Next Article in Journal
The Entomopathogenic Fungi Metarhizium anisopliae and Beauveria bassiana for Management of the Melon Fly Zeugodacus cucurbitae: Pathogenicity, Horizontal Transmission, and Compatability with Cuelure
Previous Article in Journal
Efficacy of Dust and Wettable Powder Formulation of Diatomaceous Earth (Detech®) in the Control of Tyrophagus putrescentiae (Schrank) (Acari: Acaridae)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Phylogenomic Analyses of the Tenthredinoidea Support the Familial Rank of Athaliidae (Insecta, Tenthredinoidea)

1
Laboratory of Insect Systematics and Evolutionary Biology, College of Life Sciences, Jiangxi Normal University, Nanchang 330022, China
2
Department of Molecular Biology and Genetics, Faculty of Science, Sivas Cumhuriyet University, Sivas 58140, Turkey
3
Institut de Systématique, Évolution, Biodiversité (ISYEB) Muséum National d’Histoire Naturelle, CNRS, Sorbonne Université, EPHE, Université des Antilles, CP50, 57 rue Cuvier, 75005 Paris, France
4
State Key Laboratory of Palaeobiology and Stratigraphy, Nanjing Institute of Geology and Palaeontology, Centre for Excellence in Life and Paleoenvironment, Chinese Academy of Sciences, Nanjing 210008, China
5
Univ. Rennes, CNRS, Géosciences Rennes, UMR 6118, F-35000 Rennes, France
6
CNRS, Institut des Sciences de l’Évolution de Montpellier, UMR 5554, 34090 Montpellier, France
*
Author to whom correspondence should be addressed.
Insects 2022, 13(10), 858; https://doi.org/10.3390/insects13100858
Submission received: 11 June 2022 / Revised: 7 September 2022 / Accepted: 14 September 2022 / Published: 21 September 2022
(This article belongs to the Section Insect Systematics, Phylogeny and Evolution)

Abstract

:

Simple Summary

In the current era of data explosion, the use of genetic information is increasingly being applied across numerous biological questions. One application has been to develop more robust evolutionary frameworks. Such well-resolved phylogenetic relationships are currently lacking from many of the basal branches of diversity-rich taxa. This is most pronounced at the base of the thentredinoid, especially Athalia. This study reviews earlier comparative morphological studies and advances in phylogenetic studies based on morphological characters and short sequence fragments, using both mitochondrial and nuclear genetic sequences as well as genomic structural evidence to define the family Athaliidae in several dimensions and clarify its phylogenetic position. As a result, Athaliidae no longer belongs to the Tenthredinidae but is independent and distant from it. This study clarifies a hurdle to solve the scientific problem of hymenopteran evolution.

Abstract

The systematic status of the genus Athalia and related genera is a perennial controversy in sawfly taxonomy. Several authors have hypothesized that the placement of Athalia within the Tenthredinidae is artificial, but no studies have focused on this topic. If the hypothesis that Athalia does not belong to Tenthredinidae can be supported, the taxonomic framework of Tenthredinoidea needs revision. We present a comprehensive phylogenomic study of Tenthredinoidae, focusing on the positions of Athalia and related genera by sampling 80 representatives mainly of the Tenthredinoidea, including Heptamelinae and Blasticotomidae. Our phylogenetic reconstructions based on nuclear genes and mitochondrial (mt) sequences support Athalia and related genera as a distinct clade sister to Tenthredinidae + (Cimbicidae + Diprionidae). A comparison of symphytan mitochondrial genomes reveals an innovative gene rearrangement pattern in Athaliidae, in which Dentathalia demonstrates a more ancestral pattern than Athalia and Hypsathalia. The lineage specificity of mt rRNA secondary structures also provides sufficient support to consider Athaliidae as a separate family. In summary, the phylogeny and genomic structural changes unanimously support the taxonomic treatment of Athaliidae as a family and the re-establishment of Dentathalia as a valid genus.

1. Introduction

Our current knowledge of the evolutionary relationships among insect orders is extensive and supported by numerous large-scale phylogenomic analyses [1,2,3,4,5,6,7,8,9]. However, inconsistencies persist regarding the phylogenetic relationships between a few groups (e.g., Psocoptera, Antliophora + Siphonaptera) and have challenged the perfect picture of insect evolution [1,10,11,12]. It is widely accepted that further advances in insect phylogenetics cannot rely only on deeper and broader genome and taxon sequencing. Improved modeling of the evolutionary process and a perpetually nuanced view are fundamental to clarifying and reducing the number of paraphyletic groups [13,14].
Genome sequences are the preferred data for phylogenetic relationship reconstruction because they contain abundant informative sites. Nevertheless, dissimilarity and inconsistency between data always exist. In contrast, rare genomic changes (RGCs) provide concise evidence [15]. The use of mitochondrial (mt) genomes, which are compact with nearly invariant gene contents, potentially offers lower error rates and a high potential to resolve deep nodes [16]. Even if gene orders (GOs) can be modeled as signed permutations, they cannot be treated directly as character matrices [17]. To extract phylogenetic information from GOs, specialized computational methods and tools have been developed. Among the model-based frameworks, a set of rearrangement operations known as the rearrangement model, pioneered by Watterson et al. [18], are used to measure pairwise distances. Then, it has been found that the tandem duplication random loss (TDRL) operation is more severe in rearranging GOs than other types of rearrangement operations in metazoans [19]. Several software tools that regard TDRL rearrangements were then developed for the reconstruction of evolutionary scenarios [20]. Notably, tandem duplication nonrandom loss mutations also occur [21], but they are not yet included in the abovementioned models. In addition to the model-based framework, the extraction of phylogenetic information directly from GOs without making assumptions about rearrangement operations is also useful. The results support one hypothesis among a set of competing hypotheses favored by different data [22]. Interestingly, the Hymenoptera are among the orders of insects with the highest variation in GOs [23]. The rearrangements occur in various taxa, with all three types of genes being variable [24]. As the sampling coverage increases, a large number of shared derived features are discovered. Many of these GOs provide additional support for monophyly [25] or for defining the boundaries of higher taxa [26]. These newly discovered mitochondrial genomes fill discontinuities and restore the continuous spectrum of evolutionary scenarios, calling for more active accumulation of data and a modification of the model.
In addition to GOs, other traits may be associated with strong selection, such as innovation in RNA genes, which are also considered to be useful in a phylogenetic context. Systematic studies of the evolution of non-coding RNA secondary structure have highlighted the importance of branch-specific structural insertion/deletion domains, but it appears that data accumulation was not sufficient to explore issues related to sequence signatures [27]. Moreover, the predicted secondary structures remain limited to a small number of taxa, thus preventing effective large-scale comparative studies. However, given the vast amount of data available because of next-generation sequencing (NGS) technology, the time has come to define the characteristics of predictive structures for application in phylogenetic studies.
Among the basal taxa of Hymenoptera, one of the most problematic paraphyletic groups is the Tenthredinidae (Figure 1). To clarify this issue, the position of the enigmatic genus Athalia Leach, 1817 and related genera need to be addressed.

1.1. Athalia and Its Relatives

Athalia is one of the twenty oldest described genera of Tenthredinidae andoccurs in the Eurasian and African regions [35].
Hennedyia Cameron, 1891 was described by Cameron [36]. The genus includes only one species and is found in southern Europe [37].
Benson [38] created the subgenus Dentathalia Benson, 1931, to accommodate the particular toothed claw of the species Athalia scutellariae Cameron, 1880 under the genus Athalia. However, later, Dentathalia was synonymized with Athalia by Benson [39].
Hennedyella Forsius, 1935 was described from Burma (=Myanmar) by Forsius [40]. The genus includes two species and occurs in Myanmar, Thailand [41], and the eastern Himalayas [42].
Benson [39] erected Hypsathalia Benson, 1962, for Athalia przevalskyi Jakovlev, 1887. The genus includes one known species and is distributed in Nepal, North India, and Tibet (China).

1.2. An Overview of the Systematic Position of Athalia and Its Relatives

The systematic positions of Athalia and its relatives, Hypsathalia, Hennedyia, and Hennedyella, are among the significant unresolved problems in the phylogeny of early-diverging lineages of Hymenoptera. Leach [35] classified Tenthredinidae s. lat. into nine stripes and placed the genus Athalia in stripe six with Messa Leach, 1817, Selandria Leach, 1817, and Fenusa Leach, 1817.
Hartig [43] divided Tenthredinidae s. lat. into four groups (Cimbicides, Hylotomides, Tenthredinides, and Lydides) and stated specifically that Athalia forms a “connecting link” between the Hylotomides and Tenthredinides.
Kirby [44] proposed the subfamily Athaliina, encompassing Athalia, and placed it within the family Tenthredinidae. In his system, Kirby divided the clades composing the “Symphyta” into two families: Tenthredinidae and Siricidae.
Cameron [45] recognized four families under Sessiliventris (=“Symphyta”). The family Tenthredinidae was classified into eight subfamilies, and Athalia was placed into the subfamily Tenthredina. Cameron pointed out that Athalia is a distinct and ‘ancient’ genus within the family.
Ashmead [46] divided the suborder Phytophaga into two series, Xylophaga and Phyllophaga. The genera Athalia and Hennedyia were placed into the Selandriinae of Selandriidae.
Konow [47] regarded Phyllophaga, except for Xyelidae and Lydidae, as one family: the Tenthredinidae. He subdivided the family Tenthredinidae into four subfamilies and placed Athalia into the tribe Selandriades (nearly equivalent to the current Selandriinae, Allantinae, and Cladiuchini) of Tenthredinini, and Hennedyia was placed into a heterogeneous tribe (Hoplocampides) with six other genera: Phyllotoma Fallen, 1829, Poppia Konow, 1904, Heptamelus Haliday, 1855, Anapeptamena Konow, 1898, Eriocampoides Konow, 1890, and Hoplocampa Hartig, 1837.
Rohwer [48] defined modern concepts of the family Tenthredinidae and the superfamily Tenthredinoidea. He further divided the family Tenthredinidae into eleven subfamilies, including Athaliinae, which lies between the Messinae (=Fenusinae) and the Empriinae (=Allantinae).
Benson [49] classified the Tenthredinidae into seven subfamilies, and the tribe Athaliini solely composed of Athalia was placed into the Emphytinae (=Allantinae). Note that the genera Hennedyia and Hennedyella were omitted from his system. Takeuchi [50] also placed the tribe Athaliini into Allantinae. However, Benson [51] changed his system and placed Athaliini into a broadly defined Blennocampinae (including Empriini, Allantini, Caliroini, Blennocampini, and Fenusini). Benson [39] revised the global genera and species of Athaliini. The three monotypic genera Hypsathalia, Hennedyia, and Hennedyella were placed into the Athaliini.
Lorenz and Kraus [52] studied the larval morphology of sawflies and divided the Tenthredinidae into six subfamilies (Dolerinae, Tenthredininae, Selandriinae, Blennocampinae, Nematinae, and Heterarthrinae). They placed the Athaliini into the comprehensive Blennocampinae, which corresponds now to the Allantinae, Blennocampinae, Fenusinini, Caliroini, Eriocampini, Athaliini, and Mesoneura Hartig, 1837.
Zhelochovtsev [53] divided the Tenthredinidae into two subfamilies (Tenthredininae and Nematinae), and the Athaliini were placed into the Tenthredininae.
Abe and Smith [54] divided the Tenthredinidae into eight subfamilies but without tribal arrangement. The genera Athalia, Hypsathalia, Hennedyia, and Hennedyella were placed into the Allantinae.
Wei [28] reconstructed the phylogeny of the Tenthredinidae and proposed a new system for the family. He divided the family Tenthredinidae into six families and 19 subfamilies (Figure 1a). The Athaliinae were placed into the Blennocampidae as the sister group of Lycaotinae. Wei and Nie [55] published a generic list of global Tenthredinoidea s. str. based on the data of Wei [28], first excluded the genus Athalia and its relatives from Allantinae and placed them as a subfamily, Athaliinae, under the family Blennocampidae.
Lacourt [56] classified the Tenthredinidae into 14 subfamilies, including Athaliinae, which lies between Selandriinae and Emphytinae.
Based on morphological characters, Vilhelmsen [29] reconstructed the phylogeny of the extant early-diverging lineages of Hymenoptera. Within the superfamily Tenthredinoidea, Athalia was located between the Blaticotomidae and other taxa of the superfamily (Figure 1b).
Schulmeister [30] revised the data matrix of Vilhelmsen [29], provided additional morphological data, and reconstructed the phylogeny of early-diverging Hymenoptera. This study demonstrated that the systematic position of the genus Athalia was uncertain, with one possible position being the first lineage of Tenthredinoidea except for Blasticotomidae (Figure 1c).
Since Schulmeister [31], Athalia has repeatedly been placed outside of the Tenthredinidae (Figure 1d), even though almost all of these studies were based on reduced taxon sampling (e.g., Figure 1e) [32,57,58]. In the studies of Boevé et al. (Figure 1f) [34] and Malm and Nyman (Figure 1g) [33] with more exhaustive taxon sampling, the genus Athalia was placed in an early-diverging position within the Tenthredinidae or as a rogue lineage in the tree due to the Heptamelinae, both poorly supported.
Vilhelmsen et al. [59] analyzed the phylogeny of apocritan wasps based on the structure of mesosoma. Four sawfly taxa were included in the study, and the results demonstrated that the relationship of the four taxa was (Notofenusa + (Athalia + (Monoctenus + Heteroperryia) or (Athalia + (Monoctenus + (Notofenusa + Heteroperryia).
Vilhelmsen [60] performed the most recent morphological phylogenetic analysis of Tenthredinidae. Although the generic and subfamilial relationships of Tenthredinidae were poorly resolved, his results indicated that Athalia was not a ‘normal member’ of Allantinae.
The results of these previous studies are of particular interest because they suggested that Athalia was possibly a distinct basal lineage of Tenthredinoidea (excluding Blasticotomidae). However, their scarce sampling, namely the lack of closely related genera such as Hypsathalia or Dentathalia, and limited molecular data (i.e., short DNA sequences), is not sufficient to support the monophyly of an “Athalia” clade. Such studies have led to confusion in tenthredinid classification, with some researchers placing them into Allantinae (Tenthredinidae). For example, Taeger et al. [61] compiled a world catalog of Symphyta and divided the Tenthredinidae into six subfamilies, and placed Athalia, Hypsathalia, Hennedyia, and Hennedyella into the Allantinae.
The present study aims to clarify the position of the genus Athalia and its relatives, for which we constructed a comprehensive time-calibrated phylogeny of Tenthredinoidea based on both mitochondrial and nuclear single-copy genes and by also increasing taxon coverage by including Hypsathalia and Dentathalia. We assessed the reliability of their taxonomic status by considering the evolutionary history and by investigating mt-genomic features among the “Symphyta”.

2. Materials and Methods

2.1. Taxon Sampling

The genome and/or mitogenome data of 80 sawfly species were used in this study (Table 1). The sampling includes 76 tenthredinoid species representing 53 genera from seven families and four xyeloid species as outgroups. The genome and mitogenome data of 23 and only mitogenome data of seven out of these species were sequenced and reported in this study. The specimens were provided by Jiangxi Normal University, China. The data of the remaining 44 species were retrieved from NCBI database (Table 1). From these species, RNA/DNA-seq FASTQ files of 30 species were downloaded from the GenBank SRA database, while the mitogenome sequences of 14 species were retrieved from the GenBank Nucleotide database.

2.2. DNA Extraction and Sequencing

Total genomic DNA was extracted from the thorax muscle of each ethanol-preserved specimen by the DNeasy Blood & Tissue Kit (Qiagen, Valencia, CA, USA) according to the manufacturer’s instructions, and the extracted DNA samples were quantified using a NanoDrop (Maestrogen, Inc., Waltham, MA, USA). The genomic DNA extracts were then pooled and subjected to 150-bp paired-end reads on the Illumina HiSeq 4000 platform (sequenced by Shanghai Majorbio Bio-pharm Technology Co., Ltd., Shanghai, China). Approximately 10 Gb raw data were generated for each library.

2.3. Genome Assembly and Single-Copy Assignment

The genome assemblies of NGS reads generated in this study and previously deposited to SRA database were performed using the pipeline plws v1.0.6 (https://github.com/xtmtd/PLWS, accessed on 1 October 2021) suggested by [91]. The reads were firstly grouped into clumps and duplicates were removed with clumpify.sh (BBTools suite version 38.91; [92]). The quality controls and normalizations were performed with bbduk.sh and bbnorm.sh under BBTools. The Minia v3.2.6 [93] was used for contig assembly with multiple k-mers and the REDUNDANS v0.13a5 [94] was preferred for detection and selective removal of redundant contigs. Scaffolding of the assembled contigs and gap-filling were performed with BESST v2.2.8 [95] and GAPCLOSER v1.12 [96], respectively. The scaffolds were then searched against hymenoptera_odb10.2020-08-05 database composed of 5991 genes using Benchmarking Universal Single-Copy Orthologs BUSCO v5.2.2 [97] to obtain the Single-copy Orthologs (SCOs) of each species. Orthologous sequences were combined in separate fasta files as amino acid sequences, removing duplicated and fragmentary genes.

2.4. Mitogenome Assembly, Annotation, and Structure Predictions

To construct the mitogenome assemblies of the species, the NGS reads were trimmed with BBDuk, and then the high-quality reads were assembled de novo using MIRA assembler implemented in Geneious R11 (Biomatters Ltd., Auckland, New Zealand) [98] The obtained assemblies were mapped using the reference symphytan mitogenomes (accession numbers: NC024664, NC045360, NC057102, NC056796, MG923517) under ‘medium-low sensitivity’ parameters for each species. The annotations of the mitogenomes were carried out using MITOS2 (http://mitos2.bioinf.uni-leipzig.de/index.py, accessed on 29 January 2021) [99]. The boundaries and locations of the protein-coding genes (PCGs) and rRNA genes were manually checked by comparing the reported symphytan homologous gene sequences. Considering the complexity of gene rearrangements in Hymenoptera, we performed the validation of genes and adjacent regions involved in gene rearrangements. For the same reason, the obtained reliable GOs were not used to reconstruct the ancestral state, but rather to compare the GO states for observing the fit between GO features and the phylogeny obtained by sequence-based methods. The procedure of rRNA secondary structure prediction is referred to in the previous study [100]. Instead of focusing on the variability within a taxon, this study is more concerned with the stability within family-level units and the inter-family variability, in other words, exploring potentially shared derivatives for different groups to obtain support for the taxon rank.

2.5. Phylogenomic Analyses

2.5.1. Alignment, Refinement, Supermatrix Construction, and Model Selection

Each fasta file including SCOs was aligned with MAFFT v7.471 [101] using the algorithm L-INS-i. Poorly and ambiguously aligned regions within the amino acid alignments were removed with Gblocks v0.91b [102] using default parameters except for setting the options ‘minimum number of sequences for a flank position’ to 15 and ‘minimum length of a block’ to 4. The supermatrix was then produced by concatenating the individual alignment files using concat-aln (https://github.com/aberer/concat-aln, accessed on 1 November 2021). The optimal partitioning scheme and the best-fitting substitution models were estimated by PartitionFinder v2.1.1 [103] for the SCOs dataset including the parameters of the among-site variation (+G) and parameters of amino acid frequency estimation (+F). The substitution models were further restricted to 11 amino-acid substitution models (LG + G, WAG + G, DCMUT + G, JTT + G, BLOSUM62 + G, LG + G + F, WAG + G + F, DCMUT + G + F, JTT + G + F, BLOSUM62 + G + F, LG4X) because these are the most estimated models for the insect studies [1,2].
The alignment of each PCGs of the mitochondrial gene was performed with the MAFFT algorithm implemented in Geneious R11, preferring the ‘translation align’ option. The rRNA was performed with MAFFT using the algorithm E-INS-i. The resulting alignments were concatenated using PhyloSuite [104], and the PCG, PCG + RNA, as well as AA datasets, were generated. Given the compositional heterogeneity across the datasets including PCG, the CAT-GTR model in PhyloBayes [105] was used. The cross-validation analyses were performed for AA datasets under the homogeneous models (MtArt) and heterogeneous models (CAT + GTR).

2.5.2. Phylogenomic Inferences

The phylogenomic reconstruction of Tenthredinoidea based on the supermatrix of SCOs was carried out using the maximum likelihood (ML) and Bayesian inference (BI) approaches. Two xyeloid species were selected as outgroups: Xyela alpigena (Strobl, 1895) and Megaxyela euchroma Blank, Shinohara and Wei, 2017. The ML analysis was performed in IQ-TREE v2.1.4 [106] using default parameters under the Q.plant + R10 (-lnL: 14280658.547) evolutionary model inferred by PartitionFinder v2.1.1. Node support was estimated with 10000 bootstrap replicates using the fast bootstrapping option implemented in IQTree. The BI analysis was carried out using ExaBayes v1.5.1 [107] under default parameters with automatic substitution model detection. Four coupled Markov chain Monte Carlo (MCMC) chains were run with one million generations by sampling every 500 generations. To check whether the chains had achieved stationarity, ‘burn-in’ plots were evaluated by plotting log-likelihood scores and tree lengths against generation numbers using the software Tracer v1.6 [108]. After assessing for convergence, the first 20% of trees were discarded as burn-in and a majority-rule consensus tree (BI tree) was generated using Geneious R11 from the remaining trees. Visualization of the trees was performed by FigTree v1.4.2 [109].
For the mitochondrial datasets, Bayesian analyses were run with Phylobayes-MPI 1.9 [105]. For all analyses, two runs were performed and convergence was investigated using the bpcomp option. The analysis was stopped when the conditions considered to indicate a good run were reached (maxdiff < 0.1) or sufficient effective sample size was reached (effsize > 300).

2.5.3. Divergence Time Estimation

Divergence times of the tenthredinoid lineages were estimated using the supermatrices of SCOs and mitochondrial genes by MCMCTree program implemented in the Phylogenetic Analysis by Maximum Likelihood (PAML) package v4.9 [110]. Three fossil calibration points were used: (i) Triassoxyela foveolata Rasnitsyn, 1964 for stem calibration of Xyelidae, (ii) Pseudoxyelocerus bascharagensis Nel, Petrulevicius and Henrotay, 2004 for stem calibration of Tenthredinoidea and (iii) Abrotoxyela lepida Gao, Ren and Shih, 2009 for the crown calibration of Xyelidae. Two additional molecular calibration points were set as (i) 168.5 Ma for the split between Pergidae and Argidae [111], and (ii) 73 Ma for the crown age of Cimbicidae [26]. The oldest estimated age for holometabolous insects [112] was preferred for the constraint of the root (<400 Ma). Substitution rate per site for the supermatrix of SCOs was estimated by CODEML and was used to set the prior for the mean substitution rate in the MCMCTree. The MCMC was run by 50 × 10,000 iterations with the JTT substitution model [113]. The convergence was assessed by considering effective sample sizes (ESS > 200) using Tracer v1.7 [108] and the maximum clade credibility tree was generated by TreeAnnotator v1.10.4 after removing 25% of the trees as burn-in. The trees produced by both datasets were visualized in FigTree v1.4.2 [109].

3. Results

3.1. Phylogenomic Assessment of the Placement of Athalia and Its Relatives

3.1.1. Phylogenetic Analysis Based on Sequences

The mitochondrial DNA sequence (Mt) and single-copy orthologs (SCOs), representing two different markers, yielded consistent interfamily relationships. The monophyly and the position of the ‘Athalia’ clade were supported with high support values across all analyses (Figure 2, Figure 3 and Figure 4). When using Xyelidae for rooting, the Blasticotomidae were an earlier-diverging lineage. The tree then split into two branches, one of which consisted of Argidae and Pergidae, with a predominantly Southern Hemisphere distribution. The other branch consisted of families distributed mainly in the Northern Hemisphere, with a basal ‘Athalia’ clade. Diprionidae and Cimbicidae were supported as sister clades, forming the closest relatives of Tenthredinidae s. str. All of the above clades were recovered as monophyletic with high posterior probabilities (PPs, >0.90) except for the Pergidae, which has only one representative, so monophyly of this family cannot be ensured.
The extant Athaliidae are distinct from Tenthredinidae, including Heptamelinae, and supported by the following morphological characteristics: 1. Anepimeron largely and roundly convex with only a small basin above (apomorphic; the basic state of anepimeron is flat, as shown in Xyelidae); 2. Pleural suture distinctly curved in the upper third (apomorphic); 3. Specialized penis valve (apomorphic); 4. Prepectus large and isolated (plesiomorphic); 5. vein 1 M meeting vein Rs instead of vein R (plesiomorphic; reversed in some genera of Tenthredinidae); and 6. Long and filiform antenna with more than 10 antennomeres (plesiomorphic; reversed in some genera of Tenthredinidae). The convex anepimeron is also present in the family Argidae, but it is convex overall and without a small basin above. In Cimbicidae, the anepimeron is generally more or less concave, but in several genera, the lower end of the anepimeron is distinctly convex. In Diprionidae, the whole anepimeron is elevated to form a platform.
Unlike the consistent high-level relationships, the relationships between subfamilies or equivalent levels varied with datasets and models. The SCO under ML analysis and PCG datasets supported the Heptamelinae as the early diverged group of Tenthredinidae; nevertheless, the support for this position was low (PP = 0.54) in PCGs (Figure 3a) and median in SCOs (Figure 4a, 0.86). In the analyses of AAs under both models (Figure 2) and of PCGs + RNA (Figure 3b), the Heptamelinae were recovered as the sister of (Cimbicidae + Diprionidae), with support for this position ranging from 0.59 to 0.92, and in the BI analyses of SCOs (Figure 4b), it was supported with placement at the base of (Tenthredinidae + (Cimbicidae + Diprionidae)). Whether it should be separated from the Tenthredinidae as an independent family needs a full reassessment. Another inconsistency occurs in Selandriinae and Nematinae. They are consistently located at the base of Tenthredinidae (excluding Heptamelinae); however, they are successively split out in two AA (Figure 2, PP > 0.81) and SCOs analyses (Figure 4) and exhibit a sister group relationship in the PCG-RNA (Figure 3b, PP = 0.51) analysis or form an unresolved triple branch with other species in the PCG (Figure 3a, PP = 0.99) analysis.
At least four subclades can be recognized within the ‘Athalia’ clade. In addition to Dentathalia and Hypsathalia, A. ‘tibetana’ morphotype() and A. tanaoserrula Chu & Wang, 1962 formed an independent subclade. All three cases of the relative positions of the latter two subclades were recovered, with both being sister groups to each other (Figure 3b: PCG-RNA, PP = 0.89) or diverging successively but in a different order. The subclade of A. ‘tibetana’ and A. tanaoserrula at the basal position was supported in SCOs (Figure 4), as well as AA-CAT-GTR (Figure 2a, PP > 0.72) and PCG (Figure 3a, PP > 0.98). This suggests that the diversity within the ‘Athalia’ clade has not yet been fully captured.

3.1.2. Rare Genomic Changes

To compare rare gene changes, we mapped GOs and secondary structure onto the phylogenetic backbone obtained from the PCG dataset (Figure 5).
In the ‘Athalia’ clade, trnC and trnW had moved upstream relative to those in Xyelidae, leaving a single trnY at the nad2-cox1 junction. In addition, the shuffling within the trn-IMQ cluster was also an autapomorphy. While trnT and trnP were variable across the ‘Athalia’ clade, they swapped positions at least twice. If the GO of the Xyelidae is representative of the ancestral state of Hymenoptera, then gene rearrangement seems to have occurred in two directions in the subsequent divergence. One is the minor rearrangement represented by Blasticotomidae, and the other is rather major, which involved the tRNA crossing of nad2.
The base composition and predicted secondary structures are conserved across the diverse taxa in the ‘Athalia’ clade (Figures S1 and S2). In rrnS, domain III was more conserved within the ‘Athalia’ clade. In particular, helix H939 was completely conserved among the twelve samples. However, there were only three known patterns: the shared pattern of Xyelidae and Orussidae; the pattern of Cephidae, Siricidae, and Pamphilioidea; and the pattern of the Tenthredinoidea including Blasticotomidae. Compensatory base-pair changes (CBCs) have taken place in the second and fifth pairs. The phylogenetic signals carried by this helix seem to be more suitable for defining higher-level diversity. In rrnL, helices H533, H563, H579, H736, H822, H1775, H1830, H2023, H2259, and H2064 are completely conserved. However, only some of them can be treated as signals at multiple hierarchies.

3.2. Divergence Time Estimates

There was a slight variance between the topologically constrained analyses, broadly overlapping 95% credible intervals for most nodes (Figure 6). Two inferences consistently restrict several early divergences, including the origin of Athaliidae, to the pre-Cretaceous. For Athaliidae specifically, these included 186.86 Ma (95% HPD = 167~206 Ma) inferred from mtG analysis or 168.33 Ma (95% HPD = 162~174 Ma) based on the SCO dataset. The estimates for the age of crown Athaliidae differed strongly, depending on the datasets. However, both results fell within the Cretaceous, ranging from 133.18 Ma (95% HPD = 109~156 Ma) inferred from mtG analysis to 143.46 Ma (95% HPD = 127~171 Ma) based on the SCOs dataset. The phase of species diversification could not be captured due to the scarcity of species-level samples.

4. Discussion

4.1. Sequence-Based Phylogeny, Rare Genomic Changes, and Morphology: Congruence in the Placement of Athaliidae

Resistance to the independence of Athaliidae from Tenthredinidae s. str. comes from its morphological similarity to Allantinae [61]. Our results indicate that such subjective equivalencies can mask profound evolutionary differences, as argued by Sterelny [114]. A more detailed morphological comparison reveals that the wing venation pattern originally thought to be shared by the Allantinae and Athalia is a plesiomorphic character, which does not support their monophyly, but they even have different origins and represent different evolutionary pathways. The wing venation of the Athaliidae has a consistent hint of curvature, which is a continuation of an ancient wing pattern of Hymenoptera. In contrast, the wing venation of Allantinae does not have such a bend and tends to be simplified, a derived feature in Hymenoptera [115,116].
Fossils provide complementary evidence for the placement of Athaliidae. Palaeathalia laiyangensis Zhang, 1985, discovered in eastern China, resembles extant Hennedyella by sharing numerous plesiomorphic characters, such as antenna long and slender with more than 13 antennomeres, cell 1 M broader with a short dorsal petiole, and the vein cu-a meeting cell 1 M at the middle. The main differences between Palaeathalia and Hennedyella are that cell C of the forewing in Palaeathalia is very broad, and the head is strongly dilated behind the eyes in dorsal view. Within the extant taxa of Athaliidae, only Hypsathalia shares a very broad cell C in the forewing. If we regard Palaeathalia laiyangensis as an ancient fossil representative of the former Allantinae, it pushes the divergence of this subfamily to at least 127 Ma. However, a long gap exists between it and the earliest confirmed record of the now-dominant Tenthredinidae in the early Paleocene (Tenthredo primordialis Piton, 1940, 61 Ma) [117]. If Athaliidae is located outside of Tenthredinidae and (Cimibicidae + Diprionidae), as this study reveals, then it would make sense for this clade to be represented by ancient fossils.
This speculation has received support from molecular analyses [25,31,33] and some morphological studies [28,55] demonstrating that Athalia and its relatives differ from the Allantinae. Uncertainties in their position come from the following two issues: (1) the relative positions of the Athaliidae and Heptamelinae and, (2) the rank of the monophyletic clade including Athalia and its relatives. In other words, whether the Athaliidae have a large enough gap with others to match the family rank. Phylogenomics provides the confidence needed to choose among existing hypotheses and sheds new light on genome evolution.
Sequence-based phylogenetic analyses of both mt genome data and SCOs genome data support the monophyletic Athaliidae standing between (Argidae + Pergidae) and (Cimbicidae + Diprionidae). Beyond the linear sequence comparisons, the genome-level characters also support this relationship (Figure 5). Athaliidae and all other families have their own unique ground patterns of GOs, which not only are incompatible between families but also allow the inference of reasonable GOs evolution scenarios in a broader evolutionary context (Figure 5). The GO of Heptamelinae is more similar to that of the Tenthredinidae. If it stands in place of present Athaliidae, that is, between (Argidae + Pergidae) and (Cimbicidae + Diprionidae) [33], it indicates either parallel evolution in Heptamelinae or reversal to an ancestral GO pattern in Athaliidae.
Structural inference is difficult and tedious and hence avoided in the overwhelming majority of published studies [118]. However, the intensive sampling here reduces the uncertainty in this process. The predicted structures (Figures S1 and S2) are highly conserved, while the mutation sites exhibit robust phylogenetic signals. Compared with rrnS, the pairing on the stem of rrnL shown in Figure 5 has strong lineage specificity. These discrete characters imply evolutionary independence above the species level. It could be argued that the mt-genomic distinctness of Athalia and its relatives warrant recognition at the rank of family, as could be claimed for all of the well-supported primary lineages within Tenthredinoidea.
In summary, the Athaliidae should be recognized as an isolated family that falls outside of Tenthredinidae, Cimbicidae, and Diprionidae. We suggest that the Tenthredinoidea, therefore, be considered an evolutionary clade consisting of six well-defined families and a ‘core Tenthredinidae’. Among them, three Neotropical/Southern Hemisphere families, Argidae, Pergidae, and Zenargidae, form a clade sister to the predominantly East Asian taxa, which include three small families, Cimbicidae, Diprionidae, and Athaliidae, and the ‘core Tenthredinidae’.

4.2. Spatial and Temporal Diversification of Athaliidae

The morphology of the clypeus is a crucial trait in understanding the classification of Athalia [39]. The backbone in this study corresponds in a general way to clypeus evolution. For example, A. ‘tibetana’ and A. tanaoserrula Chu and Wang, 1962 correspond to the ‘A. cordata group’ in having a subsquare clypeus, while A. icar Saini and Vasu, 1997 and all other samples of Athalia form a monophyletic lineage corresponding to the ‘A. rosae group’ [39] in having a short and weakly convexed clypeus. However, morphological studies indicate that each clypeus type is present in species from two or more distribution types. This study on the phylogeography of Athaliidae reveals that the ‘A. vollenhoveni group’ (Benson, 1962; clypeus narrowly incised) may be the sister group of the ‘A. roase group’, while the ‘A. himantopus group’ (clypeus broadly and very shallowly emarginated) is possibly related to the ‘A. bicolor group’ and ‘A. cordata group’, and Benson’s ‘rosae group’ (clypeus short with anterior margin weakly convex) is probably not monophyletic (Niu et al., in preparation). Revision of the family still requires the addition of representative samples from African, Indian, and European taxa.
Malaise [60] discussed the biogeography of Athalia in detail and its absence from America. According to his theory, Athalia was relatively ancient and already present early in the Cenozoic. This hypothesis was later supported by the description of fossil species of the genus from the Eocene by Florissant [119]. Benson [39] discussed the distribution of Athaliini. He pointed out that the two genera Hennedyia and Hennedyella were known only from the western Mediterranean and Burma, respectively, and were probably from the early Cenozoic. The high Himalayan genus Hypsathalia was likely derived from lowland Cenozoic stock, possibly the same stock that gave rise to some species of Athalia. Benson [39] also discussed the distribution of characters and the possible origin of some species groups of Athalia from Europe and Asia. The origin of the diversified species of Athalia in the Afrotropical region was uncertain in his opinion.
Our analyses suggest that the most recent common ancestor (MRCA) for the crown Athaliidae lived in the Late Cretaceous. This coincides with the origin of their primary hosts [120]. However, the Athaliidae originated before the Cretaceous (186 Ma), yet, the Brassicales and Lamiales are both attributed to the later clades in the Cretaceous radiation of the core eudicots (90–87 Ma). In other words, the origin of the hosts of the extant species emerged later than the stem group of Athaliidae. It can be speculated that the MRCA of Athaliidae may have undergone host transfer during angiosperm radiation in the Cretaceous. This may explain the uniqueness of the Athaliidae in terms of the host. Araceae and Sedum L., 1753 were previously reported as hosts for the Eurasian species. They represent Alismatales and Saxifragales, which originated approximately 150 Ma and 126 Ma, respectively. They may be candidates for the ancestral host plants of Athaliidae.
When considering that the origin of the Brassicaceae [121] and Lamiales [122] occurred nearly simultaneously at both the order level (90~87 Ma) and the family level (35 Ma), it is more likely that the early divergence of Athaliidae underwent at least two completely different evolutionary trajectories, accompanying the rapid rise of the major clades of angiosperms. This of course cannot exclude the possibility of host shifts between Lamiales and Brassicaceae [123].
Most of the East Asian taxa of Athaliidae exhibit a distinctly plain, low-mountain distribution type, feeding on cultivated crucifers. However, some branches may be adapted to high-altitude habitats and are more likely to feed on Lamiales. Clades occurring on the Qinghai–Tibetan Plateau (QTP) and in adjacent areas show phylogenetic diversity in either alpine grassland (central and western Tibet) or alpine meadows (Qinghai, northeastern Tibet). Based on a few observations during field collection, it is likely that the representative dominant taxa of the plateau, Sedum, are their new hosts.
Unlike the several Eurasian widespread basal branches of the Athaliidae, three other ancient species, Hennedyia annulitarsis Cameron, 1891, Hennedyella athaloides Forsius, 1935 and Hennedyella typica Saini and Vasu, 1996, are distributed in Gibraltar and the Eastern Himalayas. A plausible explanation for the distribution of ancient species at the edge of the diversity center is competition with other successful radiating clades. When new branches acquire ‘key innovations’ such as gene rearrangements, they likely have wider niches or richer diversity, while older branches are likely to suffer.
Such a case is prominent in the tRNA cluster upstream of nad2; for the first time in Hymenoptera, trnC was transposed upstream of nad2 from the WCY cluster downstream of nad2. Specifically, the tRNA clusters showed the pattern CWMQI in Dentathalia and CWQMI in the rest of Athaliidae. Considering the ground pattern of the members within the sister clade of Athaliidae (Cimbicidae: MQI; Tenthredinidae: MQI), it is presumed that the GO of Dentathalia represents the primitive type. Given the uniqueness of its genome structure and several primitive morphological characters, we propose that the basal lineage, Dentathalia, be recognized at the genus level. In addition, A. ‘tibetana’ and A. tanaoserrula also have a unique TP shift, and such an event has been reported only for the Diprionidae. Its phylogenetic signal and role in the mitogenome structure evolution of Hymenoptera still need to be further evaluated. However, the rearrangement was also accompanied by specificity in the CG content of the flanking PCGs. The possibility that this clade should be considered an independent genus cannot be ruled out, and it is therefore recommended that the genus-level status of Hypsathalia should be retained (with at least four undescribed species distributed on the QTP), as being supported by placement between (A. ‘tibetana’ + A. tanaoserrula) and Dentathalia.
Despite being able to confidently reestablish the genus-level status of Dentathlia and retain the genus-level status of Hypthathalia, this work was not able to provide sufficient resolution to understand the tribal classification within Athaliidae. Although some major lineages could be recognized, the absence of African species has hindered the further study of evolution and diversification patterns within and among clades. Thus, it is too early to propose a formal tribal classification of this family.
Several issues remain to be addressed. The inference that the reversals would be a small-probability event when complexity is taken into account satisfies one of the principal criteria for a more accurate phylogenetic marker [15]. However, this inference was at high risk of bias for the Hymenoptera. For example, trnTP rearrangements occur in two distant taxa, one Athalia clade, and the Diprionidae. On the other hand, gene rearrangements appear to be a key innovation that triggers diversification in Athaliidae; however, not all clades demonstrate the independent pattern. Therefore, we doubt that gene rearrangement itself was a direct driver of increased generic diversification. It can also be reasoned that the direct observation of rearrangement mechanisms acting on microevolution in genus-level diversification may be a very small-probability event. However, this does not prevent GO, as a feature of multiple hierarchical levels, from being considered as evidence supporting macroevolutionary results. The coexistence of parallelism and variability among proxies for low-level clades as well as at the family level in Hymenoptera suggests that the application of GO changes needs more exploration.
This evolutionary clade of the Athaliidae does not seem to exhibit key morphological innovations compared to the next bursts of Tenthredinidae, which is the reason why it was proposed to stay within the family. However, when gene sequences and genomic structures were observed, we were able to speculate that some traits originated long before. That is, there is a macroevolution lag [124] between the origin of the Athaliidae and the crown Athaliidae and the subsequent dramatic diversity of the Tenthredinidae. Considering the probable existence of feeding exploration in Athaliidae and their morphological similarity to the Allantiniae, we have to consider the possibility that, as a result, the Athaliidae appear to be similar to a clade of Tenthredinidae in a preadapted manner. In other words, at the origin of the Athaliidae, the Hymenoptera appeared intrinsically ready to undergo diversification until an external factor provided a ‘critical opportunity’ [125].

5. Conclusions

The present study comprehensively assessed the position of Athalia and its relatives using sequence-based methods and genomic characterization. Inferences under the phylogenomic framework revealed a monophyletic Athaliidae positioned between Blasticotomidae and the rest of Tenthredinoidea. The mitogenome characters provided strong phylogenetic information to support the independence of the respective families. The present results call for a new classification of Tenthredinoidea. The diversification within the Athaliidae dated back to at least the upper Cretaceous, with the second phase of divergence occurring shortly after the K-Pg extinction. It is reasonable to speculate that the diverse fauna has a Eurasian origin, while the African fauna is secondary and may have multiple origins. With the uplift of the Himalayas, at least two clades adapted to the cold and settled at a 3000-m altitude, probably favoring Sedum. Most East Asian species stayed in the plain region and colonized a variety of Brassicaceae. However, only one plateau branch has gene rearrangements and a specific base composition of the corresponding genes, which calls for denser sampling to clarify the intergeneric relationships of Athalia. In addition, increasing data from African and European species for phylogeographic studies will be essential to understanding the Eurasian–African distribution pattern of extant insects.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/insects13100858/s1: Figure S1, Predicted rrnS secondary structures in the Athalia icai mitochondrial genome and the alignments of Athaliidae samples using A. icai as the reference. The structure LOGO and the identity are shown above the alignment. Figure S2, Predicted rrnL secondary structures in the Athalia icai mitochondrial genome and the alignments of Athaliidae samples using A. icai as the reference. The structure LOGO and the identity are shown above the alignment.

Author Contributions

Conceptualization, M.W.; methodology, G.N. and M.W.; validation, G.N. and E.M.K.; formal analysis, S.W., Ö.D. and M.B.; investigation, G.N.; resources, M.W.; data curation, G.N.; writing—original draft, G.N. and E.M.K.; writing—review and editing, C.J., A.N. and C.C.; visualization, M.L., S.W. and G.N.; supervision, M.W.; project administration, M.W.; funding acquisition, G.N. and M.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 (31501885, 31970447) and the Jiangxi Provincial Natural Science Foundation (20202BABL213044).

Data Availability Statement

The data presented in this study are openly available in Figshare at https://figshare.com/account/home#/projects/141053, accessed on 8 June 2022.

Acknowledgments

The members of the Lab of Insect Systematics and Evolutionary Biology (LISEB) at Jiangxi Normal University and Evolutionary Bioinformatics Research Groups (EBRG) at Sivas Cumhuriyet University are thanked for their contributions to laboratory work. The authors thank Wang Jin for the assistance with figure editing. We thank the anonymous reviewers for their careful reading and many constructive comments.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Misof, B.; Liu, S.; Meusemann, K.; Peters, R.S.; Donath, A.; Mayer, C.; Frandsen, P.B.; Ware, J.; Flouri, T.; Beutel, R.G.; et al. Phylogenomics Resolves the Timing and Pattern of Insect Evolution. Science 2014, 346, 763–767. [Google Scholar] [CrossRef] [PubMed]
  2. Peters, R.S.; Krogmann, L.; Mayer, C.; Donath, A.; Gunkel, S.; Meusemann, K.; Kozlov, A.; Podsiadlowski, L.; Petersen, M.; Lanfear, R.; et al. Evolutionary History of the Hymenoptera. Curr. Biol. 2017, 27, 1013–1018. [Google Scholar] [CrossRef] [PubMed]
  3. Johnson, K.P.; Dietrich, C.H.; Friedrich, F.; Beutel, R.G.; Wipfler, B.; Peters, R.S.; Allen, J.M.; Petersen, M.; Donath, A.; Walden, K.K.O.; et al. Phylogenomics and the Evolution of Hemipteroid Insects. Proc. Natl. Acad. Sci. USA 2018, 115, 12775–12780. [Google Scholar] [CrossRef] [PubMed]
  4. Wipfler, B.; Letsch, H.; Frandsen, P.B.; Kapli, P.; Mayer, C.; Bartel, D.; Buckley, T.R.; Donath, A.; Edgerly-Rooks, J.S.; Fujita, M.; et al. Evolutionary History of Polyneoptera and Its Implications for Our Understanding of Early Winged Insects. Proc. Natl. Acad. Sci. USA 2019, 116, 3024–3029. [Google Scholar] [CrossRef] [PubMed]
  5. Evangelista, D.A.; Wipfler, B.; Béthoux, O.; Donath, A.; Fujita, M.; Kohli, M.K.; Legendre, F.; Liu, S.; Machida, R.; Misof, B.; et al. An Integrative Phylogenomic Approach Illuminates the Evolutionary History of Cockroaches and Termites (Blattodea). Proc. R. Soc. B. 2019, 286, 20182076. [Google Scholar] [CrossRef]
  6. Kawahara, A.Y.; Plotkin, D.; Espeland, M.; Meusemann, K.; Toussaint, E.F.A.; Donath, A.; Gimnich, F.; Frandsen, P.B.; Zwick, A.; Dos Reis, M.; et al. Phylogenomics Reveals the Evolutionary Timing and Pattern of Butterflies and Moths. Proc. Natl. Acad. Sci. USA 2019, 116, 22657–22663. [Google Scholar] [CrossRef]
  7. McKenna, D.D.; Shin, S.; Ahrens, D.; Balke, M.; Beza-Beza, C.; Clarke, D.J.; Donath, A.; Escalona, H.E.; Friedrich, F.; Letsch, H.; et al. The evolution and genomic basis of beetle diversity. Proc. Natl. Acad. Sci. USA 2019, 116, 24729–24737. [Google Scholar] [CrossRef]
  8. Vasilikopoulos, A.; Misof, B.; Meusemann, K.; Lieberz, D.; Flouri, T.; Beutel, R.G.; Niehuis, O.; Wappler, T.; Rust, J.; Peters, R.S.; et al. An Integrative Phylogenomic Approach to Elucidate the Evolutionary History and Divergence Times of Neuropterida (Insecta: Holometabola). BMC Evol. Biol. 2020, 20, 64. [Google Scholar] [CrossRef]
  9. Tihelka, E.; Cai, C.; Giacomelli, M.; Lozano-Fernandez, J.; Rota-Stabelli, O.; Huang, D.; Engel, M.S.; Donoghue, P.C.J.; Pisani, D. The Evolution of Insect Biodiversity. Curr. Biol. 2021, 31, R1299–R1311. [Google Scholar] [CrossRef]
  10. Tihelka, E.; Engel, M.S.; Lozano-Fernandez, J.; Giacomelli, M.; Yin, Z.; Rota-Stabelli, O.; Huang, D.; Pisani, D.; Donoghue, P.C.J.; Cai, C. Compositional Phylogenomic Modelling Resolves the ‘Zoraptera Problem’: Zoraptera Are Sister to All Other Polyneopteran Insects. bioRxiv 2021. [Google Scholar]
  11. Whiting, M.F. Phylogeny of the Holometabolous Insect Orders: Molecular Evidence. Zool. Scr. 2002, 31, 3–15. [Google Scholar] [CrossRef]
  12. Wiegmann, B.M.; Trautwein, M.D.; Winkler, I.S.; Barr, N.B.; Kim, J.-W.; Lambkin, C.; Bertone, M.A.; Cassel, B.K.; Bayless, K.M.; Heimberg, A.M.; et al. Episodic Radiations in the Fly Tree of Life. Proc. Natl. Acad. Sci. USA 2011, 108, 5690–5695. [Google Scholar] [CrossRef] [PubMed]
  13. Kapli, P.; Flouri, T.; Telford, M.J. Systematic Errors in Phylogenetic Trees. Curr. Biol. 2021, 31, R59–R64. [Google Scholar] [CrossRef] [PubMed]
  14. Donoghue, M.J. Key Innovations, Convergence, and Success: Macroevolutionary Lessons from Plant Phylogeny. Paleobiology 2005, 31, 77–93. [Google Scholar] [CrossRef]
  15. Rokas, A.; Holland, P.W.H. Rare Genomic Changes as a Tool for Phylogenetics. Trends Ecol. Evol. 2000, 15, 454–459. [Google Scholar] [CrossRef]
  16. Moret, B.M.E.; Tang, J.; Warnow, T.J. Reconstructing Phylogenies from Gene-Content and Gene-Order Data. In Proceedings of the Mathematics of Evolution and Phylogeny, Paris, France, 17–18 June 2007. [Google Scholar]
  17. Donath, A.; Stadler, P.F. 25 Molecular Morphology: Higher Order Characters Derivable from Sequence Information. In Deep Metazoan Phylogeny: The Backbone of the Tree of Life; De Gruyter: Berlin, Germany, 2014; pp. 549–562. ISBN 978-3-11-027752-4. [Google Scholar]
  18. Watterson, G.A.; Ewens, W.J.; Hall, T.E.; Morgan, A. The Chromosome Inversion Problem. J. Theor. Biol. 1982, 99, 1–7. [Google Scholar] [CrossRef]
  19. Bernt, M.; Merkle, D.; Middendorf, M.; Schierwater, B.; Schlegel, M.; Stadler, P.F. 22 Computational Methods for the Analysis of Mitochondrial Genome Rearrangements. In Deep Metazoan Phylogeny: The Backbone of the Tree of Life; Wägele, J.W., Bartolomaeus, T., Eds.; De Gruyter: Berlin, Germany, 2014; pp. 515–530. ISBN 978-3-11-027746-3. [Google Scholar]
  20. Hartmann, T.; Bernt, M.; Middendorf, M. EqualTDRL: Illustrating Equivalent Tandem Duplication Random Loss Rearrangements. BMC Bioinform. 2018, 19, 192. [Google Scholar] [CrossRef]
  21. Lavrov, D.V.; Boore, J.; Brown, W. Complete MtDNA Sequences of Two Millipedes Suggest a New Model for Mitochondrial Gene Rearrangements: Duplication and Non-Random Loss. Mol. Biol. Evol. 2002, 19, 163–169. [Google Scholar] [CrossRef]
  22. Podsiadlowski, L.; Mwinyi, A.; Lesný, P.; Bartolomaeus, T. 19 Mitochondrial Gene Order in Metazoa—Theme and Variations. In Deep Metazoan Phylogeny: The Backbone of the Tree of Life; Wägele, J.W., Bartolomaeus, T., Eds.; De Gruyter: Berlin, Germany, 2014; pp. 459–472. ISBN 978-3-11-027746-3. [Google Scholar]
  23. Moreno-Carmona, M.; Cameron, S.L.; Prada Quiroga, C.F. How Are the Mitochondrial Genomes Reorganized in Hexapoda? Differential Evolution and the First Report of Convergences within Hexapoda. Gene 2021, 791, 145719. [Google Scholar] [CrossRef]
  24. Dowton, M.; Cameron, S.L.; Dowavic, J.I.; Austin, A.D.; Whiting, M.F. Characterization of 67 Mitochondrial TRNA Gene Rearrangements in the Hymenoptera Suggests That Mitochondrial TRNA Gene Position Is Selectively Neutral. Mol. Biol. Evol. 2009, 26, 1607–1617. [Google Scholar] [CrossRef]
  25. Niu, G.; Jiang, S.; Doğan, Ö.; Korkmaz, E.M.; Budak, M.; Wu, D.; Wei, M. Mitochondrial Phylogenomics of Tenthredinidae (Hymenoptera: Tenthredinoidea) Supports the Monophyly of Megabelesesinae as a Subfamily. Insects 2021, 12, 495. [Google Scholar] [CrossRef] [PubMed]
  26. Niu, G.; Cheng, Y.; Cai, C.; Meng, G.; Yan, Y.; Sun, Z.; Wan, S.; Wei, M. Phylogenomics Illuminates the Divergence Timeline and the Dynamic Mitogenome of Cimbicidae. Insects 2022, 13, 858. [Google Scholar] [CrossRef]
  27. Donath, A.; Findeiβ, S.; Hertel, J.; Marz, M.; Otto, W.; Schulz, C.; Stadler, P.F.; Wirth, S. Noncoding RNA. In Evolutionary Genomics and Systems Biology; Caetano-Anollés, G., Ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2010; pp. 251–293. ISBN 978-0-470-57041-8. [Google Scholar]
  28. Wei, M. Studies on the Phylogeny of Tenthredinoidea (s.Str.) With the Taxonomy of Chinese Tenthredinoidea Excl. Nematidae & Tenthredininae. Ph.D. Thesis, Institute of Zoology, Academia Sinica, Beijing, China, 1994. [Google Scholar]
  29. Vilhelmsen, L. Phylogeny and Classification of the Extant Basal Lineages of the Hymenoptera (Insecta). Zool. J. Linn. Soc. 2001, 131, 393–442. [Google Scholar] [CrossRef]
  30. Schulmeister, S. Review of Morphological Evidence on the Phylogeny of Basal Hymenoptera (Insecta), with a Discussion of the Ordering of Characters. Biol. J. Linn. Soc. 2003, 79, 209–243. [Google Scholar] [CrossRef]
  31. Schulmeister, S. Simultaneous Analysis of Basal Hymenoptera (Insecta): Introducing Robust-Choice Sensitivity Analysis. Biol. J. Linn. Soc. 2003, 79, 245–275. [Google Scholar] [CrossRef]
  32. Ronquist, F.; Klopfstein, S.; Vilhelmsen, L.; Schulmeister, S.; Murray, D.L.; Rasnitsyn, A.P. A Total-Evidence Approach to Dating with Fossils, Applied to the Early Radiation of the Hymenoptera. Syst. Biol. 2012, 61, 973–999. [Google Scholar] [CrossRef] [PubMed]
  33. Malm, T.; Nyman, T. Phylogeny of the Symphytan Grade of Hymenoptera: New Pieces into the Old Jigsaw (Fly) Puzzle. Cladistics 2015, 31, 1–17. [Google Scholar] [CrossRef]
  34. Boevé, J.-L.; Blank, S.M.; Meijer, G.; Nyman, T. Invertebrate and Avian Predators as Drivers of Chemical Defensive Strategies in Tenthredinid Sawflies. BMC Evol. Biol. 2013, 13, 198. [Google Scholar] [CrossRef]
  35. Leach, W.E. The Zoological Miscellany: Being Descriptions of New, or Interesting Animals; R. and A. Taylor, Shoe-Lane: London, UK, 1817; Volume 3, p. 155. [Google Scholar]
  36. Cameron, P. Hymenopterological Notices. Mem. Proc. Manch. Lit. Philos. Soc. 1891, 4, 182–194. [Google Scholar]
  37. Taeger, A.; Blank, S.M.; Liston, A. European Sawflies (Hymenoptera: Symphyta)—A Species Checklist for the Countries. In Recent Sawfly Research—Synthesis and Prospects; Blank, S.M., Schmidt, S., Taeger, A., Eds.; Goecke & Evers: Keltern, Germany, 2006; pp. 399–504. ISBN 978-3-937783-19-2. [Google Scholar]
  38. Benson, R.B. Notes on the British Sawflies of the Genus Athalia (Hymenoptera, Tenthredinidae), with the Description of a New Species. Entomol. Mon. Mag. 1931, 67, 109–114. [Google Scholar]
  39. Benson, R.B. A Revision of the Athaliini (Hymenoptera: Symphyta). Bull. Br. Mus. (Nat. Hist.). Entomol. Ser. 1962, 11, 333–382. [Google Scholar]
  40. Forsius, R. Über Eine Neue Selandriinen-Gattung Mit Vielgliedrigen Fühlern. Not. Entomol. 1935, 15, 57–59. [Google Scholar]
  41. Togashi, I. Symphyta of Thailand (Insecta, Hymenoptera). Steenstrupia 1988, 14, 101–119. [Google Scholar]
  42. Saini, M.S.; Blank, S.M.; Smith, D.R. Checklist of the Sawflies (Hymenoptera: Symphyta) of India. In Recent Sawfly Research—Synthesis and Prospects; Blank, S.M., Schmidt, S., Taeger, A., Eds.; Goecke & Evers: Keltern, Germany, 2006; pp. 575–617. ISBN 978-3-937783-19-2. [Google Scholar]
  43. Hartig, T. Die Aderflügler Deutschlands Mit Besonderer Berücksichtigung Ihres Larvenzustandes Und Ihres Wirkens in Wäldern Und Gärten Für Entomologen, Wald-Und Gartenbesitzer; Haude und Spener: Berlin, Germany, 1837; p. 416. [Google Scholar]
  44. Kirby, W.F. List of Hymenoptera with Descriptions and Figures of the Typical Specimens in the British Museum. 1. Tenthredinidae and Siricidae Vol. 1; British Museum (Natural History): London, UK, 1882; p. 450. [Google Scholar]
  45. Cameron, P. A Monograph of the British Phytophagous Hymenoptera. (Tenthredo, Sirex and Cynips, Linné.); Ray Society: London, UK, 1882; p. 340. [Google Scholar]
  46. Ashmead, W.H. Classification of the Horntails and Sawflies, or the Sub-Order Phytophaga. Can. Entomol. 1898, 30, 249–257. [Google Scholar] [CrossRef]
  47. Konow, F.W. Hymenoptera. Fam. Tenthredinidae. Genera Insectorum 1905, 29, 1–176. [Google Scholar]
  48. Rohwer, S.A. Technical Papers on Miscellaneous Forest Insects. II. The Genotypes of Sawflies and Woodwasps, or the Superfamily Tenthredinoidea. Tech. Ser. U.S. Dep. Agric. Bur. Entomol. 1911, 20, 69–109. [Google Scholar]
  49. Benson, R.B. On the Classification of Sawflies (Hymenoptera Symphyta). Trans. R. Entomol. Soc. Lond. 1938, 87, 353–384. [Google Scholar] [CrossRef]
  50. Takeuchi, K. A Generic Classification of the Japanese Tenthredinidae (Hymenoptera: Symphyta); Shinbi Print. Co.: Kyoto, Japan, 1952; Volume 91. [Google Scholar]
  51. Benson, R.B. Hymenoptera 2. Handb. Identif. Br. Insects 1952, 6, 51–137. [Google Scholar]
  52. Lorenz, H.; Kraus, M. Die Larvalsystematik Der Blattwespen. (Tenthredinoidea Und Megalodontoidea); Akademie Verlag: Berlin, Germany, 1957; pp. 1–389. [Google Scholar]
  53. Zhelochovtsev, A.N. Otrjad Hymenoptera—Pereponchatokrylye Podotrjad Symphyta (Chalastogastra)—Sidjachebrjuhie. [Order Hymenoptera—Wasps Suborder Symphyta (Chalastogastra)—Sawflies and woodwasps.]; Opredelitel’ nasekomyh evropejskoj chasti SSSR. T. III; Nauka: Leningrad, Russia, 1988; Volume 27, ISBN 978-5-02-025709-2. [Google Scholar]
  54. Abe, M.; Smith, D.R. The Genus-Group Names of Symphyta (Hymenoptera) and Their Type Species. Esakia 1991, 31, 115. [Google Scholar] [CrossRef]
  55. Wei, M.; Nie, H. Generic List of Tenthredinoidea s. Str. in New Systematic Arrangement with Synonyms and Distribution Data. J. Cent.-South For. Coll. 1998, 18, 23–31. [Google Scholar]
  56. Lacourt, J. Répertoire Des Tenthredinidae Ouest-Paléarctiques (Hymenoptera, Symphyta). Mémoires De La Société Entomol. De Fr. 1999, 3, 1–432. [Google Scholar]
  57. Heraty, J.; Ronquist, F.; Carpenter, J.M.; Hawks, D.; Schulmeister, S.; Dowling, A.P.; Murray, D.; Munro, J.; Wheeler, W.C.; Schiff, N.; et al. Evolution of the Hymenopteran Megaradiation. Mol. Phylogenetics Evol. 2011, 60, 73–88. [Google Scholar] [CrossRef] [PubMed]
  58. Isaka, Y.; Sato, T. Was Species Diversification in Tenthredinoidea (Hymenoptera: Symphyta) Related to the Origin and Diversification of Angiosperms? Can. Entomol. 2015, 147, 443–458. [Google Scholar] [CrossRef]
  59. Vilhelmsen, L.; Mikó, I.; Krogmann, L. Beyond the Wasp-Waist: Structural Diversity and Phylogenetic Significance of the Mesosoma in Apocritan Wasps (Insecta: Hymenoptera): Phylogeny of Hymenoptera. Zool. J. Linn. Soc. 2010, 159, 22–194. [Google Scholar] [CrossRef]
  60. Vilhelmsen, L. Morphological Phylogenetics of the Tenthredinidae (Insecta: Hymenoptera). Invert. Syst. 2015, 29, 164–190. [Google Scholar] [CrossRef]
  61. Taeger, A.; Blank, S.M.; Liston, A.D. World Catalog of Symphyta (Hymenoptera). Zootaxa 2010, 2580, 1. [Google Scholar] [CrossRef]
  62. Castro, L.R.; Dowton, M. The Position of the Hymenoptera within the Holometabola as Inferred from the Mitochondrial Genome of Perga Condei (Hymenoptera: Symphyta: Pergidae). Mol. Phylogenetics Evol. 2005, 34, 469–479. [Google Scholar] [CrossRef]
  63. Du, S.; Niu, G.; Nyman, T.; Wei, M. Characterization of the Mitochondrial Genome of Arge Bella Wei & Du Sp. Nov. (Hymenoptera: Argidae). PeerJ 2018, 6, e6131. [Google Scholar] [CrossRef]
  64. Tang, P.; Zhu, J.; Zheng, B.; Wei, S.; Sharkey, M.; Chen, X.; Vogler, A.P. Mitochondrial Phylogenomics of the Hymenoptera. Mol. Phylogenetics Evol. 2019, 131, 8–18. [Google Scholar] [CrossRef]
  65. He, H.; Niu, G.; Zhang, B.; Wei, M. The Complete Mitochondrial Genome of Athalia Proxima (Hymenoptera: Tenthredinidae) and Phylogenetic Analysis. Mitochondrial DNA Part B 2019, 4, 3868–3869. [Google Scholar] [CrossRef]
  66. Yan, Y.; Li, K.; Liu, S.; Niu, G.; Wei, M. The Complete Mitochondrial Genome of Cimbex Luteus (Hymenoptera: Cimbicidae) and Phylogenetic Analysis. Mitochondrial DNA Part B 2021, 6, 2031–2032. [Google Scholar] [CrossRef]
  67. Yan, Y.; Niu, G.; Zhang, Y.; Ren, Q.; Du, S.; Lan, B.; Wei, M. Complete Mitochondrial Genome Sequence of Labriocimbex Sinicus, a New Genus and New Species of Cimbicidae (Hymenoptera) from China. PeerJ 2019, 7, e7853. [Google Scholar] [CrossRef] [PubMed]
  68. Cheng, Y.; Yan, Y.; Wei, M.; Niu, G. Characterization of Mitochondrial Genomes of Three New Species: Leptocimbex Praiaformis, L. Clavicornis, and L. Yanniae (Hymenoptera: Cimbicidae). Entomol. Res. 2021, 51, 287–304. [Google Scholar] [CrossRef]
  69. Cheng, Y.; Yan, Y.; Wei, M.; Niu, G. The Complete Mitochondrial Genome of Praia Tianmunica (Hymenoptera: Cimbicidae) with Related Phylogenetic Analysis. Mitochondrial DNA Part B 2020, 5, 3037–3038. [Google Scholar] [CrossRef] [PubMed]
  70. Song, S.-N.; Tang, P.; Wei, S.-J.; Chen, X.-X. Comparative and Phylogenetic Analysis of the Mitochondrial Genomes in Basal Hymenopterans. Sci. Rep. 2016, 6, 20972. [Google Scholar] [CrossRef] [PubMed]
  71. Chen, Y.; Wei, M.; Yang, H.; Wang, H.; Niu, G. Nearly Complete Mitochondrial Genome of Trichiosoma Vitellina Linné, 1760 (Hymenoptera: Tenthredinidae): Sequencing and Phylogenetic Analysis. Mitochondrial DNA Part B 2020, 5, 802–803. [Google Scholar] [CrossRef]
  72. Doğan, Ö.; Korkmaz, E.M. Nearly Complete Mitogenome of Hairy Sawfly, Corynis Lateralis (Brullé, 1832) (Hymenoptera: Cimbicidae): Rearrangements in the IQM and ARNS1EF Gene Clusters. Genetica 2017, 145, 341–350. [Google Scholar] [CrossRef]
  73. Aydemir, M.N.; Korkmaz, E.M. Comparative Mitogenomics of Hymenoptera Reveals Evolutionary Differences in Structure and Composition. Int. J. Biol. Macromol. 2020, 144, 460–472. [Google Scholar] [CrossRef]
  74. Wei, S.-J.; Niu, F.-F.; Du, B.-Z. Rearrangement of TrnQ-TrnM in the Mitochondrial Genome of Allantus Luctifer (Smith) (Hymenoptera: Tenthredinidae). Mitochondrial DNA 2016, 27, 856–858. [Google Scholar] [CrossRef]
  75. Yang, J.; Sun, Z.; Wei, M.; Niu, G. The Complete Mitochondrial Genome of Allantus Togatus (Panzer, 1801), in View of Possible Cryptic Species. Mitochondrial DNA Part B 2021, 6, 1114–1115. [Google Scholar] [CrossRef]
  76. Liu, J.; Wei, M.; Niu, G. The Complete Mitochondrial Genome of Hemathlophorus Brevigenatus Wei, 2005 (Hymenoptera: Tenthredinidae) with Phylogenetic Analysis. Mitochondrial DNA Part B 2021, 6, 2759–2761. [Google Scholar] [CrossRef] [PubMed]
  77. Sun, Z.; Wei, M.; Niu, G. The Complete Mitochondrial Genome of Taxoblenus Sinicus Wei & Nie, 1999 (Hymenoptera: Tenthredinidae) and Phylogenetic Analysis. Mitochondrial DNA Part B 2021, 6, 2061–2062. [Google Scholar] [CrossRef] [PubMed]
  78. Xu, M.; Tan, B.; Wei, M.; Niu, G. The Complete Mitochondrial Genome Sequence of Taxonus Zhangi Wei, 1997 (Hymenoptera: Tenthredinidae) with Phylogenetic Analysis. Mitochondrial DNA Part B 2021, 6, 3188–3189. [Google Scholar] [CrossRef] [PubMed]
  79. Li, Y.; Wei, M.; Liu, J.; Niu, G. Characterization of the Mitochondrial Genome of Eutomostethus Vegetus Konow, 1898 (Hymenoptera: Tenthredinidae) and Phylogenetic Analysis. Mitochondrial DNA Part B 2020, 5, 3033–3034. [Google Scholar] [CrossRef] [PubMed]
  80. Wu, R.; Wei, M.; Liu, M.; Niu, G. Advancement in Sequencing the Mitochondrial Genome of Birmella Discoidalisa Wei, 1994 (Hymenoptera: Tenthredinidae) and the Phylogenetic Classification of Fenusini. Mitochondrial DNA Part B 2019, 4, 4100–4101. [Google Scholar] [CrossRef] [PubMed]
  81. Wan, S.; Wei, M.; Niu, G. The Complete Mitochondrial Genome Sequence of Sinopoppia Nigroflagella Wei, 1997 (Hymenoptera: Tenthredinidae) Reveals a New Gene Order. Mitochondrial DNA Part B 2021, 6, 999–1000. [Google Scholar] [CrossRef]
  82. Wu, D.; Wang, H.; Wei, M.; Niu, G. The Nearly Complete Mitochondrial Genome of Colochela Zhongi Wei, 2016 (Hymenoptera: Tenthredinidae) and Phylogenetic Analysis. Mitochondrial DNA Part B 2020, 5, 3341–3342. [Google Scholar] [CrossRef]
  83. Luo, X.; Wei, M.; Niu, G. Nearly Complete Mitochondrial Genome of Siobla Xizangensis Xiao, Huang & Zhou, 1988 (Hymenoptera: Tenthredinidae) and Phylogenetic Analysis. Mitochondrial DNA Part B 2019, 4, 4102–4103. [Google Scholar] [CrossRef] [PubMed]
  84. Liu, M.; Li, M.; Wei, M.; Li, Z. The Complete Mitochondrial Genome of Eriocampa Ovata Linne, 1760 (Hymenoptera: Tenthredinidae) and Phylogenetic Analysis. Mitochondrial DNA Part B 2015, 29, 164–190. [Google Scholar]
  85. Song, S.-N.; Wang, Z.-H.; Li, Y.; Wei, S.-J.; Chen, X.-X. The Mitochondrial Genome of Tenthredo Tienmushana (Takeuchi) and a Related Phylogenetic Analysis of the Sawflies (Insecta: Hymenoptera). Mitochondrial DNA Part A 2016, 27, 2860–2861. [Google Scholar] [CrossRef]
  86. Liu, Y.; Wei, M.; Niu, G. The First Mitochondrial Genome of a Fern Sawfly, Strongylogaster Xanthocera Stephens, 1835 (Hymenoptera: Tenthredinidae). Mitochondrial DNA Part B 2021, 6, 902–904. [Google Scholar] [CrossRef] [PubMed]
  87. Liu, H.; Chen, Q.; Liu, Y.; Pu, D.; Chen, Z.; Liu, X. The Complete Mitochondrial Genome of a Sawfly Species, Analcellicampa Danfengensis (Hymenoptera: Tenthredinidae). Mitochondrial DNA Part B 2019, 4, 3124–3125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Wei, S.-J.; Wu, Q.-L.; Liu, W. Sequencing and Characterization of the Monocellicampa Pruni (Hymenoptera: Tenthredinidae) Mitochondrial Genome. Mitochondrial DNA 2015, 26, 157–158. [Google Scholar] [CrossRef]
  89. Tan, B.; Wei, M.; Niu, G. The Complete Mitochondrial Genome of Moricella Rufonota Rohwer, 1916 (Hymenoptera: Tenthredinidae) and Phylogenetic Analysis. Mitochondrial DNA Part B 2021, 6, 2285–2287. [Google Scholar] [CrossRef]
  90. Zheng, B.-Y.; Li, Z.-K.; Li, X.-F.; Zhu, J.-C.; Sharkey, M.; Tang, P.; Chen, X.-X. The First Mitochondrial Genome of the Living-Fossil Sawfly Macroxyela Ferruginea (Hymenoptera: Xyelidae, Macroxyelinae). Mitochondrial DNA Part B 2020, 5, 23–24. [Google Scholar] [CrossRef]
  91. Zhang, F.; Ding, Y.; Zhu, C.; Zhou, X.; Orr, M.C.; Scheu, S.; Luan, Y. Phylogenomics from Low-coverage Whole-genome Sequencing. Methods Ecol. Evol. 2019, 10, 507–517. [Google Scholar] [CrossRef]
  92. Bushnell, B. BBtools. Available online: Https://Sourceforge.Net/Projects/Bbmap/ (accessed on 1 October 2021).
  93. Chikhi, R.; Rizk, G. Space-Efficient and Exact de Bruijn Graph Representation Based on a Bloom Filter. Algorithms Mol. Biol. 2013, 8, 22. [Google Scholar] [CrossRef]
  94. Pryszcz, L.P.; Gabaldón, T. Redundans: An Assembly Pipeline for Highly Heterozygous Genomes. Nucleic Acids Res. 2016, 44, e113. [Google Scholar] [CrossRef]
  95. Sahlin, K.; Chikhi, R.; Arvestad, L. Assembly Scaffolding with PE-Contaminated Mate-Pair Libraries. Bioinformatics 2016, 32, 1925–1932. [Google Scholar] [CrossRef]
  96. Luo, R.; Liu, B.; Xie, Y.; Li, Z.; Huang, W.; Yuan, J.; He, G.; Chen, Y.; Pan, Q.; Liu, Y.; et al. SOAPdenovo2: An Empirically Improved Memory-Efficient Short-Read de Novo Assembler. GigaScience 2012, 1, 18. [Google Scholar] [CrossRef]
  97. Manni, M.; Berkeley, M.R.; Seppey, M.; Simão, F.A.; Zdobnov, E.M. BUSCO Update: Novel and Streamlined Workflows along with Broader and Deeper Phylogenetic Coverage for Scoring of Eukaryotic, Prokaryotic, and Viral Genomes. Mol. Biol. Evol. 2021, 38, 4647–4654. [Google Scholar] [CrossRef] [PubMed]
  98. 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. [Google Scholar] [CrossRef] [PubMed]
  99. Donath, A.; Jühling, F.; Al-Arab, M.; Bernhart, S.H.; Reinhardt, F.; Stadler, P.F.; Middendorf, M.; Bernt, M. Improved Annotation of Protein-Coding Genes Boundaries in Metazoan Mitochondrial Genomes. Nucleic Acids Res. 2019, 47, 10543–10552. [Google Scholar] [CrossRef]
  100. Niu, G.; Zhang, Y.; Li, Z.; Wei, M. Characterization of the Mitochondrial Genome of Analcellicampa Xanthosoma Gen. et Sp. Nov. (Hymenoptera: Tenthredinidae). PeerJ 2019, 7, e6866. [Google Scholar] [CrossRef]
  101. Katoh, K.; Standley, D.M. MAFFT Multiple Sequence Alignment Software Version 7: Improvements in Performance and Usability. Mol. Biol. Evol. 2013, 30, 772–780. [Google Scholar] [CrossRef] [PubMed]
  102. Castresana, J. Selection of Conserved Blocks from Multiple Alignments for Their Use in Phylogenetic Analysis. Mol. Biol. Evol. 2000, 17, 540–552. [Google Scholar] [CrossRef]
  103. Lanfear, R.; Frandsen, P.B.; Wright, A.M.; Senfeld, T.; Calcott, B. PartitionFinder 2: New Methods for Selecting Partitioned Models of Evolution for Molecular and Morphological Phylogenetic Analyses. Mol. Biol. Evol. 2017, 34, 772–773. [Google Scholar] [CrossRef]
  104. Zhang, D.; Gao, F.; Jakovlić, 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. 2020, 20, 348–355. [Google Scholar] [CrossRef]
  105. Lartillot, N.; Rodrigue, N.; Stubbs, D.; Richer, J. PhyloBayes MPI: Phylogenetic Reconstruction with Infinite Mixtures of Profiles in a Parallel Environment. Syst. Biol. 2013, 62, 611–615. [Google Scholar] [CrossRef]
  106. Minh, B.Q.; Schmidt, H.A.; Chernomor, O.; Schrempf, D.; Woodhams, M.D.; Von Haeseler, A.; Lanfear, R. IQ-TREE 2: New Models and Efficient Methods for Phylogenetic Inference in the Genomic Era. Mol. Biol. Evol. 2020, 37, 1530–1534. [Google Scholar] [CrossRef]
  107. Aberer, A.J.; Kobert, K.; Stamatakis, A. ExaBayes: Massively Parallel Bayesian Tree Inference for the Whole-Genome Era. Mol. Biol. Evol. 2014, 31, 2553–2556. [Google Scholar] [CrossRef] [PubMed]
  108. Rambaut, A.; Drummond, A.J.; Xie, D.; Baele, G.; Suchard, M.A. Posterior Summarization in Bayesian Phylogenetics Using Tracer 1.7. Syst. Biol. 2018, 67, 901–904. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Rambaut, A. FigTree v1.4.2, A Graphical Viewer of Phylogenetic Trees. Available online: Http://Tree.Bio.Ed.Ac.Uk/Softw./Figtree/ (accessed on 1 January 2014).
  110. Yang, Z. PAML 4: Phylogenetic Analysis by Maximum Likelihood. Mol. Biol. Evol. 2007, 24, 1586–1591. [Google Scholar] [CrossRef] [PubMed]
  111. Malagón-Aldana, L.A.; Jensen, A.R.; Smith, D.R.; Shinohara, A.; Vilhelmsen, L. Sawflies out of Gondwana: Phylogenetics and Biogeography of Argidae (Hymenoptera). Syst. Entomol. 2022, 47, 231–258. [Google Scholar] [CrossRef]
  112. Nel, A.; Roques, P.; Nel, P.; Prokin, A.A.; Bourgoin, T.; Prokop, J.; Szwedo, J.; Azar, D.; Desutter-Grandcolas, L.; Wappler, T.; et al. The Earliest Known Holometabolous Insects. Nature 2013, 503, 257–261. [Google Scholar] [CrossRef]
  113. Jones, D.T.; Taylor, W.R.; Thornton, J.M. The Rapid Generation of Mutation Data Matrices from Protein Sequences. Bioinformatics 1992, 8, 275–282. [Google Scholar] [CrossRef]
  114. Sterelny, K. Another View of Life. Stud. Hist. Philos. Sci. Part C Stud. Hist. Philos. Biol. Biomed. Sci. 2005, 36, 585–593. [Google Scholar] [CrossRef]
  115. Rasnitsyn, A.P. Proishozhdenie i Jevoljucija Nizshih Pereponchatokrylyh. [Origin and Evolution of the Lower Hymenoptera.]. Tr. Paleontol. Inst. Acad. Nauk SSSR 1969, 123, 1–196. [Google Scholar]
  116. Rasnitsyn, A.P. Proishozhdenie i Jevoljucija Pereponchatokrylyh Nasekomyh. [Origin and Evolution of Hymenoptera.]. Tr. Paleontol. Inst. Acad. Nauk SSSR 1980, 174, 1–192. [Google Scholar]
  117. Piton, L. Paléontologie Du Gisement Éocéne de Menat (Puy-de-Dôme) (Flore et Faune); Imprimeries P. Vallier: Clermont-Ferrand, France, 1940; p. 303. [Google Scholar]
  118. Stocsits, R.R.; Letsch, H.; Hertel, J.; Misof, B.; Stadler, P.F. Accurate and Efficient Reconstruction of Deep Phylogenies from Structured RNAs. Nucleic Acids Res. 2009, 37, 6184–6193. [Google Scholar] [CrossRef]
  119. Cockerell, T.D.A. Fossil Hymenoptera from Florissant, Colorado. Bull. Mus. Comp. Zoölogy Harv. Coll. 1906, 50, 3–58. [Google Scholar]
  120. Li, H.-T.; Yi, T.-S.; Gao, L.-M.; Ma, P.-F.; Zhang, T.; Yang, J.-B.; Gitzendanner, M.A.; Fritsch, P.W.; Cai, J.; Luo, Y.; et al. Origin of Angiosperms and the Puzzle of the Jurassic Gap. Nat. Plants 2019, 5, 461–470. [Google Scholar] [CrossRef] [PubMed]
  121. Cardinal-McTeague, W.M.; Sytsma, K.J.; Hall, J.C. Biogeography and Diversification of Brassicales: A 103 Million Year Tale. Mol. Phylogenetics Evol. 2016, 99, 204–224. [Google Scholar] [CrossRef]
  122. Yu, X.-Q.; Maki, M.; Drew, B.T.; Paton, A.J.; Li, H.-W.; Zhao, J.-L.; Conran, J.G.; Li, J. Phylogeny and Historical Biogeography of Isodon (Lamiaceae): Rapid Radiation in South-West China and Miocene Overland Dispersal into Africa. Mol. Phylogenetics Evol. 2014, 77, 183–194. [Google Scholar] [CrossRef] [PubMed]
  123. Opitz, S.E.W.; Boevé, J.-L.; Nagy, Z.T.; Sonet, G.; Koch, F.; Müller, C. Host Shifts from Lamiales to Brassicaceae in the Sawfly Genus Athalia. PLoS ONE 2012, 7, e33649. [Google Scholar] [CrossRef] [Green Version]
  124. Jablonski, D.; Bottjer, D.J. The Origin and Diversification of Major Groups: Environmental Patterns and Macroevolutionary Lags. In Major Evolutionary Radiation; Taylor, P.D., Larwood, G.P., Eds.; Clarendon Press: Oxford, UK, 1991; pp. 17–57. [Google Scholar]
  125. Stroud, J.T.; Losos, J.B. Ecological Opportunity and Adaptive Radiation. Annu. Rev. Ecol. Evol. Syst. 2016, 47, 507–532. [Google Scholar] [CrossRef]
Figure 1. Schematic phylogeny for basal hymenopteran lineages simplified from the reference provided below each cladogram. (a) Wei, 1994 [28], Figure 5-1, morphological data; (b) Vilhelmsen, 2001 [29], Figure 11, morphological data; (c) Schulmeister, 2003 [30], Figure 7, morphological data; (d) Schulmeister, 2003 [31], Figure 3, morphological and molecular data; (e) Ronquist et al., 2012 [32], Figure 3, morphological and molecular data; (f) Malm and Nyman, 2015 [33], Figure 4, molecular data; (g) Boevé et al. 2013 [34], Figure 3, molecular data.
Figure 1. Schematic phylogeny for basal hymenopteran lineages simplified from the reference provided below each cladogram. (a) Wei, 1994 [28], Figure 5-1, morphological data; (b) Vilhelmsen, 2001 [29], Figure 11, morphological data; (c) Schulmeister, 2003 [30], Figure 7, morphological data; (d) Schulmeister, 2003 [31], Figure 3, morphological and molecular data; (e) Ronquist et al., 2012 [32], Figure 3, morphological and molecular data; (f) Malm and Nyman, 2015 [33], Figure 4, molecular data; (g) Boevé et al. 2013 [34], Figure 3, molecular data.
Insects 13 00858 g001
Figure 2. Topological comparison of PhyloBayes trees based on the concatenated amino acid matrix under (a) the CAT-GTR model and (b) the MtArt model, showing species group sampling and points of agreement and conflict. Posterior probability support values less than 0.90 are displayed on the trees with blue dots.
Figure 2. Topological comparison of PhyloBayes trees based on the concatenated amino acid matrix under (a) the CAT-GTR model and (b) the MtArt model, showing species group sampling and points of agreement and conflict. Posterior probability support values less than 0.90 are displayed on the trees with blue dots.
Insects 13 00858 g002
Figure 3. Topological comparison of PhyloBayes trees based on the concatenated nucleotide sequences of (a) the 13 PCGs and (b) the 13 PCGs and 2 rRNAs, showing species group sampling and points of agreement and conflict. Posterior probability support values less than 0.90 are displayed on the trees with blue dots.
Figure 3. Topological comparison of PhyloBayes trees based on the concatenated nucleotide sequences of (a) the 13 PCGs and (b) the 13 PCGs and 2 rRNAs, showing species group sampling and points of agreement and conflict. Posterior probability support values less than 0.90 are displayed on the trees with blue dots.
Insects 13 00858 g003
Figure 4. Topological comparison of trees based on single-copy orthologs (SCOs). (a) Iqtree; (b) PhyloBayes. Bootstrap values and posterior probability support values less than 0.90 are displayed on the trees with blue dots.
Figure 4. Topological comparison of trees based on single-copy orthologs (SCOs). (a) Iqtree; (b) PhyloBayes. Bootstrap values and posterior probability support values less than 0.90 are displayed on the trees with blue dots.
Insects 13 00858 g004
Figure 5. High-level taxon tree, gene rearrangement, and autapomorphic secondary structures of rRNA. (a) The dendrogram was simplified from the PCG dataset under the CAT-GTR model; numbers in brackets represent the mitogenome used in phylogenetic inference; (b) outline of rearrangement events across the Tenthredinoidea, with the blue lines indicating a major translocation and the red lines indicating a minor translocation; (c) group-specific structural characters of mt rrnL.
Figure 5. High-level taxon tree, gene rearrangement, and autapomorphic secondary structures of rRNA. (a) The dendrogram was simplified from the PCG dataset under the CAT-GTR model; numbers in brackets represent the mitogenome used in phylogenetic inference; (b) outline of rearrangement events across the Tenthredinoidea, with the blue lines indicating a major translocation and the red lines indicating a minor translocation; (c) group-specific structural characters of mt rrnL.
Insects 13 00858 g005
Figure 6. Age estimation trees using the dataset of (a) mtDNA; and (b) SCOs, highlighting the early diversity of Athaliidae constrained to the pre-Cretaceous.
Figure 6. Age estimation trees using the dataset of (a) mtDNA; and (b) SCOs, highlighting the early diversity of Athaliidae constrained to the pre-Cretaceous.
Insects 13 00858 g006
Table 1. Summary information of mitogenomes and the Single-copy Orthologs (SCOs) used in phylogenetic analyses.
Table 1. Summary information of mitogenomes and the Single-copy Orthologs (SCOs) used in phylogenetic analyses.
FamilySpeciesAccession NumberReferencesBUSCOMatrixVoucher Number
BlasticotomidaeRunaria striataON964462This study//CSCS-Hym-MC0014
Runaria punctataON808427This study53214877CSCS-Hym-MC0019
Blasticotoma minutaON964461This study55765115CSCS-Hym-MC0060
Blasticotoma tegularisON840089This study54845037CSCS-Hym-MC0083
Blasticotoma filicetiON840091This study53434901CSCS-Hym-MC0140
PergidaePerga condeiAY787816[62]///
Pergagrapta polita/SRX64296122922132/
ArgidaeArge bellaMF287761[63]//CSCS-Hym-MC0008
Arge similisMG923484[64]///
Arge auroraMN913350This study52984864CSCS-Hym-MC0179
Aproceros leucopoda/SRX64291336153378/
AthaliidaeAthalia icarMN527306[65]21841934CSCS-Hym-MC0021
Athalia birmanicaON840085This study57895312CSCS-Hym-MC0077
Athalia japonicaON964466This study57135251CSCS-Hym-MC0081
Athalia sikkimensisON840087This study56725219CSCS-Hym-MC0345
Athalia scapulataON840088This study56625193CSCS-Hym-MC0080
Athalia arunachalensisON840086This study57945320CSCS-Hym-MC0355
Athalia rosae/ERX897689644804165/
Athalia ‘tibetana’ON964467This study//CSCS-Hym-MC0731
Athalia qingzangON964468This study//CSCS-Hym-MC0732
Athalia tanaoserrulaON964469This study//CSCS-Hym-MC0733
Athalia sikkimensisON964470This study//CSCS-Hym-MC0734
Hypsathalia przewalskyiON840090This study57015240CSCS-Hym-MC0187
Dentathalia scutellariaeON808426This study57805310CSCS-Hym-MC0360
CimbicidaeCimbex luteusMW136447[66]607502CSCS-Hym-MC0035
Labriocimbex sinicusMH136623[67]53384897CSCS-Hym-MC0009
Leptocimbex clavicornisMT478109[68]54234973CSCS-Hym-MC0162
Leptocimbex praiaformisMT478110[68]//CSCS-Hym-MC0167
Leptocimbex yanniaeMT478111[68]//CSCS-Hym-MC0133
Praia tianmunicaMT665975[69]40923710CSCS-Hym-MC0049
Asitrichiosoma anthracinumKT921411[70]///
Trichiosoma vitellinaMN853777[71]51864743CSCS-Hym-MC0165
Corynis lateralisKY063728[72]///
DiprionidaeNeodiprion sertiferMK994526[73]///
Neodiprion qinghaiicusON964471This study56575192CSCS-Hym-MC0198
Gilpinia baiyinaobaoaON840092This study//CSCS-Hym-MC0178
Nesodiprion biremisON964465This study//CSCS-Hym-MC0055
Nesodiprion japonicusON964464This study//CSCS-Hym-MC0010
Diprion pini/SRX64291439333651/
HeptamelinaePsedoheptamellus liiON964463This study56535202CSCS-Hym-MC0175
Heptamelus brevicercusMW632128[25]//CSCS-Hym-MC0018
TenthredinidaeAllantoides luctiferKJ713152[74]///
Allantus togatusMW464859[75]56765206CSCS-Hym-MC0142
Asiemphytus rufocephalusKR703582[70]///
Empria liiMW632124[25]47374323CSCS-Hym-MC0079
Empronus tibetanusMZ265343[25]51484703CSCS-Hym-MC0152
Hemathlophorus brevigenatusMW632125[76]47894373CSCS-Hym-MC0177
Taxoblenus sinicusMW632126[77]//CSCS-Hym-MC0193
Xenapatidea procinctaMW487928[25]//CSCS-Hym-MC0022
Taxonus zhangiMZ461490[78]//CSCS-Hym-MC0342
Beleses atrofemoratusMZ265347[25]15201311CSCS-Hym-MC0011
Eutomostethus vegetusMT663219[79]53984939CSCS-Hym-MC0184
Megatomostethus crassicornisMZ265345[25]56825222CSCS-Hym-MC0075
Phymatoceropsis melanogasterMZ265346[25]947792CSCS-Hym-MC0032
Monophadnus latus/SRX635247542373914/
Birmella discoidalisaMF197548[80]494399CSCS-Hym-MC0029
Metallus maiMW255941[25]56435168CSCS-Hym-MC0182
Sinopoppia nigroflagellaMW487927[81]54054945CSCS-Hym-MC0073
Colochela zhongiMT702984[82]57145246CSCS-Hym-MC0061
Macrophya dolichogasterMW544890[25]///
Siobla xizangensisMN562486[83]57535280CSCS-Hym-MC0150
Eriocampa ovata/[84]58115337CSCS-Hym-MC0143
Tenthredo tienmushanaKR703581[85]///
Tenthredo koehleri/SRX31489945634230/
Cladiucha magnoliaeMT295305[25]21041844CSCS-Hym-MC0015
Cladiucha punctataMT295306[25]57335266CSCS-Hym-MC0044
Megabeleses liriodendrovoraxMW255939[25]18081570CSCS-Hym-MC0030
Megabeleses magnoliaeMW255940[25]57935316CSCS-Hym-MC0042
Neostromboceros nipponicusMW632127[25]817679CSCS-Hym-MC0026
Strongylogaster xanthoceraMW324676[86]57275255CSCS-Hym-MC0070
Analcellicampa xanthosomaMH992752[25]12501066CSCS-Hym-MC0017
Analcellicampa danfengensisMN163004[87]///
Monocellicampa pruniJX566509[88]///
Moricella rufonotaMW487926[89]//CSCS-Hym-MC0068
Nematus ribesii/SRX64300135443329/
Hemibeleses tianmunicusMZ265344[25]50724631CSCS-Hym-MC0027
XyelidaeXyela sp.MG923517[64]///
Xyela alpigena/SRX64293035173308/
Megaxyela euchromaOL794667This study53934938CSCS-Hym-MC0001
Macroxyela ferrugineaMK270536[90]///
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Niu, G.; Budak, M.; Korkmaz, E.M.; Doğan, Ö.; Nel, A.; Wan, S.; Cai, C.; Jouault, C.; Li, M.; Wei, M. Phylogenomic Analyses of the Tenthredinoidea Support the Familial Rank of Athaliidae (Insecta, Tenthredinoidea). Insects 2022, 13, 858. https://doi.org/10.3390/insects13100858

AMA Style

Niu G, Budak M, Korkmaz EM, Doğan Ö, Nel A, Wan S, Cai C, Jouault C, Li M, Wei M. Phylogenomic Analyses of the Tenthredinoidea Support the Familial Rank of Athaliidae (Insecta, Tenthredinoidea). Insects. 2022; 13(10):858. https://doi.org/10.3390/insects13100858

Chicago/Turabian Style

Niu, Gengyun, Mahir Budak, Ertan Mahir Korkmaz, Özgül Doğan, André Nel, Siying Wan, Chenyang Cai, Corentin Jouault, Min Li, and Meicai Wei. 2022. "Phylogenomic Analyses of the Tenthredinoidea Support the Familial Rank of Athaliidae (Insecta, Tenthredinoidea)" Insects 13, no. 10: 858. https://doi.org/10.3390/insects13100858

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop