Hostname: page-component-7bb8b95d7b-w7rtg Total loading time: 0 Render date: 2024-09-29T20:52:19.171Z Has data issue: false hasContentIssue false

Comparative genomics reveals evolutionary drivers of the dietary shift in Hemiptera

Published online by Cambridge University Press:  15 December 2023

Guangyao Wu
Affiliation:
National Key Laboratory of Green Pesticide, Key Laboratory of Green Pesticide and Agricultural Bioengineering, Ministry of Education, Center for R&D of Fine Chemicals of Guizhou University, Guiyang 550025, China
Chunyan Wu
Affiliation:
National Key Laboratory of Green Pesticide, Key Laboratory of Green Pesticide and Agricultural Bioengineering, Ministry of Education, Center for R&D of Fine Chemicals of Guizhou University, Guiyang 550025, China
Youssef Dewer
Affiliation:
Phytotoxicity Research Department, Central Agricultural Pesticide Laboratory, Agricultural Research Center, Dokki 12618, Giza, Egypt
Peiyao Li
Affiliation:
National Key Laboratory of Green Pesticide, Key Laboratory of Green Pesticide and Agricultural Bioengineering, Ministry of Education, Center for R&D of Fine Chemicals of Guizhou University, Guiyang 550025, China
Baojun Hao
Affiliation:
School of Life and Health Science, Kaili University, Guizhou 556000, China
Liansheng Zang
Affiliation:
National Key Laboratory of Green Pesticide, Key Laboratory of Green Pesticide and Agricultural Bioengineering, Ministry of Education, Center for R&D of Fine Chemicals of Guizhou University, Guiyang 550025, China
Fengqi Li*
Affiliation:
National Key Laboratory of Green Pesticide, Key Laboratory of Green Pesticide and Agricultural Bioengineering, Ministry of Education, Center for R&D of Fine Chemicals of Guizhou University, Guiyang 550025, China
*
Corresponding author: Fengqi Li; Email: [email protected]
Rights & Permissions [Opens in a new window]

Abstract

Hemiptera insects exhibit a close relationship to plants and demonstrate a diverse range of dietary preferences, encompassing phytophagy as the predominant feeding habit while a minority engages in carnivorous or haematophagous behaviour. To counteract the challenges posed by phytophagous insects, plants have developed an array of toxic compounds, causing significant evolutionary selection pressure on these insects. In this study, we employed a comparative genomics approach to analyse the expansion and contraction of gene families specific to phytophagous insect lineages, along with their adaptive evolutionary traits, utilising representative species from the Hemiptera order. Our investigation revealed substantial expansions of gene families within the phytophagous lineages, especially in the Pentatomomorpha branch represented by Oncopeltus fasciatus and Riptortus pedestris. Notably, these expansions of gene families encoding enzymes are potentially involved in hemipteran-plant interactions. Moreover, the adaptive evolutionary analysis of these lineages revealed a higher prevalence of adaptively evolved genes in the Pentatomomorpha branch. The observed branch-specific gene expansions and adaptive evolution likely contribute significantly to the diversification of species within Hemiptera. These results help enhance our understanding of the genomic characteristics of the evolution of different feeding habits in hemipteran insects.

Type
Research Paper
Copyright
Copyright © The Author(s), 2023. Published by Cambridge University Press

Introduction

Hemiptera encompasses several crucial agricultural pests such as aphids and whiteflies, as well as significant sanitary pests including Triatoma rubrofasciata, Rhodnius prolixus, and Cimex lectularius (Rosenfeld et al., Reference Rosenfeld, Reeves, Brugler, Narechania, Simon, Durrett, Foox, Shianna, Schatz, Gandara, Afshinnekoo, Lam, Hastie, Chan, Cao, Saghbini, Kentsis, Planet, Kholodovych, Tessler, Baker, DeSalle, Sorkin, Kolokotronis, Siddall, Amato and Mason2016; Liu et al., Reference Liu, Guo, Zhang, Hu, Li, Zhu, Zhou, Wu, Chen and Zhou2019; Huang et al., Reference Huang, Ye, Ye, Yan, Wang, Wei, Chen, Li, Sun and Zhang2021; Rijal et al., Reference Rijal, Joyce and Gyawaly2021). Therefore, the study of Hemiptera holds great significance. There are 97,000 to 103,590 known species representing over 10% of all known insect species (Johnson et al., Reference Johnson, Dietrich, Friedrich, Beutel, Wipfler, Peters, Allen, Petersen, Donath, Walden, Kozlov, Podsiadlowski, Mayer, Meusemann, Vasilikopoulos, Waterhouse, Cameron, Weirauch, Swanson, Percy, Hardy, Terry, Liu, Zhou, Misof, Robertson and Yoshizawa2018). The distinguishing characteristic that differentiates Hemiptera from other insects is the variation in mouthparts, which has evolved from chewing mouthparts in their ancestors to piercing-sucking mouthparts to enhance adaptability (Li et al., Reference Li, Leavengood, Chapman, Burkhardt, Song, Jiang, Liu, Zhou and Cai2017). These specialised piercing-sucking mouthparts have undergone continuous modification and adaptation, enabling different Hemiptera species to feed on plant phloem, blood, or other fluids (Li et al., Reference Li, Leavengood, Chapman, Burkhardt, Song, Jiang, Liu, Zhou and Cai2017).

Hemipterans exhibit a remarkable biodiversity, spanning a wide range of life histories and habitats, encompassing both terrestrial and aquatic ecosystems (Li et al., Reference Li, Leavengood, Chapman, Burkhardt, Song, Jiang, Liu, Zhou and Cai2017). The diversity of environments drives their exploration of various food sources such as plants, arthropods, fungi, and vertebrate blood (Rider, Reference Rider1996). Based on previous phylogenetic study of Hemipteroid, there is a temporal correlation between the evolution of true bugs in Hemipteroids and a shift in feeding habits, implying a possible transition from phytophagy to predation (Sweet, Reference Sweet1979; Johnson et al., Reference Johnson, Dietrich, Friedrich, Beutel, Wipfler, Peters, Allen, Petersen, Donath, Walden, Kozlov, Podsiadlowski, Mayer, Meusemann, Vasilikopoulos, Waterhouse, Cameron, Weirauch, Swanson, Percy, Hardy, Terry, Liu, Zhou, Misof, Robertson and Yoshizawa2018). The successful evolution of dietary shifts in Hemiptera insects can be attributed to the interaction between trophic niche and genome-level changes (Weirauch et al., Reference Weirauch, Schuh, Cassis and Wheeler2019). The environment and prevailing conditions are primary factors that influence genetic variation within insect populations, which leads to the divergence of different species. Among various environmental factors, the trophic niche, which denotes the principal source of nourishment, plays a pivotal role in moulding phenotypic and genomic alterations in insects (Johnson et al., Reference Johnson, Dietrich, Friedrich, Beutel, Wipfler, Peters, Allen, Petersen, Donath, Walden, Kozlov, Podsiadlowski, Mayer, Meusemann, Vasilikopoulos, Waterhouse, Cameron, Weirauch, Swanson, Percy, Hardy, Terry, Liu, Zhou, Misof, Robertson and Yoshizawa2018).

Studies investigating dietary variations in Hemiptera revealed that within this insect order, Acyrthosiphon pisum, whitefly, Laodelphax striatellus, Oncopeltus fasciatus, and Riptortus pedestris are classified as phytophagous insects (Li et al., Reference Li, Leavengood, Chapman, Burkhardt, Song, Jiang, Liu, Zhou and Cai2017). On the other hand, T. rubrofasciata, R. prolixus, C. lectularius, and Gerris buenoi are carnivorous insects (Rosenfeld et al., Reference Rosenfeld, Reeves, Brugler, Narechania, Simon, Durrett, Foox, Shianna, Schatz, Gandara, Afshinnekoo, Lam, Hastie, Chan, Cao, Saghbini, Kentsis, Planet, Kholodovych, Tessler, Baker, DeSalle, Sorkin, Kolokotronis, Siddall, Amato and Mason2016; Armisén et al., Reference Armisén, Rajakumar, Friedrich, Benoit, Robertson, Panfilio, Ahn, Poelchau, Chao, Dinh, Doddapaneni, Dugan, Gibbs, Hughes, Han, Lee, Murali, Muzny, Qu, Worley, Munoz-Torres, Abouheif, Bonneton, Chen, Chiang, Childers, Cridge, Crumière, Decaras, Didion, Duncan, Elpidina, Favé, Finet, Jacobs, Cheatle Jarvela, Jennings, Jones, Lesoway, Lovegrove, Martynov, Oppert, Lillico-Ouachour, Rajakumar, Refki, Rosendale, Santos, Toubiana, van der Zee, Vargas Jentzsch, Lowman, Viala, Richards and Khila2018; Liu et al., Reference Liu, Guo, Zhang, Hu, Li, Zhu, Zhou, Wu, Chen and Zhou2019; Huang et al., Reference Huang, Ye, Ye, Yan, Wang, Wei, Chen, Li, Sun and Zhang2021; Rijal et al., Reference Rijal, Joyce and Gyawaly2021). Except G. buenoi, the habitats of these insects are primarily semi-aquatic (Armisén et al., Reference Armisén, Rajakumar, Friedrich, Benoit, Robertson, Panfilio, Ahn, Poelchau, Chao, Dinh, Doddapaneni, Dugan, Gibbs, Hughes, Han, Lee, Murali, Muzny, Qu, Worley, Munoz-Torres, Abouheif, Bonneton, Chen, Chiang, Childers, Cridge, Crumière, Decaras, Didion, Duncan, Elpidina, Favé, Finet, Jacobs, Cheatle Jarvela, Jennings, Jones, Lesoway, Lovegrove, Martynov, Oppert, Lillico-Ouachour, Rajakumar, Refki, Rosendale, Santos, Toubiana, van der Zee, Vargas Jentzsch, Lowman, Viala, Richards and Khila2018). Plants employ diverse defence strategies against phytophagous insects, they have developed various mechanisms of insect resistance specifically targeting Hemiptera to protect themselves (Santamaria et al., Reference Santamaria, Martínez, Cambra, Grbic and Diaz2013). Consequently, this dynamic interplay exerts selective pressure on phytophagous Hemiptera species.

Detoxification of plant secondary metabolites is a crucial adaptation strategy in the evolutionary dynamics of phytophagous insects, enabling them to neutralise or mitigate the effects of these compounds (Seppey et al., Reference Seppey, Ioannidis, Emerson, Pitteloud, Robinson-Rechavi, Roux, Escalona, McKenna, Misof, Shin, Zhou, Waterhouse and Alvarez2019). Several gene families play pivotal roles in this process, including cytochrome P450 monooxygenases, carboxylesterases (CEs), UDP-glycosyltransferases (UGTs), glutathione S-transferases (GSTs), and ATP-binding cassette (ABC) transporters (Voelckel Reference Voelckel2014). These genes are instrumental in the structural adjustment, modification, and transport of toxic compounds. In addition to producing compounds that are resistant to insects, plants also produce protein inhibitors targeting specific mechanisms of insect resistance. Notable gene families associated with insect resistance mechanisms include endopeptidases, such as cysteine (CYSs) and serine (SERs) proteases, along with more specialised enzymes like glycoside hydrolases (GHs) (Pauchet et al., Reference Pauchet, Wilkinson, Chauhan and Ffrench-Constant2010; McKenna et al., Reference McKenna, Scully, Pauchet, Hoover, Kirsch, Geib, Mitchell, Waterhouse, Ahn, Arsala, Benoit, Blackmon, Bledsoe, Bowsher, Busch, Calla, Chao, Childers, Childers, Clarke, Cohen, Demuth, Dinh, Doddapaneni, Dolan, Duan, Dugan, Friedrich, Glastad, Goodisman, Haddad, Han, Hughes, Ioannidis, Johnston, Jones, Kuhn, Lance, Lee, Lee, Lin, Lynch, Moczek, Murali, Muzny, Nelson, Palli, Panfilio, Pers, Poelchau, Quan, Qu, Ray, Rinehart, Robertson, Roehrdanz, Rosendale, Shin, Silva, Torson, Jentzsch, Werren, Worley, Yocum, Zdobnov, Gibbs and Richards2016). Furthermore, chemosensory systems, including the olfactory and gustatory sensors, play a pivotal role in the adaptation of phytophagous insects to their host plants (Goldman-Huertas et al., Reference Goldman-Huertas, Mitchell, Lapoint, Faucher, Hildebrand and Whiteman2015).

During lineage divergence, insects undergo genomic changes that contribute to adaptive development. Identifying these genes and their associations with phenotypic variations in different insect species is crucial for understanding the process of speciation (Hurst, Reference Hurst2009). Genomic changes such as gene duplication events lead to the expansion of gene families, resulting in the emergence of new gene copies with similar or identical functions (Kondrashov, Reference Kondrashov2012). Additionally, other genomic changes, such as point mutations in genes, alter their functions. Branch-specific gene expansion refers to the expansion of gene families in specific lineages and encompasses various mechanisms, including but not limited to adaptive evolution (Innan and Kondrashov, Reference Innan and Kondrashov2010). Despite these mechanisms are not exclusively related to adaptation, the production of different gene copies through duplication provides opportunities for natural selection to operate.

Materials and methods

Data sources and quality assessment

This study encompassed the genomes of 13 Hemiptera species carefully selected to represent five suborders of Hemiptera, ensuring a well-balanced sampling across the suborders of Cimicomorpha, Pentatomomorpha, Gerromorpha, Fulgoromorpha, and Sternorrhyncha. Sequence data of R. prolixus, C. lectularius, and A. pisum (CDS and protein sequences) was obtained from Ensembl Metazoa. Additionally, sequence data of O. fasciatus and R. pedestris (CDS and pep sequences) were acquired from InsectBase. Sequence of G. buenoi was obtained from the I5 K-pilot project, while T. rubrofasciata, L. striatellus, Sogatella furcifera, and Trialeurodes vaporariorum were obtained from the I5 K-pilot and GigaDB databases. Sequence data for the whitefly Mediterranean cryptic species (MED) and whitefly Middle East-Asia Minor 1 cryptic species (MEAM1) was obtained from the Whitefly Genome Database (http://www.whiteflygenomics.org/) and sequence data for Diaphorina citric was obtained from NCBI.

For phylogenetic and divergence time analyses, we introduced the fruit fly Drosophila melanogaster as an outgroup. The genomic data of D.melanogaster were obtained from Ensembl Metazoa. Detailed information about the genomic data sources pertinent to this study can be found in the list provided in supplementary Table S1. The quality assessment of all employed genomic datasets was conducted using BUSCO (version 5.2.2) (Waterhouse et al., Reference Waterhouse, Seppey, Simão, Manni, Ioannidis, Klioutchnikov, Kriventseva and Zdobnov2018) with the arthropoda_odb10.2020-09-10 dataset and the model tran.

Phylogenetic analysis of species

We employed the genomes of the 13 Hemiptera species and one outgroup mentioned earlier, resulting in a total of 14 species, to identify direct homologues. To limit redundancy arising from selective splice variants, we retained gene models encoding the longest protein sequence for each locus. Direct homologous gene analysis was conducted using DIAMOND in conjunction with OrthoFinder (version 2.5.4) (Emms and Kelly, Reference Emms and Kelly2019). Single-copy genes were identified based on the OrthoFinder results and subsequently utilised for downstream analyses, including the construction of phylogenetic trees, estimation of divergence times, and investigation of gene family contraction and expansion.

For the phylogenetic tree construction, a total of 22,723 protein sequences from the single-copy gene families were utilised. MUSCLE (version 3.8.95) (Edgar, Reference Edgar2004) was used to generate multiple sequence alignments with default parameters for the protein sequences within each single-copy family. The resulting alignments were combined into a super alignment matrix, which was then employed for phylogenetic tree reconstruction using the PROTGAMMAJTT model in the RAxML software (Stamatakis, Reference Stamatakis2014).

Functional annotation

The protein Pfam family was identified by InterProScan (Jones et al., Reference Jones, Binns, Chang, Fraser, Li, McAnulla, McWilliam, Maslen, Mitchell, Nuka, Pesseat, Quinn, Sangrador-Vegas, Scheremetjew, Yong, Lopez and Hunter2014). The uniref50 database was annotated using Blastp with an e-value threshold set at 1e-20. GO term annotation was performed using eggnog (Suzek et al., Reference Suzek, Wang, Huang, McGarvey and Wu2015). Orthologous groups (OGs) were preliminarily selected based on the annotation in the uniref50 database. Subsequently, candidate OGs were identified by matching the uniref50 annotation and the Pfam and GO term annotation results.

Analysis of contraction and expansion of gene families

Based on the orthogroups (gene families) information derived from the identified orthologous genes using OrthoFinder (Emms and Kelly, Reference Emms and Kelly2019), the expansion and contraction of orthologous gene families were assessed using CAFE v4.2.1 (De Bie et al., Reference De Bie, Cristianini, Demuth and Hahn2006). This software utilises birth and death processes to model the gene gain and loss events throughout the phylogenetic history process.

Adaptive evolutionary analysis

Furthermore, we conducted adaptive evolutionary selection analyses on the 13 Hemiptera species selected in this study. These analyses aimed to examine the differential occurrence of positive Darwinian selection on specific branches of the phylogeny for individual homologous genes. For this purpose, single-copy orthologous genes generated by OrthoFinder were employed for selective pressure analysis. Multiple sequence comparisons were performed using MACSE (Ranwez et al., Reference Ranwez, Douzery, Cambon, Chantret and Delsuc2018), and subsequent analyses utilised the branching site model of the codeml program within the PAMLpackage (https://github.com/Hua-CM/BatchPAML) (Yang, Reference Yang2007).

To identify positively selected genes, we employed Bonferroni correction and Benjamini-Hochberg false discovery rate (FDR) control to select genes with FDR-corrected p-values below 0.05. We performed comparative sequence recombination checks using the Pairwise Homoplasy Index (PHI) (Bruen et al., Reference Bruen, Philippe and Bryant2006), Neighbour Similarity Score (NSS), and Maximum Chi-Square tests implemented in the PhiPack program to exclude false-positive selection events resulting from sequence recombination. Recombination events were considered present when the p-value (q-value) for PHI was less than 0.05 and when at least one other test supported the occurrence of recombination. In this study, we systematically examined the branches of Pentatomomorpha, Fulgoromorpha, Sternorrhyncha, and Cimicomorpha for adaptive evolution using the aforementioned process. Enrichment analysis was conducted to identify genes associated with adaptive evolution, following the procedure described earlier. We utilised the web-based tool agriGO (systemsbiology.cau.edu.cn/agriGOv2) (Tian et al., Reference Tian, Liu, Yan, You, Yi, Du, Xu and Su2017) and the FlyBase database to assess the enrichment of genes in specific Gene Ontology (GO) terms (Thurmond et al., Reference Thurmond, Goodman, Strelets, Attrill, Gramates, Marygold, Matthews, Millburn, Antonazzo, Trovisco, Kaufman and Calvi2019). Furthermore, the KOBAS (Xie et al., Reference Xie, Mao, Huang, Ding, Wu, Dong, Kong, Gao, Li and Wei2011) and BlastKOALA software (Kanehisa et al., Reference Kanehisa, Sato and Morishima2016) were employed to statistically evaluate the enrichment of genes in the Kyoto Encyclopedia of Genes and Genomes (KEGG) pathways (Kanehisa and Goto, Reference Kanehisa and Goto2000).

Results

Representative sampling of Hemiptera insects

To investigate the adaptation of Hemiptera to host plants, we conducted a sampling of 13 Hemiptera species, representing various suborders. Specifically, our sampling included five species from the Sternorrhyncha suborder, three species from the Cimicomorpha suborder, two species from the Pentatomomorpha suborder, two species from the Fulgoromorpha suborder, and one species from the Gerromorpha suborder. These species inhabited different habitats, with G. buenoi being semi-aquatic and the remaining species terrestrial (Table S1). The feeding habits of these species varied and consisted of three categories: phytophagous, predaceous, and haematophagous. Notably, within the Cimicomorpha suborder alone, all three feeding habits were represented, among which T. rubrofasciata and R. prolixus were predaceous, C. lectularius being haematophagous. O. fasciatus and R. pedestris were phytophagous (Table S1). Therefore, our sample encompassed different Hemiptera suborders, diverse habitats, and a range of feeding habits, including a substantial number of phytophagous species (Table S1). We assessed the integrity of the genome sequences obtained from the sampled species using BUSCO analysis, which revealed sequence integrity ranging from 74.8% to 99.9% across the sampled Hemiptera species and outgroups. With these sequences, we identified 22,723 directly homologous genes and constructed a time-calibrated species phylogeny using 221 single-copy gene sequences. The sequences from these orthologous groups (OGs) were further subjected to functional annotation, particularly focusing on gene families involved in phytophagous insect-plant interactions. Among these candidate OGs, we identified 72 candidates from eight gene families, including 11 UDP-glycosyl transferases (UGTs), 61 cytochrome P450s (P450s), 13 carboxylesterases (CEs), 1 glutathione S-transferase (GST), 4 serine proteases (SERs), 2 cysteineproteases (CYSs), 21 ATP-binding cassette transporters (ABCs), and 3 glycoside hydrolases (GHs) (Table 1).

Table 1. Candidate gene categories and key terms/identifiers are used to select from the complete annotation sequences using InterProScan

InterProScan Scan-derived categories for Inclusion in OGs are considered for inclusion in candidate (OGs), OGs are required to have at least one sequence matching both UniRef and InterProScan entries. Additionally, supplementary Gene Ontology terms are taken into account.

Phytophagous insects show more frequent gene expansion

According to the CAFE analysis, the predicted λ (gain) value for gene expansions was 0.0020 gain/gene/million years, and the μ (loss) value for gene contractions was 0.0036 loss/gen e/million years (De Bie et al., Reference De Bie, Cristianini, Demuth and Hahn2006). Out of the total 22,723 orthologous groups (OGs) examined, the root of the phytophagous insect suborder exhibited a higher frequency of genes. In the root of the Sternorrhyncha suborder, 21 genes underwent expansion, while 5 genes were contracted. Similarly, within the Fulgoromorpha suborder, 64 genes experienced expansion, while 18 genes underwent contraction. In the Cimicomorpha suborder, which includes two predaceous and one haematophagous insect species, all three species are non-phytophagous, and the root exhibited expansion of 3 genes and contraction of 39 genes. Conversely, in the Pentatomomorpha suborder, which comprises two phytophagous insect species, the root showed an expansion of 82 genes and a contraction of 9 genes. The number of gene contractions and expansions in each branch is depicted in fig. 1. Significance analysis using a p-value threshold of <0.01 revealed that 11.11% of the 2525 OGs identified by CAFE exhibited significant contractions or expansions.

Figure 1. The research quantitatively analyses phylogenetics and gene family expansions using genomic datasets. The species tree is based on 221 single-copy orthologous genes, with strong bootstrap support (99% or 100%). Data are from sequenced genomes. (OGs) expand (+) and contract (-) at each subsequence root. The study included herbivorous, semi-aquatic, and blood-feeding insects.

Adaptive evolution is more widespread in the phytophagous branch Pentatomomorpha

We specifically examined the phytophagous insect branches of Pentatomomorpha, Fulgoromorpha, and Sternorrhyncha for adaptive evolution, as well as the non-phytophagous insect branch of Cimicomorpha. The results respectively a total of 18, 3, 1, and 3 genes undergoing adaptive evolution in the branches of Pentatomomorpha, Fulgoromorpha, Sternorrhyncha, and Cimicomorpha (FDR < 0.05) (Table 2). Among these genes, one member of the ABC gene family, ACYPI010103, exhibited adaptive evolution in the Pentatomomorpha branch with FDR value 7.6046E-08 (Table 2). The GO functional enrichment analysis of these adaptively evolved genes revealed that these genes were enriched in 28 different GO terms such as skeletal muscle myosin thick filament assembly, myosin filament assembly, and striated muscle myosin thick filament assembly (Table S2). In terms of KEGG functional enrichment, the adaptively evolved genes were found to be enriched in GTP-binding proteins (Table S3).

Table 2. Result of evolutionary selection pressure analysis on Pentatomomorpha, Fulgoromorpha, Sternorrhyncha, and Cimicomorpha branches

In contrast, no significantly enriched GO or KEGG terms were observed for the adaptively evolved genes in the Fulgoromorpha and Sternorrhyncha branches (P < 0.05). For the non-phytophagous Cimicomorpha branch, the three adaptively evolved genes were found to be enriched in a single GO term: brown fat cell differentiation (Table S2). However, no significantly enriched pathways were detected when analysing KEGG enrichment (Table S3).

Discussion

In this study, we employed comparative genomics to investigate the genomic characteristics, including gene family contraction, expansion and adaptive evolution, across different feeding branches of the Hemiptera suborder. We analysed whitefly MED cryptic species, whitefly MEAM cryptic species, and T. vaporariorum from the polyphagous Sternorrhyncha suborder, as well as A. pisum from the monophagous suborder. In addition, O. fasciatus and R. pedestris, belonging to the Pentatomomorpha suborder, were also monophagous. R. pedestris, commonly known as the bean bug, primarily feeds on leguminous plants, particularly G. max which is commonly referred to as soybean (Huang et al., Reference Huang, Ye, Ye, Yan, Wang, Wei, Chen, Li, Sun and Zhang2021). On the other hand, O. fasciatus, also known as the large milkweed bug, is a specialised seed feeder and has been observed to in 13 Hemiptera species (Panfilio et al., Reference Panfilio, Vargas Jentzsch, Benoit, Erezyilmaz, Suzuki, Colella, Robertson, Poelchau, Waterhouse, Ioannidis, Weirauch, Hughes, Murali, Werren, Jacobs, Duncan, Armisén, Vreede, Baa-Puyoulet, Berger, Chang, Chao, Chen, Chen, Childers, Chipman, Cridge, Crumière, Dearden, Didion, Dinh, Doddapaneni, Dolan, Dugan, Extavour, Febvay, Friedrich, Ginzburg, Han, Heger, Holmes, Horn, Hsiao, Jennings, Johnston, Jones, Jones, Khila, Koelzer, Kovacova, Leask, Lee, Lee, Lovegrove, Lu, Lu, Moore, Munoz-Torres, Muzny, Palli, Parisot, Pick, Porter, Qu, Refki, Richter, Rivera-Pomar, Rosendale, Roth, Sachs, Santos, Seibert, Sghaier, Shukla, Stancliffe, Tidswell, Traverso, van der Zee, Viala, Worley, Zdobnov, Gibbs and Richards2019). Among these, three are predaceous: T. rubrofasciata and R. prolixus from the Cimicomorpha suborder, and G. buenoi from the Gerromorpha suborder (Armisén et al., Reference Armisén, Rajakumar, Friedrich, Benoit, Robertson, Panfilio, Ahn, Poelchau, Chao, Dinh, Doddapaneni, Dugan, Gibbs, Hughes, Han, Lee, Murali, Muzny, Qu, Worley, Munoz-Torres, Abouheif, Bonneton, Chen, Chiang, Childers, Cridge, Crumière, Decaras, Didion, Duncan, Elpidina, Favé, Finet, Jacobs, Cheatle Jarvela, Jennings, Jones, Lesoway, Lovegrove, Martynov, Oppert, Lillico-Ouachour, Rajakumar, Refki, Rosendale, Santos, Toubiana, van der Zee, Vargas Jentzsch, Lowman, Viala, Richards and Khila2018; Liu et al., Reference Liu, Guo, Zhang, Hu, Li, Zhu, Zhou, Wu, Chen and Zhou2019; Huang et al., Reference Huang, Ye, Ye, Yan, Wang, Wei, Chen, Li, Sun and Zhang2021). There is also a haematophagous species, C. lectularius, which belongs to the Cimicomorpha suborder. The remaining nine species were phytophagous, consisting of L. striatellus and S. furcifera from the monophagous Fulgoromorpha suborder, which generally undergo nymphal development on a limited number of host species within the Asclepias genus. To test the hypothesis that gene expansion and adaptive evolution contribute to dietary changes in Hemiptera, we analysed gene family expansions and adaptive evolutionary events in the 13 Hemiptera species. It is widely recognised that phytophagous insects require specific genes for the detoxification of plant secondary metabolites and digestion of plant tissues on their host plants, to achieve the purpose of adapting to different environments comparing to predaceous and haematophagous insect (Heidel-Fischer and Vogel, Reference Heidel-Fischer and Vogel2015; Simon et al., Reference Simon, d'Alençon, Guy, Jacquin-Joly, Jaquiéry, Nouhaud, Peccoud, Sugio and Streiff2015). Both gene family expansion and adaptive evolution are important molecular mechanisms underlying organismal adaptation to varying environments (Innan and Kondrashov, Reference Innan and Kondrashov2010; Kondrashov, Reference Kondrashov2012). Previous studies on beetles have demonstrated the significance of gene family expansion and adaptive evolution in the dietary shift to phytophagy (Seppey et al., Reference Seppey, Ioannidis, Emerson, Pitteloud, Robinson-Rechavi, Roux, Escalona, McKenna, Misof, Shin, Zhou, Waterhouse and Alvarez2019). In our study, we observed gene family expansion in the phytophagous insect branch. However, regarding adaptive evolution, a consistent association between adaptive evolution and gene family expansion was only detected in the Pentatomomorpha suborder. Therefore, adaptive evolution may only play a minor role in the intrinsic mechanism of gene family expansion in Hemiptera. Our analysis identified the phytophagous trophic niche as a driving force behind gene family expansion in Hemiptera. However, the expansion of these genes did not lead to functional divergence (Innan and Kondrashov, Reference Innan and Kondrashov2010; Kondrashov, Reference Kondrashov2012).

There is a shared anatomical feature of a specialised piercing and sucking mouthpart within the Hemiptera insects. However, these insects have undergone diversification to exploit a wide range of food sources, including seeds and plant tissues (phytophagy), and even vertebrate blood (haematophagy) (Li et al., Reference Li, Leavengood, Chapman, Burkhardt, Song, Jiang, Liu, Zhou and Cai2017). As a result of this diversification, numerous hemipterans have become significant agricultural pests or vectors of human diseases. Consequently, extensive efforts have been devoted to genome sequencing in these species, aiming to unravel their genomic composition and elucidate their complex biological characteristics (Li et al., Reference Li, Leavengood, Chapman, Burkhardt, Song, Jiang, Liu, Zhou and Cai2017; Johnson et al., Reference Johnson, Dietrich, Friedrich, Beutel, Wipfler, Peters, Allen, Petersen, Donath, Walden, Kozlov, Podsiadlowski, Mayer, Meusemann, Vasilikopoulos, Waterhouse, Cameron, Weirauch, Swanson, Percy, Hardy, Terry, Liu, Zhou, Misof, Robertson and Yoshizawa2018).

When conducting comparative analyses of datasets from different species, it is crucial to ensure that all analysed species have comparable gene contents. To mitigate the risk of false-positive gene identification, we implemented an additional filtering step based on the method proposed by Seppey et al. (Reference Seppey, Ioannidis, Emerson, Pitteloud, Robinson-Rechavi, Roux, Escalona, McKenna, Misof, Shin, Zhou, Waterhouse and Alvarez2019). This process combined InterProScan results using Pfam or InterPro identifiers, along with Gene Ontology information. Additionally, we integrated the relevant keywords from the carefully selected UniRef dataset for further filtration (Table 1). Although this stringent filtering approach might have potentially excluded some candidate orthogroups (OGs) from our analysis, it was crucial in ensuring the accuracy and validity of the OGs used in this study.

In addition to the gene expansion in Pentatomomorpha, the positive selection analyses also identified a higher number of genes in the branch that underwent positive selection for adaptive evolution. A noteworthy gene in this context is the ABC transporter, which is known to be involved in the detoxification of plant secondary metabolites (Yazaki, Reference Yazaki2006). This finding suggests that selective pressure has played a role in the expansione expansion of the gene family associated with detoxification enzymes (Yazaki, Reference Yazaki2006; Calla, Reference Calla2021).

Emerging research indicates that ABC transporters are also involved in insect detoxification of pesticides, Bt and plant secondary metabolites (Yazaki, Reference Yazaki2006; Heckel, Reference Heckel2012; Merzendorfer, Reference Merzendorfer and Cohen2014). In addition to the ABC transporter, several other adaptively evolved genes in the Pentatomomorpha branch are associated with muscle development, including skeletal muscle myosin thick filament assembly, myosin filament assembly, and muscle cell development, and others (Table 3). The adaptive evolution of these genes provides a mechanistic foundation for the enhanced locomotor capacity of Pentatomomorpha. Several mechanisms contribute to gene family expansion, with adaptive evolution being one of the driving factors, but it may not be the predominant factor (Innan and Kondrashov, Reference Innan and Kondrashov2010; Seppey et al., Reference Seppey, Ioannidis, Emerson, Pitteloud, Robinson-Rechavi, Roux, Escalona, McKenna, Misof, Shin, Zhou, Waterhouse and Alvarez2019). Although a significant number of gene family expansions and adaptive evolutions were observed in the Pentatomomorpha branch, we did not find the same level of cooperativity of gene family expansions and adaptive evolutions in Hemiptera as observed in Coleoptera (Seppey et al., Reference Seppey, Ioannidis, Emerson, Pitteloud, Robinson-Rechavi, Roux, Escalona, McKenna, Misof, Shin, Zhou, Waterhouse and Alvarez2019). Therefore, for gene family expansion, neutral or purifying selection forces may play a more prominent role.

Table 3. Candidate OGs with CAFE p-value <0.01 were identified

OGs were categorised as Cytochrome P450 (P450), Carboxylesterase (CE), Glutathione S-Transferase (GST), and Cysteine Protease (CY5) based on OrthoDB v8 functional annotations.

Neutral or purifying selection is likely the primary mode of occurrence for gene family expansions in Hemiptera. It can be thought that genes associated with nutrition and detoxification might result from gene duplication rather than the emergence of novel functions (Panfilio et al., Reference Panfilio, Vargas Jentzsch, Benoit, Erezyilmaz, Suzuki, Colella, Robertson, Poelchau, Waterhouse, Ioannidis, Weirauch, Hughes, Murali, Werren, Jacobs, Duncan, Armisén, Vreede, Baa-Puyoulet, Berger, Chang, Chao, Chen, Chen, Childers, Chipman, Cridge, Crumière, Dearden, Didion, Dinh, Doddapaneni, Dolan, Dugan, Extavour, Febvay, Friedrich, Ginzburg, Han, Heger, Holmes, Horn, Hsiao, Jennings, Johnston, Jones, Jones, Khila, Koelzer, Kovacova, Leask, Lee, Lee, Lovegrove, Lu, Lu, Moore, Munoz-Torres, Muzny, Palli, Parisot, Pick, Porter, Qu, Refki, Richter, Rivera-Pomar, Rosendale, Roth, Sachs, Santos, Seibert, Sghaier, Shukla, Stancliffe, Tidswell, Traverso, van der Zee, Viala, Worley, Zdobnov, Gibbs and Richards2019; Volonté et al., Reference Volonté, Traverso, Estivalis, Almeida and Ons2022).

Conclusions

In this study, we analysed branch-specific gene family expansions and adaptive evolutions by comparing genomic data from representative species of Hemiptera. Our study revealed a higher occurrence rate of gene family expansions in the phytophagous branch, particularly within the Pentatomomorpha branch. Furthermore, through further analysis of adaptive evolutions within each branch, we identified a greater abundance of genes that underwent adaptive evolutions in the Pentatomomorpha branch.

Supplementary material

The supplementary material for this article can be found at https://doi.org/10.1017/S0007485323000597.

Acknowledgements

This research was funded by National Key Research and Development Plan (grant number 2023YFE0104800); the Program of Talent Introduction in Guizhou University(2023-28), and Lure Research and Application Technology of Agricultural Sucking Pests in Ningxia of China, Grant/Award Number: NGSB-2021-10-02. Thanks for the computing support of the State Key Laboratory of Public Big Data, Guizhou University.

Competing interests

None.

References

Armisén, D, Rajakumar, R, Friedrich, M, Benoit, JB, Robertson, HM, Panfilio, KA, Ahn, S-J, Poelchau, MF, Chao, H, Dinh, H, Doddapaneni, HV, Dugan, S, Gibbs, RA, Hughes, DST, Han, Y, Lee, SL, Murali, SC, Muzny, DM, Qu, J, Worley, KC, Munoz-Torres, M, Abouheif, E, Bonneton, F, Chen, T, Chiang, L-M, Childers, CP, Cridge, AG, Crumière, AJJ, Decaras, A, Didion, EM, Duncan, EJ, Elpidina, EN, Favé, M-J, Finet, C, Jacobs, CGC, Cheatle Jarvela, AM, Jennings, EC, Jones, JW, Lesoway, MP, Lovegrove, MR, Martynov, A, Oppert, B, Lillico-Ouachour, A, Rajakumar, A, Refki, PN, Rosendale, AJ, Santos, ME, Toubiana, W, van der Zee, M, Vargas Jentzsch, IM, Lowman, AV, Viala, S, Richards, S and Khila, A (2018) The genome of the water strider Gerris buenoi reveals expansions of gene repertoires associated with adaptations to life on the water. BMC Genomics 19, 832. https://doi.org/10.1186/s12864-018-5163-2.Google Scholar
Bruen, TC, Philippe, H and Bryant, D (2006) A simple and robust statistical test for detecting the presence of recombination. Genetics 172, 26652681.CrossRefGoogle ScholarPubMed
Calla, B (2021) Signatures of selection and evolutionary relevance of cytochrome P450s in plant-insect interactions. Current Opinion in Insect Science 43, 9296.CrossRefGoogle ScholarPubMed
De Bie, T, Cristianini, N, Demuth, JP and Hahn, MW (2006) CAFE: a computational tool for the study of gene family evolution. Bioinformatics (Oxford, England) 22, 12691271.Google Scholar
Edgar, RC (2004) MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Research 32, 17921797.CrossRefGoogle ScholarPubMed
Emms, DM and Kelly, S (2019) OrthoFinder: phylogenetic orthology inference for comparative genomics. Genome Biology 20, 238.CrossRefGoogle ScholarPubMed
Goldman-Huertas, B, Mitchell, RF, Lapoint, RT, Faucher, CP, Hildebrand, JG and Whiteman, NK (2015) Evolution of herbivory in Drosophilidae linked to loss of behaviors, antennal responses, odorant receptors, and ancestral diet. Proceedings of the National Academy of Sciences of the United States of America 112, 30263031.CrossRefGoogle ScholarPubMed
Heckel, DG (2012) Learning the ABCs of Bt: ABC transporters and insect resistance to Bacillus thuringiensis provide clues to a crucial step in toxin mode of action. Pesticide Biochemistry and Physiology 104, 103110.CrossRefGoogle Scholar
Heidel-Fischer, HM and Vogel, H (2015) Molecular mechanisms of insect adaptation to plant secondary compounds. Current Opinion in insect Science 8, 814.CrossRefGoogle ScholarPubMed
Huang, HJ, Ye, YX, Ye, ZX, Yan, XT, Wang, X, Wei, ZY, Chen, JP, Li, JM, Sun, ZT and Zhang, CX (2021) Chromosome-level genome assembly of the bean bug Riptortus pedestris. Molecular Ecology Resources 21, 24232436.CrossRefGoogle ScholarPubMed
Hurst, LD (2009) Fundamental concepts in genetics: genetics and the understanding of selection. Nature Reviews Genetics 10, 8393.CrossRefGoogle ScholarPubMed
Innan, H and Kondrashov, F (2010) The evolution of gene duplications: classifying and distinguishing between models. Nature Reviews Genetics 11, 97108.CrossRefGoogle ScholarPubMed
Johnson, KP, Dietrich, CH, Friedrich, F, Beutel, RG, Wipfler, B, Peters, RS, Allen, JM, Petersen, M, Donath, A, Walden, KKO, Kozlov, AM, Podsiadlowski, L, Mayer, C, Meusemann, K, Vasilikopoulos, A, Waterhouse, RM, Cameron, SL, Weirauch, C, Swanson, DR, Percy, DM, Hardy, NB, Terry, I, Liu, S, Zhou, X, Misof, B, Robertson, HM and Yoshizawa, K (2018) Phylogenomics and the evolution of hemipteroid insects. Proceedings of the National Academy of Sciences of the United States of America 115, 1277512780.CrossRefGoogle ScholarPubMed
Jones, P, Binns, D, Chang, HY, Fraser, M, Li, W, McAnulla, C, McWilliam, H, Maslen, J, Mitchell, A, Nuka, G, Pesseat, S, Quinn, AF, Sangrador-Vegas, A, Scheremetjew, M, Yong, SY, Lopez, R and Hunter, S (2014) InterProScan 5: genome-scale protein function classification. Bioinformatics (Oxford, England) 30, 12361240.Google ScholarPubMed
Kanehisa, M and Goto, S (2000) KEGG: Kyoto encyclopedia of genes and genomes. Nucleic Acids Research 28, 2730.CrossRefGoogle ScholarPubMed
Kanehisa, M, Sato, Y and Morishima, K (2016) BlastKOALA and GhostKOALA: KEGG tools for functional characterization of genome and metagenome sequences. Journal of Molecular Biology 428, 726731.CrossRefGoogle ScholarPubMed
Kondrashov, FA (2012) Gene duplication as a mechanism of genomic adaptation to a changing environment. Proceedings. Biological Sciences 279, 50485057.Google ScholarPubMed
Li, H, Leavengood, JM Jr., Chapman, EG, Burkhardt, D, Song, F, Jiang, P, Liu, J, Zhou, X and Cai, W (2017) Mitochondrial phylogenomics of Hemiptera reveals adaptive innovations driving the diversification of true bugs. Proceedings. Biological Sciences 284, 20171223. https://doi.org/10.1098/rspb.2017.1223.Google Scholar
Liu, Q, Guo, Y, Zhang, Y, Hu, W, Li, Y, Zhu, D, Zhou, Z, Wu, J, Chen, N and Zhou, XN (2019) A chromosomal-level genome assembly for the insect vector for Chagas disease, Triatoma rubrofasciata. Gigascience 8, giz089. https://doi.org/10.1093/gigascience/giz089.CrossRefGoogle ScholarPubMed
McKenna, DD, Scully, ED, Pauchet, Y, Hoover, K, Kirsch, R, Geib, SM, Mitchell, RF, Waterhouse, RM, Ahn, SJ, Arsala, D, Benoit, JB, Blackmon, H, Bledsoe, T, Bowsher, JH, Busch, A, Calla, B, Chao, H, Childers, AK, Childers, C, Clarke, DJ, Cohen, L, Demuth, JP, Dinh, H, Doddapaneni, H, Dolan, A, Duan, JJ, Dugan, S, Friedrich, M, Glastad, KM, Goodisman, MA, Haddad, S, Han, Y, Hughes, DS, Ioannidis, P, Johnston, JS, Jones, JW, Kuhn, LA, Lance, DR, Lee, CY, Lee, SL, Lin, H, Lynch, JA, Moczek, AP, Murali, SC, Muzny, DM, Nelson, DR, Palli, SR, Panfilio, KA, Pers, D, Poelchau, MF, Quan, H, Qu, J, Ray, AM, Rinehart, JP, Robertson, HM, Roehrdanz, R, Rosendale, AJ, Shin, S, Silva, C, Torson, AS, Jentzsch, IM, Werren, JH, Worley, KC, Yocum, G, Zdobnov, EM, Gibbs, RA and Richards, S (2016) Genome of the Asian longhorned beetle (Anoplophora glabripennis), a globally significant invasive species, reveals key functional and evolutionary innovations at the beetle-plant interface. Genome Biology 17, 227.CrossRefGoogle ScholarPubMed
Merzendorfer, H (2014) Chapter One - ABC transporters and their role in protecting insects from pesticides and their metabolites. In Cohen, E (ed.), Advances in Insect Physiology. New York: Academic Press, pp. 172.Google Scholar
Panfilio, KA, Vargas Jentzsch, IM, Benoit, JB, Erezyilmaz, D, Suzuki, Y, Colella, S, Robertson, HM, Poelchau, MF, Waterhouse, RM, Ioannidis, P, Weirauch, MT, Hughes, DST, Murali, SC, Werren, JH, Jacobs, CGC, Duncan, EJ, Armisén, D, Vreede, BMI, Baa-Puyoulet, P, Berger, CS, Chang, CC, Chao, H, Chen, MM, Chen, YT, Childers, CP, Chipman, AD, Cridge, AG, Crumière, AJJ, Dearden, PK, Didion, EM, Dinh, H, Doddapaneni, HV, Dolan, A, Dugan, S, Extavour, CG, Febvay, G, Friedrich, M, Ginzburg, N, Han, Y, Heger, P, Holmes, CJ, Horn, T, Hsiao, YM, Jennings, EC, Johnston, JS, Jones, TE, Jones, JW, Khila, A, Koelzer, S, Kovacova, V, Leask, M, Lee, SL, Lee, CY, Lovegrove, MR, Lu, HL, Lu, Y, Moore, PJ, Munoz-Torres, MC, Muzny, DM, Palli, SR, Parisot, N, Pick, L, Porter, ML, Qu, J, Refki, PN, Richter, R, Rivera-Pomar, R, Rosendale, AJ, Roth, S, Sachs, L, Santos, ME, Seibert, J, Sghaier, E, Shukla, JN, Stancliffe, RJ, Tidswell, O, Traverso, L, van der Zee, M, Viala, S, Worley, KC, Zdobnov, EM, Gibbs, RA and Richards, S (2019) Molecular evolutionary trends and feeding ecology diversification in the Hemiptera, anchored by the milkweed bug genome. Genome Biology 20, 64.CrossRefGoogle ScholarPubMed
Pauchet, Y, Wilkinson, P, Chauhan, R and Ffrench-Constant, RH (2010) Diversity of beetle genes encoding novel plant cell wall degrading enzymes. PLoS One 5, e15635.CrossRefGoogle ScholarPubMed
Ranwez, V, Douzery, EJP, Cambon, C, Chantret, N and Delsuc, F (2018) MACSE v2: toolkit for the alignment of coding sequences accounting for frameshifts and stop codons. Molecular Biology and Evolution 35, 25822584.CrossRefGoogle ScholarPubMed
Rider, DA (1996) The true bugs of the world (Hemiptera: Heteroptera). classification and natural history. Annals of the Entomological Society of America 89, 609610.CrossRefGoogle Scholar
Rijal, JP, Joyce, AL and Gyawaly, S (2021) Biology, ecology, and management of Hemipteran pests in almond orchards in the United States. Journal of Integrated Pest Management 12, 1-14. https://doi.org/10.1093/jipm/pmab018.Google Scholar
Rosenfeld, JA, Reeves, D, Brugler, MR, Narechania, A, Simon, S, Durrett, R, Foox, J, Shianna, K, Schatz, MC, Gandara, J, Afshinnekoo, E, Lam, ET, Hastie, AR, Chan, S, Cao, H, Saghbini, M, Kentsis, A, Planet, PJ, Kholodovych, V, Tessler, M, Baker, R, DeSalle, R, Sorkin, LN, Kolokotronis, S-O, Siddall, ME, Amato, G and Mason, CE (2016) Genome assembly and geospatial phylogenomics of the bed bug Cimex lectularius. Nature Communications 7, 10164. https://doi.org/10.1038/ncomms10164.Google Scholar
Santamaria, ME, Martínez, M, Cambra, I, Grbic, V and Diaz, I (2013) Understanding plant defence responses against herbivore attacks: an essential first step towards the development of sustainable resistance against pests. Transgenic Research 22, 697708.CrossRefGoogle ScholarPubMed
Seppey, M, Ioannidis, P, Emerson, BC, Pitteloud, C, Robinson-Rechavi, M, Roux, J, Escalona, HE, McKenna, DD, Misof, B, Shin, S, Zhou, X, Waterhouse, RM and Alvarez, N (2019) Genomic signatures accompanying the dietary shift to phytophagy in polyphagan beetles. Genome Biology 20, 98.CrossRefGoogle ScholarPubMed
Simon, JC, d'Alençon, E, Guy, E, Jacquin-Joly, E, Jaquiéry, J, Nouhaud, P, Peccoud, J, Sugio, A and Streiff, R (2015) Genomics of adaptation to host-plants in herbivorous insects. Briefings in Functional Genomics 14, 413423.CrossRefGoogle ScholarPubMed
Stamatakis, A (2014) RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics (Oxford, England) 30, 13121313.Google ScholarPubMed
Suzek, BE, Wang, Y, Huang, H, McGarvey, PB and Wu, CH (2015) UniRef clusters: a comprehensive and scalable alternative for improving sequence similarity searches. Bioinformatics (Oxford, England) 31, 926932.Google ScholarPubMed
Sweet, MH (1979) On the original feeding habits of the Hemiptera (Insecta). Annals of the Entomological Society of America 72, 575579.CrossRefGoogle Scholar
Thurmond, J, Goodman, JL, Strelets, VB, Attrill, H, Gramates, LS, Marygold, SJ, Matthews, BB, Millburn, G, Antonazzo, G, Trovisco, V, Kaufman, TC and Calvi, BR (2019) FlyBase 2.0: the next generation. Nucleic Acids Research 47, D759D765.CrossRefGoogle ScholarPubMed
Tian, T, Liu, Y, Yan, H, You, Q, Yi, X, Du, Z, Xu, W and Su, Z (2017) agriGO v2.0: a GO analysis toolkit for the agricultural community, 2017 update. Nucleic Acids Research 45, W122W129.CrossRefGoogle ScholarPubMed
Voelckel, CJG (2014) Insect-plant interactions. Annual Plant Reviews 47.CrossRefGoogle Scholar
Volonté, M, Traverso, L, Estivalis, JML, Almeida, FC and Ons, S (2022) Comparative analysis of detoxification-related gene superfamilies across five Hemipteran species. BMC Genomics 23, 757.CrossRefGoogle ScholarPubMed
Waterhouse, RM, Seppey, M, Simão, FA, Manni, M, Ioannidis, P, Klioutchnikov, G, Kriventseva, EV and Zdobnov, EM (2018) BUSCO applications from quality assessments to gene prediction and phylogenomics. Molecular Biology and Evolution 35, 543548.CrossRefGoogle ScholarPubMed
Weirauch, C, Schuh, RT, Cassis, G and Wheeler, WC (2019) Revisiting habitat and lifestyle transitions in Heteroptera (Insecta: Hemiptera): insights from a combined morphological and molecular phylogeny. Cladistics 35, 67105.CrossRefGoogle ScholarPubMed
Xie, C, Mao, X, Huang, J, Ding, Y, Wu, J, Dong, S, Kong, L, Gao, G, Li, CY and Wei, L (2011) KOBAS 2.0: a web server for annotation and identification of enriched pathways and diseases. Nucleic Acids Research 39, W316W322.CrossRefGoogle ScholarPubMed
Yang, Z (2007) PAML 4: phylogenetic analysis by maximum likelihood. Molecular Biology and Evolution 24, 15861591.CrossRefGoogle ScholarPubMed
Yazaki, K (2006) ABC transporters involved in the transport of plant secondary metabolites. FEBS Letters 580, 11831191.CrossRefGoogle ScholarPubMed
Figure 0

Table 1. Candidate gene categories and key terms/identifiers are used to select from the complete annotation sequences using InterProScan

Figure 1

Figure 1. The research quantitatively analyses phylogenetics and gene family expansions using genomic datasets. The species tree is based on 221 single-copy orthologous genes, with strong bootstrap support (99% or 100%). Data are from sequenced genomes. (OGs) expand (+) and contract (-) at each subsequence root. The study included herbivorous, semi-aquatic, and blood-feeding insects.

Figure 2

Table 2. Result of evolutionary selection pressure analysis on Pentatomomorpha, Fulgoromorpha, Sternorrhyncha, and Cimicomorpha branches

Figure 3

Table 3. Candidate OGs with CAFE p-value <0.01 were identified

Supplementary material: File

Wu et al. supplementary material

Wu et al. supplementary material
Download Wu et al. supplementary material(File)
File 18.3 KB