Abstract
The rare, vulnerable relict species Pterocarya fraxinifolia is among the last surviving tree species growing in small, scattered populations in the southern Caucasus region; P. fraxinifolia grows up to 1000 m in plain forests and is threatened by habitat loss and environmental changes. Here, we sequenced and annotated the chloroplast genome of P. fraxinifolia from Hyrcanian forests and compared it to the chloroplast genomes of five other Pterocarya species. The evolutionary relationships of P. fraxinifolia were subsequently evaluated using the chloroplast genomes and individual chloroplast loci. The chloroplast genome of P. fraxinifolia was 160,086 bp in length, comprising 128 genes and a typical quadripartite structure. A comparative analysis of the six Pterocarya species revealed limited nucleotide diversity and structural variations in genes. The bulk of the 68 loci identified by SSR analysis comprised A/T repeats. Codon bias analysis revealed strong purifying selection, with the ndhF gene showing the highest Ka/Ks ratio. Our phylogenetic analysis revealed Pterocarya as a sister to the genus Juglans and a distinct subclade within Pterocarya.
Introduction
Relict species have always excited evolutionary biologists and biogeographers who consider these species ‘living fossils’ or relics of prehistoric periods1,2. These species have great value as research models for the geographical distribution of intercontinental rifts and as species that ensure biodiversity and ecosystem balance. Relict species also provide relevant information about the adaptation of species to specific environmental changes, as well as the impact of climate change on the animal and plant kingdoms3.
Hyrcanian forests are hotspots for biodiversity and are home to numerous relict species4, including 280 endemic and subendemic species5,6,7. The genus Pterocarya Kunth (Juglandaceae), commonly referred to as wingnuts, has a disjunct distribution in East Asia and the Caucasus region with its most recent common ancestor present 40 Ma8. Pterocarya comprises six species, which are classified into two sections, Pterocarya (P. fraxinifolia, P. hupehensis, P. stenoptera, and P. tonkinensis) and Platyptera (P. macroptera and P. rhoifolia), on the basis of the presence or absence of scales on the terminal buds9. P. fraxinifolia is the only species in western Asia10. The remaining species of Pterocarya occur in eastern Asia, such as China and Japan11,12,13,14,15. Recently, a series of studies have focused on the phylogeny, biogeography, population genetics, and landscape genetics of species in this genus14,15. However, resources regarding the chloroplast genome in this genus are insufficient, and more research is still needed.
Pterocarya fraxinifolia is a deciduous tree that can reach 20–25 m in height and 1.8 m in trunk diameter and is wind-pollinated to produce wing-nut fruits12. This species is among the last surviving trees growing in small scattered populations in the southern Caucasus region, which includes northern Iran, Georgia, Armenia, Azerbaijan, and the Anatolian region in Turkey12,15. However, less than two decades ago, small populations were first recorded in western Iran in the provinces of Lorestan and Ilam in the Zagros Mountains16.
The chloroplast genome is widely used in phylogenetic studies because of its relatively conserved structure17,18 and uniparental inheritance19,20,. Chloroplast genomes can provide important information about the adaptation of species to different environmental conditions21,22,23. Despite the slow evolutionary rates of chloroplast genomes, coding and noncoding regions are useful for the identification of closely related species24,25,26 and for detecting genome-scale evolutionary patterns. Comparisons of the structure and sequence of these regions across different species within a genus can reveal important evolutionary phenomena such as gene transfer, deletion, or duplication. Recently, with the continuous application of high-throughput sequencing techniques, chloroplast DNA sequences have become readily available13,27,28. However, there is no annotated chloroplast genome available for P. fraxinifolia, which hinders the understanding of the evolution of the chloroplast genome of this species from West Asia14,15.
In this study, we aim to (1) assemble and annotate the chloroplast genome of the relict species P. fraxinifolia from Hyrcanian forests; (2) perform comparative genomics of the chloroplast genomes of six Pterocarya species; and (3) assess the systematic affinity of P. fraxinifolia using phylogenetic analysis of the assembled chloroplast genomes.
Materials and methods
Leaf material for the P. fraxinifolia sample was collected from a wild population in Mazandaran, Iran (Fig. 1). The voucher samples were deposited at the Herbarium of the Nowshahr Botanical Garden (HNBG) under voucher number 12,876.
(a) Fruits in pendulous form; (b) A mature tree; (c) Regeneration under tree canopy; (d) Seedling.
Genomic DNA was extracted using the CTAB method, and its quality and quantity were checked using a Qubit 2.0 and Agilent 2100 Bioanalyzer. Libraries were created and sequenced at Wuhan Benagen Tech Solutions Company Limited, Wuhan, China, using the DNBSEQ platform (paired-end 150 bp). SOAPnuke v1.3.0 was used to filter the raw data, yielding 20 GB of clean data29.
Chloroplast genome assembly and annotation
Raw reads were filtered using Trimmomatic v0.3930 with a quality cutoff of 15 in a 4-base sliding window; any reads that were less than 50 bp were removed, and the adapters were filtered out. The quality of the reads before and after trimming was tested using FASTQC v0.12.1. We used GetOrganelle31 v.1.7.7.0 for chloroplast genome assembly, with the embplant_pt database used as a reference and maximum extension rounds of 15 (-R). GetOrganelle produced two isomers of the whole chloroplast genome of P. fraxinifolia, and each genome had a distinct relative orientation for the small single-copy (SSC) region32. A Python script from GetOrganelle was used along with Bowtie2 v2.5.433 to determine the average read coverage throughout the chloroplast genome. GeSeq v2.0334 was used for the initial chloroplast genome annotation of P. fraxinifolia, and the output from GeSeq was imported into Geneious Prime 2025.0.3 for an additional annotation check via the “Transfer Annotation” function. Chloroplot35 was used to produce a circular representation of the plastome.
Comparative analyses of the Chloroplast genomes
Because the flanking inverted repeat (IR) regions of the chloroplast genome often vary among species, we used CPJSdraw36 to compare the IR regions of the six species. We used CUSP from EMBOSS v6.6.0.0 to calculate relative synonymous codon usage (RSCU) for protein-coding genes of P. fraxinifolia. To identify simple sequence repetitions (SSR), we used a Perl script from the Microsatellite Identification tool (MISA)37. The settings were adjusted to ten, five, and four repeats for mononucleotides, dinucleotides, and trinucleotides, respectively. Forward, reverse, palindrome, and complementary sequences with a minimum repeat length of eight bp and a maximum computed repeat of 50% were analyzed using REPuter38. The complete chloroplast genome sequences of the six Pterocarya species were aligned with Fast Statistical Alignment v1.15.938 to perform the nucleotide diversity analysis. We used a Perl script (https://github.com/xul962464/perl-Pi-nucleotide-diversity) to estimate the nucleotide diversity (PI) with a sliding window analysis with a step size of 200 bp and a window length of 800 bp.
The selection pressure on chloroplast protein-coding genes (CDSs) was evaluated by aligning the nonredundant genes from six species using MAFFT v7.52639. We ran ParaAT.pl v2.040 to compute synonymous substitution rates (Ks), nonsynonymous substitution rates (Ka), and Ka/Ks. Each CDS pair of one-to-one species combinations is used as a homolog with genetic code 11. We estimated Ka, Ks, and Ka/Ks among the six Pterocarya species with KaKs_Calculator v2.041.
Phylogenetic analysis
We constructed a maximum likelihood (ML) phylogenetic tree to understand the relationships of Pterocarya species. Chloroplast genome sequences were acquired from GenBank for the other Pterocarya and related genera in the Juglandaceae family. The multiple sequence alignment contained a total of 21 taxa. We performed our phylogenetic analysis using the full chloroplast genome alignment, treating it as a standard coalescent gene41. The chloroplast genomes were aligned using Fast Statistical Alignment v1.15.942 and then trimmed with trimAL v1.543 with the following settings: -automated1 -res overlap 0.7, -seqoverlap 65. To overcome the alignment issues, we also employed TAPER v1.0.047 with the -m N -a N parameters.
Using RAxML-NG v1.2.144, we constructed the GTR + G model and the ML tree with 500 bootstrap repetitions. The phylogenetic tree was rooted using Engelhardia roxburghiana Wall. as an outgroup. The tree was drawn using FigTree v1.4.4 (https://github.com/rambaut/figtree). To determine the genetic distance between the six Pterocarya species, the HKY85 model45 was used, and a phylogenetic network was generated using the NeighborNet approach in SplitsTree CE v6.0.046.
Results
Chloroplast genome assembly and annotation
The total numbers of raw and trimmed reads for P. fraxinifolia in this study were 143,190,876 and 141,927,817 base pairs (bp), respectively. The number of matched mapped pairs across the chloroplast genome was 393.42 ± 82.15 (Fig. S1). The complete chloroplast genome of P. fraxinifolia has a typical quadripartite structure that is 160,086 bp in length with a large single-copy region (LSC) of 89,582 bp, a small single-copy region (SSC) of 18,398 bp, and a pair of inverted repeat regions (IRs) of 26,053 bp (Fig. 2). A total of 148 genes were annotated in the chloroplast genome of P. fraxinifolia, including 103 protein-coding genes, 37 transfer RNA (tRNA) genes, and eight ribosomal RNA (rRNA) genes (Table 1 and Table S1). The GC content of the chloroplast genome was 36.17%. The annotated complete chloroplast genome of P. fraxinifolia was deposited in GenBank (accession number PV791734).
Schematic map of overall features of the chloroplast genome of P. fraxinifolia. From the center outward, the first track shows the small single-copy (SSC), inverted repeat (IRa and IRb), and large single-copy (LSC) regions. The GC content along the genome is plotted on the second track. The genes are shown on the third track. Genes are color-coded by their functional classification. The transcription directions for the inner and outer genes are clockwise and anticlockwise, respectively. The functional classification of the genes is shown in the bottom left corner.
Comparative analyses of the Chloroplast genome and nucleotide diversity
According to a comparative analysis of the chloroplast genomes of Pterocarya species, the locations of eight genes in the chloroplast maps differed among species. The rps19 gene starts at position zero of the LSC region for P. fraxinifolia, but its position has shifted three times into the IRb region in the others. However, in other species of Pterocarya, a small portion of the genes were located in the IRb region. The ndhF gene in P. fraxinifolia, P. stenoptera, P. macroptera, and P. rhoifolia is located inside the SSC and is 2226 bp in length, whereas in P. tonkinensis and P. hupehensis, it spans 69 and 145 bp, respectively, into the IRb region (Fig. 3a).
(A) Comparisons of LSC, SSC, and IR region boundaries among six Pterocarya species; (B) Nucleotide diversity (π) of CDS regions.
The average nucleotide diversity (π) value was 0.001492, with a range of 0 to 0.00556 (Fig. 3B). The CDSs with the highest π values, which were greater than 0.0031, were ndhF, infA, ycf1, rps15, and matK. The ycf1 gene is found in the SSC area, whereas ndhF, infA, rps15, and matK are found in the LSC region. Nucleotide diversity decreased in both IR zones. Furthermore, 35 CDSs had a π value of zero among the six Pterocarya species, indicating that they were conserved (Table S1).
Repeated sequence analysis
The six Pterocarya chloroplast genomes have an average of 72.6 SSR loci (Fig. 4A), with P. rhoifolia having the most SSR loci (85) of the six species (Table S2). A thorough examination of the chloroplast genome of P. fraxinifolia revealed 68 microsatellites, comprising 63 mononucleotides, four dinucleotides, and one trinucleotide simple sequence repeat. The five types of sequence repeat motifs—forward, reverse, complementary, palindromic, and tandem—are summarized in Table S3 and Fig. 4B. The analysis also revealed that the number of repetitive sequences differed across the six Pterocarya chloroplast genomes. Approximately 96.82% of the mononucleotide repeats found in P. fraxinifolia were classified as A/T (61), and 3.18% (2 repeats) were classified as C/G. In contrast, approximately 88.2% of the repeats found in P. rhoifolia were classified as A/T (75), and 3.52% (3 repeats) were classified as C/G (Fig. 4C). Dinucleotide repeats (6) for P. rhoifolia and (4) for P. fraxinifolia were the next most prevalent type of SSR. This investigation revealed no repeats of tetranucleotides, pentanucleotides, or hexanucleotides.
Analysis of perfect simple sequence repeats (SSRs) in six Pterocarya chloroplast genomes. (A) The frequency of identified SSRs in large single-copy (LSC), inverted repeat (IR,) and small single-copy (SSC) regions; (B) The number of SSR types detected in the nine sequenced chloroplast genomes; (C) The frequency of identified SSR motifs in different repeat class types.
Ka/Ks ratio and codon bias analysis
Strong purifying selection and functional limitations are indicated by the very low Ka/Ks ratios found in most CDS regions among Pterocarya species (Fig. 5A). With the exception of P. tonkinensis and P. stenoptera, the highest Ka/Ks ratio was detected in the chloroplast NADH dehydrogenase F (ndhF) gene. The GC contents for the first, second, and third codon locations were 45.30%, 38.25%, and 30.36%, respectively, whereas the overall coding GC content was 37.97%. The greatest frequencies were 42.361 for the ATT codon and 37.605 for the GAA codon. The only two codons with an RSCU value of 1 were tryptophan (TGG) and methionine (ATG) (Fig. 5B). Every codon ending in A or T had an RSCU value greater than 0.5.
Ka/Ks ratios of chloroplast protein-coding sequences across six Pterocarya species. (A) The X-axis is selected CDS with Ka/Ks ratios above 0.001. The Y-axis shows the mean Ka/Ks ratio for each gene. (B) Relative Synonymous Codon Usage (RSCU) value for each codon.
Phylogenetic analysis
The aligned multiple sequence alignment for the phylogenetic analysis consisted of 158,422 bp across 21 accessions, with 0.21% gaps and 96.19% invariant sites. The phylogenetic tree revealed Pterocarya as a sister genus to Juglans L. with 100% bootstrap support (Fig. 6A). The ML phylogenetic tree confirmed the monophyly of the genus Pterocarya with 100% bootstrap support with two subclades. P. fraxinifolia is a sister to a monophyletic subclade that include P. tonkinensis and P. macroptera and a sister to another subclade that includes P. rhoifolia, P. stenoptera, and P. hupehensis. The network analysis of the six Pterocarya species revealed a topology similar to that of the ML tree, with P. tonkinensis clustering with P. macroptera and P. rhoifolia clustering with P. stenoptera and P. hupehensis, while P. fraxinifolia branched off independently. In this study, the efficiency of two barcode regions, matK and ycf1, in the phylogeny of the genus Pterocarya was evaluated (Fig. 6B and C). The results revealed that the phylogenetic tree based on the matK region was identical to the phylogenetic tree derived from the complete chloroplast genome sequence. Pairwise distance analysis using the HKY85 method revealed that P. fraxinifolia is distantly related to Asiatic Pterocarya species (Fig. S2). The genetic distances between P. macroptera and P. tonkinensis (0.000259) and between P. stenoptera and P. hupehensis (0.000526) were the lowest, whereas the genetic distances between P. fraxinifolia and P. hupehensis (0.002153) and between P. fraxinifolia and P. tonkinensis (0.001928) were more than eightfold greater (Table S4). In the MatK dataset, P. fraxinifolia had three unique character states that differentiated it from other species of Pterocarya (Table S5).
Comparison of three phylogenetic trees based on different chloroplast sequences: (a) Whole chloroplast genome, (b) matK gene region, and (c) ndhF gene region.
Discussion
Chloroplast genomes are useful tools for studying the evolutionary relationships among species because of their preserved structure and uniparental inheritance (usually maternal in angiosperms47,48. Considering mechanisms of plant evolution49,50 and that the evolutionary history of chloroplasts is normally different from that of nuclear markers51,52, the use of genetic information from chloroplasts could reflect how seed dispersal affects the genetic makeup of wild populations and species.
This study is the first to annotate the chloroplast genome of P. fraxinifolia and compare it to that of other species. We found that the positions of eight markers, namely, rps19, rpl2, ycf1 (IRa and IRb), ndhF, trnN, rpl2, and trnH, varied among the six Pterocarya chloroplast genomes. This implies that the expansion and contraction of the IR, LSC, and SSC areas are the primary sources of fluctuations in chloroplast genome size53,54. Between 68 and 85 SSRs were found among the chloroplast genomes of the six Pterocarya species. While the number of poly(G)/(C) repeats was shown to be greater in other angiosperms, the number of poly(A)/(T) repeats was significantly greater in Pterocarya.
Five genes, ndhF, infA, ycf1, rps15 and matK, presented the greatest nucleotide variability (above 0.003). The matK and ycf1 genes have been suggested to function as barcode regions in plants55. The matK gene encodes the maturase protein, which facilitates the splicing of group II introns in several chloroplast genes and is considered a core barcode for land plants50,51. The ycf1 gene, which encodes the TIC214 protein that is essential for plant viability, is the second largest in the chloroplast genome and has recently been assessed for its DNA barcoding potential50,51,52, showing higher variability than the existing chloroplast candidate barcodes (such as rbcL, matK and trnH-psbA). Therefore, the ycf1 gene might be potentially useful as a DNA barcode for the Pterocarya genus56. With the exception of the matK region, none of the seven recommended barcode candidate genes in chloroplast genomes50 have the potential for barcoding of the Pterocarya genus because of a lack of nucleotide variation. Surprisingly, the accuracy of the matK region in resolving the phylogeny of the genus Pterocarya was identical to that of the complete chloroplast genome. Therefore, the matK gene alone is sufficient for reconstructing the phylogenetic relationships within the genus Pterocarya, eliminating the need for the additional time and financial resources required for whole-chloroplast-genome sequencing.
The genus Pterocarya consists of six species and is closely related to Juglans in terms of pollen morphology, wood anatomy and molecular phylogenetics8,9. Our phylogenetic results confirm the sister relationship of Pterocarya to Juglans. Two sections for Pterocarya have been proposed on the basis of the presence or absence of scales on the terminal buds9,13,50. P. fraxinifolia, P. hupehensis, P. stenoptera, and P. tonkinensis belong to the section Pterocarya, while P. macroptera and P. rhoifolia belong to the section Platyptera. According to our chloroplast genome-based phylogeny, this suggested morphological classification is not supported, and the Caucasian wingnut (P. fraxinifolia) is in a distant subclade from the Chinese wingnut (P. stenoptera) and the Japanese wingnut (P. rhoifolia).
The pairwise genetic distance between the Caucasian wingnut and other Asiatic Pterocarya species is greater. This distance might reflect the prolonged isolation and considerable geographic distance between Caucasian wingnut and East Asian species. Recent divergence time analyses based on fossil calibrations estimated the age of P. fraxinifolia between 9.4 and 18.4 Ma from the Miocene period and suggested the westward dispersal of Pterocarya from East Asia8. Wingnut fruit structure could facilitate the dispersal of these species by wind and water57. In this study, we collected P. fraxinifolia materials from its natural habitat in Hyrcanian forests. Our initial phylogenetic results revealed that the publicly available P. fraxinifolia in GenBank (NC046430) is not a P. fraxinifolia and is most likely a misidentified voucher that could be P. stenoptera (data not shown).
Toward conservation of P. fraxinifolia
P. fraxinifolia is classified as a vulnerable relict species on the IUCN Red List12. Our phylogenetic tree, which was constructed on the basis of chloroplast genome analysis, indicates that this species is completely distinct from other species of the genus originating from China and Japan. This distinction might highlight the species’ unique evolutionary path and specialized ecological environments. Recent studies have shown that the potentially suitable ranges of P. fraxinifolia will increase under future climate scenarios8,58, and the rapid loss of its habitat, combined with growing threats such as drought and the destruction of riparian ecosystems in Hyrcanian forests, will result in its conservation an urgent priority.
Data availability
The annotated complete chloroplast genome of P. fraxinifolia was deposited in GenBank, under accession number PV791734.1.
References
Grandcolas, P., Nattier, R. & Trewick, S. Relict species: a relict concept? Trends Ecol. Evol. 29, 655–663 (2014).
Google Scholar
Ian Milne, R. Northern hemisphere plant disjunctions: A window on tertiary land bridges and climate change? Ann. Bot. 98, 465–472 (2006).
Google Scholar
Raposo, M. & Pinto-Gomes, C. Dynamics of vegetation and climate change. Environments 9, 36 (2022).
Google Scholar
Zarandian, A. et al. Anthropogenic decline of ecosystem services threatens the integrity of the unique hyrcanian (Caspian) forests in Northern Iran. Forests 7, 51 (2016).
Google Scholar
Akhani, H., Malekmohammadi, M., Mahdavi, P., Gharibiyan, A. & Chase, M. W. Phylogenetics of the Irano-Turanian taxa of Limonium (Plumbaginaceae) based on ITS NrDNA sequences and leaf anatomy provides evidence for species delimitation and relationships of lineages: phylogenetics of IRano-TUranian Limonium. Bot. J. Linn. Soc. 171, 519–550 (2013).
Google Scholar
Noroozi, J. et al. Endemic diversity and distribution of the Iranian vascular flora across phytogeographical regions, biodiversity hotspots and areas of endemism. Sci. Rep. 9, 12991 (2019).
Google Scholar
Naqinezhad, A. et al. The combined effects of climate and canopy cover changes on understorey plants of the hyrcanian forest biodiversity hotspot in Northern Iran. Glob Change Biol. 28, 1103–1118 (2022).
Google Scholar
Yan, H. et al. Biogeographic history of Pterocarya (Juglandaceae) inferred from phylogenomic and fossil data. J. Syst. Evol. 62, 1165–1176 (2024).
Google Scholar
Manning, W. E. The classification within the Juglandaceae. Ann. Mo Bot. Gard. 65, 1058 (1978).
Google Scholar
Usher, M. B. & Edwards, M. A dipteran from South of the Antarctic circle: belgica Antarctica (Chironomidae) with a description of its larva. Biol. J. Linn. Soc. 23, 19–31 (1984).
Google Scholar
Manchester, S. R. The Fossil History of the Juglandaceae (Missouri Botanical Garden, 1987).
Kozlowski, G., Bétrisey, S. & Song, Y. G. Wingnuts (Pterocarya) & Walnut Family: Relict Trees: Linking the Past, Present and Future. (Natural History Museum Fribourg (NHMF), Department of Education, Culture and Sport of the State of Fribourg, Switzerland, Fribourg, (2018).
Song, Y. G. et al. Phylogeny, Taxonomy, and biogeography of Pterocarya (Juglandaceae). Plants 9, 1524 (2020).
Google Scholar
Wang, T. R. et al. Adaptive divergence and genetic vulnerability of relict species under climate change: a case study of Pterocarya macroptera. Ann. Bot. 132, 241–254 (2023).
Google Scholar
Lu, Z. J. et al. Phylogeography of Pterocarya hupehensis reveals the evolutionary patterns of a cenozoic relict tree around the Sichuan basin. For. Res. 4, e008 (2024).
Akhani, H. & Salimian, M. An extant disjunct stand of Pterocarya fraxinifolia (Juglandaceae) in the central Zagros Mountains, W Iran. Willdenowia 33, 113–120 (2003).
Google Scholar
Yao, X. et al. The first complete Chloroplast genome sequences in actinidiaceae: genome structure and comparative analysis. PLOS ONE. 10, e0129347 (2015).
Google Scholar
Cui, G. et al. Complete Chloroplast genome of hordeum brevisubulatum: genome organization, synonymous codon usage, phylogenetic relationships, and comparative structure analysis. PLOS ONE. 16, e0261196 (2021).
Google Scholar
Ni, Z., Zhou, P., Xin, Y., Xu, M. & Xu, L. A. Parent–offspring variation transmission in full-sib families revealed predominantly paternal inheritance of Chloroplast DNA in Pinus massoniana (Pinaceae). Tree Genet. Genomes. 17, 36 (2021).
Google Scholar
Villanueva-Corrales, S. et al. The complete Chloroplast genome of Plukenetia volubilis provides insights into the organelle inheritance. Front. Plant. Sci. 12, 667060 (2021).
Google Scholar
Pottosin, I. & Shabala, S. Transport across Chloroplast membranes: optimizing photosynthesis for adverse environmental conditions. Mol. Plant. 9, 356–370 (2016).
Google Scholar
Xia, L. et al. Chloroplast Pan-Genomes and comparative transcriptomics reveal genetic variation and temperature adaptation in the cucumber. Int. J. Mol. Sci. 24, 8943 (2023).
Google Scholar
Li, Y. et al. New insights on the phylogeny, evolutionary history, and ecological adaptation mechanism in cycle-cup Oaks based on Chloroplast genomes. Ecol. Evol. 14, e70318 (2024).
Google Scholar
Lim, L. W. K., Chung, H. H. & Hussain, H. Complete Chloroplast genome sequencing of Sago palm (Metroxylon Sagu Rottb.): molecular structures, comparative analysis and evolutionary significance. Gene Rep. 19, 100662 (2020).
Google Scholar
Turudić, A. et al. Variation in Chloroplast genome size: biological phenomena and technological artifacts. Plants 12, 254 (2023).
Google Scholar
Daniell, H., Lin, C. S., Yu, M. & Chang, W. J. Chloroplast genomes: diversity, evolution, and applications in genetic engineering. Genome Biol. 17, 134 (2016).
Google Scholar
Han, H. et al. Analysis of Chloroplast genomes provides insights into the evolution of agropyron. Front. Genet. 13, 832809 (2022).
Google Scholar
Long, L. et al. Complete Chloroplast genomes and comparative analysis of ligustrum species. Sci. Rep. 13, 212 (2023).
Google Scholar
Chen, Y. et al. SOAPnuke: a MapReduce acceleration-supported software for integrated quality control and preprocessing of high-throughput sequencing data. GigaScience 7, (2018).
Bolger, A. et al. The genome of the stress-tolerant wild tomato species solanum pennellii. Nat. Genet. 46, 1034–1038 (2014).
Google Scholar
Jin, J. J. et al. GetOrganelle: a fast and versatile toolkit for accurate de Novo assembly of organelle genomes. Genome Biol. 21, 241 (2020).
Google Scholar
Palmer, J. D., Shields, C. R., Cohen, D. B. & Orton, T. J. Chloroplast DNA evolution and the origin of amphidiploid brassica species. Theor. Appl. Genet. 65, 181–189 (1983).
Google Scholar
Langmead, B. & Salzberg, S. L. Fast gapped-read alignment with bowtie 2. Nat. Methods. 9, 357–359 (2012).
Google Scholar
Tillich, M. et al. GeSeq – versatile and accurate annotation of organelle genomes. Nucleic Acids Res. 45, W6–W11 (2017).
Google Scholar
Zheng, S., Poczai, P., Hyvönen, J., Tang, J. & Amiryousefi, A. Chloroplot: an online program for the versatile plotting of organelle genomes. Front. Genet. 11, 576124 (2020).
Google Scholar
Li, H. et al. CPJSdraw: analysis and visualization of junction sites of Chloroplast genomes. PeerJ 11, e15326 (2023).
Google Scholar
Thiel, T., Michalek, W., Varshney, R. & Graner, A. Exploiting EST databases for the development and characterization of gene-derived SSR-markers in barley (Hordeum vulgare L). Theor. Appl. Genet. 106, 411–422 (2003).
Google Scholar
Kurtz, S. REPuter: the manifold applications of repeat analysis on a genomic scale. Nucleic Acids Res. 29, 4633–4642 (2001).
Google Scholar
Katoh, K. & Standley, D. M. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol. Biol. Evol. 30, 772–780 (2013).
Google Scholar
Zhang, Z. et al. ParaAT: A parallel tool for constructing multiple protein-coding DNA alignments. Biochem. Biophys. Res. Commun. 419, 779–781 (2012).
Google Scholar
Zhang, Z. et al. KaKs_Calculator: calculating Ka and Ks through model selection and model averaging. Genomics Proteom. Bioinf. 4, 259–263 (2006).
Google Scholar
Bradley, R. K. et al. Fast statistical alignment. PLoS Comput. Biol. 5, e1000392 (2009).
Google Scholar
Capella-Gutiérrez, S., Silla-Martínez, J. M. & Gabaldón, T. TrimAl: a tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics 25, 1972–1973 (2009).
Google Scholar
Kozlov, A. M., Darriba, D., Flouri, T., Morel, B. & Stamatakis, A. RAxML-NG: a fast, scalable and user-friendly tool for maximum likelihood phylogenetic inference. Bioinformatics 35, 4453–4455 (2019).
Google Scholar
Hasegawa, M., Kishino, H. & Yano, T. Dating of the human-ape splitting by a molecular clock of mitochondrial DNA. J. Mol. Evol. 22, 160–174 (1985).
Google Scholar
Bryant, D. & Huson, D. H. NeighborNet: improved algorithms and implementation. Front. Bioinforma. 3, 1178600 (2023).
Google Scholar
Brock, J. R., Mandáková, T., McKain, M., Lysak, M. A. & Olsen, K. M. Chloroplast phylogenomics in Camelina (Brassicaceae) reveals multiple origins of polyploid species and the maternal lineage of C. sativa. Hortic. Res. 9, uhab050 (2022).
Google Scholar
Yang, L. et al. Phylogenomic analyses reveal an allopolyploid origin of core Didymocarpinae (Gesneriaceae) followed by rapid radiation. Syst. Biol. 72, 1064–1083 (2023).
Google Scholar
Wu, F. Y., Ma, S. C., Ye, P. M., Ye, H. & Ma, J. L. The complete Chloroplast genome sequence of camellia Zhaiana (Theaceae), a critically endangered species from China. Mitochondrial DNA Part. B. 6, 2425–2426 (2021).
Google Scholar
Wei, F. et al. The complete Chloroplast genome sequence of the medicinal plant sophora tonkinensis. Sci. Rep. 10, 1–13 (2020).
Google Scholar
Birky, C. W. Uniparental inheritance of mitochondrial and Chloroplast genes: mechanisms and evolution. Proc. Natl. Acad. Sci. 92, 11331–11338 (1995).
Google Scholar
Xia, M. et al. Comparative Chloroplast genome study of Zingiber in China sheds light on plastome characterization and phylogenetic relationships. Genes 15, 1484 (2024).
Google Scholar
Kim, K. J. Complete Chloroplast genome sequences from Korean ginseng (Panax schinseng Nees) and comparative analysis of sequence evolution among 17 vascular plants. DNA Res. 11, 247–261 (2004).
Google Scholar
Guo, L., Zhai, J. & Gu, Y. The complete Chloroplast genome sequence of Isoetes Baodongii (Isoetaceae). Mitochondrial DNA Part. B. 9, 667–671 (2024).
Google Scholar
Zhu, S., Liu, Q., Qiu, S., Dai, J. & Gao, X. DNA barcoding: an efficient technology to authenticate plant species of traditional Chinese medicine and recent advances. Chin. Med. 17, 1–17 (2022).
Google Scholar
Dong, W. et al. ycf1, the most promising plastid DNA barcode of land plants. Sci. Rep. 5, 1–5 (2015).
Google Scholar
Maharramova, E. et al. Phylogeography and population genetics of the riparian relict tree Pterocarya fraxinifolia (Juglandaceae) in the South Caucasus. Syst. Biodivers. 16, 14–27 (2018).
Google Scholar
Song, Y., Feng, L., Alyafei, M. A. M., Jaleel, A. & Ren, M. Function of chloroplasts in plant stress responses. Int. J. Mol. Sci. 22, 13464 (2021).
Google Scholar
Acknowledgements
Most data analysis was performed at the Smithsonian Institution Hydra cluster https://doi.org/10.25572/SIHPC. M.V. thanks the support of Rebecca B. Dikow, Matthew Kweskin, and Eric Schuettpelz.
Funding
This work was supported by a grant from the Iranian National Science Foundation (INSF), project No 4024068.
Author information
Authors and Affiliations
Contributions
H. Y. conceived and designed this study. M. V. and S. A. S conducted a formal analysis. M. B. contributed to the analytical methods. S. A. S, H. Y., and M. V. wrote the original draft. G. K. and Y. G. S. edited the manuscript. All authors have read and agreed to the published version of the manuscript.
Corresponding authors
Ethics declarations
Competing interests
The authors declare no competing interests.
Ethics approval and consent to participate
The plant material of P. fraxinifolia was collected from natural populations in northern Iran under a PhD research project approved by Tarbiat Modares University, the Ministry of Science, Research and Technology of Iran. According to national regulations, the collection of plant material for academic research within Iran does not require additional permits when conducted as part of an approved university project. All sampling was done in compliance with institutional and national guidelines. We fully acknowledge the importance of adhering to the IUCN Policy Statement on Research Involving Species at Risk of Extinction as well as the Convention on the Trade in Endangered Species of Wild Fauna and Flora (CITES). We are committed to ensuring that our research complies with these guidelines and supports the conservation of endangered species.
Additional information
Publisher’s note
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Supplementary Information
Below is the link to the electronic supplementary material.
Supplementary Material 1
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
Reprints and permissions
About this article
Cite this article
Sabbagh, S.A., Yousefzadeh, H., Vatanparast, M. et al. Characterization of the chloroplast genome of a relict tree, Pterocarya fraxinifolia (Juglandaceae), and its comparative analysis.
Sci Rep 15, 44153 (2025). https://doi.org/10.1038/s41598-025-23028-5
Received:
Accepted:
Published:
Version of record:
DOI: https://doi.org/10.1038/s41598-025-23028-5
Keywords
- Comparative genomics
- Conservation
- Hyrcanian forests
- Plastome evolution
Source: Ecology - nature.com
