INTRODUCTION
Tenebrionoidea is a large superfamily in Coleoptera, with over 34,000 known species and 28 families (Lawrence, 1995; Ślipiński et al., 2011). The species in the superfamily have the 5-5-4 tarsal formula in both sexes, with occasional 4-4-4, 3-3-3 or 3-4-4 in males. Members of the superfamily demonstrate various types of feeding strategies, the majority of which are fungivorous, xylophagous, and saprophagous. Some Tenebrionoidea species are major agricultural and forest pests that attack commercial crops or stored products (Song et al., 2018). They are widely spread throughout all terrestrial habitats from the sea shore up to dry desert and steppe habitats in all altitudinal belts and all types of forests, and species in arid environments are conspicuously diverse.
Under the combined influence of various factors, including feeding strategies, habitats, cryptic speciation, developmental processes, sexual selection, convergent evolution, and other mechanisms, the morphology of the superfamily is complex (López-Estrada et al., 2019, 2022). The species of Tenebrionoidea are of significant ecological and economic significance. For instance, the larvae of Epicauta gorhami from the Meloidae family feed on grasshopper eggs, whereas the adults take soybeans as food (Shintani et al., 2011), functioning as pest predator as well as pest in the ecological system. In addition, they produce a highly toxic substance called cantharidin, which is valuable for treating certain types of tumors (Zha et al., 2017). The Tenebrio molitor in Tenebrionidae, as a natural decomposer, has strong and rapid reproduction ability and ecological adaptability and provides food for a wide variety of birds and mammals in the field as well as are industrially bred to produce animal and even human food (Grau et al., 2017).
Taxonomic and phylogenetic studies are essential for understanding their biodiversity and biological characteristics and for further protecting, utilizing, and controlling them. The framework of the classification system for Coleoptera was firstly established in 1955 and then revised in 1982, in which the Tenebrionoidea is divided into five major lineages: (1) Tetratomidae, Melandryidae, Mordellidae, Ripiphoridae; (2) Synchroidae, Zopheridae, Prostomidae, Perimylopidae, Chalcodryidae, Tenebrionidae; (3) Oedemeridae, Stenotrachelidae, Meloidae; (4) Pythidae, Pyrochroidae, Boridae, Mycteridae, Salpingidae; and (5) Anthicidae, Aderidae, Scraptiidae (Crowson, 1955; Lawrence & Newton Jr, 1982). In past years, a number of studies have been reported to establish the taxonomy system of Tenebrionoidea using comparative morphology data (Beutel & Friedrich, 2005; Lawrence et al., 2011). Nowadays, 28 families (Mordellidae, Ripiphoridae, Anthicidae, Meloidae, Pyrochroidae, Oedemeridae, Salpingidae, Scraptiidae, Lagriidae, Tenebrionidae, and so on) are widely recognized for Tenebrionoidea (Bouchard et al., 2011). Despite its economic importance, the monophyletic status of many families and phylogenetic relationships in Tenebrionoidea need to be elucidated. A comprehensive phylogeny study of beetles showed that the two families Mordellidae and Meloidae are monophyletic, while the Scraptiidae is paraphyletic based on 18S rRNA, 16S rRNA, and COX1 gene sequences (Hunt et al., 2007). The phylogenetic analysis of Coleoptera showed that the families Tenebrionidae, Oedemeridae, Salpingidae, Mordellidae, Meloidae, and Scraptiidae are monophyletic; Pyrochroidae is polyphyletic based on four gene (16S rRNA, COX1, 28S, and 18S) sequences (Bocak et al., 2014). The higher-level phylogenetic analysis of beetles based on 95 nuclear protein-coding genes from 373 beetle species proposed that the Tenebrionidae, Oedemeridae, Salpingidae, Pyrochroidae, Meloidae, and Scraptiidae were monophyletic, whereas Anthicidae was paraphyletic (Zhang et al., 2018). The study based on 1,907,014 amino acid sites from 4818 nuclear genes across 146 beetle species proposed that Tenebrionidae, Ciidae, Ischaliidae, Salpingidae, Zopheridae, Stenotrachelidae, Scraptiidae, Prostomidae, Trictenotomidae, Mycetophagidae, Oedemeridae, Aderidae, Pyrochroidae, Chalcodryidae, Tetratomidae, Tentratomidae, Synchroidae, Archeocrypticidae, and Pterogeniidae were monophyletic. In the later study, based on the extraction of 68 single-copy nuclear protein-coding (NPC) genes from 129 families suggested that the Oedemeridae, Salpingidae, Pyrochroidae, Meloidae, Zopheridae, Ciidae, Ulodidae, Mycetophagidae, and Tenebrionidae were monophyletic (Cai et al., 2022). Almost all molecular phylogenetic analyses suggested the monophyly of Mordellidae and Meloidae; however, it is inconsistent with the results based on morphological data (Lawrence et al., 2011). Due to incongruent phylogenetic relationships from different analyses, the phylogenetic relationships and monophyly among families need to be elucidated.
With the characteristics of conservative gene content, simple genome organization, maternal inheritance, small genome size, and higher evolution rate, the complete mitochondrial genome (mtgenome) has been widely used in molecular phylogenetics, evolution, and population genetics study of insects (Cameron, 2014). Up to date (October 2021), there have been 71 species of Tenebrionoidea complete mtgenomes to be reported in GenBank, covering 15 families in Tenebrionoidea. However, there is no complete mtgenome sequence to be reported from the families Archeocrypticidae, Stenotrachelidae, and Melandryidae, and the characteristics of the mtgenomes are still little understood in the superfamily. Up to now, there have been only a few mtgenome-based phylogenetic studies at high taxa in Tenebrionoidea. Phylogenetic inference of beetles indicated that the families Tenebrionidae, Oedemeridae, Mordellidae, and Meloidae were monophyletic, whereas Scarptiidae was polyphyletic based on 245 mitochondrial sequences using Bayesian method (Timmermans et al., 2015). The phylogenetic analysis of 37 beetle mitochondrial genes supported the Mordellidae and Meloidae to be monophyletic groups and the Scraptiidae and Melandryidae to be paraphyletic groups using Bayesian inference (Song et al., 2018). The mtgenome-based phylogeny of 51 beetles found the monophyletic of Meloidae, Tenebrionidae, and the sister relationship of Meloidae and Tenebrionidae using Bayesian inference (Tang et al., 2020). As mentioned above, previous mtgenome-based phylogenetic studies have not well resolved the phylogeny of Tenebrionoidea, and phylogenetic analysis of the superfamily needs to be further evaluated.
In this study, we newly sequenced and annotated mitogenomes from 19 species belonging to Tenebrionoidea and comparatively analyzed the mtgenome characteristics of a total of 90 species in Tenebrionoidea. More importantly, we constructed and discussed the phylogenetic relationships at high taxa level based on these mtgenome sequences and inferred the divergence time at main nodes in the phylogenetics in the inference of fossil records. This study lays a foundation for further understanding of mtgenome characteristics, phylogenetics, and evolution of the Tenebrionoidea.
MATERIALS AND METHODS
Sample collection, sequencing, and mtgenome assembly
All samples were collected in China (Table S1) and then stored in 95% alcohol at −20°C until DNA extraction. These samples were identified in morphology, 19 species were selected for sequencing, and their identification was confirmed by subsequent COX1 comparison with BOLD (; Hebert et al., 2003). The total genomic DNA was extracted from the thorax and leg muscle tissue by DNeasy Blood and Tissue kit (Qiagen, Duesseldorf, Germany) according to the manufacturer's instructions. Concentration of extracted genomic DNA was determined by Qubit 2.0 (Invitrogen, Shanghai, China). The 350 bp small fragment libraries were constructed and then sequenced using the Illumina Hiseq 2500 (San Diego, CA) with 150 bp paired-end reads in Shenzhen Huitong Biotechnology Co. Ltd (Shenzhen, China). After removing the adapters and unpaired, short- and low-qualitied reads, clean reads from mtgenomes were extracted using a BLAST (Altschul et al., 1990) search against known Tenebrionoidea mtgenome sequences and then used for de novo mtgenome assembly with SPAdes v. 3.9.0 (Bankevich et al., 2012). The contigs of mtgenome were extracted and assembled into mtgenomes by searching against the reference sequences using PRICE (paired-read iterative contig extension) by NOVOPlasty version 2.6.2 (Dierckxsens et al., 2016).
Mtgenome annotation and characteristics analysis
In the present study, a total of 90 species of mtgenomes in 16 families belonging to Tenebrionoidea as ingroup and two species in Coccinellidae as outgroup were subject to mtgenome annotation, and characteristics and phylogeny analysis, in which 71 species in Tenebrionoidea and two species in Coccinellidae were selected and downloaded from GenBank in October 2021. The rough annotation of protein-coding genes (PCGs), transfer RNA genes (tRNAs), ribosomal RNA genes (rRNAs), and CR was initially identified using MITOS (; Bernt et al., 2013) and then determined in comparison of published homologous mtgenome sequences in phylogeny-close species using MEGAX (Kumar et al., 2018). The tRNAs secondary structures were predicted using tRNAscan-SE Search Server v. 1.21 (; Lowe & Eddy, 1997). The annotation of the mtgenomes was corrected manually using Geneious v. 4.8.5 (Kearse et al., 2012), and final mtgenomes were submitted to the GenBank database. The secondary structures of the tRNAs were visualized and manually edited using VARNA (; Darty et al., 2009). The mtgenomes were visualized using the Chloroplot online server with default parameters (Zheng et al., 2020). Base composition and relative synonymous codon usage (RSCU) of 90 species of mtgenome were computed with PhyloSuite desktop platform (Zhang et al., 2020). AT-skew [(A − T)/(A + T)] and GC-skew [(G − C)/(G + C)] were estimated to investigate nucleotide composition bias (Perna & Kocher, 1995), and three-dimensional scatterplots of AT-Skew, GC-Skew, and AT% were drawn using Origin Pro v. 9.0 (Mikrajuddin & Khairurrijal, 2009). The selection pressure of the 13 PCGs was analyzed by calculating Ka (nonsynonymous mutation rates) and Ks (synonymous mutation rates) values with DnaSP v. 5.0 (Librado & Rozas, 2009), and visualized using RStudio. Sequence saturation was assessed in DAMBE v. 5.0 (Xia, 2013).
Phylogenetic analysis of Tenebrionoidea
Phylogenetic relationships of 90 species of mtgenomes (including 19 sequenced in this study) in Tenebrionoidea were deduced using three data sets and two inference methods with Aiolocaria hexaspilota and Henosepilachna vigintioctopunctata (Coleoptera: Coccinelloidea) as outgroups. Taxonomic information for each species investigated and mtgenome accession numbers are listed in Table 1. Three data sets were concatenated using PhyloSuite platform: (1) amino acid sequence of 13 PCGs (AA), (2) nucleotide sequence at first and second codon position of 13 PCGs (PCG12), (3) PCG12 + 2 rRNAs, respectively, with excluding start codon, stop codon. Nucleotide sequences of 13 PCGs were aligned by codon-based multiple alignments using the L-INS-i algorithm, and the rRNAs were aligned using the Q-INS-i strategy in MAFFT v. 7.0 (Katoh & Standley, 2013), and ambiguously aligned positions were excluded using Gblocks (Talavera & Castresana, 2007). The concatenation of aligned sequences was performed using SequenceMatrix (Vaidya et al., 2011). The selection of best-fit partitioning schemes and substitution models for each data set were calculated using PartitionFinder v. 2.0 (Lanfear et al., 2016) with the settings: branch lengths as linked, model election as AICc with the greedy algorithm. Partitioning schemes and models are listed in Table S2. Two methods, maximum likelihood (ML) and Bayesian inference (BI), were employed for the deduce. ML-based phylogenetic analyses were conducted using IQ-TREE v. 1.6.8 in PhyloSuite v. 1.2.2 (Zhang et al., 2020). Nodal support values were inferred with 1000 bootstrapped replicates (BPs) (Minh et al., 2013). BI analysis was conducted using MrBayes v. 3.2.6 (Ronquist et al., 2012). A total of 2,000,000 generations with four chains were sampled every 1000 generations. Posterior probabilities (PPs) were computed after discarding the first 25% of trees as the burn-in phase. The estimated sample size (ESS) >200 and the average deviation of the split frequency of less than 0.01 indicates that the runs had converged. The phylogenetic tree was visualized using FigTree v. 1.4.4 and iTOL online tool (Letunic & Bork, 2016).
TABLE 1 Taxonomic information and GenBank accession numbers for 90 mtgenomes in Tenebrionoidea selected for characteristics and phylogenetics analysis in this study.
Family | Subfamily | Species | Accession number | References |
Tenebrionidae | Pimeliinae | Asbolus verrucosus | NC_027256 | Rider (2016) |
Stenomorpha obovata | MZ342786 | Smith et al. (2021) | ||
S. consobrina | MZ342785 | Smith et al. (2021) | ||
Diaperinae | Platydema sp. | JX412842 | Unpublished | |
Scaphidema metallicum | KX087341 | Unpublished | ||
Ulomoides dermestoides | NC_025332 | Wang et al. (2016) | ||
Stenochiinae | Strongylium Kulzeri | CT5 | This study | |
S. pinfaense | CT6 | This study | ||
S. suspicax | JX412780 | Unpublished | ||
S. nakanei | MW802591 | This study | ||
Promethis valgipes | NC_054362 | Direct Submission | ||
Morphostenophanes yunnanu | MZ298928 | Smith et al. (2021) | ||
M. sinicus | MW853764 | Direct Submission | ||
Tenebrioninae | Uloma sp. | KT876915 | Linard et al. (2016) | |
Nalassus laevioctostriatus | KT876905 | Linard et al. (2016) | ||
Tribolium castaneum | NC_003081 | Friedrich et al. (2003) | ||
T. audax | NC_024600 | Ou et al. (2016) | ||
T. confusum | NC_026702 | Ou et al. (2016) | ||
Tenebrio molitor | NC_024633 | Li-Na et al. (2014) | ||
T. obscurus | NC_037196 | Bai et al. (2018) | ||
Zophobas atratus | NC_041101 | Bai et al. (2018) | ||
Plesiophthalmus longipes | S20 | This study | ||
Paramarygmus sp. | JX412808 | Unpublished | ||
P. pallidicrus | CT8 | This study | ||
Heterotarsus carinula | CT9 | This study | ||
Gonocephalum sp. | NC_053250 | Direct Submission | ||
Gonocephalum kochi | MW822744 | Unpublished | ||
Blaps rhynchoptera | NC_047449 | Zhao et al. (2019) | ||
Alphitobius diaperinus | NC_049092 | Hong et al. (2020) | ||
Opatrum subaratum | CT11 | This study | ||
O. sabulosum | MN745102 | Feng et al. (2020) | ||
Alleculinae | Borboresthes tibialis | CT13 | This study | |
Cteniopus sp. | KX087267 | Unpublished | ||
Cteniopinus ruber | S21 | This study | ||
C. hypocrita | CT14 | This study | ||
Lagriidae | Adeliinae | Adelium sp. | NC_013554 | Sheffield et al. (2008) |
Lagriinae | Aulonogria discolora | CT41 | This study | |
Lagria ophthalmica | CT20 | This study | ||
L. nigricollis | CT25 | This study | ||
L. rufipennis | CT23 | This study | ||
Cerogria kikuchii | CT24 | This study | ||
C. popularis | S15 | This study | ||
C. janthinipennis | CT28 | This study | ||
C. pachycera | CT22 | This study | ||
Xenocerogria ruficollis | CT32 | This study | ||
Statirinae | Chlorophila portschinski | CT33 | This study | |
C. semenowi | CT34 | This study | ||
Anisostira rugipennis | CT35 | This study | ||
Exostira schroederi | MW802590 | This study | ||
Taiwanolagria merkli | CT37 | This study | ||
Casnonidea terminata | CT38 | This study | ||
Meloidae | Nemognathinae | Zonitoschema japonica | S18 | This study |
Meloinae | Epicauta gorhami | NC_036042 | Du et al. (2017) | |
E. chinensis | NC_029192 | Du et al. (2017) | ||
E. tibialis | NC_036043 | Du et al. (2017) | ||
E. aptera | NC_031820 | Jie et al. (2016) | ||
Hycleus marcipoli | NC_036044 | Du et al. (2017) | ||
H. phaleratus | NC_036045 | Du et al. (2017) | ||
H. dorsetiferus | CT18 | This study | ||
H. cichorii | NC_039657 | Unpublished | ||
Lytta caraganae | NC_033339 | Yuan et al. (2017) | ||
Mylabris aulica | NC_036046 | Du et al. (2017) | ||
M. calida | NC_050854 | Unpublished | ||
Oedemeridae | Oedemerinae | Nacerdes potanini | S44 | This study |
N. carniolica | KX087319 | Unpublished | ||
N. holzschuhi | CT1 | This study | ||
Ischnomera caerulea | JX412790 | Unpublished | ||
Ascleranoncodes sp. | CT4 | This study | ||
Scraptiidae | Scraptiinae | Scraptia sp. | JX412825 | Unpublished |
Cyrtanaspis phalerata | KX087279 | Unpublished | ||
Anaspidinae | Anaspis sp. | JX412856 | Unpublished | |
Mordellidae | Mordellinae | Mordella atrata | NC_013254 | Cameron et al. (2009) |
Tomoxia bucephala | KX087355 | Unpublished | ||
Mordellidae sp. | JX412844 | Unpublished | ||
Mordellochroa milleri | KX087318 | Unpublished | ||
Zopheridae | Zopherinae | Usechus lacerta voucher BMNH 679350 | NC_036269 | Linard et al. (2017) |
Ripiphoridae | Pelecotominae | Pelecotoma fennica | NC_036277 | Linard et al. (2017) |
Metoecus javanus | MZ618348 | Unpublished | ||
Ciidae | Ciinae | Ciidae sp. | JX412846 | Unpublished |
Tetratomidae | Tetratominae | Tetratoma fungorum | NC_036276 | Linard et al. (2017) |
Aderidae | Aderidae sp. | JX412763 | Unpublished | |
Anthicidae | Anthicinae | Stricticollis tobias | KX087350 | Unpublished |
Anthicidae sp.1 | MH789722 | Crampton-Platt et al. (2018) | ||
Anthicidae sp.2 | MG253255 | Unpublished | ||
Pyrochroidae | Pyrochroinae | Schizotus pectinicornis | KX087342 | Unpublished |
Pyrochroa sp.1 | S43 | This study | ||
Salpingidae | Salpinginae | Vincenzellus ruficollis | NC_036274 | Linard et al. (2017) |
Lissodema cursor | KX087307 | Unpublished | ||
Prostomidae | Prostomis sp. | JX412787 | Unpublished | |
Trictenotomidae | Trictenotoma davidi | MW580860 | Sheng et al. (2021) | |
Outgroup (Coccinellidae) | Aiolocaria hexaspilota | NC_042417 | Seo et al. (2019) | |
Henosepilachna vigintioctopunctata | NC_041172 | Zhang et al. (2019) |
Divergence time estimation
Divergence time was estimated using the uncorrelated relaxed clock model as implemented in BEAST v. 1.6.1 (Drummond & Rambaut, 2007). In order to limit the numbers of parameter for the estimation, Bayesian tree was used as a guide tree. The tree prior generated using a Yule speciation model, and all node calibrations were enforced using Normal distributions. Constraints on clade ages were enforced using five fossil calibrations: Protoripidius burmiticus in Mordellidae (m = 96.55, SD = 2.0, offset = 0) diverged 93.5–99.6 Mya (Cai et al., 2018), Camelomorpha longicervix in Anthicidae–Meloidae (m = 127.75, SD = 1.25, offset = 0) 125.45–130 Mya (Kirejtshuk & Azar, 2008), Alphitopsis initialis in Tenebrionidae–Lagriidae (m = 127.75, SD = 1.25, offset = 0) 125.45–130 Mya (Kirejtshuk et al., 2012), Meloe dominucanus in Lytta (logm = 10, SD = 1.25, offset = 15) 50–15 Mya (López-Estrada et al., 2022; Poinar et al, 2009), and Epicauta sanctoruensis in Epicauta (m = 4.47, SD = 0.5, offset = 0) 5.33–3.6 Mya (Caballero & de León, 2003). The posterior time estimation was conducted using an MCMC algorithm, and the MCMC run was sampled every 1000 iterations until it achieved 10,000 samples, after the first 100,000 iterations were discarded as burn-in. The effective sample sizes of every node age were confirmed using Tracer v. 1.5 until every parameter was >200. The maximum clade credibility tree was calculated using TreeAnnotator v. 1.6.1, with the node times scaled to match the mean posterior estimates.
RESULTS
Mtgenome organization
We completely sequenced mitogenomes from 19 species belonging to Tenebrionoidea in the present study (accession numbers in Table 1), with 11 of them being reported for the first time, all within Lagriidae. The complete mtgenomes from 82 species investigated contain the typical 37 genes (including 13 PCGs, 2 rRNAs, and 22 tRNAs) and one control region (CR). There are 22 genes (nine PCGs and 13 tRNAs) located on the majority coding strand (J-strand), whereas the other 15 genes (four PCGs, nine tRNAs, and two rRNAs) are on the minority strand (N-strand; Figure 1). These mtgenome sequences range in length from 14,777 bp (Cerogria kikuchii) to 16,861 bp (Heterotarsus carinula) with an average of 15,763 bp, and the length variation mainly results from the control region, intergenic overlap, and spacers. They all display obvious AT bias with A + T content ranging from 62.7% (Casnonidea terminata) to 81.6% (Pyrochroidae sp.) and an average of 73.6%. AT-skew values range from −0.141 (Paramarygmus sp.) to 0.219 (Strongylium pinfaense) and GC-skew from −0.375 (S. pinfaense) to 0.366 (Paramarygmus sp.; Figure 2).
[IMAGE OMITTED. SEE PDF]
[IMAGE OMITTED. SEE PDF]
The tRNAs sizes range from 57 to 82 bp, and all of the tRNAs can be folded into a typical clover-leaf structure except for tRNA-Ser (AGN) in which the dihydrouridine (DHU) arm is absent, and a UCU anticodon is present (Figure S1). The most frequently occurred base mismatches are U-G, U-U, and A-G, and the mismatch A-G is only occurred in tRNA-Trp. For rRNAs, rrnL is located between the trnL1 and trnV, ranging from 750 bp (Anthicidae sp.) to 1323 bp (Alcidodes juglans). The length of the rrnS ranges from 740 bp (Cerogira popularis) to 1269 bp (Uloma sp.), which is located between the trnV and the CR region. The percentage of AT content in rRNAs is 68.3%–84.3%. The CRs are located between the rrnS and trnI, and the percentage of AT content in this region is 72.2%–96.1%.
Rearrangement events
By comparing the composition and structure of the mtgenomes of these 90 Tenebrionoidea species, a total of seven Tenebrionoidea species were found to have gene rearrangement events. These rearrangement events occur in the three different families Lagriidae, Mordellidae, and Pyrochroidae (Figure 3). In other families, the gene order of the mtgenome is exactly the same as that of drosophilid. The tRNA genes have the highest frequency of rearrangement (trnW-trnC-trnY and trnA-trnR-trnN-trnS-trnE-trnF gene cluster), followed by protein-coding genes. The first one is found in the trnW-trnC-trnY gene cluster, and shuffling of trnW gene and trnC gene occurs in three species of Schizotus pectinicornis (Pyrochroidae), Pyrochroidae sp. (Pyrochroidae), and Anisostira rugipennis (Lagriidae). The second one in the A. rugipennis of Lagriidae, trnD and ATP8 transposition to the upstream of NAD2, which is the first rearrangement event of protein-coding genes found in the family Lagriidae. The last one is found in the trnA-trnR-trnN-trnS-trnE-trnF gene cluster, and shuffling of trnR gene and trnN gene occurs in four species of the family Mordellidae: Mordellidae sp., Mordellochroa milleri, Mordella atrata, and Tomoxia bucephala.
[IMAGE OMITTED. SEE PDF]
Codon usage of
Total PCGs nucleotide length ranges from 10,848 to 11,142 bp, and the AT contents range from 60.7% to 81.0%. Most of the PCGs initiate with the typical start codon ATN and TTG, whereas the special start codons AAC, AAT, AAA, and TCA are found for COX1; AAA for COX2; GTG for ND1 and ND4L; AGG and AAA for ND2 and GTG for ND4. The most frequently used stop codons are TAA and TAG, followed by the incomplete stop codons T and TA. The most frequently used codons are UUA (Leu2), UCU and UCA (Ser2), CGA (Arg), whereas AGC (Ser1), ACG (Thr), GCG (Ala), and CUG (Leu1) are the least used (Figure 4). For each PCG, the Ka/Ks ratio is less than one, and ATP8 has the highest Ka/Ks ratio (0.33–0.67), followed by seven genes (ND6, ND5, ND4, ND2, ND4L, ND1, and ND3) with Ka/Ks ratios of 0.17–0.42. Complex IV (COX1, COX2, and COX3), Complex III (CYTB), and ATP6 have low Ka/Ks ratios with a range from 0.01 to 0.17 (Figure 5). These results imply that all these 13 PCGs experienced purifying selection, especially Complex IV and Complex III.
[IMAGE OMITTED. SEE PDF]
[IMAGE OMITTED. SEE PDF]
Phylogenetic relationships
Substitution saturation tests show no saturation for three data sets AA, PCG12, and PCG12 + rRNAs (Iss < Iss.cSym or Iss.cAsym, p < .05; Table 2), which proposes that these three data sets be appropriate for phylogenetic construction based on ML and BI. Six trees generated using these three data sets and both ML and BI are slightly different in topology (Figures 6, 7 and S3–S6). The Ciidae is located at the base of all phylogenetic trees, followed by families Mordellidae + Ripiphoridae. The Mordellidae was recovered monophyletic (PP = 1; BP = 100), and Ripiphoridae is close to it. The family Aderidae was suggested to be close to the “Meloidae clade” + “Tenebrionidae clade” branch. Both “Tenebrionidae clade” and “Meloidae clade” recovered monophyletic with both of them being sister groups (Figure 6). In the “Meloidae clade,” the families Meloidae + Anthicidae were suggested to be monophyletic (PP = 1; BP = 100), and it appears sister to the “Oedemeridae clade,” and the Meloidae was recovered monophyletic (PP = 1; BP = 100). In the “Oedemeridae clade,” the Oedemeridae, Pyrochroidae, Salpingidae, and Scraptiidae seem monophyletic (PP = 1; BP = 100), and Salpingidae is the sister of Scraptiidae. In the “Tenebrionidae clade,” the family Lagriidae and Tenebrionidae are monophyletic (PP = 1; BP = 90–100), and both of them are sister groups of each other. In Lagriidae, the subfamily Adeliinae is based on the subfamilies Lagriinae and Statininae, both of which were recovered monophyletic (PP = 1; BP = 100) and are sister groups of each other. Due to the insufficient sampling of Ripiphoridae and Anthicidae, we will not discuss whether they are monophyletic here.
TABLE 2 Substitution saturation test results.
Data partition | Iss | Iss.cSyma | Psymb | Iss.cAsymc | Pasymd |
AA | 0.005 | 0.703 | 0.0000 | 0.401 | 0.0000 |
PCG12 + RNA | 0.364 | 0.818 | 0.0000 | 0.572 | 0.0000 |
PCG12 | 0.232 | 0.813 | 0.0000 | 0.570 | 0.0000 |
PCG1 | 0.309 | 0.807 | 0.0000 | 0.548 | 0.0000 |
PCG2 | 0.160 | 0.807 | 0.0000 | 0.548 | 0.0000 |
PCG3 | 0.710 | 0.807 | 0.0000 | 0.548 | 0.0000 |
[IMAGE OMITTED. SEE PDF]
[IMAGE OMITTED. SEE PDF]
In Tenebrionidae, the subfamily Pimeliinae appears to be a monophyletic group (PP = 1; BP = 100) and is located at the base of Tenebrionidae. The Alleculinae and Stenochiinae were recovered monophyletic (PP = 1; BP = 100), and the subfamilies Tenebrioninae and Diaperinae appeared in polyphyletic groups.
Six trees were generated using AA, PCG12, and PCG12 + 2 rRNAs data sets, and both ML and BI are slightly different in topology. For AA data set, the topologies using ML and BI are different in the positions of Prostomidae. Prostomidae and Tetratomidae are clustered as one clade in ML tree, but not in BI tree. For PCG12 data set, two same topologies of trees from BI and ML differ from two topologies of AA data set in phylogenetic relationship of the “Oedemeridae clade.” For PCG12 + 2 rRNAs data set, the positions of major families are the same as the four topologies of AA and PCG12 data sets, with only a few differences in the “Oedemeridae clade.”
Divergence time
The AA data set was used to estimate divergence time because AA had higher node support values than others in the initial phylogenetic assessment using Bayesian approach. Based on five fossil calibrations points of Protoripidius burmiticus (99.6–93.5 Mya), Camelomorpha longicervix (130.0–125.45 Mya), Alphitopsis initialis (130.0–125.45 Mya), Meloe dominucanus (20.44–13.82 Mya), and Epicauta sanctoruensis (5.33–3.6 Mya; filled red cycles in Figure 8), the superfamily Tenebrionoidea was inferred to originate in the late Jurassic (146.73 Mya, 95% confidence interval [CI]: 137.83–154.93 Mya), with most families subsequently diverging in the early Cretaceous (Figure 8). The family Mordellidae is the earliest diverged family in the superfamily and is estimated to originate at 69.97 Mya in the late Cretaceous. In the “Meloidae clade,” the family Meloidae is estimated to be derived at 67.84 Mya in the late Cretaceous, whereas the Oedemeridae at 53.34 Mya in the Paleogene. In the “Tenebrionidae clade,” the families Lagriidae and Tenebrionidae are estimated to originate at 111.45 and 111.49 Mya in Cretaceous, respectively. In the family Lagriidae, the subfamily Statiriinae diverged 70.62 Mya in the late Cretaceous, whereas the Lagriinae diverged 56.44 Mya in the early Paleogene. In the family Tenebrionidae, the subfamilies Pimeliinae, Alleculinae, and Stenochiinae originated at 72.25, 37.36, and 46.82 Mya between Cretaceous and Paleogene, respectively. All of these families/subfamilies are proposed to be monophyletic and confirmed in relationships in the phylogenetic analyses using different data or inferring methods, and others are not determined for their monophyly or relationships or have a few species included in the phylogenetic analyses, and therefore, they are not given an inferring of divergence time.
[IMAGE OMITTED. SEE PDF]
DISCUSSION
Characteristics of Tenebrionoidea mtgenomes
These 90 mtgenomes investigated in the present study in the superfamily Tenebrionoidea have a length variation from 14,777 to 16,861 bp, and the length variation mainly stems from CR, intergenic overlap, and spacers, which is consistent with earlier reports in Tenebrionoids (Burger et al., 2003). The nucleotide composition for all species exhibits obvious AT bias with high A + T content, similar to earlier reports in Tenebrionoids (Jie et al., 2016). All tRNA genes can form a complete clover secondary structure, except for tRNA-Ser (AGN) which lacks the DHU arm, which seems to be a common feature of Tenebrionoidea (Song et al., 2018; Zhang et al., 2016). There are some rearrangement events in some species of the family Mordellidae, Lagriidae, and Pyrochroidae. The shuffling of the trnR and trnN genes (trnA-trnR-trnN-trnS-trnE-trnF gene cluster) is found in all specials in the family Mordellidae; the translocation of the trnD and ATP8 genes is found in the A. rugipennis of Lagriidae for the first time; the shuffling of the trnC and trnW genes (trnW-trnC-trnY gene cluster) is found in Pyrochroidae and Lagriidae. These gene rearrangements may be produced by abnormal priming of mitochondrial replication by a tRNA molecule to tandem duplications, which can provide an important reference for Tenebrionoidea phylogeny inference (Boore et al., 1998; Boore & Brown, 1998; Cameron, 2014; Timmermans & Vogler, 2012). The ATN and TTG are mainly used as the start codon, and TAA and TAG as the stop codon for the 13 PCGs, which is similar to other mtgenome sequences in Tenebrionoidea (Du et al., 2017). The Ka/Ks ratio is lower than one for all PCGs, which is consistent with earlier studies in Tenebrionoidea. The COX1 gene has experienced strong evolutionary pressure in order to maintain its own functional requirements, whereas ATP8 has experienced weak evolutionary pressures with allowing more mutations to accumulate in the mtgenome (Bai et al., 2018; Ou et al., 2016).
Overview of phylogenetic relationships
A total of 16 families are included in the phylogenetics and evolution analysis of the Tenebrionoidea, in which there are 10 families with at least two representative species included. The family Ciidae seems to be the earliest derived in these families, followed Mordellidae + Ripiphoridae, and Aderidae + “Meloidae clade” + “Tenebrionidae clade”. Ciidae was historily placed in the Cucujoidea (Crowson, 1955) and then to the superfamily Tenebrionoidea mainly based on characteristics of the aedeagus and the larval abdomen (Crowson, 1960). It was proposed to be a monophyly based on 18S and COX1 genes using ML and BI methods and demonstrated to be either a sister to Nitidulidae based on the reduced sample or at the base of the cucujoid-tenebrionoid assemblage based on the entire sample (Buder et al., 2008). It was considered basal tenebrionoids based on 516 adult and larval morphological characteristics from 359 beetle taxa (Lawrence et al., 2011). The present study also suggests the family to be the basal tenebrionoids, but further investigation is necessary to elucidate its place with the inclusion of more species.
In this study, the “Mordellidae + Ripiphoridae” clade was proposed to be monophyletic, and the Mordellidae was also proposed to be monophyletic. The clade of Mordellidae + Ripiphoridae was proposed monophyletic in earlier molecular phylogeny inference based on five nuclear and mitochondrial genes with 300 genera in Tenebrionoidea using ML (Gunter et al., 2014). The Mordellidae was also proposed to be monophyletic in the study based on four molecular genes (18S rRNA, 28S rRNA, rrnL, and COX1) with 128 species in Tenebrionoidea using ML (Batelka et al., 2016). Due to the insufficient sampling of Ripiphoridae, we will not discuss whether it is monophyletic here. However, it was proposed to be paraphyletic in the molecular phylogenetic study based on four mitochondrial and four nuclear gene fragments across 404 taxa (including 250 tenebrionid species) using ML (Kergoat, Soldati, et al., 2014).
Aderidae was proposed to be a monophyletic lineage in the earlier molecular phylogenetic study (Gunter et al., 2014). There is only one species in Aderidae to be included in the present study, which was close to the clade of “Meloidae clade” + “Tenebrionidae clade”, and its position and monophyletic status are yet to be resolved with more species involved. The “Meloidae clade” + “Tenebrionidae clade,” “Meloidae clade,” and “Tenebrionidae clade” are all proposed to be monophyletic, which are consistent with earlier studies based on five nuclear and mitochondrial genes with 300 genera in Tenebrionoidea using ML (Gunter et al., 2014) and eight mitochondrial and nuclear gene with 404 taxa in Coleoptera using ML (Kergoat, Soldati, et al., 2014).
Phylogenetic relationships of “Meloidae clade”
In the “Meloidae + Anthicidae” clade, Meloidae was proposed to be monophyletic in the present work, which supported an earlier study based on 4818 nuclear genes in 146 species of beetles using ML. In the present work, Anthicidae is identified as a sister group to Meloidae; however, in an earlier study, Meloidae was included within Anthicidae. Due to only three species of Anthicidae involved in the present work, the phylogeny of “Meloidae + Anthicidae” clade needs further investigation with the inclusion of more species and more molecular data.
In the “Oedemeridae clade”, the family Prostomidae seems to be located at the base of “Oedemeridae clade”, followed Oedemeridae and Trictenotomidae + Tetratomidae. Prostomidae was proposed a sister to the “pythid-pyrochroid-lineage” (including Trictenomatidae, Pyrochroidae, Salpingidae, and so on) based on morphology characteristics of the maxillary articulatory area, the abdominal tergite IX extending to the ventral side of the segment, and the strongly pronounced prognathous condition (Schunger et al., 2003). However, Prostomidae and Tetratomidae are clustered as one clade in ML tree, and the phylogenetic position of Prostomidae remained unresolved in the present analyses.
The family Oedemeridae is proposed to be monophyletic, which is consistent with the earlier phylogenetic study based on four gene (16S rRNA, COX1, 28S and 18S rRNA) sequences from 8441 taxa of Coleoptera, which removed misplaced single specimens and minor clades (Bocak et al., 2014). However, it was proposed to be paraphyletic in the phylogenetic study based on mitochondrial and nuclear genes (Gunter et al., 2014; Zhang et al., 2018), which suggests that the monophyletic status of the Oedemeridae need be further determined with more species included. Trictenotomidae and Tetratomidae are clustered as one clade using BI in this study, whereas Trictenotomidae was a sister to Boridae in earlier phylogenetic studies based on nuclear genes. There is only species in Trictenotomidae and Tetratomidae to be included in the present study, which suggests that the monophyletic status of the Trictenotomidae and Tetratomidae remains uncertain, and its position and monophyletic status are yet to be resolved with more species to involved.
Zopheridae and Pyrochroidae are clustered as one clade, and Pyrochroidae seems monophyletic in the present study. The monophyly of Pyrochroidae was also proposed in the earlier phylogenetic study based on 95 nuclear protein-coding genes in 373 beetle species using ML and BI (Zhang et al., 2018) and in other phylogenetic studies based on mitochondrial and nuclear genes (Gunter et al., 2014; Kergoat, Soldati, et al., 2014). Zopheridae was a sister to Tetratomidae in earlier phylogenetic studies based on mitochondrial and nuclear genes (Kergoat, Soldati, et al., 2014), which suggests that the position and monophyletic status of Zopheridae and Pyrochroidae need to be further determined. The families Salpingidae and Scraptiidae seem monophyletic and are sister groups of each other in the present study. The monophyly of Salpingidae and Scraptiidae was also proposed in the earlier phylogenetic studies based on mitochondrial and nuclear genes (Bocak et al., 2014; Zhang et al., 2018), whereas Scraptiidae was proposed a paraphyletic group in other molecular phylogenetic studies (Hunt et al., 2007; Kergoat, Soldati, et al., 2014; Mckenna et al., 2015). The sister relationship of Salpingidae and Scraptiidae remains unclear due to the limited inclusion of only two or three species. Therefore, the position and monophyletic status of Salpingidae and Scraptiidae need to be further determined with more species included.
Phylogenetic relationships of “Tenebrionidae clade”
In the “Tenebrionidae clade,” the family Lagriidae and Tenebrionidae are monophyletic, and both of them are sister groups each other in the present study. The classification of Lagriidae has been debated over the years, and some scholars argue that Lagriidae should be classified as a subfamily within Tenebrionidae. However, the beetles of Lagriidae are leafivorous-like beetles of Chrysomelidae, and the morphological characteristics of Lagriidae adapt much more for free-moving and leaf-feeding than the beetles from other subfamilies in Tenebrionidae. The monophyly of Lagriidae was proposed in the earlier morphology study based on the characteristics of abdominal defense glands, female reproductive tract, mouthparts morphology, and structure of the wings in Lagriid and Tenebrionoid (Doyen & Tschinkel, 1982). The monophyly was also proposed in earlier phylogenetic inference based on mtgenomes and nuclear genes (Gunter et al., 2014; Kergoat, Soldati, et al., 2014). The mtgenome-based phylogeny of 36 species in Tenebrionidae suggested the monophyly of Lagriidae and Tenebrionidae and their sister relationships based on PCG123 data sets using BI and ML (Wu et al., 2021), which is consistent with the present study. The monophyly of Tenebrionidae was also proposed in some earlier phylogenetic studies based on mitochondrial and nuclear genes (Hunt et al., 2007; Timmermans et al., 2015).
The present study suggests the phylogenetic relationships of Adeliinae + (Lagriinae + Statininae) in the family Lagriidae and the monophyly of Lagriinae and Statininae. The monophyly of Lagriinae was proposed in earlier phylogenetic studies based on the morphology characteristics (Doyen, 1989) and also based on mitochondrial and nuclear genes (Gunter et al., 2014; Wu et al., 2021). The present study supports the monophyly of the subfamily Statiriinae for the first time. The Adeliinae is at the base of the family Lagriidae in this study, which is consistent with the earlier mtgenome-based study (Wu et al., 2021). The monophyly of Adeliinae in Lagriidae is not yet determined due to only one species to be included.
In the family Tenebrionidae, the present study supports the monophyly of the subfamily Pimeliinae, Stenochiinae, and Alleculinae, whereas Tenebrioninae and Diaperinae are recognized as polyphyly. The subfamily Pimeliinae was also proposed to be monophyletic in earlier studies based on nuclear genes and mitochondrial genes (Gunter et al., 2014; Wu et al., 2021). The present study supports the monophyly of the subfamilies Stenochiinae and Alleculinae in Tenebrionidae, which was also proposed in earlier phylogenetic studies based on mitochondrial and nuclear genes (Kergoat, Soldati, et al., 2014; Wu et al., 2021). The subfamily Tenebrioninae and Diaperinae in these molecular-based studies are found to be polyphyletic, which ineeds to be elucidated with more species involved.
Evolution of Tenebrionoidea
The Tenebrionoidea is inferred to originate in the late Jurassic (146.73 Mya, CI: 137.83–154.93 Mya) based on the mtgenomes and fossil calibrations points in the present study, which is consistent with earlier evolution studies in Coleoptera based on mitochondrial and nuclear genes (Cai et al., 2022; Mckenna et al., 2015; Zhang et al., 2016). The present results suggest that most families subsequently diverged in the Cretaceous. Angiosperms replaced the previously dominant gymnosperms during the Cretaceous, and the warm and humid environment had been produced in the Cretaceous, which provided food and habitat for the families in Tenebrionoidea. The family Ciidae seems to be the earliest derived in these families in the present study, which is inconsistent with earlier evolution studies (Kergoat, Bouchard, et al., 2014). The divergence time of Ciidae is not yet determined because only one species being included. The families Mordellidae and Ripiphoridae are among the earliest diverged families in the superfamily, Mordellidae is estimated to originate at 65.22 Mya in the late Cretaceous, which is consistent with the previous evolution study result based on 95 nuclear protein-coding genes in 373 beetle species using ML and BI (Zhang et al., 2018). In the “Meloidae clade,” the families Meloidae (67.84 Mya) and Oedemeridae (53.34 Mya) originated in the late Cretaceous and early Paleogene, which is consistent with earlier evolution studies based on mitochondrial and nuclear genes (Kergoat, Bouchard, et al., 2014; López-Estrada, 2019, 2022; Misof et al., 2014). In the “Tenebrionidae clade,” the family Lagriidae (111.45 Mya) was proposed to be derived in the middle Cretaceous, which is similar to the evolution study based on mitochondrial and nuclear genes in 404 beetle species (Kergoat, Bouchard, et al., 2014). The family Tenebrionidae (111.49 Mya) was suggested to be derived in the Cretaceous, which is consistent with the results of the previous evolution study based on 4818 nuclear genes.
In the family Lagriidae, the subfamilies Lagriinae (56.44 Mya) and Statiriinae (70.62 Mya) are proposed to be derived in the early Paleogene and late Cretaceous for the first time. In the family Tenebrionidae, the present study suggests the origin of the subfamilies Alleculinae (37.36 Mya) and Stenochiinae (46.82 Mya) in the Paleogene but is inconsistent with the results of the earlier evolution study (Kergoat, Bouchard, et al., 2014), which be due to differences in the taxa included, fossils constraints, and analysis methods applied. In further research, more accurate estimates of divergence times are necessary along with more precise fossil records for calibration and more complete sampling.
CONCLUSION
This is the first comprehensive study on the mtgenomes characteristics and mtgenome-based phylogenetics in Tenebrionoidea. In the present study, we completely sequenced mitogenomes from 19 species belonging to Tenebrionoidea. The comprehensive analysis of 90 mtgenome sequences in Tenebrionoidea suggests that the AT-skew, length variation, and codon usage are consistent with other reported mtgenomes in Tenebrionoidea. The families Mordellidae, Meloidae, Oedemeridae, Pyrochroidae, Salpingidae, Scraptiidae, Lagriidae, and Tenebrionidae were suggested to be monophyletic. Ciidae is at the base of the superfamily of Tenebrionoidea, and the Mordellidae is close to the Ripiphoridae. The “Tenebrionidae clade” and “Meloidae clade” are monophyletic, and both of them are sister groups. In the “Meloidae clade,” Meloidae is close to Anthicidae. In the “Tenebrionidae clade,” the family Lagriidae and Tenebrionidae are sister groups. In Lagriidae, the subfamily Adeliinae is based on the subfamilies Lagriinae + Statininae. In Tenebrionidae, the subfamilies Pimeliinae, Alleculinae, and Stenochiinae were recovered monophyletic, and Tenebrioninae and Diaperinae are polyphyletic. The divergence time analysis suggests that Tenebrionoidea originated in the late Jurassic, Meloidae Mordellidae, Lagriidae, and Tenebrionidae in the Cretaceous, Oedemeridae in Paleogene.
AUTHOR CONTRIBUTIONS
Yun-Jian Hu: Conceptualization (equal); software (equal); supervision (equal); validation (equal); writing – original draft (equal). Feng-Fan Jia: Data curation (equal); software (equal); visualization (equal). Li Hu: Software (equal); visualization (equal); writing – original draft (equal). Chuan Wu: Investigation (equal); supervision (equal). Tian Tian: Investigation (equal); methodology (equal). Ting-Jing Li: Conceptualization (equal); methodology (equal). Bin Chen: Conceptualization (equal); investigation (equal); methodology (equal); project administration (equal); resources (equal); software (equal); supervision (equal); validation (equal); writing – review and editing (equal).
ACKNOWLEDGMENTS
We also thank the support and assistance provided by the Daba Mountain Nature Reserve, whose cooperation enabled us to conduct field investigations smoothly.
FUNDING INFORMATION
This research was supported by the National Natural Science Foundation of China (31872262, 31672363) and the National Key Program of Science and Technology Foundation Work of China (2015FY210300).
CONFLICT OF INTEREST STATEMENT
The authors declare that they have no competing interests.
DATA AVAILABILITY STATEMENT
All data are available as tables and figures in the main paper and its supplementary files. The GenBank accession numbers for the 19 mtgenomes generated in the present study with detailed information in Table S1.
Altschul, S. F., Gish, W., Miller, W., Myers, E. W., & Lipman, D. J. (1990). Basic local alignment search tool. Journal of Molecular Biology, 215, 403–410.
Bai, Y., Li, C., Yang, M., & Liang, S. (2018). Complete mitochondrial genome of the dark mealworm Tenebrio obscurus Fabricius (Insecta: Coleoptera: Tenebrionidae). Mitochondrial DNA Part B‐Resource, 3, 171–172.
Bankevich, A., Nurk, S., Antipov, D., Gurevich, A. A., Dvorkin, M., Kulikov, A. S., Lesin, V. M., Nikolenko, S. I., Pham, S., Prjibelski, A. D., Pyshkin, A. V., Sirotkin, A. V., Vyahhi, N., Tesler, G., Alekseyev, M. A., & Pevzner, P. A. (2012). SPAdes: A new genome assembly algorithm and its applications to single‐cell sequencing. Journal of Computational Biology, 19, 455–477.
Batelka, J., Kundrata, R., & Bocak, L. (2016). Position and relationships of Ripiphoridae (Coleoptera: Tenebrionoidea) inferred from ribosomal and mitochondrial molecular markers. Annales Zoologici, 66(1), 113–123.
Bernt, M., Donath, A., Juhling, F., Externbrink, F., Florentz, C., Fritzsch, G., Putz, J., Middendorf, M., & Stadler, P. F. (2013). MITOS: Improved de novo metazoan mitochondrial genome annotation. Molecular Phylogenetics and Evolution, 69, 313–319.
Beutel, R. G., & Friedrich, F. (2005). Comparative study of larvae of Tenebrionoidea (Coleoptera: Cucujiformia). European Journal of Entomology, 102, 241–264.
Bocak, L., Barton, C., Crampton‐Platt, A., Chesters, D., Ahrens, D., & Vogler, A. P. (2014). Building the Coleoptera tree‐of‐life for >8000 species: Composition of public DNA data and fit with Linnaean classification. Systematic Entomology, 39, 97–110.
Boore, J. L., & Brown, W. M. (1998). Big trees from little genomes: Mitochondrial gene order as a phylogenetic tool. Current Opinion in Genetics and Development, 8(6), 668–674.
Boore, J. L., Lavrov, D. V., & Brown, W. M. (1998). Gene translocation links insects and crustaceans. Nature, 392(6677), 667–668.
Bouchard, P., Bousquet, Y., Davies, A., Alonso‐Zarazaga, M., Lawrence, J., Lyal, C., Newton, A., Reid, C., Schmitt, M., Slipinski, A., & Smith, A. (2011). Family‐group names in Coleoptera (Insecta). ZooKeys, 88, 1–972.
Buder, G., Grossmann, C., Hundsdoerfer, A., & Klass, K. D. (2008). A contribution to the phylogeny of the Ciidae and its relationships with other cucujoid and tenebrionoid beetles (Coleoptera: Cucujiformia). Arthropod Systematics and Phylogeny, 66(2), 165–190.
Burger, G., Gray, M. W., & Lang, B. F. (2003). Mitochondrial genomes: Anything goes. Trends in Genetics, 19, 709–716.
Caballero, S. Z., & de León, P. V. (2003). Una especie nueva de Epicauta (Coleoptera: Meloidae) del Plioceno del estado de Hidalgo, México. Revista Mexicana de Ciencias Geológicas, 20(2), 154–159.
Cai, C., Tihelka, E., Giacomelli, M., Lawrence, J. F., Ślipiński, A., Kundrata, R., Yamamoto, S., Thayer, M. K., Newton, A. F., Leschen, R. A., Gimmel, M. L., Lü, L., Engel, M. S., Huang, D., Pisani, D., & Donoghue, P. C. (2022). Integrated phylogenomics and fossil data illuminate the evolution of beetles. Royal Society Open Science, 9(3), [eLocator: 211771].
Cai, C., Yin, Z., & Huang, D. (2018). A new ripiphorid beetle from upper Cretaceous Burmese amber sheds light on early evolution of the extant subfamily Ripidiinae (Coleoptera: Ripiphoridae). Comptes Rendus Palevol, 17(6), 351–356.
Cameron, S. L. (2014). Insect mitochondrial genomics: Implications for evolution and phylogeny. Annual Review of Entomology, 59, 95–117.
Cameron, S. L., Sullivan, J., Song, H., Miller, K. B., & Whiting, M. F. (2009). A mitochondrial genome phylogeny of the Neuropterida (lace‐wings, alderflies and snakeflies) and their relationship to the other holometabolous insect orders. Zoologica Scripta, 38(6), 575–590.
Crampton‐Platt, A., Timmermans, M. J., Gimmel, M. L., Kutty, S. N., Cockerill, T. D., Vun Khen, C., & Vogler, A. P. (2015). Soup to tree: The phylogeny of beetles inferred by mitochondrial metagenomics of a Bornean rainforest sample. Molecular Biology and Evolution, 32(9), 2302–2316.
Crowson, R. A. (1955). The natural classification of the families of Coleoptera. Nathaniel Lloyd.
Crowson, R. A. (1960). The phylogeny of Coleoptera. Annual Review of Entomology, 5, 111–134.
Darty, K., Denise, A., & Ponty, Y. (2009). VARNA: Interactive drawing and editing of the RNA secondary structure. Bioinformatics, 25, 1974–1975.
Dierckxsens, N., Mardulyn, P., & Smits, G. (2016). NOVOPlasty: de novo assembly of organelle genomes from whole genome data. Nucleic Acids Research, 45(4), [eLocator: e18].
Doyen, J. T. (1989). Reconstitution of Coelometopini, Tenebrionini and related tribes of America North of Colombia (Coleoptera: Tenebrionidae). Journal of the New York Entomological Society, 97, 277–304.
Doyen, J. T., & Tschinkel, W. R. (1982). Phenetic and cladistic relationships among tenebionid beetles (Coleoptera). Systematic Entomology, 7, 127–183.
Drummond, A. J., & Rambaut, A. (2007). BEAST: Bayesian evolutionary analysis by sampling trees. BMC Evolutionary Biology, 7(1), [eLocator: 214].
Du, C., Zhang, L., Lu, T., Ma, J., Zeng, C., Yue, B., & Zhang, X. (2017). Mitochondrial genomes of blister beetles (Coleoptera, Meloidae) and two large intergenic spacers in Hycleus genera. BMC Genomics, 18(1), [eLocator: 698].
Feng, R. Q., Zhang, L. J., Li, M., Liu, J., Wen, C. L., & Yuan, M. L. (2020). Mitochondrial genome of Chrysochares punctatus (Coleoptera: Chrysomelidae: Eumolpinae) and phylogenetic analysis. Mitochondrial DNA Part B Resources, 5(1), 667–668.
Friedrich, M., & Muqim, N. (2003). Sequence and phylogenetic analysis of the complete mitochondrial genome of the flour beetle Tribolium castanaeum. Molecular Phylogenetics and Evolution, 26(3), 502–512.
Grau, T., Vilcinskas, A., & Joop, G. (2017). Sustainable farming of the mealworm Tenebrio molitor for the production of food and feed. Zeitschrift für Naturforschung. C, A Journal of Biosciences, 72(9–10), 337–349.
Gunter, N. L., Levkaničová, Z., Weir, T. H., Ślipiński, A., Cameron, S. L., & Bocak, L. (2014). Towards a phylogeny of the Tenebrionoidea (Coleoptera). Molecular Phylogenetics and Evolution, 79, 305–312.
Hebert, P. D., Cywinska, A., Ball, S. L., & Dewaard, J. R. (2003). Biological identifications through DNA barcodes. Proceedings of the Royal Society B‐Biological Sciences, 270, 313–321.
Hong, K. J., Ki, W., Park, D. S., Yang, B. K., Lee, H., Park, J., & Lee, W. (2020). The complete mitochondrial genome of Alphitobius diaperinus Panzer, 1797 (Coleoptera: Tenebrionidae). Mitochondrial DNA Part B Resources, 5(3), 2291–2293.
Hunt, T., Bergsten, J., Levkanicova, Z., Papadopoulou, A., John, O. S., Wild, R., Hammond, P. M., Ahrens, D., Balke, M., Caterino, M. S., Gomez‐Zurita, J., Ribera, I., Barraclough, T. G., Bocakova, M., Bocak, L., & Vogler, A. P. (2007). A comprehensive phylogeny of beetles reveals the evolutionary origins of a superradiation. Science, 318, 1913–1916.
Jie, H., Lei, M., Li, P., Feng, X., Zeng, D., Zhao, G., Zhu, J., Zhang, C., Yu, M., Huang, Y., & Chen, Q. (2016). The complete nucleotide sequence of the mitochondrial genome of Epicauta aptera Kaszab. Mitochondrial DNA Part B Resources, 1(1), 489–490.
Katoh, K., & Standley, D. M. (2013). MAFFT multiple sequence alignment software version 7: Improvements in performance and usability. Molecular Biology and Evolution, 30, 772–780.
Kearse, M., Moir, R., Wilson, A., Stones‐Havas, S., Cheung, M., Sturrock, S., Buxton, S., Cooper, A., Markowitz, S., Duran, C., Thierer, T., Ashton, B., Meintjes, P., & Drummond, A. (2012). Geneious basic: An integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics, 28, 1647–1649.
Kergoat, G. J., Bouchard, P., Clamens, A. L., Abbate, J. L., Jourdan, H., Jabbour‐Zahab, R., Genson, G., Soldati, L., & Condamine, F. L. (2014). Cretaceous environmental changes led to high extinction rates in a hyperdiverse beetle family. BMC Evolutionary Biology, 14, 220.
Kergoat, G. J., Soldati, L., Clamens, A. L., Jourdan, H., Jabbour‐Zahab, R., Genson, G., Bouchard, P., & Condamine, F. L. (2014). Higher level molecular phylogeny of darkling beetles (Coleoptera: Tenebrionidae). Systematic Entomology, 39, 486–499.
Kirejtshuk, A. G., & Azar, D. (2008). New taxa of beetles (Insecta, Coleoptera) from Lebanese amber with evolutionary and systematic comments. Alavesia, 2(15), [eLocator: e46].
Kirejtshuk, A. G., Nabozhenko, M. V., & Nel, A. (2012). First mesozoic representative of the subfamily tenebrioninae (Coleoptera, Tenebrionidae) from the lower cretaceous of Yixian (China, Liaoning). Entomological Review, 92(1), 97–100.
Kumar, S., Stecher, G., Li, M., Knyaz, C., & Tamura, K. (2018). MEGA‐X: Molecular evolutionary genetics analysis across computing platforms. Molecular Biology and Evolution, 35(6), 1547–1549.
Lanfear, R., Frandsen, P. B., Wright, A. M., Senfeld, T., & Calcott, B. (2016). PartitionFinder 2: New methods for selecting partitioned models of evolution formolecular and morphological phylogenetic analyses. Molecular Biology and Evolution, 34(3), 772–773.
Lawrence, J. F. (1995). Families and subfamilies of Coleoptera (with selected genera, notes, references and data on family‐group names). Biology, Phylogeny, and Classification of Coleoptera, 2, 779–1006.
Lawrence, J. F., & Newton, A. F., Jr. (1982). Evolution and classification of beetles. Annual Review of Ecology and Systematics, 13(1), 261–290.
Lawrence, J. F., Ślipiński, A., Seago, A. E., Thayer, M. K., Newton, A. F., & Marvaldi, A. E. (2011). Phylogeny of the Coleoptera based on morphological characters of adults and larvae. Annales Zoologici, 61, 1–217.
Letunic, L., & Bork, P. (2016). Interactive tree of life (iTOL) v3: An online tool for the display and annotation of phylogenetic and other trees. Nucleic Acids Research, 44(W1), W242–W245.
Librado, P., & Rozas, J. (2009). DnaSP v5: A software for comprehensive analysis of DNA polymorphism data. Bioinformatics, 25(11), 1451–1452.
Li‐Na, L. I. U., & Cheng‐Ye, W. A. N. G. (2014). Complete mitochondrial genome of yellow meal worm (Tenebrio molitor). Zoological Research, 35(6), [eLocator: 537].
Linard, B., Arribas, P., Andújar, C., Crampton‐Platt, A., & Vogler, A. P. (2016). Lessons from genome skimming of arthropod‐preserving ethanol. Molecular Ecology Resources, 16(6), 1365–1377.
López‐Estrada, E. K., Sanmartín, I., García‐París, M., & Zaldívar‐Riverón, A. (2019). High extinction rates and non‐adaptive radiation explains patterns of low diversity and extreme morphological disparity in North American blister beetles (Coleoptera, Meloidae). Molecular Phylogenetics and Evolution, 130, 156–168.
López‐Estrada, E. K., Sanmartín, I., Uribe, J. E., Abalde, S., Jiménez‐Ruiz, Y., & García‐París, M. (2022). Mitogenomics and hidden‐trait models reveal the role of phoresy and host shifts in the diversification of parasitoid blister beetles (Coleoptera: Meloidae). Molecular Ecology, 31(8), 2453–2474.
Lowe, T. M., & Eddy, S. R. (1997). tRNAscan‐SE: A program for improved detection of transfer RNA genes in genomic sequence. Nucleic Acids Research, 25, 955–964.
McKenna, D. D., Wild, A. L., Kanda, K., Bellamy, C. L., Beutel, R. G., Caterino, M. S., Farnum, C. W., Hawks, D. C., Ivie, M. A., Jameson, M. L., Leschen, R. A. B., Marvaldi, A. E., McHugh, J. V., Newton, A. F., Robertson, J. A., Thayer, M. K., Whiting, M. F., Lawrence, J. F., Ślipiński, A., … Farrell, B. D. (2015). The beetle tree of life reveals that Coleoptera survived end‐Permian mass extinction to diversify during the cretaceous terrestrial revolution. Systematic Entomology, 40, 835–880.
Mikrajuddin, A., & Khairurrijal, A. (2009). Simple method for determining surface porosity based on SEM images using origin pro software. Indonesian Journal of Physics, 20, 37–41.
Minh, B. Q., Nguyen, M. A. T., & von Haeseler, A. (2013). Ultrafast approximation for phylogenetic bootstrap. Molecular Biology and Evolution, 30, 1188–1195.
Misof, B., Liu, S., Meusemann, K., Peters, R. S., Donath, A., Mayer, C., Frandsen, P. B., Ware, J., Flouri, T., Beutel, R. G., Niehuis, O., Petersen, M., Izquierdo‐Carrasco, F., Wappler, T., Rust, J., Aberer, A. J., Aspöck, U., Aspöck, H., Bartel, D., … Zhou, X. (2014). Phylogenomics resolves the timing and pattern of insect evolution. Science, 346(6210), 763–767.
Ou, J., Liu, J. B., Yao, F. J., Wang, X. G., & Wei, Z. M. (2016). The complete mitochondrial genome of the American black flour beetle Tribolium audax (Coleoptera: Tenebrionidae). Mitochondrial DNA Part A DNA Mapping, Sequencing, and Analysis, 27(2), 958–959.
Ou, J., Yao, F. J., Li, Y., Yang, X. Y., Jin, C., & Wei, Z. M. (2016). The complete mitochondrial genome of the confused flour beetle Tribolium confusum (Coleoptera: Tenebrionidae). Mitochondrial DNA, 27(2), 958–959.
Perna, N. T., & Kocher, T. D. (1995). Patterns of nucleotide composition at fourfold degenerate sites of animal mitochondrial genomes. Journal of Molecular Evolution, 41, 353–358.
Poinar, G. O., Jr. (2009). Meloe dominicanus n. sp. (Coleoptera: Meloidae) phoretic on the bee Proplebia dominicana (Hymenoptera: Apidae) in Dominican amber. Proceedings of the Entomological Society of Washington, 111(1), 145–150.
Rider, S. D., Jr. (2016). The complete mitochondrial genome of the desert darkling beetle Asbolus verrucosus (Coleoptera, Tenebrionidae). Mitochondrial DNA Part A DNA Mapping, Sequencing, and Analysis, 27(4), 2447–2449.
Ronquist, F., Teslenko, M., Van Der Mark, P., Ayres, D. L., Darling, A., Hohna, S., Larget, B., Liu, L., Suchard, M. A., & Huelsenbeck, J. P. (2012). MrBayes 3.2: Efficient Bayesian phylogenetic inference and model choice across a large model space. Systematic Biology, 61, 539–542.
Schunger, I., Beutel, R. G., & Britz, R. (2003). Morphology of immature stages of Prostomis mandibularis (Coleoptera: Tenebrionoidea: Prostomidae). European Journal of Entomology, 100(3), 357–370.
Seo, B. Y., Park, J., Kwon, W., & Park, J. (2019). The complete mitochondrial genome of Aiolocaria hexaspilota (Hope, 1831) (Coleoptera: Coccinellidae). Mitochondrial DNA Part B Resources, 4(1), 1472–1474.
Sheffield, N. C., Song, H., Cameron, S. L., & Whiting, M. F. (2008). A comparative analysis of mitochondrial genomes in Coleoptera (Arthropoda: Insecta) and genome descriptions of six new beetles. Molecular Biology and Evolution, 25, 2499–2509.
Sheng, L., Zhou, T., Shi, Z., Pan, X., Weng, X., Ma, J., & Wu, S. (2021). The complete mitochondrial genome of Trictenotoma davidi Deyrolle, 1875 (Coleoptera: Trictenotomidae). Mitochondrial DNA Part B Resources, 6(7), 2026–2027.
Shintani, Y., Hirose, Y., & Terao, M. (2011). Effects of temperature, photoperiod and soil humidity on induction of pseudopupal diapause in the bean blister beetle Epicauta gorhami. Physiological Entomology, 36(1), 14–20.
Ślipiński, S. A., Leschen, R. A. B., & Lawrence, J. F. (2011). Order Coleoptera Linnaeus, 1758. Animal biodiversity: An outline of higher‐level classification and survey of taxonomic richness (ed. by Zhang Z. Q.). Zootaxa, 3148, 203–208.
Smith, A. D., Kamiński, M. J., Kanda, K., Sweet, A. D., Betancourt, J. L., Holmgren, C. A., Hempel, E., Alberti, F., & Hofreiter, M. (2021). Recovery and analysis of ancient beetle DNA from subfossil packrat middens using high‐throughput sequencing. Scientific Reports, 11(1), [eLocator: 12635].
Song, N., Liu, H. Y., Yang, X. J., Zhao, X. C., & Lin, A. L. (2018). Complete mitochondrial genome of the darkling beetle Gonocephalum outreyi (Coleoptera: Tenebrionidae) with phylogenetic implications. Journal of Asia‐Pacific Entomology, 21, 721–730.
Talavera, G., & Castresana, J. (2007). Improvement of phylogenies after removing divergent and ambiguously aligned blocks from protein sequence alignments. Systematic Biology, 56, 564–577.
Tang, P. A., Feng, R. Q., Zhang, L., Wang, J., Wang, X. T., Zhang, L. J., & Yuan, M. L. (2020). Mitochondrial genomes of three Bostrichiformia species and phylogenetic analysis of Polyphaga (Insecta, Coleoptera). Genomics, 112, 2970–2977.
Timmermans, M. J. T. N., Barton, C., Haran, J., Ahrens, D., Culverwell, C. L., Ollikainen, A., Dodsworth, S., Foster, P. G., Bocak, L., & Vogler, A. P. (2015). Family‐level sampling of mitochondrial genomes in Coleoptera: Compositional heterogeneity and phylogenetics. Genome Biology and Evolution, 8, 161–175.
Timmermans, M. J., & Vogler, A. P. (2012). Phylogenetically informative rearrangements in mitochondrial genomes of Coleoptera, and monophyly of aquatic elateriform beetles (Dryopoidea). Molecular Phylogenetics and Evolution, 63(2), 299–304.
Vaidya, G., Lohman, D. J., & Meier, R. (2011). SequenceMatrix: Concatenation software for the fast assembly of multi–gene datasets with character set and codon information. Cladistics, 27(2), 171–180.
Wang, C. Y., Feng, Y., & Chen, X. M. (2016). The complete mitochondrial genome of a medicinal insect Martianus dermestoides (Coleoptera, Tenebrionidae). Mitochondrial DNA Part A DNA Mapping, Sequencing, and Analysis, 27(3), 1627–1628.
Wu, C., Zhou, Y., Tian, T., Li, T. J., & Chen, B. (2021). First report of complete mitochondrial genome in the subfamily Alleculinae and mitochondrial genome‐based phylogenetics in Tenebrionidae (Coleoptera: Tenebrionoidea). Insect Science, 29(4), 1226–1238.
Xia, X. H. (2013). DAMBE5: A comprehensive software package for data analysis in molecular biology and evolution. Molecular Biology and Evolution, 30, 1720–1728.
Zha, S., Yin, Y., Wang, Y., Huang, Y., Li, Y., & Wang, Z. (2017). Cloning and functional analysis of farnesyl pyrophosphate synthase (FPPS) gene from Mylabris cichorii. Biotechnology and Applied Biochemistry, 64(5), 667–676.
Zhang, D., Gao, F. L., Jakovlic, I., Zou, H., Zhang, J., Li, W. X., & Wang, G. T. (2020). PhyloSuite: An integrated and scalable desktop platform for streamlined molecular sequence data management and evolutionary phylogenetics studies. Molecular Ecology Resources, 20, 348–355.
Zhang, H., Liu, N., Han, Z., & Liu, J. (2016). Phylogenetic analyses and evolutionary timescale of Coleoptera based on mitochondrial sequence. Biochemical Systematics and Ecology, 66, 229–238.
Zhang, S. Q., Che, L. H., Li, Y., Dan, L., Pang, H., Ślipiński, A., & Zhang, P. (2018). Evolutionary history of Coleoptera revealed by extensive sampling of genes and species. Nature Communications, 9, [eLocator: 205].
Zhao, M., He, Z., Wang, C. Y., Sun, L., & Feng, Y. (2019). Mitochondrial genome of a medicinal beetle Blaps rhynchopetera (Coleoptera, Tenebrionidae) and phylogenetic analysis. Mitochondrial DNA Part B Resources, 4(2), 2583–2584.
Zheng, S., Poczai, P., Hyvönen, J., Tang, J., & Amiryousefi, A. (2020). Chloroplot: An online program for versatile plotting of the organelle genomes. Frontiers in Genetics, 11, 1664–8021.
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
© 2024. This work is published under http://creativecommons.org/licenses/by/4.0/ (the "License"). Notwithstanding the ProQuest Terms and Conditions, you may use this content in accordance with the terms of the License.
Abstract
Despite the worldwide distribution and rich diversity of the superfamily Tenebrionoidea, the knowledge of the mitochondrial genomes (mtgenome) characteristics of the superfamily is still very limited, and its phylogenetics and evolution remain unresolved. In the present study, we newly sequenced mtgenomes from 19 species belonging to Tenebrionoidea, and a total of 90 mitochondrial genomes from 16 families of Tenebrionoidea were used for phylogenetic analysis. There exist 37 genes for all 82 species of complete mtgenomes of 16 families investigated, and their characteristics are identical as reported mtgenomes of other Tenebrionoids. The Ka/Ks analysis suggests that all 13 PCGs have undergone a strong purifying selection. The phylogenetic analysis suggests the monophyly of Mordellidae, Meloidae, Oedemeridae, Pyrochroidae, Salpingidae, Scraptiidae, Lagriidae, and Tenebrionidae, and the Mordellidae is close to the Ripiphoridae. The “Tenebrionidae clade” and “Meloidae clade” are monophyletic, and both of them are sister groups. In the “Meloidae clade,” Meloidae is close to Anthicidae. In the “Tenebrionidae clade,” the family Lagriidae and Tenebrionidae are sister groups. The divergence time analysis suggests that Tenebrionoidea originated in the late Jurassic, Meloidae Mordellidae, Lagriidae, and Tenebrionidae in the Cretaceous, Oedemeridae in Paleogene. The work lays a base for the study of mtgenome, phylogenetics, and evolution of the superfamily Tenebrionoidea.
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
Details
1 Chongqing Key Laboratory of Vector Insects, Institute of Entomology and Molecular Biology, College of Life Sciences, Chongqing Normal University, Chongqing, China