ArticlePDF Available

Comparative Mitogenomic Analysis of Two Longhorn Beetles (Coleoptera: Cerambycidae: Lamiinae) with Preliminary Investigation into Phylogenetic Relationships of Tribes of Lamiinae

MDPI
Insects
Authors:

Abstract and Figures

Simple Summary Many species of Cerambycidae are important pests in the agriculture, forestry, and fruit industries, and which have research significance. The subfamily Lamiinae is the most taxonomically diverse subfamily of Cerambycidae, but relationships between the tribes of Lamiinae are still unresolved. In order to provide a new perspective on the phylogenetic relationships among the tribes of Lamiinae, the mitogenomes of two species representing two tribes, Agapanthia amurensis (Agapanthiini) and Moechotypa diphysis (Ceroplesini), were sequenced. We present annotated, complete mitogenomes of these two species, and the results of a comparative analysis of both mitogenomes. The two new mitogenomes were found to be highly conservative, as found in other Cerambycidae. We also reconstructed the phylogenetic trees using mitogenomes of 38 species/subspecies of Lamiinae. Overall, this study explores the phylogenetic position between some tribes based on mitogenomic data and provides a further basis for studying the evolution of Lamiinae. Abstract The subfamily Lamiinae is the most taxonomically diverse subfamily of Cerambycidae, but relationships between tribes of Lamiinae are still unresolved. In order to study the characteristics of the mitogenomes of Lamiinae and the tribal-level phylogenetic relationships, we sequenced the mitogenomes of two species representing two tribes, Agapanthia amurensis (Agapanthiini) and Moechotypa diphysis (Ceroplesini), with a total length of 15,512 bp and 15,493 bp, respectively. The gene arrangements of these two new mitogenomes were consistent with the inferred ancestral insect mitogenomes. Each species contained 37 typical mitochondrial genes and a control region (A + T-rich region), including 13 protein-coding genes (PCGs), 22 transfer RNA genes (tRNAs), and two ribosomal RNA genes (rRNAs). All PCGs initiated with the standard start codon ATN, and terminated with the complete stop codons of TAA and TAG, or incomplete stop codon T. All tRNAs could be folded into a clover-leaf secondary structure except for trnS1, in which the dihydrouridine (DHU) arm was reduced. Moreover, we studied the phylogenetic relationships between some tribes of Lamiinae based in mitochondrial PCGs in nucleotides; our results show that the relationships were as follows: (Onciderini + ((Apomecynini + Acanthocinini) + ((Ceroplesini + Agapanthiini) + ((Mesosini + Pteropliini) + ((Dorcaschematini + (Saperdini 1 + (Phytoeciini + Saperdini 2))) + (Batocerini + Lamiini)))))).
This content is subject to copyright.
insects
Article
Comparative Mitogenomic Analysis of Two Longhorn Beetles
(Coleoptera: Cerambycidae: Lamiinae) with Preliminary
Investigation into Phylogenetic Relationships of Tribes
of Lamiinae
Yifang Ren 1,† , Huanhuan Lu 2, , Longyan Chen 1, Simone Sabatelli 3, Chaojie Wang 1, Guanglin Xie 1,
Ping Wang 1, Meike Liu 1,* , Wenkai Wang 1, * and Paolo Audisio 3


Citation: Ren, Y.; Lu, H.; Chen, L.;
Sabatelli, S.; Wang, C.; Xie, G.; Wang,
P.; Liu, M.; Wang, W.; Audisio, P.
Comparative Mitogenomic Analysis
of Two Longhorn Beetles (Coleoptera:
Cerambycidae: Lamiinae) with
Preliminary Investigation into
Phylogenetic Relationships of Tribes
of Lamiinae. Insects 2021,12, 820.
https://doi.org/10.3390/
insects12090820
Academic Editor: Teiji Sota
Received: 23 August 2021
Accepted: 8 September 2021
Published: 12 September 2021
Publisher’s Note: MDPI stays neutral
with regard to jurisdictional claims in
published maps and institutional affil-
iations.
Copyright: © 2021 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 (https://
creativecommons.org/licenses/by/
4.0/).
1Institute of Entomology, College of Agriculture, Yangtze University, Jingzhou 434025, China;
201973050@yangtzeu.edu.cn (Y.R.); clylyly@126.com (L.C.); 202071670@yangtzeu.edu.cn (C.W.);
xieguanglin@yangtzeu.edu.cn (G.X.); wangping1992@yangtzeu.edu.cn (P.W.)
2Chongqing Key Laboratory of Vector Insects, College of Life Sciences, Chongqing Normal University,
Chongqing 401331, China; 2019110513046@stu.cqnu.edu.cn
3Dipartimento di Biologia e Biotecnologie “Charles Darwin”, Sapienza Universitàdi Roma,
Viale dell’Università32, I-00185 Rome, Italy; simone.sabatelli@uniroma1.it (S.S.);
paolo.audisio@uniroma1.it (P.A.)
*Correspondence: liumk2009@126.com (M.L.); wwk@yangtzeu.edu.cn (W.W.)
These authors have equally contributed in this study.
Simple Summary:
Many species of Cerambycidae are important pests in the agriculture, forestry, and
fruit industries, and which have research significance. The subfamily Lamiinae is the most taxonomically
diverse subfamily of Cerambycidae, but relationships between the tribes of Lamiinae are still unresolved.
In order to provide a new perspective on the phylogenetic relationships among the tribes of Lamiinae, the
mitogenomes of two species representing two tribes, Agapanthia amurensis (Agapanthiini) and Moechotypa
diphysis (Ceroplesini), were sequenced. We present annotated, complete mitogenomes of these two
species, and the results of a comparative analysis of both mitogenomes. The two new mitogenomes
were found to be highly conservative, as found in other Cerambycidae. We also reconstructed the
phylogenetic trees using mitogenomes of 38 species/subspecies of Lamiinae. Overall, this study
explores the phylogenetic position between some tribes based on mitogenomic data and provides a
further basis for studying the evolution of Lamiinae.
Abstract:
The subfamily Lamiinae is the most taxonomically diverse subfamily of Cerambycidae,
but relationships between tribes of Lamiinae are still unresolved. In order to study the characteristics
of the mitogenomes of Lamiinae and the tribal-level phylogenetic relationships, we sequenced
the mitogenomes of two species representing two tribes, Agapanthia amurensis (Agapanthiini) and
Moechotypa diphysis (Ceroplesini), with a total length of 15,512 bp and 15,493 bp, respectively. The
gene arrangements of these two new mitogenomes were consistent with the inferred ancestral
insect mitogenomes. Each species contained 37 typical mitochondrial genes and a control region
(A + T-rich region), including 13 protein-coding genes (PCGs), 22 transfer RNA genes (tRNAs), and
two ribosomal RNA genes (rRNAs). All PCGs initiated with the standard start codon ATN, and
terminated with the complete stop codons of TAA and TAG, or incomplete stop codon T. All tRNAs
could be folded into a clover-leaf secondary structure except for trnS1, in which the dihydrouridine
(DHU) arm was reduced. Moreover, we studied the phylogenetic relationships between some tribes
of Lamiinae based in mitochondrial PCGs in nucleotides; our results show that the relationships
were as follows: (Onciderini + ((Apomecynini + Acanthocinini) + ((Ceroplesini + Agapanthiini)
+ ((Mesosini + Pteropliini) + ((Dorcaschematini + (Saperdini 1 + (Phytoeciini + Saperdini 2))) +
(Batocerini + Lamiini)))))).
Keywords: molecular; mitochondrial genome; genome structure; phylogenetic analysis
Insects 2021,12, 820. https://doi.org/10.3390/insects12090820 https://www.mdpi.com/journal/insects
Insects 2021,12, 820 2 of 13
1. Introduction
The family Cerambycidae is one of the largest families of the superfamily Chrysomeloidea
(Coleoptera: Polyphaga), consisting of a little less than 40,000 species, among which Lami-
inae have about 20,000 species [
1
]. Thus, some species of Cerambycidae are important
pests in the agriculture, forestry, and fruit industries. Some are also important quarantine
pests. For example, some Monochamus species are the main transmission vectors of the
pine wood nematode Bursaphelenchus xylophilus (Aphelenchoidae) [
2
,
3
]. However, adults
of many species of Lepturinae in the Cerambycidae family have flower-visiting behaviors
and help plants to pollinate [
4
]. All in all, Cerambycidae is believed to have great potential
for applications in forest health and ecological biodiversity assessment [57].
As the most taxonomically diverse subfamily in Cerambycidae, the monophyly of
Lamiinae is supported based on morphology and/or molecule studies [
8
13
]. However, the
phylogenetic relationships of tribes among Lamiinae are controversial. Latreille (1825) pro-
posed a modern classification of Cerambycidae, first using the term Lamiaires, which later
became Lamiinae [
14
]. Blanchard (1845) proposed names for suprageneric groups of Lami-
inae and recognized seven groups (Acanthocinites, Lamiites, Mesosites, Petrognathites,
Saperdites, Stellognathites and Tetraophtalmites) [
15
]. Then, Thomson (1860) proposed an
eight-rank taxonomic system (in French: Famille, Tribu, Sous-Tribu, Groupe, Sous-groupe,
Division, Genre, Espèce), and recorded 17 groups of Lamiinae, which later increased to
33 groups in 1864 [
16
,
17
]. This laid a foundation for the later taxonomic study of tribes
among Lamiinae. Aurivillius (1922, 1923) compiled a catalogue of Lamiinae and divided it
into 96 tribes, which was the first comprehensive summary of Lamiinae [
18
,
19
]. Breuning
(1958–1969) significantly reduced the tribes to 58 by synonymizing some controversial
tribes [
20
31
], but this system was not fully accepted by subsequent researchers. More re-
cently, Bouchard et al. (2011) synthesized the data of all known extant and fossil Coleoptera
for the first time, and recognized 80 tribes of Lamiinae [
32
]. Souza et al. (2020) performed a
study on the tribal classification of Lamiinae by molecular phylogenetic assessment [
13
].
They confirmed the monophyly of Lamiinae, and suggested some synonyms for its tribes
based on the fragments of two mitochondrial genes (cytochrome c oxidase subunit 1 and
large ribosomal RNA subunit) and three nuclear genes (wingless, carbamoyl-phosphate
synthase domain of the CAD locus, and large ribosomal rRNA subunit).
Insect mitogenomes present unique features, such as maternal inheritance, low molec-
ular weight, low recombination level, and fast evolutionary rate. Thus, they are widely
used as molecular markers in studies of classification, genetic evolution, and phylogenetic
analysis [
33
35
]. Insect mitogenomes consist of 37 genes, including 13 protein-coding genes
(PCGs), 22 transfer RNA genes (tRNAs), two ribosomal RNA genes (rRNAs), and a control
region (A + T-rich region) [
33
]. With the rapid development of next-generation sequencing,
about 40,000 mitogenomes of animals have been published in the NCBI datasets [
36
]. How-
ever, few studies have been proposed based on molecular methods at the tribal level of
Lamiinae. In this study, mitogenomes of two species of Lamiinae representing two tribes,
Agapanthia amurensis Kraatz, 1879 (Agapanthiini) and Moechotypa diphysis (Pascoe, 1871)
(Ceroplesini), were sequenced and analyzed. Of these, M. diphysis was the first complete
mitogenome of Ceroplesini. In addition, we used the available and annotated mitogenomes
from NCBI and two newly sequenced mitogenomes to infer the phylogenetic relationships
between some tribes among Lamiinae.
2. Materials and Methods
2.1. Sample Preparation and DNA Extraction
Specimens of Agapanthia amurensis Kraatz, 1879 and Moechotypa diphysis
(Pascoe, 1871)
were collected from Enshi Tujia and Miao Autonomous Prefecture, Hubei Province, China
(May 2020). The latitude and longitude of the collection sites are 30
36
0
18.6” N and
110
4
0
6.8” E. All specimens were identified by Prof. Guanglin Xie based on morpho-
logical characteristics. All specimens were immediately preserved in 100% ethanol and
stored at
20
C in the Entomological Museum of Yangtze University
(No. A.amurensis
Insects 2021,12, 820 3 of 13
YZU20200507134 and No. M.diphysis YZU20200507216). Then, total genomic DNA was
extracted from the thoracic muscle tissues using a DNeasy DNA Blood & Tissue Kit
(Qiagen, Beijing, China).
2.2. Sequence Analysis
Two mitogenome sequences were generated using the Illumina HiSeq platform with
paired ends of 2
×
251 bp at Biomarker Technologies Co. Ltd. (Beijing, China). Raw
reads were filtered, and quality was assessed using Fast-QC (http://www.bioinformatics.
babraham.ac.uk/projects/fastqc, accessed on 01 July 2021) based on Q20 (>95%) and Q30
(>90%). After quality trimming, the reads were assembled by Geneious v8.1.3 (Biomatters,
Auckland, New Zealand) [
37
] with default parameters and using the mitogenome of Aromia
bungii (Faldermann, 1835) (GenBank accession no. MT371041) as reference [
38
]. All PCGs
of two mitogenomes were identified based on finding the open reading frames (ORFs) and
translated by Geneious v8.1.3 according to the invertebrate mitochondrial genetic code. The
secondary structures of the tRNAs were identified by the MITOS Web Server [
39
] and drawn
in Adobe Illustrator CC2019. The positions of rRNAs and control region were predicted
by adjacent genes and homology alignment with A.bungii. The mitogenome maps were
drawn using CGView Server [
40
]. Nucleotide composition, AT or GC skews, and relative
synonymous codon usage (RSCU) were analyzed by PhyloSuite v1.2.2 [
41
]. The tandem
repeats of the control region were predicted by the Tandem Repeats Finder online server
(http://tandem.bu.edu/trf/trf.basic.submit.html, accessed on 01 July 2021) [42].
2.3. Phylogenetic Analysis
Two newly sequenced mitogenomes and 36 species/subspecies of Lamiinae were
selected from the NCBI, representing 12 tribes (Acanthocinini, Agapanthiini, Apome-
cynini, Batocerini, Ceroplesini, Dorcaschematini, Lamiini, Mesosini, Onciderini, Phy-
toeciini, Pteropliini, Saperdini). In addition, two species (Chrysomela vigintipunctata and
Plagiodera versicolora
) of Chrysomelidae were selected as outgroups (Supplementary Ma-
terials Table S1). The 13PCGs were aligned by MAFFT v7.0 [
43
]. Then, poorly aligned
positions and high divergence regions were removed by Gblocks v0.91b in PhyloSuite.
The potential index of substitution saturation (Iss) of each codon position of each nucleic
acid sequence was analyzed by DAMBE v7.2.1 based on default parameters. [
44
]. Phy-
logenetic analyses were conducted using two datasets: 13PCGs (nucleotides sequences
for protein-coding genes; including all codon positions) and 13PCGs_AA (amino acids of
13PCGs). The phylogenetic analyses were reconstructed by the Bayesian inference (BI) and
the Maximum likelihood (ML) methods based on these two datasets. The best evolutionary
model of BI was inferred by PartitionFinder v2.1.1 [
45
] using the greedy search algorithm
with branch lengths linked and Bayesian information criterion (BIC) in PhyloSuite. The
best fit model of ML was selected using ModelFind in IQ-TREE. ML analysis was con-
ducted by IQ-TREE [
46
], and node supports were computed via 1000 ultrafast bootstrap
replicates. BI analysis was conducted by MrBayes v3.2.6 [
47
] under four Markov chain
Monte Carlo (MCMC) chains of 1 million generations twice, sampled every 1000 genera-
tions, with the first 25% of generations removed as burn-in. MCMC analysis was stopped
when the average standard deviation of the split frequency was below 0.01. Additionally,
PhyloBayes analysis based on the CAT-GTR model was conducted by PhyloBayes v1.5a
on CIPRES [
12
,
48
]. Two chains were run until the likelihood had satisfactorily converged
(maxdiff < 0.2, minimum effective size > 50). The phylogenetic tree was displayed and
edited by Tree of Life (iTOL, http://itol.embl.de, accessed on 01 July 2021) [49].
3. Results and Discussion
3.1. General Features of the Mitogenomes of Agapanthia amurensis and Moechotypa diphysis
The complete mitogenomes of Agapanthia amurensis (GenBank accession number:
MW617354) and Moechotypa diphysis (MW617356) were 15,512 bp and 15,493 bp in length,
respectively (Table S1). Both mitogenomes included 37 typical mitogenomic genes and a
Insects 2021,12, 820 4 of 13
control region. In these two mitogenomes, most genes (nine PCGs and 14 tRNAs) were
concentrated on the J strand, and others (four PCGs, eight tRNAs, and two rRNAs) on the
N strand (Figures 1a and 2a, Table S2). Aside from the control region, six intergenic regions
were found in the mitogenomes of both A.amurensis (35 bp total) and M.diphysis (31 bp
total), and the longest region was detected between trnS2 and nad1. Sixteen overlapping
regions were found in the mitogenomes of both A.amurensis (50 bp total) and M.diphysis
(45 bp total) (Table S3). The base composition of A.amurensis was A (38.5%), T (36.7%),
G (9.3%), and C (15.5%), and the base composition of M.diphysis was A (37.6%), T (38.0%),
G (9.5%), and C (14.9%). Both sequences displayed a high AT nucleotide bias, with
A + T%
of the whole sequence of 75.2% in A.amurensis and 75.6% in M.diphysis; these were similar
to other sequenced Lamiinae species (Table S4). The composition skew analysis showed
that the two new mitogenomes presented a negative GC skew. For AT skew, the value
for A.amurensis was 0.024 and for M.diphysis was
0.005 (Table S4). Circular maps of the
mitogenomes of these two new sequences were visualized in Figures 1a and 2a.
Figure 1. Complete mitogenome of Agapanthia amurensis. (A) Circular map. (B) Organization of the
A + T-rich regions.
Insects 2021,12, 820 5 of 13
Figure 2.
Complete mitogenome of Moechotypa diphysis. (
A
) Circular map. (
B
) Organization of the
A + T-rich regions.
3.2. Genome Structure
3.2.1. Protein-Coding Genes
The PCGs ranged from 156 bp (atp8) to 1714 bp (nad5) in A.amurensis, and 156 bp
(atp8) to 1720 bp (nad5) in M.diphysis (Table S4). A.amurensis and M.diphysis exhibited
similar start and stop codons. All PCGs started with the typical ATN codon, and most
PCGs ended with TAA or TAG, while three PCGs (cox1,cox2, and nad5) in A.amurensis and
four PCGs (cox1,cox2,nad4, and nad5) in M.diphysis ended with incomplete stop codon T.
In all 38 species/subspecies of Lamiinae, the termination TAA occurred more frequently
than TAG (Table S2).
Research on relative synonymous codon usage (RSCU) showed that the frequency of
A or T is higher than that of G or C in the third codon position (Figure 3). For example,
the third codon position of the seven most used codons (TTA, TCT, AGA, CCT, GTA, ACT,
and TGT) in the mitogenome of M.diphysis was A or T, whereas codons with G or C in the
third position (GCG, CCG, CGC, and GGC) were seldom presented in the mitogenome of
M.diphysis (Table S5).
3.2.2. Transfer and Ribosomal RNA Genes
Both new mitogenomes included 22 typical tRNAs genes. The size of these genes
ranged from 64 bp (trnE,trnG,trnL1, and trnP) to 70 bp (trnK, and trnY) in A. amurensis,
and from 61 bp (trnC) to 70 bp (trnK) in M. diphysis (Table S2). The whole tRNA region of
these two mitogenomes was 1462 bp in A. amurensis and 1435 bp in M. diphysis
(Table S4)
.
Insects 2021,12, 820 6 of 13
Except for trnS1, which lacked a dihydrouracil (DHU) arm and formed a simple loop,
other tRNAs could be folded into the classic clover-leaf secondary structure (Figure 4a,b).
This phenomenon often occurs in the mitogenomes of Cerambycidae [
50
,
51
]. Based on the
secondary structure of tRNAs, we found five types of mismatched bases (U-U, G-U, A-G,
A-C, and A-A) in A.amurensis, and three types (U-U, G-U, and A-G) in M.diphysis.
Figure 3.
Relative synonymous codon usage (RSCU) of PCGs in mitogenomes of Lamiinae. X- and Y-axis represent codon
type and species name, respectively.
Int. J. Environ. Res. Public Health 2021, 18, x. https://doi.org/10.3390/xxxxx www.mdpi.com/journal/ijerph
(A)
Figure 4. Cont.
Insects 2021,12, 820 7 of 13
(B)
Figure 4.
(
A
) Secondary structures of tRNAs in mitogenome of Agapanthia amurensis; bonds of A-U, G-C are marked with
purple and red lines, and G-U mismatched bases with green lines and solid dots. (
B
) Secondary structures of tRNAs in
mitogenome of Moechotypa diphysis; bonds of A-U, G-C are marked with purple and red lines, and G-U mismatched bases
with green lines and solid dots.
There were two rRNAs in both new mitogenome sequences: rrnL (16SrRNA) was
located between trnL and trnV, and rrnS (12SrRNA) was located between trnV and the
control region. The length of rrnL was 1281 bp in A.amurensis and 1280 bp in M.diphysis,
and the length of rrnS was 781 bp (A.amurensis) and 777 bp (M.diphysis), respectively
(Table S2). These two rRNAs displayed a heavy AT nucleotide bias, with A + T content of
78.7% in A.amurensis and 80.4% in M.diphysis. Both rRNAs showed a negative AT skew
and a positive GC skew in these two new sequences (Table S4).
3.2.3. Control Region
The control region is normally the largest non-coding region in insect mitogenomes,
also known as the A + T-rich region owing to the high AT content. This region was thought
to have a function in regulating the transcription and replication of insect genes [
52
54
].
Insects 2021,12, 820 8 of 13
Tandem repeats in the control region were considered as the sequences that mainly affect the
size of the mitogenome [
52
,
55
]. In this study, the control region of both new mitogenomes
was located between rrnS and trnI, and the size was 857 bp in A.amurensis and 874 bp in
M.diphysis (Table S2). One tandem repeat-like region was found in the control region of
each. One Poly (A) was found in the non-repeat region of each. Three Poly (T) were found
in non-repeat regions of M.diphysis (Figures 1b and 2b); Poly (T) stretch was considered as
an origin of transcription and replication [52].
3.3. Phylogenetic Analysis
The substitution saturation analyses (Table S6) showed that the index of substitu-
tion saturation (Iss) was less than Iss.cSym (p<0.05) regardless of whether datasets were
trimmed by Gblocks. This result suggested that all nucleotide positions of PCGs could pro-
vide useful information for the phylogenetic analysis. Based on the 13PCGs and 13PCGs_AA
datasets, the phylogenetic relationships between the 12 tribes of Lamiinae were inferred
from ML and BI analyses. The best model of the phylogenetic relationship is presented
in Table S7. Inconsistent topologies were generated by ML and BI analyses based on the
13PCGs_AA dataset (Supplementary Materials Figure S1). The positions of some tribes
(Agapanthiini, Acanthocinini, Onciderini, Apomecynini, and Phytoeciini) in ML and BI
trees were different, and there were some clade nodes had low support values
(Figure S1)
.
It may require more mitogenomes and other molecular evidence of Lamiinae to solve
these problems [
56
,
57
]. Except for the position of Psacothea hilaris, the ML (Figure 5a)
and Bl (Figure 5b) analyses based on the 13PCGs dataset produced basically consistent
topologies, and the support value of the BI tree was generally higher than that of the
ML tree. They all indicated that the relationships among the 12 tribes of Lamiinae were:
(Onciderini + ((Apomecynini + Acanthocinini) + ((Ceroplesini + Agapanthiini) + ((Mesosini
+ Pteropliini) + ((Dorcaschematini + (Saperdini 1 + (Phytoeciini + Saperdini 2))) + (Bato-
cerini + Lamiini)))))). Due to the heterogeneity in base composition and evolutionary
rates, phylogenetics research based on mitogenomes is still controversial [
58
60
]. Previ-
ous studies argued that using a CAT-GTR model in PhyloBayes was more suitable than
other methods to reconstruct relationships at the subfamily level and above [
12
,
59
,
60
].
In this study, the PhyloBayes analysis based on the 13PCGs_AA dataset (Figure S2) us-
ing the CAT-GTR model produced almost congruent topologies with the phylogenetic
tree based on the 13PCGs. The only differences were the position of clades Onciderini
and
(Mesosini + Pteropliini)
. Therefore, only the phylogenetic trees constructed using the
13PCGs dataset are shown in this paper (Figure 5a,b).
In this study, the monophyly of Lamiinae was highly supported, which had been
confirmed in previous studies [
8
13
,
51
,
61
]. We also studied the phylogenetic relationships
between some tribes among Lamiinae, and the monophyly of Lamiini was supported
(BS = 100; PP = 1)
. The boundary of Lamiini is controversial; some researchers have sup-
ported the retention of Lamiini and Monochamini [
32
,
62
,
63
], while others have proposed
that the tribe Monochamini should be included in Lamiini [
64
66
]. Souza et al. (2020)
proposed that Monochamini is a synonym of Lamiini based on the fragments of two mito-
chondrial genes (cox1 and rrnL) and three nuclear genes (Wg,CPS, and LSU) [
13
], which
is also supported by our present result. Moreover, in our study, the monophyly of the
tribe Saperdini was not supported as Phytoeciini was nested within the Saperdini clade
(BS = 76.7; PP = 1). Based on this result, we considered that Phytoeciini should be included
in Saperdini. Souza et al. (2020), who performed the first, relatively dense phylogenetic
systematic assessment of Lamiinae, supported Phytoeciini as a synonym of Saperdini [
13
].
Further, the phylogenetic tree (Figure 3in Zhang et al. (2021)) of 30 species of Lamiinae
based on the nucleotide dataset of the 13PCGs of mitogenomes showed the same result [
61
].
Insects 2021,12, 820 9 of 13
Figure 5. Cont.
Figure 5.
(
A
) Maximum likelihood (ML) phylogenetic tree based on 13PCGs dataset. Numbers on
branches are bootstrap values. (
B
) Bayesian inference (BI) phylogenetic tree based on 13PCGs dataset.
Numbers on branches are Bayesian posterior probabilities.
4. Conclusions
This study describes two complete mitogenomes of Agapanthia amurensis and
Moechotypa diphysis in Lamiinae. The gene arrangements of these new mitogenomes
were consistent with other longhorn beetles. Each species contained 37 typical mitochon-
drial genes and a control region. All tRNAs could be folded into a clover-leaf secondary
structure except for trnS1, in which the dihydrouridine (DHU) arm was reduced. In all
38 species/subspecies of Lamiinae in this study, the termination TAA occurred more than
Insects 2021,12, 820 10 of 13
TAG. Our phylogenetic tree inferences of Lamiinae based on mitogenomes confirm the
monophyly of Lamiinae. Regarding the relationships between the tribes among Lami-
inae, our results support Lamiini being monophyletic and Phytoeciini as a synonym of
Saperdini. In this study, the Bayesian analysis based on the 13PCGs dataset is better than
other analyses, which of course requires more samples and studies to verify. All in all, in
order to better understand the phylogenetic relationships between tribes among Lamiinae,
more mitochondrial datasets and other molecular evidence are still needed. Using more
taxon samples and molecular markers may help to determine the higher-level relationships
among Cerambycidae.
Supplementary Materials:
The following are available online at https://www.mdpi.com/article/
10.3390/insects12090820/s1. Figure S1: Phylogenetic relationship of ML and Bl trees based on
13PCGs_AA dataset. Figure S2: Phylogenetic relationship of PhyloBayes tree based on 13PCGs_AA
dataset. Table S1: Summary of mitogenome data used in this study. Table S2: Comparison of
annotated mitogenomes of Lamiinae. Table S3: Intergenic region statistics of mitogenomes of
Lamiinae. Table S4: Nucleotide composition of mitogenomes of Agapanthia amurensis and Moechotypa
diphysis. Table S5: Relative synonymous codon usage (RSCU) of PCGs in mitogenomes. Table S6:
Saturation substitution tests for 13PCGs dataset. Table S7: Best model of phylogenetic relationships.
Author Contributions:
Conceptualization, M.L., W.W., Y.R., and P.A.; methodology, Y.R., H.L., M.L.,
L.C. and S.S.; validation, M.L., W.W. and P.A.; formal analysis, H.L. and S.S.; investigation, Y.R., H.L.,
and M.L.; resources, C.W., G.X. and P.W.; data curation, C.W., G.X. and P.W.; writing—original draft
preparation, Y.R., H.L., L.C. and C.W.; writing—review and editing, M.L., W.W., S.S. and P.A.; funding
acquisition, W.W. All authors have read and agreed to the published version of the manuscript.
Funding:
This research was funded by the National Natural Science Foundation of China (31672327).
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement:
All mitogenomic sequences in this study are available in the Gen-
Bank database (https://www.ncbi.nlm.nih.gov/nuccore, accessed on 01 July 2021), and the newly
generated sequences are deposited in GenBank with accession numbers MW617354 and MW617356.
Acknowledgments:
The authors greatly appreciate the valuable comments on an earlier version of
this manuscript received from Professor Alessandro Bruno Biscaccianti (Laboratorio di Entomologia
ed Ecologia Applicata, UniversitàMediterranea di Reggio Calabria, Dipartimento PAU, Repubblica
Italiana), and other anonymous reviewers.
Conflicts of Interest: All authors declare no conflict of interest.
References
1.
Tavakilian, G.L.; Chevillotte, H. Titan: Base de données internationales sur les Cerambycidae ou Longicornes. Version 3.0.
Available online: http://titan.gbif.fr/index.html (accessed on 1 May 2021).
2.
Ikeda, T.; Oda, K. The occurrence of attractiveness for Monochamus alternatus Hope (Coleoptera: Cerambycidae) in nematode-
infected pine trees. J. Jpn. For. Soc. 1980,62, 432–434. [CrossRef]
3.
Evans, H.F.; McNamara, D.G.; Braasch, H.; Chadoeuf, J.; Magnusson, C. Pest risk analysis (PRA) for the territories of the European
Union (as PRA area) on Bursaphelenchus xylophilus and its vectors in the genus Monochamus.EPPO Bull.
1996
,26, 199–249.
[CrossRef]
4.
Jiang, S.N. Coleoptera, Fauna Sinica Insecta.Vol 21 Coleoptera Cerambycidae Lepturinae; Science Press: Beijing, China, 2001; pp. 1–299.
ISBN 7-03-008163-3/Q934.
5.
Allison, J.D.; Borden, J.H.; Seybold, S.J. A review of the chemical ecology of the Cerambycidae (Coleoptera). Chemoecology
2004
,
14, 123–150. [CrossRef]
6.
Raje, K.R.; Abdel-Moniem, H.E.M.; Farlee, L.; Ferris, V.B.; Holland, J.D. Abundance of pest and benign Cerambycidae both
increase with decreasing forest productivity. Agric. For. Entomol. 2012,14, 165–169. [CrossRef]
7.
Karpi´nski, L.; Maák, I.; Wegierek, P. The role of nature reserves in preserving saproxylic biodiversity: Using longhorn beetles
(Coleoptera: Cerambycidae) as bioindicators. Eur. Zool. J. 2021,88, 487–504. [CrossRef]
Insects 2021,12, 820 11 of 13
8.
Napp, D.S. Phylogenetic relationships among the subfamilies of Cerambycidae (Coleoptera, Chrysomeloidea). Rev. Bras. Entomol.
1994,38, 265–419.
9.
Marvaldi, A.E.; Duckett, C.N.; Kjer, K.M.; Gillespie, J.J. Structural alignment of 18S and 28S rDNA sequences provides insights
into the phylogeny of Phytophaga (Coleoptera: Curculionoidea and Chrysomeloidea). Zool. Scripta 2008,38, 63–77. [CrossRef]
10.
Wang, B.; Ma, J.Y.; McKenna, D.D.; Yan, E.V.; Zhang, H.C.; Jarzembowski, E.A. The earliest known longhorn beetle (Cerambycidae:
Prioninae) and implications for the early evolution of Chrysomeloidea. J. Syst. Palaeontol. 2013,12, 565–574. [CrossRef]
11.
Haddad, S.; Shin, S.; Lemmon, A.R.; Lemmon, E.M.; Svacha, P.; Farrell, B.; ´
Slipi´nski, A.; Windsor, D.; Mckenna, D.D. Anchord
hybrid enrichment provides new insights into the phylogeny and evolution of longhorned beetles (Cerambycidae). Syst. Entomol.
2017,43, 68–89. [CrossRef]
12.
Nie, R.E.; Vogler, A.P.; Yang, X.K.; Lin, M.Y. Higher-level phylogeny of longhorn beetles (Coleoptera: Chrysomeloidea) inferred
from mitochondrial genomes. Syst. Entomol. 2020,46, 56–70. [CrossRef]
13.
Souza, D.D.S.; Marinoni, L.; Monné, M.L.; Gómez-Zurita, J. Molecular phylogenetic assessment of the tribal classification of
Lamiinae (Coleoptera: Cerambycidae). Mol. Phylogenetics Evol. 2020,145, 106736. [CrossRef]
14.
Latreille, P.A. Familles Naturelles du Règne Animal, Exposées Succinctement et Dans un Ordre Analytique, Avec l’Indication de Leurs
Genres; Imprimerie DE J.: Paris, France, 1825; pp. 1–570.
15.
Blanchard, C.É.Histoire dês Insectes, Traitant de Leurs Moeurs et Leurs Métamorphoses em General et Comprenant une Nouvelle
Classification Fondée sur Leurs Rapports Naturels; Hymenoptères et Coléoptères; Librairie de Firmin Didot Frères: Paris, France,
1845; pp. 154–161.
16.
Thomson, J. Essai d’Une Classification de la Famille des Cérambycides et Matériaux pour Servir a Une Monographie de Cette Famille;
Bouchard-Huzard: Paris, France, 1860; pp. 1–128.
17.
Thomson, J. Systema Cerambycidarum ou Exposéde Tous les Genres Compris Dans La famille dês Cérambycides et Familles Limitrophes;
Sociétéroyale des sciences de Liège: Paris, France, 1864; pp. 1–540.
18. Aurivillius, C. Coleopterorum Catalogus, Pars 73, Cerambycidae: Lamiinae, I; W, Junk: Berlin, Germany, 1922; pp. 1–322.
19. Aurivillius, C. Coleopterorum Catalogus, Pars 74, Cerambycidae: Lamiinae II; W, Junk: Berlin, Germany, 1923; pp. 322–704.
20.
Breuning, S. Catalogue des Lamiaires du Monde (Col. Céramb.). 1; Verlag des Museum, G. Frey: Tutzing bei Munich, Germany, 1958;
pp. 1–48.
21.
Breuning, S. Catalogue des Lamiaires du Monde (Col. Céramb.). 2.; Verlag des Museum, G. Frey: Tutzing bei Munich, Germany, 1959;
pp. 109–182.
22.
Breuning, S. Catalogue des Lamiaires du Monde (Col. Céramb.). 3.; Verlag des Museum, G. Frey: Tutzing bei Munich, Germany, 1960;
pp. 183–284.
23. Breuning, S. Révision des Pteropliini (Col. Cerambycidae). Pesquisas Zoologia 1961,9, 5–60.
24.
Breuning, S. Catalogue des Lamiaires du Monde (Col. Céramb.). 4; Verlag des Museum, G. Frey: Tutzing bei Munich, Germany, 1961;
pp. 287–382.
25.
Breuning, S. Catalogue des Lamiaires du Monde (Col. Céramb.). 5; Verlag des Museum, G. Frey: Tutzing bei Munich, Germany, 1961;
pp. 387–459.
26.
Breuning, S. Catalogue des Lamiaires du Monde (Col. Céramb.). 6; Verlag des Museum, G. Frey: Tutzing bei Munich, Germany, 1962;
pp. 463–555.
27.
Breuning, S. Catalogue des Lamiaires du Monde (Col. Céramb.). 7; Verlag des Museum, G. Frey: Tutzing bei Munich, Germany, 1963;
pp. 49–107.
28.
Breuning, S. Catalogue des Lamiaires du Monde (Col. Céramb.). 8; Verlag des Museum, G. Frey: Tutzing bei Munich, Germany, 1965;
pp. 559–655.
29.
Breuning, S. Catalogue des Lamiaires du Monde (Col. Céramb.). 9; Verlag des Museum, G. Frey: Tutzing bei Munich, Germany, 1966;
pp. 659–765.
30.
Breuning, S. Catalogue des Lamiaires du Monde (Col. Céramb.). 10; Verlag des Museum, G. Frey: Tutzing bei Munich, Germany, 1967;
pp. 771–864.
31.
Breuning, S. Catalogue des Lamiaires du Monde (Col. Céramb.). 11; Verlag des Museum, G. Frey: Tutzing bei Munich, Germany, 1969;
pp. 865–1069.
32.
Bouchard, P.; Bousquet, Y.; Davies, A.; Alonso-Zarazaga, M.A.; Lawrence, J.F.; Lyal, C.H.C.; Newton, A.F.; Reid, C.A.M.; Schmitt,
M.; ´
Slipi´nski, S.A.; et al. Family-group names in Coleoptera (Insecta). ZooKeys 2011,88, 1–972. [CrossRef] [PubMed]
33.
Saccone, C.; Giorgi, C.D.; Gissi, C.; Pesole, G.; Reyes, A. Evolutionary genomics in Metazoa: The mitochondrial DNA as a model
system. Gene 1999,238, 195–209. [CrossRef]
34.
Abascal, F.; Posada, D.; Knight, R.D.; Zardoya, R. Parallel evolution of the genetic code in arthropod mitochondrial genomes.
PLoS Biol. 2006,4, e127. [CrossRef] [PubMed]
35.
Simon, C.; Buckley, T.R.; Frati, F.; Stewart, J.B.; Beckenbach, A.T. Incorporating molecular evolution into phylogenetic analysis,
and a new compilation of conserved polymerase chain reaction primers for animal mitochondrial DNA. Annu. Rev. Ecol. Evol.
Syst. 2006,37, 545–579. [CrossRef]
36.
Tan, M.H.; Gan, H.M.; Lee, Y.P.; Poore, G.C.; Austin, C.M. Digging deeper: New gene order rearrangements and distinct patterns
of codons usage in mitochondrial genomes among shrimps from the Axiidea, Gebiidea and Caridea (Crustacea: Decapoda). PeerJ
2017,5, e2982. [CrossRef]
Insects 2021,12, 820 12 of 13
37.
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]
38.
Li, R.M.; Song, X.; Du, Y.M. Mitochondrial genome of Aromia bungii (Coleoptera: Chrysomeloidea: Cerambycidae) and phyloge-
netic analysis. Mitochondrial DNA Part B 2021,6, 71–72. [CrossRef] [PubMed]
39.
Bernt, M.; Donath, A.; Jühling, F.; Externbrink, F.; Florentz, C.; Fritzsch, G.; Pütz, J.; Middendorf, M.; Stadler, P.F. MITOS:
Improved de novo metazoan mitochondrial genome annotation. Mol. Phylogenet. Evol. 2012,69, 313–319. [CrossRef]
40.
Grant, J.R.; Stothard, P. The CGView Server: A comparative genomics tool for circular genomes. Nucleic Acids Res.
2008
,36,
181–184. [CrossRef]
41.
Zhang, D.; Gao, F.L.; Jakovli´c, I.; Zou, H.; Zhang, J.; Li, W.X.; Wang, G.T. PhyloSuite: An integrated and scalable desktop platform
for streamlined molecular sequence data management and evolutionary phylogenetics studies. Mol. Ecol. Resour.
2019
,20,
348–355. [CrossRef] [PubMed]
42.
Benson, G. Tandem repeats finder: A program to analyze DNA sequences. Nucleic Acids Res.
1999
,27, 573–580. [CrossRef]
[PubMed]
43.
Katoh, K.; Standley, D.M. MAFFT multiple sequence alignment software version 7: Improvements in performance and usability.
Mol. Biol. Evol. 2013,30, 772–780. [CrossRef]
44.
Xia, X. DAMBE7: New and improved tools for data analysis in molecular biology and evolution. Mol. Biol. Evol.
2018
,35,
1550–1552. [CrossRef]
45.
Lanfear, R.; Frandsen, P.B.; Wright, A.M.; Senfeld, T.; Calcott, B. Partition finder 2: New Methods for selecting partitioned models
of evolution for molecular and morphological phylogenetic analyses. Mol. Biol. Evol. 2016,34, 772–773. [CrossRef]
46.
Nguyen, L.T.; Schmidt, H.A.; Von Haeseler, A.; Minh, B.Q. IQ-TREE: A fast and effective stochastic algorithm for estimating
maximum-likelihood phylogenies. Mol. Biol. Evol. 2015,32, 268–274. [CrossRef]
47.
Ronquist, F.; Teslenko, M.; Van Der Mark, P.; Ayres, D.L.; Darling, A.; Höhna, S.; Larget, B.; Liu, L.; Suchard, M.A.; Huelsenbeck,
J.P. MrBayes 3.2: Efficient Bayesian phylogenetic inference and model choice across a large model space. Syst. Biol.
2012
,61,
539–542. [CrossRef]
48.
Timmermans, M.J.; Barton, C.; Haran, J.; Ahrens, D.; Culverwell, C.L.; Ollikainen, A.; Vogler, A.P. Family-level sampling of
mitochondrial genomes in Coleoptera: Compositional heterogeneity and phylogenetics. Genome Biol. Evol.
2016
,8, 161–175.
[CrossRef]
49.
Letunic, I.; Bork, P. Interactive tree of life (iTOL) v3: An online tool for the display and annotation of phylogenetic and other trees.
Nucleic Acids Res. 2016,44, 242–245. [CrossRef]
50.
Liu, Y.Q.; Chen, D.B.; Liu, H.H.; Hu, H.L.; Bian, H.X.; Zhang, R.S.; Yang, R.S.; Jiang, X.F.; Shi, S.L. The complete mitochondrial
genome of the longhorn Beetle Dorysthenes paradoxus (Coleoptera: Cerambycidae: Prionini) and the implication for the
phylogenetic relationships of the Cerambycidae species. J. Insect Sci. 2018,18, 1–8. [CrossRef] [PubMed]
51.
Wang, J.; Dai, X.Y.; Xu, X.D.; Zhang, Z.Y.; Zhang, J.Y. The complete mitochondrial genomes of five longicorn beetles (coleoptera:
Cerambycidae) and phylogenetic relationships within cerambycidae. PeerJ 2019,7, e7633. [CrossRef] [PubMed]
52.
Zhang, D.X.; Hewitt, G.M. Insect mitochondrial control region: A review of its structure, evolution and usefulness in evolutionary
studies. Biochem. Syst. Ecol. 1997,25, 99–120. [CrossRef]
53. Boore, J.L. Animal mitochondrial genomes. Nucleic Acids Res. 1999,27, 1767–1780. [CrossRef]
54.
Cameron, S.L. Insect mitochondrial genomics: Implications for evolution and phylogeny. Annu. Rev. Èntomol.
2014
,59, 95–117.
[CrossRef]
55. Rand, D.M. Endotherms, ectotherms, and mitochondrial genome size variation. J. Mol. Evol. 1993,37, 281–295. [CrossRef]
56. Rydin, C.; Källersjö, M. Taxon sampling and seed plant phylogeny. Cladistics 2002,18, 485–513. [CrossRef]
57. Sanderson, M.J.; Donoghue, M.J. Patterns of variation in levels of homoplasy. Evolution 1989,43, 1781–1795. [CrossRef]
58.
Cai, C.; Tihelka, E.; Pisani, D.; Donoghue, P.C. Data curation and modeling of compositional heterogeneity in insect phylogenomics:
A case study of the phylogeny of Dytiscoidea (Coleoptera: Adephaga). Mol. Phylogenetics Evol. 2020,147, 106782. [CrossRef]
59.
Ai, D.; Peng, L.; Qin, D.; Zhang, Y. Characterization of three complete mitogenomes of Flatidae (Hemiptera: Fulgoroidea)
and compositional heterogeneity analysis in the Planthoppers’ mitochondrial phylogenomics. Int. J. Mol. Sci.
2021
,22, 5586.
[CrossRef]
60.
Nie, R.E.; Andújar, C.; Gómez-Rodríguez, C.; Bai, M.; Xue, H.J.; Tang, M.; Vogler, A.P. The phylogeny of leaf beetles (Chrysomeli-
dae) inferred from mitochondrial genomes. Syst. Entomol. 2020,45, 188–204. [CrossRef]
61.
Zhang, Z.Y.; Guan, J.Y.; Cao, Y.R.; Dai, X.Y.; Storey, K.B.; Yu, D.N.; Zhang, J.Y. Mitogenome analysis of four lamiinae species
(Coleoptera: Cerambycidae) and gene expression responses by monochamus alternatus when infected with the parasitic
nematode, bursaphelenchus mucronatus. Insects 2021,12, 453. [CrossRef] [PubMed]
62.
Cherepamov, A.I.; Zolotarenko, G.S.; Kothekar, V.S. Cerambycidae of Northern Asia. Vollume 3, Lamiinae, Part 1 (English Edition);
Amerind Publishing: New Delhi, India, 1990; pp. 1–300.
63.
Löbl, I.; Smetana, A. Catalogue of Palaearctic Coleoptera, Vol. 6: Chrysomeloidae; Apollo Books: Stenstrup, Denmark, 2010; pp. 1–924.
64.
Gressitt, J.L. Longicorn Beetles of China. Longicornia, Études et Notes sur les Longicornes; Paul Lechevalier: Paris, Germany, 1951;
pp. 1–667. Volume 2.
Insects 2021,12, 820 13 of 13
65. Ohbayashi, N.; Niisato, T. Longicorn Beetles of Japan; Tokai University Press: Kanagawa, Japan, 2007; pp. 1–818.
66.
Toki, W.; Kubota, K. Molecular phylogeny based on mitochondrial genes and evolution of host plant use in the long-horned
beetle tribe Lamiini (Coleoptera: Cerambycidae) in Japan. Environ. Entomol. 2010,4, 1336–1343. [CrossRef] [PubMed]
... However, they retained Batocerini as a separate tribe, considering it the sister clade to Lamiini. The placement of Batocerini has since become a topic of considerable debate, with various studies either including it within Lamiini s. l. (Gorring, 2019;Nie et al., 2020;Bai et al., 2022;Pu et al., 2022;Soydabaş-Ayoub & Uçkan, 2023) or treating it as a sister group (Souza et al., 2020;Ashman et al., 2021;Ren et al., 2021;Zhang et al., 2021;Li et al., 2023). ...
... This comprehensive dataset has, on one hand, significantly clarified the internal relationships among different clades, but on the other hand, it has highlighted problematic issues in classification and intergeneric relationships that remain unresolved. A key finding of the study is the confirmation of the monophyly of Lamiini s. l., as suggested by several studies with much smaller taxa samples and using different genetic markers and evolutionary models (Nie et al., 2020;Souza et al., 2020;Ren et al., 2021;Pu et al., 2022). My research demonstrated that Lamiini s. l. consists of at least eight distinct evolutionary clades: TAENI, DORCA, ACALO, BATOC, PSACO, ANOPL, MONOC, and LAMIA, which are successively nested on the phylogenetic tree. ...
... The most variable results are observed with the basal clades, as well as the position of Batocerini. In many studies, Batocerini is defined as a sister group to Lamiini (Souza et al., 2020;Ashman et al., 2021;Ren et al., 2021;Zhang et al., 2021;Li et al., 2023), while several other studies place Batocerini within Lamiini (Gorring, Gorring (2019) in his dissertation, proposing that this group should be considered a part of the Lamiini tribe. The results of the current study align with Gorring's conclusions (2019), so I emphasize that Batocerini is indeed part of Lamiini. ...
Article
Full-text available
In this study I conduct the phylogenetic and biogeographical analysis of Lamiini sensu lato (Coleoptera, Cerambycidae, Lamiinae), confirming it as a monophyletic group that is broader than previously assumed and supporting the redefinition of it as a unified tribe-Lamiini sensu novo. The group consist of eight major clades (TAENI, DORCA, ACALO, BATOC, PSACO, ANOLPL, MONOC, LAMIA) and integrates together several traditionally recognized tribes (Dorcaschematini, Batocerini, Petrognathini, Rhodopini, Monochamini, Dorcadionini, Phrissomini). Most of them, including Monochamini, Petrognathini, Dorcadionini, and Phrissomini are polyphyletic, representing multiple lineages with intricate evolutionary history. Phylogeographic analysis suggests a South Gondwanan origin of Lamiini s. l., with its ancestral lineage (LaCA) emerging in the Antarctica-South America suture zone before global expansion through three key migration routes: northeast Pantalassic, Trans-Tethyan, and southwest Pantalassic. Continental drift and climate oscillations influenced this process. Initial diversification (~70-60 Ma) led to vicariant taxa due to continental isolation, with basal clades (TAENI, DORCA, ACALO) distributed across South America, North America, and Australia. Extinction events, including the K-T boundary (~66 Ma) and Cenozoic glaciations, further shaped diversification. The rise of the BATOC and PSACO clades marked the early diversification of crown-group Lamiini s. l. within the island continent of Greater India and the multiple archipelagos of the Trans-Tethyan Arc (~60-50 Ma). The collision of Greater India with Asia (~50-45 Ma) facilitated faunal exchanges, aiding dispersal into Africa and Southeast Asia and driving ANOPL diversification (45-35 Ma). The ACALO clade underwent secondary diversification during the initial collision of Australia and Eurasia (~25-20 Ma), coinciding with the uplift of New Guinea in the Miocene. The LAMIA clade likely originated in East Asia (~25-20 Ma). During the Miocene Climatic Optimum, it expanded into Africa via the Gomphotherium Land Bridge (~16 Ma) and colonized Europe. Later cooling and aridification (~5-11 Ma) drove further diversification, particularly in Eurasian steppes. The MONOC clade also originated in East Asia, adapting to montane coniferous forests during the Miocene cooling. It spread across Eurasia during the Miocene glaciations (~20 Ma, ~14 Ma) and entered North America via Beringia. Expansion into Mesoamerica, northern South America, and the Caribbean likely occurred during the Pleistocene glaciations (~0.01-2 Ma) when climate and fluctuating sea levels enabled dispersal. In summary, the findings refine Lamiini s. l. phylogeny and highlight the impact of geological and climatic events on its evolution. This study clarifies taxonomic ambiguities and provides a framework for future research on diversification and biogeographic patterns.
... de Santana Souza et al. (2020) rendered ten out of 46 tribes monophyletic, including Batocerini, Dorcadionini, Lamiini, and Mesosini and 15 tribes polyphyletic, including some widely distributed or diverse tribes, such as Acanthocinini, Acanthoderini, Agapanthiini, Monochamini, Phytoeciini, Pogonocherini, and Saperdini, based on mitochondrial and nuclear marker DNA sequences. Ren et al. (2021) presented the sisterships of Saperdini and Phytoeciini and Batocerini and Lamiini based on the analyzed 13 protein-coding genes of mitogenomes of 12 tribes' representatives. Ashman et al. (2022) also reported polyphyly of Acanthocinini and Lamiini, relying on the phylogenetic analysis of hundreds of nuclear gene regions, including representatives of 14 tribes, primarily Australian. ...
... Our results were conformable with the results of Ashman et al. (2022) and de Santana Souza et al. (2020) on the position of Monochamus and Batocera. Ren et al. (2021) also supported the sisterships of Batocerini and Lamiini. Contemplating the results of the present study and the previous studies mentioned above, the expectations of Lacordaire (1869) and Pascoe (1866) (as per de Santana Souza et al. 2020) regarding the close relationship between Batocerini and Monochamini seem to be plausible, and Dorcadionini, Gnomini, and Monochamini (at least in terms of Monochamus) should be revised. ...
Article
Lamiinae (Cerambycidae, Coleoptera) is a striking subfamily due to its members' economic importance and role in the forest ecosystem. Morphological diversity, worldwide distribution and species richness complicate its already intricate phylogenetic relationships. We implemented Maximum Likelihood (ML) and time-scaled Bayesian Inference (BI) analyses to the species from East of Marmara Basin, Türkiye, from the tribes Acanthocinini, Acanthoderini, Agapanthiini, Batocerini, Dorcadionini, Lamiini, Mesosini, Monochamini, Phytoeciini, Phrynetini, Pogonocherini (including Exocentrini) and Saperdini using partial mitochondrial cytochrome c oxidase-I (COI) and 16S rRNA and nuclear 28S rRNA gene regions (2257 base pair alignment length) and Neighbor-Joining (NJ) and ML analysis to the global COI gene region dataset (658 bp). The most recent common ancestor (MRCA) of Lamiinae members included in the analyses was dated ~ 127 million years ago (Mya) in the Cretaceous. The MRCA of Dorcadionini, Lamiini and Monochamini was younger than the common ancestors of the other close tribes. There was a concurrence between resolutions of ML and BI on the affiliations of Dorcadionini and Monochamini to Lamiini and the proximity of Batocerini to Lamiini, Acanthocinini to Acanthoderini, Phrynetini to Pogonocherini, and Phytoeciini to Saperdini. The COI-based NJ and ML gene trees suggest that the closest relatives of most of the sampled Lamiinae species from the East of Marmara Basin were the European conspecifics or congeners. Our results support Dorcadionini and Monochamini as synonyms of Lamiini; and Phytoeciini of Saperdini. Also, they suggest that the emergence of the living tribes included in this study was during the Paleogene, and their intrageneric diversifications occurred during the Cenozoic, mostly the Neogene.
... However, the high degree of subjectivity in the classification of tribes within Lamiinae indicates that the tribal relationships within this group require further investigation [8,24]. Furthermore, the monophyly of proposed tribes has not been tested using molecular data, with the exception of a few recent studies [8,25,26]. The subfamily Lamiinae contains various forest pests, such as Anoplophora glabripennis, which is a major international quarantine pest, and species in the genus Monochamus Dejean, which are the primary vectors of Bursaphelenchus xylophilus (Steiner and Buhrer) (Nematoda: Aphelenchoididae) [27]. ...
... Mitochondrial genomes have been widely used for species identification, as well as population genetic, phylogenetic, and evolutionary studies because their structure is evolutionarily conserved, they are maternally inherited, they accumulate mutations at a high rate, they lack introns, and they have a negligible rate of recombination [28,29]. Mitochondrial genomes have been used to resolve the phylogenetic relationships among various taxa within Cerambycidae [4,26,30,31]. However, relationships among tribe-level taxa in Lamiinae have still been controversial. ...
Article
The genus Monochamus within the subfamily Lamiinae is the main vector of Bursaphelenchus xylophilus, which causes pine wilt disease and induces substantial economic and ecological losses. Only three complete mitochondrial genomes of the genus Monochamus have been sequenced to date, and no comparative mitochondrial genomic studies of Lamiinae have been conducted. Here, the mitochondrial genomes of two Monochamus species, M. saltuarius and M. urussovi, were newly sequenced and annotated. The composition and order of genes in the mitochondrial genomes of Monochamus species are conserved. All transfer RNAs exhibit the typical clover-leaf secondary structure, with the exception of trnS1. Similar to other longhorn beetles, Lamiinae mitochondrial genomes have an A + T bias. All 13 protein-coding genes have experienced purifying selection, and tandem repeat sequences are abundant in the A + T-rich region. Phylogenetic analyses revealed congruent topologies among trees inferred from the five datasets, with the monophyly of Acanthocinini, Agapanthiini, Batocerini, Dorcaschematini, Pteropliini, and Saperdini receiving high support. The findings of this study enhance our understanding of mitochondrial genome evolution and will provide a basis for future studies of population genetics and phylogenetic investigations in this group.
... The use of mitogenomic data is a common approach to exploring phylogenetic relationships among different insect groups [68,69]. The sister relationship between Formica and Polyergus ants confirmed in the present work was largely congruent with the results of previous studies [6]. ...
Article
Full-text available
Formica is a large genus in the family Formicidae with high diversity in its distribution, morphology, and physiology. To better understand evolutionary characteristics of Formica, the complete mitochondrial genomes (mitogenomes) of two Formica species were determined and a comparative mitogenomic analysis for this genus was performed. The two newly sequenced Formica mitogenomes each included 37 typical mitochondrial genes and a large non-coding region (putative control region), as observed in other Formica mitogenomes. Base composition, gene order, codon usage, and tRNA secondary structure were well conserved among Formica species, whereas diversity in sequence size and structural characteristics was observed in control regions. We also observed several conserved motifs in the intergenic spacer regions. These conserved genomic features may be related to mitochondrial function and their highly conserved physiological constraints, while the diversity of the control regions may be associated with adaptive evolution among heterogenous habitats. A negative AT-skew value on the majority chain was presented in each of Formica mitogenomes, indicating a reversal of strand asymmetry in base composition. Strong codon usage bias was observed in Formica mitogenomes, which was predominantly determined by nucleotide composition. All 13 mitochondrial protein-coding genes of Formica species exhibited molecular signatures of purifying selection, as indicated by the ratio of non-synonymous substitutions to synonymous substitutions being less than 1 for each protein-coding gene. Phylogenetic analyses based on mitogenomic data obtained fairly consistent phylogenetic relationships, except for two Formica species that had unstable phylogenetic positions, indicating mitogenomic data are useful for constructing phylogenies of ants. Beyond characterizing two additional Formica mitogenomes, this study also provided some key evolutionary insights into Formica.
... The tribe Lamiini Latreille, 1925 includes more than 220 genera and 1400 species widespread in all bioregions excluding poles (Zicha, 2021). Beside the genera traditionally included (Breuning, 1943), recent genetic analyses imply that also other ones, once classified in some closely related tribes, i.e., Gnomini Thomson, 1864and Rhodopinini Gressitt, 1951, belong to Lamiini (Souza et al., 2020Ren et al., 2021). Despite this great number of taxa, new ones are continuously discovered, especially in Tropical regions. ...
Article
Full-text available
Pseudobrachyhammus sulawensis n. gen. n. sp. is described from Sulawesi. Differential characters with other genera of Lamiini, especially those present in Indonesia, are yielded.
Article
Cerambycidae (Coleoptera: Chrysomeloidea) are a widely distributed group of insects with significant economic importance. Despite substantial efforts and some advancements in inferring the phylogeny of Cerambycidae, high-level phylogenetic relationships within the family, including subfamily and tribe-level classification, remain contentious. In this study, we performed whole-genome sequencing on a total of 65 species. We integrated these comprehensive genomic data with existing whole-genome and transcriptome data to conduct a phylogenetic analysis of Cerambycidae s.s. The results supported the monophyly of Lamiinae, Cerambycinae, and Spondylidinae, whereas Lepturinae and Prioninae were found to be non-monophyletic. Lamiinae were recovered as the sister group to all other Cerambycidae s.s. Spondylidinae were found to be sister to the clade comprising Lepturinae and Necydalinae, and Cerambycinae were identified as the sister group to the clade consisting of Prioninae and Parandrinae. Within Lamiinae, the tribes Mesosini, Saperdini, and Dorcaschematini were found to be monophyletic; however, Acanthocinini, Agapanthiini, Pogonocherini, Pteropliini, Lamiini, and Monochamini were non-monophyletic. This study supported the classification integrating the current Monochamini into Lamiini. Molecular dating analysis suggests that diversification within Cerambycidae s.s. began at the boundary between the Jurassic and Cretaceous periods, ~145 Mya.
Article
Full-text available
Simple Summary Lamiinae is the largest subfamily among the Cerambycidae (longhorn beetles), and its members are distributed worldwide. The monophyly of Lamiinae is generally recognized, but there are still diverse ideas as to whether the tribes belonging to Lamiinae are monophylic. Ambiguous classification boundaries and the existence of synonyms are major issues leading to controversies over Lamiinae classification. It is not enough to conduct research solely on the morphological characteristics and simple molecular loci of longhorn beetles. Mitochondrial genomes have proven to be reliable markers and can shed more light on phylogenetic relationships among Lamiinae. The present study resolved infra-subfamilial relationships among Lamiinae and provides more mitochondrial data for further phylogenetic research on longhorn beetles. Abstract Lamiinae is the largest subfamily of the Cerambycidae (longhorn beetles), with approximately 21,863 described species. Previous phylogenetic studies of Lamiinae showed that this subfamily was monophyletic, but the relationship between the tribes of Lamiinae is still controversial. Partial molecular data and species morphological characteristics are not sufficient to resolve species phylogenetic studies perfectly. At the same time, the full mitochondrial genome contains more comprehensive genetic data. Benefiting from the development of next-generation sequencing (NGS), mitochondrial genomes can be easily acquired and used as reliable molecular markers to investigate phylogenetic relationships within Cerambycidae. Using NGS technology, we obtained 11 mitochondrial genome sequences of Lamiinae species. Based on this newly generated mitochondrial genome dataset matrix, we reconstructed the phylogeny of Lamiinae. The Bayesian Inference and Maximum Likelihood analyses strongly support the monophyly of four tribes (Lamiini, Batocerini, Mesosini, and Saperdini), whereas the tribe Acanthocinini was identified as paraphyletic. Other mitochondrial structural features were also observed: the start codon in the nad1 gene of all 11 mitochondrial genomes is TTG; 17–22 bp intergenic spacers (IGS) with a ‘TACTA’ motif were found between trnS2 and nad1. Moreover, two long IGS were found in Mesosa myops and Batocera sp. Tandem repeats were found in the IGS of Batocera sp.
Preprint
Full-text available
The subfamily Lamiinae (Cerambycidae, Coleoptera) is striking due to its morphological diversity and species richness with intricate phylogenetic relationships. We inferred the phylogeny and evolutionary history of extant species of East of Marmara Basin, Türkiye, from the tribes Acanthocinini, Acanthoderini, Agapanthiini, Batocerini, Dorcadionini, Lamiini, Mesosini, Monochamini, Phytoeciini, Phrynetini, Pogonocherini (including Exocentrini) and Saperdini using partial mitochondrial cytochrome c oxidase-I ( COI ) and 16S rRNA and nuclear 28S rRNA gene regions (2257 base pair alignment length). The most recent common ancestor (MRCA) of Lamiinae members included in the analyses was dated ~ 127 million years ago (Mya) in the Cretaceous. The MRCA of Dorcadionini, Lamiini and Monochamini was younger than the common ancestors of the other close tribes. There was a concurrence between resolutions of Maximum Likelihood (ML) and Bayesian analyses on the affiliations of Dorcadionini and Monochamini to Lamiini and the proximity of Batocerini to Lamiini, Acanthocinini to Acanthoderini, Phrynetini to Pogonocherini, and Phytoeciini to Saperdini. The COI -based Neighbor-Joining and ML gene trees suggest that the closest relatives of the extant Lamiinae species of East of Marmara Basin were the European conspecifics or congeners. Moreover, Paraleprodera and Lamia (Lamiini) were sisters to Imantocera (Gnomini), Oberea (Obereini) to Phytoecia Phytoeciini), and Hippopsis (Agapanthiini) to Omosarotes singularis Pascoe,1860 (Acanthomerosternoplini). Our results support Dorcadionini, Gnomini and Monochamini as synonyms of Lamiini; and Obereini and Phytoeciini of Saperdini and suggest that the emergence of the living tribes included in this study was during Paleogene, and their intrageneric diversifications occurred during Cenozoic, mostly Neogene.
Preprint
Lamiinae is a tremendous subfamily of Cerambycidae, with around 20,000 members dispersed across continents. The knowledge of the evolutionary history of the subfamily is scarcely, and there are growing doubts about the phylogenetic relationships due to the recognised illusion caused by the convergence of the morphological characters. The present study contributes to the evolutionary history and phylogenetic relationships of the tribes Acanthocinini, Acanthoderini, Agapanthiini, Batocerini, Dorcadionini, Lamiini, Mesosini, Monochamini, Phytoeciini, Phrynetini, Pogonocherini (including Exocentrini) and Saperdini with Neighbor-Joining (NJ), Maximum Likelihood (ML) and time-scaled Bayesian analyses based on partial mitochondrial COI and 16S rRNA and nuclear 28S rRNA gene regions (2257 base pair alignment length). The most recent common ancestor (MRCA) of the taxa included in the analyses appeared during the Middle or Late Cretaceous, and the MRCAs of the closely related tribes emerged in Paleogene. The MRCA of Dorcadionini, Lamiini and Monochamini was younger than the common ancestors of the other close tribes. The hypothetical ML phylogram was consistent with the Bayesian chronogram in the proximity of Batocerini to Lamiini, Acanthocinini to Acanthoderini, Phyretrini to Pogonocherini, and Phytoeciini to Saperdini, in addition to the affiliation of Lamiini, Dorcadionini and Monochamini. At the COI-based NJ and ML gene trees, Paraleprodera and Lamia (Lamiini) were sisters to Imantocera (Gnomini), Oberea (Obereini) to Phytoecia (Phytoeciini), and Hippopsis (Agapanthiini) to Omosarotes (Acanthomerosternoplini). The present results support Dorcadionini Gnomini and Monochamini as synonyms of Lamiini; and Obereini and Phytoeciini of Saperdini. We suggest that the emergence of the living tribes included in this study was during Paleogene, and the intrageneric diversifications occurred in Cenozoic, mostly during Neogene.
Chapter
The longhorn beetles are cosmopolitan in nature, with diverse host range and species diversity. Identification of the longhorn beetles based on the conventional taxonomy is time consuming and tough to adopt due to scarce availability of the type specimens, lingual problems, and scanty literature. There is a wide scope for alternative tool for identification of these beetles. One such technique is molecular-based DNA barcoding. This technique is rapid, accurate, and species-specific. Due to less information available related to the sequence information and improper data submission of longhorn beetles, there is a chance of misidentification. The identifications based on the COI, RAPD, RFLP, AFLP, microsatellite, and LAMP are the promising tools for longhorn beetles. Development of an independent database for store and retrieving the information for faster identification and simpler work nature is the need of the hour. The adoption of the molecular techniques alone will not give conclusive results in the identification of the longhorn beetles. Hence, integration of the morphological, ecological, physiological, and molecular barcoding information provides the more complete and comprehensive results.
Article
Full-text available
Although sequences of mitogenomes have been widely used for investigating phylogenetic relationship, population genetics, and biogeography in many members of Fulgoroidea, only one complete mitogenome of a member of Flatidae has been sequenced. Here, the complete mitogenomes of Cerynia lineola, Cromna sinensis, and Zecheuna tonkinensis are sequenced. The gene arrangements of the three new mitogenomes are consistent with ancestral insect mitogenomes. The strategy of using mitogenomes in phylogenetics remains in dispute due to the heterogeneity in base composition and the possible variation in evolutionary rates. In this study, we found compositional heterogeneity and variable evolutionary rates among planthopper mitogenomes. Phylogenetic analysis based on site-homogeneous models showed that the families (Delphacidae and Derbidae) with high values of Ka/Ks and A + T content tended to fall together at a basal position on the trees. Using a site-heterogeneous mixture CAT + GTR model implemented in PhyloBayes yielded almost the same topology. Our results recovered the monophyly of Fulgoroidea. In this study, we apply the heterogeneous mixture model to the planthoppers’ phylogenetic analysis for the first time. Our study is based on a large sample and provides a methodological reference for future phylogenetic studies of Fulgoroidea.
Article
Full-text available
Simple Summary The longicorn beetle, Monochamus alternatus, is a major vector for the transmission of pine wilt disease, which is caused by a nematode pathogen, Bursaphelenchus xylophilus (or also possibly by B. mucronatus) that is spread by the beetle as it feeds on pine trees. In this study, the mitochondrial genome sequences of four longicorn species (Coleoptera: Cerambycidae: Lamiinae) were determined to further elaborate the phylogenetic relationships of Lamiinae. RT-qPCR was also used to assess the expression of eight mitochondrial protein-coding genes in M. alternatus when carrying B. mucronatus or not, so as to explore the relationship between these two species. The results showed that expression of mitochondria-encoded genes was elevated in M. alternatus beetles that were infected with B. mucronatus, suggesting that B. mucronatus putatively activates an immune response, which significantly affects the metabolic processes of M. alternatus. These results are of significance for further understanding the phylogenetic relationships of longicorn beetles and controlling the spread of pine wilt disease. Abstract We determined the mitochondrial gene sequence of Monochamus alternatus and three other mitogenomes of Lamiinae (Insect: Coleoptera: Cerambycidae) belonging to three genera (Aulaconotus, Apriona and Paraglenea) to enrich the mitochondrial genome database of Lamiinae and further explore the phylogenetic relationships within the subfamily. Phylogenetic trees of the Lamiinae were built using the Bayesian inference (BI) and maximum likelihood (ML) methods and the monophyly of Monochamus, Anoplophora, and Batocera genera was supported. Anoplophora chinensis, An. glabripennis and Aristobia reticulator were closely related, suggesting they may also be potential vectors for the transmission of the pine wood pathogenic nematode (Bursaphelenchus xylophilus) in addition to M. alternatus, a well-known vector of pine wilt disease. There is a special symbiotic relationship between M. alternatus and Bursaphelenchus xylophilus. As the native sympatric sibling species of B. xylophilus, B. mucronatus also has a specific relationship that is often overlooked. The analysis of mitochondrial gene expression aimed to explore the effect of B. mucronatus on the energy metabolism of the respiratory chain of M. alternatus adults. Using RT-qPCR, we determined and analyzed the expression of eight mitochondrial protein-coding genes (COI, COII, COIII, ND1, ND4, ND5, ATP6, and Cty b) between M. alternatus infected by B. mucronatus and M. alternatus without the nematode. Expression of all the eight mitochondrial genes were up-regulated, particularly the ND4 and ND5 gene, which were up-regulated by 4–5-fold (p < 0.01). Since longicorn beetles have immune responses to nematodes, we believe that their relationship should not be viewed as symbiotic, but classed as parasitic.
Article
Full-text available
The potential of forest nature reserves as refuges for biodiversity seems to be overlooked probably due to their small size. These, however, may constitute important safe havens for saproxylic organisms since forest reserves are relatively numerous in Europe. Saproxylic beetles are among the key groups for the assessment of biodiversity in forest habitats and longhorn beetles may play an important role in bioindication as they are ecologically associated with various microhabitats and considered a very heterogeneous family of insects. To study the role of forest reserves as important habitats for saproxylic beetles, we compared cerambycid assemblages in corresponding pairs of sites (nature reserves and managed stands) in a forest region under high anthropogenic pressure (Upper Silesia, Poland, Central Europe). Moreover, we also intended to assess the role played by these beetles as bioindicators in the different forest types from this area. According to the obtained diversity index values, the most valuable stands are located in nature reserves, whilst sites with the lowest value included managed forests together with two homogeneous and relatively recently established nature reserves. Our analyses demonstrated a positive correlation between deadwood volume and biodiversity, for both species richness and abundance. Our results indicate that the decisive factor is the type of a given habitat, whose characteristics can be mainly influenced/determined by forest management. The potential role of longhorn beetles as bioindicators is highlighted and the effectiveness of using traps in this family, as well as general issues regarding the use of non-selective lethal traps in the study of single invertebrate groups in protected areas are discussed.
Article
Full-text available
The red-necked longhorn beetle Aromia bungii is a major pest of peach orchards. In this study, we sequenced and analyzed the complete mitochondrial genome (mitogenome) of A. bungi,i. This mitogenome was 15,760 bp long and encoded 13 protein-coding genes (PCGs), 22 transfer RNA genes (tRNAs) and two ribosomal RNA unit genes (rRNAs). Gene order was conserved and identical to most other previously sequenced Cerambycidae. Most PCGs of A. bungii have the conventional start codons ATN (six ATT, five ATG and one ATC), with the exception of nad1 (TTG). Except for three genes (cox1, cox2 and nad5) end with the incomplete stop codon T−, all other PCGs terminated with the stop codon TAA or TAG. The whole mitogenome exhibited heavy AT nucleotide bias (74.3%). Phylogenetic analysis positioned A. bungii in a well-supported clade within the subfamily Cerambycinae with Xystrocera globosa (tribe Xystrocerini). These results support the currently accepted taxonomy and provide a better understanding of the phylogenetic analysis of the Cerambycidae.
Article
Full-text available
Cerambycidae (longhorn beetles) and related families in the superfamily Chrysomeloidea are important components of forest ecosystems and play a key role in nutrient cycling and pollination. Using full mitochondrial genomes and dense taxon sampling, the phylogeny of Chrysomeloidea with a focus on Cerambycidae and allied families was explored. We used 151 mitochondrial genomes (75 newly sequenced) covering all families and 29 subfamilies of Chrysomeloidea. Our results reveal that (i) Chrysomelidae (leaf beetles) are sister to all other chrysomeloid families; (ii) Cerambycidae sensu stricto (s.s.) is polyphyletic due to the inclusion of other families that split Cerambycidae into a ‘lamiine’ clade comprising Lepturinae sensu lato (s.l.) + (Lamiinae + Spondylidinae) and a ‘cerambycine’ clade comprising Dorcasominae + (Cerambycinae + Prioninae s. l.); (iii) the subfamilies within the two clades of Cerambycidae s. s. were monophyletic, except for the placement of Necydalinae nested in Lepturinae, and the placement of Parandrinae within Prioninae (now considered as tribes Necydalini and Parandrini, respectively); (iv) smaller families were grouped into two major clades: one composed of Disteniidae+Vesperidae and the other composed of Orsodacnidae + (Megalopodidae + Oxypeltidae); (v) relationships among the four major clades were poorly supported but were resolved as ((cerambycines + (Disteniidae + Vesperidae) + Orsodacnidae + (Megalopodidae + Oxypeltidae)) + lamiines. Divergence time analyses estimated that Chrysomeloidea originated ca. 154.1 Mya during the late Jurassic, and most subfamilies of Cerambycidae originated much earlier than subfamilies of Chrysomelidae. The diversification of families within Chrysomeloidea was largely coincident with the radiation of angiosperms during the Early Cretaceous.
Article
Full-text available
Multigene and genomic data sets have become commonplace in the field of phylogenetics, but many existing tools are not designed for such data sets, which often makes the analysis time‐consuming and tedious. Here, we present phylosuite, a (cross‐platform, open‐source, stand‐alone Python graphical user interface) user‐friendly workflow desktop platform dedicated to streamlining molecular sequence data management and evolutionary phylogenetics studies. It uses a plugin‐based system that integrates several phylogenetic and bioinformatic tools, thereby streamlining the entire procedure, from data acquisition to phylogenetic tree annotation (in combination with iTOL). It has the following features: (a) point‐and‐click and drag‐and‐drop graphical user interface; (b) a workplace to manage and organize molecular sequence data and results of analyses; (c) GenBank entry extraction and comparative statistics; and (d) a phylogenetic workflow with batch processing capability, comprising sequence alignment (mafft and macse), alignment optimization (trimAl, HmmCleaner and Gblocks), data set concatenation, best partitioning scheme and best evolutionary model selection (partitionfinder and modelfinder), and phylogenetic inference (mrbayes and iq‐tree). phylosuite is designed for both beginners and experienced researchers, allowing the former to quick‐start their way into phylogenetic analysis, and the latter to conduct, store and manage their work in a streamlined way, and spend more time investigating scientific questions instead of wasting it on transferring files from one software program to another.
Article
Full-text available
Cerambycidae is one of the most diversified groups within Coleoptera and includes nearly 35,000 known species. The relationships at the subfamily level within Cerambycidae have not been convincingly demonstrated and the gene rearrangement of mitochondrial genomes in Cerambycidae remains unclear due to the low numbers of sequenced mitogenomes. In the present study, we determined five complete mitogenomes of Cerambycidae and investigated the phylogenetic relationship among the subfamilies of Cerambycidae based on mitogenomes. The mitogenomic arrangement of all five species was identical to the ancestral Cerambycidae type without gene rearrangement. Remarkably, however, two large intergenic spacers were detected in the mitogenome of Pterolophia sp. ZJY-2019. The origins of these intergenic spacers could be explained by the slipped-strand mispairing and duplication/random loss models. A conserved motif was found between trnS2 and nad1 gene, which was proposed to be a binding site of a transcription termination peptide. Also, tandem repeat units were identified in the A + T-rich region of all five mitogenomes. The monophyly of Lamiinae and Prioninae was strongly supported by both MrBayes and RAxML analyses based on nucleotide datasets, whereas the Cerambycinae and Lepturinae were recovered as non-monophyletic.
Article
Diving beetles and their allies are a virtually ubiquitous group of freshwater predators. Knowledge of the phylogeny of the adephagan superfamily Dytiscoidea has significantly improved since the advent of molecular phylogenetics. However, despite recent comprehensive phylogenomic studies, some phylogenetic relationships among the constituent families remain elusive. In particular, the position of the family Hygrobiidae remains uncertain. We address these issues by re-analyzing recently published phylogenomic datasets for Dytiscoidea, using approaches to reduce compositional heterogeneity and adopting a site-heterogeneous mixture model. We obtained a consistent, well-resolved, and strongly supported tree. Consistent with previous studies, our analyses support Aspidytidae as the monophyletic sister group of Amphizoidae, and more importantly, Hygrobiidae as the sister of the diverse Dytiscidae, in agreement with morphology-based phylogenies. Our analyses provide a backbone phylogeny of Dytiscoidea, which lays the foundation for better understanding the evolution of morphological characters, life habits, and feeding behaviors of dytiscoid beetles.
Article
Lamiinae is the most diverse subfamily of longhorned beetles, with about 20,000 described species classified into 80 tribes. Most of the tribes of Lamiinae were proposed during the 19th century and the suprageneric classification of the subfamily has never been assessed under phylogenetic criteria. In this study, we present the first tribal-level phylogeny of Lamiinae, inferred from 130 terminals (representing 46 tribes, prioritizing generic type species of the tribes) and fragments of two mitochondrial and three nuclear markers (cox1, rrnL, Wg, CPS and LSU; 5,024 aligned positions total). Analyses were performed under Maximum Likelihood and Bayesian methods based on two datasets: a dataset including all taxa available for the study, and a reduced dataset with 111 terminals where taxa only contributing with mitochondrial markers were excluded from the matrix. The monophyly of Lamiinae was corroborated in three of the four analyses and 11 of the 35 tribes with more than one species represented in the analyses were consistently recovered as monophyletic. However, 15 tribes were not retrieved as monophyletic, requiring a revision of their boundaries: Acanthocinini, Acanthoderini, Agapanthiini, Apomecynini, Desmiphorini, Dorcaschematini, Enicodini, Hemilophini, Monochamini, Onciderini, Parmenini, Phytoeciini, Pogonocherini, Pteropliini and Saperdini. Based on these results, when strong support values for paraphyly were recovered, we argue a number of tribe synonymies, including Moneilemini as synonym of Acanthocinini; Onocephalini of Onciderini; Dorcadionini, Gnomini, Monochamini and Rhodopinini of Lamiini; and Obereini and Phytoeciini of Saperdini. Other taxonomic changes proposed in this study based on the criterion of monophyly and supported by morphological characters include the transfer of Tricondyloides and Stenellipsis to Enicodini, and of Dylobolus stat. rest., which is removed as subgenus of Mecas and restituted as genus, to Hemilophini. Furthermore, our analyses suggest that Ostedes and Neohoplonotus should be removed from Acanthocinini and Parmenini, respectively, and Colobotheini should be redefined to encompass several genera currently placed in Acanthocinini.