Next Article in Journal
Comparative Study of the Antioxidant Constituents, Activities and the GC-MS Quantification and Identification of Fatty Acids of Four Selected Helichrysum Species
Next Article in Special Issue
Phylogenomic Analysis of the Plastid Genome of the Peruvian Purple Maize Zea mays subsp. mays cv. ‘INIA 601’
Previous Article in Journal
Exploring New Sources of Bioactive Phenolic Compounds from Western Balkan Mountains
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Plastid Genome of Equisetum xylochaetum from the Atacama Desert, Chile and the Relationships of Equisetum Based on Frequently Used Plastid Genes and Network Analysis

by
Anchittha Satjarak
1,*,
Linda E. Graham
2,
Marie T. Trest
2 and
Patricia Arancibia-Avila
3
1
Plants of Thailand Research Unit, Department of Botany, Faculty of Science, Chulalongkorn University, Bangkok 10330, Thailand
2
Department of Botany, University of Wisconsin, Madison, WI 53706-1313, USA
3
Department of Basic Sciences, University of Bío-Bío, Chillan 3780000, Chile
*
Author to whom correspondence should be addressed.
Plants 2022, 11(7), 1001; https://doi.org/10.3390/plants11071001
Submission received: 13 March 2022 / Revised: 2 April 2022 / Accepted: 4 April 2022 / Published: 6 April 2022
(This article belongs to the Special Issue Plant Molecular Phylogenetics and Evolutionary Genomics III)

Abstract

:
The modern pteridophyte genus Equisetum is the only survivor of Sphenopsida, an ancient clade known from the Devonian. This genus, of nearly worldwide distribution, comprises approximately 15 extant species. However, genomic information is limited. In this study, we assembled the complete chloroplast genome of the giant species Equisetum xylochaetum from a metagenomic sequence and compared the plastid genome structure and protein-coding regions with information available for two other Equisetum species using network analysis. Equisetum chloroplast genomes showed conserved traits of quadripartite structure, gene content, and gene order. Phylogenetic analysis based on plastome protein-coding regions corroborated previous reports that Equisetum is monophyletic, and that E. xylochaetum is more closely related to E. hyemale than to E. arvense. Single-gene phylogenetic estimation and haplotype analysis showed that E. xylochaetum belonged to the subgenus Hippochaete. Single-gene haplotype analysis revealed that E. arvense, E. hyemale, E. myriochaetum, and E. variegatum resolved more than one haplotype per species, suggesting the presence of a high diversity or a high mutation rate of the corresponding nucleotide sequence. Sequences from E. bogotense appeared as a distinct group of haplotypes representing the subgenus Paramochaete that diverged from Hippochaete and Equisetum. In addition, the taxa that were frequently located at the joint region of the map were E. scirpoides and E. pratense, suggesting the presence of some plastome characters among the Equiseum subgenera.

1. Introduction

Equisetum L. is a genus of vascular plants that represents ancient Sphenopsida, a long-enduring clade known from fossils of the Devonian and later ages and, therefore, is considered useful in understanding the evolution of vascular plants. This genus is comprised of approximately 15 extant species, with a nearly worldwide distribution [1,2]. Previous studies have examined the evolutionary relationships among stem and crown Equisetum species using both morphology and genomic data. However, because morphology can vary as the result of hybridization and climate differences, molecular approaches have become popular. Recent studies have indicated three subgenera, including the primitive subgenus Paramochaete, and the later diverging subgenera Equisetum and Hippochaete. However, such relationships were estimated from relatively few plastid genes, e.g., rbcL, rps4, and trnL-F e.g., [3,4,5,6].
Among reported pteridophyte plastome sequences, only three were from Equisetum species: one from E. hyemale [7] and two from US and Korean E. arvense [8,9], which were placed in subgenera Equisetum and Hippochaete, respectively e.g., [3,4,5,6]. These reports revealed plastome variation among Equisetum species. The two E. arvense genomes differed by 417 bp and the E. hyemale genome was about 1.5 kbp smaller than that of E. arvense. In addition, rpl16 of E. arvense had an intron that is not present in E. hyemale [7,8,9]. These observations indicate that additional chloroplast genomes would be useful in evaluating evolutionary trends in this long-enduring genus.
Previous Equisetum plastome information was obtained using PCR amplification or from organelle-enriched DNA. The advancement of sequencing technologies and computational techniques allowed us to obtain complete organelle genome sequences from the shotgun metagenomic data we have archived for E. xylochaetum, presenting an additional technical option for obtaining Equisetum plastid genomes. Therefore, in this study, we assembled the complete plastid genome of E. xylochaetum, a giant species endemic to the Atacama Desert of Chile, South America, and used the information obtained to explore the evolution of Equisetum plastid genomes and the phylogenetic relationship of Equisetum species, as well as to determine whether the phylogeny estimated by using the popular plastid conserved regions was congruent with the haplotype mapping results of the corresponding sequences. Results showed that we successfully constructed the de novo plastid genome of E. xylochaetum using the shotgun metagenomic data. Phylogenetic estimations and comparison of Equisetum plastid genomes showed that E. xylochaetum was in the subgenus Hippochaete and that the Equisetum plastid genomes from subgenera Hippochaete and Equisetum were conserved in terms of genome structure, gene content, and gene order. Furthermore, results from TCS haplotype mapping showed that some of the taxa had a higher level of nucleotide diversity and some of the taxa shared common nucleotide haplotypes. Therefore, more conserved nucleotide regions and complete plastid genomes are needed for a better understanding of the evolutionary relationships of Equisetum.

2. Results

The chloroplast genome of E. xylochaetum displayed a quadripartite structure. The single-copy regions were 93,902 bp and 9726 bp, with two reverse repeated regions (IRa and IRb) of 14,386 bp in length. The GC contents of the LSC, SSC, and IR regions individually, and of the cp genome as a whole, were 31.5%, 30.9%, 48.4%, and 33.9%, respectively. The E. xylochaetum plastome encoded a total of 119 unique genes, of which nine were duplicated in the IR regions. Seventy-eight were protein-coding genes, 38 were tRNA genes, and eight were rRNA genes. Fourteen genes contain introns (atpF, clpP, ndhA, ndhB, petB, petD, rpl2, rpoC1, rps12, ycf3, trnK(uuu), trnL(uaa), trnV(uac), and trnI(gau)) as shown in Figure 1.
A comparison of Equisetum plastid genomes showed a collinear relationship, forming only one syntenic block in whole genome alignment. The genomes had similar genome size, % GC, gene content, gene length, and had identical gene order (Table 1 and Table 2). The protein-coding regions of E. xylochaetum plastid genes were subjected to purifying selection when compared against the corresponding protein-coding genes of E. hyemale and E. arvense.
Protein-coding regions of Equisetum species were similar in size, ranging between having the same length in atpB, E, F, H, I, clpP, infA, ndhB, C, D, E, H, I, petB, D, G, L, N, psaA, B, C, I, J, M, psbA, B, C, D, E, F, H, I, J, K, L, M, N, Z, rbcL, rpl14, 16, 22, 23, 32, 33, 36, rps2, 3, 4, 7, 8, 11, 12, 14, 15, 18, and 19 to having a 195 bp or 64 amino acids difference in accD. These genes have a similar number and position of introns except for the presence of 753 bp of intron in rpl16 in E. arvense. The percentage of the identical nucleotide of the aligned sites ranged from 88.7 percent in matK to 99.2 percent in psbJ, while the percentage of the identical derived amino acid of the aligned sites ranged from 73.9 percent in atpF to 100 percent in atpH, ndhE, petB, D, G, N, psaJ, psbA, E, F, I, J, L, Z, rpl36, and rps2 (Table 1). Phylogenetic estimation of Equisetum using plastome protein-coding sequences suggested that the known complete plastid genomes of Equisetum species formed a monophyletic clade of the two subgenera, Hippochaete and Equisetum. The newly assembled E. xylochaetum plastome indicates placement within Hippochaete with E. hyemale (Figure 2).
Single-gene ML phylogenetic analysis of atpB, matK, rpoB, rps4, and trnL-F resolved the known subgenera of Equisetum, including Paramochaete, Hippochaete, and Equisetum (Figure 3, Figure 4, Figure 5, Figure 6 and Figure 7). The majority of the Equisetum species were resolved with ML bootstrap values of at least 50. However, the monophyly of some Equisetum species could not be resolved. The monophyly of E. arvense and E. variegatum was not resolved in the matK tree, the monophyly of E. bogotense, E. laevigatum, E. myriochaetum, E. hyemale, and E. giganteum was not resolved in the rps4 tree, and the monophyly of E. hyemale, E. praealtum, E. ramosissimum, E. trachyodon, and E. xylochaetum was not resolved in the trnL-F tree. All hybrid taxa were phylogenetically placed within the clade consisting of the majority of their maternal parent, if the monophyly of the taxa was absent. In the case of the rps4 tree, these hybrids included Equisetum x fontqueri isolate 26093 located within the clade of E. telmateia, Equisetum x litorale isolates 41084 and 41085 with E. arvense, Equisetum x schaffneri isolates 40813 and 40824 with E. giganteum, and Equisetum x schaffneri isolate 40814 with E. myriochaetum. For trnL-F, the hybrid taxa Equisetum x ferrissii (AY226113) located in the clade with E. laevigatum, Equisetum x litorale isolates 41084 and 41085 with E. arvense, and Equisetum x schaffneri isolate 40814 with E. myriochaetum.
TCS haplotype network analyses using atpB, matK, and rpoB resolved distinct clades representing each of the Equisetum subgenera. At the species level, haplotype networks constructed using atpB and rpoB showed one haplotype for each Equisetum species. In contrast, maps of matK, rps4, and trnL-F resolved more than one haplotype for some species and resolved some haplotypes that consisted of more than one species. For matK, there was more than one haplotype for E. arvense and E. hyemale and there was one haplotype that consisted of sequences from E. arvense and E. variegatum (Figure 4).
Haplotype maps of rps4 and trnL-F seemed to be more complex compared to those of atpB, matK, and rpoB. In the map of rps4 (Figure 6), we observed 10 haplotypes, of which, haplotypes 1–3 of E. bogotense appeared as a distinct group representing subgenus Paramochaete. A few haplotypes consisted of only one Equisetum species, which were haplotype 4 for E. palustre, haplotype 5 for E. diffusum, and haplotype 8 for E. scirpoides. The hybrid taxa were embedded within the same haplotypes as their maternal taxa. These included Equisetum x fontqueri isolate 26093 that was in haplotype 6 with E. telmateia, Equisetum x litorale isolates 41084 and 41085 in haplotype 7 with E. arvense, Equisetum x schaffneri isolates 40813 and 40824 in haplotype 9 with E. giganteum, and Equisetum x schaffneri isolate 40814 in haplotype 9 with E. myriochaetum. Some haplotypes consisted of many plant species, i.e., haplotype 7 and 9, where the majority of Equisetum and Hippochaete were placed together, respectively. Interestingly, a rps4 sequence from E. hyemale grouped with other sequences of that species but also was present as a unique haplotype, as haplotype 10 with E. praealtum isolate 41501.
The map of trnL-F (Figure 7) resolved two distinct groups of haplotypes representing subgenus Paramochaete (haplotype 1) and subgenus Equisetum (haplotypes 2–8). Many of the Equisetum species were present as unique haplotypes, including E. bogotense (haplotype 1), E. palustre (haplotype 2), E. pratense (haplotype 3), E. telmateia (haplotype 4), E. sylvaticum (haplotype 5), E. fluviatile (haplotype 6), Equisetum x dycei (haplotype 7), and E. scirpoides (haplotype 9).
Some Equisetum species were resolved as more than a single haplotype. E. hyemale isolate 20201 was resolved as a unique haplotype 14 while E. hyemale isolate 1273o was located in haplotype 10 with E. variegatum. For E. variegatum, in addition to its member in haplotype 10, E. variegatum isolates 40820 and 40823 were resolved as additional unique haplotypes 11 and 12. In addition, E. myriochaetum isolate 40826 was present as haplotype 15, while most members were located in haplotype 13.
Most of the hybrid taxa in the trnL-F map were placed in the same haplotypes as their maternal taxa. Equisetum x ferrissii (AY226113) was located in haplotype 16 with its maternal taxon, E. laevigatum, Equisetum x ferrissii in haplotype 13 with the majority of E. hyemale, Equisetum x litorale isolates 41084 and 41085 in haplotype 8 with E. hyemale, Equisetum x schaffneri isolates 40813 and 40824 in haplotype 13 with E. giganteum, and Equisetum x schaffneri isolate 40814 in haplotype 13 with the majority of E. myriochaetum.

3. Discussion

In this study, we assembled the complete plastid genome of E. xylochaetum from shotgun metagenomes of E. xylochaetum sampled from two Atacama Desert locales exhibiting different degrees of disturbance. Results showed that the plastid genomes constructed from these two E. xylochaetum metagenome accessions were identical, suggesting that the Equisetum samples were from the same Equisetum population. Comparison of nucleotide, and their derived protein, sequences of this newly assembled E. xylochaetum plastid genome to those of E. hyemale and E. arvense showed that the Equisetum plastid genomes were highly conserved in terms of structure and function, even though the two subgenera (Hippochaete and Equisetum) might have diverged as early as 135 Mya during the early Cretaceous [4,6]. All the plastid protein-coding sequences were subjected to purifying selection, with genes of the same type having identical nucleotide percentages and having nucleotide identity ranging from 88.7–99.2 percent. The only major difference in gene structure was presence of the intron in E. arvense rpl16. To determine where and when the intron of rp116 originated in the Equisetum lineage, more Equisetum rpl16 sequences or complete plastid genomes are required.
In a broad sense, the phylogenetic positions of Equisetum species inferred by using all protein-coding sequences along with their derived proteins and the single-gene analysis present in this study were congruent with results from previous studies that used a single-gene approach [11] or a combination of multi-genes and morphological characters e.g., [4,5,6], where Equisetum formed monophyletic clades of each subgenus and placed E. xylochaetum in Hippochaete. Despite the presence of the high conservation level of Equisetum plastid genes, it was surprising to us that the single-gene phylogenetic approach was not sufficient to resolve relationships among Equisetum taxa, especially those closely related taxa placed in subgenus Hippochaete, e.g., E. giganteum, E. variegatum, and E. hyemale. Therefore, it is evident that more Equisetum plastid genomes, plus additional molecular information from other genetic compartments, are needed.
The addition of haplotype mapping provided in this study enhanced the understanding of how plastid genes from each taxon are related. In general, the haplotype maps reflected the relationship resolved from phylogenetic estimation using the corresponding nucleotide regions. Even so, these new maps aid the visualisation of how these plastome nucleotide data were interrelated to each other at the level of isolate, species, and subgenus. The presence of only one shared distinct haplotype of an Equisetum species, though its samples were collected from different locales, suggested a high conservation level of the corresponding genes within its plastid genomes. On the other hand, the presence of more than one haplotype at the specific level suggested the presence of nucleotide diversity, indicating the need to further examine the populations of E. arvense, E. bogotense, E. hyemale, E. variegatum, and E. myriochaetum. In addition, the presence of a haplotype consisting of more than one Equisetum species, e.g., haplotypes 8 and 9 of the rps4 map and haplotypes 8 and 13 of the trnL-F map, suggested that these conserved regions alone were not sufficient for studying the relationship and diversity of Equisetum taxa. These findings emphasize the need for more Equisetum plastid genomes.
The presence of distinct haplotype(s) in the early-diverging species E. bogotense in rps4 and trnL-F suggested that these plastid sequences might not represent the ancestral characters of Equisetum. Instead, these E. bogotense samples may only represent the survival representatives of the extinct members that also evolved during the course of time. In contrast, according to the rps4 and trnL-F maps, the taxa that frequently occurred at the junction region between each subgenus were E. scirpoides and E. pratense, suggesting that these taxa might be particularly helpful for understanding how the Equisetum subgenera diverged.

4. Materials and Methods

Nucleotide data for Equisetum xylochaetum Mett. were obtained from GenBank BioProject PRJNA555713 [12], generated by metagenomic shotgun sequencing of the microbiome of giant Equisetum xylochaetum sampled from two streambed locales in the Atacama Desert of northern Chile that differed in the degree of human disturbance. The two raw data sets, separately archived in accessions SRX6486516 and SRX6486517, each represented pooled replicate DNA extractions from both above-ground green and below-ground non-green tissues. To obtain the complete chloroplast genome of E. xylochaetum, metagenomic sequences were trimmed using Trimmomatic v. 0.39 [13] using the parameter sliding window:4:30. Next, the trimmed sequences from the two raw data sets were independently assembled using MEGAHIT ver. 1.2.9 [14] with the parameter “bubble-level equal to 0” in order to prevent the merging of sequences that were highly similar, e.g., sequences from closely related species or sequences that display single nucleotide polymorphisms. Each assembly yielded a contig of the complete plastid genome of E. xylochaetum, and these two contigs were identical in sequence. To validate the assembly, we calculated the coverage of the plastid genomes using the methods described in Satjarak and Graham [15]. One of the two contigs, which had the mean coverage of 706 fold, was then selected for annotation of protein-coding genes using proteins inferred from E. arvense [8,9] and E. hyemale [7] as references. The tRNAs and rRNA genes were annotated using tRNAscan-SE On-line [16] and the RNAmmer 1.2 Server [17], respectively. The complete plastid genome of E. xylochaetum was deposited in GenBank under accession number MW282958. A representative plant specimen has been deposited at the University of Concepción herbarium under accession number CONC-CH 6005.
We compared the plastid genome of E. xylochaetum obtained from this study to other complete Equisetum plastid genomes, including E. hyemale (KC117177), E. arvense from the US (GU191334), and E. arvense from Korea (JN968380). We examined general characteristics of the genomes, including the genome size, %GC, the gene content, gene length, gene order, and polymorphism of nucleotides within coding regions and their derived proteins. To consider nucleotide polymorphisms, we aligned the protein-coding sequences (Table 1) using Geneious translation alignment: global alignment with free end gap, standard genetic code, and Identity (1.0/0.0) cost matrix (Geneious ver. 9.1.3; https://www.geneious.com; accessed in 31 January 2022).
The mode of evolution of protein-coding regions was performed using the method described in Mekvipad and Satjarak [18]. For the polymorphism of protein sequences, we aligned the derived protein sequences using MAFFT alignment: auto algorithm and Blosum62 scoring matrix [19]. To investigate the relationship of E. xylochaetum and other Equisetum complete chloroplast genomes, Psilotum nudum (NC_003386.1) was used as an outgroup. The protein-coding sequences and protein sequences (Table 1) were similarly aligned, trimmed using Trimal ver. 1.2 [20], and concatenated. The nucleotide data matrix was 60,987 bp and the protein data matrix consisted of 19,435 amino acids. Phylogenetic relationships were estimated using maximum likelihood and Bayesian frameworks as described in Satjarak and Graham [15].
To investigate whether the Equisetum relationship resolved from frequently-used nucleotide sequences reported in previous studies exhibited grades or evolutionary intermediates, we performed haplotype network analysis of selected, frequently-used Equisetum conserved regions. These included atpB, matK, rpoB, rps4, and trnL-F (Table 3). To prepare the data matrices, the conserved nucleotide regions were extracted from the complete plastid genomes and from DNA sequences from other published studies (Table 3). Next, the data were aligned, and the phylogenetic trees were estimated using the methods described above. The haplotype network analysis was calculated using (Templeton, Crandall, and Sing; TCS) [21] and visualized in PopArt v1.7 [22].

5. Conclusions

In summary, our study demonstrated that metagenomic data can be a useful way to obtain plastid genomes. The comparison of the de novo plastid genome of E. xylochaetum with other reported Equisetum plastomes showed a high degree of conservation in terms of structure, gene content, gene order, and nucleotide polymorphisms. Even so, this new plastid genome provided additional information about the evolution and diversity of Equisetum, e.g., the presence of an intron in rpl16. Haplotype analyses of the selected conserved nucleotides showed that some Equisetum species were distantly related to other taxa, inferred from the presence of distinct haplotypes. Many of the taxa appeared as shared haplotypes, suggesting that the molecular data we currently have might not be sufficient for a full understanding of the evolutionary relationship of Equisetum and that more Equisetum plastid genomes are needed.

Author Contributions

Conceptualization, A.S.; methodology, A.S.; validation, A.S.; formal analysis, A.S.; resources, L.E.G., M.T.T. and P.A.-A.; data curation, A.S.; writing—original draft preparation, A.S.; writing—review and editing, A.S., L.E.G., M.T.T. and P.A.-A.; visualization, A.S.; funding acquisition, L.E.G. and P.A.-A. All authors have read and agreed to the published version of the manuscript.

Funding

This work was partially supported by US NSF grant DEB1119944 (to L.E.G.) and Chilean CONICYT-FONDECYT grant 1120619 (to P.A.-A.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The complete plastid genome of Equisetum xylochaetum is publicly available in NCBI GenBank (https://www.ncbi.nlm.nih.gov accessed on 12 March 2022) accession number MW282958. Nucleotide data for analysis are available at GenBank BioProject PRJNA555713 accessions SRX6486516 and SRX6486517.

Acknowledgments

We thank Karnjana Ruen-pham for the illustration.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hauke, R.L. A taxonomical monograph of the genus Equisetum subgenus. Nova Hedwig. 1963, 8, 1–123. [Google Scholar]
  2. Husby, C. Biology and functional ecology of Equisetum with emphasis on the giant horsetails. Bot. Rev. 2013, 79, 147–177. [Google Scholar] [CrossRef]
  3. Guillon, J.M. Phylogeny of horsetails (Equisetum) based on the chloroplast rps4 gene and adjacent noncoding sequences. Syst. Bot. 2004, 29, 251–259. [Google Scholar] [CrossRef]
  4. Elgorriaga, A.; Escapa, I.H.; Rothwell, G.W.; Tomescu, A.M.; Rubén Cúneo, N. Origin of Equisetum: Evolution of horsetails (Equisetales) within the major euphyllophyte clade Sphenopsida. Am. J. Bot. 2018, 105, 1286–1303. [Google Scholar] [CrossRef]
  5. Christenhusz, M.J.; Bangiolo, L.; Chase, M.W.; Fay, M.F.; Husby, C.; Witkus, M.; Viruel, J. Phylogenetics, classification and typification of extant horsetails (Equisetum, Equisetaceae). Bot. J. Linn. Soc. 2019, 189, 311–352. [Google Scholar] [CrossRef]
  6. Christenhusz, M.J.; Chase, M.W.; Fay, M.F.; Hidalgo, O.; Leitch, I.J.; Pellicer, J.; Viruel, J. Biogeography and genome size evolution of the oldest extant vascular plant genus, Equisetum (Equisetaceae). Ann. Bot. 2021, 127, 681–695. [Google Scholar] [CrossRef]
  7. Grewe, F.; Guo, W.; Gubbels, E.A.; Hansen, A.K.; Mower, J.P. Complete plastid genomes from Ophioglossum californicum, Psilotum nudum, and Equisetum hyemale reveal an ancestral land plant genome structure and resolve the position of Equisetales among monilophytes. BMC Evol. Biol. 2013, 13, 8. [Google Scholar] [CrossRef] [Green Version]
  8. Karol, K.G.; Arumuganathan, K.; Boore, J.L.; Duffy, A.M.; Everett, K.D.; Hall, J.D.; Hansen, S.K.; Kuehl, J.V.; Mandoli, D.F.; Mishler, B.D.; et al. Complete plastome sequences of Equisetum arvense and Isoetes flaccida: Implications for phylogeny and plastid genome evolution of early land plant lineages. BMC Evol. Biol. 2010, 10, 321. [Google Scholar] [CrossRef] [Green Version]
  9. Kim, H.T.; Kim, K.J. Chloroplast genome differences between Asian and American Equisetum arvense (Equisetaceae) and the origin of the hypervariable trnY-trnE intergenic spacer. PLoS ONE 2014, 9, e103898. [Google Scholar]
  10. Greiner, S.; Lehwark, P.; Bock, R. OrganellarGenomeDRAW (OGDRAW) version 1.3. 1: Expanded toolkit for the graphical visualization of organellar genomes. Nucleic Acids Res. 2019, 47, W59–W64. [Google Scholar] [CrossRef] [Green Version]
  11. Guillon, J.M. Molecular phylogeny of horsetails (Equisetum) including chloroplast atpB sequences. J. Plant Res. 2007, 120, 569–574. [Google Scholar] [CrossRef] [PubMed]
  12. Satjarak, A.; Piotrowski, M.J.; Graham, L.E.; Trest, M.T.; Wilcox, L.W.; Knack, J.J.; Cook, M.E.; Arancibia-Avila, P. Microscopic and metagenomic evidence for eukaryotic microorganisms associated with Atacama Desert populations of giant Equisetum. Am. Fern. J. 2021, 111, 86–109. [Google Scholar] [CrossRef]
  13. Bolger, A.M.; Lohse, M.; Usadel, B. Trimmomatic: A flexible trimmer for Illumina sequence data. Bioinformatics 2014, 30, 2114–2120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Li, D.; Liu, C.M.; Luo, R.; Sadakane, K.; Lam, T.W. MEGAHIT: An ultra-fast single-node solution for large and complex metagenomics assembly via succinct de Bruijn graph. Bioinformatics 2015, 31, 1674–1676. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Satjarak, A.; Graham, L.E. Comparative DNA sequence analyses of Pyramimonas parkeae (Prasinophyceae) chloroplast genomes. J. Phycol. 2017, 53, 415–424. [Google Scholar] [CrossRef] [PubMed]
  16. Lowe, T.M.; Chan, P.P. tRNAscan-SE On-line: Integrating search and context for analysis of transfer RNA genes. Nucleic Acids Res. 2016, 44, W54–W57. [Google Scholar] [CrossRef]
  17. Lagesen, K.; Hallin, P.; Rødland, E.A.; Stærfeldt, H.H.; Rognes, T.; Ussery, D.W. RNAmmer: Consistent and rapid annotation of ribosomal RNA genes. Nucleic Acids Res. 2007, 35, 3100–3108. [Google Scholar] [CrossRef]
  18. Mekvipad, N.; Satjarak, A. Evolution of organellar genes of chlorophyte algae: Relevance to phylogenetic inference. PLoS ONE 2019, 14, e0216608. [Google Scholar] [CrossRef]
  19. Katoh, K.; Standley, D.M. MAFFT multiple sequence alignment software version 7: Improvements in performance and usability. Mol. Biol. Evol. 2013, 30, 772–780. [Google Scholar] [CrossRef] [Green Version]
  20. Capella-Gutiérrez, S.; Silla-Martínez, J.M.; Gabaldón, T. trimAl: A tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics 2009, 25, 1972–1973. [Google Scholar] [CrossRef]
  21. Crandall, M.C.D.P.K.; Clement, M.; Posada, D. TCS: A computer program to estimate gene genealogies. Mol. Ecol. 2000, 9, 1657–1660. [Google Scholar]
  22. Leigh, J.W.; Bryant, D. POPART: Full-feature software for haplotype network construction. Methods Ecol. Evol. 2015, 6, 1110–1116. [Google Scholar] [CrossRef]
  23. Murdock, A.G. Phylogeny of marattioid ferns (Marattiaceae): Inferring a root in the absence of a closely related outgroup. Am. J. Bot. 2008, 95, 626–641. [Google Scholar] [CrossRef] [PubMed]
  24. Pryer, K.M.; Schneider, H.; Smith, A.R.; Cranfill, R.; Wolf, P.G.; Hunt, J.S.; Sipes, S.D. Horsetails and ferns are a monophyletic group and the closest living relatives to seed plants. Nature 2001, 409, 618–622. [Google Scholar] [CrossRef]
  25. Cook, R.; Hennell, J.R.; Lee, S.; Khoo, C.S.; Carles, M.C.; Higgins, V.J.; Govindaraghavan, S.; Sucher, N.J. The Saccharomyces cerevisiae transcriptome as a mirror of phytochemical variation in complex extracts of Equisetum arvense from America, China, Europe and India. BMC Genom. 2013, 14, 1–18. [Google Scholar] [CrossRef] [Green Version]
  26. Hausner, G.; Olson, R.; Simon, D.; Johnson, I.; Sanders, E.R.; Karol, K.G.; McCourt, R.M.; Zimmerly, S. Origin and evolution of the chloroplast trnK (matK) intron: A model for evolution of group II intron RNA structures. Mol. Biol. Evol. 2006, 23, 380–391. [Google Scholar] [CrossRef]
  27. Knie, N.; Fischer, S.; Grewe, F.; Polsakiewicz, M.; Knoop, V. Horsetails are the sister group to all other monilophytes and Marattiales are sister to leptosporangiate ferns. Mol. Phylogenet. Evol. 2015, 90, 140–149. [Google Scholar] [CrossRef]
  28. Fazekas, A.J.; Burgess, K.S.; Kesanakurti, P.R.; Graham, S.W.; Newmaster, S.G.; Husband, B.C.; Percy, D.M.; Hajibabaei, M.; Barrett, S.C. Multiple multilocus DNA barcodes from the plastid genome discriminate plant species equally well. PLoS ONE 2008, 3, e2802. [Google Scholar] [CrossRef] [Green Version]
  29. Magallón, S.; Hilu, K.W.; Quandt, D. Land plant evolutionary timeline: Gene effects are secondary to fossil constraints in relaxed clock estimation of age and substitution rates. Am. J. Bot. 2013, 100, 556–573. [Google Scholar] [CrossRef] [Green Version]
  30. Scharn, R.; Little, C.J.; Bacon, C.D.; Alatalo, J.M.; Antonelli, A.; Björkman, M.P.; Molau, U.; Nilsson, R.H.; Björk, R.G. Decreased soil moisture due to warming drives phylogenetic diversity and community transitions in the tundra. Environ. Res. Lett. 2021, 16, 064031. [Google Scholar] [CrossRef]
  31. Kuo, L.Y.; Li, F.W.; Chiou, W.L.; Wang, C.N. First insights into fern matK phylogeny. Mol. Phylogenet. Evol. 2011, 59, 556–566. [Google Scholar] [CrossRef] [PubMed]
  32. Gao, L.; Zhou, Y.; Wang, Z.W.; Su, Y.J.; Wang, T. Evolution of the rpoB-psbZ region in fern plastid genomes: Notable structural rearrangements and highly variable intergenic spacers. BMC Plant Biol. 2011, 11, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Newton, A.E.; Cox, C.J.; Duckett, J.G.; Wheeler, J.A.; Goffinet, B.; Hedderson, T.A.; Mishler, B.D. Evolution of the major moss lineages: Phylogenetic analyses based on multiple gene sequences and morphology. Bryologist 2000, 103, 187–211. [Google Scholar] [CrossRef]
  34. Des Marais, D.L.; Smith, A.R.; Britton, D.M.; Pryer, K.M. Phylogenetic relationships and evolution of extant horsetails, Equisetum, based on chloroplast DNA sequence data (rbcL and trnL-F). Int. J. Plant Sci. 2003, 164, 737–751. [Google Scholar] [CrossRef] [Green Version]
  35. Lang, A.; Naciri, Y. New chloroplast primers for intraspecific variation in Dicranum scoparium Hedw. (Dicranaceae) and amplification success in other bryophyte species. Mol. Ecol. Resour. 2010, 10, 735–737. [Google Scholar] [CrossRef]
  36. Hiiesalu, I.; Oepik, M.; Metsis, M.; Lilje, L.; Davison, J.; Vasar, M.; Moora, M.; Zobel, M.; Wilson, S.D.; Paertel, M. Plant species richness belowground: Higher richness and new patterns revealed by next-generation sequencing. Mol. Ecol. 2012, 21, 2004–2016. [Google Scholar] [CrossRef]
  37. Soininen, E.M.; Valentini, A.; Coissac, E.; Miquel, C.; Gielly, L.; Brochmann, C.; Brysting, A.K.; Sønstebø, J.H.; Ims, R.A.; Yoccoz, N.G.; et al. Analysing diet of small herbivores: The efficiency of DNA barcoding coupled with high-throughput pyrosequencing for deciphering the composition of complex plant mixtures. Front Zool. 2009, 6, 1–9. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Circular map of the Equisetum xylochaetum plastid genome, NCBI accession MW282958, drawn by OGDRAW version 1.3.1 [10]. Genes positioned on the outside of the map are transcribed counterclockwise and those inside the map are transcribed clockwise. The thick lines indicate the extent of the inverted repeat regions.
Figure 1. Circular map of the Equisetum xylochaetum plastid genome, NCBI accession MW282958, drawn by OGDRAW version 1.3.1 [10]. Genes positioned on the outside of the map are transcribed counterclockwise and those inside the map are transcribed clockwise. The thick lines indicate the extent of the inverted repeat regions.
Plants 11 01001 g001
Figure 2. Maximum-likelihood tree inferred from all Equisetum plastome protein-coding regions using a GTR+I+F model. The scale bar represents the estimated number of nucleotide substitutions per site. The bootstrap and posterior probability values are reported at the respective nodes. The values include the ML bootstrap values of nucleotide and protein data and the BI posterior probability of the nucleotide and protein data, respectively.
Figure 2. Maximum-likelihood tree inferred from all Equisetum plastome protein-coding regions using a GTR+I+F model. The scale bar represents the estimated number of nucleotide substitutions per site. The bootstrap and posterior probability values are reported at the respective nodes. The values include the ML bootstrap values of nucleotide and protein data and the BI posterior probability of the nucleotide and protein data, respectively.
Plants 11 01001 g002
Figure 3. Phylogenetic estimation and TCS network of Equisetum atpB sequences. The scale bar of the tree represents the estimated number of nucleotide substitutions per site. The maximum-likelihood bootstrap values are reported at the respective nodes. The colours of taxa present in the tree correspond with the colours in the TCS haplotype map. The size of the circle represents the number of the taxa that share the same haplotype.
Figure 3. Phylogenetic estimation and TCS network of Equisetum atpB sequences. The scale bar of the tree represents the estimated number of nucleotide substitutions per site. The maximum-likelihood bootstrap values are reported at the respective nodes. The colours of taxa present in the tree correspond with the colours in the TCS haplotype map. The size of the circle represents the number of the taxa that share the same haplotype.
Plants 11 01001 g003
Figure 4. Phylogenetic estimation and TCS network of Equisetum matK sequences. The scale bar of the tree represents the estimated number of nucleotide substitutions per site. The maximum-likelihood bootstrap values are reported at the respective nodes. The colours of taxa present in the tree correspond with the colours in the TCS haplotype map. The size of the circle represents the number of the taxa that share the same haplotype.
Figure 4. Phylogenetic estimation and TCS network of Equisetum matK sequences. The scale bar of the tree represents the estimated number of nucleotide substitutions per site. The maximum-likelihood bootstrap values are reported at the respective nodes. The colours of taxa present in the tree correspond with the colours in the TCS haplotype map. The size of the circle represents the number of the taxa that share the same haplotype.
Plants 11 01001 g004
Figure 5. Phylogenetic estimation and TCS network of Equisetum rpoB sequences. The scale bar of the tree represents the estimated number of nucleotide substitutions per site. The maximum-likelihood bootstrap values are reported at the respective nodes. The colours of taxa present in the tree correspond with the colours in the TCS haplotype map. The size of the circle represents the number of the taxa that share the same haplotype.
Figure 5. Phylogenetic estimation and TCS network of Equisetum rpoB sequences. The scale bar of the tree represents the estimated number of nucleotide substitutions per site. The maximum-likelihood bootstrap values are reported at the respective nodes. The colours of taxa present in the tree correspond with the colours in the TCS haplotype map. The size of the circle represents the number of the taxa that share the same haplotype.
Plants 11 01001 g005
Figure 6. Phylogenetic estimation and TCS network of Equisetum rps4 sequences. The scale bar of the tree represents the estimated number of nucleotide substitutions per site. The maximum-likelihood bootstrap values are reported at the respective nodes. The colours of taxa present in the tree correspond with the colours in the TCS haplotype map. The size of the circle represents the number of the taxa that share the same haplotype.
Figure 6. Phylogenetic estimation and TCS network of Equisetum rps4 sequences. The scale bar of the tree represents the estimated number of nucleotide substitutions per site. The maximum-likelihood bootstrap values are reported at the respective nodes. The colours of taxa present in the tree correspond with the colours in the TCS haplotype map. The size of the circle represents the number of the taxa that share the same haplotype.
Plants 11 01001 g006
Figure 7. Phylogenetic estimation and TCS network of Equisetum trnL-trnF. The scale bar of the tree represents the estimated number of nucleotide substitutions per site. The maximum-likelihood bootstrap values are reported at the respective nodes. The colours of taxa present in the tree correspond with the colours in the TCS haplotype map. The size of the circle represents the number of the taxa that share the same haplotype.
Figure 7. Phylogenetic estimation and TCS network of Equisetum trnL-trnF. The scale bar of the tree represents the estimated number of nucleotide substitutions per site. The maximum-likelihood bootstrap values are reported at the respective nodes. The colours of taxa present in the tree correspond with the colours in the TCS haplotype map. The size of the circle represents the number of the taxa that share the same haplotype.
Plants 11 01001 g007
Table 1. Comparison of general characters of Equisetum plastid genomes.
Table 1. Comparison of general characters of Equisetum plastid genomes.
E. xylochaetumE. hyemaleE. arvense (US)E. arvense (Korea)
AccessionMW282958KC117177GU191334JN968380
genome size132,400131,760133,309132,726
LSC93,90292,58093,54292,961
SSC9,72618,99419,46919,477
IRs14,38610,09310,14910,144
%GC33.933.733.433.4
Table 2. Protein-coding gene content and introns of Equisetum plastid genomes. Comparison showed percent identity and size of the gene and its derived proteins.
Table 2. Protein-coding gene content and introns of Equisetum plastid genomes. Comparison showed percent identity and size of the gene and its derived proteins.
DNAProtein
No.Gene (# of Intron)Identical Site (%)Mean (bp)SD (bp)min (bp)max (bp)Identical Site (%)Mean (aa)SD (aa)min (aa)max (aa)
1.accD92.81021.581948114389.934026.5316380
2.atpA95.5153920.815271575985126.9508524
3.atpB94.4147001470147096.34890489489
4.atpE93.4396039639693.11310131131
5.atpF (1)97.5555055555573.91840184184
6.atpH96.324602462461008108181
7.atpI95.4747074774799.12480284248
8.ccsA93.3943.51.594294591.1313.50.5313314
9.cemA89.51438.521.71425147682.7478.57.2474491
10.chlB94.31549.54.51545155493.2515.51.5514517
11.chlL91.6879687388593.92922290294
12.chlN92.81300.510.51290131190.4432.53.5429436
13.clpP (1)95.1615061561598.52040204204
14.infA94.7243024324396.38008080
15.matK88.7147031467147381.44891488490
16.ndhA (1)92.71101.81.31101110491.8366.30.4366367
17.ndhB (1)94.8147301473147394.34900490490
18.ndhC95.9363036336398.31200120120
19.ndhD95.1149701497149795.24980498498
20.ndhE98.330303033031001000100100
21.ndhF92.82221.51.52220222392.2739.50.5739740
22.ndhG91.460617.258563385.72015.7194210
23.ndhH95.3118201182118297.23930393393
24.ndhI97.3549054954998.41820182182
25.ndhJ93.9520.54.551652594.8172.51.5171174
26.ndhK90.6747.89.173275386.8248.33243250
27.petA92.4955.57.594896393.8317.52.5315320
28.petB (1)9664806486481002150215215
29.petD (1)96.948304834831001600160160
30.petG97.411401141141003703737
31.petL93.8960969693.53103131
32.petN9996096961003103131
33.psaA96.2225302253225399.67500750750
34.psaB95.7220502205220599.27340734734
35.psaC95.1246024624698.88108181
36.psaI91.9111011111194.43603636
37.psaJ97.712901291291004204242
38.psaM96990999996.93203232
39.psbA98.110620106210621003530353353
40.psbB96.11527015271527995080508508
41.psbC95142201422142299.44730473473
42.psbD95.6106201062106287.33530353353
43.psbE97.224602462461008108181
44.psbF98.312001201201003903939
45.psbH94.7225022522589.27407474
46.psbI97.311101111111003603636
47.psbJ99.212301231231004004040
48.psbK97168016816896.45505555
49.psbL98.311701171171003803838
50.psbM98.2111011111194.43603636
51.psbN95.51320132132934304343
52.psbT97.4112.51.511111497.336.50.53637
53.psbZ94.218901891891006206262
54.rbcL96.1142801428142899.24750475475
55.rpl1497.6369036936999.21220122122
56.rpl16 (1 in E. arvense)93.14230423423951400140140
57.rpl2 (1)94.3834.81.383483795.3277.30.4277278
58.rpl2089.7347.31.334534885.2114.80.4114115
59.rpl2191.3364.51.536336686120.50.5120121
60.rpl2294.1372037237296.61230123123
61.rpl2394.9273027327393.39009090
62.rpl3295.3171017117198.25605656
63.rpl3395.5201020120190.96606666
64.rpl3693.911401141141003703737
65.rpoA93.51.18.54.51014102393.5338.51.5337340
66.rpoB93.93235.533.83216329492.91.00.511.310711097
67.rpoC1 (1)93.32060.33.92058206791.3685.81.3685688
68.rpoC292.34143214122416487.21380713731387
69.rps1194.7396039639695.41310131131
70.rps1298.137203723721001230123123
71.rps1492.2306030630693.11010101101
72.rps1595.2270027027092.18908989
73.rps1896.5228022822898.77507575
74.rps1995.7279027927998.99209292
75.rps295.27080708708972350235235
76.rps395.4657065765796.32180218218
77.rps494.1624062462492.32070207207
78.rps794.9468046846894.81550155155
79.rps895.7399039939995.51320132132
Table 3. Nucleotide sequences used in the single gene phylogenetic analysis and TCS haplotype mapping.
Table 3. Nucleotide sequences used in the single gene phylogenetic analysis and TCS haplotype mapping.
No.NameGenBank AccessionLocalityReferences
atpB
1.E. arvenseGU191334USA[8]
2.E. arvenseJN968380Korea[9]
3.E. hyemaleKC117177unknown[7]
4.E. ramosissimum subsp. debileEU439074unknown[23]
5.E. telmateiaAF313542unknown[24]
6.E. xylochaetumMW282958ChileThis study
7.Equisetum x ferrissiiAF313541unknown[24]
matK
1.E. arvenseJX392862China[25]
2.E. arvenseJX392863Europe[25]
3.E. arvenseAY348551unknown[26]
4.E. arvenseGU191334USA[8]
5.E. arvenseJN968380Korea[9]
6.E. bogotenseKP757846unknown[27]
7.E. hyemaleEU749486unknown[28]
8.E. hyemaleEU749485unknown[28]
9.E. hyemaleEU749484unknown[28]
10.E. hyemaleEU749487unknown[28]
11.E. hyemaleHF585136unknown[29]
12.E. hyemaleKC117177unknown[7]
13.E. palustreMZ400482Sweden[30]
14.E. ramosissimumJF303895unknown[31]
15.E. scirpoidesMZ400480Sweden[30]
16.E. variegatumMZ400481Sweden[30]
17.E. xylochaetumMW282958ChileThis study
rpoB
1.E. arvenseHQ658110China[32]
2.E. arvenseGU191334USA[8]
3.E. arvenseJN968380Korea[9]
4.E. hyemaleKC117177Unknown[7]
5.E. ramossissimumHQ658109China[32]
6.E. xylochaetumMW282958ChileThis study
rps4
1.E. arvense subsp. arvense isolate 41072MH750111Finland [5]
2.E. arvenseAJ583677unknown[3]
3.E. arvenseJN968380Korea[9]
4.E. arvenseGU191334USA[8]
5.E. arvense subsp. arvense isolate 26084MH750108India (Himachal Pradesh) [5]
6.E. arvense subsp. arvense isolate 40833MH750109USA (California) [5]
7.E. arvense subsp. arvense isolate 41071MH750110Finland [5]
8.E. arvense subsp. boreale isolate 41073MH750112Finland/Norway (border) [5]
9.E. arvense subsp. boreale isolate 41074MH750113Finland/Norway (border)[5]
10.E. arvense x E. telmateia subsp. braunii isolate 40834MH750114USA (California) [5]
11.E. bogotenseAF231898unknown[33]
12.E. bogotenseAF313603unknown[24]
13.E. bogotenseAJ583678unknown[3]
14.E. bogotense isolate 40800MH750115Argentina[5]
15.E. bogotense isolate 40802MH750116Ecuador[5]
16.E. bogotense isolate 40827MH750117Colombia[5]
17.E. diffusumAJ583679unknown[3]
18.E. diffusum isolate 40804MH750118 India [5]
19.E. fluviatileAJ583680unknown[3]
20.E. fluviatile isolate 41075MH750119Finland [5]
21.E. fluviatile isolate 41076MH750120Finland [5]
22.E. giganteumAJ583681unknown[3]
23.E. giganteum isolate 40806MH750121Chile[5]
24.E. hyemaleAJ583682unknown[3]
25.E. hyemaleKC117177unknown[7]
26.E. hyemale isolate 23252MH750123Norway[5]
27.E. laevigatumAJ583683unknown[3]
28.E. laevigatum isolate 40812MH750125USA (California) [5]
29.E. myriochaetumAJ583684unknown[3]
30.E. myriochaetum isolate 40825MH750126Mexico [5]
31.E. myriochaetum isolate 40936MH750127El Salvador [5]
32.E. palustreAJ583685unknown[3]
33.E. palustre isolate 17671MH750128UK (England, Norfolk) [5]
34.E. palustre isolate 39349MH750129UK (England, Surrey) [5]
35.E. praealtum isolate 41501MH750122USA (Ohio) [5]
36.E. pratenseAJ583686unknown[3]
37.E. pratense isolate 39348MH750130Finland [5]
38.E. ramosissimum subsp. debileAJ583687unknown[3]
39.E. ramosissimum subsp. debileEU439173unknown[23]
40.E. ramosissimum subsp. debile isolate 24579MH750131Sri Lanka [5]
41.E. ramosissimum subsp. debile isolate 40837MH750132New Caledonia [5]
42.E. ramosissimum subsp. ramosissimum isolate 36802MH750133Spain (Andalucia) [5]
43.E. scirpoidesAJ583688unknown[3]
44.E. scirpoides isolate 26090MH750134Greenland [5]
45.E. scirpoides isolate 40830MH750124Russia (Kamtschatka) [5]
46.E. sylvaticumAJ583689unknown[3]
47.E. telmateia subsp. brauniiAJ583690unknown[3]
48.E. telmateia subsp. braunii isolate 40817 MH750136USA (California) [5]
49.E. telmateia subsp. braunii isolate 40828MH750137Canada (British Columbia)[5]
50.E. telmateia subsp. braunii isolate 40832MH750138USA (California) [5]
51.E. telmateia subsp. braunii isolate 40836MH750139USA (California) [5]
52.E. telmateia subsp. telmateia isolate 41082MH750140Ireland[5]
53.E. variegatumAJ583691unknown[3]
54.E. variegatum isolate 11639MH750141UK (Wales) [5]
55.E. variegatum isolate 40819MH750148Ireland [5]
56.E. variegatum isolate 40820MH750142France (Pyrenees) [5]
57.E. variegatum isolate 40823MH750143USA (Keweenaw, Michigan) [5]
58.E. variegatum isolate 41083MH750144Ireland [5]
59.E. variegatum subsp. alaskanum isolate 40818MH750145USA (Alaska) [5]
60.E. variegatum subsp. alaskanum isolate 40821MH750146Canada (British Columbia)[5]
61.E. variegatum subsp. alaskanum isolate 40822MH750147Canada (Banff)[5]
62.E. xylochaetumMW282958ChileThis study
63.Equisetum scirpoides isolate 41089MH750135Finland[5]
64.Equisetum x ferrissiiAF313590unknown[24]
65.Equisetum x fontqueri isolate 26093
(E. telmateia x E. palustre)
MH750149UK (Scotland) [5]
66.Equisetum x litorale isolate 41084
(E. arvense x E. fluviatile)
MH750150Ireland [5]
67.Equisetum x litorale isolate 41085
(E. arvense x E. fluviatile)
MH750151Ireland [5]
68.Equisetum x schaffneri isolate 40813
(E. giganteum x E. myriochaetum)
MH750152Mexico [5]
69.Equisetum x schaffneri isolate 40814
(E. myriochaetum x E. giganteum)
MH750153Peru (cult RBG Edinburgh) [5]
70.Equisetum x schaffneri isolate 40824
(E. giganteum x E. myriochaetum)
MH750154Mexico [5]
trnL-trnF
1.E. arvenseJN968380Korea[9]
2.E. arvenseGU191334USA[8]
3.E. arvenseAY226125Franc[34]
4.E. arvenseGQ428069unknown[35]
5.E. arvenseHM590277Estonia[36]
6.E. arvenseGQ244921unknown[37]
7.E. arvense subsp boreale isolate 41074MH750043Finland/Norway[5]
8.E. arvense subsp. arvense isolate 26084MH750038India[5]
9.E. arvense subsp. arvense isolate 26085MH750039UK[5]
10.E. arvense subsp. arvense isolate 40833MH750040USA[5]
11.E. arvense subsp. arvense isolate 41071MH750041Finland[5]
12.E. arvense subsp. boreale isolate 41073MH750042Finland/Norway[5]
13.E. arvense x E. telmateia subsp. braunii isolate 40834MH750044USA[5]
14.E. bogotenseAY226124Colombia[34]
15.E. bogotense isolate 40800MH750045Argentina[5]
16.E. bogotense isolate 40801MH750046Chile[5]
17.E. bogotense isolate 40802MH750047Ecuador[5]
18.E. bogotense isolate 40803MH750048Chile[5]
19.E. bogotense isolate 40827MH750049Colombia[5]
20.E. diffusumAY226126India[34]
21.E. diffusum isolate 40804MH750050India[5]
22.E. fluviatileAY226121Canada[34]
23.E. fluviatileGQ244922unknown[37]
24.E. fluviatile isolate 41075MH750051Finland[5]
25.E. fluviatile isolate 41076MH750052Finland[5]
26.E. giganteumAY226118Ecuador[34]
27.E. giganteum isolate 40805MH750053Jamaica[5]
28.E. giganteum isolate 40806MH750054Chile[5]
29.E. giganteum isolate 40807MH750055Peru[5]
30.E. giganteum isolate 40810MH750057Argentina[5]
31.E. giganteum isolate 40811MH750058Argentina[5]
32.E. hyemaleKC117177unknown[7]
33.E. hyemaleAY327837unknown[34]
34.E. hyemale isolate 0796gGQ244923unknown[37]
35.E. hyemale isolate 1273oGQ244924unknown[37]
36.E. hyemale isolate 20201MH750061France[5]
37.E. hyemale isolate 23252MH750062Norway[5]
38.E. hyemale isolate 41088MH750063Finland[5]
39.E. hyemale subsp. affineAY226110USA[34]
40.E. iganteum isolate 40809MH750056Argentina[5]
41.E. laevigatumAY226112USA[34]
42.E. laevigatum isolate 40812MH750065USA[5]
43.E. myriochaetumAY226114USA[34]
44.E. myriochaetum isolate 40815MH750066USA[5]
45.E. myriochaetum isolate 40816MH750067USA[5]
46.E. myriochaetum isolate 40825MH750068Mexico[5]
47.E. myriochaetum isolate 40826MH750069Ecuador[5]
48.E. myriochaetum isolate 40936MH750070El Savador[5]
49.E. myriochaetum isolate 41080MH750071Guatemala[5]
50.E. palustreAY226123Canada[34]
51.E. palustreGQ244925unknown[37]
52.E. palustre isolate 39349MH750072UK[5]
53.E. praealtum isolate 40831MH750059USA[5]
54.E. praealtum isolate 41501MH750060USA[5]
55.E. pratenseAY226122Canada[34]
56.E. pratenseGQ244926unknown[37]
57.E. pratenseHM590278Estonia[36]
58.E. pratense isolate 39348MH750073Finland[5]
59.E. pratense isolate 41086MH750074Finland[5]
60.E. pratense isolate 41087MH750075Finland[5]
61.E. ramosissimum subsp. debileAY226115Taiwan[34]
62.E. ramosissimum subsp. debile isolate 23679MH750076Reunion[5]
63.E. ramosissimum subsp. debile isolate 24579MH750077Sri Lanka[5]
64.E. ramosissimum subsp. debile isolate 40837MH750078New Caledonia[5]
65.E. ramosissimum subsp. ramosissimum isolate 36802MH750079Spain[5]
66.E. ramosissimum subsp. ramosissimum isolate 40829MH750080Turkey[5]
67.E. scirpoidesAY226116Canada[34]
68.E. scirpoidesGQ244927unknown[37]
69.E. scirpoides isolate 26090MH750082Greenland[5]
70.E. scirpoides isolate 40830MH750064Russia[5]
71.E. scirpoides isolate10933MH750081UK[5]
72.E. sylvaticumMH750083UK[5]
73.E. sylvaticumAY226120France[34]
74.E. sylvaticumGQ244928unknown[37]
75.E. sylvaticum isolate 41081MH750084Finland[5]
76.E. telmateia isolate 11642MH750089China[5]
77.E. telmateia isolate 41082MH750090Ireland[5]
78.E. telmateia subsp. brauniiAY226119USA[34]
79.E. telmateia subsp. braunii isolate 40817MH750085USA[5]
80.E. telmateia subsp. braunii isolate 40828MH750086Canada[5]
81.E. telmateia subsp. braunii isolate 40832MH750087USA[5]
82.E. telmateia subsp. braunii isolate 40836MH750088USA[5]
83.E. trachyodon isolate 41092MH750106Finland[5]
84.E. variegatumAY226117USA[34]
85.E. variegatum isolate 0584gGQ244929unknown[37]
86.E. variegatum isolate 0977oGQ244930unknown[37]
87.E. variegatum isolate 11639MH750091UK[5]
88.E. variegatum isolate 40819MH750098Ireland[5]
89.E. variegatum isolate 40820MH750092France[5]
90.E. variegatum isolate 40823MH750093USA[5]
91.E. variegatum isolate 41083MH750094Ireland[5]
92.E. variegatum subsp. alaskanum isolate 40818MH750095USA[5]
93.E. variegatum subsp. alaskanum isolate 40821MH750096Canada[5]
94.E. variegatum subsp. alaskanum isolate 40822MH750097Canada[5]
95.E. xylochaetumMW282958ChileThis study
96.E. xylochaetum isolate 40614MH750107Chile[5]
97.Equisetum sp.AY327838unknown[34]
98.Equisetum x dycei isolate 26083MH750099UK[5]
99.Equisetum x ferrissii (E. laevigatum x E. hyemale)AY226113USA[34]
100.Equisetum x ferrissii (Equisetum hyemale x laevigatum) AY226111Canada[34]
101.Equisetum x litorale isolate 41084MH750101Ireland[5]
102.Equisetum x litorale isolate 41085MH750102Ireland[5]
103.Equisetum x schaffneri isolate 40813MH750103Mexico[5]
104.Equisetum x schaffneri isolate 40814MH750104Peru[5]
105.Equisetum x schaffneri isolate 40824MH750105Mexico[5]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Satjarak, A.; Graham, L.E.; Trest, M.T.; Arancibia-Avila, P. Plastid Genome of Equisetum xylochaetum from the Atacama Desert, Chile and the Relationships of Equisetum Based on Frequently Used Plastid Genes and Network Analysis. Plants 2022, 11, 1001. https://doi.org/10.3390/plants11071001

AMA Style

Satjarak A, Graham LE, Trest MT, Arancibia-Avila P. Plastid Genome of Equisetum xylochaetum from the Atacama Desert, Chile and the Relationships of Equisetum Based on Frequently Used Plastid Genes and Network Analysis. Plants. 2022; 11(7):1001. https://doi.org/10.3390/plants11071001

Chicago/Turabian Style

Satjarak, Anchittha, Linda E. Graham, Marie T. Trest, and Patricia Arancibia-Avila. 2022. "Plastid Genome of Equisetum xylochaetum from the Atacama Desert, Chile and the Relationships of Equisetum Based on Frequently Used Plastid Genes and Network Analysis" Plants 11, no. 7: 1001. https://doi.org/10.3390/plants11071001

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