Next Article in Journal
Resisting the Resistance: Navigating BTK Mutations in Chronic Lymphocytic Leukemia (CLL)
Previous Article in Journal
The Stilbene Synthase Family in Arachis: A Genome-Wide Study and Functional Characterization in Response to Stress
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Comparative Analysis of Six Chloroplast Genomes in Chenopodium and Its Related Genera (Amaranthaceae): New Insights into Phylogenetic Relationships and the Development of Species-Specific Molecular Markers

1
College of Life Sciences, Yantai University, Yantai 264005, China
2
College of Grassland Science, Qingdao Agricultural University, Qingdao 266109, China
*
Authors to whom correspondence should be addressed.
Genes 2023, 14(12), 2183; https://doi.org/10.3390/genes14122183
Submission received: 31 October 2023 / Revised: 27 November 2023 / Accepted: 4 December 2023 / Published: 6 December 2023
(This article belongs to the Section Plant Genetics and Genomics)

Abstract

:
Species within the genus Chenopodium hold significant research interest due to their nutritional richness and salt tolerance. However, the morphological similarities among closely related species and a dearth of genomic resources have impeded their comprehensive study and utilization. In the present research, we conduct the sequencing and assembly of chloroplast (cp) genomes from six Chenopodium and related species, five of which were sequenced for the first time. These genomes ranged in length from 151,850 to 152,215 base pairs, showcased typical quadripartite structures, and encoded 85 protein-coding genes (PCGs), 1 pseudogene, 37 tRNA genes, and 8 rRNA genes. Compared with the previously published sequences of related species, these cp genomes are relatively conservative, but there are also some interspecific differences, such as inversion and IR region contraction. We discerned 929 simple sequence repeats (SSRs) and a series of highly variable regions across 16 related species, predominantly situated in the intergenic spacer (IGS) region and introns. The phylogenetic evaluations revealed that Chenopodium is more closely related to genera such as Atriplex, Beta, Dysphania, and Oxybase than to other members of the Amaranthaceae family. These lineages shared a common ancestor approximately 60.80 million years ago, after which they diverged into distinct genera. Based on InDels and SNPs between species, we designed 12 pairs of primers for species identification, and experiments confirmed that they could completely distinguish 10 related species.

1. Introduction

In the most recent APG IV (Angiosperm Phylogeny Group) plant classification system, the genus Chenopodium is categorized under the Amaranthaceae family, which encompasses approximately 250 species globally [1,2]. Chenopodium plants are mostly annual or perennial herbs, rarely semi-shrubs, with cystic hairs (powder) or cylindrical hairs, a small quantity of glandular hairs or completely hairless, and very low odor. They generally grow in arid or semi-arid areas and have tolerance to multiple stresses, including a strong salt tolerance. These characteristics make Chenopodium plants have extremely high scientific research value [3,4,5,6]. Chenopodium plants also have a high nutritional value. They can not only be grown as pseudocereals but also eaten as leafy vegetables [7]. Among them, the leaves of C. album are an important source of vitamins and trace elements in certain areas [8]. C. quinoa, known as a “superfood”, is not only rich in fatty acids, vitamins, minerals, dietary fiber, and proteins, containing many amino acids, but also rich in carotenoids, vitamin C, and phenolic compounds. These compounds contribute to the prevention of conditions such as allergies and cardiovascular diseases [9,10,11,12,13]. In some arid areas, they are also used as medicinal plants, fodder, and dye raw materials [14].
In addition, some plants of the related genera of Chenopodium also have various utilization values. For example, some plants from Atriplex can revegetate arsenic-contaminated saline–alkali soils, while effectively accumulating and preventing the spread of arsenic and reducing the risk of human exposure to arsenic [15]. Some compounds extracted from Atriplex can protect the chromosome damage caused by gamma radiation and can be used to develop new radioprotective drugs [16]. D. ambrosioides exhibits high antimicrobial and antioxidant activities and can enhance the antibacterial activity with the synergistic effect of antibiotics. At the same time, some compounds contained in it can achieve the purpose of killing mosquitoes by destroying the nervous system of insects and inhibiting the activity of key enzymes [17,18,19].
However, due to the high similarity and polymorphism of leaf shape, floral structure, and seed morphology and the existence of possible hybrid species, the classification and identification of Chenopodium and its related genera have become extremely complex and controversial, which severely hinders the development of related scientific research and the efficient utilization of Chenopodium plants [20,21,22,23].
Previous investigations in Chenopodium primarily relied on conventional molecular markers for phylogenetic analysis and species identification. These markers include nuclear DNA amplification fragment length polymorphism, ribosomal DNA (rDNA) internally transcribed spacer sequence (ITS) variation, and chloroplast DNA (cpDNA) fragments [24,25,26,27,28]. However, these research methods have limitations due to the restricted information they provide regarding mutation sites. Consequently, these approaches may not fully address the requirements of related research [29].
Several scholars have suggested employing the complete cp genome as a prospective solution to address this issue [30,31,32,33]. The cp genome of closely related species is highly conserved in structure, yet it exhibits subtle differences. Furthermore, the longer sequences provide a wealth of mutation site information, thereby enhancing its capability for species identification [34,35,36,37]. Simultaneously, the cp genome of angiosperms is mostly maternally inherited, reducing gene pollution and ensuring the accurate transmission and storage of genetic information [38,39]. Some researchers have used the cp genome to mine mutation sites to distinguish closely related species (especially some weed relatives), or have used the cp genome to elaborate new perspectives on plant phylogenetic relationships and taxonomy. The importance and efficiency of chloroplast genomes in related research have been confirmed by more and more studies [40,41].
In this study, we obtain the complete cp genomes of six closely related species of Chenopodium and describe their basic characteristics in detail. By comparing with the sequence information of other species that have been published, we studied their phylogenetic relationship and divergence time, which provided valuable clues to understand the evolutionary history of the genus Chenopodium. In addition, we also developed a set of highly efficient molecular marker primers that can be used to accurately identify these closely related species; these molecular markers could be widely applied to assist in screening and identifying Chenopodium plants with specific genomic features, thereby accelerating the breeding process for new varieties of crops, such as C. quinoa.

2. Materials and Methods

2.1. Plant Material and DNA Isolation

We collected the young leaves of 6 plant species (A. patens, A. prostrata, C. album, C. berlandieri, C. bryoniifolium, and L. franchetii) from Yantai, Shandong Province, China (121°44′ E, 37°51’ N, 264,003). The identification of samples was based on the classification and external morphological diagnosis of the samples according to the Chinese Virtual Herbarium (CVH) and Flora of China [42]. The specimen voucher materials were preserved in the College of Life Sciences, Yantai University. The total DNA was obtained from plants through the modified CTAB method [43]. The quality of DNA was assessed using a NanoDrop spectrophotometer 2000 (Thermo Scientific, Waltham, MA, USA), and DNA integrity was verified through 1% gel electrophoresis.

2.2. DNA Sequencing, Assembly, and Annotation

To create libraries, 350 bp fragments were generated using paired-end sequencing on the Illumina HiSeq 2500 platform (Illumina Inc., San Diego, CA, USA).
To obtain clean data, Fastp 0.23.1 was used to remove ambiguous bases or low-quality bases from raw sequencing reads [44]. GetOrganelle v1.7.1 was utilized for the assembly of the cp genome, with the k-mer length parameter configured as follows: −k 21, 45, 65, 85, and 105 [45]. Bandage v0.81 and Gepard v1.40 were used to examine the cp gene composition loop and collinearity, respectively [46,47].
The cp genome was annotated by employing CPGAVAS2 and GeSeq. Subsequently, Apollo was used for proofreading [48,49,50]. To locate and annotate tRNA genes, tRNAscan-SE v2.0.7 was employed for analysis, and CPgview was employed to check the annotation of the rps12 gene [51,52]. Finally, the complete cp genome map was generated through Chloroplot [53]. The annotated cp genome sequence was stored in GenBank with the accession number OR374019-OR374024.

2.3. Genome Structure and Comparison Analysis

Combined with the published sequence information in the NCBI database, a sum of 16 complete cp genomes of Chenopodium and related species were used for comparative analysis. The analysis of SSRs was conducted using MISA v2.1 [54]. The parameters were set as follows: mononucleotides ≥ 10; dinucleotides ≥ 5; trinucleotides ≥ 4; and tetranucleotides, pentanucleotides, and hexanucleotides ≥ 3. REPuter was used to detect the four types of interspersed repeats, including forward repeats, palindromic repeats, reverse repeats, and complementary repeats [55]. The minimum repeat fragment size was set to 30 bp, and the Hamming distance was set to 3. The relative synonymous codon usage (RSCU) was calculated by CodonW v1.4.2, and the results were subsequently visualized by Tbtools v1.113 [56,57].
The Perl script CPJSdraw v1.0 was employed to compare and visualize the inverted repeat (IR) region boundary characteristics between the cp genomes of the 16 species of Chenopodium and its related genera [58]. Using A. centralasiatica as the reference genome, in the Shuffle-LAGAN model of mVISTA, the cp genomes were compared to understand the sequence variations among closely related species [59,60]. DnaSP v5.10 was employed to analyze the nucleotide diversity of 111 gene regions and 135 intergenic regions (including introns) shared by all closely related species to obtain high information content sites [61].

2.4. Phylogenomic Analysis and Divergence Time Estimation

For the purpose of acquiring an in-depth understanding of the phylogenetic relationship among the species of Chenopodium and its related genera, data were downloaded from other species under the Amaranthaceae from the NCBI database as a supplement. In total, 155 complete cp genome sequences from 69 species in 24 genera were obtained, and their sequence starting points and small single copy (SSC) region directions were corrected and unified. Among them, five species of Plumbaginaceae were employed as outgroups. The alignment of all cp genome sequences was performed using MAFFT v7.50 and then trimmed using trimAl v1.41; the phylogenetic relationships were inferred based on the maximum likelihood (ML) and Bayesian inference (BI) [62,63]. The best nucleic acid replacement model was selected using Model Finder, and the ML tree was reconstructed using IQ-TREE 2.03 under the TVM + F + R3 model [64,65]. The ML branch support was evaluated by 1000 ultrafast guided repetitions.
The BI tree was reconstructed using MrBayes 3.2.7 [66]. A total of 20 million generations were iterated with the GTR + F + I + G4 model, with one cold chain and three heated chains, and the initial 25% of the sampled data was discarded as burn-in. Finally, the average standard deviation of split frequencies (ASDSF) < 0.01, the effective sample size was >200, and the potential scale reduction factor (PSRF) was close to 1, reaching convergence.
Based on the ML tree, the divergence time of the 16 Chenopodium related species (each species retains only one sequence) in Clade I was calculated using the Markov chain Monte Carlo (MCMC) tree program in PAML [67]. According to previous studies, five fossil calibration points were used to limit each node, of which 81.3–97.0 Ma was used as the age of the root node [68]. The independent rates that follow the lognormal distribution were selected as the clock model. We selected HKY85 as the nucleic acid replacement model, in which the alpha for gamma rates at the sites was set to 0.5. The birth–death process was used to generate uniform node age priors in the tree, using the default parameters (l = 1, m = 1, and s = 0.1). We ran the Markov chain Monte Carlo program for 200 million generations, sampling every 100 cycles. The first 25% of the samples was discarded as burn-in. Finally, Tracer v1.72 was used to check the output file to confirm the convergence based on ESS > 200.

2.5. Development and Validation of InDel and dCAPS Markers

Based on the variable regions in the cp genomes of Chenopodium and its related species, a series of molecular markers were developed to distinguish a total of 10 related species. Primer3 4.1.0 and dCAPS Finder 2.0 were used to design primers for the amplification of InDel markers and SNP-based dCAPS markers in the conserved regions near the variant sites [69,70].
Polymerase chain reaction (PCR) was performed in a total volume of 10 μg. It contained 5 μg 2 × Taq Master Mix (Vazyme, Nanjing, China), 1 μg DNA template, 0.5 μg of upstream and downstream primers, and 3 μg of ddH2O. The PCR was executed with a T100TM thermal cycler (Bio-Rad, Hercules, CA, USA). The settings were established as: pre-denaturation at 95 °C for 5 min, followed by 35 cycles of denaturation at 95 °C for 15 s, annealing at 55–58 °C for 15 s, extension at 72 °C for 1–3 min, and finally extension at 72 °C for 7 min. The PCR products were electrophoresed in 2–3% agarose gel, and the PCR products of the dCAPS marker were first reacted with the restriction enzyme (BsrI) at 65 °C for 1 h. Finally, Gel DocTM XR + gel imager (Bio-Rad, Hercules, CA, USA) was used for visualization.

3. Results

3.1. Genome Structure

We sequenced and assembled the cp genomes of six Chenopodium species and related genera, and then compared these to the existing cp genomes of the associated genera in the NCBI database. The lengths of these newly sequenced cp genomes varied from 151,850 bp (A. patens) to 152,215 bp (A. prostrata). Both Chenopodium and Atriplex cp genomes presented a typical quadripartite structure (Figure 1). The IR region (varying from 25,161 bp (A. patens) to 25,192 bp (C. bryoniifolium)) were separated by a large single copy (LSC) region (ranging from 83,675 bp (C. bryoniifolium) to 83,887 bp (A. prostrata)) and an SSC region (ranging from 17,811 bp (A. patens) to 18,131 bp (C. berlandieri)). The overall GC content in the cp genomes of Chenopodium and Atriplex was found to range between 37.25% and 37.3%. Nevertheless, there was a non-uniform distribution of the GC content in distinct genomic regions. Precisely, the GC content within the IR region (42.72–42.81%) exceeded that of both the LSC (35.30–35.40%) and the SSC regions (30.82–31.02%), which is consistent with other Chenopodium and related species previously reported [71,72]. Several studies proposed that the elevated GC content observed in the IR region possibly stems from the higher prevalence of GC-rich genes, such as rRNA and tRNA, within these genomic sections [73,74,75].
The cp genomes of Chenopodium and Atriplex encode 131 genes (totaling 113 unique genes), including 85 protein-coding genes (PCGs), 1 pseudogene (a shorter incomplete copy of ycf1), 37 transfer RNA (tRNA) genes, and 8 ribosomal RNA (rRNA) genes. The order and orientation of these genes were the same in all species, which is in accordance with the outcomes documented in the related genera, such as Beta, Dysphania, and Oxybasis (Table 1). Among the PCGs, 48 genes were associated with photosynthesis, while 75 genes were involved in transcription and translation processes (Table 2). In total, 17 genes contained one intron (rps16, atpF, rpoC1, petB, petD, rpl16, ndhB (×2), ndhA, trnK-UUU, trnG-UCC, trnL-UAA, trnI-GAU (×2), trnA-UGC (×2), and trnV-UAC) and 3 genes contained two introns (ycf3, clpP, and rps12). Additionally, rps12 underwent trans-splicing, involving the concatenation of two exons. The 5′ exon is situated within the LSC region, while the 3′ exon is positioned in the IR region. A total of 17 genes, including 6 PCGs (rpl2, rpl23, ycf2, ndhB, rps7, and rps12), 7 tRNA genes (trnI, trnL, trnV, trnI, trnA, trnR, and trnN), and all 4 rRNA genes (rrn16, rrn23, rrn4.5, and rrn5), were found to be duplicated in the inverted repeat (IR) region.

3.2. Repeat Sequence Analysis

SSRs with 1–6 bp repeat nucleotide units are widely distributed in plant cp genomes [76]. The number of SSRs in the 16 species of Chenopodium and related genera ranged from 44 (C. quinoa and C. petiolare) to 74 (A. centralasiatica). Except for pentanucleotide SSRs, which were not found in A. prostrata, A. patens, and A. gmelinii, all species contain mononucleotide, dinucleotide, trinucleotide, tetranucleotide, and pentanucleotide SSRs. However, hexanucleotide SSRs only exist in A. centralasiatica. Among all SSRs, A/T and AT/AT were the most abundant, accounting for 65.30% and 7.53% respectively. Research has demonstrated that a high AT content in the cp genome of higher plants is a prevalent phenomenon [77,78,79]. SSRs were predominantly located within the intergenic spacer (IGS) of the LSC and SSC regions, with a smaller fraction being distributed within the coding regions of ycf1, ycf2, and rrn23 in the IR region (Figure 2).
Within the cp genomes of 16 species, the count of interspersed repeats varied between 36 (D.ambrosioides) and 87 (O.glauca). Forward and palindrome repeats were identified across all species, whereas reverse and complementary repeats were exclusively observed in C. quinoa and D. ambrosioides, respectively. The majority of the interspersed repeats were forward repeats (n = 14–41) and palindrome repeats (n = 21–46), accounting for 47.62% and 52.15%, respectively, and the size was mainly concentrated in 30–39 bp (Figure 3).

3.3. Analysis of Codon Usage Bias

The gene coding sequences in the 16 closely related Chenopodium species consist of 64 codons, with 61 encoding 20 amino acids, while the remaining 3 (UAA, UAG, and UGA) serve as stop codons. Leucine exhibits the highest frequency, while cysteine demonstrates the lowest frequency among all amino acids. The RSCU indicates the codon usage frequency relative to the expected frequency [80]. In this study, 30 codons exhibited RSCU values exceeding 1, indicating values higher than expected, whereas 32 codons showed RSCU values below 1, signifying values lower than expected. Furthermore, the RSCU values of the remaining two codons, AUG and UGG, equaled 1. In addition, the bases of the high-frequency codons in the third position showed a strong A/T preference (Figure 4).

3.4. Genome Comparison and Nucleotide Diversity

In the field of plant evolution research, the dynamic alteration of the IR region within the cp genome stands as a pivotal event. It holds the potential to induce variations in both the size and genetic composition of the cp genome [81]. The comparative analysis of seven species of Chenopodium and nine species of related genera showed that, except for the obvious contraction of the IR region of Beta, the cp genome structure and gene sequence were substantially conserved within and among the genera, except for a subtle variation in the IR/SC region junctions. The LSC/IRb (JLB) junctions of all 16 species are located within rps19, but the part of rps19 located in the IRb region is less in Chenopodium (79 bp), while all other closely related species, except for D. botrys (79 bp), are more (142–156 bp). The IRb/SSC (JSB) junctions are located within ndhF and range from 2198 to 2229 bp, except for D. ambrosioides (811 bp from ndhF). The SSC/IRa (JSA) junction is located at 1443–1481 bp inside ycf1. The IRa/LSC (JLA) junction is located between rpl2 and trnH and is 0–4 bp away from trnH (Figure 5).
Taking A. centralasiatica as a reference, the cp genomes of the 16 species of Chenopodium and related genera were compared and analyzed (Figure 6). The cp genome is extremely conserved within the same plant genus, yet intergeneric differences persist. On the whole, the variation level of IGS and intron regions is higher than that of exon regions. A nucleotide diversity (Pi) analysis was further performed on 42 cp genomes of the above 16 species (Figure 7). The results show that the IGS region (average Pi = 0.04430) had higher polymorphism than the gene region (average Pi = 0.01810). The most variable gene regions included: matK, accD, psaI, rps11, rps8, rpl22, ycf1, ndhF, ccsA, and ndhD (PI > 0.035), six of which were situated in the LSC region, while four were positioned in the SSC region. The IGS regions with a large variation include: trnH-GUG-psbA, psbI-trnS-GCU, trnS-GCU-trnG-UCC, psbM-trnD-GUC, rps4-trnT-UGU, ycf4-cemA, rpl36-infA, rpl22-rps19, ndhF-rpl32, rpl32-trnL-UAG, ccsA-ndhD, and ndhG-ndhI (Pi > 0.080). Among them, eight were situated in the LSC region and four in the SSC region.

3.5. Phylogenomic Analysis and Divergence Time Estimation

The cp genome plays a pivotal role in revealing the evolutionary relationships among plants [82]. Taking several species of Limonium as the outgroup, we reconstructed the phylogenetic trees of 63 species of Amaranthaceae, including Chenopodium (Table S2). The maximum likelihood (ML) trees and Bayesian inference (BI) trees had almost the same topological structure, with a strong bootstrap support (BS) and posterior probabilities (Figure 8). In the ingroup, all species were classified into three main clades (clade I, clade II, and clade III). Based on morphological studies (Flora of China), all species in clade I belong to Subfam. Amaranthaceae, while all species in clade II and clade III belong to Subfam. Chenopodiaceae. Among them, all the species in clade II are distributed in three tribes (Trib. Atripliceae, Trib. Beteae, and Trib. Chenopodieae), while the species in clade III are distributed in four other tribes (Trib. Camphorosmeae, Trib. Salicornieae, Trib. Salsoleae, and Trib. Suaedeae). In addition, all the species of Chenopodium and its four related genera (Atriplex, Beta, Dysphania, and Oxybasis) in clade II were clustered together as genera, forming five monophyletic groups. This result further supports the inference of other studies that Chenopodium is a monophyletic group [7].
Based on the ML tree, we estimated the divergence time of the 16 species in clade I (Figure 9). The outcomes demonstrate that the 16 species of Chenopodium and related genera had a common ancestor in the Paleocene period at 60.80 Mya (95% highest posterior densities (HPD): 56.27–64.52 Mya). At the same time, Beta also split from other genera. After that, Dysphania and Oxybasis split from other species at 51.47 Mya (95% HPD: 49.91–52.51 Mya) and 47.63 Mya (95% HPD: 46.54–49.17 Mya), respectively. Atriplex was the closest relative to Chenopodium in all related genera; they were sister clades and split at 30.28 Mya (95% HPD: 28.22–33.62 Mya). The seven species of Chenopodium began to split in the Miocene period at about 9.13 Mya (95% HPD: 4.61–14.89 Mya) and separated into two branches represented by C. quinoa and C. album. C. album and C. bryoniifolium are two species that were extremely difficult to distinguish in morphological taxonomy. Their split occurred 20,000 years ago (95% HPD: 0–0.06 Mya) and is also the most recent.

3.6. Development and Validation of the InDel and dCAPS Markers

Based on the interspecific analysis, we designed a series of primers on the conserved region near the variant sites to amplify fragments with genetic marker information (Table S3). The results show that 12 pairs of primers could successfully amplify the expected polymorphic sequences and could completely distinguish 10 species of Chenopodium and related genera (Figure 10). Most of the primers were designed based on InDel markers. Among them, primers 1 and 2 can be used to distinguish Atriplex from other genus species, and primers 3 and 4 can further distinguish A. patens and A. prostrata. Primer 5 and primer 6 can be used to distinguish Beta species. Primer 7 can distinguish Chenopodium species, except for distinguishing C. quinoa from other species. Primer 8 can be used to distinguish C. acuminatum from other Chenopodium species, and primer 8 and primer 9 can further distinguish C. berlandieri. In addition to directly distinguishing D. ambrosioides, primers 10 and 11 can also distinguish C. ficifolium from other Chenopodium species. At the same time, the combination of primers 7 can further distinguish C. quinoa. Primer 12 is the only pair of primers based on dCAPs markers, which are more cumbersome than InDel markers. Because there are only two SNPs in the cp genome of C. album and C. bryoniifolium, we assembled and annotated their ITS sequences, but there was no difference after comparison.

4. Discussion

After sequencing and assembling, we obtained the complete cp genomes of six closely related Chenopodium species, five of which are published in this paper for the first time, which greatly enriches the genome database of Chenopodium species and facilitates their classification, evolutionary relationship inference, and species identification.
Similar to previous studies, their lengths ranged from 151,850 to 152,215 bp and exhibited typical quadripartite structures [83,84]. The cp genomes of most terrestrial plants exhibit a nearly collinear sequence and genetic information, but there are also other occurrences, such as sequence inversion, gene loss, and IR region boundary contraction or expansion [85,86,87]. In the comparison of the cp genome of Chenopodium with other related genera, we found that there was an inversion of almost 6 kb in the LSC region of Beta, Dysphania, and Oxybasis (Figure 11). Studies have shown that this phenomenon is caused by tRNA activity or a high G + C content [88,89]. There was a trnV gene on the flanking of the region where the inversion occurs, but no significant change in the GC content was found.
The primary drivers behind cp genome size alterations are the contraction or expansion of the IR boundary [90,91,92]. The IR boundaries of Chenopodium and its related species are relatively conserved on most species, except for the obvious contraction on Beta, which also leads to the cp genome of B. vulgaris, which is about 2000 bp smaller than that of the other related species.
Repeat sequences exert a pivotal influence on the cp genome rearrangement and stabilization and population genetic diversity [93,94]. In this study, interspersed repeats and SSRs were widely distributed in all related species, and the proportion of repeat sequences in the IGS region was significantly greater than that in the CDS region, which also confirms the previous research results.
Previous studies on the phylogeny of Chenopodium based on cpDNA fragments (trnL-trnF, matK-trnK, atpB, atpB-rbcL, and rbcL) and ITS in rDNA presented some taxa conclusions that are different to our results, and the relationship was relatively vague [2,95,96]. Consequently, we reconstructed the phylogenetic tree utilizing the complete cp genome sequence. The results show that Chenopodium, Atriplex, Beta, Dysphania, and Oxybasis were clustered into a clade, and each of them became a monophyletic group. This also confirmed that the genetic distance between Chenopodium and Beta was closer than that of Suaeda and Haloxylon. In addition, the traditional plant classification method based on morphological characteristics included Chenopodium in the same tribe as Dysphania and Oxybasis, while Atriplex was classified as Trib. Atripliceae, which is obviously contrary to the phylogenetic outcomes derived from the complete cp genome, because Chenopodium and Atriplex have a closer genetic distance.
We used Limonium franchetii, which is also Caryophyllales, as the outgroup to estimate the evolutionary time of Chenopodium and its related species. The findings indicate that the divergence time between Amaranthaceae and the other families of Caryophyllaceae was 90.81 Mya, and the divergence time between Chenopodium and other genera was 30.28 Mya, which was different from the results of previous studies.
Previously, rbcL, trnH-psbA, and matK were regarded as critical barcodes for plant identification, but their resolution at the species level was inadequate [97]. The complete cp genome has more variable regions and has a higher resolution when used as a DNA barcode. Dependent on the complete cp genome, we developed InDel and dCAPS that can completely distinguish the 10 related species of Chenopodium. At the same time, a series of primers were designed based on these markers and verified by PCR experiments. These primers have a good repeatability and can be efficiently used for species identification and germplasm evaluation among the related species of Chenopodium. The ITS sequences of Chenopodium species were also assembled and we attempted to utilize these sequences for molecular marker development, but unfortunately, there are less variable sequences than the complete cp genome. For example, the ITS sequences of C. album and C. bryoniifolium are exactly the same and cannot be used for species discrimination.
In conclusion, this study introduced a significantly more efficient means for the identification of species within the Chenopodium genus and their closely related species.

5. Conclusions

In this investigation, the complete cp genomes of six species were assembled and annotated, and compared with the published sequences of related species. The outcomes reveal that the cp genomes of Chenopodium and its related species were highly conserved in structure and gene content, but there were also subtle variances. We identified a series of SSRs and highly variable regions from all Chenopodium and related species, and found that most of them were distributed in IGS regions and introns. The phylogenetic analysis showed that Chenopodium had the closest relationship with Atriplex, Beta, Dysphania, and Oxybasis and belonged to a monophyletic group compared with other Amaranthaceae species. They had a common ancestor 60.80 Mya, and since then, it gradually divided into different genera. Based on the interspecific variation sites, we designed 12 pairs of primers for species identification, and the experiments confirmed that they could completely distinguish 10 related species.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/genes14122183/s1. Table S1: Comparison of the nucleotide diversity (Pi) of the cp genomes of the 16 Chenopodium and related species. Table S2: The GenBank information of the chloroplast genomes used to infer the amaranthaceae phylogeny. Table S3: Information on the primers used to identify the 10 closely related species of Chenopodium.

Author Contributions

Conceptualization, Z.W. and S.C.; validation, F.C., H.D. and W.L.; data curation, Z.W. and B.Y.; writing—original draft preparation, Z.W. and J.G.; writing—review and editing, S.C.; funding acquisition, S.C. and S.G. 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 (31370296); The “Bohai Sea Granary” Science and Technology Demonstration Project of Shandong Province (2019BHLC001); and the Yantai City School and Local Integration Development Project (SM21JY05).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All WGS data employed in this research were stored at the National Center for Biotechnology Information (NCBI), BioProject: PRJNA1027242, and the assembled cp genomes were stored under Genbank accession: OR374019-OR374024.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Haston, E. An update of the Angiosperm Phylogeny Group classification for the orders and families of flowering plants: APG IV. Bot. J. Linn. Soc. 2019, 181, 1–20. [Google Scholar]
  2. Fuentes-Bazan, S.; Mansion, G.; Borsch, T. Towards a species level tree of the globally diverse genus Chenopodium (Chenopodiaceae). Mol. Phylogenetics Evol. 2012, 62, 359–374. [Google Scholar] [CrossRef] [PubMed]
  3. Lin, P.-H.; Chao, Y.-Y. Different Drought-Tolerant Mechanisms in Quinoa (Chenopodium quinoa Willd.) and Djulis (Chenopodium formosanum Koidz.) Based on Physiological Analysis. Plants 2021, 10, 2279. [Google Scholar] [CrossRef] [PubMed]
  4. Miranda-Apodaca, J.; Agirresarobe, A.; Martínez-Goñi, X.S.; Yoldi-Achalandabaso, A.; Pérez-López, U. N metabolism performance in Chenopodium quinoa subjected to drought or salt stress conditions. Plant Physiol. Biochem. 2020, 155, 725–734. [Google Scholar] [CrossRef] [PubMed]
  5. Zhang, Y.; Mutailifu, A.; Lan, H. Structure, development, and the salt response of salt bladders in Chenopodium album L. Front. Plant Sci. 2022, 13, 989946. [Google Scholar] [CrossRef]
  6. Ding, S.; Zhang, H.; Zhou, C.; Bao, Y.; Xu, X.; Chen, Y.; Shen, Z.; Chen, C. Transcriptomic, epigenomic and physiological comparisons reveal key factors for different manganese tolerances in three Chenopodium ambrosioides L. populations. Plant Physiol. Biochem. 2023, 201, 107883. [Google Scholar] [CrossRef]
  7. Hong, S.-Y.; Cheon, K.-S.; Yoo, K.-O.; Lee, H.-O.; Cho, K.-S.; Suh, J.-T.; Kim, S.-J.; Nam, J.-H.; Sohn, H.-B.; Kim, Y.-H. Complete Chloroplast Genome Sequences and Comparative Analysis of Chenopodium quinoa and C. album. Front. Plant Sci. 2017, 8, 1696. [Google Scholar] [CrossRef] [PubMed]
  8. Bhargava, A.; Shukla, S.; Ohri, D. Evaluation of foliage yield and leaf quality traits in Chenopodium spp. in multiyear trials. Euphytica 2006, 153, 199–213. [Google Scholar] [CrossRef]
  9. Pathan, S.; Siddiqui, R.A. Nutritional Composition and Bioactive Components in Quinoa (Chenopodium quinoa Willd.) Greens: A Review. Nutrients 2022, 14, 558. [Google Scholar] [CrossRef]
  10. Gómez-Caravaca, A.M.; Iafelice, G.; Verardo, V.; Marconi, E.; Caboni, M.F. Influence of pearling process on phenolic and saponin content in quinoa (Chenopodium quinoa Willd). Food Chem. 2014, 157, 174–178. [Google Scholar] [CrossRef]
  11. Tang, Y.; Li, X.; Zhang, B.; Chen, P.X.; Liu, R.; Tsao, R. Characterisation of phenolics, betanins and antioxidant activities in seeds of three Chenopodium quinoa Willd. genotypes. Food Chem. 2015, 166, 380–388. [Google Scholar] [CrossRef] [PubMed]
  12. Nowak, V.; Du, J.; Charrondière, U.R. Assessment of the nutritional composition of quinoa (Chenopodium quinoa Willd.). Food Chem. 2016, 193, 47–54. [Google Scholar] [CrossRef]
  13. Vilcacundo, R.; Hernández-Ledesma, B. Nutritional and biological value of quinoa (Chenopodium quinoa Willd.). Curr. Opin. Food Sci. 2017, 14, 1–6. [Google Scholar] [CrossRef]
  14. Vysochina, G.I. Flavonoids of the Chenopodium, L. genus of world flora. Russ. J. Bioorganic Chem. 2010, 36, 787–792. [Google Scholar] [CrossRef]
  15. Tapia, Y.; Diaz, O.; Pizarro, C.; Segura, R.; Vines, M.; Zúñiga, G.; Moreno-Jiménez, E. Atriplex atacamensis and Atriplex halimus resist As contamination in Pre-Andean soils (northern Chile). Sci. Total Environ. 2013, 450–451, 188–196. [Google Scholar] [CrossRef]
  16. Gođevac, D.; Stanković, J.; Novaković, M.; Anđelković, B.; Dajić-Stevanović, Z.; Petrović, M.; Stanković, M. Phenolic Compounds from Atriplex littoralis and Their Radiation-Mitigating Activity. J. Nat. Prod. 2015, 78, 2198–2204. [Google Scholar] [CrossRef]
  17. Kandsi, F.; Elbouzidi, A.; Lafdil, F.Z.; Meskali, N.; Azghar, A.; Addi, M.; Hano, C.; Maleb, A.; Gseyra, N. Antibacterial and Antioxidant Activity of Dysphania ambrosioides (L.) Mosyakin and Clemants Essential Oils: Experimental and Computational Approaches. Antibiotics 2022, 11, 482. [Google Scholar] [CrossRef]
  18. Pavela, R.; Maggi, F.; Lupidi, G.; Mbuntcha, H.; Woguem, V.; Womeni, H.M.; Barboni, L.; Tapondjou, L.A.; Benelli, G. Clausena anisata and Dysphania ambrosioides essential oils: From ethno-medicine to modern uses as effective insecticides. Environ. Sci. Pollut. Res. 2017, 25, 10493–10503. [Google Scholar] [CrossRef]
  19. Almadiy, A.A. Chemical profile, mosquitocidal, and biochemical effects of essential oil and major components of Dysphania ambrosioides against Culex quinquefasciatus Say. Environ. Sci. Pollut. Res. 2020, 27, 41568–41576. [Google Scholar] [CrossRef]
  20. Kurashige, N.S. Phenotypic Plasticity to Light Competition and Herbivory in Chenopodium Album; University of Toronto: Toronto, ON, Canada, 2004. [Google Scholar]
  21. La Duke, J.; Crawford, D.J. Character Compatibility and Phyletic Relationships in Several Closely Related Species of Chenopodium of the Western United States. Taxon 2019, 28, 307–314. [Google Scholar] [CrossRef]
  22. Wilson, H.D. Artificial Hybridization Among Species of Chenopodium sect. Chenopodium. Syst. Bot. 1980, 5, 253. [Google Scholar] [CrossRef]
  23. Zhou, D.; Wang, T.; Valentine, I. Phenotypic plasticity of life-history characters in response to different germination timing in two annual weeds. Can. J. Bot. 2005, 83, 28–36. [Google Scholar] [CrossRef]
  24. Rana, T.S.; Narzary, D.; Ohri, D. Molecular differentiation of Chenopodium album complex and some related species using ISSR profiles and ITS sequences. Gene 2012, 495, 29–35. [Google Scholar] [CrossRef] [PubMed]
  25. Chrungoo, N.K.; Jashmi Devi, R.; Goel, S.; Das, K. Deciphering species relationships and evolution in Chenopodium through sequence variations in nuclear internal transcribed spacer region and amplified fragment-length polymorphism in nuclear DNA. J. Genet. 2019, 98, 37. [Google Scholar] [CrossRef]
  26. Kolano, B.; Siwinska, D.; McCann, J.; Weiss-Schneeweiss, H. The evolution of genome size and rDNA in diploid species of Chenopodium s.l.(Amaranthaceae). Bot. J. Linn. Soc. 2015, 179, 218–235. [Google Scholar] [CrossRef]
  27. Mandák, B.; Krak, K.; Vít, P.; Lomonosova, M.N.; Belyayev, A.; Habibi, F.; Wang, L.; Douda, J.; Štorchová, H. Hybridization and polyploidization within the Chenopodium album aggregate analysed by means of cytological and molecular markers. Mol. Phylogenet. Evol. 2018, 129, 189–201. [Google Scholar] [CrossRef] [PubMed]
  28. Devi, R.J.; Chrungoo, N.K. Evolutionary divergence in Chenopodium and validation of SNPs in chloroplast rbcL and matk genes by allele-specific PCR for development of Chenopodium quinoa-specific markers. Crop J. 2017, 5, 32–42. [Google Scholar] [CrossRef]
  29. Liu, Z.-J.; Xu, S.; Li, D.; Li, J.; Xiang, X.; Jin, W.; Huang, W.; Jin, X.; Huang, L. Evaluation of the DNA Barcodes in Dendrobium (Orchidaceae) from Mainland Asia. PLoS ONE 2015, 10, e0115168. [Google Scholar] [CrossRef]
  30. Kane, N.C.; Cronk, Q. Botany without borders: Barcoding in focus. Mol. Ecol. 2008, 17, 5175–5176. [Google Scholar] [CrossRef]
  31. Chen, J.; Wang, F.; Zhao, Z.; Li, M.; Liu, Z.; Peng, D. Complete Chloroplast Genomes and Comparative Analyses of Three Paraphalaenopsis (Aeridinae, Orchidaceae) Species. Int. J. Mol. Sci. 2023, 24, 11167. [Google Scholar] [CrossRef]
  32. Fang, H.; Dai, G.; Liao, B.; Zhou, P.; Liu, Y. Application of chloroplast genome in the identification of Phyllanthus urinaria and its common adulterants. Front. Plant Sci. 2023, 13, 1099856. [Google Scholar] [CrossRef] [PubMed]
  33. Chong, X.; Li, Y.; Yan, M.; Wang, Y.; Li, M.; Zhou, Y.; Chen, H.; Lu, X.; Zhang, F. Comparative chloroplast genome analysis of 10 Ilex species and the development of species-specific identification markers. Ind. Crops Prod. 2022, 187, 115408. [Google Scholar] [CrossRef]
  34. Cai, Z.; Guisinger, M.; Kim, H.-G.; Ruck, E.; Blazier, J.C.; McMurtry, V.; Kuehl, J.V.; Boore, J.; Jansen, R.K. Extensive Reorganization of the Plastid Genome of Trifolium subterraneum (Fabaceae) Is Associated with Numerous Repeated Sequences and Novel DNA Insertions. J. Mol. Evol. 2008, 67, 696–704. [Google Scholar] [CrossRef] [PubMed]
  35. Lee, H.L.; Jansen, R.K.; Chumley, T.W.; Kim, K.J. Gene Relocations within Chloroplast Genomes of Jasminum and Menodora (Oleaceae) Are Due to Multiple, Overlapping Inversions. Mol. Biol. Evol. 2007, 24, 1161–1180. [Google Scholar] [CrossRef] [PubMed]
  36. Guisinger, M.M.; Chumley, T.W.; Kuehl, J.V.; Boore, J.L.; Jansen, R.K. Implications of the Plastid Genome Sequence of Typha (Typhaceae, Poales) for Understanding Genome Evolution in Poaceae. J. Mol. Evol. 2010, 70, 149–166. [Google Scholar] [CrossRef] [PubMed]
  37. Martin, G.E.; Rousseau-Gueutin, M.; Cordonnier, S.; Lima, O.; Michon-Coudouel, S.; Naquin, D.; de Carvalho, J.F.; Aïnouche, M.; Salmon, A.; Aïnouche, A. The first complete chloroplast genome of the Genistoid legume Lupinus luteus: Evidence for a novel major lineage-specific rearrangement and new insights regarding plastome evolution in the legume family. Ann. Bot. 2014, 113, 1197–1210. [Google Scholar] [CrossRef]
  38. Quan, Z.; Yang, L.; Plant, S.J.; Physiology, C. Examination of the cytoplasmic DNA in male reproductive cells to determine the potential for cytoplasmic inheritance in 295 angiosperm species. Plant Cell Physiol. 2003, 44, 941–951. [Google Scholar]
  39. Corriveau, J.L.; Coleman, A.W. Rapid Screening Method to Detect Potential Biparental Inheritance of Plastid DNA and Results for over 200 Angiosperm Species. Am. J. Bot. 1988, 75, 1443–1458. [Google Scholar] [CrossRef]
  40. Li, L.; Wang, W.; Zhang, G.; Wu, K.; Fang, L.; Li, M.; Liu, Z.; Zeng, S. Comparative analyses and phylogenetic relationships of thirteen Pholidota species (Orchidaceae) inferred from complete chloroplast genomes. BMC Plant Biol. 2023, 23, 269. [Google Scholar] [CrossRef]
  41. Zhang, Y.; Iaffaldano, B.J.; Zhuang, X.; Cardina, J.; Cornish, K. Chloroplast genome resources and molecular markers differentiate rubber dandelion species from weedy relatives. BMC Plant Biol. 2017, 17, 34. [Google Scholar] [CrossRef]
  42. Columbus, J.T.; Wu, Z.; Raven, P.H.; Hong, D.J.T. Flora of China. Taxon 2008, 57, 682. [Google Scholar] [CrossRef]
  43. Doyle, J.J.P.B. A rapid DNA isolation procedure for small quantities of fresh leaf tissue. Phytochem. Bull. 1987, 19, 11–15. [Google Scholar]
  44. Chen, S.; Zhou, Y.; Chen, Y.; Gu, J. fastp: An ultra-fast all-in-one FASTQ preprocessor. Bioinformatics 2018, 34, i884–i890. [Google Scholar] [CrossRef] [PubMed]
  45. Jin, 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]
  46. Wick, R.R.; Schultz, M.B.; Justin, Z.; Holt, K.E.J.B. Bandage: Interactive visualization of de novo genome assemblies. Bioinformatics 2015, 31, 3350–3352. [Google Scholar] [CrossRef]
  47. Krumsiek, J.; Arnold, R.; Rattei, T. Gepard: A rapid and sensitive tool for creating dotplots on genome scale. Bioinformatics 2007, 23, 1026–1028. [Google Scholar] [CrossRef]
  48. Dunn, N.A.J.P.; Genome, A. WebApollo 2.0: A Scalable, Distributed Sequence Annotation Editor for Multiple Genomes. In Proceedings of the Plant & Animal Genome Conference, San Diego, CA, USA, 9–13 January 2015. [Google Scholar]
  49. Shi, L.; Chen, H.; Jiang, M.; Wang, L.; Wu, X.; Huang, L.; Liu, C. CPGAVAS2, an integrated plastome sequence annotator and analyzer. Nucleic Acids Res. 2019, 47, W65–W73. [Google Scholar] [CrossRef]
  50. Michael, T.; Pascal, L.; Tommaso, P.; Ulbricht-Jones, E.S.; Axel, F.; Ralph, B.; Stephan, G. GeSeq—Versatile and accurate annotation of organelle genomes. Nucleic Acids Res. 2017, 45, W6–W11. [Google Scholar]
  51. Chan, P.P.; Lin, B.Y.; Mak, A.J.; Lowe, T.M. tRNAscan-SE 2.0: Improved detection and functional classification of transfer RNA genes. Nucleic Acids Res. 2021, 49, 9077–9096. [Google Scholar] [CrossRef]
  52. Liu, S.; Ni, Y.; Li, J.; Zhang, X.; Yang, H.; Chen, H.; Liu, C. CPGView: A package for visualizing detailed chloroplast genome structures. Mol. Ecol. Resour. 2023, 23, 694–704. [Google Scholar] [CrossRef]
  53. Zheng, S.; Poczai, P.; Hyvönen, 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]
  54. 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]
  55. Stefan, K.; Choudhuri, J.V.; Enno, O.; Chris, S.; Jens, S.; Robert, G.J. REPuter: The manifold applications of repeat analysis on a genomic scale. Nucleic Acids Res. 2001, 29, 4633–4642. [Google Scholar]
  56. Sharp, P.M.; Tuohy, T.M.F.; Mosurski, K.R.J.N.A.R. Codon usage in yeast: Cluster analysis clearly differentiates highly and lowly expressed genes. Nucleic Acids Res. 1986, 14, 5125–5143. [Google Scholar] [CrossRef] [PubMed]
  57. Chen, C.; Chen, H.; Zhang, Y.; Thomas, H.R.; Xia, R.J.M.P. TBtools: An Integrative Toolkit Developed for Interactive Analyses of Big Biological Data. Mol. Plant 2020, 13, 1194–1202. [Google Scholar] [CrossRef]
  58. Li, H.; Guo, Q.; Xu, L.; Gao, H.; Liu, L.; Zhou, X. CPJSdraw: Analysis and visualization of junction sites of chloroplast genomes. PeerJ 2023, 11, e15326. [Google Scholar] [CrossRef]
  59. Brudno, M.; Do, C.B.; Cooper, G.M.; Kim, M.F.; Davydov, E.; NISC Comparative Sequencing Program; Green, E.D.; Sidow, A.; Batzoglou, S. LAGAN and Multi-LAGAN: Efficient Tools for Large-Scale Multiple Alignment of Genomic DNA. Genome Res. 2003, 13, 721. [Google Scholar] [CrossRef]
  60. Frazer, K.A.; Lior, P.; Alexander, P.; Rubin, E.M.; Inna, D. VISTA: Computational tools for comparative genomics. Nucleic Acids Res. 2004, 32, W273. [Google Scholar] [CrossRef]
  61. Rozas, J.; SánchezDelBarrio, J.C.; Messeguer, X.; Rozas, R.J.B. DnaSP, DNA polymorphism analyses by the coalescent and other methods. Bioinformatics 2003, 19, 2496–2497. [Google Scholar] [CrossRef] [PubMed]
  62. 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]
  63. Capella-Gutierrez, S.; Silla-Martinez, J.M.; Gabaldon, T.J.B. trimAl: A tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics 2009, 25, 1972–1973. [Google Scholar] [CrossRef] [PubMed]
  64. Kalyaanamoorthy, S.; Minh, B.Q.; Wong, T.K.F.; von Haeseler, A.; Jermiin, L.S. ModelFinder: Fast model selection for accurate phylogenetic estimates. Nat. Methods 2017, 14, 587–589. [Google Scholar] [CrossRef] [PubMed]
  65. Lam-Tung, N.; Schmidt, H.A.; Arndt, V.H.; Quang, M.B.J.M.B. Evolution. IQ-TREE: A Fast and Effective Stochastic Algorithm for Estimating Maximum-Likelihood Phylogenies. Mol. Biol. Evol. 2015, 32, 268–274. [Google Scholar]
  66. Huelsenbeck, J.P.J.S.B. MrBayes 3.2: Efficient Bayesian Phylogenetic Inference and Model Choice Across a Large Model Space. Syst. Biol. 2012, 61, 539–542. [Google Scholar]
  67. Puttick, M.N.J.B. MCMCtreeR: Functions to prepare MCMCtree analyses and visualize posterior ages on trees. Bioinformatics 2019, 35, 5321–5322. [Google Scholar] [CrossRef]
  68. Sudhir, K.; Michael, S.; Craig, J.M.; Kasprowicz, A.E.; Maxwell, S.; Michael, L.; Glen, S.; Blair, H.S.J.M.B. Evolution. TimeTree 5: An Expanded Resource for Species Divergence Times. Mol. Biol. Evol. 2022, 39, msac174. [Google Scholar]
  69. Untergasser, A.; Cutcutache, I.; Koressaar, T.; Ye, J.; Faircloth, B.C.; Remm, M.; Rozen, S.G. Primer3—New capabilities and interfaces. Nucleic Acids Res. 2012, 40, e115. [Google Scholar] [CrossRef]
  70. Kalishman, T.M.J.T.i.G. Web-based primer design for single nucleotide polymorphism analysis. Trends Genet. 2002, 18, 613–615. [Google Scholar] [CrossRef]
  71. Park, I.; Yang, S.; Kim, W.J.; Noh, P.; Lee, H.O.; Moon, B.C. The Complete Chloroplast Genomes of Six Ipomoea Species and Indel Marker Development for the Discrimination of Authentic Pharbitidis Semen (Seeds of I. nil or I. purpurea). Front. Plant Sci. 2018, 9, 965. [Google Scholar] [CrossRef]
  72. Li, B.; Zheng, Y. Dynamic evolution and phylogenomic analysis of the chloroplast genome in Schisandraceae. Sci. Rep. 2018, 8, 9285. [Google Scholar] [CrossRef]
  73. Hu, Y.; Chen, X.; Zhou, T.; Hou, N.; Zhao, P. Characterization of the complete chloroplast genome of an endangered species dwarf birch (Betula nana L.). Conserv. Genet. Resour. 2016, 9, 55–58. [Google Scholar] [CrossRef]
  74. Li, Y.; Yang, Y.; Yu, L.; Du, X.; Ren, G. Plastomes of nine hornbeams and phylogenetic implications. Ecol. Evol. 2018, 8, 8770–8778. [Google Scholar] [CrossRef]
  75. Wang, S.; Yang, C.; Zhao, X.; Chen, S.; Qu, G.-Z. Complete chloroplast genome sequence of Betula platyphylla: Gene organization, RNA editing, and comparative and phylogenetic analyses. BMC Genom. 2018, 19, 950. [Google Scholar] [CrossRef]
  76. Powell, W.; Morgante, M.; McDevitt, R.; Vendramin, G.G.; Rafalski, J.A. Polymorphic simple sequence repeat regions in chloroplast genomes: Applications to the population genetics of pines. Proc. Natl. Acad. Sci. USA 1995, 92, 7759–7763. [Google Scholar] [CrossRef] [PubMed]
  77. Lei, W.; Ni, D.; Wang, Y.; Shao, J.; Liu, C.J.S.R. Intraspecific and heteroplasmic variations, gene losses and inversions in the chloroplast genome of Astragalus membranaceus. Sci. Rep. 2016, 6, 21669. [Google Scholar] [CrossRef] [PubMed]
  78. Tao, Z.; Jian, W.; Yun, J.; Wenli, L.; Fusheng, X.; Xumei, W. Comparative Chloroplast Genome Analyses of Species in Gentiana section Cruciata (Gentianaceae) and the Development of Authentication Markers. Int. J. Mol. Sci. 2018, 19, 1962. [Google Scholar]
  79. Yang, Z.; Huang, Y.; An, W.; Zheng, X.; Huang, S.; Liang, L. Sequencing and Structural Analysis of the Complete Chloroplast Genome of the Medicinal Plant Lyciumchinense Mill. Plants 2019, 8, 87. [Google Scholar] [CrossRef]
  80. Nie, L.; Cui, Y.; Wu, L.; Zhou, J.; Xu, Z.; Li, Y.; Li, X.; Wang, Y.; Yao, H. Gene Losses and Variations in Chloroplast Genome of Parasitic Plant Macrosolen and Phylogenetic Relationships within Santalales. Int. J. Mol. Sci. 2019, 20, 5812. [Google Scholar] [CrossRef] [PubMed]
  81. Long, L.; Li, Y.; Wang, S.; Liu, Z.; Wang, J.; Yang, M. Complete chloroplast genomes and comparative analysis of Ligustrum species. Sci. Rep. 2023, 13, 212. [Google Scholar] [CrossRef]
  82. Wan, T.; Qiao, B.-x.; Zhou, J.; Shao, K.-s.; Pan, L.-y.; An, F.; He, X.-s.; Liu, T.; Li, P.-k.; Cai, Y.-L. Evolutionary and phylogenetic analyses of 11 Cerasus species based on the complete chloroplast genome. Front. Plant Sci. 2023, 14, 1070600. [Google Scholar] [CrossRef]
  83. Yang, L.; Li, Q.; Zhao, G. Characterization of the complete chloroplast genome of Chenopodium sp. (Caryophyllales: Chenopodiaceae). Mitochondrial DNA B Resour. 2019, 4, 2574–2575. [Google Scholar] [CrossRef]
  84. Kim, Y.; Chung, Y.; Park, J. The complete chloroplast genome of Chenopodium ficifolium Sm. (Amaranthaceae). Mitochondrial DNA B Resour. 2019, 4, 872–873. [Google Scholar] [CrossRef]
  85. Cho, K.S.; Yun, B.K.; Yoon, Y.H.; Hong, S.Y.; Mekapogu, M.; Kim, K.H.; Yang, T.J. Complete Chloroplast Genome Sequence of Tartary Buckwheat (Fagopyrum tataricum) and Comparative Analysis with Common Buckwheat (F. esculentum). PLoS ONE 2015, 10, e0125332. [Google Scholar] [CrossRef] [PubMed]
  86. Fu, P.C.; Zhang, Y.Z.; Geng, H.M.; Chen, S.L. The complete chloroplast genome sequence of Gentiana lawrencei var. farreri (Gentianaceae) and comparative analysis with its congeneric species. PeerJ 2016, 4, e2540. [Google Scholar] [CrossRef] [PubMed]
  87. Choi, K.S.; Chung, M.G.; Park, S. The Complete Chloroplast Genome Sequences of Three Veroniceae Species (Plantaginaceae): Comparative Analysis and Highly Divergent Regions. Front Plant Sci. 2016, 7, 355. [Google Scholar] [CrossRef] [PubMed]
  88. Walker, J.F.; Zanis, M.J.; Emery, N.C. Comparative analysis of complete chloroplast genome sequence and inversion variation in Lasthenia burkei (Madieae, Asteraceae). Am. J. Bot. 2014, 101, 722–729. [Google Scholar] [CrossRef] [PubMed]
  89. Fullerton, S.M.; Antonio, B.C.; Clark, A.G. Local Rates of Recombination Are Positively Correlated with GC Content in the Human Genome. Mol. Biol. Evol. 2001, 18, 1139–1142. [Google Scholar] [CrossRef] [PubMed]
  90. Cho, K.S.; Park, T.H. Complete chloroplast genome sequence of Solanum nigrum and development of markers for the discrimination of S. nigrum. Hortic. Environ. Biotechnol. 2016, 57, 69–78. [Google Scholar] [CrossRef]
  91. 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]
  92. Ni, L.; Zhao, Z.; Xu, H.; Chen, S.; Dorje, G.J.G. The complete chloroplast genome of Gentiana straminea (Gentianaceae), an endemic species to the Sino-Himalayan subregion. Gene 2016, 577, 281–288. [Google Scholar] [CrossRef]
  93. Song, S.L.; Lim, P.E.; Phang, S.M.; Lee, W.W.; Hong, D.D.; Prathep, A. Development of chloroplast simple sequence repeats (cpSSRs) for the intraspecific study of Gracilaria tenuistipitata (Gracilariales, Rhodophyta) from different populations. BMC Res. Notes 2014, 7, 77. [Google Scholar] [CrossRef] [PubMed]
  94. Cavalier-Smith, T.J.C.B. Chloroplast Evolution: Secondary Symbiogenesis and Multiple Losses. Curr. Biol. 2002, 12, R62–R64. [Google Scholar] [CrossRef] [PubMed]
  95. Kadereit, G.; Newton, R.J.; Vandelook, F. Evolutionary ecology of fast seed germination—A case study in Amaranthaceae/Chenopodiaceae. Perspect. Plant Ecol. Evol. Syst. 2017, 29, 1–11. [Google Scholar] [CrossRef]
  96. Kadereit, G.; Mavrodiev, E.V.; Zacharias, E.H.; Sukhorukov, A.P. Molecular phylogeny of Atripliceae (Chenopodioideae, Chenopodiaceae): Implications for systematics, biogeography, flower and fruit evolution, and the origin of C4 photosynthesis. Am. J. Bot. 2010, 97, 1664–1687. [Google Scholar] [CrossRef]
  97. Song, Y.; Wang, S.; Ding, Y.; Xu, J.; Li, M.F.; Zhu, S.; Chen, N. Chloroplast Genomic Resource of Paris for Species Discrimination. Sci. Rep. 2017, 7, 3427. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Complete cp genome map of Atriplex and Chenopodium. Genes with different functions were divided by color. Genes inside and outside the circle were transcribed clockwise and counterclockwise, respectively.
Figure 1. Complete cp genome map of Atriplex and Chenopodium. Genes with different functions were divided by color. Genes inside and outside the circle were transcribed clockwise and counterclockwise, respectively.
Genes 14 02183 g001
Figure 2. Types and numbers of simple sequence repeats in the cp genomes of 16 Chenopodium and related species.
Figure 2. Types and numbers of simple sequence repeats in the cp genomes of 16 Chenopodium and related species.
Genes 14 02183 g002
Figure 3. Interspersed repeats in the cp genomes of 16 Chenopodium and related species. (a) Number of 4 types of repeats. (b) Number of different repetition repeats.
Figure 3. Interspersed repeats in the cp genomes of 16 Chenopodium and related species. (a) Number of 4 types of repeats. (b) Number of different repetition repeats.
Genes 14 02183 g003
Figure 4. RSCU values of the chloroplast genomes of the 16 Chenopodium and related species. Red and green represent higher and lower RSCU values, respectively.
Figure 4. RSCU values of the chloroplast genomes of the 16 Chenopodium and related species. Red and green represent higher and lower RSCU values, respectively.
Genes 14 02183 g004
Figure 5. Comparison of the boundaries of the LSC, SSC, and IR regions among the 16 Chenopodium and related species.
Figure 5. Comparison of the boundaries of the LSC, SSC, and IR regions among the 16 Chenopodium and related species.
Genes 14 02183 g005
Figure 6. Based on the reference cp genome of A.centralasiatica, the cp genomes of the 16 Chenopodium and related species were compared using mVISTA. The y-axis represents the percent identity within 50–100%. Genome regions were color-coded as exon (purple), Trna/rRNA (blue), and intergenic spacer region and intron (pink). Gray arrows display the gene orientation.
Figure 6. Based on the reference cp genome of A.centralasiatica, the cp genomes of the 16 Chenopodium and related species were compared using mVISTA. The y-axis represents the percent identity within 50–100%. Genome regions were color-coded as exon (purple), Trna/rRNA (blue), and intergenic spacer region and intron (pink). Gray arrows display the gene orientation.
Genes 14 02183 g006
Figure 7. Comparison of the nucleotide diversity (Pi) of the cp genomes of the 16 Chenopodium and related species. (a) and (b) represent the gene region and intergenic spacer region (IGS), respectively, with 0.035 and 0.08 as the screening thresholds for the gene region and IGS, respectively (Table S1).
Figure 7. Comparison of the nucleotide diversity (Pi) of the cp genomes of the 16 Chenopodium and related species. (a) and (b) represent the gene region and intergenic spacer region (IGS), respectively, with 0.035 and 0.08 as the screening thresholds for the gene region and IGS, respectively (Table S1).
Genes 14 02183 g007
Figure 8. Maximum likelihood phylogenetic tree based on the complete cp genome. Several species of Limonium were outgrouped.
Figure 8. Maximum likelihood phylogenetic tree based on the complete cp genome. Several species of Limonium were outgrouped.
Genes 14 02183 g008
Figure 9. Estimation of the divergence time of the 16 related species. The gray bar and the number above it represent the 95% highest posterior densities and divergence time, respectively.
Figure 9. Estimation of the divergence time of the 16 related species. The gray bar and the number above it represent the 95% highest posterior densities and divergence time, respectively.
Genes 14 02183 g009
Figure 10. Verification of the primers based on the InDel and dCAPS markers. The lane M represents the 2000/5000 bp marker; A1–A2 are A. patens and A. prostrata, respectively; C1–C6 are C. acuminatum, C. album, C. berlandieri, C. bryoniifolium, C. ficifolium, and C. quinoa, respectively; and B and D are B. vulgaris and D. ambrosioides, respectively. The rectangles composed of solid lines and dotted lines represent the insertion and deletion of InDel, respectively; the circles and triangles represent SSRs and SNPs, respectively; the horizontal solid line and dotted line represent the conserved and variant sequences, respectively; the arrow indicates the direction of the primer; and the blue fork indicates that the primers are not applicable.
Figure 10. Verification of the primers based on the InDel and dCAPS markers. The lane M represents the 2000/5000 bp marker; A1–A2 are A. patens and A. prostrata, respectively; C1–C6 are C. acuminatum, C. album, C. berlandieri, C. bryoniifolium, C. ficifolium, and C. quinoa, respectively; and B and D are B. vulgaris and D. ambrosioides, respectively. The rectangles composed of solid lines and dotted lines represent the insertion and deletion of InDel, respectively; the circles and triangles represent SSRs and SNPs, respectively; the horizontal solid line and dotted line represent the conserved and variant sequences, respectively; the arrow indicates the direction of the primer; and the blue fork indicates that the primers are not applicable.
Genes 14 02183 g010
Figure 11. Dot plot of the inversions in the cp genomes of Beta, Dysphania, and Oxybasis. (a,b) represent the collinearity between Chenopodium and Atriplex, and between Chenopodium and Beta, Dysphania, Oxybasis, respectively. The interior of the red circle is the area where inversions occur.
Figure 11. Dot plot of the inversions in the cp genomes of Beta, Dysphania, and Oxybasis. (a,b) represent the collinearity between Chenopodium and Atriplex, and between Chenopodium and Beta, Dysphania, Oxybasis, respectively. The interior of the red circle is the area where inversions occur.
Genes 14 02183 g011
Table 1. Summary of the cp genomes for the 16 Chenopodium and related species.
Table 1. Summary of the cp genomes for the 16 Chenopodium and related species.
AccessionSpeciesTotalLSCIRSSCTotalLSCIRSSCTotal Protein-Coding rRNAtRNA
NumberLength (bp)(bp)(bp)(bp)GC Content (%)GC Content (%)GC Content (%)GC Content (%)GenesGenesGenesGenes
NC_045304A. centralasiatica152,23783,72125,20918,09537.2735.3842.7230.8513186837
NC_059062A. gmelinii151,85283,71625,16117,81137.2935.3542.8130.8613186837
OR374023A. patens *151,85083,71425,16117,81137.2935.3542.8130.8613186837
OR374024A. prostrata *152,21583,88725,18617,95337.335.442.7730.8213186837
ON641300B. vulgaris149,72483,10724,40917,7963734.9742.8430.4413186837
NC_054154C. acuminatum152,20083,68325,19218,13037.2535.342.7231.0113186837
OR374020C. album *152,16683,67525,19218,10437.2535.342.7230.9913186837
OR374022C. berlandieri *152,19483,67625,19218,13137.2535.342.7231.0213186837
OR374021C. bryoniifolium *152,16683,67525,19218,10437.2535.342.7230.9913186837
NC_041200C. ficifolium151,82383,66825,10817,93637.2635.3142.753113186837
OQ957163C. petiolare152,06483,52025,21718,10737.2435.2842.7630.9213186837
NC_034949C. quinoa152,09983,58225,20418,10637.2435.2942.7430.9613186837
NC_041201D. ambrosioides151,68983,42125,10218,06136.9234.8642.7430.2913186837
NC_042166D. botrys152,05583,76925,18417,91536.8434.7442.7230.1413186837
NC_041159D. pumilio151,96283,75825,23017,74136.9534.8442.7430.4413186837
NC_047226O. glauca151,65583,57925,08017,91336.8734.7342.7730.3513186837
* Represents the newly obtained chloroplast genome data in this study.
Table 2. Gene composition of the cp genome in Chenopodium and its related species.
Table 2. Gene composition of the cp genome in Chenopodium and its related species.
CategoryGroup of GenesGenes
Photosynthesis Photosystem IpsaA, psaB, psaC, psaI, psaJ
Related genesPhotosystem IIpsbA, psbB, psbC, psbD, psbE, psbF, psbH, psbI, psbJ, psbK,
psbL, psbM, psbN, psbT, psbZ
NADH oxidoreductasendhA *, ndhB *#, ndhC, ndhD, ndhE, ndhF, ndhG, ndhH,
ndhI, ndhJ, ndhK
Cytochrome b6/f complexpetA, petB *, petD *, petG, petL, petN
ATP synthaseatpA, atpB, atpE, atpF *, atpH, atpI
RubiscorbcL
Assembly/stabilityycf3 **, ycf4
Cytochrome c synthesisccsA
Transcription andLarge subunit ribosomal proteinsrpl14, rpl16 *, rpl2 #, rpl20, rpl22, rpl23 #, rpl32, rpl33, rpl36
Translation-related genesSmall subunit ribosomal proteinsrps11, rps12 **#, rps14, rps15, rps16 *, rps18, rps19, rps2, rps3,
rps4, rps7 #, rps8
TranscriptionrpoA, rpoB, rpoC1 *, rpoC2
Translation initiation factorinfA
Ribosomal RNAsrrn16 #, rrn23 #, rrn4.5 #, rrn5 #
Transfer RNAstrnA-UGC *#, trnC-GCA, trnD-GUC, trnE-UUC, trnF-GAA,
trnG-GCC, trnG-UCC *, trnH-GUG, trnI-CAU #, trnI-GAU *#,
trnK-UUU *, trnL-CAA #, trnL-UAA *, trnL-UAG, trnM-CAU,
trnN-GUU #, trnP-UGG, trnQ-UUG, trnR-ACG #, trnR-UCU,
trnS-GCU, trnS-GGA, trnS-UGA, trnT-GGU, trnT-UGU,
trnV-GAC #, trnV-UAC *, trnW-CCA, trnY-GUA, trnfM-CAU
Other genesRNA processingmatK
ProteolysisclpP **
Carbon metabolismcemA
Fatty acid synthesisaccD
Proteins with unknown functionsycf1 #, ycf2 #
Gene *: Gene containing one intron; Gene **: Gene containing two introns; Gene #: Two gene copies in the IRs (ycf1 located at the boundary of SSC and IRb is a pseudogene).
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wei, Z.; Chen, F.; Ding, H.; Liu, W.; Yang, B.; Geng, J.; Chen, S.; Guo, S. Comparative Analysis of Six Chloroplast Genomes in Chenopodium and Its Related Genera (Amaranthaceae): New Insights into Phylogenetic Relationships and the Development of Species-Specific Molecular Markers. Genes 2023, 14, 2183. https://doi.org/10.3390/genes14122183

AMA Style

Wei Z, Chen F, Ding H, Liu W, Yang B, Geng J, Chen S, Guo S. Comparative Analysis of Six Chloroplast Genomes in Chenopodium and Its Related Genera (Amaranthaceae): New Insights into Phylogenetic Relationships and the Development of Species-Specific Molecular Markers. Genes. 2023; 14(12):2183. https://doi.org/10.3390/genes14122183

Chicago/Turabian Style

Wei, Zixiang, Fangjun Chen, Hongxia Ding, Wenli Liu, Bo Yang, Jiahui Geng, Shihua Chen, and Shanli Guo. 2023. "Comparative Analysis of Six Chloroplast Genomes in Chenopodium and Its Related Genera (Amaranthaceae): New Insights into Phylogenetic Relationships and the Development of Species-Specific Molecular Markers" Genes 14, no. 12: 2183. https://doi.org/10.3390/genes14122183

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