ArticlePDF Available

Complete Plastid Genome Sequencing of Eight Species from Hansenia, Haplosphaera and Sinodielsia (Apiaceae): Comparative Analyses and Phylogenetic Implications

Authors:

Abstract and Figures

Hansenia Turcz., Haplosphaera Hand.-Mazz. and Sinodielsia H.Wolff are three Apiaceae genera endemic to the Hengduan Mountains and the Himalayas, which usually inhabit elevations greater than 2000 m. The phylogenetic relationships between and within the genera were uncertain, especially the placement of Hap. himalayensis and S. microloba. Therefore, we aimed to conduct comparative (simple sequence repeat (SSR) structure, codon usage bias, nucleotide diversity (Pi) and inverted repeat (IR) boundaries) and phylogenetic analyses of Hansenia, Haplosphaera and Sinodielsia (also compared with Chamaesium and Bupleurum) to reduce uncertainties in intergeneric and interspecific relationships. We newly assembled eight plastid genomes from Hansenia, Haplosphaera and Sinodielsia species, and analyzed them with two plastid genomes from GenBank of Hap. phaea,S. yunnanensis. Phylogenetic analyses used these ten genomes and another 22 plastid genome sequences of Apiaceae. We found that the newly assembled eight genomes ranged from 155,435 bp to 157,797 bp in length and all had a typical quadripartite structure. Fifty-five to 75 SSRs were found in Hansenia, Haplosphaera and Sinodielsia species, and the most abundant SSR was mononucleotide, which accounted for 58.47% of Hansenia, 60.21% of Haplosphaera and 48.01% of Sinodielsia. There was no evident divergence of codon usage frequency between the three genera, where codons ranged from 21,134 to 21,254. The Pi analysis showed that trnE(UUC)-trnT(GGU), trnH(GUG)-psbA and trnE(UUC)-trnT(GGU) spacer regions had the highest Pi values in the plastid genomes of Hansenia (0.01889), Haplosphaera (0.04333) and Sinodielsia (0.01222), respectively. The ndhG-ndhI spacer regions were found in all three genera to have higher diversity values (Pi values: 0.01028–0.2), and thus may provide potential DNA barcodes in phylogenetic analysis. IR boundary analysis showed that the length of rps19 and ycf1 genes entering IRs were usually stable in the same genus. Our phylogenetic tree demonstrated that Hap. himalayensis is sister to Han. weberbaueriana; meanwhile, Haplosphaera and Hansenia are nested together in the East Asia clade, and S. microloba is nested within individuals of S. yunnanensis in the Acronema clade. This study will enrich the complete plastid genome dataset of the Apiaceae genera and has provided a new insight into phylogeny reconstruction using complete plastid genomes of Hansenia, Haplosphaera and Sinodielsia.
Content may be subject to copyright.
plants
Article
Complete Plastid Genome Sequencing of Eight
Species from Hansenia,Haplosphaera and Sinodielsia
(Apiaceae): Comparative Analyses and
Phylogenetic Implications
Wei Gou 1,, Sheng-Bin Jia 1,, Megan Price 2, Xian-Lin Guo 1, Song-Dong Zhou 1
and Xing-Jin He 1, *
1Key Laboratory of Bio-Resources and Eco-Environment of Ministry of Education, College of Life Sciences,
Sichuan University, Chengdu 610065, China; gouwei1@stu.scu.edu.cn (W.G.);
sdjiashengbin@gmail.com (S.-B.J.); xlguo@stu.scu.edu.cn (X.-L.G.); zsd@scu.edu.cn (S.-D.Z.)
2Sichuan Key Laboratory of Conservation Biology on Endangered Wildlife, College of Life Sciences,
Sichuan University, Chengdu 610065, China; meganprice@scu.edu.cn
*Correspondence: xjhe@scu.edu.cn
Equal contributions to this work.
Received: 12 October 2020; Accepted: 6 November 2020; Published: 9 November 2020


Abstract:
Hansenia Turcz., Haplosphaera Hand.-Mazz. and Sinodielsia H.Wolare three Apiaceae
genera endemic to the Hengduan Mountains and the Himalayas, which usually inhabit elevations
greater than 2000 m. The phylogenetic relationships between and within the genera were uncertain,
especially the placement of Hap. himalayensis and S. microloba. Therefore, we aimed to conduct
comparative (simple sequence repeat (SSR) structure, codon usage bias, nucleotide diversity (Pi)
and inverted repeat (IR) boundaries) and phylogenetic analyses of Hansenia,Haplosphaera and
Sinodielsia (also compared with Chamaesium and Bupleurum) to reduce uncertainties in intergeneric and
interspecific relationships. We newly assembled eight plastid genomes from Hansenia,Haplosphaera
and Sinodielsia species, and analyzed them with two plastid genomes from GenBank of Hap. phaea,
S. yunnanensis. Phylogenetic analyses used these ten genomes and another 22 plastid genome
sequences of Apiaceae. We found that the newly assembled eight genomes ranged from 155,435 bp to
157,797 bp in length and all had a typical quadripartite structure. Fifty-five to 75 SSRs were found
in Hansenia,Haplosphaera and Sinodielsia species, and the most abundant SSR was mononucleotide,
which accounted for 58.47% of Hansenia, 60.21% of Haplosphaera and 48.01% of Sinodielsia. There was
no evident divergence of codon usage frequency between the three genera, where codons ranged
from 21,134 to 21,254. The Pi analysis showed that trnE(UUC)-trnT(GGU),trnH(GUG)-psbA and
trnE(UUC)-trnT(GGU) spacer regions had the highest Pi values in the plastid genomes of Hansenia
(0.01889), Haplosphaera (0.04333) and Sinodielsia (0.01222), respectively. The ndhG-ndhI spacer regions
were found in all three genera to have higher diversity values (Pi values: 0.01028–0.2), and thus may
provide potential DNA barcodes in phylogenetic analysis. IR boundary analysis showed that the
length of rps19 and ycf1 genes entering IRs were usually stable in the same genus. Our phylogenetic
tree demonstrated that Hap. himalayensis is sister to Han. weberbaueriana; meanwhile, Haplosphaera
and Hansenia are nested together in the East Asia clade, and S. microloba is nested within individuals
of S. yunnanensis in the Acronema clade. This study will enrich the complete plastid genome dataset of
the Apiaceae genera and has provided a new insight into phylogeny reconstruction using complete
plastid genomes of Hansenia,Haplosphaera and Sinodielsia.
Keywords: Apiaceae; Hansenia;Haplosphaera; phylogeny; plastid genome; Sinodielsia
Plants 2020,9, 1523; doi:10.3390/plants9111523 www.mdpi.com/journal/plants
Plants 2020,9, 1523 2 of 17
1. Introduction
Hansenia Turcz., Haplosphaera Hand.-Mazz. and Sinodielsia H.Wolare three endemic high-elevation
(typically >2000 m) genera of Apiaceae, mainly distributed in the Hengduan Mountains and the
Himalayas (Figure 1) [
1
]. According to the latest Apiaceae taxonomy, Hansenia,Haplosphaera and
Sinodielsia comprise five, two and four species, respectively [
2
5
]. Previous phylogenetic studies of
Hansenia,Haplosphaera and Sinodielsia based on two plastid genome regions (rpl16 and rps16 introns)
and nuclear internal transcribed spacers (nrITSs) found that Sinodielsia is within the Acronema clade,
while the closely related Hansenia and Haplosphaera are located in the East Asia clade, albeit from
limited sampling [
6
11
]. One study of 106 nrITS sequences representing 100 species from 52 genera
of Chinese Apiaceae found that the Acronema clade and the East Asia clade were well-supported
(posterior probability both valued 100% by Bayesian inference) [
8
]. Additionally, Hap. himalayensis and
S. microloba were unknown at the time of these previous studies and remain as two little-known species.
Plants 2020, 9, x FOR PEER REVIEW 2 of 19
1. Introduction
Hansenia Turcz., Haplosphaera Hand.-Mazz. and Sinodielsia H.Wolff are three endemic high-
elevation (typically >2000 m) genera of Apiaceae, mainly distributed in the Hengduan Mountains
and the Himalayas (Figure 1) [1]. According to the latest Apiaceae taxonomy, Hansenia, Haplosphaera
and Sinodielsia comprise five, two and four species, respectively [2–5]. Previous phylogenetic studies
of Hansenia, Haplosphaera and Sinodielsia based on two plastid genome regions (rpl16 and rps16
introns) and nuclear internal transcribed spacers (nrITSs) found that Sinodielsia is within the Acronema
clade, while the closely related Hansenia and Haplosphaera are located in the East Asia clade,
albeit fr
Figure 1. Plants of Hansenia, Haplosphaera and Sinodielsia. (A) Han. forbesii, (B) Han. forrestii, (C) Han.
oviformis, (D) Han. weberbaueriana, (E) Hap. himalayensis, (F) Hap. phaea, (G) S. microloba and (H) S.
yunnanensis.
With the development of second-generation sequencing technology, more plastid genomes have
been used in phylogeny and comparative studies, and fairly good results have been obtained [12].
Generally the circular genome consists of two inverted repeats (IRs) divided by two regions, the large
(LSC) and small single-copy (SSC) regions [13,14], and most angiosperm complete plastid genomes
are between 115 and 165 kb in length [15]. The gene content and order of plastid genomes are usually
highly conserved, and the substitution rate in plastid DNA is much lower than in plant nuclear DNA
[16]. The similarity of gene length and the low substitution rate of plant plastids make them valuable
sources of genetic markers for phylogenetic studies [17].
In our previous studies, we sequenced two plastid genomes of Hap. phaea [18] and S. yunnanensis
(HB: Zhongdian population) [19] and provided preliminary phylogenetic positions for the two
species. Following on from this, we aimed to conduct comparative (simple sequence repeat (SSR),
Figure 1.
Plants of Hansenia,Haplosphaera and Sinodielsia. (
A
)Han. forbesii, (
B
)Han. forrestii,
(
C
)Han. oviformis, (
D
)Han. weberbaueriana, (
E
)Hap. himalayensis, (
F
)Hap. phaea, (
G
)S. microloba and
(H)S. yunnanensis.
With the development of second-generation sequencing technology, more plastid genomes have
been used in phylogeny and comparative studies, and fairly good results have been obtained [
12
].
Generally the circular genome consists of two inverted repeats (IRs) divided by two regions, the large
(LSC) and small single-copy (SSC) regions [
13
,
14
], and most angiosperm complete plastid genomes are
between 115 and 165 kb in length [
15
]. The gene content and order of plastid genomes are usually
highly conserved, and the substitution rate in plastid DNA is much lower than in plant nuclear
DNA [
16
]. The similarity of gene length and the low substitution rate of plant plastids make them
valuable sources of genetic markers for phylogenetic studies [17].
In our previous studies, we sequenced two plastid genomes of Hap. phaea [
18
] and S. yunnanensis
(HB: Zhongdian population) [
19
] and provided preliminary phylogenetic positions for the two species.
Following on from this, we aimed to conduct comparative (simple sequence repeat (SSR), codon
Plants 2020,9, 1523 3 of 17
usage bias, nucleotide diversity (Pi) and IR) and phylogenetic analyses of Hansenia,Haplosphaera
and Sinodielsia to reduce uncertainties in intergeneric and interspecific relationships. Analyses were
conducted using a 32 complete plastid genome dataset, which was compiled from eight newly
assembled plastid genomes from Han. forbesii,Han. forrestii,Han. oviformis,Han. weberbaueriana,
Hap. himalayensis,S. microloba,S. yunnanensis (EY: Eryuan County pop.) and S. yunnanensis (KM:
Kunming pop.), as well as previously sequenced genomes. Our study provides newly complete plastid
genomes of Hansenia,Haplosphaera and Sinodielsia species, and since these three genera are endemic
to Pan-Himalayan regions, the information provided herein is indispensable for Apiaceae plastid
evolutionary and phylogenetic studies.
2. Results and Discussion
2.1. Phylogenetic Analysis
The phylogenetic tree (Figure 2) showed that Hansenia and Haplosphaera form a strongly supported
monophyly (Maximum Likelihood-Bootstrap Support (ML-BS) =100%) in the East Asia clade.
Hap. himalayensis is sister to Han. weberbaueriana (ML-BS =67%), and Hap. phaea, the type species of
Haplosphaera, is sister to Han. forrestii (ML-BS =100%), showing that Haplosphaera is nested within
Hansenia. This suggests that Hansenia and Haplosphaera should be combined to a single genus, as already
proposed in previous phylogenetic studies. [8,9,20].
Plants 2020, 9, x FOR PEER REVIEW 3 of 19
codon usage bias, nucleotide diversity (Pi) and IR) and phylogenetic analyses of Hansenia,
Haplosphaera and Sinodielsia to reduce uncertainties in intergeneric and interspecific relationships.
Analyses were conducted using a 32 complete plastid genome dataset, which was compiled from
eight newly assembled plastid genomes from Han. forbesii, Han. forrestii, Han. oviformis, Han.
weberbaueriana, Hap. himalayensis, S. microloba, S. yunnanensis (EY: Eryuan County pop.) and S.
yunnanensis (KM: Kunming pop.), as well as previously sequenced genomes. Our study provides
newly complete plastid genomes of Hansenia, Haplosphaera and Sinodielsia species, and since these
three genera are endemic to Pan-Himalayan regions, the information provided herein is
indispensable for Apiaceae plastid evolutionary and phylogenetic studies.
2. Results and Discussion
2.1. Phylogenetic Analysis
The phylogenetic tree (Figure 2) showed that Hansenia and Haplosphaera form a strongly
supported monophyly (Maximum Likelihood-Bootstrap Support (ML-BS) = 100%) in the East Asia
clade. Hap. himalayensis is sister to Han. weberbaueriana (ML-BS = 67%), and Hap. phaea, the type species
of Haplosphaera, is sister to Han. forrestii (ML-BS = 100%), showing that Haplosphaera is nested within
Hansenia. This suggests that Hansenia and Haplosphaera should be combined to a single genus, as
already proposed in previous phylogenetic studies. [8,9,20].
Figure 2. Maximum Likelihood (ML) phylogenetic tree of Apiaceae using the 32 plastid genomes
dataset. The studied taxa are bold. The numbers above the nodes are Maximum Likelihood-Bootstrap
Figure 2.
Maximum Likelihood (ML) phylogenetic tree of Apiaceae using the 32 plastid genomes
dataset. The studied taxa are bold. The numbers above the nodes are Maximum Likelihood-Bootstrap
Support (ML-BS) presented as percentages (>50%). The names of the clades follow the study of
Zhou et al. [8,9].
Plants 2020,9, 1523 4 of 17
The tree also showed that S. microloba is sister to S. yunnanensis (HB) (ML-BS =100%) and allied with
other S. yunnanensis (EY and KM) (ML-BS =100%) populations instead of other species in the Acronema
clade. The extremely close distance between S. microloba and S. yunnanensis is unusual compared
to other Apiaceae species, even though they are congeneric [
21
23
]. However, their morphological
characters are distinctive (Figure 1). We speculate there may have been a hybridization phenomenon
(i.e., recent or ongoing) between the two species or other intrageneric species, which have to also use
nuclear markers to disentangle the true phylogeny of the species. Otherwise, the observed closeness
may be caused by incomplete lineage sorting. More populations of S. microloba and other Sinodielsia
species are needed to explore their relationship and evolution.
2.2. The Plastid Genomes of Hansenia, Haplosphaera and Sinodielsia Species
The complete plastid genomes of Hansenia, Haplosphaera and Sinodielsia species exhibited a typical
quadripartite organization of a single circular DNA molecule (Figure 3). The lengths of the ten
genomes (four Hansenia, two Haplosphaera and four Sinodielsia species, including the additional two
populations of S. yunnanensis) ranged from 154,670 bp (S. yunnanensis: HB) to 157,797 bp (Han. forbesii)
(Table 1). Two identical IRs (including IRa and IRb, with lengths 26,404–26,542 bp) were found in the
plastid genomes, which were separated by LSC (85,233–86,968 bp) and SSC (17,370–17,891 bp) regions.
The quadripartite organization was found in most plastid genomes of higher plants [
13
,
14
], while IRs
were absent in Taxus chinensis var. mairei,Erodium species, Pisum sativum and Vicia faba [
21
,
24
,
25
],
which may lead to a reduction in the number of duplicated genes and the length of whole plastid
genomes. In Apiaceae, the IRa and IRb were both present in the studied plastid genomes from
GenBank, Bupleurum species [
26
] and Chamaesium species [
27
] ranging from 26,280–26,303 bp and
25,727–26,147 bp, respectively, and the dierence of length is mainly caused by the loss or insertion of
the spacer regions. The Hansenia,Haplosphaera and Sinodielsia plastid genomes had almost identical GC
content (37.5–37.7%) to whole plastid genomes. Higher GC content (42.7–42.8%) was detected in the
IR regions compared to the average of a whole genome, which was possibly due to the presence of
rRNA sequences with high GC content (55.2–55.3%) in IR regions. Similarly, the high GC content of
rRNA sequences also occurs in other Apiaceae species [
26
,
27
]. The ten new plastid genomes contained
133 genes, including eight rRNA, 37 tRNA and 85 protein-coding genes (PCGs) (Table 2). Among these
133 genes, 95 genes only had one copy, while 19 genes were duplicated in the IRa and IRb regions,
including four rRNA genes (rrn4.5,rrn5,rrn16 and rrn23), seven tRNA genes (trnA-UGC,trnI-CAU,
trnI-GAU,trnL-CAA,trnN-GUU,trnR-ACG and trnV-GAC) and eight PCGs (ndhB,rpl2,rpl23,rps7,
rps12,rps19,ycf1 and ycf2). Four pseudogenes—
ψ
rps19,
ψ
ycf1 and two
ψ
ycf15—were found in all the
ten genomes. In comparison, Hansenia, Haplosphaera, Sinodielsia,Bupleurum [
26
] and Chamaesium [
27
]
plastid genes are consistent in their total number of predicted coding regions.
The PCGs in the Hansenia,Haplosphaera and Sinodielsia plastid genomes included five genes (psaA,
psaB,psaC,psaI and psaJ) encoding photosystem I subunits, while 15 genes (psbA,psbB,psbC,psbD,psbE,
psbF,psbH,psbI,psbJ,psbK,psbL,psbM,psbN,psbT and psbZ) were related to photosystem II subunits.
Nine (rpl2,rpl14,rpl16,rpl20,rpl22,rpl23,rpl32,rpl33 and rpl36) encoding large ribosomal protein
genes and 12 (rps2,rps3,rps4,rps7,rps8,rps11,rps12,rps14,rps15,rps16,rps18 and rps19) encoding
small ribosomal protein genes were detected. Additionally, six genes (atpA,atpB,atpE,atpF,atpH and
atpI) of ATP synthase subunits were detected. These PCGs were also detected in Bupleurum [
26
] and
Chamaesium species [27] and are generally involved in the important process of plant growth.
Plants 2020,9, 1523 5 of 17
Plants 2020, 9, x FOR PEER REVIEW 5 of 19
Figure 3. Plastid genome map of eight Hansenia, Haplosphaera and Sinodielsia species (for a better view,
these eight maps were combined into one derived from the map of H. forbesii because they have the
same order and composition of genes). The genes shown inside and outside of the circle indicate those
transcribed in the clockwise and counterclockwise direction, respectively. Genes of different
functional groups are colored differently. The GC contents are shown in the inner circle with darker
grey.
Figure 3.
Plastid genome map of eight Hansenia,Haplosphaera and Sinodielsia species (for a better view,
these eight maps were combined into one derived from the map of H. forbesii because they have the
same order and composition of genes). The genes shown inside and outside of the circle indicate those
transcribed in the clockwise and counterclockwise direction, respectively. Genes of dierent functional
groups are colored dierently. The GC contents are shown in the inner circle with darker grey.
Plants 2020,9, 1523 7 of 17
Table 2. List of genes encoded in the ten Hansenia,Haplosphaera and Sinodielsia plastid genomes.
Category Group of Genes Name of Genes
Self-replication
transfer RNAs (tRNAs)
trnA-UGC *, trnC-GCA, trnD-GUC, trnE-UUC, trnF-GAA, trnfM-CAU,
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
ribosomal RNAs
(rRNAs) rrn4.5 *, rrna5 *, rrn16 *, rrn23 *
RNA polymerase rpoA, rpoB, rpoC1, rpoC2
Small subunit of
ribosomal proteins
(SSU)
rps2, rps3, rps4, rps7 *, rps8, rps11, rps12, rps14, rps15, rps16, rps18, rps19 *
Large subunit of
ribosomal proteins
(LSU)
rpl2 *, rpl14, rpl16, rpl20, rpl22, rpl23 *, rpl32, rpl33, rpl36
Genes involved in
photosynthesis
Subunits of
NADH-dehydrogenase
ndhA, ndhB *, ndhC, ndhD, ndhE, ndhF, ndhG, ndhH, ndhI, ndhJ, ndhK
Subunits in
photosystem I psaA, psaB, psaC, psaI, psaJ
Subunits in
photosystem II
psbA, psbB, psbC, psbD, psbE, psbF, psbH, psbI, psbJ,psbK, psbL, psbM, psbN,
psbT, psbZ
Subunits of cytochrome
b/f complex petA, petB, petD, petG, petL, petN
Subunits of ATP
synthase atpA, atpB, atpE, atpF, atpH, atpI
Large subunit of
rubisco rbcL
Other genes
Translational initiation
factor infA
Protease clpP
Maturase matK
Subunit of
Acetyl-CoA-carboxylase
accD
Envelope membrane
protein cemA
C-type cytochrome
synthesis gene ccsA
Conserved reading
frames
Conserved open
reading frames ψrps19, ycf1 * (ycf1, ψycf1), ycf2 *, ycf3, ycf4, ψycf15
* Duplicated genes.
2.3. IR Boundaries and Simple Sequence Repeats (SSRs) Structure Analysis
The IR boundaries of the ten Hansenia,Haplosphaera and Sinodielsia plastid genomes were compared
to analyze the fluctuations (expansion or contraction) in these regions (Figure 4). Although the ten
plastid genomes showed a similar structure and content, some variations were still identified. The rps19
gene entered the IRb region with 46 bp, 46 bp, 46 bp, 41 bp, 54 bp and 46 bp in the plastid genomes
of Han. forbesii,Han. forrestii,Han. oviformis,Han. weberbaueriana,Hap. himalayensis and Hap. phaea
(respectively), while 102 bp, 102 bp, 102 bp and 102 bp were entered in the plastid genomes of
S. microloba,S. yunnanensis (EY), S. yunnanensis (HB) and S. yunnanensis (KM), respectively. The ndhF
genes of Han. forrestii,Han. oviformis,Han. weberbaueriana,Hap. himalayensis,Hap. phaea,S. microloba
and S. yunnanensis (HB) are entirely within the SSC region, and a 6–48 bp intergenic region exists
between the ndhF gene and the JSB line (the border between IRb and SSC), while the ndhF genes of
Han. forbesii,S. yunnanensis (EY) and S. yunnanensis (KM) are partly included in IRb regions with 7 bp,
Plants 2020,9, 1523 8 of 17
5 bp and 5 bp, respectively. The ycf1 gene of the ten plastid genomes occupies the JSA line (the border
between SSC and IRa), with a length ranging from 1961 to 1997 bp in IRa regions, and ranging from 3475
to 3523 bp in close SSC regions. This also created a pseudogene ycf1 in the IRb regions. The trnH genes
of Han. forbesii,Han. forrestii,Han. oviformis,Han. weberbaueriana,Hap. himalayensis and Hap. phaea are
entirely within the LSC region. A 2–65 bp length of the intergenic region exists between the trnH genes
and JLA line (the border between IRa and SSC), while the trnH gene of S. microloba,S. yunnanensis (EY),
S. yunnanensis (HB) and S. yunnanensis (KM) is included in IRa regions with 1 bp.
Plants 2020, 9, x FOR PEER REVIEW 8 of 19
2.3. IR Boundaries and Simple Sequence Repeats (SSRs) Structure Analysis
The IR boundaries of the ten Hansenia, Haplosphaera and Sinodielsia plastid genomes were
compared to analyze the fluctuations (expansion or contraction) in these regions (Figure 4). Although
the ten plastid genomes showed a similar structure and content, some variations were still identified.
The rps19 gene entered the IRb region with 46 bp, 46 bp, 46 bp, 41 bp, 54 bp and 46 bp in the plastid
genomes of Han. forbesii, Han. forrestii, Han. oviformis, Han. weberbaueriana, Hap. himalayensis and Hap.
phaea (respectively), while 102 bp, 102 bp, 102 bp and 102 bp were entered in the plastid genomes of
S. microloba, S. yunnanensis (EY), S. yunnanensis (HB) and S. yunnanensis (KM), respectively. The ndhF
genes of Han. forrestii, Han. oviformis, Han. weberbaueriana, Hap. himalayensis, Hap. phaea, S. microloba
and S. yunnanensis (HB) are entirely within the SSC region, and a 6–48 bp intergenic region exists
between the ndhF gene and the JSB line (the border between IRb and SSC), while the ndhF genes of
Han. forbesii, S. yunnanensis (EY) and S. yunnanensis (KM) are partly included in IRb regions with 7
bp, 5 bp and 5 bp, respectively. The ycf1 gene of the ten plastid genomes occupies the JSA line (the
border between SSC and IRa), with a length ranging from 1961 to 1997 bp in IRa regions, and ranging
from 3475 to 3523 bp in close SSC regions. This also created a pseudogene ycf1 in the IRb regions. The
trnH genes of Han. forbesii, Han. forrestii, Han. oviformis, Han. weberbaueriana, Hap. himalayensis and
Hap. phaea are entirely within the LSC region. A 2–65 bp length of the intergenic region exists between
the trnH genes and JLA line (the border between IRa and SSC), while the trnH gene of S. microloba, S.
yunnanensis (EY), S. yunnanensis (HB) and S. yunnanensis (KM) is included in IRa regions with 1 bp.
Figure 4. Comparisons of inverted repeat (IR) boundaries between ten Hansenia, Haplosphaera and
Sinodielsia plastid genomes.
We found that the lengths of rps19 and trnH genes entering IRs are similar between Hansenia
(36–41 bp) and Haplosphaera (41–49 bp), which is different from these genes in Sinodielsia (102 bp). Not
Figure 4.
Comparisons of inverted repeat (IR) boundaries between ten Hansenia,Haplosphaera and
Sinodielsia plastid genomes.
We found that the lengths of rps19 and trnH genes entering IRs are similar between Hansenia
(36–41 bp) and Haplosphaera (41–49 bp), which is dierent from these genes in Sinodielsia (102 bp).
Not surprisingly, Hansenia is more closely related to Haplosphaera but distantly related to Sinodielsia,
which has also been confirmed by nrITS sequences and transcriptome data [
8
,
20
]. A collective analysis
of our results and previous studies [
26
29
] found that the lengths of rps19 and trnH genes entering IRs
are usually stable in the same genus or phylogenetically related groups. For example, the length of
rps19 gene entering IRs is 57–96 bp in Chamaesium species [
27
], and 49–84 bp in Bupleurum species [
26
],
which are the phylogenetically basal groups in the Chamaesium clade and Bupleureae. Whereas in
Apieae members, such as Anethum,Apium and Petroselium, the rps19 genes are not in the IRs. Instead,
the rpl2 genes are partly in the IRs [
26
,
30
]. Although the rpl2 genes are partly duplicated in the IRs of
some Apieae members, they are duplicated completely in the Hansenia,Haplosphaera and Sinodielsia
species. As for the two pseudogenes
ψ
rps19 and
ψ
ycf1 detected in the plastid genomes of Hansenia,
Haplosphaera and Sinodielsia, their incomplete duplication may be caused by the fluctuations in the
Plants 2020,9, 1523 9 of 17
IR boundary. The fluctuations may lead to gene loss (or pseudogene loss), which is a common
phenomenon in Apiaceae species [
26
,
27
]. Nevertheless, the length of land plants IRs can vary from
10 to 76 kbp, with most species having an IR of about 25 kbp, and often less than 15 kbp in lower
land plants and fern species [
31
]. Fluctuations in IR regions are the main reason for the dierences in
plastid genome lengths in most species, which also leads to several genes entering the IR regions or the
single-copy sequences [32].
Sixty-six, 72, 65, 62, 65, 75, 61, 57, 55 and 56 SSRs were identified in the plastid genomes of
Han. forbesii,Han. forrestii,Han. oviformis,Han. weberbaueriana,Hap. himalayensis,Hap. phaea,S. microloba,
S. yunnanensis (EY), S. yunnanensis (HB) and S. yunnanensis (KM), respectively. The results hint that
plastids of related groups share similar numbers of SSRs, while that is not always the result [
26
].
Six repeat motifs (mono-, di-, tri-, tetra-, penta- and hexanucleotide repeats) of microsatellites were
detected in the plastid genome of the Hansenia,Haplosphaera and Sinodielsia species. Mononucleotide
repeats were the most abundant SSR, which accounted for 58.47% of Hansenia, 60.21% of Haplosphaera
and 48.01% of Sinodielsia. Dinucleotide SSRs were the second most abundant SSR, which accounted
for 18.02% of Hansenia, 17.69% of Haplosphaera and 29.29% of Sinodielsia. This was followed by
tetranucleotide repeats of 8.69%, 9.28% and 4.79%, and trinucleotide repeats of 10.65%, 11.38% and
14.41% in Hansenia, Haplosphaera and Sinodielsia, respectively. Pentanucleotide and hexanucleotide
were the least abundant SSR (average of the three genera: 2.73% and 0.62%, respectively) (Figure 5).
The repeat motifs with the highest content in SSRs were all mononucleotide, which is similar to
most species including Apiaceae plants [
26
,
27
], but dierent from Forthysia (dinucleotide) [
33
] and
Nitotiana (trinucleotide) species [
34
]. The similar contents of mononucleotide and dinucleotide repeats
of Hansenia and Haplosphaera indicates that both genera seem to belong together. In all Hansenia,
Haplosphaera and Sinodielsia species the repeats were composed almost entirely of A/T, except for
mononucleotide repeat motifs that also had G/C contents of 10.53%, 2.33%, 10.81%, 7.32% and 4.65 in
Han. forbesii,Han. forrestii,Han. weberbaueriana,Hap. himalayensis and Hap. phaea. The phylogenetic
closeness of these species suggests that the relatives may share similar mononucleotide repeated
contents. Most dinucleotide repeats were AT/TA, except for one TC repeat that was in the plastid
genome of S. yunnanensis (HB). Across all SSR loci, 94 SSRs (14.83%) were found in the IRs, 417 SSRs
(65.77%) in LSC regions and 123 SSRs (19.40%) in SSC regions of the plastid genomes (Figure 6).
Plants 2020, 9, x FOR PEER REVIEW 10 of 19
Figure 5. Frequency of detected simple sequence repeat (SSR) motifs in different repeat types in ten
Hansenia, Haplosphaera and Sinodielsia plastid genomes.
Figure 5.
Frequency of detected simple sequence repeat (SSR) motifs in dierent repeat types in ten
Hansenia,Haplosphaera and Sinodielsia plastid genomes.
Plants 2020,9, 1523 10 of 17
Plants 2020, 9, x FOR PEER REVIEW 11 of 19
Figure 6. Frequency of detected SSR motifs in LSC (blue), IR (orange) and SSC (green) regions of the
ten Hansenia, Haplosphaera and Sinodielsia plastid genomes.
2.4. Codon Usage Analysis
The codon usage bias and relative synonymous codon usage (RSCU) values were calculated
using 53 PCGs in the ten Hansenia, Haplosphaera and Sinodielsia plastid genomes. There was no evident
divergence of the codon usage frequency when we compared the three genera (Figure 7). The number
of codons of PCGs ranged from 21,134 (S. yunnanensis: KM) to 21,254 (Han. weberbaueriana)
(Supplementary Table S2). Among these codons, leucine was encoded by 2227–2236 and cysteine was
encoded by 214–224 codons, which presented the maximum and the minimum number of codons per
amino acids in our study species, respectively. AUU (850–872) involved in encoding isoleucine and
UAG (13–14) involved in encoding a terminator were the most and least used codons. Furthermore,
30 codons of Hansenia and Haplosphaera, and 31 codons of Sinodielsia plastid genomes had RSCU
values larger than 1, indicating that they were the preferred codons in those ten plastid genomes.
Among these 30/31 preferred codons, most codons terminated in A/T, except that UUG ended with
G, and C was not found at the third position. This demonstrated that the codon usages of the ten
plastid genomes were biased towards A/T at the third position of codons, which is generally
consistent with other reported genomes of angiosperm [35,36], including the Bupleurum [26] and
Chamaesium [27] species.
Figure 6.
Frequency of detected SSR motifs in LSC (blue), IR (orange) and SSC (green) regions of the
ten Hansenia,Haplosphaera and Sinodielsia plastid genomes.
2.4. Codon Usage Analysis
The codon usage bias and relative synonymous codon usage (RSCU) values were calculated
using 53 PCGs in the ten Hansenia,Haplosphaera and Sinodielsia plastid genomes. There was no
evident divergence of the codon usage frequency when we compared the three genera (Figure 7).
The number of codons of PCGs ranged from 21,134 (S. yunnanensis: KM) to 21,254 (Han. weberbaueriana)
(Supplementary Table S2). Among these codons, leucine was encoded by 2227–2236 and cysteine was
encoded by 214–224 codons, which presented the maximum and the minimum number of codons per
amino acids in our study species, respectively. AUU (850–872) involved in encoding isoleucine and
UAG (13–14) involved in encoding a terminator were the most and least used codons. Furthermore,
30 codons of Hansenia and Haplosphaera, and 31 codons of Sinodielsia plastid genomes had RSCU values
larger than 1, indicating that they were the preferred codons in those ten plastid genomes. Among these
30/31 preferred codons, most codons terminated in A/T, except that UUG ended with G, and C was not
found at the third position. This demonstrated that the codon usages of the ten plastid genomes were
biased towards A/T at the third position of codons, which is generally consistent with other reported
genomes of angiosperm [35,36], including the Bupleurum [26] and Chamaesium [27] species.
Codon usage bias is an important indicator for studying the evolutionary relationship of
genomes [
37
]. Studies have shown that many biological factors aect the preference of synonymous
codon usage, such as gene expression level [
38
], gene sequence length [
39
], tRNA abundance [
40
]
and GC distribution position [
41
]. Codon bias seems to be maintained by selection–mutation–drift
balance [
42
]. There is a general agreement that the strong bias towards highly expressed genes is due
to selection for speed or translational eciency [
43
]. Our studied species of Hansenia,Haplosphaera and
Sinodielsia are from the Himalayan and Hengduan regions, usually inhabiting alpine areas over 3000 m,
and one of the reasons for their similar RSCU values may be the shared natural selection pressures,
even though they are not all closely related. This may provide another analytical technique to use
when studying nuclear genomes and can assist in understanding how species of separate lineages,
yet similar environments, have undergone parallel evolution. Increased numbers and improved
methods of analytical techniques are especially useful when studying many Pan-Himalayan region
Apiaceae species that are morphologically similar.
Plants 2020,9, 1523 11 of 17
Plants 2020, 9, x FOR PEER REVIEW 12 of 19
Figure 7. Codon content of 20 amino acids and the stop codon present in all 53 studied protein-coding
genes (PCGs) of the (A) Hansenia, (B) Haplosphaera and (C) Sinodielsia plastid genome. Different colors
of the histogram correspond to the different codons below.
Codon usage bias is an important indicator for studying the evolutionary relationship of
genomes [37]. Studies have shown that many biological factors affect the preference of synonymous
codon usage, such as gene expression level [38], gene sequence length [39], tRNA abundance [40] and
GC distribution position [41]. Codon bias seems to be maintained by selection–mutation–drift balance
[42]. There is a general agreement that the strong bias towards highly expressed genes is due to
selection for speed or translational efficiency [43]. Our studied species of Hansenia, Haplosphaera and
Sinodielsia are from the Himalayan and Hengduan regions, usually inhabiting alpine areas over 3000
m, and one of the reasons for their similar RSCU values may be the shared natural selection pressures,
even though they are not all closely related. This may provide another analytical technique to use
when studying nuclear genomes and can assist in understanding how species of separate lineages,
yet similar environments, have undergone parallel evolution. Increased numbers and improved
methods of analytical techniques are especially useful when studying many Pan-Himalayan region
Apiaceae species that are morphologically similar.
Figure 7.
Codon content of 20 amino acids and the stop codon present in all 53 studied protein-coding
genes (PCGs) of the (
A
)Hansenia, (
B
)Haplosphaera and (
C
)Sinodielsia plastid genome. Dierent colors
of the histogram correspond to the dierent codons below.
2.5. Nucleotide Diversity Analysis
Nucleotide diversity (Pi) values of the plastid genomes from the Hansenia,Haplosphaera and
Sinodielsia species were calculated to evaluate their sequence divergence level (Figure 8). In the four
Hansenia genomes, Pi values in the LSC regions ranged from 0 to 0.01889, with an average of 0.00371,
and ranged from 0 to 0.01472 in the SSC regions, with an average of 0.00263. The Pi values of IR regions
ranged from 0 to 0.00639 and had an average of 0.00089, the lowest Pi values of the three regions for
these four genomes. In the two Haplosphaera genomes, Pi values ranged from 0 to 0.04333 and averaged
0.00531 in the LSC regions, and from 0 to 0.02 in the SSC regions, with an average value of 0.0028.
The IR region values in the two Haplosphaera genomes were similarly low, with an average of 0.0015
and ranging from 0 to 0.01333. The two Haplosphaera species with high Pi values seem to be distant
species within the genus Hansenia (see further under phylogenetic studies). In the two Sinodielsia
genomes, Pi values ranged from 0 to 0.01222 with an average of 0.0021 in the LSC regions, and from
0 to 0.01028, averaging 0.00149 in the SSC regions. The Pi values of IR regions were again low for
the two Sinodielsia genomes and ranged from 0 to only 0.00833, with a value of 0.00112. Pi values
indicated mutations in the respective regions [
26
]. The Pi values of Sinodielsia were the lowest of the
three genera, indicating that S. microloba is more closely related to S. yunnanensis than the other species
within each genus.
Plants 2020,9, 1523 12 of 17
Plants 2020, 9, x FOR PEER REVIEW 14 of 19
Figure 8. The nucleotide diversity of (A) the four Hansenia, (B) the two Haplosphaera and (C) the four
Sinodielsia plastid genomes. The positions of 0.02 in all three graphs were marked by a line. Ten or
seven regions with the highest Pi values were named out.
We found that high Pi values of sequences were usually detected in spacer regions between
genes. Among these genomic spacer regions of the Hansenia species, trnK (UUU)-rps16, psbK-psbI,
trnE (UUC)-trnT (GGU), rps4-trnL (UAA), petL-psaJ, rpl22-rps19 and ndhG-ndhI had the highest Pi
values, ranging from 0.01083 to 0.01889. The highest Pi values of spacer regions of the Haplosphaera
species ranged between 0.015 and 0.04333 for trnH (GUG)-psbA, atpH-atpI, rps4-trnT (UGU), ycf4-
cemA, petB-petD and ndhG-ndhI. Within the spacer regions of Sinodielsia species, matK-rps16, trnE
(UUC)-trnT (GGU), ycf3-trnS (GGA), trnL (UAA)-ndhJ, rpl33-rps18, rps11-infA, rps3-rps19, ycf1-ndhF,
ndhF-rpl32 and ndhG-ndhI had the highest Pi values, ranging from 0.00722 to 0.01222. Although the
Figure 8.
The nucleotide diversity of (
A
) the four Hansenia, (
B
) the two Haplosphaera and (
C
) the four
Sinodielsia plastid genomes. The positions of 0.02 in all three graphs were marked by a line. Ten or
seven regions with the highest Pi values were named out.
We found that high Pi values of sequences were usually detected in spacer regions between
genes. Among these genomic spacer regions of the Hansenia species, trnK (UUU)-rps16,psbK-psbI,
trnE (UUC)-trnT (GGU),rps4-trnL (UAA),petL-psaJ,rpl22-rps19 and ndhG-ndhI had the highest Pi
values, ranging from 0.01083 to 0.01889. The highest Pi values of spacer regions of the Haplosphaera
species ranged between 0.015 and 0.04333 for trnH (GUG)-psbA,atpH-atpI,rps4-trnT (UGU),ycf4-cemA,
petB-petD and ndhG-ndhI. Within the spacer regions of Sinodielsia species, matK-rps16,trnE (UUC)-trnT
(GGU),ycf3-trnS (GGA),trnL (UAA)-ndhJ,rpl33-rps18,rps11-infA,rps3-rps19,ycf1-ndhF,ndhF-rpl32 and
Plants 2020,9, 1523 13 of 17
ndhG-ndhI had the highest Pi values, ranging from 0.00722 to 0.01222. Although the spacer regions
were diverse among the three genera, we found the spacer regions ndhG-ndhI occurred in all three
genera with relatively high values. Pending more specimen samples and plastid genome studies of
Chinese Apiaceae, the spacer regions (e.g., high Pi value ndhG-ndhI) may provide DNA barcodes for
molecular identification and phylogenetic studies in the large-scale clades such as East Asia clade and
Acronema clade, where there are many parallel branches whose support rate is less than 50% using ITS
or plastid DNA introns (rpl16 and rps16 genes) [8,9].
3. Materials and Methods
3.1. Material, DNA Extraction and Complete Genome Sequencing
The materials of Han. forbesii,Han. forrestii,Han. oviformis,Han. weberbaueriana,Hap. himalayensis,
S. microloba, S. yunnanensis (EY) and S. yunnanensis (KM) were newly obtained for this study, which were
collected from the type localities or their adjacent areas during 2018–2019 (Supplementary Table S1).
The dierent populations of S. yunnanensis are represented by EY: from Eryuan County, Yunnan; KM:
from Kunming, and Yunnan; HB: from Zhongdian, Yunnan (HB previously sequenced [
19
]). The total
genomic DNA was extracted from silica gel-dried leaves according to the protocols of the plant genomic
DNA kit (cwbio, Beijing, China), then sequenced using the Illumina Novaseq 6000 platform (Illumina,
San Diego, CA, USA) with Novaseq 150 sequencing strategy by Novogene (Beijing, China).
3.2. Genome Construction and Annotation
The clean data (removed connectors and low-quality reads) were assembled using NOVOPlasty
2.7.1 [
44
] with K-mer 39, where the rbcL gene of Han. oviformis and S. yunnanensis (amplificated and
sequenced beforehand) was used as a seed input and the reference sequence. The assembled complete
plastid genomes were checked then aligned with the reference plastid genome of Chuanminshen violaceum
(KU921430) using GENEIOUS R11 [
45
] to select the best option, which was then annotated using
PGA [
46
]. GENEIOUS R11 was then used to manually adjust the annotation for uncertain start and
stop codons based on the comparison with homologous genes from other annotated plastid genomes.
The eight annotated plastid genomes were submitted to GenBank, and their accession numbers are in
Supplementary Table S1. Their genome maps were drawn using OGDRAW version 1.3.1 [47].
3.3. Phylogenetic Analysis
To better infer phylogenetic relationships between Hansenia,Haplosphaera and Sinodielsia,
the 32 plastid genomes were applied for reconstructing the phylogenetic tree, of which 24 plastid
genomes were obtained from GenBank (Supplementary Table S1). All the plastid genomes were aligned
using MAFFT v7.308 [
48
,
49
]. Maximum Likelihood was then conducted for phylogenetic analyses
using RAxML version 8.2.4 [
50
] under the model GTR+G with 1000 rapid bootstraps. The scientific
names of plants followed the International Plant Names Index (https://www.ipni.org). Chamaesium and
Bupleurum species were selected as outgroups [8,9].
3.4. IRs Boundaries and SSR Analysis
The boundaries between the LSC, SSC and IR regions of the ten plastid genomes were compared
and drawn using the online program: IRscope (https://irscope.shinyapps.io/irapp/) [
51
]. The SSRs
were identified using MISA [
52
] with the repeat threshold settings: 10 repeats for mono-nucleotide,
5 for di-nucleotide, 4 for tri-nucleotide and 3 repeats for tetra-, penta-and hexanucleotide SSRs.
3.5. Codon Usage Bias Analysis
Codon usage bias analysis and calculation of RSCU values were conducted using the program
CodonW [
53
]. Fifty-three PCGs (more than 300 bp in length) of each Hansenia,Haplosphaera and
Sinodielsia plastid genomes were filtered. The codon adaptation index (CAI), the codon bias index (CBI),
Plants 2020,9, 1523 14 of 17
the eective number of codons (ENC), the frequency of optimal codons (Fop) and the GC content of the
synonymous third codons positions (GC3s) were calculated to assess the extent of the codon usage bias.
The RSCU values of the four Hansenia, two Haplosphaera and four Sinodielsia plastid genomes were also
calculated to assess their codon usages, including Hap. phaea and S. yunnanensis (HB) downloaded
from GenBank (Supplementary Table S1).
3.6. Nucleotide Diversity Analysis
The plastid genomes of Hansenia,Haplosphaera and Sinodielsia species were aligned using MAFFT
v7.308 [
48
,
49
]. DNA polymorphism analysis was then conducted to calculate the Pi values in DnaSP
v5 [
54
] in the sliding window. The setting parameters were as follows: (1) windows length: 600 sites;
(2) step size: 200 sites.
4. Conclusions
In this study, the plastid genomes of Han. forbesii,Han. forrestii,Han. oviformis,Han. weberbaueriana,
Hap. himalayensis,S. microloba,S. yunnanensis (KM) and S. yunnanensis (EY) were newly assembled
and annotated. All the eight plastid genomes exhibited a typical circular quadripartite organization
with similar whole length (155,435 bp to 157,797 bp) and gene contents. The IR boundary analysis
showed that length of rps19 and ycf1 genes entering IRs are usually stable in the same genus.
Hansenia shared relatively similar mononucleotide SSRs to Haplosphaera. Additionally, Hansenia,
Haplosphaera and Sinodielsia species had similar codon usage. Although Hansenia and Haplosphaera are
not phylogenetically close to Sinodielsia, there was no significant dierence in their plastid genomes.
Furthermore, we found that the ndhG-ndhI spacer regions possessed higher nucleotide diversity in
the three genera and, therefore, may provide DNA barcodes for intra- and inter-genus identification
in Apiaceae. The phylogeny of the 32 plastid genomes, including the eight taxa mentioned above,
showed a close relationship between Hansenia and Haplosphaera, and S. microloba may be a species
of hybrid origin. This study will enrich the complete plastid genome dataset of the genera Hansenia,
Haplosphaera and Sinodielsia, and has provided a new insight into comparisons of distant taxa and
phylogeny reconstruction using complete plastid genomes of Apiaceae.
Supplementary Materials:
The following are available online at http://www.mdpi.com/2223-7747/9/11/1523/s1,
Table S1: Voucher details and GenBank accession numbers of taxa used in this study, Table S2: The indexes of the
codon usage bias in the Hansenia,Haplosphaera and Sinodielsia species.
Author Contributions:
Conceptualization, W.G., S.-B.J., X.-L.G., S.-D.Z. and X.-J.H.; Data curation, W.G. and
S.-B.J.; Formal analysis, W.G. and S.-B.J.; Funding acquisition, S.-D.Z. and X.-J.H.; Investigation, W.G. and S.-B.J.;
Methodology, W.G. and S.-B.J.; Resources, W.G. and S.-B.J.; Software, W.G. and S.-B.J.; Supervision, S.-D.Z. and
X.-J.H.; Writing-original draft, W.G. and S.-B.J.; Writing-review & editing, W.G., S.-B.J., M.P., X.-L.G., S.-D.Z. and
X.-J.H. All authors have read and agreed to the published version of the manuscript.
Funding:
This work was supported by the National Natural Science Foundation of China (Grant Nos. 31872647).
The Chinese Ministry of Science and Technology throng the National Science and Technology Infrastructure
Platform Project (Grant No. 2005DKA21403-JK). The fourth national survey of traditional Chinese medicine
resources (Grant No. 2019PC002).
Acknowledgments: We would like to thank Yan Yu, Dan-Mei Su and Fu-Min Xie for the help in software use.
Conflicts of Interest: The authors declare no conflict of interest.
References
1.
Pu, F.D.; Watson, M.F. Apiaceae. In Flora of China; Science Press: Beijing, China; Missouri Botanical garden
Press: Saint Louis, MO, USA, 2005; Volume 14, pp. 1–205.
2.
Pimenov, M.G.; Kljuykov, E.V.; Ostroumova, T.A. Reduction of Notopterygium to Hansenia (Umbelliferae).
Willdenowia 2008,38, 155–172. [CrossRef]
3.
Jia, S.B.; Guo, X.L.; Zhou, S.D.; He, X.J. Hansenia pinnatiinvolucellata is conspecific with H. weberbaueriana
(Apiaceae) based on morphology and molecular data. Phytotaxa 2019,418, 203–210. [CrossRef]
Plants 2020,9, 1523 15 of 17
4.
Pimenov, M.G.; Kljuykov, E.V. New nomenclatural combinations for Chinese Umbelliferae. Feddes Repert.
1999,110, 481–491. [CrossRef]
5.
Pimenov, M.G. Updated checklist of Chinese Umbelliferea: Nomenclature, synonymy, typification,
distribution. Turcz 2017,20, 106–239.
6.
Valiejo-Roman, C.M.; Terentieva, E.I.; Samigullin, T.H.; Pimenov, M.G. nrDNA ITS sequences and anities
of Sino-Himalayan Apioideae (Umbelliferae). Taxon 2002,51, 685–701.
7.
Valiejo-Roman, C.M.; Shneyer, V.S.; Samigullin, T.H.; Terentieva, E.I.; Pimenov, M.G. An attempt to clarify
taxonomic relationships in “Verwandtschaftskreis der Gattung Ligusticum” (Umbelliferae-Apioideae) by
molecular analysis. Plant Syst. Evol. 2006,257, 25–43. [CrossRef]
8.
Zhou, J.; Peng, H.; Downie, S.R.; Liu, Z.W.; Gong, X. A molecular phylogeny of Chinese Apiaceae
subfamily Apioideae inferred from nuclear ribosomal DNA internal transcribed spacer sequences. Taxon
2008,57, 402–416.
9.
Zhou, J.; Gong, X.; Downie, S.R.; Peng, H. Towards a more robust molecular phylogeny of Chinese Apiaceae
subfamily Apioideae: Additional evidence from nrDNA ITS and cpDNA intron (rpl16 and rps16) sequences.
Mol. Phylogenetics Evol. 2009,53, 56–68. [CrossRef]
10.
Wang, Z.X.; Downie, S.R.; Tan, J.B.; Liao, C.Y.; Yu, Y.; He, X.J. Molecular phylogenetics of Pimpinella and
allied genera (Apiaceae), with emphasis on Chinese native species, inferred from nrDNA ITS and cpDNA
intron sequence data. Nord. J. Bot. 2014,32, 642–657. [CrossRef]
11.
Liu, Z.W.; Zhou, J.; Gao, Y.z.; Wei, J.; Downie, S.R. Molecular phylogenetics of Ligusticum (Apiaceae) based
on nrDNA ITS sequences: Placement of the Chinese endemic species and a reduced circumscription of the
genus. Int. J. Plant Sci. 2019.
12.
Downie, S.R.; Palmer, J.D. Use of Chloroplast DNA Rearrangements in Reconstructing Plant Phylogeny; Molecular
Systematics of Plants; Soltis, P.S., Soltis, D.E., Doyle, J.J., Eds.; Springer: Boston, MA, USA, 1992.
13.
Ravi, V.; Khurana, J.P.; Tyagi, A.K.; Khurana, P. An update on chloroplast genomes. Plant Syst. Evol.
2008,271, 101–122. [CrossRef]
14.
Yu, X.L.; Tan, W.; Zhang, H.Y.; Gao, H.; Wang, W.X.; Tian, X.X. Complete chloroplast genomes of
Ampelopsis humulifolia and Ampelopsis japonica: Molecular structure, comparative analysis, and phylogenetic
analysis. Plants 2019,8, 410. [CrossRef]
15.
Jansen, R.K.; Raubeson, L.A.; Boore, J.L.; Depamphilis, C.W.; Cui, L. Methods for Obtaining and Analyzing
Whole Chloroplast Genome Sequences. Methods Enzymol. 2005,395, 348–384. [PubMed]
16.
Korpelainen, H. The evolutionary processes of mitochondrial and chloroplast genomes dier from those of
nuclear genomes. Sci. Nat. 2004,91, 505–518. [CrossRef]
17.
Huotari, T.; Korpelainen, H. Complete chloroplast genome sequence of Elodea canadensis and comparative
analyses with other monocot plastid genomes. Gene 2012,508, 96–105. [CrossRef]
18.
Jia, S.B.; Guo, X.L.; Xie, D.F.; Yu, Y.; Zhou, S.D.; He, X.J. The complete chloroplast genome of Haplosphaera phaea
(Apiaceae). Mitochondrial Dna B 2019,4, 1969–1970. [CrossRef]
19.
Gou, W.; Guo, X.L.; Yu, Y.; Zhou, S.D.; He, X.J. The complete chloroplast genome of Meeboldia yunnanensis
(Apiaceae). Mitochondrial Dna B 2019,4, 4176–4177. [CrossRef]
20.
Wen, J.; Yu, Y.; Xie, D.F.; Peng, C.; Liu, Q.; Zhou, S.D.; He, X.J. A transcriptome-based study on the phylogeny
and evolution for taxonomic controversial subfamily Apioideae (Apiaceae). Ann. Bot.
2020
,125, 937–953.
[CrossRef]
21.
Zhang, Y.Z.; Ma, J.; Yang, B.X.; Li, R.Y.; Zhu, W.; Sun, L.L.; Tian, J.K.; Zhang, L. The complete chloroplast
genome sequence of Taxus chinensis var. mairei (Taxaceae): Loss of an inverted repeat region and comparative
analysis with related species. Gene 2014,540, 201–209. [CrossRef]
22.
Guo, X.L.; Wang, C.B.; Wen, J.; Zhou, S.D.; He, X.J. Phylogeny of Chinese Chamaesium (Apiaceae: Apioideae)
inferred from ITS, cpDNA and morphological characters. Phytotaxa 2018,376, 1–16. [CrossRef]
23.
Xiao, Q.Y.; Yu, Y.; Xie, D.F.; Guo, X.L.; He, X.J. Taxonomic revision of Angelica oncosepala and Heracleum
yunnanense.Nord. J. Bot. 2018,36, 1–10.
24.
Guisinger, M.M.; Kuehl, J.V.; Boore, J.L.; Jansen, R.K. Extreme Reconfiguration of Plastid Genomes in
the Angiosperm Family Geraniaceae: Rearrangements, Repeats, and Codon Usage. Mol. Biol. Evol.
2011,28, 583–600. [CrossRef]
25.
Hu, Y.J. Chloroplast genome structure and chloroplast genes. Plant. Physiol. Commun.
1985
,2, 65–71.
(In Chinese)
Plants 2020,9, 1523 16 of 17
26.
Li, J.; Xie, D.F.; Guo, X.L.; Zheng, Z.Y.; He, X.J.; Zhou, S.D. Comparative Analysis of the Complete Plastid
Genome of Five Bupleurum Species and New Insights into DNA Barcoding and Phylogenetic Relationship.
Plants 2020,9, 543. [CrossRef]
27.
Guo, X.L.; Zheng, H.Y.; Price, M.; Zhou, S.D.; He, X.J. Phylogeny and Comparative Analysis of Chinese
Chamaesium Species Revealed by the Complete Plastid Genome. Plants 2020,9, 965. [CrossRef]
28.
Mustafina, F.U.; Yi, D.K.; Choi, K.; Shin, C.H.; Tojibaev, K.S.; Downie, S.R. A comparative analysis of complete
plastid genomes from Prangos fedtschenkoi and Prangos lipskyi (Apiaceae). Ecol. Evol. 2018, 1–14. [CrossRef]
29. Kang, L.; Xie, D.F.; Xiao, Q.Y.; Peng, C.; Yu, Y.; He, X.J. Sequencing and analyses on chloroplast genomes of
Tetrataenium candicans and two allies give new insights on structural variants, DNA barcoding and phylogeny
in Apiaceae subfamily Apioideae. PeerJ 2019,7, e8063. [CrossRef]
30.
Peery, R.M. Understanding Angiosperm Genome Interactions and Evolution: Insights from Sacred
Lotus (Nelumbo nucifera) and the Carrot Family (Apiaceae). Ph.D. Thesis, University of Illinois at
Urbana-Champaign, Champaign, IL, USA, 2015.
31.
Aii, J.; Kishima, Y.; Mikami, T.; Adachi, T. Expansion of the IR in the chloroplast genomes of buckwheat
species is due to incorporation of an SSC sequence that could be mediated by an inversion. Curr. Genet.
1997,31, 276–279. [CrossRef]
32.
Chung, H.J.; Jung, J.D.; Park, H.W.; Kim, J.H.; Cha, H.W. The complete chloroplast genome sequences of
Solanum tuberosum and comparative analysis with Solanaceae species identified the presence of a 241-bp
deletion in cultivated potato chloroplast DNA sequence. Plant Cell Rep. 2006,25, 1369–1379. [CrossRef]
33.
Wang, W.B.; Yu, H.; Wang, J.H.; Lei, W.J.; Gao, J.H.; Qiu, X.P.; Wang, J.S. The complete chloroplast genome
sequences of the medicinal plant Forsythia suspensa (Oleaceae). Int. J. Mol. Sci.
2017
,18, 2288. [CrossRef]
[PubMed]
34.
Asaf, S.; Khan, A.L.; Khan, A.R.; Waqas, M.; Kang, S.M.; Khan, M.A.; Lee, S.M.; Lee, I.J. Complete chloroplast
genome of Nicotiana otophora and its comparison with related species. Front. Plant. Sci.
2016
,7, 447.
[CrossRef]
35.
Sharp, P.M. An evolutionary perspective on synopnymous codon usage in unicellular organisms. J. Mol. Evol.
1986,24, 28–38.
36.
Du, Y.; Bi, Y.; Chen, X.; Yang, F.; Xue, J.; Zhang, X. The complete chloroplast genome of Lilium cernuum:
Genome structure and evolution. Conserv. Genet. Resour. 2016,8, 375–378. [CrossRef]
37.
Grantham, R.; Gautier, C.; Gouy, M. Codon frequencies in 119 individual genes confirm consistent choices of
degenerate bases according to genome type. Nucleic Acids Res. 1980,8, 1893–1912. [CrossRef]
38.
Rao, Y.S.; Wu, G.Z.; Wang, Z.F.; Chai, X.W.; Nie, Q.H.; Zhang, X.Q. Mutation Bias is the Driving Force of
Codon Usage in the genome. DNA Res. 2011,18, 499–512. [CrossRef]
39.
Rensing, S.A.; Fritzowsky, D.; Lang, D.; Reski, R. Protein encoding genes in an ancient plant: Analysis of
codon usage, retained genes and splice sites in a moss, Physcomitrella patens. BMC Genom.
2005
,6, 43.
[CrossRef] [PubMed]
40.
Novoa, E.M.; Pouplana, L.R.D. Speeding with control: Codon usage, TRNAs, and ribosomes. Trends Genet.
2012,28, 574–581. [CrossRef]
41.
Harrison, R.J.; Charlesworth, B. Biased Gene Conversion Aects Patterns of Codon Usage and Amino Acid
Usage in the Saccharomyces sensu stricto Group of Yeasts. Mol. Biol. Evol. 2011,28, 117–129. [CrossRef]
42.
Clark, A.; Eisen, M.; Smith, D.; Singh, N.D.; Sackton, T.B.; Larracuente, A.M.; Bergman, C.M.; Gelbart, W.M.;
Oliver, B.; Markow, T.A.; et al. Evolution of genes and genomes on the Drosophila phylogeny. Nature
2007,450, 203–218.
43.
Bulmer, M.G. The selection-mutation-drift theory of synonymous codon usage. Genetics
1991
,129, 897–907.
44.
Dierckxsens, N.; Mardulyn, P.; Smits, G. NOVOPlasty: De novo assembly of organelle genomes from whole
genome data. Nucleic Acids Res. 2017,4, 4.
45.
Kearse, M.; Moir, R.; Wilson, A.; Stones-Havas, S.; Cheung, M.; Sturrock, S.; Buxton, S.; Cooper, A.;
Markowitz, S.; Duran, C.; et al. Geneious Basic: An integrated and extendable desktop software platform for
the organization and analysis of sequence data. Bioinformatics 2012,28, 1647–1649. [CrossRef]
46.
Qu, X.J.; Moore, M.J.; Li, D.Z.; Yi, T.S. PGA: A software package for rapid, accurate, and flexible batch
annotation of plastomes. Plant Methods 2019,15, 50. [CrossRef]
47.
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, 59–64. [CrossRef]
Plants 2020,9, 1523 17 of 17
48.
Amiryousefi, A.; Hyvönen, J.; Poczai, P. Irscope: An online program to visualize the junction sites of
chloroplast genomes. Bioinformatics 2018,34, 3030–3031. [CrossRef] [PubMed]
49.
Thiel, T.; Michalek, W.; Varshney, R.; Graner, A. Exploiting est databases for the development and characterization
of gene-derived ssr-markers in barley (hordeum vulgarel.). Theor. Appl. Genet. 2003,106, 411–422. [CrossRef]
50. Peden, J.F. CodonW. Ph.D. Thesis, University of Nottingham, Nottinghamshire, UK, 1999.
51.
Katoh, K.; Misawa, K.; Kuma, K.i.; Miyata, T. MAFFT: A novel method for rapid multiple sequence alignment
based on fast Fourier transform. Nucleic Acids Res. 2002,30, 3059–3066. [CrossRef]
52.
Katoh, K.; Standley, D.M. MAFFT Multiple Sequence Alignment Software Version 7: Improvements in
Performance and Usability. Mol. Phylogenet. Evol. 2013,30, 772–780. [CrossRef]
53.
Librado, P.; Rozas, J. DnaSP v5: A software for comprehensive analysis of DNA polymorphism data.
Bioinformatics 2009,25, 1451–1452. [CrossRef] [PubMed]
54.
Stamatakis, A. RAxML version 8: A tool for phylogenetic analysis and post-analysis of large phylogenies.
Bioinformatics 2014,30, 1312–1313. [CrossRef]
Publisher’s Note:
MDPI stays neutral with regard to jurisdictional claims in published maps and institutional
aliations.
©
2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
... For these obvious advantages, the plastome has become a powerful tool to generate highly supported and resolved phylogenies and to explore more efficient specific DNA barcodes [25,27,28,[31][32][33]. In recent years, with the rapid development of next-generation sequencing technology and bioinformatics technology, a large number of sequence data of plastomes has become more accessible at a much lower cost [34], and it has been extensively and successfully applied to solve the plants phylogenies, especially for those taxonomically difficult taxa within the family Apiaceae [30][31][32][35][36][37][38][39][40]. ...
... IR contraction and expansion is a common phenomenon in the Apiaceae plastomes [20,[30][31][32][35][36][37][38][39][40]. This Fig. 9 The divergence time estimation based on 79 shared genes. ...
... Therefore, we supported the hypothesis that the genome size variation was caused by IR contraction and expansion. Furthermore, we also examined the inverted repeats types described in the Apiaceae [20,[30][31][32][35][36][37][38][39][40] and found that the situation observed in the genus Sanicula was also detected in the other genera of Apiaceae, such as in Hansenia Turcz., Haplosphaera Hand.-Mazz., Sinodielsia H. Wolff [37]. ...
Article
Full-text available
Background The genus Sanicula L. is a unique perennial herb that holds important medicinal values. Although the previous studies on Sanicula provided us with a good research basis, its taxonomic system and interspecific relationships have not been satisfactorily resolved, especially for those endemic to China. Moreover, the evolutionary history of this genus also remains inadequately understood. The plastid genomes possessing highly conserved structure and limited evolutionary rate have proved to be an effective tool for studying plant phylogeny and evolution. Results In the current study, we newly sequenced and assembled fifteen Sanicula complete plastomes. Combined with two previously reported plastomes, we performed comprehensively plastid phylogenomics analyses to gain novel insights into the evolutionary history of this genus. The comparative results indicated that the seventeen plastomes exhibited a high degree of conservation and similarity in terms of their structure, size, GC content, gene order, IR borders, codon bias patterns and SSRs profiles. Such as all of them displayed a typical quadripartite structure, including a large single copy region (LSC: 85,074–86,197 bp), a small single copy region (SSC: 17,047–17,132 bp) separated by a pair of inverted repeat regions (IRs: 26,176–26,334 bp). And the seventeen plastomes had similar IR boundaries and the adjacent genes were identical. The rps19 gene was located at the junction of the LSC/IRa, the IRa/SSC junction region was located between the trnN gene and ndhF gene, the ycf1 gene appeared in the SSC/IRb junction and the IRb/LSC boundary was located between rpl12 gene and trnH gene. Twelve specific mutation hotspots (atpF, cemA, accD, rpl22, rbcL, matK, ycf1, trnH-psbA, ycf4-cemA, rbcL-accD, trnE-trnT and trnG-trnR) were identified that can serve as potential DNA barcodes for species identification within the genus Sanicula. Furthermore, the plastomes data and Internal Transcribed Spacer (ITS) sequences were performed to reconstruct the phylogeny of Sanicula. Although the tree topologies of them were incongruent, both provided strong evidence supporting the monophyly of Saniculoideae and Apioideae. In addition, the sister groups between Saniculoideae and Apioideae were strongly suggested. The Sanicula species involved in this study were clustered into a clade, and the Eryngium species were also clustered together. However, it was clearly observed that the sections of Sanicula involved in the current study were not respectively recovered as monophyletic group. Molecular dating analysis explored that the origin of this genus was occurred during the late Eocene period, approximately 37.84 Ma (95% HPD: 20.33–52.21 Ma) years ago and the diversification of the genus was occurred in early Miocene 18.38 Ma (95% HPD: 10.68–25.28 Ma). Conclusion The plastome-based tree and ITS-based tree generated incongruences, which may be attributed to the event of hybridization/introgression, incomplete lineage sorting (ILS) and chloroplast capture. Our study highlighted the power of plastome data to significantly improve the phylogenetic supports and resolutions, and to efficiently explore the evolutionary history of this genus. Molecular dating analysis explored that the diversification of the genus occurred in the early Miocene, which was largely influenced by the prevalence of the East Asian monsoon and the uplift of the Hengduan Mountains (HDM). In summary, our study provides novel insights into the plastome evolution, phylogenetic relationships, taxonomic framework and evolution of genus Sanicula.
... transiliensis) within the 22 plastomes ( Fig. 2A). The most abundant were mononucleotide repeats (32)(33)(34)(35)(36)(37)(38)(39)(40)(41)(42)(43)(44)(45)(46)(47)(48), followed by dinucleotides (14)(15)(16)(17)(18)(19), tetranucleotides (8)(9)(10)(11)(12), trinucleotides (3)(4)(5), and pentanucleotides (0-3). Only F. songarica and F. kingdon-wardii had one hexanucleotide ( Fig. 2A). ...
... Additionally, the gene numbers, type and distribution of large repeats, number and type of SSRs, and codon usage were rather similar among these plastomes. This circumstance is common across other genera in the family Apiaceae [47,48]. Therefore, these results demonstrated that the plastomes were highly conserved in terms of structure, gene number, type and distribution of large repeat, number and type of SSRs, and codon usage in Ferula, Talassia and Soranthus. ...
Article
Full-text available
Background The Ferula genus encompasses 180–185 species and is one of the largest genera in Apiaceae, with many of Ferula species possessing important medical value. The previous studies provided more information for Ferula, but its infrageneric relationships are still confusing. In addition, its genetic basis of its adaptive evolution remains poorly understood. Plastid genomes with more variable sites have the potential to reconstruct robust phylogeny in plants and investigate the adaptive evolution of plants. Although chloroplast genomes have been reported within the Ferula genus, few studies have been conducted using chloroplast genomes, especially for endemic species in China. Results Comprehensively comparative analyses of 22 newly sequenced and assembled plastomes indicated that these plastomes had highly conserved genome structure, gene number, codon usage, and repeats type and distribution, but varied in plastomes size, GC content, and the SC/IR boundaries. Thirteen mutation hotspot regions were detected and they would serve as the promising DNA barcodes candidates for species identification in Ferula and related genera. Phylogenomic analyses with high supports and resolutions showed that Talassia transiliensis and Soranthus meyeri were nested in the Ferula genus, and thus they should be transferred into the Ferula genus. Our phylogenies also indicated the monophyly of subgenera Sinoferula and subgenera Narthex in Ferula genus. Twelve genes with significant posterior probabilities for codon sites were identified in the positively selective analysis, and their function may relate to the photosystem II, ATP subunit, and NADH dehydrogenase. Most of them might play an important role to help Ferula species adapt to high-temperatures, strong-light, and drought habitats. Conclusion Plastome data is powerful and efficient to improve the support and resolution of the complicated Ferula phylogeny. Twelve genes with significant posterior probabilities for codon sites were helpful for Ferula to adapt to the harsh environment. Overall, our study supplies a new perspective for comprehending the phylogeny and evolution of Ferula.
... We estimated and compared the SSRs in the ten plastomes. The total number of SSRs altered from 71 to 84, of which the LSC region had the richest repeats (46)(47)(48)(49)(50)(51)(52)(53)(54)(55)(56)(57)(58)(59)(60)(61)(62) and the SSC region (9)(10)(11)(12)(13)(14) had the poorest following IR regions (8)(9)(10)(11)(12)(13)(14) (Table S5, Figure 5A). Figure 5B displayed the number of different repeat types in ten plastomes. ...
... We reported the newly sequenced and assembled complete plastome of P. franchetii and performed a comparative analysis with nine other Ligusticopsis species. The results indicated that ten plastomes exhibited a typical quadripartite structure, which was in accordance with the plastome structure of Apiaceae published previously [42][43][44][45][46][47][48]. Furthermore, all plastomes were similar in genome size, gene order, and GC content. ...
Article
Full-text available
Peucedanum franchetii is a famous folk medicinal plant in China. However, the taxonomy of the P. franchetii has not been sufficiently resolved. Due to similar morphological features between P. franchetii and Ligusticopsis members, the World Flora Online (WFO) Plant List suggested that this species transformed into the genus Ligusticopsis and merged with Ligusticopsis likiangensis. However, both species are obviously diverse in leaf shape, bracts, and bracteoles. To check the taxonomic position of P. franchetii, we newly sequenced and assembled the plastome of P. franchetii and compared it with nine other plastomes of the genus Ligusticopsis. Ten plastomes were highly conserved and similar in gene order, codon bias, RNA editing sites, IR borders, and SSRs. Nevertheless, 10 mutation hotspot regions (infA, rps8, matK, ndhF, rps15, psbA-trnH, rps2-rpoC2, psbA-trnK, ycf2-trnL, and ccsA-ndhD) were still detected. In addition, both phylogenetic analyses based on plastome data and ITS sequences robustly supported that P. franchetii was not clustered with members of Peucedanum but nested in Ligusticopsis. P. franchetii was sister to L. likiangensis in the ITS topology but clustered with L. capillacea in the plastome tree. These findings implied that P. franchetii should be transferred to genus Ligusticopsis and not merged with L. likiangensis, but as an independent species, which was further verified by morphological evidences. Therefore, transferring P. franchetii under the genus Ligusticopsis as an independent species was reasonable, and a new combination was presented.
... In most angiosperms, plastids are usually considered to be inherited from the maternal parent and have low nucleotide substitution rates (Wicke et al., 2011;Wataru and Tsuneaki, 2023). Thus, the plastid genomes (plastomes) have been widely and successfully used for plant phylogenetic analyses (Duminil et al., 2012;Miller et al., 2014;Razafimandimbison et al., 2014;Zhang et al., 2018;Schneider et al., 2021;Xu and Hong., 2021;Ji et al., 2022;Scatigna et al., 2022;Xiang et al., 2022;Baldwin et al., 2023;Fu et al., 2023), especially for those taxonomically controversial taxa within the family Apiaceae (Gou et al., 2020;Ren et al., 2020Ren et al., , 2022Cai et al., 2022;Liu et al., 2022;Guo et al., 2023;Liu et al., 2023a;Lei et al., 2022;Gui et al., 2023;Peng et al., 2023;Qin et al., 2023;Song et al., 2023;Tian et al., 2023;Song et al., 2024). For example, Song et al. (2023) transferred Peucedanum franchetii C.Y.Wu & F.T.Pu under the genus Ligusticopsis Leute based on phylogenetic analysis of ten plastomes. ...
Article
Full-text available
Introduction The genus Sanicula L. is a taxonomically complicated taxa within Apiaceae, as its high variability in morphology. Although taxonomists have performed several taxonomic revisions for this genus, the interspecific relationships and species boundaries have not been satisfactorily resolved, especially for those endemic to China. This study mainly focused on S. giraldii var. ovicalycina, S. tienmuensis var. pauciflora, and S. orthacantha var. stolonifera and also described two new members of the genus. Methods We newly sequenced sixteen plastomes from nine Sanicula species. Combined with eleven plastomes previously reported by us and one plastome downloaded, we performed a comprehensively plastid phylogenomics analysis of 21 Sanicula taxa. Results and Discussion The comparative results showed that 21 Sanicula plastomes in their structure and features were highly conserved and further justified that two new species were indeed members of Sanicula. Nevertheless, eleven mutation hotspot regions were still identified. Phylogenetic analyses based on plastome data and the ITS sequences strongly supported that these three varieties were clearly distant from three type varieties. The results implied that these three varieties should be considered as three independent species, which were further justified by their multiple morphological characters. Therefore, revising these three varieties into three independent species was reasonable and convincing. Moreover, we also identified and described two new Sanicula species (S. hanyuanensis and S. langaoensis) from Sichuan and Shanxi, China, respectively. Based on their distinct morphological characteristics and molecular phylogenetic analysis, two new species were included in Sanicula. In summary, our study impelled the revisions of Sanicula members and improved the taxonomic system of the genus.
... Complete cp genomes from 20 Quercus section Cyclobalanopsis species were aligned using the multiple sequence alignment program MAFFT v7.487 [46]. Sliding window analysis was performed using DnaSP v6.12.03 software [47], with a step size of 200 bp and window length of 800 bp, to calculate nucleotide diversity (Pi values) and detect highly variant hotspots in the cp genomes [48]. ...
Article
Full-text available
Quercus is a valuable genus ecologically, economically, and culturally. They are keystone species in many ecosystems. Species delimitation and phylogenetic studies of this genus are difficult owing to frequent hybridization. With an increasing number of genetic resources, we will gain a deeper understanding of this genus. In the present study, we collected four Quercus section Cyclobalanopsis species (Q. poilanei, Q. helferiana, Q. camusiae, and Q. semiserrata) distributed in Southeast Asia and sequenced their complete genomes. Following analysis, we compared the results with those of other species in the genus Quercus. These four chloroplast genomes ranged from 160,784 bp (Q. poilanei) to 161,632 bp (Q. camusiae) in length, with an overall guanine and cytosine (GC) content of 36.9%. Their chloroplast genomic organization and order, as well as their GC content, were similar to those of other Quercus species. We identified seven regions with relatively high variability (rps16, ndhk, accD, ycf1, psbZ—trnG-GCC, rbcL—accD, and rpl32—trnL-UAG) which could potentially serve as plastid markers for further taxonomic and phylogenetic studies within Quercus. Our phylogenetic tree supported the idea that the genus Quercus forms two well-differentiated lineages (corresponding to the subgenera Quercus and Cerris). Of the three sections in the subgenus Cerris, the section Ilex was split into two clusters, each nested in the other two sections. Moreover, Q. camusiae and Q. semiserrata detected in this study diverged first in the section Cyclobalanopsis and mixed with Q. engleriana in the section Ilex. In particular, 11 protein-coding genes (atpF, ndhA, ndhD, ndhF, ndhK, petB, petD, rbcL, rpl22, ycf1, and ycf3) were subjected to positive selection pressure. Overall, this study enriches the chloroplast genome resources of Quercus, which will facilitate further analyses of phylogenetic relationships in this ecologically important tree genus.
... The organization and CDS order (Table S1) of these cp. genomes were highly identical and similar to those of other plants in Apioideae [21][22][23]. ...
Article
Full-text available
Background: The Sinodielsia clade of the subfamily Apioideae (Apiacieae) was established in 2008, and it is composed of 37 species from 17 genera. Its circumscription is still poorly delimited and unstable, and interspecific relationships in the clade lack comprehensive analysis. Chloroplast (cp.) genomes provide valuable and informative data sources for evolutionary biology and have been widely used in studies on plant phylogeny. To infer the phylogenetic history of the Sinodielsia clade, we assembled complete cp. genomes of 39 species and then performed phylogenetic analysis based on these cp. genome sequence data combined with 66 published cp. genomes from 16 genera relative to the Sinodielsia clade. Results: These 39 newly assembled genomes had a typical quadripartite structure with two inverted repeat regions (IRs: 17,599-31,486 bp) separated by a large single-copy region (LSC: 82,048-94,046 bp) and a small single-copy region (SSC: 16,343-17,917 bp). The phylogenetic analysis showed that 19 species were clustered into the Sinodielsia clade, and they were divided into two subclades. Six mutation hotspot regions were detected from the whole cp. genomes among the Sinodielsia clade, namely, rbcL-accD, ycf4-cemA, petA-psbJ, ycf1-ndhF, ndhF-rpl32 and ycf1, and it was found that ndhF-rpl32 and ycf1 were highly variable in the 105 sampled cp. genomes. Conclusion: The Sinodielsia clade was subdivided into two subclades relevant to geographical distributions, except for cultivated and introduced species. Six mutation hotspot regions, especially ndhF-rpl32 and ycf1, could be used as potential DNA markers in the identification and phylogenetic analyses of the Sinodielsia clade and Apioideae. Our study provided new insights into the phylogeny of the Sinodielsia clade and valuable information on cp. genome evolution in Apioideae.
... In this study, we conducted comprehensively comparative analyses for the plastomes of P. nanum, P. violaceum, and Ligusticopsis species. The thirteen plastomes showed typical quadripartite structures, including a pair of inverted repeat regions divided by a large single-copy region and a small single-copy region, which is the same as the other plastomes of Apiaceae [7,8,19,[36][37][38][39][40][41]. Although gene loss and rearrangement have been reported in the plastomes of Apiaceae [19,38,39], the gene content and order in the thirteen studied plastomes were identical. ...
Article
Full-text available
Peucedanum nanum and P. violaceum are recognized as members of the genus Peucedanum because of their dorsally compressed mericarps with slightly prominent dorsal ribs and narrowly winged lateral ribs. However, these species are not similar to other Peucedanum taxa but resemble Ligusticopsis in overall morphology. To check the taxonomic positions of P. nanum and P. violaceum, we sequenced their complete plastid genome (plastome) sequences and, together with eleven previously published Ligusticopsis plastomes, performed comprehensively comparative analyses. The thirteen plastomes were highly conserved and similar in structure, size, GC content, gene content and order, IR borders, and the patterns of codon bias, RNA editing, and simple sequence repeats (SSRs). Nevertheless, twelve mutation hotspots (matK, ndhC, rps15, rps8, ycf2, ccsA-ndhD, petN-psbM, psbA-trnK, rps2-rpoC2, rps4-trnT, trnH-psbA, and ycf2-trnL) were selected. Moreover, both the phylogenetic analyses based on plastomes and on nuclear ribosomal DNA internal transcribed spacer (ITS) sequences robustly supported that P. nanum and P. violaceum nested in Ligusticopsis, and this was further confirmed by the morphological evidence. Hence, transferring P. nanum and P. violaceum into Ligusticopsis genus is reasonable and convincing, and two new combinations are presented.
Article
Full-text available
Melanosciadium is considered a monotypic genus and is also endemic to the southwest of China. No detailed phylogenetic studies or plastid genomes have been identified in Melanosciadium. In this study, the plastid genome sequence and nrDNA sequence were used for the phylogenetic analysis of Melanosciadium and its related groups. Angelica tsinlingensis was previously considered a synonym of Hansenia forbesii. Similarly, Ligusticum angelicifolium was previously thought to be the genus Angelica or Ligusticopsis. Through field observations and morphological evidence, we believe that the two species are more similar to M. pimpinelloideum in leaves, umbel rays, and fruits. Meanwhile, we found a new species from Anhui Province (eastern China) that is similar to M. pimpinelloideum and have named it M. Jinzhaiensis. We sequenced and assembled the complete plastid genomes of these species and another three Angelica species. The genome comparison results show that M. pimpinelloideum, A. tsinlingensis, Ligusticum angelicifolium, and M. jinzhaiensis have similarities to each other in the plastid genome size, gene number, and length of the LSC and IR regions; the plastid genomes of these species are distinct from those of the Angelica species. In addition, we reconstruct the phylogenetic relationships using both plastid genome sequences and nrDNA sequences. The phylogenetic analysis revealed that A. tsinlingensis, M. pimpinelloideum, L. angelicifolium, and M. jinzhaiensis are closely related to each other and form a monophyletic group with strong support within the Selineae clade. Consequently, A. tsinlingensis and L. angelicifolium should be classified as members of the genus Melanosciadium, and suitable taxonomical treatments have been proposed. Meanwhile, a comprehensive description of the new species, M. jinzhaiensis, is presented, encompassing its habitat environment and detailed morphological traits.
Article
Full-text available
Chamaesium H. Wolff (Apiaceae, Apioideae) is a small genus mainly distributed in the Hengduan Mountains and the Himalayas. Ten species of Chamaesium have been described and nine species are distributed in China. Recent advances in molecular phylogenetics have revolutionized our understanding of Chinese Chamaesium taxonomy and evolution. However, an accurate phylogenetic relationship in Chamaesium based on the second-generation sequencing technology remains poorly understood. Here, we newly assembled nine plastid genomes from the nine Chinese Chamaesium species and combined these genomes with eight other species from five genera to perform a phylogenic analysis by maximum likelihood (ML) using the complete plastid genome and analyzed genome structure, GC content, species pairwise Ka/Ks ratios and the simple sequence repeat (SSR) component. We found that the nine species’ plastid genomes ranged from 152,703 bp (C. thalictrifolium) to 155,712 bp (C. mallaeanum), and contained 133 genes, 34 SSR types and 585 SSR loci. We also found 20,953–21,115 codons from 53 coding sequence (CDS) regions, 38.4–38.7% GC content of the total genome and low Ka/Ks (0.27–0.43) ratios of 53 aligned CDS. These results will facilitate our further understanding of the evolution of the genus Chamaesium.
Article
Full-text available
Bupleurum L. (Apiaceae) is a perennial and herbal genus, most species of which have high medicinal value. However, few studies have been performed using plastome data in this genus, and the phylogenetic relationships have always been controversial. In this study, the plastid genomes of Bupleurum chinense and Bupleurum commelynoideum were sequenced, and their gene content, order, and structure were counted and analyzed. The only three published Bupleurum species (B. boissieuanum, B. falcatum, and B. latissimum) and other fifteen allied species were selected to conduct a series of comparative and phylogenetic analyses. The genomes of B. chinense and B. commelynoideum were 155,869 and 155,629 bp in length, respectively, both of which had a typical quadripartite structure. The genome length, structure, guanine and cytosine (GC) content, and gene distribution were highly similar to the other three Bupleurum species. The five Bupleurum species had nearly the same codon usages, and eight regions (petN-psbM, rbcL-accD, ccsA-ndhD, trnK(UUU)-rps16, rpl32-trnL(UAG)-ccsA, petA-psbJ, ndhF-rpl32, and trnP(UGG)-psaJ-rpl33) were found to possess relatively higher nucleotide diversity, which may be the promising DNA barcodes in Bupleurum. Phylogenetic analysis revealed that all Bupleurum species clustered into a monophyletic clade with high bootstrap support and diverged after the Chamaesium clade. Overall, our study provides new insights into DNA barcoding and phylogenetic relationship between Bupleurum and its related genera, and will facilitate the population genomics, conservation genetics, and phylogenetics of Bupleurum in Apiaceae.
Article
Full-text available
Background and aims: The long-standing controversy in Apioideae concerns relationships among the major lineages of this subfamily, which led to comprehensive study for fruits and evolutionary history of the whole subfamily cannot be implemented accurately. Here we attempt to use single copy genes (SCGs) generated from transcriptome datasets to generate a reliable species tree and explore the evolutionary history of Apioideae. Methods: Total of 3351 SCGs were generated from 27 transcriptome datasets and one genome, and further used for phylogenetic analysis using coalescent-based methods. Morphology and anatomy of the fruits were studied in combination with the species tree. Eleven SCGs were screened out for dating analysis with two fossils selected for calibration. Key results: A highly supported species tree were generated with topology [Chamaesieae, (Bupleureae, (Pleurospermeae, (Physospermopsis Clade, (Group C, (Group A, Group B)))))] differing from the previous. Daucinae and Torilidinae skipped out of tribe Scandiceae and existed as sister groups to Acronema Clade. Five branches (I~V) of species tree showed low quartet supports but strong local posterior probabilities. Dating analysis suggested that Apioideae originated around 56.64 Ma (95 % HPD, 45.18~73.53 Ma). Conclusions: This study resolves a controversial phylogenetic relationship in Apioideae based on 3351 SCGs and coalescent-based species tree estimate methods. Gene trees that contributed to the species tree may undergoing rapid evolutionary divergence and incomplete lineage sorting. Fruits of the Apioideae might evolve in two directions, anemochorous and hydrochorous, with epizoochorous as a derived mode. Molecular and morphological evidence suggested Daucinae and Torilidinae should be restored as tribe. Our results provide new insights into the morphological evolution of this subfamily, which may contribute to a better understanding of species diversification in Apioideae. Molecular dating analysis suggests that uplift of the QTP and climate changes probably drive rapid radiation speciation and diversification of Apioideae in the QTP region.
Article
Full-text available
Meeboldia yunnanensis Wolff (Apiaceae) is a perennial species naturally distributed in Yunnan and Xizang. The complete chloroplast genome sequence of M. yunnanensis was generated by de novo assembly using whole-genome next-generation sequencing data. The complete chloroplast genome of M. yunnanensis was 154,865 bp in total sequence length and divided into four distinct regions: small single-copy region (17,370 bp), large single-copy region (84,641 bp), and a pair of inverted repeat regions (26,427 bp). The genome annotation displayed a total of 130 genes, including 85 protein-coding genes, 37 tRNA genes, and 8 rRNA genes. Phylogenetic analysis with the reported chloroplast genomes revealed that M. yunnanensis has close relationship to Pterygopleurum neurophyllum.
Article
Full-text available
Background Tetrataenium candicans is a traditional Chinese folk herbal medicine used in the treatment of asthma and rheumatic arthritis. Alongside several Tordyliinae species with fleshy roots, it is also regarded as a substitute for a Chinese material medicine called ‘Danggui’. However, a lack of sufficient sampling and genomic information has impeded species identification and the protection of wild resources. Methods The complete chloroplast genomes of T. candicans from two populations, Tetrataenium yunnanense and Semenovia transilliensis , were assembled from two pipelines using data generated from next generation sequencing (NGS). Pseudogenes, inverted repeats (IRs) and hyper-variable regions were located by Geneious 11.1.5. Repeat motifs were searched using MISA and REPuter. DNA polymorphism and segment screening were processed by DNAsp5, and PCR product was sequenced with Sanger’s sequencing method. Phylogeny was inferred by MEGA 7.0 and PhyML 3.0. Results The complete chloroplast genomes of T. candicans from two populations, T. yunnanense and S. transilliensis , were 142,261 bp, 141,985 bp, 142,714 bp and 142,145 bp in length, respectively, indicating conservative genome structures and gene categories. We observed duplications of trnH and psbA caused by exceptional contractions and expansions of the IR regions when comparing the four chloroplast genomes with previously published data. Analyses on DNA polymorphism located 29 candidate cp DNA barcodes for the authentication of ‘Danggui’ counterfeits. Meanwhile, 34 hyper-variable markers were also located by the five Tordyliinae chloroplast genomes, and 11 of them were screened for population genetics of T. candicans based on plastome information from two individuals. The screening results indicated that populations of T.candicans may have expanded. Phylogeny inference on Apiaceae species by CDS sequences showed most lineages were well clustered, but the five Tordyliinae species failed to recover as a monophyletic group, and the phylogenetic relationship between tribe Coriandreae, tribe Selineae, subtribe Tordyliinae and Sinodielsia clade remains unclear. Discussion The four chloroplast genomes offer valuable information for further research on species identification, cp genome structure, population demography and phylogeny in Apiaceae subfamily Apioideae.
Article
Full-text available
Ampelopsis humulifolia (A. humulifolia) and Ampelopsis japonica (A. japonica), which belong to the family Vitaceae, are valuably used as medicinal plants. The chloroplast (cp) genomes have been recognized as a convincing data for marker selection and phylogenetic studies. Therefore, in this study we reported the complete cp genome sequences of two Ampelopsis species. Results showed that the cp genomes of A. humulifolia and A. japonica were 161,724 and 161,430 bp in length, respectively, with 37.3% guanine-cytosine (GC) content. A total of 114 unique genes were identified in each cp genome, comprising 80 protein-coding genes, 30 tRNA genes, and 4 rRNA genes. We determined 95 and 99 small sequence repeats (SSRs) in A. humulifolia and A. japonica, respectively. The location and distribution of long repeats in the two cp genomes were identified. A highly divergent region of psbZ (Photosystem II reaction center protein Z) -trnG (tRNA-Glycine) was found and could be treated as a potential marker for Vitaceae, and then the corresponding primers were designed. Additionally, phylogenetic analysis showed that Vitis was closer to Tetrastigma than Ampelopsis. In general, this study provides valuable genetic resources for DNA barcoding marker identification and phylogenetic analyses of Ampelopsis.
Article
Full-text available
Organellar (plastid and mitochondrial) genomes play an important role in resolving phylogenetic relationships, and next-generation sequencing technologies have led to a burst in their availability. The ongoing massive sequencing efforts require software tools for routine assembly and annotation of organellar genomes as well as their display as physical maps. OrganellarGenomeDRAW (OGDRAW) has become the standard tool to draw graphical maps of plastid and mitochondrial genomes. Here, we present a new version of OGDRAW equipped with a new front end. Besides several new features, OGDRAW now has access to a local copy of the organelle genome database of the NCBI RefSeq project. Together with batch processing of (multi-)GenBank files, this enables the user to easily visualize large sets of organellar genomes spanning entire taxonomic clades. The new OGDRAW server can be accessed at https://chlorobox.mpimp-golm.mpg.de/OGDraw.html.
Article
Full-text available
Haplosphaera phaea Handel-Mazzetti (Apiaceae) is an endangered species naturally distributed in China. The complete chloroplast genome of H. pheae was determined in this study. The chloroplast genome of H. pheae was 157,271 bp in length and divided into four distinct regions, such as large single copy region was 86,489 bp, small single copy (SSC) region was 17,892 bp, and a pair of inverted repeat regions is 26,446 bp in each length. The chloroplast genome detected a total of 130 genes including 85 protein-coding genes, 37 tRNA genes, and eight rRNA genes. Phylogenetic analysis with the reported chloroplast genomes revealed that H. pheae is nested in the genus Hansenia.
Article
Hansenia pinnatiinvolucellata and H. weberbaueriana have been recorded to co-occur in Hengduan Mountains and adjacent regions. However, there were no reliable distinguishing characters to identify them. In our research, we explored key morphological characters through field work, specimen examinations and original literature reviews and reconstructed the phylogeny based on nuclear ITS regions, to determine the relationship between H. pinnatiinvolucellata and H. weberbaueriana. Morphological results indicated that the key, unique morphological characters (entire or pinnati bracteoles longer than pedicels and fruit with 3 ribs) of H. pinnatiinvolucellata were occurs simultaneously in some populations of H. weberbaueriana. Phylogenetic analyses also reveal that H. pinnatiinvolucellata and H. weberbaueriana are nested together. So, H. pinnatiinvolucellata should be a synonym of H. weberbaueriana.