Next Article in Journal
State of the Art of Microwave Treatment of Wood: Literature Review
Next Article in Special Issue
Cryopreservation and Micropropagation Methods for Conservation of Genetic Resources of Ulmus laevis and Ulmus glabra
Previous Article in Journal
Filamentous Fungi and Yeasts Associated with Mites Phoretic on Ips typographus in Eastern Finland
Previous Article in Special Issue
A Comparison of Three Circular Mitochondrial Genomes of Fagus sylvatica from Germany and Poland Reveals Low Variation and Complete Identity of the Gene Space
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Comparative Chloroplast Genomics of Litsea Lam. (Lauraceae) and Its Phylogenetic Implications

1
College of Life Sciences, Nanjing University, Nanjing 210023, China
2
Centre for Rainforest Studies, The School for Field Studies, Yungaburra, QLD 4884, Australia
3
Shanghai Chenshan Botanical Garden, Shanghai 201602, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Forests 2021, 12(6), 744; https://doi.org/10.3390/f12060744
Submission received: 11 May 2021 / Revised: 30 May 2021 / Accepted: 2 June 2021 / Published: 4 June 2021

Abstract

:
Litsea Lam. is an ecological and economic important genus of the “core Lauraceae” group in the Lauraceae. The few studies to date on the comparative chloroplast genomics and phylogenomics of Litsea have been conducted as part of other studies on the Lauraceae. Here, we sequenced the whole chloroplast genome sequence of Litsea auriculata, an endangered tree endemic to eastern China, and compared this with previously published chloroplast genome sequences of 11 other Litsea species. The chloroplast genomes of the 12 Litsea species ranged from 152,132 (L. szemaois) to 154,011 bp (L. garrettii) and exhibited a typical quadripartite structure with conserved genome arrangement and content, with length variations in the inverted repeat regions (IRs). No codon usage preferences were detected within the 30 codons used in the chloroplast genomes, indicating a conserved evolution model for the genus. Ten intergenic spacers (psbE–petL, trnH–psbA, petA–psbJ, ndhF–rpl32, ycf4–cemA, rpl32–trnL, ndhG–ndhI, psbC–trnS, trnE–trnT, and psbM–trnD) and five protein coding genes (ndhD, matK, ccsA, ycf1, and ndhF) were identified as divergence hotspot regions and DNA barcodes of Litsea species. In total, 876 chloroplast microsatellites were located within the 12 chloroplast genomes. Phylogenetic analyses conducted using the 51 additional complete chloroplast genomes of “core Lauraceae” species demonstrated that the 12 Litsea species grouped into four sub-clades within the Laurus-Neolitsea clade, and that Litsea is polyphyletic and closely related to the genera Lindera and Laurus. Our phylogeny strongly supported the monophyly of the following three clades (Laurus–Neolitsea, Cinnamomum–Ocotea, and Machilus–Persea) among the above investigated “core Lauraceae” species. Overall, our study highlighted the taxonomic utility of chloroplast genomes in Litsea, and the genetic markers identified here will facilitate future studies on the evolution, conservation, population genetics, and phylogeography of L. auriculata and other Litsea species.

1. Introduction

The Lauraceae or the Laurel family is a large monophyletic family in the order Laurales and encompasses approximately 2500 to 3000 species from around 50 genera [1,2]. Systematic classification schemes and the tribe- and genus-level phylogenetic relationships within this family have long been controversial [1,3,4,5,6]. The focal genus of the present study, Litsea Lam., belongs to the “core Lauraceae” or “core Laureae” group in the sense of Chanderbali et al. [1] and Rohwer and Rudolph [7]. As currently circumscribed, Litsea contains around 408 species, with Litsea cubeba (Lour.) Pers. as its type species [8], and this genus is commonly characterized by the following morphological traits: evergreen/deciduous trees or shrubs; leaves alternate, rarely opposite or verticillate; umbel inflorescence surrounded by an involucre consisted of persistent, alternate, and opposite bracts; anthers four-celled, all introrse, cells opening by lids [3,9]. The majority of Litsea species inhabit tropical and subtropical Asia, while a few Litsea species are distributed in Australia and from North America to subtropical South America [9]. Regarding the fossil records of Litsea, Dai et al. [10] reported a well-preserved fossil leaf of Litsea cf. chunii Cheng discovered in the Late Pliocene’s sediments of the Mangbang formation in Tengchong county, Yunnan Province, China, and a mummified fossil wood of Litseoxylon nanningensis gen. et sp. Nov. was identified from the Upper Oligocene Yongning Formation of the Nanning Basin, Guangxi Province, South China, by Huang et al. [11]. Moreover, the origin and evolution of Litsea genera group sensu lato was discussed by Li [12], who pointed that this group originated at South Laurasia and North Gondwana, as well as the tropical coast of Tethys sea after Mid-Cretaceous; furthermore, the core genera of this group (Litsea and Lindera) probably originated and speciated ranging from South China to Indo-Malaysia, from where they migrated into tropical America and Australasia. Moreover, many of the species of Litsea have ethnobotanical or economic uses [13,14,15,16,17,18]. For instance, Litsea cubeba is an important source of May Chang essential oil (roughly 95% terpenoid), which is broadly used for perfumes and cosmetics [13,14,15], and Litsea mollis is used as a traditional Chinese medicine for the treatment of inflammation, poor blood circulation, lumbocrural pain, quadriplegia, and dysmenorrhea [18]. Additionally, many Litsea species are in urgent need of conservation: 98 species of this genus are listed in the International Union for Conservation of Nature (IUCN) Red List of Threatened Species (https://www.iucnredlist.org/ (accessed on 2 May 2021)) [19].
Despite the diversity and ecological and economic importance of Litsea, there are still uncertainties surrounding the systematic placement of the genus. Even though there have been many previous phylogenetic studies on Litsea that are based on morphological and anatomical characteristics as well as molecular data, systematic placement of Litsea remains perplexed. Richter [20] showed that Litsea, Apollonias, Laurus, Lindera, Sassafras, and Umbellularia shared similar wood structure. Li [21] found that Litsea and Lindera exhibited many similarities in terms of morphology and distribution. Specifically, Li [21] showed that the ancestral phenotype of the above two genera was characterized by umbellate flowers, elongated flower branches in epiphytic inflorescences, and evergreen leaves with three veins. Additionally, both genera originated from tropical Asia and followed identical evolutionary processes: from evergreen to deciduous, from three veins to pinnate veins, from more to less flowers, and from axillary to terminal inflorescences [21]. Li [21] also further suggested that Litsea and Lindera can be merged into a group if they did not display the variations in the number of anther cells (four-celled versus two-celled). In addition, Raj and Werff [22] made the case that Litsea was related to Lindera from the investigation of pollen morphology. Specifically, the size of pollen grains, the number of spinules per pollen grain, and the ultra-structural details of the pollen wall of Litsea and Lindera have been found to be identical [22]. On the basis of these morphological characters, Li [21] concluded that the number of anther cells could be a good taxonomic basis for the monophyly of Litsea within Lauraceae. However, Werff and Richter [23] discovered the homoplasy or instability of this character in distinguishing Lauraceae, thus resulting in the disputed systematics and delimitation of Litsea.
Rohwer [4] first used the universal chloroplast DNA gene matK to construct a phylogenetic tree of Lauraceae, finding that the genera with umbellate, involucrate inflorescences (Litsea, Actinodaphne, Laurus, Lindera, Neolitsea and Umbellularia) had a low degree of matK sequence differentiation and Litsea had a close relationship with Lindera. Chanderbali et al. [1] rebuilt the phylogenetic tree among 122 species of Lauraceae representing 44 genera employing the combined analysis of the chloroplast DNA genes (trnL–trnF, trnT–trnL, psbA–trnH, and rpl16) and ribosomal DNA sequences: 26S, internal transcribed spacer (ITS) and 5.8S, and their results indicated that Litsea grouped with Actinodaphne, Lindera, and Neolitsea. Due to the unsettled generic delimitation of Litsea and Lindera, they cautioned against assessing the delimitations of these genera from morphology. On the basis of the matK gene of chloroplast DNA and nuclear internal transcribed spacer (nrITS), Li et al. [3] analyzed the phylogenetic relationship of “core Laureae” group and found that Litsea, Actinodaphne, Neolitsea, and Lindera were polyphyletic but poorly bootstrap-supported. Fijridiyanto and Murakami [24] reconstructed the phylogenetic tree of Litsea and its related genera using the nuclear taxonomic marker rpb2, which supported the fact that Litsea was not monophyletic and closer to Lindera.
Over the last decade, the rapid advances of high-throughput sequencing or next-generation sequencing (NGS) technologies have yielded tremendous genome-scale data for angiosperm species and have improved the development of plastid phylogenomics and phylogenetic resolution of many land plants [25,26,27]. Sequencing complete chloroplast genome, for instance, has proven to be informative and effective in resolving complex phylogenetic relationships at a wide range of taxonomic levels [25,26,27]. Specific to the intrageneric phylogeny of the Lauraceae, these advances have revealed previously unknown phylogenetic relationships among genera. For example, Song et al. [6] employed the method of plastid phylogenomics to construct a comprehensive and robust phylogenetic tree of Lauraceae, finding that species of Litsea, Actinodaphne, Iteadaphne, Laurus, Lindera, Neolitsea, and Parasassafras grouped into one clade. Within the clade, four Litsea species (L. glutinosa, L. monopetala, L. magnoliifolia, and L. tsilingensis), Laurus nobilis, and four Lindera species (L. communis, L. glauca, L. megaphylla, and L. nacusua) were located in one sub-clade, while Lindera obtusiloba with other three Litsea species (L. cubeba, L. panamonja, and L. pierrei) formed another sub-clade. Tian et al. [28] investigated the phylogeny of the “core Lauraceae” group with their whole chloroplast genomes and found that three Litsea species (L. monopetala, L. glutinosa, and L. tsinlingensis), Laurus nobilis, and Lindera megaphylla formed one clade with high bootstrap. Using whole plastome genomes, Zhao et al. [29] also found strong support for a close relationship between the genera Litsea, Laurus, and Lindera, and suggested that the clade containing Litsea, Laurus, and Lindera was a sister to the “CinnamomumOcotea clade” consisting of the genera Cinnamomum, Nectandra, and Sassafras. These studies have strongly indicated the polyphyletic nature of Litsea and their unresolved phylogenetic positions within the Lauraceae.
Litseaauriculata Chien et Cheng, the focal species of this study, is a rare deciduous tree and precious medicinal tree restricted and endemic to montane regions in the Zhejiang and Anhui provinces of eastern China [30,31]. Due to its narrow distribution and small number of wild populations, L. auriculata has been listed as “near threatened” by the IUCN and “Grade III Key Protected Wild Plant” by the Chinese Plant Red Book [32]. Thus far, research on L. auriculata has primarily focused on its seed germination, ecophysiology, and population genetics [30,33,34,35], while studies on the conservation genetics and genomics of the species is lacking. Therefore, we sequenced and reported the complete chloroplast genome of Litsea auriculata for the first time and conducted comparative chloroplast genomics of the genus Litsea with published chloroplast genomes of 11 other Litsea species. Moreover, we developed potential genetic markers such as microsatellites and DNA barcodes for the genus Litsea. Furthermore, to shed light on the systematic placement of Litsea within the “core Lauraceae” or “core Laureae” group, we reconstructed maximum likelihood (ML) and Bayesian inference (BI) phylogenetic trees of these Litsea species and an additional 51 published plastomes of “core Lauraceae” species. The outcomes here will provide useful genomic resources for further studies on conservation and utilization of L. auriculata and other Litsea species and can lay a solid foundation for understanding the phylogeny of Litsea.

2. Materials and Methods

2.1. Plant Sampling and DNA Extraction of L. auriculata

We sampled fresh young leaves of L. auriculata from a cultivated tree at the Shanghai Chenshan Botanical Garden, Shanghai, China (31°4.2609′ N, 121°10.9040′ E), and lodged a voucher specimen (accession number: TMMJZ20200904) at the herbarium of Nanjing University. Total genomic DNA of L. auriculata was extracted from approximately 5 mg of the silica-dried leaf tissue using a modified CTAB method [36]. Subsequently, the integrity of DNA was evaluated by agarose gel electrophoresis and validated using an Agilent 2100 Bioanalyzer (Agilent Technologies, Santa Clara, CA, USA). NanoDrop LITE spectrophotometer (Thermo Fisher Scientific, Wilmington, DE, USA) was employed to measure the concentration of DNA.

2.2. Illumina Paired-End Sequencing, De Novo Assembly, and Annotation of the Chloroplast Genome of L. auriculata

The high-quality genomic DNA of L. auriculata was used to construct Illumina paired-end (2 × 150 bp) library and implement the low-coverage shotgun sequencing in a lane of HiSeq Xten platform (Illumina, San Diego, CA, USA) at The Beijing Genomics Institute (BGI, Shenzhen, China). Approximately 6 Gb of raw data were sequenced, and the clean data were obtained employing NGS QC Tool Kit by removing adapter sequences and low-quality reads with a Q-value ≤ 20. We de novo assembled the complete chloroplast genome of L. auriculata via the GetOrganelle pipeline [37] using the clean data. The chloroplast genome was then automatically annotated using Plastid Genome Annotator [38], with manual adjusting and confirmation of the annotated protein coding genes (CDS) in Geneious Prime (http://www.geneious.com/ (accessed on 10 May 2021), Biomatters Ltd., Auckland, New Zealand). The annotated tRNA genes were further verified using the tRNAscan-SE [39] with default parameters. Afterwards, the resulting annotated cp genome of L. auriculata was submitted to The National Center for Biotechnology Information (NCBI; https://www.ncbi.nlm.nih.gov/ (accessed on 11 May 2021)). The circular cp genome physical map of L. auriculata was drawn employing the Chloroplot (https://irscope.shinyapps.io/chloroplot/ (accessed on 5 May 2021)) [40], with subsequent manual editing.

2.3. Comparative Chloroplast Genome Analysis of Litsea

Chloroplast genome comparison among the 12 Litsea species was carried out under the Shuffle-LAGAN mode via mVISTA program (http://genome.lbl.gov/vista/mvista/ (accessed on 8 May 2021)) [41] to elucidate the level of sequence divergence, using the annotation of L. acutivena chloroplast genome as a reference.
To further investigate the chloroplast genome-wide evolutionary dynamics and structural variations across these 12 Litsea species, we used MAUVE (http://darlinglab.org/mauve/ (accessed on 10 May 2021)) [42] to identify the following evolutionary events: gene loss, duplication, re-arrangements, or translocations in multiple alignments. IRscope (https://irscope.shinyapps.io/irapp/ (accessed on 10 May 2021)) [43] was employed to trace the size variation in the boundary regions of chloroplast genome among inverted repeat regions (IRs), small single copy region (SSC), and large single copy region (LSC).
Additionally, codon usage together with relative synonymous codon usage (RSCU) [44] value was estimated for all protein-coding genes (the genes with sequence length less than 300 bp and repeated genes were eliminated) of 12 whole chloroplast genomes of Litsea via CodonW v1.4.2 (http://codonw.sourceforge.net/ (accessed on 8 May 2021)) [45]. The two unique codons (AUG and UGG) and the three stop codons (TAA, TAG, and TGA) have no degeneracy and were deleted from the data before analysis.

2.4. Mining of cp Microsatellite Markers and Hypervariable Regions of Litsea

We used the MIcroSAtellite (MISA) perl script [46] to exploit simple sequence repeats (SSRs) within the chloroplast genomes of the studied 12 Litsea species, setting the parameters with thresholds of 10 repeat units for mononucleotide SSRs; 6 repeat units for dinucleotide SSRs; and 5 repeat units for tri-, tetra-, penta-, and hexa-nucleotide SSRs. Genus-targeted polymorphic SSRs among these 12 Litsea species were selected under the following three criteria: (1) SSRs located in the homologous regions, (2) SSRs that possessed the same repeat units, and (3) the number of repeat units was variable.
Finally, we used DnaSP v6.0 (http://www.ub.edu/dnasp/ (accessed on 8 May 2021)) [47] to calculate the nucleotide variability (π) of both coding regions and non-coding regions (including intergenic spacers and introns) with aligned length > 200 bp and mutation site > 0 after sequence alignment via MAFFT v7 (https://mafft.cbrc.jp/alignment/software/ (accessed on 10 May 2021)) [48] for identifying the highly variable regions within the chloroplast genomes of 12 Litsea species for their future population genetics and species delineation studies. The calculated nucleotide variability values were then plotted in R v4.0.2 (https://www.r-project.org/ (accessed on 2 May 2021)).

2.5. Phylogenetic Analysis

To ascertain the phylogenetic position of Litsea within the “core Lauraceae” or “core Laureae” group, we employed 12 Litsea together with other 50 “core Lauraceae” species representing the Laurus–Neolitsea, Cinnamomum–Ocotea, and Machilus–Persea clades as ingroups and Caryodaphnopsis tonkinensis as an outgroup according to the newly updated classification of Lauraceae by Song et al. [6]. Phylogenetic analyses were conducted on the whole chloroplast genome sequences of the above 63 species. The nucleotide sequences were aligned using the default parameters in the MAFFT v7 (https://mafft.cbrc.jp/alignment/software/ (accessed on 10 May 2021)) [48] and followed by some manual adjustments.
We inferred the phylogenetic relationships among the above-studied species using the maximum likelihood (ML) method as implemented in RAxML-HPC v8.2.8 [49] on the CIPRES Science Gateway website (https://www.phylo.org/ (accessed on 11 May 2021)) and the Bayesian inference (BI) method as implemented in MrBayes v3.1.2 [50] under the unpartitioned strategy. A corrected Akaike information criterion (AIC) value was used to determine the best-fitting model of nucleotide substitution and sequence evolution via jModelTest v2.1.10 [51,52], resulting the optimal model of GTR + I + G for both subsequent ML and BI phylogenetic analysis.
Bayesian analyses were conducted using two separated runs of the Markov chain Monte Carlo (MCMC) algorithm for 1 million generations and tree sampling every 1000 generations. The first 25% of sampled trees were discarded as burn-in, and the 25% best-scoring trees were used to construct the consensus tree and to estimate the posterior probabilities (PPs). Convergence was determined by estimating the average standard deviation of the split frequencies (<0.01). To construct the ML tree, we also ran two searches to ensure identical topologies and ML nodal support was calculated with 1000 bootstrap (BS) replicates for each run. Topologies of the above phylogenetic trees were visualized using the Interactive Tree of Life (iTOL) v4 (https://itol.embl.de (accessed on 11 May 2021)) [53].

3. Results and Discussion

3.1. Conservation of Litsea Chloroplast Genomes

The Illumina HiSeq Xten platform produced 40,992,320 clean paired-end reads for the chloroplast genome de novo assembly of L. auriculata, and the mean sequencing coverage of the chloroplast genome estimated by the GetOrganelle pipeline was 159×. The complete chloroplast genome of L. auriculata was 152,377 bp in length and displayed a quadripartite structure consisting of a pair of inverted repeat regions (IR with 20,015 bp) divided by two single-copy regions (LSC, 93,533 bp, and SSC, 18,814 bp; Figure 1, Table 1). The overall GC content of the chloroplast genome was 39.2%, with the corresponding values of 37.9%, 33.9%, and 44.4% for the LSC, SSC, and IR regions, respectively. Moreover, there were a total of 113 unique genes, including 79 protein-coding genes (CDS), 30 transfer RNA (tRNA) genes, and 4 ribosomal RNA (rRNA) genes. Among these genes, 10 protein-coding genes and six tRNA genes contained a single intron, while three protein-coding genes possessed two introns. The gene rps12 was trans-spliced: the exon at the 5′ end was located in the LSC region, whereas the 3′ exon and intron were located in the IR regions. Moreover, the ψycf1 and ψycf2 were identified as pseudogenes because of the partial duplication (Table 1). The chloroplast genome of L. auriculata was deposited in GenBank (NCBI) with the accession number MW355498.
Comparative chloroplast genomics can provide insights into the mechanism of chloroplast evolution of plant species, including the structural rearrangements and gene features [54,55,56,57]. In our results, all the 12 Litsea chloroplast genomes exhibited the typical and canonical quadripartite structure akin to the majority of angiosperms and other taxa in the Lauraceae [5,6,58]. Variation in chloroplast genome size was very small among 12 Litsea species, ranging from 152,132 (L. szemaois) to 154,011 bp (L. garrettii). The length of their LSC region varied from 93,119 (L. szemaois) to 93,827 bp (L. elongata), their SSC region from 18,799 (L. monopetala) to 18,936 bp (L. mollis), and their IR region from 20,015 (L. auriculata) to 20,744 bp (L. garrettii). The overall GC content was 39.2% across all species except for L. japonica and L. elongata (39.1%) (Table 1). Gene contents of all 12 chloroplast genomes of Litsea species were also highly conserved, and they all encoded an identical set of 113 genes with 30 tRNA genes. Specifically, 15 genes possessed one intron (atpF, ndhA, ndhB, petB, petD, rpl2, rpl16, rpoC1, rps16, trnA–UGC, trnG–UCC, trnI–GAU, trnK–UUU, trnL–UAA, and trnV–UAC), and three genes had two introns (clpP, rps12, and ycf3) (Table 2).
We found a high level of similarity in genome-wide organization and evidence of close evolutionary relationships between the 12 Litsea species. The chloroplast genome comparisons of the 12 Litsea in mVISTA revealed a high sequence similarity between the species (Figure S1) and a lower sequence divergence observed in the IRs than LSC and SSC regions. This probably resulted from the copy correction between IR sequences by gene conversion [59] and indicated a conserved evolution process of these chloroplast genomes. The non-coding regions possessed higher variation than the coding regions, which was generally consistent with many previous studies on chloroplast genomes [5,6,9,27,28,29]. Moreover, the MAUVE alignment with the algorithm of progressive Mauve based on the 12 Litsea chloroplast genomes showed only one locally collinear block between all analyzed cp genomes, and all of the genes exhibited the same and consistent sequence order, with no gene re-arrangements or inversion events being detected in these genomes (Figure 2).
Additionally, the LSC/IRb (JLB), IRb/SSC (JSB), SSC/IRa (JSA), and IRa/LSC (JLA) boundaries/junctions of the 12 Litsea chloroplast genomes (Figure 3) across the 12 Litsea species were highly conserved with minor variations. Specifically, the genes ycf2, ndhF, ycf1, trnH–GUG, and psbA were distributed at the boundaries of LSC/IR and SSC/IR in all 12 Litsea species. The entire ycf2 crossed the LSC/IRb boundary corresponding to the pseudogene fragment ψycf2 with 3051–3186 bp truncated at the IRa/LSC border, and the gene ycf1 was located in the junctions of SSC/IRa among all 12 Litsea species, meaning their IRa boundaries all extended into ycf1 gene with the extension of the pseudogene fragment ψycf1 ranging from 1371 (L. mollis) to 1397 bp (L. monopetala) in the IRb/SSC (JSB) boundaries. In addition, the distance between IRb and gene ndhF ranged from 16 (L. garrettii) to 57 bp (L. japonica), and the length between IRa and gene trnH–GUG varied from 3 (L. mollis) to 23 bp (L. cubeba). Correspondingly, the expansion and contraction of the IR regions further explained the length variations in the chloroplast genomes of 12 Litsea species, and this phenomenon was quite common in most previously studied angiosperms and the other Lauraceae species [60,61,62].
The comparison results of codon preference indexed by the value of the relative synonymous codon usage (RSCU) showed that 12 Litsea species had a common preference for the usage of 30 codons within the 61 shared codons used in the chloroplast genomes (RSCU > 1), and the majority of them were characterized with adenine–thymine ending (Table S1), which was in line with the findings in other Lauraceae plants such as Cinnamomum bodinieri Levl. [63], indicating the relative conservation of the cp genes in the genus Litsea. However, the preference of the usage of three codons (AUA, GCC, and GGU) varied among the different species of this genus: Litsea auriculata, L. garrettii, L. glutinosa, L. monopetala, L. pungens, and L. szemaois seldom selected AUA when encoding isoleucine; L. cubeba, L. garrettii, L. glutinosa, and L. pungens preferred using GCC than other Litsea species; and L. elongata, L. mollis, and L. monopetala had bias on GGU (Table S1). This phenomenon might be explained by the biological characteristics of species or genes formed in the process of long-term evolution and adaptation to the environment and the result of the combined effects of selection, mutation, and drift. Other influencing factors may also include genome size, number of introns, tRNA abundance, and gene expression levels [64,65].

3.2. Enrichment of Chloroplast DNA Genetic Resources of Litsea

Microsatellites or SSRs are ubiquitous short tandem repeats often consisting of repetitive sequences/motifs of 1–6 bp in length and can be commonly found in the genomes of diverse organisms [66]. Owing to high level of polymorphisms, reproducibility, and abundance in plant genomes, they are widely used as one of the most important and valuable molecular markers for plant population genetic analysis, evolutionary studies, and molecular marker-assisted selection in plant breeding programs [58,67]. Here, we detected a total of 876 chloroplast SSRs within the chloroplast genomes of the 12 Litsea species by MISA, with the number of chloroplast SSRs ranging from 67 (L. mollis) to 80 (L. auriculata). The majority of these chloroplast SSRs (548 out of 876 or 62.56%) were mononucleotides composed mainly of short polyadenine (polyA) or polythymine (polyT) repeats and much less often contained guanine (G) or cytosine (C) tandem repeats. The remaining chloroplast SSRs were comprised of complex nucleotide repeats (13.81%), dinucleotides (11.99%), tetranucleotides (9.25%), trinucleotides (1.83%), pentanucleotides (0.34%), and hexanucleotides (0.11%) (Table 3).
Consistent with many previous reports [6,9], the chloroplast SSRs in Litsea were mainly located in the LSC regions (77.51%). A smaller percentage of the chloroplast SSRs occurred in the SSC (16.55%) and IR (5.94%) regions, respectively. These chloroplast SSRs were also richer in the non-coding regions (84.7%), such as intergenic spacers and introns, than coding regions (15.3%). The complete detail of repeat types and locations of cpSSRs in each Litsea species are listed in Table S2. Moreover, chloroplast SSRs exhibited high diversity and variations in copy numbers [56,68], and thus we identified 10 (6 mononucleotides, 3 trinucleotides, 1 tetranucleotides) polymorphic chloroplast SSRs within the 12 Litsea species (Table S2). These chloroplast SSR molecular markers developed in our study will be useful in population genetics and evolutionary studies of the genus Litsea as well as the molecular breeding and conservation of this and related genera.
Chloroplast DNA molecular markers have been extensively used for research on plant population genetics, phylogeny, phylogeography, and DNA barcodes for species identification and delimitation [67,69,70] by virtue of their advantages of low rates of nucleotide substitutions, usually uniparental inheritance and recombination [58,70]. On the basis of our comparative chloroplast genomics results, we found that the average nucleotide diversity in the intergenic spacer regions (mean π = 0.00835) was significantly higher than that in the protein coding gene regions (mean π = 0.00326) and introns (mean π = 0.00305) (Figure 4, Table S3). For the intergenic spacer (IGS) and intron regions, π values varied from 0.0006 (IGS, trnL–ndhB) to 0.05523 (IGS, psbE–petL), and the top 10 hypervariable regions were psbE–petL, trnH–psbA, petA–psbJ, ndhF–rpl32, ycf4–cemA, rpl32–trnL, ndhG–ndhI, psbC–trnS, trnE–trnT, and psbM–trnD (π > 0.0094) (Figure 4, Table S3). Regarding the protein coding gene regions, pairwise sequence divergence values (π) for each region ranged from 0.00021 (ndhB) to 0.00832 (ndhF), whereas five regions (ndhD, matK, ccsA, ycf1, ndhF) had remarkably high values (π > 0.005). Compared with other previous DNA barcoding studies of Lauraceae, for instance, the fragments of rpl32–trnL, ndhF–rpl32, and ycf1 had also been identified as variable regions in the tribes Laureae, Cinnamoneae, and Perseeae [28,63,71,72,73]. The rpl32–trnL region could also be used to further distinguish Alseodaphne, Alseodaphnopsis, and Dehaasia [72,73,74,75], while the psbA–trnH, matK, and psbC–trnS had proven to be effective for resolving the taxonomy and phylogeny of Machilus and Phoebe also [5,59,60]. In our study, the fragments ndhD, matK, ccsA, ycf1, ndhF, psbE–petL, trnH–psbA, petA–psbJ, ndhF–rpl32, ycf4–cemA, rpl32–trnL, ndhG–ndhI, psbC–trnS, trnE–trnT, and psbM–trnD were found to be particularly variable loci among Litsea chloroplast genomes, which could be employed for DNA barcoding or intraspecific delimitations, as well as for facilitating a better-resolved molecular phylogeny of Litsea species.

3.3. Phylogenetic Analysis

The phylogenetic trees resulting from the maximum likelihood and Bayesian inference analyses yielded identical tree topologies and showed that the studied Litsea species can be divided into the following four sub-clades on the basis of the node supports (Figure 5). Sub-clade I (BS = 100, PP = 1.00) contained two sibling Litsea species (L. mollis and L. cubeba) and Lindera obtusiloba. The sub-clade II, which contained Actinodaphne lancifolia (L. coreana), L. auriculata, L. szemaois, L. dilleniifolia, L. monopetala, L. garrettii, L. elongata, and L. japonica, was sister to sub-clade I. Within sub-clade II, the grouping of L. auriculata, as well as its sister taxon Actinodaphne lancifolia, was well supported with a of BS value 80 and PP of 1.00 (Figure 5). Six Litsea species (L. szemaois, L. dilleniifolia, L. monopetala, L. garrettii, L. elongata, L. japonica) formed a cluster into within sub-clade II, which was sister to the L. auriculataA. lancifolia pair. The other two Litsea species (L. glutinosa, L. acutivena), Lindera megaphylla, and Laurus nobilis formed sub-clade III with full support of BS 100 and PP 1.00, within which Litsea acutivena was sister to Lindera megaphylla (Figure 5). The sub-clade IV was sister to sub-clade III and contained the highly supported sister taxa of Litsea pungens and Lindera floribunda (with a BS of 100 and PP of 1.00), which were sister taxa to five other Lindera taxa (L. rubronervia, L. praecox, L. neesiana, L. sericea, L. reflexa) (Figure 5).
The new chloroplast genome phylogeny of Litsea enables us to elucidate the phylogenetic relationships among Litsea and related genera. First and foremost, the phylogenetic relationships between Litsea, Lindera, and Laurus have long been complex and controversial [3,76]. All three genera share similar morphological traits such as having umbellate inflorescences subtended by large involucral bracts, introrse anthers, the presence of two stipitate glands at the third whorl, and oval or spherical fruits [3,76]. Previously, Lindera and Litsea were differentiated on the basis of the number of anther cells [21], but our tree topology and recent work by Tian et al. [28] showed that this character may not be as taxonomically useful as previously thought for differentiating these genera. It is likely that future work with more comprehensive sampling of Litsea and Lindera will reveal the extend of the polyphyly of both genera.
Our phylogenetic tree also revealed the surprising relationship between Litsea auriculata and Actinodaphne lancifolia. A deeper look at the nomenclature of Actinodaphne lancifolia revealed that Litsea coreana H. Lév. is a synonym of this species [77]. The sessile umbels, persistent bracts, and scattered leaves of A. lancifolia in conjunction with our molecular data supports the reinstatement of Litsea coreana Lévl. as the correct name for the taxon. Additionally, our phylogenetic trees demonstrated full support (BS = 100, PP = 1.00) that the genus Actinodaphne was most related to the genus Neolitsea, specifically that Actinodaphne trichocarpa was sister to Neolitsea sericea and Neolitsea pallens, and Actinodaphne obovata was closely related with them, which is consistent with the research of Song et al. [6].
To sum up, our phylogenetic study resolved here revealed that the Litsea group (among the 12 species investigated) containing in the LaurusNeolitsea clade was polyphyletic and closely related to the genera Lindera and Laurus. Moreover, our ML and BI analyses with complete plastid genomes grouped all ingroup taxa into three clades (Laurus–Neolitsea clade, Cinnamomum–Ocotea clade, Machilus–Persea clade), and our phylogenetic topologies further supported the monophyly of the above three clades, which was also in accordance with the previously comprehensive phylogenetic tree constructed by Song et al. [6] on the basis of the cp genome data. Therefore, whole plastome sequencing has been proven to display higher resolution and reliability in resolving phylogenetic relationships at a wide range of taxonomic levels. On the other hand, to illuminate the more transparent systematic placements of Litsea and its related genera, we recommend more intensive and comprehensive taxa sampling and the combination of other sources of molecular data such as the transcriptomes and whole genomes [78,79]. Considering hybridization, gene introgression, polyploidization, and stochastic and systematic errors may also complicate the use of molecular data as the taxonomic proof of phylogeny [72,73,74,75], and thus we propose an integrated analysis of phenotypic traits such as inflorescence characters and molecular data for better resolving and clarifying the Lauraceae systematic research.

4. Conclusions

The chloroplast genome of Litsea auriculata was determined and charactered for the first time in this study, and we conducted the comparative plastid genomics and phylogenomics with that of 11 other Litsea species and 51 “core Lauraceae” species. Our results showed the high conservativeness of the plastid genomes of these 12 Litsea species regarding the canonical angiosperm quadripartite structure with no structural arrangement or gene inversion and the unbiased codon usage preferences. In the meantime, abundant genetic resources including 10 intergenic spacers (psbE–petL, trnH–psbA, petA–psbJ, ndhF–rpl32, ycf4–cemA, rpl32–trnL, ndhG–ndhI, psbC–trnS, trnE–trnT, and psbM–trnD), 5 protein coding genes (ndhD, matK, ccsA, ycf1, and ndhF), and 876 cpSSRs were developed here as potential DNA barcodes and variable molecular markers for further research of delimitation, phylogenetics, population genetics, and evolution of the genus Litsea, as well as the molecular breeding and conservation of this and the related genus. Our chloroplast genome phylogeny revealed that the 12 Litsea species investigated were polyphyletic, closely related to the genera Lindera and Laurus, and nested within the LaurusNeolitsea clade. In addition, our analyses with complete chloroplast genomes grouped all ingroup taxa into three clades (Laurus–Neolitsea clade, Cinnamomum–Ocotea clade, Machilus–Persea clade), and our phylogenetic topologies further supported the monophyly of the above three clades previously delimited by Song et al. [6]. We conclude that whole chloroplast genome sequencing enables the development of useful molecular markers and allows for higher resolution and reliability in resolving phylogenetic relationships of the Lauraceae at a wide range of taxonomic levels.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/f12060744/s1, Figure S1: Comparison of the cp genomes among the 12 Litsea species via mVISTA using annotation of L. acutivena as a reference. Table S1: RSCU analysis of protein-coding regions in 12 Litsea species. Table S2: The identified cpSSR loci and genus-targeted polymorphic cpSSRs in 12 Litsea species. Table S3: The nucleotide variability value (π) of both coding regions and non-coding regions (including intergenic spacers and introns) in 12 Litsea species.

Author Contributions

Conceptualization, P.L. and Z.W. (Zhongsheng Wang); methodology, Y.Z. (Yunyan Zhang), Y.T., and Y.Z. (Yuntian Zhang); software, Y.T., Y.Z. (Yunyan Zhang), D.Y.P.T., and Y.Z. (Yuntian Zhang); validation, Y.Z. (Yunyan Zhang), D.Y.P.T., and Z.W.; formal analysis, Z.W. (Zhengwei Wang), Y.Z. (Yunyan Zhang), and J.Z.; investigation, Y.Z. (Yuntian Zhang) and J.Z.; resources, P.L. and Z.W. (Zhongsheng Wang); data curation, Y.T., Y.Z. (Yuntian Zhang), and D.Y.P.T.; writing—original draft preparation, Y.Z. (Yunyan Zhang) and Y.T.; writing—review and editing, D.Y.P.T., P.L., and Z.W. (Zhengwei Wang); visualization, Y.Z. (Yunyan Zhang), Y.T., Y.Z. (Yuntian Zhang), and J.Z.; supervision, Z.W. (Zhengwei Wang), Y.Z. (Yuntian Zhang), and J.Z.; project administration, P.L. and Z.W. (Zhengwei Wang); funding acquisition, P.L. and Z.W. (Zhongsheng Wang) All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (grant no. 30970512) and the State Key Basic Research and Development Plan of China (grant no. 2017YFA0605104).

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; accession number: MW355498 (accessed on 11 May 2021).

Acknowledgments

The authors sincerely thank Mengyuan Zhang for his great help in collecting plant materials, Ruisen Lu for his kind instructions on drawing high-quality pictures, and Shan Lu for his helpful instruction for preserving the plant specimen at the herbarium of Nanjing University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chanderbali, A.S.; Werff, H.V.D.; Renner, S.S. Phylogeny and historical biogeography of Lauraceae: Evidence from the chloroplast and nuclear genomes. Ann. Mo. Bot. Gard. 2001, 88, 104–134. [Google Scholar] [CrossRef] [Green Version]
  2. Renner, S.S. Circumscription and phylogeny of the Laurales: Evidence from molecular and morphological data. Am. J. Bot. 1999, 86, 1301–1315. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Li, J.; Christophel, D.C.; Conran, J.G.; Li, H.W. Phylogenetic relationships within the ‘core’ Laureae (Litsea complex, Lauraceae) inferred from sequences of the chloroplast gene matK and nuclear ribosomal DNA ITS regions. Plant Syst. Evol. 2004, 246, 19–34. [Google Scholar] [CrossRef]
  4. Rohwer, J.G. Toward a phylogenetic classification of the Lauraceae: Evidence from matK sequences. Syst. Bot. 2000, 25, 60–71. [Google Scholar] [CrossRef]
  5. Song, Y.; Yu, W.B.; Tan, Y.H.; Liu, B.; Yao, X.; Jin, J.; Michael, P.; Yang, J.B.; Corlett, R.T. Evolutionary comparisons of the chloroplast genome in Lauraceae and insights into loss events in the Magnoliids. Genome Biol. Evol. 2017, 9, 2354–2364. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Song, Y.; Yu, W.B.; Tan, Y.H.; Jin, J.J.; Wang, B.; Yang, J.B.; Liu, B.; Corlett, R.T. Plastid phylogenomics improve phylogenetic resolution in the Lauraceae. J. Syst. Evol. 2020, 58, 423–439. [Google Scholar] [CrossRef]
  7. Rohwer, J.G.; Rudolph, B. Jumping genera: The phylogenetic positions of Cassytha, Hypodaphnis, and Neocinnamomum (Lauraceae) based on different analyses of trnK intron sequences. Ann. Mo. Bot. Gard. 2005, 92, 153–178. [Google Scholar]
  8. Chen, Y.C.; Li, Z.; Zhao, Y.X.; Gao, M.; Wang, J.Y.; Liu, K.W.; Wang, X.; Wu, L.W.; Jiao, Y.L.; Xu, Z.L.; et al. The Litsea genome and the evolution of the Laurel family. Nat. Commun. 2020, 11, 16–75. [Google Scholar] [CrossRef] [Green Version]
  9. Li, H.W. Lauraceae. In Flora of China; Science Press: Beijing, China, 1982; Volume 21, pp. 1–463. [Google Scholar]
  10. Dai, J.; Sun, B.N.; Xie, S.P.; Lin, Z.C.; Wen, W.W.; Wu, J.Y. Cuticular microstructure of Litsea cf. chunii from the Pliocene of Tengchong, Yunnan province. J. Lanzhou Univ. Nat. Sci. 2010, 46, 22–28. [Google Scholar]
  11. Huang, L.L.; Sun, J.; Jin, J.H.; Quan, C.; Alexei, A.O. Litseoxylon gen. nov. (Lauraceae): The most ancient fossil angiosperm wood with helical thickenings from southeastern Asia. Rev. Palaeobot. Palynol. 2018, 258, 223–233. [Google Scholar] [CrossRef]
  12. Li, H.W. The origin and evolution of  Litsea genera group (Laureae) in Lauraceae. Acta Bot. Yunnanica 1995, 17, 251–254. [Google Scholar]
  13. Huang, X.W.; Feng, Y.C.; Huang, Y. Potential cosmetic application of essential oil extracted from Litsea cubeba fruits from China. J. Essent. Oil Res. 2013, 25, 112–119. [Google Scholar] [CrossRef]
  14. Su, Y.C.; Ho, C.L. Essential oil compositions and antimicrobial activities of various parts of Litsea cubeba from Taiwan. Nat. Prod. Commun. 2016, 11, 515–518. [Google Scholar] [CrossRef] [Green Version]
  15. Nguyen, H.V.; Meile, J.C.; Lebrun, M.; Caruso, D.; Chu, K.S.; Sarter, S. Litsea cubeba leaf essential oil from Vietnam: Chemical diversity and its impacts on antibacterial activity. Lett. Appl. Microbiol. 2018, 66, 207–214. [Google Scholar] [CrossRef]
  16. Hasan, H.; Azad, M.S.A.; Islam, M.Z.; Rahman, S.M.; Islam, M.R.; Rahman, S.; Rahmatullah, M. Antihyperglycemic activity of methanolic extract of Litsea monopetala (Roxb.) Pers. leaves. Adv. Nat. Appl. Sci. 2014, 8, 51–55. [Google Scholar]
  17. Kim, C.S.; Lee, I.S.; Lee, Y.M.; Sohn, E.; Jo, K.; Kim, J.H.; Kim, J.S. Extract of Litsea japonica ameliorates blood–retinal barrier breakdown in db/db mice. Endocrine 2014, 46, 462–469. [Google Scholar] [CrossRef] [PubMed]
  18. Cai, X.T.; Wang, X.; Chen, Y.C.; Wang, Y.D.; Song, D.F.; Gu, Q. A natural biopreservative: Antibacterial action and mechanisms of Chinese Litsea mollis Hemsl. extract against Escherichia coli DH5α and Salmonella spp. J. Dairy Sci. 2019, 102, 9663–9673. [Google Scholar] [CrossRef] [PubMed]
  19. Bhuinya, T.; Singh, P.; Mukherjee, S.K. An account of the species of Litsea Lam. (Lauraceae) endemic to India. Bangladesh J. Plant Taxon. 2010, 17, 183–191. [Google Scholar] [CrossRef] [Green Version]
  20. Richter, H.G. Anatomie des sekundaren xylems und der rinde der Lauraceae. Sonderbd. Naurwissenschaftlichen Ver. Hambg. 1981, 5, 1–148. [Google Scholar]
  21. Li, H.W. Parallel evolution in Litsea and Lindera of Lauraceae. Acta Bot. Yunnanica 1985, 7, 129–135. [Google Scholar]
  22. Raj, B.; Werff, H.V.D. A contribution to the pollen morphology of Neotropical Lauraceae. Ann. Mo. Bot. Gard. 1988, 75, 130–167. [Google Scholar] [CrossRef]
  23. Werff, H.V.D.; Richter, H.G. Toward an improved classification of Lauraceae. Ann. Mo. Bot. Gard. 1996, 83, 409–418. [Google Scholar] [CrossRef]
  24. Fijridiyanto, I.A.; Murakami, N. Phylogeny of Litsea and related genera (Laureae–Lauraceae) based on analysis of rpb2 gene sequences. J. Plant Res. 2009, 122, 283–298. [Google Scholar] [CrossRef] [PubMed]
  25. Jung, J.; Kim, C.; Kim, J.H. Insights into phylogenetic relationships and genome evolution of subfamily Commelinoideae (Commelinaceae Mirb.) inferred from complete chloroplast genomes. BMC Genom. 2021, 22, 231. [Google Scholar] [CrossRef] [PubMed]
  26. Huang, R.; Xie, X.N.; Li, F.; Tian, E.W.; Chao, Z. Chloroplast genomes of two Mediterranean Bupleurum species and the phylogenetic relationship inferred from combined analysis with East Asian species. Planta 2021, 253, 81–84. [Google Scholar] [CrossRef]
  27. Ping, J.; Feng, P.; Li, J.; Zhang, R.; Su, Y.; Wang, T. Molecular evolution and SSRs analysis based on the chloroplast genome of Callitropsis funebris. Ecol. Evol. 2021, 11, 4786–4802. [Google Scholar] [CrossRef]
  28. Tian, X.Y.; Ye, J.W.; Song, Y. Plastome sequences help to improve the systematic position of trinerved Lindera species in the family Lauraceae. PeerJ 2019, 7, e7662. [Google Scholar] [CrossRef] [Green Version]
  29. Zhao, M.L.; Song, Y.; Ni, J.; Yao, X.; Tan, Y.H.; Xu, Z.F. Comparative chloroplast genomics and phylogenetics of nine Lindera species (Lauraceae). Sci. Rep. 2018, 8, 8844. [Google Scholar] [CrossRef] [Green Version]
  30. Geng, Q.F.; Sun, L.; Zhang, P.H.; Wang, Z.S.; Qiu, Y.X.; Liu, H.; Lian, C.L. Understanding population structure and historical demography of Litsea auriculata (Lauraceae), an endangered species in east China. Sci. Rep. 2017, 7, 17343. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Tanaka, H.; Yatsuhashi, S.; Yasuda, T.; Sato, M.; Sakai, E.; Xiao, C.; Murata, H.; Murata, J. A new amide from the leaves and twigs of Litsea auriculata. J. Nat. Med. 2009, 63, 331–334. [Google Scholar] [CrossRef]
  32. Fu, L.G.; Jin, J.M. China Plant Red Data Book: Rare and Endangered Plants; Science Press: Beijing, China, 1992; Volume 1, pp. 350–351. [Google Scholar]
  33. Deng, Y.; Pu, F.G. Conservation actualities of rare and endangered plants in Anhui Tianma National Nature Reserve and coping strategies. Anhui For. Sci. Technol. 2015, 41, 35–38. [Google Scholar]
  34. Cheng, H.Y.; Hu, X.L.; Zhou, M.Y.; Li, M.Q. Seedling techniques by sowing of Litsea auriculata. Anhui For. Sci. Technol. 2004, 4, 33–34. [Google Scholar]
  35. Wang, F.Z.; Xie, F.; Zhan, M.D.; Yuan, S.H. Preliminary study on Litsea auriculata Chien et Cheng tree twig cottage under full sunshine and automatic spray. J. Henan For. Sci. Technol. 2011, 31, 7–17. [Google Scholar]
  36. Doyle, J.J.; Doyle, J.L. A rapid DNA isolation procedure for small quantities of fresh leaf tissue. Phytochem. Bull. 1987, 19, 11–15. [Google Scholar]
  37. Jian, J.J.; Yu, W.B.; Yang, J.B.; Song, Y.; Li, D.Z. GetOrganelle: A fast and versatile toolkit for accurate de novo assembly of organelle genomes. Genome Biol. 2020, 21, 241. [Google Scholar] [CrossRef] [PubMed]
  38. Qu, X.J.; Moore, M.J.; Li, D.Z.; Yi, T.S. PGA: A software package for rapid, accurate, and flexible batch annotation of plastomes. Plant Methods 2019, 15, 50. [Google Scholar] [CrossRef] [Green Version]
  39. Schattner, P.; Brooks, A.N.; Lowe, T.M. The tRNAscan-SE, snoscan and snoGPS web servers for the detection of tRNAs and snoRNAs. Nucleic Acids Res. 2005, 33, 686–689. [Google Scholar] [CrossRef] [PubMed]
  40. Zheng, S.Y.; Poczai, P.; Hyvonen, J.; Tang, J.; Amiryousefi, A. Chloroplot: An online program for the versatile plotting of organelle genomes. Front. Genet. 2020, 11, 576124. [Google Scholar] [CrossRef] [PubMed]
  41. Chris, M.; Michael, B.; Schwartz, J.R.; Alexander, P.; Rubin, E.M.; Frazer, K.A.; Pachter, L.S.; Inna, D. VISTA: Visualizing global DNA sequence alignments of arbitrary length. Bioinformatics 2000, 16, 1046–1047. [Google Scholar]
  42. Darling, A.C.E.; 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] [Green Version]
  43. Amiryousefi, A.; Hyvonen, J.; Poczai, P. IRscope: An online program to visualize the junction sites of chloroplast genomes. Bioinformatics 2018, 34, 3030–3031. [Google Scholar] [CrossRef]
  44. Sharp, P.M.; Li, W.H. The codon adaptation index–a measure of directional synonymous codon usage bias, and its potential applications. Nucleic Acids Res. 1987, 15, 1281–1295. [Google Scholar] [CrossRef] [Green Version]
  45. John, F.P. Analysis of Codon Usage. Ph.D. Thesis, The University of Nottingham, Nottingham, UK, 1999. [Google Scholar]
  46. Beier, S.; Thiel, T.; Münch, T.; Scholz, U.; Mascher, M. MISA-web: A web server for microsatellite prediction. Bioinformatics 2017, 33, 2583–2585. [Google Scholar] [CrossRef] [Green Version]
  47. Rozas, J.; Ferrer-Mata, A.; Sánchez-DelBarrio, J.C.; Guirao-Rico, S.; Librado, P.; Ramos-Onsins, S.E.; Sánchez-Gracia, A. DnaSP 6: DNA sequence polymorphism analysis of large datasets. Mol. Biol. Evol. 2017, 34, 3299–3302. [Google Scholar] [CrossRef]
  48. Nakamura, T.; Yamada, K.D.; Tomii, K.; Katoh, K. Parallelization of MAFFT for large-scale multiple sequence alignments. Bioinformatics 2018, 34, 2490–2492. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Stamatakis, A. RAxML version 8: A tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 2014, 30, 1312–1313. [Google Scholar] [CrossRef]
  50. Ronquist, F.; Huelsenbeck, J.P. MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics 2003, 19, 1572–1574. [Google Scholar] [CrossRef] [Green Version]
  51. Guindon, S.; Gascuel, O. A simple, fast, and accurate algorithm to estimate large phylogenies by maximum likelihood. Syst. Biol. 2003, 52, 696–704. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Darriba, D.; Taboada, G.L.; Doallo, R.; Posada, D. jModelTest2: More models, new heuristics and parallel computing. Nat. Methods 2012, 9, 772. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Letunic, I.; Bork, P. Interactive Tree of Life (iTOL) v4: Recent updates and new developments. Nucleic Acids Res. 2019, 47, 256–259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Zhao, Y.B.; Yin, J.L.; Guo, H.Y.; Zhang, Y.Y.; Xiao, W.; Sun, C.; Wu, J.Y.; Qu, X.B.; Yu, J.; Wang, X.M.; et al. The complete chloroplast genome provides insight into the evolution and polymorphism of Panax ginseng. Front. Plant Sci. 2015, 5, 696. [Google Scholar] [CrossRef] [Green Version]
  55. Li, P.; Lu, R.S.; Xu, W.Q.; Ohi-Toma, T.; Cai, M.Q.; Qiu, Y.X.; Cameron, K.M.; Fu, C.X. Comparative genomics and phylogenomics of East Asian tulips (Amana, Liliaceae). Front. Plant Sci. 2017, 8, 451. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Wu, Y.; Liu, F.; Yang, D.G.; Li, W.; Zhou, X.J.; Pei, X.Y.; Liu, Y.G.; He, K.L.; Zhang, W.S.; Ren, Z.Y.; et al. Comparative chloroplast genomics of Gossypium species: Insights into repeat sequence variations and phylogeny. Front. Plant Sci. 2018, 9, 376. [Google Scholar] [CrossRef] [Green Version]
  57. She, R.; Zhao, P.; Zhou, H.; Yue, M.; Zhang, S. Complete chloroplast genomes of Liliaceae (s.l.) species: Comparative genomic and phylogenetic analyses. Nord. J. Bot. 2020, 38, e02477. [Google Scholar] [CrossRef]
  58. Zhang, Y.Y.; Shi, E.; Yang, Z.P.; Geng, Q.F.; Qiu, Y.X.; Wang, Z.S. Development and application of genomic resources in an endangered palaeoendemic tree, Parrotia subaequalis (Hamamelidaceae) from eastern China. Front. Plant Sci. 2018, 9, 246. [Google Scholar] [CrossRef] [Green Version]
  59. Khakhlova, O.; Bock, R. Elimination of deleterious mutations in plastid genomes by gene conversion. Plant J. 2006, 46, 85–94. [Google Scholar] [CrossRef] [PubMed]
  60. Song, Y.; Dong, W.; Bing, L.; Chao, X.; Yao, X.; Gao, J.; Corlett, R.T. Comparative analysis of complete chloroplast genome sequences of two tropical trees Machilus yunnanensis and Machilus balansae in the family Lauraceae. Front. Plant Sci. 2015, 6, 662. [Google Scholar] [CrossRef] [Green Version]
  61. Song, Y.; Yao, X.; Tan, Y.; Gan, Y.; Corlett, R.T. Complete chloroplast genome sequence of the avocado: Gene organization, comparative analysis, and phylogenetic relationships with other Lauraceae. Can. J. For. Res. 2016, 46, 1293–1301. [Google Scholar] [CrossRef]
  62. Kim, K.J.; Lee, H.L. Complete chloroplast genome sequences from Korean ginseng (Panax schinseng Nees) and comparative analysis of sequence evolution among 17 vascular plants. DNA Res. 2004, 11, 247–261. [Google Scholar] [CrossRef] [PubMed]
  63. Huang, J.F.; Li, L.; Werff, H.V.D.; Li, H.W.; Rohwer, J.H.; Crayn, D.M.; Meng, H.H.; Merwe, M.V.D.; Conran, J.H.; Li, J. Origins and evolution of cinnamon and camphor: A phylogenetic and historical biogeographical analysis of the Cinnamomum group (Lauraceae). Mol. Phylogenet. Evol. 2016, 96, 33–44. [Google Scholar] [CrossRef] [PubMed]
  64. Plotkin, J.B.; Kudla, G. Synonymous but not the same: The causes and consequences of codon bias. Nat. Rev. Genet. 2011, 12, 32–42. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Goodman, D.B.; Church, G.M.; Kosuri, S. Causes and effects of N-terminal codon bias in bacterial genes. Science 2013, 342, 475–479. [Google Scholar] [CrossRef] [Green Version]
  66. Xia, E.H.; Yao, Q.Y.; Zhang, H.B.; Jiang, J.J.; Zhang, L.P.; Gao, L.Z. CandiSSR: An efficient pipeline used for identifying candidate polymorphic SSRs based on multiple assembled sequences. Front. Plant Sci. 2016, 6, 1171. [Google Scholar] [CrossRef] [Green Version]
  67. Jansen, R.K.; Cai, Z.; Raubeson, L.A.; Daniell, H.; de Pamphilis, C.W.; Leebens-Mack, J. Analysis of 81 genes from 64 plastid genomes resolves relationships in angiosperms and identifies genome-scale evolutionary patterns. Proc. Natl. Acad. Sci. USA 2007, 104, 19369–19374. [Google Scholar] [CrossRef] [Green Version]
  68. Pauwels, M.; Vekemans, X.; Godé, C.; Frérot, H.; Castric, V.; Saumitou-Laprade, P. Nuclear and chloroplast DNA phylogeog-raphy reveals vicariance among European populations of the model species for the study of metal tolerance, Arabidopsis halleri (Brassicaceae). New Phytol. 2012, 193, 916–928. [Google Scholar] [CrossRef] [PubMed]
  69. Moore, M.J.; Soltis, P.S.; Bell, C.D.; Burleigh, J.G.; Soltis, D.E. Phylogenetic analysis of 83 plastid genes further resolves the early diversification of eudicots. Proc. Natl. Acad. Sci. USA 2010, 107, 4623–4628. [Google Scholar] [CrossRef] [Green Version]
  70. Shaw, J.; Shafer, H.L.; Leonard, O.R.; Kovach, M.J.; Schorr, M.; Morris, A.B. Chloroplast DNA sequence utility for the lowest phylogenetic and phylogeographic inferences in angiosperms: The tortoise and the hare IV. Am. J. Bot. 2014, 101, 1987–2004. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Hinsinger, D.D.; Strijk, J.S. Toward phylogenomics of Lauraceae: The complete chloroplast genome sequence of Litsea glutinosa (Lauraceae), an invasive tree species on Indian and Pacific Ocean islands. Plant Gene 2017, 9, 71–79. [Google Scholar] [CrossRef]
  72. Rohwer, J.G.; Lj, J.; Rudolph, B.; Schmidt, S.A.; Werff, H.V.D.; Li, H.W. Is Persea (Lauraceae) monophyletic? Evidence from nuclear ribosomal ITS sequences. Taxon 2009, 58, 1153–1167. [Google Scholar] [CrossRef]
  73. Li, L.; Li, J.; Rohwer, J.G.; Werff, H.V.D.; Wang, Z.H.; Li, H.W. Molecular phylogenetic analysis of the Persea group (Lauraceae) and its biogeographic implications on the evolution of tropical and subtropical Amphi-Pacifi disjunctions. Am. J. Bot. 2011, 98, 1520–1536. [Google Scholar] [CrossRef] [PubMed]
  74. Mo, Y.Q.; Li, L.; Li, J.W.; Rohwer, J.G.; Li, H.W.; Li, J. Alseodaphnopsis: A new genus of Lauraceae based on molecular and morphological evidence. PLoS ONE 2017, 12, e0186545. [Google Scholar]
  75. Song, Y.; Yao, X.; Lu, B.; Tan, Y.H.; Corlett, R.T. Complete plastid genome sequences of three tropical Alseodaphne trees in the family Lauraceae. Holzforschung 2018, 72, 337–345. [Google Scholar] [CrossRef]
  76. Li, J.; Li, H.W. Advances in Lauraceae systematic research on the world scale. Acta Bot. Yunnanica 2004, 26, 1–11. [Google Scholar]
  77. Friedrich, F. Centralblatt für Sammlung und Veroffentlichung von Einzeldiagnosen neuer Pflanzen. Repert. Spec. Nov. Regni Veg. 1912, 10, 370. [Google Scholar]
  78. Zhang, Z.; He, Z.W.; Xu, S.H.; Li, X.N.; Guo, W.X.; Yang, Y.C.; Zhong, C.R.; Zhou, R.C.; Shi, S.H. Transcriptome analyses provide insights into the phylogeny and adaptive evolution of the mangrove fern genus. Acrostichum. Sci. Rep. 2016, 6, 35634. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Wang, N.; Kelly, L.J.; McAllister, H.A.; Zohren, J.; Buggs, R.J.A. Resolving phylogeny and polyploid parentage using genus-wide genome-wide sequence data from birch trees. Mol. Phylogenet. Evol. 2021, 160, 107126. [Google Scholar] [CrossRef]
Figure 1. Circular map of chloroplast genome of Litsea auriculata with annotated genes. Genes shown inside and outside of the circle are transcribed in clockwise and counterclockwise directions, respectively. Genes belonging to different functional groups are color-coded. The GC and AT content are denoted by the dark gray and light gray colors in the inner circle, respectively. LSC, SSC, and IR are large single-copy region, small single-copy region, and inverted repeat region, respectively.
Figure 1. Circular map of chloroplast genome of Litsea auriculata with annotated genes. Genes shown inside and outside of the circle are transcribed in clockwise and counterclockwise directions, respectively. Genes belonging to different functional groups are color-coded. The GC and AT content are denoted by the dark gray and light gray colors in the inner circle, respectively. LSC, SSC, and IR are large single-copy region, small single-copy region, and inverted repeat region, respectively.
Forests 12 00744 g001
Figure 2. Alignment of 12 Litsea chloroplast genomes. The L. acutivena genome is shown at the top as the reference genome. Within each of the alignment, local collinear blocks are represented by blocks of the same color connected by lines.
Figure 2. Alignment of 12 Litsea chloroplast genomes. The L. acutivena genome is shown at the top as the reference genome. Within each of the alignment, local collinear blocks are represented by blocks of the same color connected by lines.
Forests 12 00744 g002
Figure 3. Comparison of the borders of the IR, SSC, and LSC regions among 12 chloroplast genomes of Litsea. JLB, JSB, JSA, and JLA represent the junctions of LSC/IRb, IRb/SSC, SSC/IRa, and IRa/LSC, respectively.
Figure 3. Comparison of the borders of the IR, SSC, and LSC regions among 12 chloroplast genomes of Litsea. JLB, JSB, JSA, and JLA represent the junctions of LSC/IRb, IRb/SSC, SSC/IRa, and IRa/LSC, respectively.
Forests 12 00744 g003
Figure 4. Nucleotide variability (π) values of 12 Litsea chloroplast genomes.
Figure 4. Nucleotide variability (π) values of 12 Litsea chloroplast genomes.
Forests 12 00744 g004
Figure 5. Complete chloroplast genome phylogenetic tree of 63 species inferred from maximum likelihood (ML) and Bayesian inference (BI) analysis. Numbers above the lines represent ML bootstrap values (BS) and BI posterior probabilities (PP).
Figure 5. Complete chloroplast genome phylogenetic tree of 63 species inferred from maximum likelihood (ML) and Bayesian inference (BI) analysis. Numbers above the lines represent ML bootstrap values (BS) and BI posterior probabilities (PP).
Forests 12 00744 g005
Table 1. Comparison of complete plastid genomes of 12 Litsea species.
Table 1. Comparison of complete plastid genomes of 12 Litsea species.
SpeciesGenBank IDWhole Sequence Length (bp)Length of LSC Region (bp)Length of IR Region (bp)Length of SSC Region (bp)Total GC Content (%)Total Number of GenesTotal Number of CDS GenesTotal Number of tRNA GenesTotal Number of rRNA Genes
L. acutivenaNC_050362152,71893,67720,06618,90939.211379304
L. auriculataMW355498152,37793,53320,01518,81439.211379304
L. cubebaNC_048954152,72593,67420,06418,92339.211379304
L. dilleniifoliaNC_050363152,29893,21820,09418,89239.211279294
L. elongataNC_050364152,79393,82720,06618,83439.111379304
L. garrettiiMN698967154,01193,69820,74418,82539.211379304
L. glutinosaKU382356152,61893,69020,06118,80639.211379304
L. japonicaNC_045267152,71893,69720,06618,88939.111379304
L. mollisNC_050366152,73693,65520,06318,93639.211379304
L. pungensNC_050368152,65593,52020,13118,87339.211379304
L. szemaoisNC_050369152,13293,11920,09018,83339.211379304
L.monopetalaNC_050367152,70593,75820,07418,79939.211379304
Table 2. List of genes in the chloroplast genomes of Litsea.
Table 2. List of genes in the chloroplast genomes of Litsea.
Groups of GenesNames of Genes
Ribosomal RNAsrrn4.5 (×2), rrn5 (×2), rrn16 (×2), rrn23 (×2)
Transfer RNAs* trnA-UGC (×2), ^ trnC-GCA, trnD-GUC, trnE-UUC, trnF-GAA, trnG-GCC, * trnG-UCC, trnH-GUG, trnI-CAU, * trnI-GAU (×2), * trnK-UUU, trnL-CAA (×2), * trnL-UAA, trnL-UAG, trnfM-CAU, 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
Photosystem IpsaA, psaB, psaC, psaI, psaJ
Photosystem IIpsbA, psbB, psbC, psbD, psbE, psbF, psbH, psbI, psbJ, psbK, psbL, psbM, psbN, psbT, psbZ
CytochromepetA, * petB, * petD, petG, petL, petN
ATP synthaseatpA, atpB, atpE, * atpF, atpH, atpI
RubiscorbcL
NADH dehydrogenease* ndhA, * ndhB (×2), ndhC, ndhD, ndhE, ndhF, ndhG, ndhH, ndhI, ndhJ, ndhK
ATP-dependent protease subunit P** clpP
Chloroplast envelop membrane proteincemA
Large units* rpl2 (×2), rpl14, * rpl16, rpl20, rpl22, rpl23, rpl32, rpl33, rpl36
Small unitsrps2, rps3, rps4, rps7 (×2), rps8, rps11, ** rps12, rps14, rps15, * rps16, rps18, rps19
RNA polymeraserpoA, rpoB, * rpoC1, rpoC2
Translational initiation factorinfA
Miscellaneous proteinsmatK, accD, ccsA
Hypothetical proteins and conserved reading frame** ycf3, ycf4, ycf1, ycf2
Pseudogeneψycf1, ψycf2
Note: Asterisks (*) before gene names indicate one intron containing genes, and double asterisks (**) indicate two introns in the gene. ×2 indicates genes duplicated in the IR regions. ^ indicates does not exist in L. dilleniifolia. Pseudogene is represented by ψ.
Table 3. Summary of the simple sequence repeats (SSR) in 12 Litsea species.
Table 3. Summary of the simple sequence repeats (SSR) in 12 Litsea species.
SpeciesSSR NumbersP1
Loci (N)
P2
Loci (N)
P3
Loci (N)
P4
Loci (N)
P5
Loci (N)
P6
Loci (N)
Pc
Loci (N)
LSCSSCIR
L. acutivena7246918855134
L. auriculata80559151062144
L. cubeba71439181955124
L. dilleniifolia71429281055124
L. elongata74468261258124
L. garrettii75469171258134
L. glutinosa72451117175796
L. japonica72447161456124
L. mollis67418171953104
L. pungens7348926858114
L. szemaois6943827954114
L. monopetala804991621358166
Total/
Percentage
876
(100%)
548
(62.56%)
105
(11.99%)
16
(1.83%)
81
(9.25%)
1
(0.11%)
4
(0.46%)
121
(13.81%)
679
(77.51%)
145
(16.55%)
52
(5.94%)
Note: P1–P6 represent SSRs of the mononucleotide, dinucleotide, trinucleotide, tetranucleotide, pentanucleotide, and hexanucleotide types, respectively. Pc represents a complex nucleotide repeat. LSC and SSC denote large and small single copy regions, respectively. The last row represents the percentage for each type of SSR. “–“ indicates no data.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, Y.; Tian, Y.; Tng, D.Y.P.; 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. https://doi.org/10.3390/f12060744

AMA Style

Zhang Y, Tian Y, Tng DYP, Zhou J, Zhang Y, Wang Z, Li P, Wang Z. Comparative Chloroplast Genomics of Litsea Lam. (Lauraceae) and Its Phylogenetic Implications. Forests. 2021; 12(6):744. https://doi.org/10.3390/f12060744

Chicago/Turabian Style

Zhang, Yunyan, Yongjing Tian, David Y. P. Tng, Jingbo Zhou, Yuntian Zhang, Zhengwei Wang, Pengfu Li, and Zhongsheng Wang. 2021. "Comparative Chloroplast Genomics of Litsea Lam. (Lauraceae) and Its Phylogenetic Implications" Forests 12, no. 6: 744. https://doi.org/10.3390/f12060744

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