Next Article in Journal
Childhood Hearing Impairment in Senegal
Previous Article in Journal
LhANS-rr1, LhDFR, and LhMYB114 Regulate Anthocyanin Biosynthesis in Flower Buds of Lilium ‘Siberia’
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Comparative Analyses of Chloroplast Genomes for Parasitic Species of Santalales in the Light of Two Newly Sequenced Species, Taxillus nigrans and Scurrula parasitica

Key Laboratory of Bio-Resource and Eco-Environment, Ministry of Education, College of Life Sciences, Sichuan University, Chengdu 610064, China
*
Author to whom correspondence should be addressed.
Genes 2023, 14(3), 560; https://doi.org/10.3390/genes14030560
Submission received: 30 January 2023 / Revised: 16 February 2023 / Accepted: 19 February 2023 / Published: 23 February 2023
(This article belongs to the Section Plant Genetics and Genomics)

Abstract

:
When a flowering plant species changes its life history from self-supply to parasite, its chloroplast genomes may have experienced functional physical reduction, and gene loss. Most species of Santalales are hemiparasitic and few studies focus on comparing the chloroplast genomes of the species from this order. In this study, we collected and compared chloroplast genomes of 12 species of Santalales and sequenced the chloroplast genomes of Taxillus nigrans and Scurrula parasitica for the first time. The chloroplast genomes for these species showed typical quadripartite structural organization. Phylogenetic analysis suggested that these 12 species of Santalales clustered into three clades: Viscum (4 spp.) and Osyris (1 sp.) in the Santalaceae and Champereia (1 sp.) in the Opiliaceae formed one clade, while Taxillus (3 spp.) and Scurrula (1 sp.) in the Loranthaceae and Schoepfia (1 sp.) in the Schoepfiaceae formed another clade. Erythropalum (1 sp.), in the Erythropalaceae, appeared as a third, most distant, clade within the Santalales. In addition, both Viscum and Taxillus are monophyletic, and Scurrula is sister to Taxillus. A comparative analysis of the chloroplast genome showed differences in genome size and the loss of genes, such as the ndh genes, infA genes, partial ribosomal genes, and tRNA genes. The 12 species were classified into six categories by the loss, order, and structure of genes in the chloroplast genome. Each of the five genera (Viscum, Osyris, Champereia, Schoepfia, and Erythropalum) represented an independent category, while the three Taxillus species and Scurrula were classified into a sixth category. Although we found that different genes were lost in various categories, most genes related to photosynthesis were retained in the 12 species. Hence, the genetic information accorded with observations that they are hemiparasitic species. Our comparative genomic analyses can provide a new case for the chloroplast genome evolution of parasitic species.

1. Introduction

Parasitic plants obtain all or most nutrients and water from their host plants and are often divided into hemiparasitic and holoparasitic species [1,2]. Hemiparasitic species can obtain nutrients and water from their host species and can produce organic matter by conducting photosynthesis. Holoparasitic species, however, cannot conduct photosynthesis as their leaves are usually degraded into squamous and there is insufficient chlorophyll [3]. On the tree of life, most parasitic plant species are nestled in different clades of autotrophic plant species, suggesting that they most likely experienced a transformation in life history from autotroph to parasite [4]. During this process, many morphological changes may have occurred, including the reduction of roots into haustoria [5], the decrease of photosynthetic tissues, etc. At the same time, at the genetic level, the chloroplast genomes of parasitic plant species experience a functional and physical reduction, gene loss, etc. [6,7,8]. Since parasitic plants use their haustoria to penetrate their host plants’ tissues and uptake nutrients, their ability to produce carbohydrates via photosynthesis may have degraded, their photosynthetic tissues may have decreased, and their chloroplast genomes may have degraded [3,9]. The chloroplast genomes of parasitic plants have attracted the attention of many researchers [10], but there are few comparative analyses of the chloroplast genomes of different families of hemiparasitic plants in Santalales.
Chloroplast is a key organelle of green plants used for photosynthesis [11], during which organic matter is generated and energy is stored [12]. Chloroplast is thought to be an endosymbiotic organelle and it has an independent and conserved genome [13]. The chloroplast genome is presented as multiple copies of covalent, closed, and conserved circular double-stranded DNA that is single-parentally inherited. However, many studies have shown that it can also be biparentally inherited [14,15,16]. The specific double-stranded circular chloroplast DNA molecule ranges in length from 115 to 165 kb. Usually it consists of four parts: a short single copy (SSC), a long single copy (LSC), and two inverted repeats (IR) [17]. The chloroplast genome contains around 110 genes in most plant species, but several genes are absent in the chloroplast genomes of some plant species [9]. It is prevalent in parasitic plant species whose chloroplast genomes have lost many photosynthetic genes [18]. For many parasitic plant species, their photosynthesis function is reduced, or they cannot photosynthesize, leading to changes in their chloroplast genomes [13,19,20]. Therefore, compared to photosynthetic plant species, the genetic content of the chloroplast genome of parasitic plant species is significantly reduced [13]. These genes are primarily associated with photosynthesis, such as photorespiration genes [21], protein coding genes [22], ribosomal protein genes [23], and transfer RNA (tRNA) genes [22,23]. A few genes remain as pseudogenes. Genes associated with genetic mechanisms are usually retained [19,20,24]. The first holoparasitic plant species sequenced chloroplast genome was Epifagus virginiana, whose genome is only 70 kb and contains only 42 protein-encoding genes and 13 tRNA genes [25]. This species has lost all photosynthesis and energy-producing genes; only a few of those genes remain as pseudogenes, and the entire chloroplast cannot conduct photosynthesis [13].
Santalales is an order of woody flowering plant species. Its members are primarily found in subtropical regions, most of which are hemiparasitic plant species that can produce organic matter through photosynthesis. Still, those plants need to obtain water and minerals via haustoria used to penetrate the stems or roots of their host plants. In this study, we sequenced and assembled the chloroplast genomes of two species in the family Loranthaceae in the order Santalales: T. nigrans and S. parasitica. We compiled ten species of Santalales whose chloroplast genomes were available on NCBI Genbank before April 2018. We compared the size, structure, presence, and absence of genes of the chloroplast genomes in the 12 species of Santalales and constructed the maximum likelihood tree of those species. A previous study of T. nigrans and S. parasitica found that the EST-SSR markers developed based on transcriptomic data of T. nigrans were successfully identified in S. parasitica individuals, suggesting a close relationship between those two species [26]. In this study, we aimed: (a) to figure out the variation of size and assembly of the chloroplast genomes of the 12 species of Santalales, (b) to infer the evolution of the chloroplast genome structure of the 12 species, and (c) to infer the phylogenetic relationship of the 12 species within the order Santalales. Our comparative genomic analyses can provide a new case for the chloroplast genome evolution of parasitic species.

2. Materials and Methods

2.1. DNA Sequencing and Chloroplast Genome Assembly

The experimental samples we collected were the fresh leaves of T. nigrans and S. parasitica, that parasitize on Platanus acerifolia and Ligustrum lucidum, respectively. The samples were collected at Sichuan University. We immediately froze those leaves in a liquid nitrogen tank and sent the frozen leaves to the Novogene Company for DNA extraction and genome sequencing. The modified CTAB method [27] was used to extract the total DNA of the leaf samples and Nanodrop was applied to detect DNA purity (OD 260/280 ratio). Sequencing was performed using the Illumina High-throughput Sequencing Platform (HiSeq/MiSeq). Finally, we obtained raw reads of 6.73 G and 6.85 G from T. nigrans and S. parasitica samples, respectively, and their GC contents were 42.3% and 41.6%, respectively.
We used the Trimmomatic v0.36 [28] to filter the raw reads and obtain high-quality clean reads. The BWA-MEM V0.7.12 [29] was used to compare the clean reads using T. chinensis as the reference chloroplast genome (Genbank: NC_036306.1). The read sequence was mapped to the corresponding reference genome. We used NOVOPlasty v2.6.3 [30] and Velvet v1.2.07 [31] to assemble and splice chloroplast genomes. We spliced contigs into scaffold sequences and then used them to assemble the chloroplast genome.

2.2. Gene Annotation and Sequence Analyses

We used Geneious v11.0.3 [32] to check and then save the assembly results as fasta files. We used Plann V1.1 [33] to annotate the T. nigrans and S. parasitica chloroplast by referring to the effects of T. chinensis chloroplast genome annotation. We then used Geneious v8.1.4 [32] and Sequin v15.10 to correct the annotation results. The final chloroplast genomes sequence of T. nigrans and S. parasitica was submitted to GenBank. The chloroplast gene map was obtained using the online program Organellar Genome DRAW (OGDRAW) v1.2 (http://ogdraw.mpimp-golm.mpg.de/ (accessed on 12 April 2018). To analyze the characteristics of the variations in synonymous codon usage, we used MEGA6 [34] to obtain the relative synonymous codon usage values (RSCU) and codon usage by neglecting the influence of amino acid composition.

2.3. Genome Comparison

The genomes of the 12 species of Santalales were compared using the online program mVISTA (http://genome.lbl.gov/vista/mvista/submit.shtml (accessed on 15 April 2018)). In addition to T. nigrans and S. parasitica, chloroplast genomes data for another 10 species of Santalales were downloaded from the NCBI database, including five Santalaceae species: Viscum album (NC_028012.1), V. crassulae (NC_027959.1), V. coloratum (NC_035414.1), V. minimum (NC_027829.1), Osyris alba (NC_027960.1); two Loranthaceae species: T. chinensis (NC_036306.1), T. sutchuenensis (NC_036307.1), Schoepfiaceae: Schoepfia jasminodora (NC_034228.1); Olacaceae: Erythropalum scandens (NC_036759.1) and Opiliaceae: Champereia manillana (NC_034931.1). We labeled the gene exon, intron sites, and transcriptional direction for T. nigrans and S. parasitica, and submitted that information to the online program mVISTA (http://genome.lbl.gov/vista/mvista/submit.shtml(accessed on 15 April 2018). We ran mVISTA twice: once using T. nigrans as the reference, and again using S. parasitica as the reference.

2.4. Phylogenetic Analysis

Phylogenetic trees were constructed based on the chloroplast genomes of 12 species of Santalales to analyze their phylogenetic relationships. We used MAFFT V7.158 [35] and MEGA v6.0 [34] to extract and align the amino acid sequences of the proteins encoded by their common genes. An ML tree was constructed by the RAxML V8.2.11 software [36] based on the PROTGAMMAJTT model. The outgroups consist of two holophytes (Pvrola. rotundifolia (KU833271.1) and Vaccinium. macrocarpon (NC019616.1). Then we used FigTree v1.4.3 to check the results.

3. Results

3.1. Characteristics of T. nigrans and S. parasitica Chloroplast Genomes

The chloroplast genomes of T. nigrans and S. parasitica are circular molecules that retained the typical structure (Figure 1). The lengths of the chloroplast genomes of those two species were 121,419 bp and 121,750 bp, respectively. Both chloroplast genomes comprised IR regions (T. nigrans, 22,569 bp; S. parasitica, 22,687 bp) that were separated by the LSC region (T. nigrans, 70,181 bp; S. parasitica, 70,270 bp) and the SSC region (T. nigrans, 6100 bp; S. parasitica, 6106 bp) (Figure 1). The GC contents of the T. nigrans and S. parasitica chloroplast DNA were 37.4% and 37.2%, respectively. These were unevenly distributed throughout their chloroplast genomes (Table 1). A total of 106 genes were annotated in both the T. nigrans and S. parasitica, including four pseudogenes (Table 2), eight rRNA genes, 28 tRNA genes, and 66 protein-coding genes.
In T. nigrans, 3993 codons (9.9%) were encoded for leucine, while 581 (1.4%) were encoded for tryptophan. Similarly, in S. parasitica, 4027 codons (9.9%) were encoded for leucine and 524 (1.3%) were encoded for tryptophan. In both species, leucine was the most prevalent and tryptophan was the least prevalent of these amino acids. The complete chloroplast genome sequence of T. nigrans and S. parasitica has been deposited in GenBank under accession numbers MH095982 and MH101514, respectively.

3.2. Comparative Chloroplast Genomic Analysis

A comparative analysis of the chloroplast genomes of the 12 species of Santalales (Table 3) demonstrated that the lengths of the genomes varied from 118 kb to 156 kb. The length of the S. jasminodora chloroplast genome was the shortest and the length of the E. scandens chloroplast genome was the longest. E. scandens had the largest LSC (84,799 bp) and SSC (18,567 bp) of the 12 species. However, E. scandens also had the smallest proportion of LSC and the largest proportion of SSC of the 12 species. The length of IR varied from 22 k to 28 kbp. S. jasminodora had the smallest IR, much shorter than the IRs of the other 11 species, but its LSC had the largest proportion of chloroplast genomes. C. manillana had the largest IR (28,075 bp) and the largest proportion of chloroplast genomes.
By comparing the lengths of LSC, SSC, and IR (Table 3), the chloroplast genomes from the 12 species can be classified into six categories: Viscum, Osyris, Champereia, Schoepfia, and Erythropalum. Each represented an independent category, while the three Taxillus species and Scurrula represented a sixth category. The three Taxillus species and Scurrula had a minimal chloroplast genome size and composition differences and the smallest proportion of the SSC. Meanwhile, the lengths of the chloroplast genome, LSC, SSC, and IR of the four Viscum species were similar in gene numbers, position, and mVista analysis. Yet, their total number of genes, number of protein-coding genes, and tRNA genes were different. When comparing the above two categories, the three Taxillus species, one Scurrula species, and the four Viscum species all showed slight differences. However, among all 12 species, the remaining four categories (Champereia, Erythropalum, Osyris, and Schoepfia) all showed apparent differences, especially concerning the lengths of the chloroplast genome, SSC, and IR (Table 3).
Overall, among the chloroplast genomes of the 12 species, the coding areas were more conservative than the non-coding areas. The IR had a lower divergence than the LSC and SSC. Eight rRNA genes did not have relatively large indels and were highly conserved. TRNA genes and protein-coding genes, such as rpoC2 and ycf2, had large indels. The difference between the chloroplast genomes of T. nigrans and T. sutchuenensis was the smallest. The chloroplast genomes of the three species of Taxillus and S. parasitica were very similar to each other (Figure 2 and Figure 3).
In the chloroplast genomes of the 12 species, eight rRNA genes, several rps genes (rps2, 3, 4, 7, 8, 11, 12, 14, 18, and 19), rpl genes (rpl2, 14, 20, 22, 23, and 36), and trn (E and fM) genes were all present and were relatively conservative. The DNA sequences from gene trnT-GGU to gene trnQ-UUG in V. minimum were found to be in a reverse direction compared to the other 11 species. The trnR-ACG and trnN-GUU genes were absent in O. alba; at the same time, the DNA sequences of the ccsA plus trnL-UAG genes were found to be in a reverse direction when compared to the other 11 species.
By comparing and analyzing the chloroplast genomes of the 12 species of Santalales, we found that, in the three Taxillus species and the one Scurrula species, some genes were missing, including several NAD(P)H dehydrogenase complex subunits (ndh gene), four ribosomal protein genes (rpl32, rps15, rps16, and rps33), one ycf gene (ycf1), and the initiation factor gene (infA). Ten tRNA genes (trnL-UAA, trnI-GAU, trnK-UUU, trnP-GGG, trnP-TGG, trnH-GUG, trnQ-UUG, trnG-UCC, trnV-UAC, and trnA-UGC) were also missing in the chloroplast genomes of these four species. Two ribosomal protein genes (rpl16 and rpl2) and the repeat gene ycf15 degenerated into pseudogenes in the same four species as their gene coding regions were interrupted by deletions, insertions, or internal stop codons. The pseudogene rpl2 was in the IRb region. The 12 species of Santalales retained several photosystem genes (psaA, B, C, I, J, ycf3, 4 and psbA, B, C, D, E, F, H, I, J, K, L, M, N, T, Z), pet genes (petA, B, D, G, L, N), atp genes (atpA, B, E, F, H, I), rpo genes (rpoA, B, C1, C2), and the rbcL gene.
At the LSC/IRb junction, rpl2 genes locate here and extend to LSC regions with different lengths for the four Loranthaceae species (T. sutchuenensis, 217 bp; T. nigrans, 360 bp; T. chinensis, 220 bp; and S. parasitica, 215 bp). In S. jasminodora, the rpl2 gene was in the LSC region, 7677 bp away from the IRb region. The rpl2 genes in E. scandens, C. manillana, and the five species of Santalaceae (O. alba and the 4 Viscum species) were in the IRb region (E. scandens, 79 bp; C. manillana, 134 bp; O. alba, 50 bp; V. album, 975 bp V. coloratum, 980 bp; V. crassulae, 800 bp; and V. minimum, 750 bp). In C. manillana, the junction of LSC/IRb was the rps19 gene, extending 205 bp to the LSC region. In the O. alba, S. Jasminodora, and the 4 Loranthaceae species, the rps19 gene was in the LSC region (E. scandens, 25 bp; O. alba, 63 bp; S. jasminodora, 9220 bp; S. parasitica, 375 bp; T. chinensis, 379 bp; T. nigrans, 537 bp; and T. sutchuenensis, 394 bp). Among the 4 Viscum species, the rps19 gene was in the IRb region (V. album, 638 bp; V. coloratum, 643 bp; V. crassulae, 463 bp; V. minimum, 407 bp). In S. jasminodora, the trnL-CAA gene was located at the junction of LSC/IRb and extended 26 bp into the LSC region. The trnL-CAA gene was in the IRb region for the other 11 species (V. album, 10,320 bp; V. coloratum, 10,334 bp; V. crassulae, 10,040 bp; V. minimum, 10,113 bp; O. alba, 10,081 bp; C. manillana, 10,059 bp; T. chinensis, 9596 bp; T. nigrans, 9484 bp; T. sutchuenensis, 9522 bp; S. parasitica, 9644 bp; and E. scandens, 10,147 bp). In the IRb/SSC junction, the 4 Loranthaceae species were trnL-UAG genes, extending 65 bp to the SSC region. The trnL-UAG genes of the other eight species are all located in the SSC region (C. manillana, 185 bp; E. scandens, 4246 bp; O. alba, 917 bp; S. jasminodora, 733 bp; V. album, 561 bp; V. coloratum, 478 bp; V. crassulae, 646 bp; V. minimum, 620 bp). For C. manillana, E. scandens, and O. alba, the ycf1 gene was in the IRb region, 0 bp, 1 bp, and 0 bp away from the junction; the other nine species have lost the ycf1 gene near IRb/SSC junction. For C. manillana, E. scandens, S. jasminodora, V. album, V. coloratum, and V. crassulae, the ycf1 gene were located at the SSC/IRa junction, which extends to SSC regions 2701 bp, 4516 bp, 5284 bp, 5390 bp, 5441 bp, and 5486 bp, respectively. For O. alba, the ycf1 gene was in the SSC region, 1004 bp away from the junction. In other species, the ycf1 gene has been lost from the chloroplast genome. All species had no genes at the IRa/LSC junction. The trnH-GUG gene was in the LSC region in C. manillana, E. scandens, O. alba, V. album, V. crassulae, V. minimum species, 28 bp, one bp, two bp, two bp, two bp, and two bp from junctions, respectively. Both in S. jasminodora and V. coloratum are in the IRa area, 185 bp and 127 bp from the junction (Figure 4). The trnH-GUG gene was lost from the chloroplast genomes in the other four species.

3.3. Phylogenetic Analysis

The chloroplast genome can provide essential data for evolution, taxonomy, and phylogenetic studies. 11 out of the 12 nodes in the maximum likelihood tree received 94% to 100% bootstrap support (Figure 5). Phylogenetic analysis revealed that 11 out of the 12 species of Santalales clustered into two highly supported clades. One clade includes Viscum (4 spp.), Osyris (Santalaceae), and Champereia (Opiliaceae), while the other clade includes Taxillus (3 spp.), Scurrula (Loranthaceae) and Schoepfia (Schoepfiaceae). In addition, a third clade is the most distant clade among the Santalales, and only includes Erythropalum (Erythropalaceae). While the monophyly of Viscum was strongly sustained, the monophyly of Santalaceae received moderate bootstrap support (75%), as indicated by the apparent sister relationship between Viscum (Visceae) and Osyris (Santaleae). In this way, the monophyly of Santalaceae deserves closer research.

4. Discussion

At the third codon position, it prefers A and T. Also, the preference for A and T appears at the stop codons. We found that the usage of A-ending and U-ending is generally excessive. Other than trnL-CAA and trnS-GGA, all types of synonymous codons (RSCU > 1) prefer to end with A or U (Table 4). Commonly, it prefers A and T in plant chloroplast genomes at the third codon position [37]. This universal law can differentiate chloroplast DNA from mitochondrial and nuclear DNA [37].
Advances in phylogenetic studies indicate that the chloroplast genome’s evolution involves nucleotide substitutions and changes in genomic structure [37,38]. Examples of the latter include the loss of genes and introns. Previous studies have shown that introns significantly regulate gene expression and selective splicing, enhancing exogenous gene expression at particular sites in plants at specific times. It has been noted that introns can significantly stabilize transcription in some eukaryotes [13,39]. The chloroplast genomes of T. nigrans and S. parasitica contained seven intron-existing genes, including atpF, rpoC1, ycf3, rps12, petB, petD, and rpl2 genes (Table 5). Among them, the ycf3 gene, located in the LSC, contained two introns and three exons. The rps12 gene was dedicated to trans-splicing, with the 5′ exon in the LSC and the 3′ exon in the IR. Comparative analysis revealed that the size of the rps12 gene was also decreased due to the loss of cis-spliced introns. However, the ability of the rpsl2 gene to code and express could remain intact; none of the gene’s coding region has been degraded due to its reduced length or frameshift mutation [40].
The loss of genes often occurs during the life cycle transition from autotrophy to parasitism in plants [13,41]. Overall, all 12 species of Santalales in this study retained some photosystem genes, indicating that these species have a relatively complete photosynthetic capacity despite some loss of photosynthesis-related genes. The comparative analysis of chloroplast genomes showed that both T. nigrans and S. parasitica lost the ndh gene and infA gene in that chloroplast genome, which is consistent with general models of plastome degradation [42]. The ndh gene plays an essential role in plant photoautotrophy. Its expression marks significant plant transition evolution [43]. However, cases where the ndh gene was missing or degenerated into pseudogenes have been found in the chloroplast genomes of many parasitic plant species [13,37,38,44,45]. By comparing gene content, plastome structure, and selection pressure, Li, et al. [45] found that hemiparasitism accelerates the pseudogenization and loss of the plastid ndh gene of Orobanchaceae plant species. Genetic changes in parasitic plants of Santalales are usually characterized by pseudogenization or loss of the ndh complex gene [42].
Among the 12 species of Santalales in this study, the ndhA gene located in the SSC of S. jasminodora and the repeat gene ndhB located in the IR of the species V. minimum all degenerated into pseudogenes. More importantly, however, except for E. scandens (an autotrophic plant), all ndh genes were found missing in the other nine species of Santalales, include autotrophic plant C. manillana. These results indicate that the photosynthetic capacity of these species gradually degraded during the evolution to heterotrophy, which may also reflect the increased host dependence of these species. As for E. scandens, the ndh gene in its chloroplast genome is more intact than the other 11 Santalales species. The ndhJ, K, and C genes located in the LSC and the repeat gene ndhB located in the IR, the ndhF, D, E, G, I, A, and H in the SSC of the E. scandens chloroplast genome were retained and the chloroplast genome is the longest. These results demonstrated that the photosynthetic capacity of E. scandens was stronger than that of hemiparasites such as T. nigrans and S. parasitica, and that it is less dependent on its host. Our findings suggest that the lifestyle transition of parasitic plants is accompanied by the relaxation of chloroplast gene purifying selection [7,42]. However, the chloroplast gene of hemiparasitic plants evolution is comparatively conserved in the hemiparasitic plants. Our study supports the idea that hemiparasitic plants still reserve the ability of photosynthesis and can produce organic matter by conducting photosynthesis.

5. Conclusions

We used high-throughput sequencing technology to sequence the chloroplast genome sequences of two hemiparasitic species: T. nigrans and S. parasitica. The sequencing, assembly, annotation, and comparative analysis showed that the T. nigrans chloroplast genome was 121,419 bp and the S. parasitica chloroplast genome was 121,750 bp. A total of 106 genes of T. nigrans and S. parasitica were annotated, including 66 protein-coding genes, 28 tRNA genes, eight rRNA genes, and four pseudogenes. In the comparison of the chloroplast genomes of the 12 species of Santalales, E. scandens chloroplast DNA was the largest and the S. jasminodora chloroplast DNA was the smallest. All ndh genes associated with NAD(P)H dehydrogenase have become pseudogenes or have been completely lost, while some tRNA genes have been lost, a few ribosomal protein genes and ycf genes have been degraded, and the loss of these genes has significantly reduced the size of SSC and LSC of T. nigrans and S. parasitica chloroplast DNA. Phylogenetic analysis showed that 11 of the 12 species were clustered into two clades with high bootstrap support. In agreement with phylogenetic analyses, the loss of genes, the order of genes, and the structure of genes in the chloroplast genomes of these species can be assigned to six categories: Viscum (4 spp.), Osyris (Santalaceae), and Champereia (Opiliaceae) formed one clade, and Taxillus (3 spp.), Scurrula (Loranthaceae), and Schoepfia (Schoepfiaceae) formed another clade. Erythropalum (Erythropalaceae) was the most distant clade within the Santalales. Our phylogenetic relationship among the families of the 12 species of Santalales, based on the chloroplast genomes, is consistent with recently reported phylogenetic trees (e.g., Angiosperm Phylogeny Website, Version 14, 2017). Although different genes are lost in various categories, most genes related to photosynthesis are retained in the 12 species. Hence, the genetic information from chloroplast genomes accorded with observations that they are hemiparasitic plants. This study will provide information for further research about the chloroplast DNA evolution and phylogenetic and molecular ecology of the family Santalales, and our comparative genomic analyses provide a new case for the chloroplast genome evolution of parasitic plants.

Author Contributions

Conceptualization, N.M. and K.M.; methodology, X.Y. and N.M.; software, X.Y. and Z.F.; data curation, N.M. and K.M.; writing—original draft preparation, X.Y. and N.M.; writing—review and editing, N.M., X.Y., Z.F. and K.M; supervision, N.M.; project administration, N.M.; funding acquisition, N.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key Research and Development Program of China, grant number 2016YFD0600203.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available upon request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Heide-Jørgensen, H.S. Parasitic Flowering Plants, 1st ed.; Brill: Leiden, The Netherlands, 2008; Volume 10, pp. 6–7. [Google Scholar]
  2. Xu, Y.; Zhang, J.; Ma, C.; Lei, Y.; Shen, G.; Jin, J.; Eaton, D.A.R.; Wu, J. Comparative genomics of orobanchaceous species with different parasitic lifestyles reveals the origin and stepwise evolution of plant parasitism. Mol. Plant 2022, 15, 1384–1399. [Google Scholar] [CrossRef] [PubMed]
  3. Young, N.D.; Steiner, K.E.; dePamphilis, C.W. The Evolution of Parasitism in Scrophulariaceae/Orobanchaceae: Plastid Gene Sequences Refute an Evolutionary Transition Series. Ann. Mo. Bot. Gard. 1999, 86, 876–893. [Google Scholar] [CrossRef]
  4. Twyford, A.D. Parasitic plants. Curr. Biol. 2018, 28, R857–R859. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Westwood, J.H.; Yoder, J.I.; Timko, M.P.; dePamphilis, C.W. The evolution of parasitism in plants. Trends Plant Sci. 2010, 15, 227–235. [Google Scholar] [CrossRef] [PubMed]
  6. Barrett, C.F.; Wicke, S.; Sass, C. Dense infraspecific sampling reveals rapid and independent trajectories of plastome degradation in a heterotrophic orchid complex. New Phytol. 2018, 218, 1192–1204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Barrett, C.F.; Freudenstein, J.V.; Li, J.; Mayfield-Jones, D.R.; Perez, L.; Pires, J.C.; Santos, C. Investigating the path of plastid genome degradation in an early-transitional clade of heterotrophic orchids, and implications for heterotrophic angiosperms. Mol. Biol. Evol. 2014, 31, 3095–3112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Petersen, G.; Anderson, B.; Braun, H.-P.; Meyer, E.H.; Møller, I.M. Mitochondria in parasitic plants. Mitochondrion 2020, 52, 173–182. [Google Scholar] [CrossRef]
  9. Su, H.J.; Liang, S.L.; Nickrent, D.L. Plastome variation and phylogeny of Taxillus (Loranthaceae). PLoS ONE 2021, 16, e0256345. [Google Scholar] [CrossRef]
  10. Yoshida, S.; Kee, Y.J. Large-scale sequencing paves the way for genomic and genetic analyses in parasitic plants. Curr. Opin. Biotechnol. 2021, 70, 248–254. [Google Scholar] [CrossRef]
  11. Li, X.; Zhang, T.C.; Qiao, Q.; Ren, Z.M.; Zhao, J.Y.; Yonezawa, T.; Hasegawa, M.; Crabbe, M.J.C.; Li, J.Q.; Zhong, Y. Complete chloroplast genome sequence of holoparasite Cistanche deserticola (Orobanchaceae) reveals gene loss and horizontal gene transfer from its host Haloxylon ammodendron (Chenopodiaceae). PLoS ONE 2013, 8, e58747. [Google Scholar] [CrossRef] [Green Version]
  12. Palmer, J.D. Comparative organization of chloroplast genomes. Ann. Rev. Genet. 1985, 19, 325–354. [Google Scholar] [CrossRef] [PubMed]
  13. Li, Y.; Zhou, J.g.; Chen, X.l.; Cui, Y.x.; Xu, Z.c.; Li, Y.h.; Song, J.y.; Duan, B.z.; Yao, H. Gene losses and partial deletion of small single-copy regions of the chloroplast genomes of two hemiparasitic Taxillus species. Sci. Rep. 2017, 7, 12834. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Mogensen, H.L. The hows and whys of cytoplasmic inheritance in seed plants. Am. J. Bot. 1996, 83, 383–404. [Google Scholar] [CrossRef]
  15. Nagata, N. Mechanisms for independent cytoplasmic inheritance of mitochondria and plastids in angiosperms. J. Plant Res. 2010, 123, 193–199. [Google Scholar] [CrossRef]
  16. Barnard-Kubow, K.B.; McCoy, M.A.; Galloway, L.F. Biparental chloroplast inheritance leads to rescue from cytonuclear incompatibility. New Phytol. 2017, 213, 1466–1476. [Google Scholar] [CrossRef] [Green Version]
  17. Zheng, H.; Fan, L.; Wang, T.; Zhang, L.; Ma, T.; Mao, K. The complete chloroplast genome of Populus rotundifolia (Salicaceae). Conserv. Genet. Resour. 2016, 8, 399–401. [Google Scholar] [CrossRef]
  18. Wang, L.; Dong, W.P.; Zhou, S.L. Structural mutations and reorganizations in chloroplast genomes of flowering plants. Acta Bot. Boreali-Occident. Sin. 2012, 32, 1282–1288. (In Chinese) [Google Scholar]
  19. Bungard, R.A. Photosynthetic evolution in parasitic plants: Insight from the chloroplast genome. Bioessays 2004, 26, 235–247. [Google Scholar] [CrossRef]
  20. Gruzdev, E.V.; Mardanov, A.V.; Beletsky, A.V.; Kochieva, E.Z.; Ravin, N.V.; Skryabin, K.G. The complete chloroplast genome of parasitic flowering plant Monotropa hypopitys: Extensive gene losses and size reduction. Mitochondrial DNA B 2016, 1, 212–213. [Google Scholar] [CrossRef] [Green Version]
  21. Guo, X.R.; Zhang, G.F.; Fan, L.Y.; Liu, C.K.; Ji, Y.H. Highly degenerate plastomes in two hemiparasitic dwarf mistletoes: Arceuthobium chinense and A. pini (Viscaceae). Planta 2021, 253, 125. [Google Scholar] [CrossRef]
  22. Petersen, G.; Cuenca, A.; Seberg, O. Plastome evolution in hemiparasitic mistletoes. Genome Biol. Evol. 2015, 7, 2520–2532. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Wolfe, K.H.; Morden, C.W.; Ems, S.C.; Palmer, J.D. Rapid evolution of the plastid translational apparatus in a nonphotosynthetic plant—Loss or accelerated sequence evolution of transfer-rna and ribosomal-protein genes. J. Mol. Evol. 1992, 35, 304–317. [Google Scholar] [CrossRef] [PubMed]
  24. dePamphilis, C.W.; Palmer, J.D. Loss of photosynthetic and chlororespiratory genes from the plastid genome of a parasitic flowering plant. Nature 1990, 348, 337. [Google Scholar] [CrossRef] [PubMed]
  25. Wolfe, K.H.; Morden, C.W.; Palmer, J.D. Function and evolution of a minimal plastid genome from a nonphotosynthetic parasitic plant. Proc. Natl. Acad. Sci. USA 1992, 89, 10648–10652. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Miao, N.; Zhang, L.; Li, M.; Fan, L.; Mao, K. Development of EST-SSR markers for Taxillus nigrans (Loranthaceae) in southwestern China using next-generation sequencing. Appl. Plant Sci. 2017, 5, 1700010. [Google Scholar] [CrossRef] [PubMed]
  27. Doyle, J.J.; Doyle, J.L.; Doyle, J.A.; Doyle, F.J. A rapid DNA isolation procedure for small amounts of fresh leaf tissue. Phytochem. Bull. 1987, 19, 11–15. [Google Scholar]
  28. Bolger, A.M.; Lohse, M.; Usadel, B. Trimmomatic: A flexible trimmer for Illumina sequence data. Bioinformatics 2014, 30, 2114–2120. [Google Scholar] [CrossRef] [Green Version]
  29. Li, H.; Durbin, R. Fast and accurate short read alignment with Burrows–Wheeler transform. Bioinformatics 2009, 25, 1754–1760. [Google Scholar] [CrossRef] [Green Version]
  30. Dierckxsens, N.; Mardulyn, P.; Smits, G. NOVOPlasty: De novo assembly of organelle genomes from whole genome data. Nucleic Acids Res. 2017, 45, e18. [Google Scholar]
  31. Zerbino, D.R.; Birney, E. Velvet: Algorithms for de novo short read assembly using de Bruijn graphs. Genome Res 2008, 18, 821–829. [Google Scholar] [CrossRef] [Green Version]
  32. Kearse, M.; Moir, R.; Wilson, A.; Stones-Havas, S.; Cheung, M.; Sturrock, S.; Buxton, S.; Cooper, A.; Markowitz, S.; Duran, C.; et al. Geneious Basic: An integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics 2012, 28, 1647–1649. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Huang, D.I.; Cronk, Q.C. Plann: A command-line application for annotating plastome sequences. Appl. Plant Sci. 2015, 3, 1500026. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Tamura, K.; Stecher, G.; Peterson, D.; Filipski, A.; Kumar, S. MEGA6: Molecular evolutionary genetics analysis version 6.0. Mol. Biol. Evol. 2013, 30, 2725–2729. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Katoh, K.; Standley, D.M. MAFFT multiple sequence alignment software version 7: Improvements in performance and usability. Mol. Biol. Evol. 2013, 30, 772–780. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Stamatakis, A. RAxML version 8: A tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 2014, 30, 1312–1313. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. He, L.; Qian, J.; Li, X.; Sun, Z.; Xu, X.; Chen, S. Complete chloroplast genome of medicinal plant Lonicera japonica: Genome rearrangement, intron gain and loss, and implications for phylogenetic studies. Molecules 2017, 22, 249. [Google Scholar] [CrossRef]
  38. Braukmann, T.W.A.; Broe, M.B.; Stefanović, S.; Freudenstein, J.V. On the brink: The highly reduced plastomes of nonphotosynthetic Ericaceae. New Phytol. 2017, 216, 254–266. [Google Scholar] [CrossRef] [Green Version]
  39. Han, R.L.; Hao, G.; Zhang, D.X. Interfamilial relationships of Santalales as revealed by chloroplast trnL intron sequences. J. Trop. Subtrop. Bot. 2004, 12, 393–398. (In Chinese) [Google Scholar]
  40. Freyer, R.; Neckermann, K.; Maier, R.M.; Kossel, H. Structural and Functional-Analysis of Plastid Genomes from Parasitic Plants—Loss of an Intron within the Genus Cuscuta. Curr. Genet. 1995, 27, 580–586. [Google Scholar] [CrossRef]
  41. Wu, C.S.; Wang, T.J.; Wu, C.W.; Wang, Y.N.; Chaw, S.M. Plastome Evolution in the Sole Hemiparasitic Genus Laurel Dodder (Cassytha) and Insights into the Plastid Phylogenomics of Lauraceae. Genome Biol. Evol. 2017, 9, 2604–2614. [Google Scholar] [CrossRef] [Green Version]
  42. Wicke, S.; Muller, K.F.; dePamphilis, C.W.; Quandt, D.; Bellot, S.; Schneeweiss, G.M. Mechanistic model of evolutionary rate variation en route to a nonphotosynthetic lifestyle in plants. Proc. Natl. Acad. Sci. USA 2016, 113, 9045–9050. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Ruhlman, T.A.; Chang, W.J.; Chen, J.J.W.; Huang, Y.T.; Chan, M.T.; Zhang, J.; Liao, D.C.; Blazier, J.C.; Jin, X.H.; Shih, M.C.; et al. NDH expression marks major transitions in plant evolution and reveals coordinate intracellular gene loss. BMC Plant Biol. 2015, 15, 100. [Google Scholar] [CrossRef] [Green Version]
  44. Xue, Z.Q.; Xue, J.H.; Victorovna, K.M.; Ma, K.P. The complete chloroplast DNA sequence of Trapa maximowiczii Korsh. (Trapaceae), and comparative analysis with other Myrtales species. Aquat. Bot. 2017, 143, 54–62. [Google Scholar] [CrossRef]
  45. Li, X.; Yang, J.B.; Wang, H.; Song, Y.; Corlett, R.T.; Yao, X.; Li, D.Z.; Yu, W.B. Plastid NDH Pseudogenization and Gene Loss in a Recently Derived Lineage from the Largest Hemiparasitic Plant Genus Pedicularis (Orobanchaceae). Plant Cell Physiol. 2021, 62, 971–984. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Gene map of T. nigrans and S. parasitica chloroplast genome (the darker gray in the inner circle corresponds to GC content, while the lighter gray corresponds to AT content).
Figure 1. Gene map of T. nigrans and S. parasitica chloroplast genome (the darker gray in the inner circle corresponds to GC content, while the lighter gray corresponds to AT content).
Genes 14 00560 g001
Figure 2. Comparison of the chloroplast genomes of the 12 species, using T. nigrans as a reference in the mVISTA program. (Grey arrows and thick black lines above the alignment indicate the orientation of genes. A cut-off of 70% identity was used for the plots. The Y-scale axis represents the percent identity within 50%–100%. Genomic regions are color-coded as either protein-coding exons, rRNAs, tRNAs, or conserved non-coding sequences (CNS), the same as below).
Figure 2. Comparison of the chloroplast genomes of the 12 species, using T. nigrans as a reference in the mVISTA program. (Grey arrows and thick black lines above the alignment indicate the orientation of genes. A cut-off of 70% identity was used for the plots. The Y-scale axis represents the percent identity within 50%–100%. Genomic regions are color-coded as either protein-coding exons, rRNAs, tRNAs, or conserved non-coding sequences (CNS), the same as below).
Genes 14 00560 g002
Figure 3. Comparison of the chloroplast genomes of the 12 species, using S. parasitica as a reference in mVISTA program.
Figure 3. Comparison of the chloroplast genomes of the 12 species, using S. parasitica as a reference in mVISTA program.
Genes 14 00560 g003
Figure 4. Comparison of border distance between adjacent genes and junctions of the LSC, SSC, and two IR regions among the chloroplast genomes of the 12 species of Santalales (The figure shows relative changes at or near the IR/SSC borders).
Figure 4. Comparison of border distance between adjacent genes and junctions of the LSC, SSC, and two IR regions among the chloroplast genomes of the 12 species of Santalales (The figure shows relative changes at or near the IR/SSC borders).
Genes 14 00560 g004
Figure 5. Maximum likelihood phylogenetic tree of Santalales based on 39 protein-coding genes. (Hemiparasitic plants are labeled with circles (○), and autotrophic plants are marked with white squares (□)).
Figure 5. Maximum likelihood phylogenetic tree of Santalales based on 39 protein-coding genes. (Hemiparasitic plants are labeled with circles (○), and autotrophic plants are marked with white squares (□)).
Genes 14 00560 g005
Table 1. Elemental composition in the T. nigrans and S. parasitica chloroplast genomes.
Table 1. Elemental composition in the T. nigrans and S. parasitica chloroplast genomes.
Species T(U) (%)C (%)A (%)G (%)Length (bp)
T. nigransLSC33.418.031.816.870,181
SSC41.213.332.513.06100
IRa28.820.628.322.322,569
IRb28.322.328.720.722,569
Total32.019.030.618.4121,419
S. parasiticaLSC33.517.731.916.770,270
SSC41.512.832.613.16106
IRa28.820.528.322.422,687
IRb28.322.428.820.522,687
Total32.118.930.718.3121,750
Table 2. List of genes in the T. nigrans and S. parasitica chloroplast genomes.
Table 2. List of genes in the T. nigrans and S. parasitica chloroplast genomes.
Genes
1Photosystem IpsaA, B, C, I, J, ycf3, 4
2Photosystem IIpsbA, B, C, D, E, F, H, I, J, K, L, M, N, T, Z
3Cytochrome b6/fpetA, B, D, G, L, N
4ATP synthaseatpA, B, E, F, H, I
5RubiscorbcL
6RNA polymeraserpoA, B, C1, C2
7Small subunit ribosomal proteinsrps2, 3, 4, 7, 8, 11, 12, 14, 18, 19
8Large subunit ribosomal proteinsrpl2, 14, 20, 22, 23, 33, 36
9Other proteinsaccD, ccsA, cemA, clpP, matK
10Proteins of unknown functionycf2
11Ribosomal RNAsrrn4.5S, rrn5S, rrn16S, rrn23S
12Transfer RNAstrnC(GCA), trnD(GUC), trnE(UUC), trnF(GAA), trnfM(CAU), trnG(GCC), trnI(CAU), trnL(CAA), trnL(UAG), trnM(CAU), trnN(GUU), trnP(UGG), trnQ(UUG), trnR(ACG), trnR(UCU), trnS(GGA), trnS(GCU), trnS(UGA), trnT(GGU), trnT(UGU), trnV(GAC), trnW(CCA), trnY(GUA)
Table 3. Comparative Chloroplast Genomic Analysis in Santalales.
Table 3. Comparative Chloroplast Genomic Analysis in Santalales.
SpeciesE. scandensS. jasminodoraS. parasiticaT. chinensisT. nigransT. sutchuenensisC. manillanaO. albaV. albumV. coloratumV. crassulaeV. minimum
FamilyErythropalaceaeSchoepfiaceaeLoranthaceaeLoranthaceaeLoranthaceaeLoranthaceaeOpiliaceaeSantalaceaeSantalaceaeSantalaceaeSantalaceaeSantalaceae
Accession No.NC_036759.1NC_034228.1MH101514NC_036306.1MH095982NC_036307.1NC_034931.1NC_027960.1NC_028012.1NC_035414.1NC_027959.1NC_027829.1
Genome size (bp)156,154118,743121,750121,363121,419122,562147,461147,253128,921128,744126,064131,016
LSC length (bp)84,79984,16870,27070,35770,18170,63083,50584,46673,89373,68473,22575,814
LSC length (%)54.370.957.758.057.857.656.657.457.357.258.157.9
SSC length (bp)18,56797636106608261006102780613,9728632863086289014
SSC length (%)11.98.25.05.05.05.05.39.56.76.76.86.9
IR length (bp)26,39412,40622,68722,46222,56922,91528,07524,34023,19823,21522,10523,094
IR length (%)33.820.937.337.037.237.438.133.136.036.135.135.2
GC content (%)38.038.137.237.337.437.337.437.736.436.336.436.2
Number of genes130121106106106106120117115119116104
Number of protein coding genes867266666666737471707366
Number of tRNAs363428282828373536363529
Number of rRNAs888888888888
Table 4. Codon–anticodon recognition patterns and codon usage of the T. nigrans and S. parasitica chloroplast genomes.
Table 4. Codon–anticodon recognition patterns and codon usage of the T. nigrans and S. parasitica chloroplast genomes.
SpeciesAmino AcidCodonNo.RSCUtRNAAmino AcidCodonNo.RSCUtRNAAmino AcidCodonNo.RSCUtRNA
T. nigransPheUUU18211.22 TyrUAU12271.38 StopUAG5840.75
PheUUC11530.78trnF-GAATyrUAC5450.62trnY-GUALeuUUG8631.3trnL-CAA
LeuUUA8771.32 StopUAA9611.24 LeuCUC4970.75
LeuCUU8071.21 HisCAU7731.41 HisCAC3210.59
LeuCUA5790.87trnL-UAGGlnCAA7851.33trnQ-UUGAsnAAC6000.59trnN-GUU
LeuCUG3700.56 GlnCAG3970.67 AsnAAU14191.41
IleAUU13871.21 IleAUC9050.79 IleAUA11481
AspGAU7891.44 LysAAA16171.37 LysAAG7480.63
MetAUG7261trn(f)M-CAU
trnI-CAU
ValGUC3690.74trnV-GACAspGAC3040.56trnD-GUC
ValGUU6591.32 ValGUA5991.2 ValGUG3700.74
SerUCC5931.03trnS-GGAGluGAA9971.37trnE-UUCCysUGC3160.74trnC-GCA
SerUCU8161.42 CysUGU5351.26 GluGAG4550.63
SerUCG4900.85 StopUGA7771 ProCCU5361.08
ProCCC4991.01 ArgCGC2280.56 ArgCGA4611.14
TrpUGG5811trnW-CCAArgCGU2670.66trnR-ACGSerUCA6841.19trnS-UGA
ProCCA5891.19trnP-UGGThrACC4781.01trnT-GGUSerAGC3670.64trnS-GCU
ProCCG3580.72 ArgCGG3010.75 SerAGU4910.86
ThrACU5321.13 ArgAGG4661.15 ThrACG3310.7
ThrACA5451.16trnT-UGUArgAGA7001.73trnR-UCUGlyGGC2670.63trnG-GCC
AlaGCU3761.28 AlaGCC2770.94 AlaGCA3141.07
AlaGCG2060.7 GlyGGA5271.23 GlyGGU4441.04
GlyGGG4691.1
S. parasiticaPheUUU18421.23 TyrUAU11571.33 StopUAA9851.32
PheUUC11510.77trnF-GAATyrUAC5880.67trnY-GUAArgAGA7301.76trnR-UCU
LeuUUA8831.32 StopUAG5420.72 LeuCUU7771.16
LeuCUC4770.71 HisCAC3250.66 HisCAU6641.34
LeuCUA5780.86trnL-UAGGlnCAA7381.35trnQ-UUGLeuUUG9091.35trnL-CAA
LeuCUG4030.6 GlnCAG3540.65 IleAUU14681.27
IleAUC7810.67 AsnAAU14381.39 LysAAG7320.6
MetAUG7011trn(f)M-CAU
trnI-CAU
AsnAAC6240.61trnN-GUUGluGAA9411.34trnE-UUC
IleAUA12231.06 LysAAA17251.4 ValGUA5711.19
ValGUU6731.4 AspGAU8321.4 ValGUG3500.73
ValGUC3280.68trnV-GACAspGAC3600.6trnD-GUCSerUCA7091.15trnS-UGA
SerUCU8611.4 CysUGU6031.28 GluGAG4680.66
SerUCC6851.11trnS-GGACysUGC3360.72trnC-GCATrpUGG5241trnW-CCA
SerUCG5080.83 StopUGA7160.96 ArgCGA4571.1
ProCCU4801.01 ArgCGU3050.74trnR-ACGProCCA5691.2trnP-UGG
ProCCC5191.1 ProCCG3240.68 ArgCGC2170.52
ThrACU5101.11 SerAGU5400.88 ArgCGG3170.77
ThrACC4701.03trnT-GGUSerAGC3880.63trnS-GCUGlyGGC2780.65trnG-GCC
ThrACA5301.16trnT-UGUAlaGCC2790.94 ArgAGG4601.11
ThrACG3210.7 AlaGCG1860.62 AlaGCA3521.18
AlaGCU3741.26 GlyGGA5631.31 GlyGGU4210.98
GlyGGG4621.07
Table 5. Genes with introns in the T. nigrans and S. parasitica chloroplast genomes, including the exon and intron length.
Table 5. Genes with introns in the T. nigrans and S. parasitica chloroplast genomes, including the exon and intron length.
SpeciesGeneLocationExon I(bp)Intron I(bp)Exon II(bp)Intron II(bp)Exon III(bp)
T. nigransatpFLSC150786390
rpoC1LSC4507561602
ycf3LSC127755230785153
rps12 *LSC, IR114-23254326
petBLSC6798642
petDLSC9696483
rpl2LSC, IR394652434
S. parasiticaatpFLSC150759390
rpoC1LSC4567781626
ycf3LSC153712230759127
petBLSC6700642
petDLSC9652483
rpl2LSC, IR394665431
rps12 *LSC, IR114-23253426
* The rps12 gene is divided into 5’-rps12 in the LSC region and 3’-rps12 in the IR region.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yue, X.; Miao, N.; Fan, Z.; Mao, K. Comparative Analyses of Chloroplast Genomes for Parasitic Species of Santalales in the Light of Two Newly Sequenced Species, Taxillus nigrans and Scurrula parasitica. Genes 2023, 14, 560. https://doi.org/10.3390/genes14030560

AMA Style

Yue X, Miao N, Fan Z, Mao K. Comparative Analyses of Chloroplast Genomes for Parasitic Species of Santalales in the Light of Two Newly Sequenced Species, Taxillus nigrans and Scurrula parasitica. Genes. 2023; 14(3):560. https://doi.org/10.3390/genes14030560

Chicago/Turabian Style

Yue, Ximing, Ning Miao, Zilu Fan, and Kangshan Mao. 2023. "Comparative Analyses of Chloroplast Genomes for Parasitic Species of Santalales in the Light of Two Newly Sequenced Species, Taxillus nigrans and Scurrula parasitica" Genes 14, no. 3: 560. https://doi.org/10.3390/genes14030560

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop