Nothing Special   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (714)

Search Parameters:
Keywords = chloroplast genome

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 14586 KiB  
Article
Chloroplast Genome and Description of Borodinellopsis insignis sp. nov. (Chlamydomonadales, Chlorophyta), a Rare Aerial Alga from China
by Qiufeng Yan, Benwen Liu and Guoxiang Liu
Plants 2024, 13(22), 3199; https://doi.org/10.3390/plants13223199 - 14 Nov 2024
Abstract
The genus Borodinellopsis is extremely rare and is the subject of limited research and reports. It currently comprises only two species, Borodinellopsis texensis and Borodinellopsis oleifera, which differ from other globose algae due to their unique centrally radiating chloroplasts. In this study, [...] Read more.
The genus Borodinellopsis is extremely rare and is the subject of limited research and reports. It currently comprises only two species, Borodinellopsis texensis and Borodinellopsis oleifera, which differ from other globose algae due to their unique centrally radiating chloroplasts. In this study, we describe a new specimen in detail based on morphological data and phylogenetic analysis and identify it as B. insignis. B. insignis and B. texensis exhibit a high degree of similarity, likely due to their shared characteristics of centrally radiating chloroplasts and flagella that are significantly longer than the cell body. A phylogenetic tree constructed based on the 18S rDNA sequence indicates that B. insignis and B. texensis form a branch that is distinct from other genera, such as Tetracystis, Spongiococcum, and Chlorococcum. Phylogenetic analysis of the ITS sequence, the rbcL gene, and the tufA gene reveals that B. insignis is significantly different from B. texensis, in that it has oil droplets, smaller vegetative cells and zoospores, and distinct habitats. It is also different from B.oleifera as it has smaller vegetative cells and zoospores, turns red after cultivation, has longer flagella, and resides in different habitats. The chloroplast genomes of B. texensis and B. insignis further show significant differences, with the phylogenetic tree constructed based on the analysis of 49 protein-coding genes forming two separate branches. The collinearity of the chloroplast genomes in B. texensis and B. insignis is poor, with 15 out of the 31 homologous modules displaying inversions and complex rearrangements. Given these differences, we classify this alga as a new species and named it Borodinellopsis insignis sp. nov. Full article
(This article belongs to the Section Plant Systematics, Taxonomy, Nomenclature and Classification)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>,<b>B</b>) The habitat. (<b>C</b>) Vegetative cells. (<b>D</b>) Spherical vegetative cells. (<b>E</b>) Near-spherical vegetative cells. (<b>F</b>) Ellipsoidal vegetative cells. (<b>G</b>) Cell wall thickening. (<b>H</b>) A single chloroplast. (<b>I</b>) Centrally radiating chloroplasts. (<b>J</b>) Chloroplasts of ellipsoidal vegetative cells. (<b>K</b>) A sporangium containing 2 autospores. (<b>L</b>) A sporangium containing 4 autospores. (<b>M</b>,<b>N</b>) Zoospores. (<b>O</b>) Chloroplasts of zoospores. (<b>P</b>,<b>Q</b>) A large number of orange oil droplets. Scale bar: 10 μm. The cell in (<b>N</b>) was fixed and photographed by using Lugol’s solution.</p>
Full article ">Figure 2
<p>The phylogenetic tree constructed with the Bayesian method based on the 18S rDNA sequences. The Bayesian posterior probabilities and maximum likelihood bootstrap values are shown at the nodes, and the new species from this study is indicated.</p>
Full article ">Figure 3
<p>The phylogenetic tree constructed with the Bayesian method based on the ITS sequences. The Bayesian posterior probabilities and maximum likelihood bootstrap values are shown at the nodes, and the new species from this study is indicated.</p>
Full article ">Figure 4
<p>The phylogenetic tree constructed with the Bayesian method based on the <span class="html-italic">rbc</span>L sequences. The Bayesian posterior probabilities and maximum likelihood bootstrap values are shown at the nodes, and the new species from this study is indicated.</p>
Full article ">Figure 5
<p>The phylogenetic tree constructed with the Bayesian method based on the <span class="html-italic">tuf</span>A sequences. The Bayesian posterior probabilities and maximum likelihood bootstrap values are shown at the nodes, and the new species from this study is indicated.</p>
Full article ">Figure 6
<p>The chloroplast genome map of <span class="html-italic">Borodinellopsis insignis</span>, with genes that have different functions indicated by the colors shown in the legend.</p>
Full article ">Figure 7
<p>Chloroplast genome collinearity alignment of three species in <span class="html-italic">Borodinellopsis</span> and related taxa. <span class="html-italic">Pleurastrum insigne</span> (NC042182), <span class="html-italic">Chlorosarcinopsis insigne</span> (NC042250), <span class="html-italic">Dunaliella salina</span> (GQ250046), <span class="html-italic">Borodinellopsis insignis</span> (PQ144585), and <span class="html-italic">Borodinellopsis texensis</span> (MG778121).</p>
Full article ">Figure 8
<p>The phylogenetic tree constructed with the Bayesian method based on 49 shared protein-coding genes. Bayesian posterior probabilities and maximum likelihood bootstrap values are shown at the nodes, and the new species from this study is indicated.</p>
Full article ">
17 pages, 8454 KiB  
Article
Complete Chloroplast Genomes and Phylogenetic Analysis of Woody Climbing Genus Phanera (Leguminosae)
by Yuan Chen, Yanlin Zhao, Wei Wu, Pengwei Li, Jianwu Li, Chang An, Yanfang Zheng, Mingqing Huang, Yanxiang Lin and Quan Yan
Genes 2024, 15(11), 1456; https://doi.org/10.3390/genes15111456 - 12 Nov 2024
Viewed by 395
Abstract
Background: Phanera Lour., a genus in the subfamily Cercidoideae of the family Leguminosae, is characterized by woody liana habit, tendrils, and distinctive bilobate or bifoliolate leaves. The genus holds important medicinal value and constitutes a complex group characterized by morphological diversity and unstable [...] Read more.
Background: Phanera Lour., a genus in the subfamily Cercidoideae of the family Leguminosae, is characterized by woody liana habit, tendrils, and distinctive bilobate or bifoliolate leaves. The genus holds important medicinal value and constitutes a complex group characterized by morphological diversity and unstable taxonomic boundaries. However, limited information on the chloroplast genomes of this genus currently available constrains our understanding of its species diversity. Hence, it is necessary to obtain more chloroplast genome information to uncover the genetic characteristics of this genus. Methods: We collected and assembled the complete chloroplast genomes of nine representative Phanera plants, including Phanera erythropoda, Phanera vahlii, Phanera aureifolia, Phanera bidentata, Phanera japonica, Phanera saigonensis, Phanera championii, Phanera yunnanensis, and Phanera apertilobata. We then conducted a comparative analysis of these genomes and constructed phylogenetic trees. Results: These species are each characterized by a typical quadripartite structure. A total of 130–135 genes were annotated, and the GC content ranged from 39.25–42.58%. Codon usage analysis indicated that codons encoding alanine were dominant. We found 82–126 simple sequence repeats, along with 5448 dispersed repeats, mostly in the form of forward repeats. Phylogenetic analysis revealed that 16 Phanera species form a well-supported monophyletic group, suggesting a possible monophyletic genus. Furthermore, 10 hypervariable regions were detected for identification and evolutionary studies. Conclusions: We focused on comparing chloroplast genome characteristics among nine Phanera species and conducted phylogenetic analyses, laying the foundation for further phylogenetic research and species identification of Phanera. Full article
(This article belongs to the Special Issue Advances in Evolution of Plant Organelle Genome—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Species morphological diversity within <span class="html-italic">Phanera</span>. (<b>a</b>) <span class="html-italic">P. yunnanensis</span>; (<b>b</b>) <span class="html-italic">P. aureifolia</span>; (<b>c</b>) <span class="html-italic">P. japonica</span>; (<b>d</b>) <span class="html-italic">P. saigonensis</span>; (<b>e</b>) <span class="html-italic">P. erythropoda</span>; (<b>f</b>) <span class="html-italic">P. championii</span>.</p>
Full article ">Figure 2
<p>Physical map of the cp genomes in nine <span class="html-italic">Phanera</span> species. The gray column chart in the inner circle represents the CG content. The inner and outer sides of the outer circle represent the genes in clockwise and counterclockwise directions, respectively. Different color blocks represent gene groups with different functions, and the functions corresponding to the colors are annotated in the lower left corner.</p>
Full article ">Figure 3
<p>The RSCU values of each amino acid in the chloroplast genomes of nine <span class="html-italic">Phanera</span> species.</p>
Full article ">Figure 4
<p>Identification of dispersed repeat sequences and SSRs in the cp genomes of nine <span class="html-italic">Phanera</span> species. (<b>a</b>) the number of dispersed repeat sequences of varying lengths within the cp genomes of these species; (<b>b</b>) the number of SSRs in the cp genomes of these species; (<b>c</b>) the number of SSRs with different motifs; (<b>d</b>) The number of SSRs in LSC, SSC, and IRs of the cp genomes of these species.</p>
Full article ">Figure 5
<p>Comparison of LSC, SSC, and IRs boundaries in the cp genome of nine <span class="html-italic">Phanera</span> species.</p>
Full article ">Figure 6
<p>Sequence similarity of the cp genomes between nine <span class="html-italic">Phanera</span> species was analyzed using mVISTA with <span class="html-italic">P. erythropoda</span> as a reference. The horizontal axis represents the gene position in the reference genome of <span class="html-italic">P. erythropoda</span>, while the vertical axis indicates the similarity ranging from 50% to 100%. Exons, t/rRNAs, and conserved noncoding sequences (CNS) are represented by different colors in the lower left corner.</p>
Full article ">Figure 7
<p>Nucleotide polymorphism (Pi) in the cp genomes of 16 <span class="html-italic">Phanera</span> species. (<b>a</b>) Pi of the common CDS regions in the cp genomes. (<b>b</b>) Pi of the common IGS regions.</p>
Full article ">Figure 8
<p>The phylogenetic relationships of the subfamily Cericidoideae were inferred using the complete cp genomes based on BI and ML methods. The support values on the branches are presented in the order of PP<sub>BI</sub>/BS<sub>ML</sub> (with 1.00/100 support values). The pink box indicates species belonging to <span class="html-italic">Phanera.</span> The lineages of these species are labeled on the right. Outgroups are <span class="html-italic">C. chinensis</span> (MZ128523) and <span class="html-italic">C. canadensis</span> (KF856619). The Cercidoideae species used to construct this phylogenetic tree include <span class="html-italic">Adenolobus garipensis</span> (KY806280), <span class="html-italic">Barklya syringifolia</span> (MF135594), <span class="html-italic">Bauhinia brachycarpa</span> (OQ701639), <span class="html-italic">Bauhinia purpurea</span> (MW899157), <span class="html-italic">Bauhinia racemosa</span> (ON456405), <span class="html-italic">Cheniella didyma</span> (OQ701647), <span class="html-italic">Cheniella glauca</span> (OQ701649), <span class="html-italic">Cheniella lakhonensis</span> (OQ701654), <span class="html-italic">Cheniella longipes</span> (OQ701657), <span class="html-italic">Cheniella tianlinensis</span> (OQ701670), <span class="html-italic">Griffonia simplicifolia</span> (MF135596), <span class="html-italic">Lysiphyllum binatum</span> (MF135597), <span class="html-italic">Lysiphyllum hookeri</span> (MF135601), <span class="html-italic">P. apertilobata</span> (OQ701676), <span class="html-italic">Phanera aurea</span> (OQ701682), <span class="html-italic">Phanera carcinophylla</span> (OQ701677), <span class="html-italic">Phanera cardinalis</span> (OQ701678), <span class="html-italic">Phanera cercidifolia</span> (OQ701679), <span class="html-italic">Phanera macrostachya</span> (OQ701685), <span class="html-italic">Phanera ornata</span> var. <span class="html-italic">kerrii</span> (OQ701681), <span class="html-italic">Phanera venustula</span> (OQ701684), <span class="html-italic">Piliostigma thonningii</span> (MF135598), <span class="html-italic">Schnella trichosepala</span> (MF135599), <span class="html-italic">Tylosema esculentum</span> (OP271860), <span class="html-italic">Tylosema fassoglense</span> (MF135600), as well as nine <span class="html-italic">Phanera</span> species assembled in this study. The cp genomes from nine <span class="html-italic">Phanera</span> plants assembled here are marked with a star.</p>
Full article ">
20 pages, 11054 KiB  
Article
Complete Mitogenome Assembly and Comparative Analysis of Vaccinium bracteatum (Ericaceae), a Rich Source of Health-Promoting Molecules
by Peng Zhou, Fei Li, Qiang Zhang and Min Zhang
Int. J. Mol. Sci. 2024, 25(22), 12027; https://doi.org/10.3390/ijms252212027 - 8 Nov 2024
Viewed by 349
Abstract
Vaccinium bracteatum is a valuable plant used both as food and medicine in China, but low production limits the development of its industry. As such, it is important to develop genetic resources for the high-value species for preservation of wild populations and utilization. [...] Read more.
Vaccinium bracteatum is a valuable plant used both as food and medicine in China, but low production limits the development of its industry. As such, it is important to develop genetic resources for the high-value species for preservation of wild populations and utilization. The complete chloroplast and nuclear genomes have already been available; however, its mitogenome has not yet been characterized. Here, the V. bracteatum mitogenome was assembled using HiFi reads, and a comparative analysis was conducted. The mitogenome was a circular sequence of 708,384 bp with a GC content of 45.28%, in which 67 genes were annotated, including 36 protein-coding genes, 26 tRNA genes, 3 rRNA genes, and 2 pseudogenes. Overall, 370 dispersed repeats, 161 simple repeats, and 42 tandem repeats were identified, and 360 RNA editing sites were predicted. There was extensive DNA migration among the three genomes. In addition, most of the protein-coding genes underwent purifying selection throughout evolution, and the nucleotide diversity was highly variable. In addition, comparative analysis indicated that the sizes, structures, and gene contents of the mitogenomes differed significantly, but the GC contents and functional genes were relatively conserved among the Ericales species. Mitogenome-based phylogenetic analysis indicated the precise. evolutionary and taxonomic status of V. bracteatum. The complete mitogenome represents the last link of the reference genome of V. bracteatum and lays the foundation for effective utilization and molecular breeding of this plant. Full article
(This article belongs to the Section Molecular Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>Circular map of the <span class="html-italic">V. bracteatum</span> mitogenome. Genes shown on the outside and inside of the circle are transcribed clockwise and counterclockwise, respectively. The dark gray region in the inner circle depicts the GC content.</p>
Full article ">Figure 2
<p>The length distribution of interspersed repeats in the <span class="html-italic">V. bracteatum</span> mitogenome.</p>
Full article ">Figure 3
<p>RNA editing sites in the PCGs of the <span class="html-italic">V. bracteatum</span> mitogenome.</p>
Full article ">Figure 4
<p>Codon usage bias analysis in the <span class="html-italic">V. bracteatum</span> mitogenome. The block below represents all of the codons that encode each amino acid.</p>
Full article ">Figure 5
<p>Characteristics of nuclear/mitochondrial sequences in <span class="html-italic">V. bracteatum</span>. (<b>a</b>) Homologous fragments between mitogenome and nuclear genome. (<b>b</b>) Distributions of nuclear/mitochondrial matches. (<b>c</b>) Distributions of lengths of nuclear/mitochondrial matches.</p>
Full article ">Figure 6
<p>A schematic representation of the homologous fragments between the mt and cp genomes in <span class="html-italic">V. bracteatum</span>. The purple arcs of the circle represent the cp genome and the blue arcs represent the mitogenome. The lines between the arcs correspond to the genomic fragments that are homologous.</p>
Full article ">Figure 7
<p>Maximum likelihood tree based on the mitogenomes of 22 Ericales species. <span class="html-italic">Nicotiana tabacum</span> and <span class="html-italic">Lactuca serriola</span> were used as outgroups. The number on each node is bootstrap support value. The evolutionary branch length represents the degree of branch variation. The number after the species name is the GenBank accession number. Colors indicate the groups to which the species belong.</p>
Full article ">Figure 8
<p>Boxplots of the Ka/Ks values of the shared PCGs among seven Ericales species.</p>
Full article ">Figure 9
<p>Nucleotide diversity (Pi) among Ericales mitogenomes.</p>
Full article ">Figure 10
<p>The composition of PCGs in the 7 Ericales mitogenomes.</p>
Full article ">Figure 11
<p>Dot-plot graphs indicating syntenic sequences between mitogenomes in Ericales species.</p>
Full article ">Figure 12
<p>Collinearity analysis of the mitogenomes between <span class="html-italic">V. bracteatum</span> and <span class="html-italic">V. macrocarpon</span>. The same color blocks represented homologous regions.</p>
Full article ">
15 pages, 9295 KiB  
Article
Hybrid Origin of ×Leymotrigia bergrothii (Poaceae) as Revealed by Analysis of the Internal Transcribed Spacer ITS1 and trnL Sequences
by Elizaveta O. Punina, Alexander A. Gnutikov, Nikolai N. Nosov, Victoria S. Shneyer and Alexander V. Rodionov
Int. J. Mol. Sci. 2024, 25(22), 11966; https://doi.org/10.3390/ijms252211966 - 7 Nov 2024
Viewed by 243
Abstract
×Leymotrigia bergrothii is a presumed hybrid of Leymus arenarius and Elytrigia repens. This article investigates the hybrid origin and genome composition of this species. These plants are sterile, do not undergo pollination, and do not produce seeds; occasionally, underdeveloped stamens containing [...] Read more.
×Leymotrigia bergrothii is a presumed hybrid of Leymus arenarius and Elytrigia repens. This article investigates the hybrid origin and genome composition of this species. These plants are sterile, do not undergo pollination, and do not produce seeds; occasionally, underdeveloped stamens containing abortive pollen grains form in individual spikelets. The karyotype analysis of root meristem cells revealed a diploid chromosome number of 49 in ×L. bergrothii, reported here for the first time. Subsequently, we examined the intragenomic polymorphism of the transcribed spacer ITS1 in several species of Elytrigia, Elymus, Leymus, Hordeum, and Psathyrostachys, and compared the ribotype patterns of these species with those of ×L. bergrothii. It is shown that the St-ribotype variants found in Elytrigia repens and Elytrigia pseudocaesia, as well as the ribotypes of the La family, which dominate in the genome of Leymus arenarius, correspond to major ribotypes in ×L. bergrothii. The ribotypes of the St and La families are present in the nuclear genome of ×L. bergrothii in almost equal proportions. A comparison of intron and exon sequences of the trnL gene in the chloroplast DNA of Leymus arenarius, Elytrigia repens, and ×L. bergrothii showed that this region in ×L. bergrothii is identical or very close to that of Elytrigia repens, suggesting that Elytrigia repens was the cytoplasmic donor to ×L. bergrothii. Thus, our study confirms the hypothesis that this species represents a sterile first-generation hybrid of Leymus arenarius and Elytrigia repens, reproducing vegetatively. Full article
(This article belongs to the Section Molecular Plant Sciences)
Show Figures

Figure 1

Figure 1
<p>Deformed and uncolored pollen grains in the anthers of ×<span class="html-italic">Leymotrigia bergrothii.</span> Acetic-orcein staining. Bar = 100 μm.</p>
Full article ">Figure 2
<p>Karyotype of ×<span class="html-italic">Leymotrigia bergrothii,</span> 2n = 49, root meristem cell, acetic-orcein staining. Bar = 10 μm.</p>
Full article ">Figure 3
<p>Phylogenetic tree of chloroplast gene <span class="html-italic">trnL</span> sequences based on the Minimal Evolution method and Maximum Composite Likelihood model, using 1000 replicates in the Bootstrap test of phylogeny and the GTR + G substitution model. Bootstrap values are indicated above and below branches. Green color box indicates common origin of ×<span class="html-italic">Leymotrigia bergrothii, Elytrigia, Elymus</span>, and <span class="html-italic">Pseudoroegneria</span> chloroplast sequences. Pink, brownish, and blue colour boxes indicate common origin of <span class="html-italic">Agropyron</span>, <span class="html-italic">Leymus</span>/<span class="html-italic">Psathyrostachys</span>, and <span class="html-italic">Hordeum</span> respectively.</p>
Full article ">Figure 4
<p>The grass ribotype networks showing the level of intragenome and interspecific variation based on the ITS1 sequences. The network was made by the probabilistic method of Statistical Parsimony [<a href="#B31-ijms-25-11966" class="html-bibr">31</a>] by using the TCS 1.21 software [<a href="#B32-ijms-25-11966" class="html-bibr">32</a>] and the tcsBU program network visualization [<a href="#B33-ijms-25-11966" class="html-bibr">33</a>]. The filled circles represent ZOTUs (ribotypes). Colors represent different species and the sizes of wedges and circles are proportional to the relative abundance of a variant in a specimen.</p>
Full article ">Figure 5
<p>Ribotype composition diagram reflecting intragenomic polymorphism of the ITS1 sequences in ×<span class="html-italic">Leymotrigia bergrothii</span> and their parent species.</p>
Full article ">Figure 6
<p>(<b>A</b>) Ribotype tree of the studied species. Index above the branch shows Bayesian support (PP); index below the branch is a bootstrap index. (<b>B</b>) A more detailed picture of the relationships between St, L, and Ns subgenomes. (<b>C</b>) A more detailed picture of the relationships between H, St (Elytrigia repens-variant), and P subgenomes.</p>
Full article ">Figure 6 Cont.
<p>(<b>A</b>) Ribotype tree of the studied species. Index above the branch shows Bayesian support (PP); index below the branch is a bootstrap index. (<b>B</b>) A more detailed picture of the relationships between St, L, and Ns subgenomes. (<b>C</b>) A more detailed picture of the relationships between H, St (Elytrigia repens-variant), and P subgenomes.</p>
Full article ">Figure 7
<p>The geographic locations of sample collection.</p>
Full article ">Figure 8
<p>Spikelets of (<b>a</b>) <span class="html-italic">Elytigia repens</span> (voucher WS16-21); (<b>b</b>) × <span class="html-italic">Leymotrigia bergrothii</span> (voucher Pur19-27); (<b>c</b>) <span class="html-italic">L. arenarius</span> (voucher Pur19-21).</p>
Full article ">
18 pages, 4385 KiB  
Article
Comparative Analyses of the Complete Mitogenomes of Two Oxyria Species (Polygonaceae) Provide Insights into Understanding the Mitogenome Evolution Within the Family
by Lijuan Li, Zhuo Jiang, Ye Xiong, Caleb Onoja Akogwu, Olutayo Mary Tolulope, Hao Zhou, Yanxia Sun, Hengchang Wang and Huajie Zhang
Int. J. Mol. Sci. 2024, 25(22), 11930; https://doi.org/10.3390/ijms252211930 - 6 Nov 2024
Viewed by 356
Abstract
Oxyria (Polygonaceae) is a small genus only comprising two species, Oxyria digyna and O. sinensis. Both species have well-documented usage in Chinese herbal medicine. We sequenced and assembled the complete mitogenomes of these two species and conducted a comparative analysis of the [...] Read more.
Oxyria (Polygonaceae) is a small genus only comprising two species, Oxyria digyna and O. sinensis. Both species have well-documented usage in Chinese herbal medicine. We sequenced and assembled the complete mitogenomes of these two species and conducted a comparative analysis of the mitogenomes within Polygonaceae. Both O. digyna and O. sinensis displayed distinctive multi-branched conformations, consisting of one linear and one circular molecule. These two species shared similar gene compositions and exhibited distinct codon preferences, with mononucleotides as the most abundant type of simple sequence repeats. In the mitogenome of O. sinensis, a pair of long forward repeat sequences can mediate the division of molecule 1 into two sub-genomic circular molecules. Homologous sequence analysis revealed the occurrence of gene transfer between the chloroplast and mitochondrial genomes within Oxyria species. Additionally, a substantial number of homologous collinear blocks with varied arrangements were observed across different Polygonaceae species. Phylogenetic analysis suggested that mitogenome genes can serve as reliable markers for constructing phylogenetic relationships within Polygonaceae. Comparative analysis of eight species revealed Polygonaceae mitogenomes exhibited variability in gene presence, and most protein-coding genes (PCGs) have undergone negative selection. Overall, our study provided a comprehensive overview of the structural, functional, and evolutionary characteristics of the Polygonaceae mitogenomes. Full article
(This article belongs to the Special Issue Advances in Plant Genomics and Genetics: 2nd Edition)
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">O. digyna</span> (<b>A</b>) and <span class="html-italic">O. sinensis</span> (<b>B</b>) mitogenome gene map. Genes shown on the outside and inside of the circle are transcribed clockwise and counterclockwise, respectively. The dark grey region within the inner circle represents the GC content.</p>
Full article ">Figure 2
<p><span class="html-italic">O. digyna</span> (<b>A</b>) and <span class="html-italic">O. sinensis</span> (<b>B</b>) mitogenome repeated sequence diagram. The colored lines in inner circle shows the dispersed repeats with a length greater than or equal to 50 bp in which blue represents forward repeats and yellow represents palindromic repeats. The two outer circles show tandem repeats and simple sequence repeats, both represented as short bars. (<b>C</b>) The long repeat sequence in the mitogenome of <span class="html-italic">O. sinensis</span> molecule 1 mediates the potential conformations generated from the recombination. (<b>D</b>) The gel electrophoresis results of PCR products were amplified using primers. The amplified sequences correspond to the four sites of the recombination conformation. For repeat sequences that were too long to be fully amplified by PCR, we selected portions of the repeat sequences along with 200 bp of their flanking regions for the PCR experiments.</p>
Full article ">Figure 3
<p><span class="html-italic">O. digyna</span> (<b>A</b>) and <span class="html-italic">O. sinensis</span> (<b>B</b>) mitogenome relative synonymous codon usage. The codon families are shown on the <span class="html-italic">X</span>-axis. The RSCU values indicate how frequently a specific codon is observed compared to its expected frequency under uniform synonymous codon usage. RSCU values greater than 1 suggest a preference for specific amino acids in codon usage.</p>
Full article ">Figure 4
<p>Schematic diagram of gene transfer between chloroplast and mitogenomes in <span class="html-italic">O. digyna</span> (<b>A</b>) and <span class="html-italic">O. sinensis</span> (<b>B</b>). The orange and green arcs represent the mitogenome and chloroplast genomes, respectively, with the yellow lines between the arcs corresponding to homologous genomic fragments.</p>
Full article ">Figure 5
<p>The phylogenetic relationships (<b>A</b>) and mitogenomes synteny (<b>B</b>) and mitochondrial genes distribution (<b>C</b>) of <span class="html-italic">O. digyna</span> and <span class="html-italic">O. sinensis</span> with the 8 closely related species. The Maximum Likelihood tree was constructed based on the sequences of 20 conserved PCGs. Regarding mitogenome synteny, bars indicate the mitogenomes, and the ribbons display the homologous sequences between the adjacent species. The red areas indicate where the reversal occurred; the grey areas indicate regions of good homology. Common blocks less than 500 bp in length are not retained, and regions without a common block indicate that they are peculiar to the species. Regarding mitochondrial gene distribution, the colors of the boxes indicate the number of copies that exist in the mitogenomes.</p>
Full article ">Figure 6
<p>Boxplots of pairwise dN, dS values and their ratio among 20 mitochondrial genes in <span class="html-italic">O. digyna</span> and <span class="html-italic">O. sinensis</span> and 8 closely related species mitogenomes.</p>
Full article ">
21 pages, 8277 KiB  
Article
Identification and Expression Analysis of TCP Transcription Factors Under Abiotic Stress in Phoebe bournei
by Wenzhuo Lv, Hao Yang, Qiumian Zheng, Wenhai Liao, Li Chen, Yiran Lian, Qinmin Lin, Shuhao Huo, Obaid Ur Rehman, Wei Liu, Kehui Zheng, Yanzi Zhang and Shijiang Cao
Plants 2024, 13(21), 3095; https://doi.org/10.3390/plants13213095 - 3 Nov 2024
Viewed by 554
Abstract
The TCP gene family encodes plant transcription factors crucial for regulating growth and development. While TCP genes have been identified in various species, they have not been studied in Phoebe bournei (Hemsl.). This study identified 29 TCP genes in the P. bournei genome, [...] Read more.
The TCP gene family encodes plant transcription factors crucial for regulating growth and development. While TCP genes have been identified in various species, they have not been studied in Phoebe bournei (Hemsl.). This study identified 29 TCP genes in the P. bournei genome, categorizing them into Class I (PCF) and Class II (CYC/TB1 and CIN). We conducted analyses on the PbTCP gene at both the protein level (physicochemical properties) and the gene sequence level (subcellular localization, chromosomal distribution, phylogenetic relationships, conserved motifs, and gene structure). Most P. bournei TCP genes are localized in the nucleus, except PbTCP9 in the mitochondria and PbTCP8 in both the chloroplast and nucleus. Chromosomal mapping showed 29 TCP genes unevenly distributed across 10 chromosomes, except chromosome 8 and 9. We also analyzed the promoter cis-regulatory elements, which are mainly involved in plant growth and development and hormone responses. Notably, most PbTCP transcription factors respond highly to light. Further analysis revealed three subfamily genes expressed in five P. bournei tissues: leaves, root bark, root xylem, stem xylem, and stem bark, with predominant PCF genes. Using qRT-PCR, we examined six representative genes—PbTCP16, PbTCP23, PbTCP7, PbTCP29, PbTCP14, and PbTCP15—under stress conditions such as high temperature, drought, light exposure, and dark. PbTCP14 and PbTCP15 showed significantly higher expression under heat, drought, light and dark stress. We hypothesize that TCP transcription factors play a key role in growth under varying light conditions, possibly mediated by auxin hormones. This work provides insights into the TCP gene family’s functional characteristics and stress resistance regulation in P. bournei. Full article
(This article belongs to the Special Issue Molecular Biology and Bioinformatics of Forest Trees)
Show Figures

Figure 1

Figure 1
<p>Conceptual framework of regulation of auxin signaling by <span class="html-italic">TCP</span> transcription factors. Note: The role of <span class="html-italic">TCP</span> transcription factors in regulating auxin biosynthesis and plant responses to abiotic stress. <span class="html-italic">TCP5</span>, <span class="html-italic">TCP13</span>, and <span class="html-italic">TCP17</span> enhance auxin synthesis by upregulating PIF, which in turn increases the expression of YUC enzymes, key players in the auxin biosynthesis pathway. Elevated auxin levels contribute to plant growth and stress adaptation. <span class="html-italic">TCP14</span> and <span class="html-italic">TCP15</span> specifically promote plant elongation by regulating auxin-induced genes associated with cell expansion. Together, these <span class="html-italic">TCP</span> factors support plant resilience under various abiotic stresses.</p>
Full article ">Figure 2
<p>Chromosomal localization analysis of the <span class="html-italic">TCP</span> gene family of <span class="html-italic">Phoebe bournei</span> (Hemsl.). Distribution of <span class="html-italic">PbTCP</span> genes in the <span class="html-italic">P. bournei</span> chromosome. (<b>A</b>) Each chromosome figure shows the chromosome number at the top. The scale on the left can be used to assess chromosome length and gene position. (<b>B</b>) The number of <span class="html-italic">TCP</span> genes on the chromosome.</p>
Full article ">Figure 3
<p>Phylogenetic analysis of <span class="html-italic">TCP</span> protein. (<b>A</b>) Genome and protein sequences <span class="html-italic">P. bournei</span>. (<b>B</b>) The percentage of 3 subfamilies of PbTCP genes. (<b>C</b>) Phylogenetic tree of <span class="html-italic">PbTCP</span> and <span class="html-italic">AtTCP</span> proteins. The arcs of different colors indicate a subfamily of the TCP family. One thousand times with MEGA11 and Bootstrap respectively. The tree was constructed by 29 <span class="html-italic">PbTCPs</span> identified in <span class="html-italic">P. bournei</span> and 25 <span class="html-italic">AtTCPs</span> identified in <span class="html-italic">A. thaliana</span>.</p>
Full article ">Figure 4
<p>Multiple sequence alignment of TCP domains. Note: TCP domain serial alignment of <span class="html-italic">P. bournei</span> TCP family members. At the bottom, the highly conserved amino acid position is indicated by the length of the rectangle; The serial indicator is displayed at the bottom.</p>
Full article ">Figure 5
<p><span class="html-italic">PbTCP</span> protein structure analysis. Note: The categories of three branches are marked on the left, and the confidence level of the protein’s secondary structure is indicated by different colors, and the four levels of confidence are shown in the lower right corner.</p>
Full article ">Figure 6
<p><span class="html-italic">PbTCP</span> conserved domain and motif analysis. (<b>A</b>) Phylogenetic tree of <span class="html-italic">PbTCPS</span>. (<b>B</b>) The motif of <span class="html-italic">PbTCPS</span>. Patterns 1–10 are displayed in rectangles of different colors. Protein length can be estimated using the scale at the bottom. (<b>C</b>) <span class="html-italic">PbTCP</span> protein with conserved domains. (<b>D</b>) Gene structure of the <span class="html-italic">PbTCPS</span> gene. Yellow boxes indicate exons (CDS), black lines indicate introns, and blue boxes indicate 5′ and 3′ untranslated regions. (<b>E</b>) The sequence logo of Motif1.The colored letters indicate the specific sequence of motif1.</p>
Full article ">Figure 7
<p>Genomic location, replication events, and homology of the <span class="html-italic">PbTCP</span> gene. Note: Synteny analysis of the <span class="html-italic">PbTCP</span> family in <span class="html-italic">P. bournei</span>. The gray line represents all isotope blocks in the <span class="html-italic">P. bournei</span> genome, while the red line represents the gene pairs of the duplicate <span class="html-italic">PbTCP</span>. The chromosome number is displayed in a rectangular box for each chromosome.</p>
Full article ">Figure 8
<p>Orthologous analysis of TCP genes in <span class="html-italic">A. thaliana</span>, <span class="html-italic">O. sativa</span>, <span class="html-italic">P. trichocarpa</span> and <span class="html-italic">P. bournei.</span> (<b>A</b>) Genome homology analysis of <span class="html-italic">A. thaliana</span> and <span class="html-italic">P. trichocarpus</span>. The grey line represents the genome pairs between homologous blocks, and the blue line highlights the <span class="html-italic">TCPS</span> gene pairs synthesized in the three species. (<b>B</b>) Number of genome pairs of three clades of different species. (<b>C</b>) Number of shared genes of three genome pairs of three species.</p>
Full article ">Figure 9
<p>Analysis of the cis-acting element of the gene for the promoter. (<b>A</b>) Cis-component predictions of 29 <span class="html-italic">PbTCP</span> gene promoter serial (−2000 bp) were analyzed using PlantCARE technology. Here are the 19 categories of cis-elements. (<b>B</b>) Number of 19 cis-components for the 29 <span class="html-italic">PbTCP</span> genes.</p>
Full article ">Figure 10
<p>Expression spectrum of <span class="html-italic">PbTCP</span>. Note: Different colors are used to indicate the level of expression, and there is an expression value on the right. At the bottom, there are three sub-categories with gene names.</p>
Full article ">Figure 11
<p>The expression of <span class="html-italic">PbTCPs</span> under high temperature, drought, light stress and dark stress was detected by qRT-PCR. (<b>A</b>) Relative gene expression levels at high temperature (40 °C) and control (25 °C). (<b>B</b>) Relative gene expression levels at the same point (4, 8, 12, and 24 h) were treated with 10% PEG nutrient solution in a simulated arid environment. The control group is treated in distilled water. (<b>C</b>) Relative gene expression levels under light stress. (<b>D</b>) Relative gene expression levels under dark stress. (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.0005, **** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 12
<p>The different responses of PIFs and HY5 under Pfr and Pr form. Note: Light promotes the degradation of PIF and the expression of HY5 in the nucleus through photosensitive pigments (Pfr); Under dark conditions, the Pr form promotes the transcription and accumulation of PIFs and the degradation of HY5, which together regulate the normal physiological state of plants.</p>
Full article ">
18 pages, 5629 KiB  
Article
Comparison of the Complete Chloroplast Genomes of Astilbe: Two Korean Endemic Plant Species
by Sang-Chul Kim, Beom Kyun Park and Hyuk-Jin Kim
Genes 2024, 15(11), 1410; https://doi.org/10.3390/genes15111410 - 31 Oct 2024
Viewed by 396
Abstract
Background: Astilbe, consisting of about 18 species, is distributed throughout East Asia and Northeastern America, and most Astilbe species are widely cultivated as ornamental plants. A total of four species of Astilbe have been confirmed to be distributed throughout Korea, two of [...] Read more.
Background: Astilbe, consisting of about 18 species, is distributed throughout East Asia and Northeastern America, and most Astilbe species are widely cultivated as ornamental plants. A total of four species of Astilbe have been confirmed to be distributed throughout Korea, two of which are endemic to Korea. Methods: In this study, we sequenced and assembled the complete chloroplast genomes of two endemic Korean plants using Illumina sequencing technology, identified simple sequence repeats (SSRs) and repetitive sequences, and compared them with three previously reported chloroplast genomes. Results: The chloroplast genomes of the two species were 156,968 and 57,142 bp in length and had a four-part circular structure. They consisted of a large single-copy region of 87,223 and 87,272 bp and a small single-copy region of 18,167 and 18,138 bp, separated by a pair of inverted repeats (IRa and IRb, 25,789 and 25,866 bp). The genomes contained 130 genes, 49 SSRs, and 49 long repetitive sequences. Comparative analysis with the chloroplast genomes of five Astilbe species indicated that A. uljinensis was closely related to A. chinensis and A. taquetii to A. koreana. Conclusions: This study provides valuable references for the identification of two endemic Korean Astilbe species and contributes to a deeper understanding of the phylogeny and evolution of the genus Astilbe. Full article
(This article belongs to the Section Plant Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>Chloroplast genome map of <span class="html-italic">A. uljinensis</span> and <span class="html-italic">A. taquetii</span>. The different colors represent genes in each group. Genes that are transcribed counterclockwise are placed on the outside, and genes that are transcribed clockwise are placed on the inside, depending on the direction of gene transcription. * : Indicates genes that contain introns.</p>
Full article ">Figure 2
<p>Aligned sequence plots for the five <span class="html-italic">Astilbe</span> species using the <span class="html-italic">A. rivularis</span> chloroplast genome as reference.</p>
Full article ">Figure 3
<p>Comparison of four junctions in the chloroplast genome sequences of the five <span class="html-italic">Astilbe</span> species (large single-copy (LSC) region and inverted repeats (IR) and small single-copy (SSC) region).</p>
Full article ">Figure 4
<p>Nucleotide polymorphism analysis of the chloroplast genomes of 5 <span class="html-italic">Astilbe</span> species.</p>
Full article ">Figure 5
<p>Heat map for RSCU analysis of the five <span class="html-italic">Astilbe</span> species. The RSCU values of 61 codons were used for tree clustering. Each column represents a different codon. Each row represents a different <span class="html-italic">Astilbe</span> species. The darker the blue, the lower the RSCU value. The darker the red, the higher the RSCU value.</p>
Full article ">Figure 6
<p>The Ka/Ks ratio of 79 CDSs of 4 <span class="html-italic">Astilbe</span> chloroplast genomes in comparison with <span class="html-italic">A. rivularis</span>. Ka/Ks ratio &gt; 1 indicates strong positive selection.</p>
Full article ">Figure 7
<p>Analysis of SSRs in the five <span class="html-italic">Astilbe</span> cp genomes. (<b>A</b>) Frequencies of different types of SSRs. (<b>B</b>) Number of five SSRs.</p>
Full article ">Figure 8
<p>Analysis of long repeats in the chloroplast genomes of the five <span class="html-italic">Astilbe</span> species. (<b>A</b>) The number of different types of long repeats. (<b>B</b>) The number of repeats of each length.</p>
Full article ">Figure 9
<p>Maximum likelihood (ML) and Bayesian phylogenetic tree based on 79 protein-coding genes from 42 Saxifragales and 2 Cornales species. The ML posterior probability and BI posterior probability are mentioned above the lines. In red are species reported in this study.</p>
Full article ">
18 pages, 9697 KiB  
Article
Characterization of Cytoskeletal Profilin Genes in Plasticity Elongation of Mesocotyl and Coleoptile of Maize Under Diverse Abiotic Stresses
by Xiaoqiang Zhao, Siqi Sun, Zhenzhen Shi, Fuqiang He, Guoxiang Qi, Xin Li and Yining Niu
Int. J. Mol. Sci. 2024, 25(21), 11693; https://doi.org/10.3390/ijms252111693 - 30 Oct 2024
Viewed by 373
Abstract
The plasticity elongation of mesocotyl (MES) and coleoptile (COL) largely determines the morphology of maize seedlings under abiotic stresses. The profilin (PRF) proteins play a pivotal role in cytoskeleton dynamics and plant development via regulating actin polymerization. However, little is known about whether [...] Read more.
The plasticity elongation of mesocotyl (MES) and coleoptile (COL) largely determines the morphology of maize seedlings under abiotic stresses. The profilin (PRF) proteins play a pivotal role in cytoskeleton dynamics and plant development via regulating actin polymerization. However, little is known about whether and how the expression of the ZmPRF gene family regulates MES and COL elongation in maize under adverse abiotic stresses. Here, a total of eight ZmPRF gene members were identified in the maize genome. They were mainly located in the cytoplasm, chloroplast, and mitochondrion, and clearly divided into four classes, based on phylogenetic analysis. Segmental duplication was the main driver for the expansion of ZmPRF genes. Ka/Ks analysis indicated that most ZmPRF genes were intensely purified and selected. Promoter cis-element analysis suggested their potential roles in response to growth and development, stress adaption, hormone response, and light response. The protein–protein interaction network and two independent RNA-sequencing analyses revealed that eight ZmPRF genes and their thirty-seven interacting genes showed varied expression patterns in MES and COL of three maize genotypes under different sowing depths, 24-epibrassinolide application, and light spectral-quality treatments, of which ZmPRF3.3 was a potential core conserved gene for breeding application. Moreover, the quantitative real-time PCR (qRT-PCR) verified that the relative expression levels of most ZmPRF genes in MES and COL under above treatments were significantly correlated with the plasticity elongation of MES and COL in maize. Therefore, these results perform a comprehensive overview of the ZmPRF family and will provide valuable information for the validation of the function of ZmPRF genes in maize development under diverse abiotic stress. Full article
(This article belongs to the Special Issue Advanced Plant Molecular Responses to Abiotic Stresses)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Chromosomal locations of eight profilin (PRF) gene members detected in maize. (<b>B</b>) Phylogenic tree of PRF proteins from seven plant species, including <span class="html-italic">Arabidopsis thaliana</span> (At), <span class="html-italic">Glycine max</span> (Gm), <span class="html-italic">Triticum aestivum</span> (Ta), <span class="html-italic">Gossypium hirsutum</span> (Gh), <span class="html-italic">Sorghum bicolor</span> (Sb), <span class="html-italic">Zea mays</span> (Zm), and <span class="html-italic">Oryza sativa</span> (Os).</p>
Full article ">Figure 2
<p>(<b>A</b>) Phylogenic tree of eight profilin (PRF) protein members in maize. (<b>B</b>) The exon–intron structure of these <span class="html-italic">PRF</span> gene members in maize. (<b>C</b>) Conserved motif distribution of all PRF protein members in maize. (<b>D</b>) The sequence logos of the 15 conserved motifs.</p>
Full article ">Figure 3
<p>(<b>A</b>) Number statistics and element categories of predicted cis-elements in maize <span class="html-italic">profilin</span> (<span class="html-italic">PRF</span>) gene promoter regions. (<b>B</b>) Number statistics of the top 12 cis-elements in maize <span class="html-italic">PRF</span> gene promoter regions.</p>
Full article ">Figure 4
<p>Protein–protein interaction network between profilin (PRF) proteins and other proteins in maize. Nodes represent proteins, and empty nodes are proteins with unknown 3D structures. Connections between nodes represent interactions between proteins, with edge thickness indicating the confidence level of the interaction. Query proteins are depicted in red-colored nodes.</p>
Full article ">Figure 5
<p>(<b>A</b>) Heatmaps of expression patterns from eight <span class="html-italic">profilin</span> (<span class="html-italic">PRF</span>) genes and their thirty-seven interacting genes in mesocotyl (MES) and coleoptile (COL) of W64A and K12 seedlings under three treatments by RNA-sequencing (RNA-Seq). W3M: MES of W64A seedlings at 3 cm sowing depth; W20M: MES of W64A seedlings at 20 cm deep-sowing stress; W20EM: MES of W64A seedlings were treated with 2.0 mg g<sup>−1</sup> 24-epibrassinolide (EBR) application at 20 cm deep-sowing stress; W3C: COL of W64A seedlings at 3 cm sowing depth; W20C: COL of W64A seedlings at 20 cm deep-sowing stress; W20EC: COL of W64A seedlings were treated with 2.0 mg g<sup>−1</sup> EBR application at 20 cm deep-sowing stress; K3M: MES of K12 seedlings at 3 cm sowing depth; K20M: MES of K12 seedlings at 20 cm deep-sowing stress; K20EM: MES of K12 seedlings were treated with 2.0 mg g<sup>−1</sup> EBR application at 20 cm deep-sowing stress; K3C: COL of K12 seedlings at 3 cm sowing depth; K20C: COL of K12 seedlings at 20 cm deep-sowing stress; K20EC: COL of K12 seedlings were treated with 2.0 mg g<sup>−1</sup> EBR application at 20 cm deep-sowing stress. TPM: transcripts per million. (<b>B</b>) The relative expression levels of eight <span class="html-italic">PRF</span> genes in MES and COL of W64A and K12 seedlings under three treatments by quantitative real-time PCR (qRT-PCR). Different lowercase letters in MES or COL in W64A/K12 seedlings under three treatments represent significant differences (<span class="html-italic">p</span> &lt; 0.05) by analysis of variance (ANOVA). (<b>C</b>) Linear relationship between qRT-PCR and RNA-Seq for eight <span class="html-italic">PRF</span> genes in COL of two genotypes under three treatments. *** represents significant difference (<span class="html-italic">p</span> &lt; 0.001) by ANOVA. (<b>D</b>) The Pearson correlation among eight <span class="html-italic">PRF</span> genes and two coleoptile phenotypes in two genotypes under three treatments. COLL: coleoptile length; COLC: coleoptile coarse. * represents significant correlations at <span class="html-italic">p</span> &lt; 0.01 level. (<b>E</b>) Linear relationship between qRT-PCR and RNA-Seq for eight <span class="html-italic">PRF</span> genes in MES of two genotypes under three treatments. *** represents significant difference (<span class="html-italic">p</span> &lt; 0.001) by ANOVA. (<b>F</b>) The Pearson correlation among eight <span class="html-italic">PRF</span> genes and two mesocotyl phenotypes in two genotypes under three treatments. MESL: mesocotyl length; MESC: mesocotyl coarse. * represents significant correlations at <span class="html-italic">p</span> &lt; 0.01 level.</p>
Full article ">Figure 6
<p>(<b>A</b>) Heatmaps of expression patterns from eight <span class="html-italic">profilin</span> (<span class="html-italic">PRF</span>) genes and their thirty-seven interacting genes in mesocotyl (MES) and coleoptile (COL) of Zheng58 seedlings under four treatments by RNA-sequencing (RNA-Seq). MES.Red: mesocotyl was cultured in red light; MES.Blue: mesocotyl was cultured in blue light; MES.White: mesocotyl was cultured in white light; MES.Dark: mesooctyl was cultured in darkness; COL.Red: coleoptile was cultured in red light; COL.Blue: coleoptile was cultured in blue light; COL.White: coleoptile was cultured in white light; COL.Dark: coleoptile was cultured in darkness. TPM: transcripts per million. (<b>B</b>) The relative expression levels of eight <span class="html-italic">PRF</span> genes in MES and COL of Zheng58 seedlings under four treatments by quantitative real-time PCR (qRT-PCR). Different lowercase letters in MES or COL in Zheng58 seedlings under four treatments represent significant differences (<span class="html-italic">p</span> &lt; 0.05) by analysis of variance (ANOVA). (<b>C</b>) Linear relationship between qRT-PCR and RNA-Seq for eight <span class="html-italic">PRF</span> genes in COL of Zheng58 seedlings under four treatments. *** represents significant difference (<span class="html-italic">p</span> &lt; 0.001) by ANOVA. (<b>D</b>) The Pearson correlation among eight <span class="html-italic">PRF</span> genes and three coleoptile phenotypes in Zheng58 seedlings under four treatments. COLL: coleoptile length; COLC: coleoptile coarse; COLW: coleoptile weight. * represents significant correlations at <span class="html-italic">p</span> &lt; 0.01 level. (<b>E</b>) Linear relationship between qRT-PCR and RNA-Seq for eight <span class="html-italic">PRF</span> genes in MES of Zheng58 seedlings under four treatments. *** represents significant difference (<span class="html-italic">p</span> &lt; 0.001) by ANOVA. (<b>F</b>) The Pearson correlation among eight <span class="html-italic">PRF</span> genes and three mesocotyl phenotypes in Zheng58 seedlings under four treatments. MESL: mesocotyl length; MESC: mesocotyl coarse, MESW: mesooctyl weight. * represents significant correlations at <span class="html-italic">p</span> &lt; 0.01 level.</p>
Full article ">Figure 7
<p>(<b>A</b>) Based on the fragments-per-kilobase-per-million mapped (FPKM) values of RNA-sequencing (RNA-Seq) for mesocotyl (MES) and coleoptile (COL) in W64A and K12 seedlings under three treatments, including 3 cm sowing depth, 20 cm sowing depth, 2.0 mg g<sup>−1</sup> 24-epibrassinolide (EBR) was applied at 20 cm sowing depth, the interaction network mapping was constructed between <span class="html-italic">profilin</span> (<span class="html-italic">PRF</span>) and other corresponding genes in maize. (<b>B</b>) Based on the FPKM values of RNA-Seq for MES and COL in Zheng58 seedlings under four treatments, including red light, blue light, white light, and darkness, the interaction networks mapping was constructed between <span class="html-italic">PRF</span> and other corresponding genes in maize.</p>
Full article ">
13 pages, 4890 KiB  
Article
Complete Chloroplast Genome of Crassula aquatica: Comparative Genomic Analysis and Phylogenetic Relationships
by Kyu Tae Park and OGyeong Son
Genes 2024, 15(11), 1399; https://doi.org/10.3390/genes15111399 - 30 Oct 2024
Viewed by 382
Abstract
Background/Objectives: Crassula aquatica (L.) Schonl. is a very small annual plant growing along riverbanks. Chloroplast (cp) genomes, crucial for photosynthesis, are highly conserved and play a key role in understanding plant evolution. In this study, we conducted cp genome analysis of C. aquatica [...] Read more.
Background/Objectives: Crassula aquatica (L.) Schonl. is a very small annual plant growing along riverbanks. Chloroplast (cp) genomes, crucial for photosynthesis, are highly conserved and play a key role in understanding plant evolution. In this study, we conducted cp genome analysis of C. aquatica, aiming to elucidate its phylogenetic position and structural variations. We analyzed and described the features of the complete cp genome of C. aquatica and conducted comparative analysis with the cp genomes of closely related taxa. Rsults: The cp genome was 144,503 bp in length and exhibited the typical quadripartite structure, consisting of a large single-copy region (LSC; 77,993 bp), a small single-copy region (SSC; 16,784 bp), and two inverted repeats (24,863 bp). The cp genome of C. aquatica comprised 113 unique genes, including 79 protein-coding genes (PCGs), 30 tRNAs, and 4 rRNA genes. Comparative genomic analysis of 13 other Crassula species and six outgroups demonstrated highly conserved gene content and order among Crassula species. However, notable differences were observed, including the complete loss of the rpoC1 intron in C. aquatica and several closely related species, which may serve as a synapomorphic trait supporting the monophyly of the subgenus Disporocarpa. We analyzed the nucleotide diversity among 14 Crassula cp genomes and identified five highly variable regions (pi > 0.08) in the IGS regions. Phylogenetic analysis based on 78 PCGs confirmed the monophyly of Crassula and its division into two subgenera: Crassula and Disporocarpa. Although the phylogenetic tree supported the subgeneric classification system, the sectional classification system requires reassessment. Conclusions: In this study, we conducted a comparative analysis of the cp genome of the genus Crassula. We inferred evolutionary trends within the Crassula cp genome and provided molecular evidence supporting the integration of the genus Tillaea into the genus Crassula. However, as this study does not represent all species within the genus Tillaea, further comprehensive phylogenetic analyses are requrired. Full article
(This article belongs to the Topic Plant Chloroplast Genome and Evolution)
Show Figures

Figure 1

Figure 1
<p>Gene map of the <span class="html-italic">C. aquatica</span> chloroplast genome. Genes inside the circle are transcribed clockwise, and genes outside are transcribed counterclockwise. The dark gray inner circle corresponds to the GC content, and the light gray circle corresponds to the AT content.</p>
Full article ">Figure 2
<p>Nucleotide diversity analysis of 14 <span class="html-italic">Crassula</span> chloroplast genomes (window length: 600 bp; step size: 200 bp).</p>
Full article ">Figure 3
<p>Alignment of the <span class="html-italic">rpoC1</span> intron loss in <span class="html-italic">Crassula</span>.</p>
Full article ">Figure 4
<p>Analyses of repeated sequences in 14 <span class="html-italic">Crassula</span> cp genomes. (<b>A</b>) Distributions of tandem repeat types in <span class="html-italic">Crassula</span> cp genomes. (<b>B</b>) Frequencies of tandem repeat types in <span class="html-italic">Crassula</span> cp genomes. (<b>C</b>) Distributions of SSR motifs in <span class="html-italic">Crassula</span> cp genomes. (<b>D</b>) Frequencies of SSR motifs in <span class="html-italic">Crassula</span> cp genomes.</p>
Full article ">Figure 5
<p>Relative synonymous codon usage (RSCU) values of 20 amino acid and stop codons in all protein-coding genes of the chloroplast genome of <span class="html-italic">C. aquatica</span>.</p>
Full article ">Figure 6
<p>The phylogenetic tree reconstruction of 39 Crassulaceae taxa maximum likelihood based on the concatenated sequence of 78 PCGs. Numbers above the branches indicate bootstrap values and posterior probabilities.</p>
Full article ">
12 pages, 3819 KiB  
Article
Pan-Chloroplast Genomes Reveal the Accession-Specific Marker for Gastrodia elata f. glauca
by Jiaxue Li, Daichuan Pan, Junfei Wang, Xu Zeng and Shunxing Guo
Int. J. Mol. Sci. 2024, 25(21), 11603; https://doi.org/10.3390/ijms252111603 - 29 Oct 2024
Viewed by 337
Abstract
Gastrodia elata rhizomes have been applied as traditional medicinal materials for thousands of years. In China, G. elata f. elata (red flower and stem, Ge), G. elata f. viridis (green, Gv), and G. elata f. glauca (black, Gg) represent the primary cultivars in [...] Read more.
Gastrodia elata rhizomes have been applied as traditional medicinal materials for thousands of years. In China, G. elata f. elata (red flower and stem, Ge), G. elata f. viridis (green, Gv), and G. elata f. glauca (black, Gg) represent the primary cultivars in artificial cultivation. Although the annual output of G. elata amounts to 150,000 tons, only 20% is Gg. The long production period, low yield, and high quality of Gg led to its extremely high market prices. However, an effective method to identify this crude drug based solely on its morphological or chemical characteristics is lacking. In this study, the complete chloroplast genomes of three G. elata variants were sequenced using the Illumina HiSeq 2500 platform. Another 21 chloroplast genomes from Gastrodia species, which have published in previous reports, were combined and analyzed together. Our results showed that larger genomic sizes, fewer long tandem repeats, and more simple sequence repeats were the major features of the Gg chloroplast genomes. Phylogenetic analysis showed that the Gg samples were separately clustered in a subclade. Moreover, an accession-specific marker was successfully developed and validated for distinguishing additional rhizome samples. Our study provides investigations of the taxonomic relationships of Gastrodia species. The molecular marker will be useful for differentiating Gastrodia products on the market. Full article
(This article belongs to the Section Molecular Plant Sciences)
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">Gastrodia elata</span> variant sample (inflorescence and rhizome) and chloroplast genome map (outer circle shows gene names, and dark grey bar on inner circle indicates GC content).</p>
Full article ">Figure 2
<p>Heat map shows the differences in synonymous codon usage among <span class="html-italic">Gastrodia</span> species based on RSCU values. * represents termination codon.</p>
Full article ">Figure 3
<p>Maximum likelihood phylogenetic tree of <span class="html-italic">Gastrodia</span> species based on 18 shared chloroplast protein-coding sequences. Bootstrap scores were calculated from 1000 replicates.</p>
Full article ">Figure 4
<p>Nucleotide diversity (Pi) across the chloroplast genomes of <span class="html-italic">Gastrodia</span> species.</p>
Full article ">Figure 5
<p>The development and validation of the accession-specific marker for the <span class="html-italic">G. elata</span> variants. The peak maps of the flower and stem samples (<b>up</b>) and the sequencing results of the rhizome samples (<b>down</b>) from each <span class="html-italic">G. elata</span> variant are shown.</p>
Full article ">
14 pages, 21485 KiB  
Article
Comparative Chloroplast Genome Analysis in High-Yielding Pinus kesiya var. langbianensis
by Dong Wang, Yi Wang, Xiaolong Yuan, Wei Chen and Jiang Li
Diversity 2024, 16(11), 665; https://doi.org/10.3390/d16110665 - 29 Oct 2024
Viewed by 438
Abstract
Pinus kesiya var. langbianensis, a species endemic to Yunnan, China, accounts for over 90% of Yunnan’s Pinus resin production. However, there is significant variation in resin yield among individuals, and molecular markers for identifying high-yielding individuals have yet to be developed. In [...] Read more.
Pinus kesiya var. langbianensis, a species endemic to Yunnan, China, accounts for over 90% of Yunnan’s Pinus resin production. However, there is significant variation in resin yield among individuals, and molecular markers for identifying high-yielding individuals have yet to be developed. In this study, a comparative analysis of complete chloroplast genomes of P. kesiya var. langbianensis was conducted to perform a phylogenetic analysis and differentiate high-yielding individuals. Both high-yielding (HY) and low-yielding (LY) trees possess a typical quadripartite structure, with respective genome sizes of 119,812 bp and 119,780 bp. Each chloroplast genome contains 112 genes, including 72 protein-coding genes, 36 tRNAs, and 4 rRNAs. Furthermore, HY and LY trees contain 30 and 34 SSRs, respectively, with mononucleotide repeats being predominant; neither genome exhibited trinucleotide or pentanucleotide repeats. Six highly variable regions were identified: trnI-CAU-psbA, trnH-GUG-trnI-CAU, rpl16, rrn4.5-rrn5, petG-petL, and psaJ. Phylogenetic analysis based on 72 Pinus species revealed that HY and LY trees clustered separately, with the HY tree grouping with P. kesiya and the LY tree with P. yunnanensis. This study provides a theoretical foundation for the molecular identification of high-yield P. kesiya var. langbianensis individuals and enriches the understanding of its phylogenetic relationships. Full article
Show Figures

Figure 1

Figure 1
<p>Chloroplast genome map of HY and LY <span class="html-italic">P. keisya</span> var. <span class="html-italic">langbianensis</span>. The genes depicted inside the circle are transcribed clockwise, while the genes shown on the outside of the circle are transcribed counterclockwise. Different colors of genes that represent various functions are shown in the left corner of the bottom panel.</p>
Full article ">Figure 2
<p>Analysis of simple sequence repeats (SSRs) in the chloroplast genomes of HY and LY trees. (<b>A</b>) Number of different types of SSRs identified in the chloroplast genomes. (<b>B</b>) Number of SSRs in regular and compound formations in the chloroplast genomes. (<b>C</b>) SSR distributions in the LSC, SSC, and IR regions.</p>
Full article ">Figure 3
<p>Comparison of the borders of large single-copy (LSC), small single-copy (SSC), and inverted repeat (IR) regions in the chloroplast genomes of four <span class="html-italic">Pinus</span> individuals. JLB: junction of LSC and IRb; JSB: junction of SSC and IRb; JSA: junction of SSC and IRa; JLA: junction of LSC and IRa.</p>
Full article ">Figure 4
<p>Relative synonymous codon usage (RSCU) in the chloroplast genomes of 4 <span class="html-italic">Pinus</span> individuals and the amino acids encoded by these codons. The different colors within each group represent the various codons that code for that amino acid. The order of the two columns is <span class="html-italic">P. kesiya</span>, <span class="html-italic">P. yunnanensis</span>, HY <span class="html-italic">P. kesiya</span> var. <span class="html-italic">langbianensis</span>, LY <span class="html-italic">P. kesiya</span> var. <span class="html-italic">langbianensis</span>.</p>
Full article ">Figure 5
<p>Alignment of chloroplast genome sequences in <span class="html-italic">P. kesiya</span>, <span class="html-italic">P. yunnanensis</span>, HY <span class="html-italic">P. kesiya</span> var. <span class="html-italic">langbianensis</span>, LY <span class="html-italic">P. kesiya</span> var. <span class="html-italic">langbianensis</span>.</p>
Full article ">Figure 6
<p>Comparative analysis of the nucleotide diversity (p<sub>i</sub>) values among the four individual <span class="html-italic">Pinus</span> plastomes: (<b>A</b>) protein-coding genes; (<b>B</b>) non-coding and intron regions of each window. The x-axis represents the genes, while the y-axis represents nucleotide diversity.</p>
Full article ">Figure 7
<p>Phylogenetic tree based on whole chloroplast genome datasets of 74 <span class="html-italic">Pinus</span> species using ML methods. The HY and LY trees are highlighted in red.</p>
Full article ">Figure 8
<p>Phylogenetic reconstruction of section <span class="html-italic">Pinus</span> using maximum likelihood (ML) methods based on the whole chloroplast genome (<b>left</b>) and protein-coding genes (<b>right</b>), respectively. The orange branches represent the different topologies between the whole chloroplast genome and protein-coding gene trees.</p>
Full article ">
17 pages, 5522 KiB  
Article
Comparative and Phylogenetic Analysis of Anthurium andraeanum Hybridization Based on Molecular and Morphological Traits
by Yingwen Pan, Jiatong Li and Chaozu He
Horticulturae 2024, 10(11), 1146; https://doi.org/10.3390/horticulturae10111146 - 28 Oct 2024
Viewed by 363
Abstract
Hybridization is considered an important mode of species evolution, but the genetic evolutionary process of Anthurium andraeanum hybridization is still poorly characterized. In order to provide the molecular and morphological basis for phylogenetic analysis in A. andraeanum hybridization, we analyzed the morphological, nuclear [...] Read more.
Hybridization is considered an important mode of species evolution, but the genetic evolutionary process of Anthurium andraeanum hybridization is still poorly characterized. In order to provide the molecular and morphological basis for phylogenetic analysis in A. andraeanum hybridization, we analyzed the morphological, nuclear genomic, and chloroplast genomic data of five A. andraeanum cultivars and explored the correlations between different traits and nuclear and chloroplast genome characteristics. A. andraeanum hybrid 1 is an A. andraeanum ‘Baron’ (♀) × A. andraeanum ‘Spice’ (♂) cross, and A. andraeanum hybrid 2 is an A. andraeanum ‘Cheers’ (♀) × A. andraeanum hybrid 1 (♂) cross. The A. andraeanum hybrids reflected their parents’ heterozygous features in their morphologies, nuclear genomes, and chloroplast genomes. The morphological traits in the F1 generation were widely separated, showing continuous variation. Based on cluster analysis, the five A. andraeanum cultivars could be divided into two groups. The ISSR analysis results were highly correlated with the spathe color. Among the five A. andraeanum cultivars, the composition and structure features of chloroplast genomes were completely the same or highly similar, respectively. Phylogenetic analysis based on complete chloroplast genome data showed that the genetic stability of the chloroplast is high in A. andraeanum, manifested as uniparental maternal inheritance, where the chloroplast genome composition and structural features of hybrids are highly similar to those of the maternal parent. Full article
Show Figures

Figure 1

Figure 1
<p>The spathe and spadix of the five <span class="html-italic">A. andraeanum</span> cultivars investigated in this study. (<b>a</b>) <span class="html-italic">A. andraeanum</span> ‘Baron’; (<b>b</b>) <span class="html-italic">A. andraeanum</span> ‘Spice’; (<b>c</b>) <span class="html-italic">A. andraeanum</span> ‘Cheers’; (<b>d</b>) <span class="html-italic">A. andraeanum</span> hybrid 1; (<b>e</b>) <span class="html-italic">A. andraeanum</span> hybrid 2.</p>
Full article ">Figure 2
<p>Chloroplast genome map of the five <span class="html-italic">A. andraeanum</span> cultivars.</p>
Full article ">Figure 3
<p>Amino acid frequency among the Araceae chloroplast genomes. The X-axis shows amino acids, whereas the Y-axis shows the amino acid frequency as a percentage.</p>
Full article ">Figure 4
<p>Gene arrangement analysis based on Mauve alignment. White boxes represent protein-coding genes, red boxes represent rRNA, black boxes represent tRNA, green boxes represent intron-containing tRNA, the line between two white boxes indicates intron-containing protein-coding genes, and the vertical line on the collinearity analysis graph indicates the collinear region among the Araceae chloroplast genomes.</p>
Full article ">Figure 5
<p>Comparison of the borders of large single-copy (LSC), small single-copy (SSC), and inverted-repeat (IR) regions among the Araceae chloroplast genomes.</p>
Full article ">Figure 6
<p>UPGMA clustering analysis of the five <span class="html-italic">A. andraeanum</span> cultivars based on ISSR markers.</p>
Full article ">Figure 7
<p>Maximum likelihood phylogenetic tree of the Araceae family reconstructed from chloroplast genome data.</p>
Full article ">
18 pages, 5747 KiB  
Article
Comparative Transcriptome Analysis of Non-Organogenic and Organogenic Tissues of Gaillardia pulchella Revealing Genes Regulating De Novo Shoot Organogenesis
by Yashika Bansal, A. Mujib, Mahima Bansal, Mohammad Mohsin, Afeefa Nafees and Yaser Hassan Dewir
Horticulturae 2024, 10(11), 1138; https://doi.org/10.3390/horticulturae10111138 - 25 Oct 2024
Viewed by 558
Abstract
Gaillardia pulchella is an important plant species with pharmacological and ornamental applications. It contains a wide array of phytocompounds which play roles against diseases. In vitro propagation requires callogenesis and differentiation of plant organs, which offers a sustainable, alternative synthesis of compounds. The [...] Read more.
Gaillardia pulchella is an important plant species with pharmacological and ornamental applications. It contains a wide array of phytocompounds which play roles against diseases. In vitro propagation requires callogenesis and differentiation of plant organs, which offers a sustainable, alternative synthesis of compounds. The morphogenetic processes and the underlying mechanisms are, however, known to be under genetic regulation and are little understood. The present study investigated these events by generating transcriptome data, with de novo assembly of sequences to describe shoot morphogenesis molecularly in G. pulchella. The RNA was extracted from the callus of pre- and post-shoot organogenesis time. The callus induction was optimal using leaf segments cultured onto MS medium containing α-naphthalene acetic acid (NAA; 2.0 mg/L) and 6-benzylaminopurine (BAP; 0.5 mg/L) and further exhibited a high shoot regeneration/caulogenesis ability. A total of 68,366 coding sequences were obtained using Illumina150bpPE sequencing and transcriptome assembly. Differences in gene expression patterns were noted in the studied samples, showing opposite morphogenetic responses. Out of 10,108 genes, 5374 (53%) were downregulated, and there were 4734 upregulated genes, representing 47% of the total genes. Through the heatmap, the top 100 up- and downregulating genes’ names were identified and presented. The up- and downregulated genes were identified using the Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway. Important pathways, operative during G. pulchella shoot organogenesis, were signal transduction (13.55%), carbohydrate metabolism (8.68%), amino acid metabolism (5.11%), lipid metabolism (3.75%), and energy metabolism (3.39%). The synthesized proteins displayed phosphorylation, defense response, translation, regulation of DNA-templated transcription, carbohydrate metabolic processes, and methylation activities. The genes’ product also exhibited ATP binding, DNA binding, metal ion binding, protein serine/threonine kinase -, ATP hydrolysis activity, RNA binding, protein kinase, heme and GTP binding, and DNA binding transcription factor activity. The most abundant proteins were located in the membrane, nucleus, cytoplasm, ribosome, ribonucleoprotein complex, chloroplast, endoplasmic reticulum membrane, mitochondrion, nucleosome, Golgi membrane, and other organellar membranes. These findings provide information for the concept of molecular triggers, regulating programming, differentiation and reprogramming of cells, and their uses. Full article
(This article belongs to the Special Issue Plant Tissue and Organ Cultures for Crop Improvement in Omics Era)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Non-organogenic callus, and (<b>B</b>) organogenic callus of <span class="html-italic">G. pulchella</span> with arrow indicating the origin of shoot from the callus mass.</p>
Full article ">Figure 2
<p>Workflow and tools used for mRNA sequence analysis of non-organogenic and organogenic callus of <span class="html-italic">G. pulchella</span>.</p>
Full article ">Figure 3
<p>Length distribution of primary assembly and unigenes of <span class="html-italic">G. pulchella</span>.</p>
Full article ">Figure 4
<p>KEGG pathway classification for <span class="html-italic">G. pulchella</span>.</p>
Full article ">Figure 5
<p>Top hit species distribution pattern showing the number of genes identified in <span class="html-italic">G. pulchella</span> matching with the other plant species.</p>
Full article ">Figure 6
<p>Gene ontology annotation for all a ssembled unigenes in the <span class="html-italic">G. pulchella</span> transcriptome.</p>
Full article ">Figure 7
<p>Volcano plot showing the comparison of differential expressed genes.</p>
Full article ">Figure 8
<p>Heatmap representing the gene expression of the top 100 differentially expressed genes in the non-organogenic and organogenic calluses of <span class="html-italic">G. pulchella</span>.</p>
Full article ">Figure 9
<p>Principal component analysis (PCA) plot showing the relationship between the non-organogenic and organogenic calluses of <span class="html-italic">G. pulchella</span>.</p>
Full article ">
13 pages, 2513 KiB  
Article
Mitochondrial Genome Assembly and Structural Characteristics Analysis of Gentiana rigescens
by Zongyi Xie, Yingmin Zhang, Lixin Wu and Guodong Li
Int. J. Mol. Sci. 2024, 25(21), 11428; https://doi.org/10.3390/ijms252111428 - 24 Oct 2024
Viewed by 482
Abstract
Gentiana rigescens, an alpine plant with significant medicinal value, possesses a complex genetic background. However, comprehensive genomic research on G. rigescens is still lacking, particularly concerning its organelle genome. In this study, G. rigescens was studied to sequence the mitochondrial genome (mitogenome) and [...] Read more.
Gentiana rigescens, an alpine plant with significant medicinal value, possesses a complex genetic background. However, comprehensive genomic research on G. rigescens is still lacking, particularly concerning its organelle genome. In this study, G. rigescens was studied to sequence the mitochondrial genome (mitogenome) and ascertain the assembly, informational content, and developmental expression of the mitogenome. The mitogenome of G. rigescens was 393,595 bp in length and comprised four circular chromosomes ranging in size from 6646 bp to 362,358 bp. The GC content was 43.73%. The mitogenome featured 30 distinct protein-coding genes, 26 tRNA genes, and 3 rRNA genes. The mitogenome of G. rigescens also revealed 70 SSRs, which were mostly tetra-nucleotides. In addition, 48 homologous fragments were found between the mitogenome and the chloroplast genome, with the longest measuring 23,330 bp. The documentation of the mitochondrial genome of G. rigescens is instrumental in advancing the understanding of its physiological development. Decoding the G. rigescens mitogenome will offer valuable genetic material for phylogenetic research on Gentianaceae and enhance the use of species germplasm resources. Full article
(This article belongs to the Section Molecular Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>Mitochondrial genome assembly diagram, with node ID marked in the diagram. (<b>A</b>) Mitochondrial genome assembly primitive diagram. (<b>B</b>) Mitochondrial genome assembly final simplified diagram.</p>
Full article ">Figure 2
<p>Mitochondrial genome characterization of <span class="html-italic">G. rigescens</span>. Different colors indicate various functional groups.</p>
Full article ">Figure 3
<p>Codon usage bias of mitochondrial PCGs of <span class="html-italic">G. rigescens</span>.</p>
Full article ">Figure 4
<p>Repeat sequence analysis histogram. The abscissa represents the mitochondrial molecule and the ordinate represents the number of repeated fragments. (<b>A</b>) Grey legend represents monomeric SSRs, orange legend represents dimeric SSRs, green legend represents trimeric SSRs, blue legend represents tetrameric SSRs, purple legend represents pentameric SSRs, and red legend represents hexameric SSRs. (<b>B</b>) Blue legend represents tandem repeats, yellow legend represents palindromic repeats, green legend represents forward repeats, purple legend represents reverse repeats, and red legend represents complementary repeats.</p>
Full article ">Figure 5
<p>Analysis of gene transfer between the mitochondrial and chloroplast genomes of <span class="html-italic">G. rigescens</span>. The purple lines connecting the arcs indicate the homologous fragments shared between the two genomes.</p>
Full article ">Figure 6
<p>Synteny analysis of 7 mitogenomes.</p>
Full article ">Figure 7
<p>Phylogenetic tree constructed from the shared mitochondrial PCGs of 24 species.</p>
Full article ">Figure 8
<p>The number of RNA editing sites predicted by each mitochondrial PCGs.</p>
Full article ">
13 pages, 26304 KiB  
Article
Assessing Genetic Diversity in Endangered Plant Orchidantha chinensis: Chloroplast Genome Assembly and Simple Sequence Repeat Marker-Based Evaluation
by Yiwei Zhou, Jianjun Tan, Lishan Huang, Yuanjun Ye and Yechun Xu
Int. J. Mol. Sci. 2024, 25(20), 11137; https://doi.org/10.3390/ijms252011137 - 17 Oct 2024
Viewed by 447
Abstract
Orchidantha chinensis T. L. Wu, an endemic species in China, is listed as a key protected wild plant in Guangdong Province. However, the lack of reports on the chloroplast genome and simple sequence repeat (SSR) markers has hindered the assessment of its genetic [...] Read more.
Orchidantha chinensis T. L. Wu, an endemic species in China, is listed as a key protected wild plant in Guangdong Province. However, the lack of reports on the chloroplast genome and simple sequence repeat (SSR) markers has hindered the assessment of its genetic diversity and conservation strategies. The limited number of molecular markers to assess the genetic diversity of this species, and thus develop proper conservation strategies, highlighted the urgent need to develop new ones. This study developed new SSR markers and investigated genetic variation using 96 samples of O. chinensis from seven populations. Through high-throughput sequencing, a complete chloroplast genome of 134,407 bp was assembled. A maximum-likelihood phylogenetic tree, based on the chloroplast genome, showed that O. chinensis is closely related to Ravenala madagascariensis. The study identified 52 chloroplast SSRs (cpSSRs) and 5094 expressed sequence tag SSRs (EST-SSRs) loci from the chloroplast genome and leaf transcriptome, respectively. Twenty-one polymorphic SSRs (seven cpSSRs and fourteen EST-SSRs) were selected to evaluate the genetic variation in 96 accessions across seven populations. Among these markers, one cpSSR and 11 EST-SSRs had high polymorphism information content (>0.5). Cluster, principal coordinate, and genetic structure analyses indicated that groups G1 and G6 were distinct from the other five groups. However, an analysis of molecular variance showed greater variation within groups than among groups. The genetic distance among the populations was significantly positively correlated with geographical distance. These findings provide new markers for studying the genetic variability of O. chinensis and offer a theoretical foundation for its conservation strategies. Full article
(This article belongs to the Special Issue Plant Phylogenomics and Genetic Diversity (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>The chloroplast genome of <span class="html-italic">O. chinensis</span> and the phylogenetic tree of representative species from eight families in Zingiberales. (<b>A</b>) Plants of <span class="html-italic">O. chinensis</span>. (<b>B</b>) The inflorescence of <span class="html-italic">O. chinensis</span>. (<b>C</b>) The chloroplast genome map of <span class="html-italic">O. chinensis</span>. (<b>D</b>) The maximum-likelihood phylogeny tree obtained from eight complete chloroplast sequences in Zingiberales. The position of <span class="html-italic">O. chinensis</span> has been highlighted in red.</p>
Full article ">Figure 2
<p>Genetic distance between different individuals (<b>A</b>) and populations (<b>B</b>) of <span class="html-italic">O. chinensis</span>.</p>
Full article ">Figure 3
<p>Population structure analysis of 96 <span class="html-italic">O. chinensis</span> accessions across seven groups.</p>
Full article ">Figure 4
<p>AMOVA analysis and general linear regression analysis between genetic distance and geographical distance. (<b>A</b>) Percentages of molecular variance. (<b>B</b>) General linear regression analysis. “**” indicates <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
Back to TopTop