Phylogenic study of Lemnoideae (duckweeds) through complete chloroplast genomes for eight accessions

View article
PeerJ

Introduction

Three kinds of genomes with different evolutionary origins and histories coexist in plant cells: nuclear, chloroplastic and mitochondrial. Generally, mitochondrial genomes are not the best choice for phylogenetic studies in plants, because their rate of rearrangements is extraordinarily fast compared to chloroplast (cp) genomes (cpDNAs) (Palmer & Herbon, 1988). Meanwhile, phylogenetic studies using nuclear genomes are restricted by their complex and infeasibility of enough data (Olsen et al., 2016; Wang et al., 2014). A single, independent genealogical history can be readily obtained from cpDNAs (Kim et al., 2015; Whittall et al., 2010), since their inheritance differs from that of nuclear genomes, such as vegetative segregation, uniparental inheritance, haploid status, and general absence of recombination (Birky Jr, 2001; Hansen et al., 2007; Petit et al., 2003). Therefore, cpDNAs are particularly useful for phylogenetic and phylogeographic studies.

However, phylogenetic and phylogeographic analyses based on cpDNAs are typically limited by DNA sequencing costs and genomes assembling methods (Parks, Cronn & Liston, 2009). In previous studies, the cpDNAs were directly obtained by primer walking (based on closely related cpDNAs) and by shotgun sequencing (Jansen et al., 2005; Shi et al., 2012; Whittall et al., 2010). Recently, owing to the lower costs of next-generation sequencing (NGS), a new cost-effective method has arisen: indirectly assembling complete cpDNA by filtering from total DNA data (including DNA data of nuclei, cps and mitochondria) (Wang & Messing, 2011; Zhang et al., 2011). However, few studies had compared the chloroplast genomes obtained from this method with those obtained from primer walking or shotgun sequencing.

Lemnoideae (duckweeds), a kind of small aquatic flowering monocotyledonous plants, has got increasingly more attention due to its asexual reproduction, rapid propagation and potential values in eutrophic wastewater treatment, starch production and bioenergy transformation (Zhao et al., 2012). Lemnoideae is a subfamily of the Araceae family and includes five genera and 37 species (Sree, Bog & Appenroth, 2016). Traditionally, morphological characters are inspected for identifying the taxonomy of the species and phylogenetic studies. The phylogenetic relationship within the different genera of Lemnoideae has not been well resolved mainly because of their small sizes and highly morphological degeneration. In addition, the confidence values of phylogenetic trees were not strongly supported when using traditional markers (Rothwell et al., 2004). Given that neither morphological characters nor traditional markers are sufficient for phylogenetic study within Lemnoideae, we used the whole cpDNAs obtained from NGS to clarify the phylogenetic relationships within this group.

In this paper, we compared two different cpDNA extraction and assembly methods and verified that assembling the cpDNA based on the filtration of the total DNA was reliable. Then we built a phylogenetic tree on the basis of the whole cpDNA sequences which clarified the phylogenetic relationship within different genera of Lemnoideae. This study can help to resolve the phylogeny of Lemnoideae. Meanwhile, it highlights that the whole cpDNA is a feasible and effective option for phylogenetic studies, and demonstrates the possibilities that the NGS can elucidate those phylogenies which traditionally are not well solved.

Materials & Methods

Duckweeds strains

Eight duckweeds strains were used in this study. They were Landoltia punctata strain ZH0202, Landoltia punctata strain 0086, Landoltia punctata strain 0062, Lemna minor strain 9532, Lemna gibba strain 9584, Lemna japonica strain 0234, Lemna japonica strain 8695, and Wolffia australiana strain 7317, respectively. All the strains were stored and cultured in the Chengdu Institute of Biology, Chinese Academy of Sciences (Chengdu, China) under controlled conditions in the greenhouse. Four announced duckweeds cpDNAs data (Spirodela polyrhiza strain 7498, Lemna minor Renner 2188, Wolffiella lingulata strain 7289, and Wolffia australiana strain 7733) were also included in the study (Mardanov et al., 2008; Wang & Messing, 2011).

DNA extraction and sequencing

The cp DNA of L. punctata strain ZH0202 and L. japonica strain 8695 was isolated from 1 g of tissue of young duckweeds produced from a single mother, by using a Plant Cp DNA Isolation Kit (Genmed Scientific Inc., Arlington, USA). The DNA concentration and purity were checked with NanoDrop 2000c. Paired-end (PE) libraries with a 300-bp insert size were constructed and sequenced with Illumina HiSeq 2500 platform by the Beijing Genomics Institute (Shenzhen, China).

Pipeline of chloroplast genome assembly.

Figure 1: Pipeline of chloroplast genome assembly.

The total DNA of L. punctata strain ZH0202, L. punctata strain 0086, L. punctata strain 0062, L. minor strain 9532, L. gibba strain 9584, L. japonica strain 0234, and W. australiana strain 7317 was extracted by using the CTAB method (Jansen et al., 2005). The DNA concentration and purity were also checked with NanoDrop 2000c. PE libraries with 500-bp insert size were constructed and sequenced with Illumina HiSeq 2000 platform by the Beijing Genomics Institute (Shenzhen, China).

CpDNA assembly and annotation

When using the total DNA, we assembled the cpDNAs as follows (Fig. 1): (1) filtering of the data using FastQC 0.11.3 with the default parameters; (2) pre-assembly using SOAPdenovo 2 with diverse K-mer values (23–89) (Luo et al., 2012); (3) isolation of the cp contigs; (4) addition of the inverted repeats (IRs, including IRa and IRb) sequence; (5) extension of the contigs with SSPACE 2.02 (Boetzer et al., 2011); (6) closing of the gaps with GapCloser 1.12 (Kim et al., 2015); (7) mapping of the reads to the draft genome by using SOAP 2.21 to identify and correct any errors (Zhang et al., 2012). When isolating the cp contigs (step (3)), the best assembly result (K-mer 87) was first aligned with the most closely related reference genome (S. polyrhiza strain 7498, GenBank Accession: JN160603; L. minor, GenBank Accession: DQ400350; W. australiana strain 7733, GenBank Accession: JN160605) using BLAST to identify contigs with high identity (>95%, e-value 10−5) (McGinnis & Madden, 2004). Then, the read depth of those high-identity contigs was calculated using SOAP 2.21 and SOAPcoverage 2.7.7 (Li et al., 2009). The cp contigs with high coverage (more than 1,000×) were isolated. Subsequently, the contigs were aligned with the reference genome by using MAUVE 2.3.1 to ensure that they belonged to the cpDNA (Bennett & Triemer, 2015). Meanwhile, when using cp DNA, all the steps were applied to assemble the cpDNAs, except the isolation of cp contigs (step(3)).

CpDNAs were annotated using DOGMA with default parameters (Wyman, Jansen & Boore, 2004). The tRNA genes were further identified using tRNAscan-SE under default parameters (Sedlar et al., 2015). The single sequence repeats (SSRs) and tandem repeats in the cpDNAs of 12 strains of duckweeds were detected using Phobos 3.3.12 with the default parameters, except that the maximum unit length was set as 30 (Raman & Park, 2015). Map of the circular plastome was drawn with OGDraw 1.2 (Lohse et al., 2013).

Sequence polymorphisms of duckweeds cpDNAs

DnaSP v5 was used to calculate the DNA polymorphism (Librado & Rozas, 2009). To study the sequence divergence in the whole genome level, whole-genome alignments were carried out using mVISTA with “LAGAN” alignment program (Global multiple alignment of finished sequences) (Kim et al., 2015). We also compared the large single copy (LSC)/IRa/ small single copy (SSC) /IRb boundary regions of the duckweeds cpDNAs to study IRs contraction or expansion.

Phylogeny of Lemnoideae based on whole cpDNAs

ClustalW was used to align the cpDNAs sequences under default parameters (Larkin et al., 2007), and the alignment was checked manually. The Bayesian Inference (BI) and Maximum-likelihood (ML) methods were performed for the genome-wide phylogenetic analyses using MrBayes v.3.2.6 (Huelsenbeck & Ronquist, 2001; Ronquist & Huelsenbeck, 2003) and PhyML 3.0 (Guindon et al., 2010), respectively. Nucleotide substitution model selection was estimated with jModelTest 2.1.10 (Darriba et al., 2012) and Smart Model Selection in PhyML 3.0. The model GTR + G was selected as the best-fitting model for both BI and ML analyses. Bayesian Inference partitioning analysis followed the programs with calculating a majority rule consensus tree with 1 × 107 generations of Markov chain Monte Carlo (MCMC), with frequency of tree sampling every 1,000 generations and the first 2,500 trees discarding as burn-in, and starting from a random tree. After performing two independent runs, the output trees were combined to estimate the Bayesian posterior probabilities (BPP) in 50% majority rule for each node. For ML analysis, PhyML 3.0 was performed with 1,000 bootstrap replicates to calculate the bootstrap values (BS) of the topology. In addition, the significant nodes’ supports were considered with 95% BPP and 75% BS (Hillis & Bull, 1993) in BI and ML analysis, respectively. The results were treated with iTOL 3.4.3 (Letunic & Bork, 2016). Colocasia esculenta share the same family as the members of Lemnoideae and was included as an outgroup (Luo et al., 2016).

Results

Reliability of assembly of the cpDNA on the basis of the total DNA

To test the reliability of assembling the cpDNA based on the filtration of the total DNA, the cpDNA of the same strain (L. punctata strain ZH0202) was also assembled by using cp DNA directly. As a result, the identity of L. punctata strain ZH0202 cpDNAs assembled through the two methods was 100%, and the sequence and length were completely identical. This experiment revealed no nucleotide variability between the two cpDNA assemblies of L. punctata strain ZH0202 (Table 1).

Table 1:
Duckweed cp genomes assembly results.
Assembly method Sample name Strain Latin name CGS (bp) IRs (bp) LSC (bp) SSC (bp) GC content (%) Collection area GenBank accession number
ECD ZH0051 ZH0202 L. punctata 171,013 31,899 92,742 14,473 35.46 Xinjin, China KY993962
D0101 8695 L. japonica 166,424 31,571 89,277 14,005 35.74 Kyoto, Japan KY993955
FTD ZH0051 ZH0202 L. punctata 171,013 31,899 92,742 14,473 35.46 Xinjin, China KY993962
ZH0086 0086 L. punctata 170,994 31,900 92,721 14,473 35.49 Leshan, China KY993960
ZH0062 0062 L. punctata 171,152 31,894 92,726 14,635 35.44 Yaan, China KY993959
D0107 9532 L. minor 165,775 31,218 89,735 13,604 35.75 Ohrid, Macedonia KY993956
D0289 9584 L. gibba 166,553 31,763 89,408 13,619 35.73 Perebel River, Poland KY993957
ZH0234 0234 L. japonica 165,436 31,468 88,635 13,866 35.74 Kunming, China KY993961
M170 7317 W. australiana 168,270 31,990 90,871 13,419 35.86 Australia KY993958
DOI: 10.7717/peerj.4186/table-1

Notes:

CGS

Cp genome size

IRs

Inverted repeats

LSC

Large single copy

SSC

Small single copy

ECD

Extraction of the cp DNA

FTD

Filtering of the total DNA.

Assembly and annotation of Lemnoideae chloroplast genomes

The Illumina HiSeq system yielded 59 Gb of total clean data (Table S1). After the data were filtered, assembled and validated, eight complete duckweeds cpDNAs were obtained with coverage of over 1, 000 ×. Together with four reported duckweeds cpDNAs, all of the duckweeds cpDNAs were within a range of 165,775 bp to 171,152 bp in length (Table 1), and carried two copies of IRs separated by a SSC and LSC (Fig. 2). Their lengths were variable: IR (312,18 bp–31,990 bp); SSC (13,392 bp–14,635 bp) (Wang & Messing, 2011); and LSC (88,635 bp–92,742 bp) (Table 1, Table S2).

Chloroplast genome of Landoltia punctate strain ZH0202.

Figure 2: Chloroplast genome of Landoltia punctate strain ZH0202.

The outer circle shows positions of genes in the large single copy (LSC), small single copy (SSC), and two inverted repeat (IRa and IRb) regions. The inner circle is a graph depicting GC content across the genome. Plastome maps were generated in OGDraw 1.2.

Each of the duckweeds cpDNAs contained 80 protein-coding sequences (CDS), four rRNAs, 30 tRNAs and two pseudogenes, including two new annotated genes in duckweeds: orf42 and orf56 (Fig. 2, Table S3). All of the duckweeds cpDNAs had the same gene content and order.

We annotated the repeat sequences and compared their type, length and number among the cpDNAs of 12 strains of duckweeds (Table 2). The total length of repeat sequences was in a range of 10,004 bp (L. japonica strain 0234) to 12,832 bp (L. punctata strain 0062), and the percentages in cpDNAs were in a range of 6.03% (W. australiana strain 7733) to 7.59% (S. polyrhiza strain 7498) (Table 2). In all 12 cpDNAs of duckweeds, homopolymers were the most frequent, followed by hexa-, penta-, and tetrapolymers. However, there were relatively more pentapolymers (approximately 180) compared to hexapolymers (152) in L. punctata (Table 2). Interestingly, the longest tandem repeats (AAAAATATATAATAATATTAATAAAAT × 2) in the known duckweeds cpDNAs were found in L. japonica strain 0234, which had the shortest total length of repeat sequences and the smallest total number of repeat sequences among the duckweeds (Table 2).

Table 2:
The type, length and number of repeat sequence in the cp genomes of 12 strains of duckweed.
S. polyrhiz a strain 7498a L. punctata strain ZH0202 L. punctata strain 0086 L. punctata strain 0062 L. japonica strain 8695 L. minor strain 9532 L. minor Renner2188a L. gibba strain 9584 L. japonica strain 0234 W. lingulata strain 7289a W. australiana strain 7317 W. australiana strain 7733a
Repeats sequence (bp) 12,810 12,810 12,719 12,832 11,119 10,966 11,085 11,222 10,004 11,558 10,424 10,178
Percent(%) 7.59 7.49 7.44 7.50 6.68 6.61 6.68 6.74 6.05 6.83 6.19 6.03
Mono- 335 373 373 373 356 355 355 357 355 354 363 359
Di- 64 71 74 74 63 62 63 63 63 71 70 71
Tri- 83 70 69 69 69 69 69 69 65 74 59 60
Tetra- 123 121 121 121 106 104 105 105 101 112 115 117
Penta- 178 180 178 180 125 122 124 125 117 138 123 124
Hexa- 183 152 152 152 155 155 155 157 136 161 154 155
7 45 38 38 38 40 38 38 37 32 35 23 24
8 21 22 22 22 22 23 23 23 15 20 15 14
9 21 15 15 15 22 22 22 23 17 19 17 15
10 9 6 6 6 6 5 6 6 7 10 7 8
11 3 2 2 2 3 3 3 3 2 3 1 1
12 1 2 2 2 4 3 4 4 1 2 0 0
13 1 7 6 7 0 0 0 1 1 1 1 1
14 1 4 4 4 2 3 3 3 1 2 0 0
15 2 4 3 3 2 1 1 1 2 3 1 1
16 0 1 1 1 3 3 3 3 0 0 0 0
17 1 3 3 3 1 1 1 1 0 0 0 0
18 1 1 1 1 1 2 2 2 1 0 0 0
19 3 1 1 1 1 1 1 1 3 0 1 0
20 0 2 2 2 1 1 1 1 0 0 0 0
21 0 0 1 1 0 0 0 0 1 1 0 0
22 2 2 2 2 0 0 0 0 0 0 0 0
23 1 1 1 1 1 1 1 1 0 1 0 0
24 0 4 4 4 1 1 1 1 0 2 2 0
25 1 1 1 1 1 1 1 0 0 0 1 0
26 0 0 0 0 0 0 0 1 1 0 1 0
27 0 0 0 0 0 0 0 0 1 0 0 0
DOI: 10.7717/peerj.4186/table-2

Notes:

Duckweed cp genomes from previous studies.

Sequence polymorphisms of duckweeds cpDNAs

A total of 17,438 polymorphic sites were found among duckweeds cpDNAs by using ClustalW and DnaSP. Most of the divergent sequences were in intergenic regions, whereas some were in introns (Fig. S1). For instance, L. punctata strain ZH0202 has a 564-bp insertion in the 32-kb petN-psbM region and an 88-bp insertion in the intron of atpF. Five strains of Lemna all had approximately 300-bp deletions at the 105-kb regions (Figs. S1; S2). In comparison to intergenic regions, the sequence divergence frequency for regions of coding genes was low. IRs were more conserved than LSC and SSC. Although there was less sequence divergence in IRs, more sequence polymorphisms appeared at the junctions of IRs and LSC/SSC (Fig. S1). We furthermore found that the locations of some genes in the LSC/IRa/SSC/IRb boundary regions were different (Fig. 3), although the gene content and order were the same. The most comprehensive variation was found in the boundary of the LSC and IRa regions, where the rps19 sequence completely shifted position towards the LSC region in L. minor and S. polyrhiza. Additionally, a 385 bp of the rpl2 sequence relocated from the IRa toward the LSC region. Furthermore, the rpl2 gene at the end of the IRb region was incomplete in L. minor, thus annotated as a pseudogene (Fig. 3).

Boundary gene-flow and IR region expansion/contraction events.

Figure 3: Boundary gene-flow and IR region expansion/contraction events.

Comparison of the junction positions of IR boundaries among 7 duckweed cp genomes. The genes near the junction positions are identified by color: red, rps19; light blue, rpl2; yellow, rps15; light green, ndhF; green, ndhH; purple, trnH-GUG.

Phylogeny of Lemnoideae based on whole cpDNAs

A phylogenetic tree was generated using BI and ML methods, which consistently supported the uniform topology. The topology was reliable with 1.00 BPP and 100% BS for nine out of ten nodes (Fig. 4). The phylogenetic analysis demonstrated that Spirodela was derived first from the lineage of the remaining members of the subfamily, then Landoltia, Lemna, Wolffiella and Wolffia (Fig. 4). Surprisingly, we found that L. japonica strain 8695 and L. japonica strain 0234 were in separate branches.

The Bayesian Inference (BI) and Maximum likelihood (ML) tree on the basis of cp genome sequences.

Figure 4: The Bayesian Inference (BI) and Maximum likelihood (ML) tree on the basis of cp genome sequences.

The Numbers upon each node indicate Bayesian posterior probabilities and ML Bootstrap, respectively (showed in BPP/BS).

Discussion

Reliability of assembling the cpDNA based on the filtration of the total DNA

For previous studies, the cpDNAs were assembled from independently extracted cp DNA (Jansen et al., 2005; Shinozaki et al., 1986). Assembling the cpDNA based on indirect filtration of the total DNA was first described in 2011 (Zhang et al., 2011). To our knowledge, no study had compared these two assembly approaches until now. In this study, we assembled the cpDNA of L. punctata strain ZH0202 with the two methods simultaneously. The sequence and length of L. punctata strain ZH0202 cpDNAs assembled through the two methods were completely identical (Table 1). This result verified that indirectly assembling the cpDNA based on the filtration of the total DNA is reliable. Directly assembling cpDNAs from cp still has advantages especially when the reference genomes are unavailable (Jansen et al., 2005). Meanwhile, the sequencing costs decrease every year, indirectly assembling cpDNAs from the total DNA becomes more and more attractive (Wang & Messing, 2011).

Two genes in duckweeds cpDNAs: orf42 and orf56

Here, two genes were annotated: orf 42 and orf 56, which were not found in previous studies of duckweeds cpDNAs (Mardanov et al., 2008; Wang & Messing, 2011). These two genes are located 200 bp apart in the intron of trnA-UGC. Their sequences are conserved, and it has previously been reported that they are related to mitochondrial genes (Do, Kim & Kim, 2013). However, functions of orf 42 and orf 56 are still unknown (Bodin, Kim & Kim, 2013).

Differences in cpDNA sizes among the genera of Lemnoideae

The known cpDNA sizes are conservative within some subfamilies of higher plants. For instance, all the members of the Maloideae have a similar cpDNA size, ranging from 15,9161 bp in Pydus spinosa to 16,0041 bp in Malus pdunifolia voucher MPRUN20160302 (Korotkova et al., 2014). However, the cpDNA sizes are more variable within Lemnoideae, ranging from 166 kb in Lemna to 171 kb in Landoltia. In the former (Lemna), the IR size was the smallest among the known duckweeds cpDNAs, indicating the IRs have contracted in this species (Fig. 3). While in the latter (Landoltia), it was found that the lengths of the IRs, LSC and SSC regions were longer than those of the other genera except IRs of Wolffia. Our results contrast with a previous study carried out on species of Gossypium genus, in which most of the cp size differences reflected indels in the LSC region (Chen et al., 2017). The results of this study supported the previous finding that changes in the length of the IR can account for the size variation among plant cpDNAs (Jansen et al., 2005; Wakasugi, Tsudzuki & Sugiura, 2001).

We furthermore found the differences in the cpDNA sizes among the genera of Lemnoideae resulted primarily from variation in the non-coding regions, while the lengths of coding regions were almost the same. The results of the whole cpDNA alignments also supported this (Fig. S1). This finding was consistent with Zheng’s study in seed plants, in which their results pointing to intergenic regions having a great role in the variations in chloroplast genome size among closely related species (Zheng et al., 2017).

In addition, variation in the repeat sequences were found to partially account for the difference in cpDNA sizes. For example, Landoltia had 12.8 kb of repeat sequences, which were approximately 2.2 kb longer than those of Lemna. Moreover, 5.0 kb of sequence in L. minor strain 9532 did not exist in the cpDNA of L. punctata strain ZH0202, 25.57% of which comprised repeat sequences. This percentage was greater than the average percentage (7.49%) of repeat sequences in L. punctata strain ZH0202. Similar as above, 1.9 kb of sequence in W. australiana strain 7317 did not exist in the cpDNA of Wolffiella lingulata strain 7289, 38.10% of which consisted of repeat sequences. This percentage was greater than the average percentage (6.83%) of repeat sequences in W. lingulata strain 7289. There were more repeat sequences in the expanded portion, meaning that repeat sequences influenced the expansion or contraction of the cpDNA and partially led to the difference in cpDNA sizes. Our results were consistent with Wu’s study in rice, which indicates the main source of cp length variation is coming from mononucleotide SSRs (Wu et al., 2017). The variation in cpDNA size may influence energy generation and ecological strategy (Zheng et al., 2017) and provide more information for phylogenic study.

Phylogeny of lemnoideae

It was surprising that L. japonica strain 8695 and L. japonica strain 0234 are in separate branches (Fig. 4). However, similar cases were found in the phylogenetic study of apples: Malus sieversii were scattered across branches containing other wild species, which justified its splitting into at least two species (Nikiforova et al., 2013). In our study, L. japonica strain 8695 and L. japonica strain 0234 were scattered across branches containing L.minor, indicating that genetic diversity of cpDNAs within L. japonica exceeds that between other species. In addition, they have phenotypic differences: L. japonica strain 0234 cultured in Kunming have air spaces on the back of the fronds, whereas the air spaces were not found on the fronds of L. japonica strain 8695 (Fig. S3). Moreover, their collection areas are far apart (Table 1). Considering all these evidences, we think the most probable reasons of why two strains of L. japonica are in separate branches may be a wrong identification or they are two different species. Further research would be necessary to elucidate the cause of why these two strains are present in separate branches.

On the basis of the duckweeds cpDNAs, we confirmed that the evolutionary branching order of Lemnoideae was as follows: Spirodela, Landoltia, Lemna, Wolffiella, Wolffia (Fig. 4). The evolutionary divergence of Landoltia came after that of Spirodela and before that of Lemna. These results were consistent with a previous finding by Les et al. (2002). In this study, authors applied more than 4,700 characters including data on morphology and anatomy, flavonoids, allozymes, and DNA sequences from cp genes (rbcL, matK) and introns (trnK, rpl16). In our study, higher bootstrap values were obtained that support the evolutionary order of Lemnoideae (Fig. 4). However, our results were different from those of Rothwell et al. (2004). In this study, authors applied the cp trnL - trn F intergenic spacer and found the evolutionary order of Lemnoideae was Spirodela, Landoltia, Wolffia, Wolffiella, and Lemna. Our results were also different from Lidia I. Cabrera’s study. In this study, authors applied coding regions (rbcL, matK) and non-coding plastid DNA (partial trnK intron, trnL intron, trnL - trnF spacer), and found the evolutionary order of Lemnoideae was Spirodela, Lemna, Landoltia, Wolffia, and Wolffiella (Cabrera et al., 2008). It was suggested that the evolutionary studies based on whole cpDNAs were more reliable than those based on several cp coding and/or non-coding regions, because the whole cpDNAs contain more rich genetic information (Hansen et al., 2007; Sree, Bog & Appenroth, 2016). Our finding supported Matthew Parks’ study (Parks, Cronn & Liston, 2009), in which they compared the resolution obtained when using the cpDNAs and two cp markers, and found the increase in phylogenetic resolution is primarily due to the increase in data matrix length. Therefore, our results indicated that the whole cpDNA is a feasible and effective option for phylogenetic studies, especially for inferring phylogenies at low taxonomic levels (Parks, Cronn & Liston, 2009; Whittall et al., 2010).

Conclusion

We indicated that assembly of the cp genome based on the filtration of the total DNA was reliable. Our study suggested that the whole cpDNA was appropriate for phylogenetic studies, especially for inferring phylogenies at low taxonomic levels, and it showed the possibilities that the NGS can offer to elucidate those phylogenies that traditionally have not been well solved. In this study, we demonstrated the evolutionary order of Lemnoideae was as follows: Spirodela, Landoltia, Lemna, Wolffiella, Wolffia.

Supplemental Information

The cp genomes of Lemnoideae subfamily alignment using mVISTA

DOI: 10.7717/peerj.4186/supp-1

The cp genomes of Lemna genera alignment using mVISTA

DOI: 10.7717/peerj.4186/supp-2

Lemna japonica strain 8695 and Lemna japonica strain 0234

Lemna japonica strain 8695: (A) The back of the fronds; (B) The front of the fronds; Lemna japonica strain 0234: (C, D) The back of the fronds; (E) The front of the fronds.

DOI: 10.7717/peerj.4186/supp-3

Statistics of clean data

DOI: 10.7717/peerj.4186/supp-4

Reported duckweed chloroplast genomes

CGS, Cp genome size; IRs, Inverted repeats; LSC, Large single copy; SSC, Small single copy.

DOI: 10.7717/peerj.4186/supp-5

Genes of Landoltia punctata ZH0202 chloroplast genome (116 genes)

Note: genes with one or two intron are noted with “*” or “**”, genes in IR are noted with “x2”.

DOI: 10.7717/peerj.4186/supp-6

Clustalw alignment result

DOI: 10.7717/peerj.4186/supp-8
13 Citations   Views   Downloads