAFFT online service: multiple sequence alignment, interactive sequence choice and visualization. Briefings in Bioinformatics. 2019; 20: 1160–1166.
[33]
Wang D, Zhang Y, Zhang Z, Zhu J, Yu J. KaKs_Calculator 2.0: a toolkit incorporating gamma-series methods and sliding window strategies. Genomics, Proteomics & Bioinformatics. 2010; 8: 77–80.
[34]
Kumar S, Stecher G, Tamura K. MEGA7: Molecular Evolutionary Genetics Analysis Version 7.0 for Bigger Datasets. Molecular Biology and Evolution. 2016; 33: 1870–1874.
[35]
Mehrotra S, Goyal V. Repetitive sequences in plant nuclear DNA: types, distribution, evolution and function. Genomics, Proteomics & Bioinformatics. 2014; 12: 164–171.
[36]
Zhang X, Shan Y, Li J, Qin Q, Yu J, Deng H. Assembly of the Complete Mitochondrial Genome of Pereskia aculeata Revealed That Two Pairs of Repetitive Elements Mediated the Recombination of the Genome. International Journal of Molecular Sciences. 2023; 24: 8366.
[37]
Group TAP, Chase MW, Christenhusz MJM, Fay MF, Byng JW, Judd WS, et al. An update of the Angiosperm Phylogeny Group classification for the orders and families of flowering plants: APG IV. Botanical Journal of the Linnean Society. 2016; 181: 1–20.
[38]
Silva-Pinheiro P, Minczuk M. The potential of mitochondrial genome engineering. Nature Reviews. Genetics. 2022; 23: 199–214.
[39]
Zhang X, Liu S, Takano T. Overexpression of a mitochondrial ATP synthase small subunit gene (AtMtATP6) confers tolerance to several abiotic stresses in Saccharomyces cerevisiae and Arabidopsis thaliana. Biotechnology Letters. 2008; 30: 1289–1294.
[40]
Moghadam AA, Ebrahimie E, Taghavi SM, Niazi A, Babgohari MZ, Deihimi T, et al. How the nucleus and mitochondria communicate in energy production during stress: nuclear MtATP6, an early-stress responsive gene, regulates the mitochondrial F₁F₀-ATP synthase complex. Molecular Biotechnology. 2013; 54: 756–769.
[41]
Zhang X, Takano T, Liu S. Identification of a mitochondrial ATP synthase small subunit gene (RMtATP6) expressed in response to salts and osmotic stresses in rice (Oryza sativa L.). Journal of Experimental Botany. 2006; 57: 193–200.
[42]
Trávníček P, Čertner M, Ponert J, Chumová Z, Jersáková J, Suda J. Diversity in genome size and GC content shows adaptive potential in orchids and is closely linked to partial endoreplication, plant life-history traits and climatic conditions. The New Phytologist. 2019; 224: 1642–1656.
[43]
Li X, Zhe M, Huang Y, Fan W, Yang J, Zhu A. The Evolution of Mitochondrial Genomes between Two Cymbidium Sister Species: Dozens of Circular Chromosomes and the Maintenance and Deterioration of Genome Synteny. Genes. 2023; 14: 864.
[44]
Kim CK, Kim YK. The multipartite mitochondrial genome of Fallopia multiflora (Caryophyllales: Polygonaceae). Mitochondrial DNA. Part B, Resources. 2018; 3: 155–156.
[45]
Yuan F, Lan X. Sequencing the organelle genomes of Bougainvillea spectabilis and Mirabilis jalapa (Nyctaginaceae). BMC Genomic Data. 2022; 23: 28.
[46]
Bowers JE, Tang H, Burke JM, Paterson AH. GC content of plant genes is linked to past gene duplications. PloS One. 2022; 17: e0261748.
[47]
Sliwinska E, Loureiro J, Leitch IJ, Šmarda P, Bainard J, Bureš P, et al. Application-based guidelines for best practices in plant flow cytometry. Cytometry. Part a: the Journal of the International Society for Analytical Cytology. 2022; 101: 749–781.
[48]
Chaudhary J, Patil GB, Sonah H, Deshmukh RK, Vuong TD, Valliyodan B, et al. Expanding Omics Resources for Improvement of Soybean Seed Composition Traits. Frontiers in Plant Science. 2015; 6: 1021.
[49]
Shapiro JA, von Sternberg R. Why repetitive DNA is essential to genome function. Biological Reviews of the Cambridge Philosophical Society. 2005; 80: 227–250.
[50]
Lynch M, Conery JS. The evolutionary fate and consequences of duplicate genes. Science (New York, N.Y.). 2000; 290: 1151–1155.
[51]
Nardi F, Carapelli A, Frati F. Repeated regions in mitochondrial genomes: distribution, origin and evolutionary significance. Mitochondrion. 2012; 12: 483–491.
[52]
Baloch FS, Altaf MT, Liaqat W, Bedir M, Nadeem MA, Cömertpay G, et al. Recent advancements in the breeding of sorghum crop: current status and future strategies for marker-assisted breeding. Frontiers in Genetics. 2023; 14: 1150616.
[53]
Amiteye S. Basic concepts and methodologies of DNA marker systems in plant molecular breeding. Heliyon. 2021; 7: e08093.
[54]
Jia HM, Jiao Y, Wang GY, Li YH, Jia HJ, Wu HX, et al. Genetic diversity of male and female Chinese bayberry (Myrica rubra) populations and identification of sex-associated markers. BMC Genomics. 2015; 16: 394.
[55]
Yan J, Zhang Q, Yin P. RNA editing machinery in plant organelles. Science China. Life Sciences. 2018; 61: 162–169.
[56]
Simpson L, Emeson RB. RNA editing. Annual Review of Neuroscience. 1996; 19: 27–52.
[57]
Sheridan C. Shoot the messenger: RNA editing is here. Nature Biotechnology. 2023; 41: 306–308.
[58]
Hao W, Liu G, Wang W, Shen W, Zhao Y, Sun J, et al. RNA Editing and Its Roles in Plant Organelles. Frontiers in Genetics. 2021; 12: 757109.
[59]
Suzuki H, Yu J, Ness SA, O’Connell MA, Zhang J. RNA editing events in mitochondrial genes by ultra-deep sequencing methods: a comparison of cytoplasmic male sterile, fertile and restored genotypes in cotton. Molecular Genetics and Genomics: MGG. 2013; 288: 445–457.
[60]
Browning SR, Browning BL. Haplotype phasing: existing methods and new developments. Nature Reviews. Genetics. 2011; 12: 703–714.
[61]
Bellavista E, Santoro A, Galimberti D, Comi C, Luciani F, Mishto M. Current understanding on the role of standard and immunoproteasomes in inflammatory/immunological pathways of multiple sclerosis. Autoimmune Diseases. 2014; 2014: 739705.
[62]
Androsiuk P, Paukszto Ł, Jastrzębski JP, Milarska SE, Okorski A, Pszczółkowska A. Molecular Diversity and Phylogeny Reconstruction of Genus Colobanthus (Caryophyllaceae) Based on Mitochondrial Gene Sequences. Genes. 2022; 13: 1060.
[63]
Fourdrilis S, de Frias Martins AM, Backeljau T. Relation between mitochondrial DNA hyperdiversity, mutation rate and mitochondrial genome evolution in Melarhaphe neritoides (Gastropoda: Littorinidae) and other Caenogastropoda. Scientific Reports. 2018; 8: 17964.
[64]
Liu SL, Adams K. Molecular adaptation and expression evolution following duplication of genes for organellar ribosomal protein S13 in rosids. BMC Evolutionary Biology. 2008; 8: 25.
[65]
Zhu ZX, Wang JH, Chen CR, Zhao KK, Wang HF. Complete plastome sequence of Nepenthes mirabilis (Nepenthaceae): a ”vulnerable” herb in China. Mitochondrial DNA. Part B, Resources. 2018; 3: 732–733.
[66]
Choi KS, Kwak M, Lee B, Park S. Complete chloroplast genome of Tetragonia tetragonioides: Molecular phylogenetic relationships and evolution in Caryophyllales. PloS One. 2018; 13: e0199626.
[67]
Yang Y, Moore MJ, Brockington SF, Soltis DE, Wong GKS, Carpenter EJ, et al. Dissecting Molecular Evolution in the Highly Diverse Plant Clade Caryophyllales Using Transcriptome Sequencing. Molecular Biology and Evolution. 2015; 32: 2001–2014.
[68]
Yao G, Jin JJ, Li HT, Yang JB, Mandala VS, Croley M, et al. Plastid phylogenomic insights into the evolution of Caryophyllales. Molecular Phylogenetics and Evolution. 2019; 134: 74–86.
[69]
Daniell H, Lin CS, Yu M, Chang WJ. Chloroplast genomes: diversity, evolution, and applications in genetic engineering. Genome Biology. 2016; 17: 134.
[70]
Wang D, Wu YW, Shih ACC, Wu CS, Wang YN, Chaw SM. Transfer of chloroplast genomic DNA to mitochondrial genome occurred at least 300 MYA. Molecular Biology and Evolution. 2007; 24: 2040–2048.
[71]
Ala KG, Zhao Z, Ni L, Wang Z. Comparative analysis of mitochondrial genomes of two alpine medicinal plants of Gentiana (Gentianaceae). PloS One. 2023; 18: e0281134.
[72]
Richardson AO, Palmer JD. Horizontal gene transfer in plants. Journal of Experimental Botany. 2007; 58: 1–9.

Publisher’s Note: IMR Press stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share
Back to top