Next Article in Journal
Overexpression of BQ323636.1 Modulated AR/IL-8/CXCR1 Axis to Confer Tamoxifen Resistance in ER-Positive Breast Cancer
Previous Article in Journal
The Influence of an Acute Bout of Aerobic Exercise on Vascular Endothelial Function in Moderate Stages of Chronic Kidney Disease
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Complete Chloroplast Genome Sequence of Fagus longipetiolata Seemen (Fagaceae): Genome Structure, Adaptive Evolution, and Phylogenetic Relationships

Institute for Forest Resources & Environment of Guizhou, Key Laboratory of Forest Cultivation in Plateau Mountain of Guizhou Province, College of Forestry, Guizhou University, Guiyang 550025, China
*
Author to whom correspondence should be addressed.
Life 2022, 12(1), 92; https://doi.org/10.3390/life12010092
Submission received: 25 November 2021 / Revised: 26 December 2021 / Accepted: 5 January 2022 / Published: 9 January 2022
(This article belongs to the Section Genetics and Genomics)

Abstract

:
Fagus longipetiolata Seemen is a deciduous tree of the Fagus genus in Fagaceae, which is endemic to China. In this study, we successfully sequenced the cp genome of F. longipetiolata, compared the cp genomes of the Fagus genus, and reconstructed the phylogeny of Fagaceae. The results showed that the cp genome of F. longipetiolata was 158,350 bp, including a pair of inverted repeat (IRA and IRB) regions with a length of 25,894 bp each, a large single-copy (LSC) region of 87,671 bp, and a small single-copy (SSC) region of 18,891 bp. The genome encoded 131 unique genes, including 81 protein-coding genes, 37 transfer RNA genes (tRNAs), 8 ribosomal RNA genes (rRNAs), and 5 pseudogenes. In addition, 33 codons and 258 simple sequence repeats (SSRs) were identified. The cp genomes of Fagus were relatively conserved, especially the IR regions, which showed the best conservation, and no inversions or rearrangements were found. The five regions with the largest variations were the rps12, rpl32, ccsA, trnW-CCA, and rps3 genes, which spread over in LSC and SSC. The comparison of gene selection pressure indicated that purifying selection was the main selective pattern maintaining important biological functions in Fagus cp genomes. However, the ndhD, rpoA, and ndhF genes of F. longipetiolata were affected by positive selection. Phylogenetic analysis revealed that F. longipetiolata and F. engleriana formed a close relationship, which partially overlapped in their distribution in China. Our analysis of the cp genome of F. longipetiolata would provide important genetic information for further research into the classification, phylogeny and evolution of Fagus.

1. Introduction

The Fagus genus is mainly composed of deciduous trees and is a genus of the Fagaceae family [1]. Fagus species have important economic value and are important components of temperate and subtropical deciduous broad-leaved forests in the Northern Hemisphere; there are approximately 10–13 species of the genus worldwide, including five to seven species found in China [2,3]. Because Fagus tree species are widely distributed around Eurasia, the study of the evolutionary history of the genus is conducive to revealing the impact of global climate change on vegetation, as well as the geological and biological connections of this group [4]. Most of the previous studies on the classification of Fagus were based on external morphological characteristics, such as the total stem length, involucre bract type and shape, and leaf textural and structural characteristics [5,6]. However, Fagus is a relatively specialized group with a rich diversity and complex genetic relationships [7]. The external morphological characteristics of the members of the genus Fagus often vary greatly under the influence of environmental factors, and some characteristics are crossed, leading to confusion in their classification [5]. There is still a debate about the number of species in the genus.
In recent years, genome sequencing has been widely used to analyze the genetic variability and evolution of species [8]. Chloroplasts (cp), mitochondria, and nuclei contain independent genomes, which can provide important genetic information for phylogenetic analysis [9]. Chloroplast cells are unique plant cells and the main site of photosynthesis; they have a complete cp genome independent of the nuclear genome [10]. The cp genome is maternally dependent and presents the advantages of a shorter length, easier extraction and purification, a highly conserved sequence and a large number of simple sequence repeat (SSR) loci [11]. When compared with the mitochondrial genome, variations in cp genome size in different plants are relatively low (the mitochondrial genome size of most angiosperms is 300–600 kb, and the cp genome size is 115–165 kb) [12]. Therefore, the cp genome has more important reference value for revealing the origin and evolution of species and determining the genetic relationships between different species. Thus far, the cp genomes of Fagus engleriana, Fagus sylvatica, Fagus crenata, and Fagus japonica var. multinervis have been published [13,14,15,16]. However, no complete genome sequence of Fagus longipetiolata, the Fagus species with the most widespread distribution in China [17], has been reported.
In this study, we reported the whole chloroplast genome of F. longipetiolata and compared it with the published chloroplast genome of Fagus. Moreover, its annotations, codon bias, scattered repeat sequences, and SSRs were analyzed. Our data will be a valuable genetic resource for the study of Fagus species.

2. Materials and Methods

2.1. Sampling, DNA Extraction and Genome Sequencing

Leaves were taken from the F. longipetiolata seedlings cultivated at Guizhou University, Guizhou Province, China (26°4.504′ N, 106°6.568′ E), and lodged a voucher specimen (accession number FL-GZU-001) in the Institute for Forest Resources & Environment of Guizhou at Guizhou.
A Plant Genomic DNA Kit (TIANGEN, Beijing, China) was used to extract total genomic DNA from 100 mg of the leaves. The purified DNA was then fragmented by mechanical disruption (sonication). Then, the paired-end (PE) library was constructed using VAHTS Multiplex Oligos set 4 for Illumina (Vazyme, Nanjing, China) and VAHTS Universal DNA Library Prep Kit for Illumina V3 (Vazyme, Nanjing, China), according to the manufacturer’s protocols. Finally, the qualified libraries were sequenced on the Illumina platform, according to the paired-end PE150 sequencing strategy. Approximately 6 Gb of raw data were sequenced. All of the above works were conducted by Genepioneer Biotechnologies Co. Ltd. (Nanjing, China).

2.2. Initial Assembly and Annotation of the cp Genome

The cp genome of F. longipetiolata was assembled using SPAdes software (v3.10.1) [18] with k-mers of 55, 87, and 121, and the assembly was independent of the reference genome. After assembly, quality control was carried out according to the sequence of F. japonica (accession no. MT762295) [19]. We used two methods to annotate the cp genome to improve the accuracy of annotation. First, Prodigal software (v2.6.3) was used to annotate the cp coding sequences (CDSs), Hmmer software (v3.1b2) was used to predict ribosomal RNAs (rRNAs), and Aragorn software (v1.2.38) was used to predict transfer RNAs (tRNAs). Second, the gene sequences were extracted according to the sequences of related species already published in NCBI, and BLAST v2.6 was used to compare the assembled sequence to obtain the second annotation result. Then, the results of the two annotations for different genes were checked manually, erroneous and redundant annotations were removed, and multiexon boundaries were determined to obtain the final annotation. We then mapped the entire genome using OGDRAW software [20].

2.3. Codon Usage and Repeat Sequence Analysis

According to the CDSs of 81 protein-coding genes, unique CDSs (one copy selected for CDSs with multiple copies) were screened by using Perl scripts, and the RSCU of each codon was estimated with CodonW 1.4.2 software [21].
The forward repeat sequences, reverse repeat sequences, complementary repeat sequences, and palindromic repeat sequences were analyzed by online Vmatch v2.3.0 software and Perl scripts, with a minimal repeat size of 30 bp and a Hamming distance of 3. MISA v1.0 software was used to search for SSR markers in the cp genome, setting the minimum number of mononucleotide repeats to 8, the minimum number of dinucleotide repeats to 5, and the minimum number of trinucleotide, tetranucleotide, pentanucleotide, and hexanucleotide SSR repeats to 3 [22].

2.4. Genome Comparison

Four reported cp genomes of Fagus and one exogenous species were downloaded from the NCBI database (F. sylvatica (MK598696), F. engleriana (KX852398), F. crenata (MH171101), F. japonica var. multinervis (MN894556), and Arabidopsis thaliana (AP000423)). The cp genome structures of six plants were analyzed with CGView software [23]. The homology and collinearity of the cp sequences were analyzed with Mauve software [24]. MAFFT software (--auto mode) [25] was used for a global comparison of homologous gene sequences of different plants. DNAsp 5.0 [26] was used to calculate the Pi value of each gene. The boundary information of inverted repeat (IR), small single-copy (SSC), and large single-copy (LSC) regions was visualized by using the SVG module in Perl.

2.5. Adaptive Evolution and Phylogenetic Analyses

Based on the five cp genomes of Fagus used in this study, the Ka/Ks values of each functional protein-coding gene were calculated by KaKs_Calculator v2.0 software with the default settings [27].
We constructed a phylogenetic tree with the newly sequenced F. longipetiolata cp genome and 19 cp genomes (from 1 family, 5 tribes, and 1 outgroup (Populus trichocarpa)) downloaded from NCBI. MAFFT (v7.427, --auto mode) was used for multiple sequence alignment [25]. Then, the aligned data were analyzed by RAxML v8.2.10 with the maximum likelihood (MJ) method to construct a phylogenetic tree (1000 bootstraps) [28,29].

3. Results

3.1. Features of the F. longipetiolata Chloroplast Genome

A total of 20,928,581 paired-end reads were obtained from the Illumina NovaSeq platform, and the Q20 and Q30 values were 97.53 and 92.98%, respectively. The complete cp genome sequence of F. longipetiolata was assembled de novo and uploaded to the NCBI database (GenBank accession number MZ562567). The cp genome of F. longipetiolata was a 158,350 bp long circular genome, including a pair of reverse repeats, IRA and IRB (25,894 bp), a small single copy region (SSC, 18,891 bp), and a large single copy region (LSC, 87,671 bp) (Table 1 and Figure 1). The GC content of the IR region sequence was the highest (42.70%), while the GC content of the SSC region was the lowest (31.19%). The average GC content of the whole genome was 37.09%. There were 131 predicted functional genes in the F. longipetiolata cp genome, including 81 protein-coding genes, 37 tRNA genes, 8 rRNA genes, and 5 pseudogenes.
Subsequently, we annotated the assembled genes, and all the genes were annotated with gene functions (Table 2). These genes belonged to four types: photosynthesis-related; self-replication-related; genes of unknown function; and maturase (matK), protease (clpP), and other genes. A total of 18 of the annotated genes were double-copy genes, including three protein-coding genes, seven tRNAs, and eight rRNAs. Fifteen genes (ndhB, petB, petD, ndhA, atpF, trnK-UUU, trnL-UAA, trnA-UGC, rpl16, rpl2, rps12, rpoC1, trnG-UCC, trnI-GAU, and trnV-UAC) had one intron each, and two genes (ycf3, clpP) comprised two introns each (Table 2 and Table S1). The longest intron (2524 bp) was located in the turnK-UUU gene, which completely encompassed the matK gene, and the smallest intron (535 bp) was found in the trnL-UAA gene.

3.2. Codon Usage Bias

There are great differences in codon usage among different species and organisms because each amino acid corresponds to at least one codon and, at most, six codons [30]. This inequality of synonymous codon usage is referred to as codon preference (RSCU) [31,32]. Natural selection, species mutation, and genetic drift are considered to be the reasons for this preference [33,34,35]. We screened the unique CDSs and calculated the codon preference (Figure 2 and Table S2). The results showed that the CDS of the F. longipetiolata cp genome encoded a total of 24,169 amino acids (including stop codons). Leucine was the most abundant, with 2553 codons (10.56%) detected, followed by isoleucine (2108 codons, 8.72%), and serine (1863 codons, 7.71%), while the rarest amino acid was cysteine (276 codons, 1.14%). Among the codons, 33 (60.94%) preferred codons (RSCU > 1). Twenty-nine preferred codons ended in A/U, but the most preferred codon was AUG, encoding methionine (Met), with an RSCU value of 2.9892.

3.3. Detection of Chloroplast Repeat Sequences and SSRs

In the F. longipetiolata cp genome, we discovered 37 repeat sequences. Palindromic repeats were the most common type (17 repeats), accounting for 45.95% of all the repeats, followed by forward (14 repeats, approximately 37.84%), reverse (five repeats, approximately 13.51%), and complementary (one repeat, approximately 2.70%) repeats (Figure 3 and Table S3). A total of 30 repeats were between 30–38 bp in length, and the other repeats were within 40–46 bp. The LSC region had the greatest number of repetitions (26 repeats), followed by the IRs (16 repeats), and the SSC region (seven repeats). In addition, most of the repeats were located in genes (24, 64.87%), and a minority were found in intergenic spacer regions (21, 56.76%).
We detected 258 SSR loci in the F. longipetiolata cp genome. The SSRs were largely distributed in the LSC region (176, 68.22%), followed by the IR regions (46 SSRs), and the SSC region (36 SSRs) (Figure 4 and Table S4). Additionally, 159 SSRs were located in intergenic spaces, and 99 SSRs were located in genes such as matK, atpF, trnG-UCC, trnK-UUU, atpI, rpoB, rpoC2, rpoC1, psbC, rps14, psaB, psaA, ycf3, trnL-UAA, and ndhK. The SSRs consisted of 12 complex nucleotide repeats, 156 mononucleotides, 12 dinucleotides, 66 trinucleotides, 5 tetranucleotides, 5 pentanucleotides, and 2 hexanucleotides, overall. Polyadenine (poly A) and polythymine (poly T) repetitions made up the majority of mononucleotide SSRs (95.51%), whereas C and G mononucleotides were uncommon (4.49%). The larger proportion of A or T bases in certain cpSSRs corresponded to the total A/T content (62.91%) of the F. longipetiolata cp genome.

3.4. Comparison of Complete Chloroplast Genomes

The sequences from six cp genomes were compared using multigenome comparative analysis, employing the cp genome of F. longipetiolata as the reference genome (Figure 5). The cp genomes of these six plants (A. thaliana, F. crenata, F. engleriana, F. japonica var. multinervis, F. longipetiolata, and F. sylvatica) ranged in length from 154,478 bp to 158,462 bp, according to the findings. There was a high degree of similarity between F. longipetiolata and the other four Fagus cp genome sequences. They also had similarity with A. thaliana in IR regions (90~110 kbp and 130~155 kbp, Figure 5) but showed heterogeneity in other regions.
The nucleotide diversity (Pi) values of 106 loci in the chloroplast genome of F. longipetiolata glauca and four other Fagus plants (F. engleriana, F. crenata, F. sylvatica, and F. japonica var. multinervis) were calculated to determine the divergent hotspots (Figure 6). The minimum and maximum values of the entire genome sequence were between 0 and 0.01345, and the average value was 0.00099. The SSC area showed the maximum nucleotide diversity (average Pi = 0.00262), followed by the LSC region (average Pi = 0.00090), and the IR regions had the lowest Pi value (average Pi = 0.00009), indicating that the IR regions were substantially more conserved. In addition, five highly divergent regions were detected, including rps12 (0.01345), rpl32 (0.00641), ccsA (0.00543), trnW-CCA (0.00541), and rps3 (0.00450). The LSC region contained three of these divergent regions (rps12, trnW-CCA, and rps3), and the SSC region contained two divergent regions (rpl32 and ccsA).

3.5. IR Expansion and Contraction

We analyzed the binding regions of IR/LSC and IR/SSC of F. longipetiolata and five reference cp genomes (F. sylvatica, MK598696; F. engleriana, KX852398; F. crenata, MH171101; F. japonica var. multinervis, MN894556; A. thaliana, AP000423), as well as the length of genes located in the binding region. The genes located at the binding regions of the LSC/ IRB, IRB/SSC, SSC/IRA, and IRA/LSC regions were rsp19, rpl2, ycf1, ndhF, ycf1, trnN, rpl2, and trnH, respectively (Figure 7). The locations of the rps19 genes of all Fagus plants were similar, occurring in the LSC region, 8–10 bp distant from the binding regions between the LSC and IRB, differing from the rps19 gene location in A. thaliana, which spanned the LSC and IRB binding regions. This may be the reason that the rpl2 gene of A. thaliana (166 bp) was farther from the LSC and IRB binding regions than that of Fagus (65–67 bp). The trnH gene was found to be located in the LSC region of Fagus species and 22 bp from the IRA/LSC boundary, except for F. engleriana (16 bp) and F. crenata (24 bp). The genes at IR-SSC junctions in all species were ycf1 genes. The ndhF genes of all species were located mainly in the SSC region but also crossed the IRB/SSC boundary to some extent. The ndhF genes of Fagus extended into the IRB region by 13–14 bp, differing from that in A. thaliana by 37 bp. Based on the above results, the IRs and two SC regions of the five species of Fagus, in which the numbers and sequences of genes were conserved, showed slight differences at the borders.

3.6. Adaptive Evolution Analysis

Using F. longipetiolata as a reference, synonymous and nonsynonymous alterations in the five Fagus cp genomes were examined to uncover patterns of selection among protein-coding genes (Figure 8). In the five cp genomes, the Ka/Ks ratios of 80 protein-coding genes were determined by comparison. The ratio of Ka to Ks of most coding genes was less than one or could not be computed because one of the Ka or Ks values was zero, indicating that they were relatively conserved; in particular, the Ka/KS values of all the genes of F. longipetiolata and F. engleriana glauca were less than one. However, the Ka/Ks values of the rpoA gene between F. longipetiolata and F. sylvatica, the ndhF gene between F. longipetiolata and F. japonica var. multinervis, and the ndhD gene between F. longipetiolata and F. crenata were greater than one.

3.7. Phylogenetic Inference

The cp genome is of great significance for system development research [36]. To determine the phylogenetic status of F. longipetiolata within Fagaceae, a phylogenetic tree was constructed using the ML method, using cp sequences of 18 Fagaceae species, with Populus trichocarpa as the outgroup (Figure 9). The analysis showed that the phylogenetic tree had a total of 18 nodes, 15 of which presented support rates ≥86%, and 10 presented support rates of 100%, which indicated that the reliability of the clustering results was high. In the phylogenetic tree, the 20 species of plants could be divided into two large groups and six small groups. P. trichocarpa of Salicaceae was located in one large group, and the 19 species of Fagaceae were in the other group. Moreover, the 19 Fagaceae species could be divided into Fagus, Quercus, Castanea, Castanopsis, and Trigonobalanus. Within the Fagaceae family, Fagus and Trigonobalanus were sister groups with high credibility. F. longipetiolata and F. engleriana were located on the same branch with 100% support, and this small branch belonged to the same branch as F. japonica var. multinervis. In addition, F. sylvatica was relatively distantly related to the other four Fagus species. This result shows that F. longipetiolata is highly homologous with F. engleriana but has a distant relationship with other plants in the genus.

4. Discussion

In general, the complete cp genome of F. longipetiolata showed great similarities to the other reported cp genomes of Fagus plants in terms of genome length, structure, and gene composition. No rearrangement phenomenon was observed, and a good collinearity relationship was found. Thus, the cp genome of Fagus is relatively well conserved [13,14]. Nevertheless, we observed minor differences in IR/SC border areas, which might be due to IR contraction and expansion. The contraction and expansion of the IR region is a common phenomenon in the process of evolution [37], and it is also the main reason for the differences in cp genome length [38]. However, the expansion and contraction of IR boundaries has not been shown to cause the transfer, gain or deletion of genes in the SC and IR regions of the cp genome, and consistent findings were obtained in the genomes of Fagaceae [13]. In angiosperms, the pseudogenes ycf1 or rps19 are produced by contraction and expansion of the IR region [39]. The rps19 gene usually crosses the boundary between LSC/IR and SSC/IR [39,40]. In Fagus, the rps19 coding gene was located in the LSC region, which is consistent with the results of other Fagaceae plants [41]. In this study, ycf1 across the junction of IR/SSC, indicating that ycf1 gene has no phylogenetic significance [42].
Nucleotide diversity (Pi) can indicate the magnitude of variation in various species’ nucleic acid sequences, and locations with higher variability can be used as molecular markers in population genetics [43,44]. In this study, the results of nucleotide diversity (Pi) assessment showed that the gene sequences of the LSC/SSC region were more variable than those in the IR region, which was consistent with the results found in other genera [13,41,45,46,47]. The same conclusion has been reached in the study of Lagerstroemia and Adrinandra plants [48,49]. Through cp genome sequence variation analysis, we discovered five hypervariable regions in the LSC (rps12, trnW-CCA, and rps3) and SSC regions (rpl32 and ccsA). A previous study showed that point mutations in conserved regions of the rps12 gene would affect the folding of 16S rRNA and the interaction with streptomycin in Nicotiana plumbaginifolia [50]. Moreover, the deletion/transfer of the rpl32 gene in the plastid genome offers crucial phylogenetic data for the monophyletic evolution of the Thalictroideae subfamily [51]. It is also considered to be an ideal genetic marker for new Glycine varieties and Diospyros species [52,53]. The ccsA gene has been considered a locus to understand the evolution of the cp genome in Litsea [54], Pterocarpus [55], and Prosopis genera [56]. In this study, the Pi values of the five divergent regions were higher than 0.004, corresponding to highly variable regions. We suggest that the rps12, rpl32, ccsA, trnW-CCA, and rps3 genes be used to study the molecular phylogeny of Fagus. They might be applicable for further analyses of phylogenetic relationships and population genetics and for species identification in Fagus.
The study of codon preference can not only aid in the interpretation of species evolution but can also be used to optimize the expression of foreign genes and to predict gene functions and gene expression levels [57]. In plant cp genomes, codons tend to end in an A or U base [58]. In this study, we found 33 high-frequency codons in F. longipetiolata dentata, 29 of which ended in A or U. This result may be caused by natural selection and mutation [59]. In addition, the amount of leucine was the highest and that of cysteine was the lowest among the amino acids. The same results were obtained in previous studies on the cp genomes of angiosperms [60].
In the evolution of species, as well as the inheritance and variation of genes within species, repeated sequences play a significant role [61,62]. In general, most of the repeated sequences in the genome are distributed in noncoding regions because of the process of species evolution. An organism retains the smallest amount of genetic information as much as possible to improve its genetic efficiency [63]. In this study, a total of 38 repeat sequences were discovered in the cp genome of F. longipetiolata, the majority of which were found in genes. It was indicated that the cp genome of F. longipetiolata retained abundant genetic information. SSRs of the plant plastid genome have been frequently employed in phylogenetic investigations on account of their unique maternal genetic advantages [64,65]. We found a total of 258 SSR sites in the cp genome of F. longipetiolata, among which single nucleotide repeats and trinucleotide repeats were the most common, with frequencies of 60.47 and 25.58%, respectively. These results were consistent with previous studies reported in the F. crenata and F. engleriana cp genomes [14]. It was suggested that the single nucleotide repetition of Fagus plants may play a more important role than other SSRs in genetic variation, which is similar to that of Lythraceae species [40]. In this study, the single nucleotides within the SSRs were almost A/T bases (95.51%), and AT/TA accounted for the majority of the two- to six-base repeats. F. longipetiolata had a relatively high A or T content and A/T polymerization at the SSR sites of cpDNA. These might be reasons for the high content of AT in the cp genome of F. longipetiolata, similar to the rich results of AT in other cp genomes [66,67]. Consistent with previous reports [68,69], the SSRs found in the cp genome of F. longipetiolata were mainly located in the LSC region and were enriched in the non-coding region.
If a base mutation leads to an amino acid change, it is referred to as a nonsynonymous mutation; otherwise, it is a synonymous mutation, and nonsynonymous mutations are usually influenced by natural selection [70]. The selection effect of genes is usually expressed by the ratio of Ka to Ks. When Ka/Ks is greater than one, it indicates a positive selection effect, and when Ka/Ks is less than one, it indicates a purification selection effect [71]. In this study, the Ka/Ks of most genes (77 of 80) was less than one in the comparisons between F. longipetiolata and the other four Fagus species, showing that purifying selection plays an important role in the cp genes of the five species of Fagus species. However, in the three control groups, the Ka/Ks ratios of the ndhD, rpoA, and ndhF genes were greater than one, which showed that the three genes of F. longipetiolata were positively selected to adapt to the living environment. Positive selection of ndhD (Pterocarpus and Leguminosae) [50,72], rpoA (Trifolium alexandrinum and Trifolium resupinatum), and ndhF (T. alexandrinum, T. resupinatum and Cardamine genus) [73,74] genes has also been reported in previous studies. It is also reported that NADPH dehydrogenase genes (ndhD and ndhA) tend to evolve at a higher rate than other genes [75]. The low Ka/Ks ratio (ka/ks = 0) of the ClpP gene in this study may be due to the fact that the ClpP gene contains two introns, which is similar to that of Zingibereae [42]. The evolution rate of the ClpP gene is species-specific, which would lead to the loss of introns in the process of rapid evolution [76].
Here, phylogenetic analysis relying on the cp genome revealed that Fagus is a sister genus of all other Fagaceae plants and forms a monophyletic branch, which was similar to the results of previous studies [15,16,77]. Research using fossil records also suggests that the Fagus is closer to the ancestral group of Fagaceae than Quercus [78]. We concluded that Fagus and Trigonobalanus were located at the base of the phylogenetic tree (Figure 9), which was consistent with the fossil records [78]. An earlier study on the cp genome of Quercus also obtained similar results, although only one of the cp genomes (F. engleriana) was used in Fagus [41]. In this study, we showed that F. longipetiolata was closely related to F. engleriana. Moreover, they are sympatric, with distributions partially overlapping in geographical areas [17]. It indicates potential introgressive effects in these two species. However, the cp genomes of a number of species within Fagus have not yet been published. Further research of the Fagus cp genomes may provide more evidence to clarify the relationship between chloroplast phylogeny and geographic distribution.

5. Conclusions

In this study, we published the complete cp genome sequence of F. longipetiolata for the first time and compared it with those of other Fagus species, providing a useful reference for the phylogeny of Fagus. Although the cp genomes of F. longipetiolata and other Fagus were substantially similar in terms of genome structure, gene content, and gene sequences, some hot spots could be found in LSC and SSC regions, which would provide informative markers for the phylogenetic analysis of Fagus. The analysis of the selection pressure on the Fagus cp genome showed that the ndhD, rpoA, and ndhF genes of F. longipetiolata were affected by positive choices. Phylogenetic research showed a tight connection between F. longipetiolata and F. engleriana, which partially overlapped in their distribution in China. The phylogenetic relationship of Fagus with Fagaceae has been well resolved and strongly supports that Fagus is a monophyletic group. The new genome information obtained in this study can only contribute to the better development and utilization of F. longipetiolata but also provide reference data for population genome research, phylogenetic analysis and genetic engineering research.

Supplementary Materials

The following supporting information can be downloaded at: www.mdpi.com/article/10.3390/life12010092/s1, Table S1: The lengths of the introns and exons of the genes of the F. longipetiolata cp genome, Table S2: Codon usage and amino acid type statistics of the cp genome of F. longipetiolata, Table S3: Repeated sequences in the F. longipetiolata cp genome, Table S4: Simple sequence repeats (SSRs) in the F. longipetiolata cp genome.

Author Contributions

F.W. designed and directed this study and revised the paper. D.L. data analysis and writing of the first draft of the paper. H.W., J.Z. and Y.Z. Collect research materials and conduct experiments. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Guizhou Science and Technology Department under Grant ([2018] 5261, [2018] 2304, [2017] 5788); The first-class discipline construction project of Guizhou Province under Grant (GNYL [2017] 007).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the article and Supplementary Materials. The whole chloroplast genomes data of this study are openly available in GenBank of NCBI at https://www.ncbi.nlm.nih.gov (accessed on 13 June 2021) under accession number: MZ562567. The associated BioProject, BioSample and SRA numbers are PRJNA792394, SAMN24423804 and SRX13508090, respectively.

Acknowledgments

The authors sincerely thank Mingchang Wu for his great help in collecting seeds and cultivating seedlings.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Renner, S.S.; Grimm, G.W.; Kapli, P.; Denk, T. Species relationships and divergence times in beeches: New insights from the inclusion of 53 young and old fossils in a birth–death clock model. Philos. Trans. R. Soc. B 2016, 371, 20150135. [Google Scholar] [CrossRef] [Green Version]
  2. Guo, K.; Werger, M.-J. Effect of prevailing monsoons on the distribution of beeches in continental East Asia. For. Ecol. Manag. 2010, 259, 2197–2203. [Google Scholar] [CrossRef]
  3. Peters, R. Ecology of Beech Forests in the Northern Hemisphere. Ph.D. Thesis, Wageningen Agricultural University, Wageningen, The Netherlands, 1992. [Google Scholar]
  4. Fang, J.; Lechowicz, M.-J. Climatic limits for the present distribution of beech (Fagus L.) species in the world. J. Biogeogr. 2006, 33, 1804–1819. [Google Scholar] [CrossRef]
  5. Denk, T. Phylogeny of Fagus L. (Fagaceae) based on morphological data. Plant Syst. Evol. 2003, 240, 55–81. [Google Scholar] [CrossRef]
  6. Li, J.-Q. On the phylogeny of the Fagaceae. Acta Phytotaxon. Sin. 1996, 34, 597–609. (In Chinese) [Google Scholar]
  7. Ying, L.-X.; Zhang, T.-T.; Chiu, C.-A.; Chen, T.-Y.; Luo, S.-J.; Chen, X.-Y.; Shen, Z.-H. The phylogeography of Fagus hayatae (Fagaceae): Genetic isolation among populations. Ecol. Evol. 2016, 6, 2805–2816. [Google Scholar] [CrossRef] [PubMed]
  8. Nater, A.; Burri, R.; Kawakami, T.; Smeds, L.; Ellegren, H. Resolving evolutionary relationships in closely related species with whole-genome sequencing data. Syst. Biol. 2015, 64, 1000–1017. [Google Scholar] [CrossRef] [Green Version]
  9. Ravi, V.; Khurana, J.-P.; Tyagi, A.-K.; Khurana, P. An update on chloroplast genomes. Plant Syst. Evol. 2008, 271, 101–122. [Google Scholar] [CrossRef]
  10. Shinozaki, K.; Ohme, M.; Tanaka, M.; Wakasugi, T.; Hayashida, N.; Matsubayashi, T.; Sugiura, M. The complete nucleotide sequence of the tobacco chloroplast genome: Its gene organization and expression. EMBO J. 1986, 5, 2043–2049. [Google Scholar] [CrossRef]
  11. Dobrogojski, J.; Adamiec, M.; Luciński, R. The chloroplast genome: A review. Acta Physiol. Plant. 2020, 42, 1–13. [Google Scholar] [CrossRef]
  12. Shaw, J.; Lickey, E.-B.; Beck, J.-T.; Farmer, S.-B.; Liu, W.; Miller, J.; Siripun, K.-C.; Winder, C.-T.; Schilling, E.-E.; Small, R.-L. The tortoise and the hare II: Relative utility of 21 noncoding chloroplast DNA sequences for phylogenetic analysis. Am. J. Bot. 2005, 92, 142–166. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Yang, Y.; Zhu, J.; Feng, L.; Zhou, T.; Bai, G.; Yang, J.; Zhao, G. Plastid genome comparative and phylogenetic analyses of the key genera in Fagaceae: Highlighting the effect of codon composition bias in phylogenetic inference. Plant Sci. 2018, 9, 82. [Google Scholar] [CrossRef]
  14. Worth, J.-R.; Liu, L.; Wei, F.-J.; Tomaru, N. The complete chloroplast genome of Fagus crenata (subgenus Fagus) and comparison with F. engleriana (subgenus Engleriana). Peer J. 2019, 7, e7026. [Google Scholar] [CrossRef]
  15. Mader, M.; Liesebach, H.; Liesebach, M.; Kersten, B. The complete chloroplast genome sequence of Fagus sylvatica L. (Fagaceae). Mitochondrial DNA Part B 2019, 4, 1818–1819. [Google Scholar] [CrossRef]
  16. Park, J.; Oh, S.-H. A second complete chloroplast genome sequence of Fagus multinervis Nakai (Fagaceae): Intraspecific variations on chloroplast genome. Mitochondrial DNA Part B 2020, 5, 1868–1869. [Google Scholar] [CrossRef] [Green Version]
  17. Fang, J.; Wang, Z.; Tang, Z. Atlas of Woody Plants in China: Distribution and Climate; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2011; p. 162. [Google Scholar]
  18. Bankevich, A.; Nurk, S.; Antipov, D.; Gurevich, A.-A.; Dvorkin, M.; Kulikov, A.-S.; Pevzner, P.-A. SPAdes: A new genome assembly algorithm and its applications to single-cell sequencing. J. Comput. Biol. 2012, 19, 455–477. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Yang, J.; Takayama, K.; Youn, J.-S.; Pak, J.-H.; Kim, S.-C. Plastome characterization and phylogenomics of East Asian beeches with a special emphasis on Fagus multinervis on Ulleung Island, Korea. Genes 2020, 11, 1338. [Google Scholar] [CrossRef] [PubMed]
  20. Wyman, S.-K.; Jansen, R.-K.; Boore, J.-L. Automatic annotation of organellar genomes with DOGMA. Bioinformatics 2004, 20, 3252–3255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Wong, E.-H.; Smith, D.-K.; Rabadan, R.; Peiris, M.; Poon, L.-L. Codon usage bias and the evolution of influenza A viruses. Codon usage biases of influenza virus. BMC Evol. Biol. 2010, 10, 1–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Beier, S.; Thiel, T.; Munch, T.; Scholz, U.; Mascher, M. MISA-web: A web server for microsatellite prediction. Bioinformatics 2017, 33, 2583–2585. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Stothard, P.; Grant, J.R.; Van Domselaar, G. Visualizing and comparing circular genomes using the CGView family of tools. Brief. Bioinf. 2019, 20, 1576–1582. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Darling, A.-C.; Mau, B.; Blattner, F.-R.; Perna, N.-T. Mauve: Multiple alignment of conserved genomic sequence with rearrangements. Genome Res. 2004, 14, 1394–1403. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. 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] [Green Version]
  26. Librado, P.; Rozas, J. DnaSP v5: A software for comprehensive analysis of DNA polymorphism data. Bioinformatics 2009, 25, 1451–1452. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Wang, D.; Zhang, Y.; Zhang, Z.; Zhu, J.; Yu, J. KaKs Calculator 2.0: A toolkit incorporating gamma-series methods and sliding window strategies. Genom. Proteom. Bioinform. 2010, 8, 77–80. [Google Scholar] [CrossRef] [Green Version]
  28. 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]
  29. Felsenstein, J. Confidence limits on phylogenies: An approach using the bootstrap. Evolution 1985, 39, 783–791. [Google Scholar] [CrossRef]
  30. Alberti, S. The origin of the genetic code and protein synthesis. J. Mol. Evol. 1997, 45, 352–358. [Google Scholar] [CrossRef]
  31. Grantham, R.; Gautier, C.; Gouy, M.; Mercier, R.; Pave, A. Codon catalog usage and the genome hypothesis. Nucleic Acids Res. 1980, 8, 197. [Google Scholar] [CrossRef] [PubMed]
  32. Martin, A.; Bertranpetit, J.; Oliver, J.-L.; Medina, J.-R. Variation in G + C-content and codon choice: Differences among synonymous codon groups in vertebrate genes. Nucleic Acids Res. 1989, 17, 6181–6189. [Google Scholar] [CrossRef] [Green Version]
  33. Duret, L.; Mouchiroud, D. Expression pattern and, surprisingly, gene length shape codon usage in Caenorhabditis, Drosophila, and Arabidopsis. Proc. Natl. Acad. Sci. USA 1999, 96, 4482–4487. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Gu, W.; Zhou, T.; Ma, J.; Sun, X.; Lu, Z. Analysis of synonymous codon usage in SARS Coronavirus and other viruses in the Nidovirales. Virus Res. 2004, 101, 155–161. [Google Scholar] [CrossRef] [PubMed]
  35. Van der Linden, M.-G.; de Farias, S.-T. Correlation between codon usage and thermostability. Extremophiles 2006, 10, 479–481. [Google Scholar] [CrossRef] [PubMed]
  36. Hu, Y.; Woeste, K.-E.; Zhao, P. Completion of the chloroplast genomes of five Chinese Juglans and their contribution to chloroplast phylogeny. Front. Plant Sci. 2017, 7, 1955. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Huang, H.; Shi, C.; Liu, Y.; Mao, S.-Y.; Gao, L.-Z. Thirteen Camellia chloroplast genome sequences determined by high-throughput sequencing: Genome structure and phylogenetic relationships. BMC Evol. Biol. 2014, 14, 151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Raubeson, L.-A.; Peery, R.; Chumley, T.-W.; Dziubek, C.; Fourcade, H.-M.; Boore, J.-L.; Jansen, R.-K. Comparative chloroplast genomics: Analyses including new sequences from the angiosperms Nuphar advena and Ranunculus macranthus. BMC Genom. 2007, 8, 174. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Gu, C.; Ma, L.; Wu, Z.; Chen, K.; Wang, Y. Comparative analyses of chloroplast genomes from 22 Lythraceae species: Inferences for phylogenetic relationships and genome evolution within Myrtales. BMC Plant Biol. 2019, 19, 281. [Google Scholar] [CrossRef] [PubMed]
  40. Kim, S.C.; Lee, J.W.; Choi, B.K. Seven Complete Chloroplast Genomes from Symplocos: Genome Organization and Comparative Analysis. Forests 2021, 12, 608. [Google Scholar] [CrossRef]
  41. Liu, X.; Chang, E.-M.; Liu, J.-F.; Huang, Y.-N.; Wang, Y.; Yao, N.; Jiang, Z.-P. Complete chloroplast genome sequence and phylogenetic analysis of Quercus bawanglingensis Huang, Li et Xing, a vulnerable oak tree in China. Forests 2019, 10, 587. [Google Scholar] [CrossRef] [Green Version]
  42. Liang, H.; Chen, J. Comparison and Phylogenetic Analyses of Nine Complete Chloroplast Genomes of Zingibereae. Forests 2021, 12, 710. [Google Scholar] [CrossRef]
  43. Ding, S.; Dong, X.; Yang, J.; Guo, C.; Cao, B.; Guo, Y.; Hu, G. Complete chloroplast genome of Clethra fargesii Franch., an original sympetalous plant from central China: Comparative analysis, adaptive evolution, and phylogenetic relationships. Forests 2021, 12, 441. [Google Scholar] [CrossRef]
  44. Mehmood, F.; Shahzadi, I.; Waseem, S.; Mirza, B.; Ahmed, I.; Waheed, M.-T. Chloroplast genome of Hibiscus rosa-sinensis (Malvaceae): Comparative analyses and identification of mutational hotspots. Genomics 2020, 112, 581–591. [Google Scholar]
  45. Yang, Y.; Hu, Y.; Ren, T.; Sun, J.; Zhao, G. Remarkably conserved plastid genomes of Quercus group Cerris in China: Comparative and phylogenetic analyses. Nord. J. Bot. 2018, 36, e01921. [Google Scholar] [CrossRef]
  46. Yang, Y.; Zhou, T.; Duan, D.; Yang, J.; Feng, L.; Zhao, G. Comparative analysis of the complete chloroplast genomes of five Quercus species. Front. Plant Sci. 2016, 7, 959. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Zhang, X.; Hu, Y.; Liu, M.; Lang, T. Optimization of assembly pipeline may improve the sequence of the chloroplast genome in Quercus spinosa. Sci. Rep. 2018, 8, 8906. [Google Scholar] [CrossRef] [Green Version]
  48. Dong, W.; Xu, C.; Liu, Y.; Shi, J.; Li, W.; Suo, Z. Chloroplast phylogenomics and divergence times of Lagerstroemia (Lythraceae). BMC Genom. 2021, 22, 434. [Google Scholar] [CrossRef] [PubMed]
  49. Nguyen, H.Q.; Nguyen, T.N.L.; Doan, T.N.; Nguyen, T.T.N.; Phạm, M.H.; Le, T.L.; Sy, T.D.; Chu, H.H.; Chu, H.M. Complete chloroplast genome of novel Adrinandra megaphylla Hu species: Molecular structure, comparative and phylogenetic analysis. Sci. Rep. 2021, 11, 11731. [Google Scholar] [CrossRef]
  50. Hsu, C.-M.; Yang, W.-P.; Chen, C.-C.; Lai, Y.-K.; Lin, T.-Y. A point mutation in the chloroplast rps12 gene from Nicotiana plumbaginifolia confers streptomycin resistance. Plant Mol. Biol. 1993, 23, 179–183. [Google Scholar] [CrossRef] [PubMed]
  51. Park, S.; Jansen, R.-K.; Park, S. Complete plastome sequence of Thalictrum coreanum (Ranunculaceae) and transfer of the rpl32 gene to the nucleus in the ancestor of the subfamily Thalictroideae. BMC Plant. Biol. 2015, 15, 40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Sahoo, D.-K.; Das, A.; Huang, X.; Cianzio, S.; Bhattacharyya, M.-K. Tightly linked Rps12 and Rps13 genes provide broad-spectrum Phytophthora resistance in soybean. Sci. Rep. 2021, 11, 16907. [Google Scholar] [CrossRef]
  53. Li, W.; Liu, Y.; Yang, Y.; Xie, X.; Lu, Y.; Yang, Z.; Jin, X.; Suo, Z. Interspecific chloroplast genome sequence diversity and genomic resources in Diospyros. BMC Plant Biol. 2018, 18, 210. [Google Scholar] [CrossRef] [PubMed]
  54. Zhang, Y.; Tian, Y.; Tng, D.-Y.; Zhou, J.; Zhang, Y.; Wang, Z.; Li, P.; Wang, Z. Comparative chloroplast genomics of Litsea Lam. (Lauraceae) and its phylogenetic implications. Forests 2021, 12, 744. [Google Scholar] [CrossRef]
  55. Hong, Z.; Wu, Z.; Zhao, K.; Yang, Z.; Zhang, N.; Guo, J.; Xu, D. Comparative analyses of five complete chloroplast genomes from the genus Pterocarpus (Fabacaeae). Int. J. Mol. Sci. 2020, 21, 3758. [Google Scholar] [CrossRef] [PubMed]
  56. Asaf, S.; Khan, A.L.; Khan, A.; Al-Harrasi, A. Unraveling the chloroplast genomes of two prosopis species to identify its genomic information, comparative analyses and phylogenetic relationship. Int. J. Mol. Sci. 2020, 21, 3280. [Google Scholar] [CrossRef] [PubMed]
  57. Li, Y.-F.; Sylvester, S.-P.; Li, M.; Zhang, C.; Li, X.; Duan, Y.-F.; Wang, X.-R. The complete plastid genome of Magnolia zenii and genetic comparison to Magnoliaceae species. Molecules 2019, 24, 261. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Zhou, M.; Long, W.; Li, X. Patterns of synonymous codon usage bias in chloroplast genomes of seed plants. For. Stud. China 2008, 10, 235–242. [Google Scholar] [CrossRef]
  59. Necşulea, A.; Lobry, J.-R. A new method for assessing the effect of replication on DNA base composition asymmetry. Mol. Biol Evol. 2007, 24, 2169–2179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Orešič, M.; Shalloway, D. Specific correlations between relative synonymous codon usage and protein secondary structure. J. Mol. Biol. 1998, 281, 31–48. [Google Scholar] [CrossRef]
  61. Chen, Y.; Hu, N.; Wu, H. Analyzing and characterizing the chloroplast genome of Salix wilsonii. BioMed Res. Int. 2019, 2019, 5190425. [Google Scholar] [CrossRef] [Green Version]
  62. Khan, A.; Asaf, S.; Khan, A.-L.; Al-Harrasi, A.; Al-Sudairy, O.; AbdulKareem, N.-M.; Shinwari, Z.-K. First complete chloroplast genomics and comparative phylogenetic analysis of Commiphora gileadensis and C. foliacea: Myrrh producing trees. PLoS ONE 2019, 14, e0208511. [Google Scholar] [CrossRef] [Green Version]
  63. Li, B.; Zheng, Y. Dynamic evolution and phylogenomic analysis of the chloroplast genome in Schisandraceae. Sci. Rep. 2018, 8, 9285. [Google Scholar] [CrossRef] [PubMed]
  64. Takahashi, D.; Sakaguchi, S.; Isagi, Y.; Setoguchi, H. Comparative chloroplast genomics of series Sakawanum in genus Asarum (Aristolochiaceae) to develop single nucleotide polymorphisms (SNPs) and simple sequence repeat (SSR) markers. J. For. Res. 2018, 23, 387–392. [Google Scholar] [CrossRef]
  65. Yamane, K.; Kawahara, T. Size homoplasy and mutational behavior of chloroplast simple sequence repeats (cpSSRs) inferred from intra-and interspecific variations in four chloroplast regions of diploid and polyploid Triticum and Aegilops species. Genet. Resour. Crop. Evol. 2018, 65, 727–743. [Google Scholar] [CrossRef]
  66. Alanazi, K.M.; Ali, M.A.; Kim, S.Y.; Rahman, M.O.; Farah, M.A.; Alhemaid, F.; Elangbam, M.; Gurung, A.B.; Lee, J. The cp genome characterization of Adenium obesum: Gene content, repeat organization and phylogeny. Saudi J. Biol. Sci. 2021, 28, 3768–3775. [Google Scholar] [CrossRef]
  67. Wu, L.; Cui, Y.; Wang, Q.; Xu, Z.; Wang, Y.; Lin, Y.; Song, J.; Yao, H. Identification and phylogenetic analysis of five Crataegus species (Rosaceae) based on complete chloroplast genomes. Planta 2021, 254, 14. [Google Scholar] [CrossRef]
  68. Li, D.M.; Zhao, C.Y.; Liu, X.F. Complete Chloroplast Genome Sequences of Kaempferia galanga and Kaempferia elegans: Molecular Structures and Com-parative Analysis. Molecules 2019, 24, 474. [Google Scholar] [CrossRef] [Green Version]
  69. Mo, Z.; Lou, W.; Chen, Y.; Jia, X.; Zhai, M.; Guo, Z.; Xuan, J. The chloroplast genome of Carya illinoinensis: Genome structure, adaptive evolution, and phylogenetic analysis. Forests 2020, 11, 207. [Google Scholar] [CrossRef] [Green Version]
  70. Loewe, L.; Charlesworth, B.; Bartolomé, C.; Noel, V. Estimating selection on nonsynonymous mutations. Genetics 2006, 172, 1079–1092. [Google Scholar] [CrossRef] [Green Version]
  71. Yang, Z.; Nielsen, R. Estimating synonymous and nonsynonymous substitution rates under realistic evolutionary models. Mol. Biol. Evol. 2000, 17, 32–43. [Google Scholar] [CrossRef] [Green Version]
  72. Tian, S.; Lu, P.; Zhang, Z.; Wu, J.-Q.; Zhang, H.; Shen, H. Chloroplast genome sequence of Chongming lima bean (Phaseolus lunatus L.) and comparative analyses with other legume chloroplast genomes. BMC Genom. 2021, 22, 194. [Google Scholar] [CrossRef]
  73. Xiong, Y.; Xiong, Y.; He, J.; Yu, Q.; Zhao, J.; Lei, X.; Ma, X. The complete chloroplast genome of two important annual clover species, Trifolium alexandrinum and T. resupinatum: Genome structure, comparative analyses and phylogenetic relationships with relatives in Leguminosae. Plants 2020, 9, 478. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Hu, S.; Sablok, G.; Wang, B.; Qu, D.; Barbaro, E.; Viola, R.; Varotto, C. Plastome organization and evolution of chloroplast genes in Cardamine species adapted to contrasting habitats. BMC Genom. 2015, 16, 306. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Xie, X.; Huang, R.; Li, F.; Tian, E.; Li, C.; Chao, Z. Phylogenetic position of Bupleurum sikangense inferred from the complete chloroplast genome sequence. Gene 2021, 798, 145801. [Google Scholar] [CrossRef] [PubMed]
  76. Williams, A.M.; Friso, G.; Van Wijk, K.J.; Sloan, D.B. Extreme variation in rates of evolution in the plastid Clp protease complex. Plant J. 2019, 98, 243–259. [Google Scholar] [CrossRef]
  77. Oh, S.-H.; Manos, P.-S. Molecular phylogenetics and cupule evolution in Fagaceae as inferred from nuclear CRABS CLAW sequences. Taxon 2008, 57, 434–451. [Google Scholar]
  78. Zhou, Z.-K. Fossils of the Fagaceae and their implications in systematics and biogeography. J. Syst. Evol. 1999, 37, 369–385. [Google Scholar]
Figure 1. Gene map of the F. longipetiolata cp genome. The forward coding genes are located on the outside of the circle, and the reverse coding genes are located on the inside of the circle. Genes with different functions are colour-coded. The deeper grey in the inner circle represents the GC content, while the lighter grey represents the AT content.
Figure 1. Gene map of the F. longipetiolata cp genome. The forward coding genes are located on the outside of the circle, and the reverse coding genes are located on the inside of the circle. Genes with different functions are colour-coded. The deeper grey in the inner circle represents the GC content, while the lighter grey represents the AT content.
Life 12 00092 g001
Figure 2. Counting of relative synonymous codon usage (RSCU) of amino acids in the chloroplast genome of F. longipetiolata. The colors of the histogram correspond to the colors of the codons.
Figure 2. Counting of relative synonymous codon usage (RSCU) of amino acids in the chloroplast genome of F. longipetiolata. The colors of the histogram correspond to the colors of the codons.
Life 12 00092 g002
Figure 3. Repeated sequences of the F. longipetiolata cp genome. The abscissa is the length of the repeat sequence, and the ordinate is the number of repeat sequences. F stands for forward repetition, P for palindromic repetition, R for reverse repetition, and C for complementary repetition.
Figure 3. Repeated sequences of the F. longipetiolata cp genome. The abscissa is the length of the repeat sequence, and the ordinate is the number of repeat sequences. F stands for forward repetition, P for palindromic repetition, R for reverse repetition, and C for complementary repetition.
Life 12 00092 g003
Figure 4. Counting of the types of SSRs. The abscissa represents SSR repetition units, and the ordinate represents the number of SSRs of each type.
Figure 4. Counting of the types of SSRs. The abscissa represents SSR repetition units, and the ordinate represents the number of SSRs of each type.
Life 12 00092 g004
Figure 5. Comparative analysis of cp genome structure between Fagus and A. thaliana. The outer two circles describe the gene length and direction of the genome. The five circles inside represent the similarity results of genome alignment between F. longipetiolata and A. thaliana (AP000423), F. engleriana (KX852398), F. crenata (MH171101), F. sylvatica (MK598696), and F. japonica var. multinervis (MN894556). Black circles represent GC content; green circles indicate G > C; purple circles indicate G < C.
Figure 5. Comparative analysis of cp genome structure between Fagus and A. thaliana. The outer two circles describe the gene length and direction of the genome. The five circles inside represent the similarity results of genome alignment between F. longipetiolata and A. thaliana (AP000423), F. engleriana (KX852398), F. crenata (MH171101), F. sylvatica (MK598696), and F. japonica var. multinervis (MN894556). Black circles represent GC content; green circles indicate G > C; purple circles indicate G < C.
Life 12 00092 g005
Figure 6. Line chart of gene Pi values calculated from 106 loci based on the five Fagus cp genomes. x-axis: gene name, y-axis: Pi value.
Figure 6. Line chart of gene Pi values calculated from 106 loci based on the five Fagus cp genomes. x-axis: gene name, y-axis: Pi value.
Life 12 00092 g006
Figure 7. Comparisons of IR-SC boundary locations across the cp genomes of the five Fagus and A. thaliana.
Figure 7. Comparisons of IR-SC boundary locations across the cp genomes of the five Fagus and A. thaliana.
Life 12 00092 g007
Figure 8. The 80 protein-coding genes of the F. longipetiolata cp genome and four Fagus species were used for ka/ka analysis.
Figure 8. The 80 protein-coding genes of the F. longipetiolata cp genome and four Fagus species were used for ka/ka analysis.
Life 12 00092 g008
Figure 9. The phylogenetic tree was constructed by using the complete cp sequences of F. longipetiolata and 18 Fagaceae species. P. trichocarpa was used as the outgroup.
Figure 9. The phylogenetic tree was constructed by using the complete cp sequences of F. longipetiolata and 18 Fagaceae species. P. trichocarpa was used as the outgroup.
Life 12 00092 g009
Table 1. Characteristics of F. longipetiolata cp genome.
Table 1. Characteristics of F. longipetiolata cp genome.
CategoryItemDescribe
Chloroplast genome structureCp gene/bp158,350
LSC/bp87,671
SSC/bp18,891
IRA/IRB/bp25,894
Gene compositionCp gene131
CDS81
tRNA37
rRNA8
pseudo5
GC Content (%)Cp gene37.09
LSC35.05
SSC31.19
IRA/IRB42.70
Table 2. Genes in cp genome of F. longipetiolata.
Table 2. Genes in cp genome of F. longipetiolata.
CategoryGene GroupGene Name
PhotosynthesisSubunits of photosystem IpsaA, psaB, psaC, psaI, psaJ
Subunits of photosystem IIpsbA, psbB, psbC, psbD, psbE, psbF, psbH, psbI, psbJ, psbK, psbL, psbM, psbN, psbT, psbZ
Subunits of NADH dehydrogenasendhA *, ndhB * (2), ndhC, ndhD, ndhE, ndhF, ndhG, ndhH, ndhI, ndhJ, ndhK
Subunits of cytochrome b/f complexpetA, petB *, petD *, petG, petL, petN
Subunits of ATP synthaseatpA, atpB, atpE, atpF *, atpH, atpI
Large subunit of rubiscorbcL
Subunits photochlorophyllide reductase-
Self-replicationProteins of large ribosomal subunit# rpl22, rpl14, rpl16 *, rpl2 * (2), rpl20, rpl23 (2), rpl32, rpl33, rpl36
Proteins of small ribosomal subunit# rps16, rps11, rps12 * (2), rps14, rps15, rps18, rps19, rps2, rps3, rps4, rps7 (2), rps8
Subunits of RNA polymeraserpoA, rpoB, rpoC1 *, rpoC2
Ribosomal RNAsrrn16 (2), rrn23 (2), rrn4.5 (2), rrn5 (2)
Transfer RNAstrnA-UGC * (2), trnC-GCA, trnD-GUC, trnE-UUC, trnF-GAA, trnG-GCC, trnG-UCC *, trnH-GUG, trnI-CAU (2), trnI-GAU * (2), trnK-UUU *, trnL-CAA (2), trnL-UAA *, trnL-UAG, trnM-CAU, trnN-GUU (2), trnP-UGG, trnQ-UUG, trnR-ACG (2), trnR-UCU, trnS-GCU, trnS-GGA, trnS-UGA, trnT-GGU, trnT-UGU, trnV-GAC (2), trnV-UAC *, trnW-CCA, trnY-GUA, trnfM-CAU
Other genesMaturasematK
ProteaseclpP **
Envelope membrane proteincemA
Acetyl-CoA carboxylaseaccD
c-type cytochrome synthesis geneccsA
Translation initiation factor# infA
other-
Genes of unknown functionConserved hypothetical chloroplast ORF# ycf1 (2), ycf2 (2), ycf3 **, ycf4
Note: # Gene, Pseudo gene; Gene (2), Multiple copy gene, the number of copies in parenthesis; Gene *, Gene with one intron; Gene **, Genes containing two introns.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liang, D.; Wang, H.; Zhang, J.; Zhao, Y.; Wu, F. Complete Chloroplast Genome Sequence of Fagus longipetiolata Seemen (Fagaceae): Genome Structure, Adaptive Evolution, and Phylogenetic Relationships. Life 2022, 12, 92. https://doi.org/10.3390/life12010092

AMA Style

Liang D, Wang H, Zhang J, Zhao Y, Wu F. Complete Chloroplast Genome Sequence of Fagus longipetiolata Seemen (Fagaceae): Genome Structure, Adaptive Evolution, and Phylogenetic Relationships. Life. 2022; 12(1):92. https://doi.org/10.3390/life12010092

Chicago/Turabian Style

Liang, Daqu, Haoyun Wang, Jun Zhang, Yuanxiang Zhao, and Feng Wu. 2022. "Complete Chloroplast Genome Sequence of Fagus longipetiolata Seemen (Fagaceae): Genome Structure, Adaptive Evolution, and Phylogenetic Relationships" Life 12, no. 1: 92. https://doi.org/10.3390/life12010092

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