ABSTRACT
Bumblebees are among the most abundant and important pollinators for sub-alpine and alpine flowering plant species in the Northern Hemisphere, but little is known about their adaptations to high elevations. In this article, we focused on two bumblebee species, Bombus friseanus and Bombus prshewalskyi, and their respective gut microbiota. The two species, distributed through the Hengduan Mountains of southwestern China, show species replacement at different elevations. We performed genome sequencing based on 20 worker bee samples of each species. Applying evolutionary population genetics and metagenomic approaches, we detected genes under selection and analyzed functional pathways between bumblebees and their gut microbes. We found clear genetic differentiation between the two host species and significant differences in their microbiota. Species replacement occurred in both hosts and their bacteria (Snodgrassella) with an increase in elevation. These extremely high-elevation bumblebees show evidence of positive selection related to diverse biological processes. Positively selected genes involved in host immune systems probably contributed to gut microbiota changes, while the butyrate generated by gut microbiota may influence both host energy metabolism and immune systems. This suggests a close association between the genomes of the host species and their microbiomes based on some degree of natural selection.
IMPORTANCE
Two closely related and dominant bumblebee species, distributed at different elevations through the Hengduan Mountains of southwestern China, showed a clear genomic signature of adaptation to elevation at the molecular level and significant differences in their respective microbiota. Species replacement occurred in both hosts and their bacteria (Snodgrassella) with an increase in elevation. Bumblebees’ adaptations to higher elevations are closely associated with their gut microbiota through two biological processes: energy metabolism and immune response. Information allowing us to understand the adaptive mechanisms of species to extreme conditions is implicit if we are to conserve them as their environments change.
KEYWORDS: high elevation, adaptive selection, Bombus, bumblebee microbiome, gut microbiome, metagenome, population structure
INTRODUCTION
Environments at high elevations are characterized by low temperature, high radiation, and hypoxia, presenting strong selective pressures on their resident fauna (1). Genetic variants, contributing to animals’ adaptations to high elevations, have been found to be present in Tibetans (2), yaks (3), Tibetan pigs (4), and some insects including honeybees (5) and locusts (6). While bumblebees (Apidae: Bombus) remain one of the most important pollinators of wildflowers and crops, few studies focus on their adaptation to high elevations. The most recent research work provides a comparative study focusing on the genomes of 17 Bombus species within 15 subgenera related to high-elevation adaptation (7). However, the authors found no positive selection genes in common within four high-elevation species. This can be attributed to multiple factors. First, each bumblebee species may have a different genetic architecture of adaptation. Second, bumblebees’ adaptation to high elevations may be a polygenic trait, with different loci selected in each species. Besides, phylogenetic analyses on multispecies divergence data identify long-term genetic variations that have been fixed in each species, which may lead to missed detection of recent adaptive signals. Therefore, to understand bumblebees’ high-elevation adaptation strategies, more species need to be investigated using more sensitive methods. Comparatively investigating closely related species that occupy different elevation ranges based on population approaches may be useful in identifying more candidate genes under adaptive selection (8, 9).
Accumulating evidence indicates that gut microbes regulate their host’s physiology and health. In particular, specific gut microbes are now associated with high-elevation adaptations, including the people of Tibet (10, 11) and some of their domesticated mammals (12, 13). However, the composition of the gut microbiota of eusocial honeybees and bumblebees appears simpler compared to that of mammals (14, 15). Just a handful of taxa of bacteria dominate the bumblebee gut floras, including Snodgrassella, Gilliamella, Schmidhempelia, Bifidobacteriaceae, and two clusters within the Lactobacillaceae. Previous studies showed that these bacteria exhibited host specificity, substantial strain-level diversity, differences in genome structure and function, and spatial niche partitioning within individual hosts (15–17). Therefore, it is reasonable to assume that the gut microbiota of high-elevation bumblebees may possess a unique composition and function that may be indicative of their adaptation to certain extreme conditions.
The Hengduan Mountains of southwestern China are classified as a global hotspot (18) and one of several centers of bumblebee diversity. In general, bumblebees remain the most abundant and important pollinator in the Order Hymenoptera for sub-alpine and alpine plant species (19, 20). As the elevation zones increase from 2,000 to 4,000 m, closely related bumblebee species are replaced (20). The most notable replacement occurs between two species in subgenus Melanobombus: Bombus friseanus and Bombus prshewalskyi. These two species shared a common ancestor 4.5 million years ago (18). At these elevations, B. friseanus is distributed from 1,500 to 3,500 m, while B. prshewalskyi is found above the tree line from 3,500 to 5,000 m.
B. friseanus is endemic to the Hengduan Mountains, foraging in subalpine meadows bordering conifer forests, and is abundant between 2,700 and 3,300 m on the Yulong Snow Mountain (20). At 3,200 m, B. friseanus accounted for 68% of all bumblebee species in situ, dominating bumblebee–plant networks (19). In contrast, B. prshewalskyi is distributed from the Hengduan Mountain westward through the Tibetan Plateau and the Eastern Himalayas (18). It is abundant above 4,200 m on the Baima Snow Mountain. Species replacements along elevation gradients are common in montane sites (21, 22), suggesting different adaptations to localized biotic and abiotic factors among closely related species (23). Therefore, we propose that B. friseanus and B. prshewalskyi provide a representative model for understanding the adaptation mechanism of wild bumblebees to extremely high elevations. We hypothesize that, compared to B. friseanus, B. prshewalskyi may have unique characteristics in its genome and gut microbiota, allowing this insect to survive at extremely high elevations.
We collected workers of B. friseanus and B. prshewalskyi and extracted their guts. Deep-coverage metagenomic shotgun sequencing was performed for host and gut microbiomes. We conducted evolutionary population genetic analyses to identify genetic variants related to bumblebee adaptations to extremely high elevations. Comparative metagenomic analyses were applied to provide an overview of the gut microbiota diversity and functional differences. Integration of these results should deepen our understanding of ecophysiological strategies employed by bumblebees to survive in extreme environmental conditions.
RESULTS
Metagenomic sequencing of bumblebees
Deep-coverage metagenomic shotgun sequencing performed on all 20 samples for each species (Fig. 1; Table S1) resulted in an average of 36 Gbp for each sample. Subsequent processing divided each sample datum into two parts (Table S2). One part derived from the bumblebee host showed an average of 28.36 Gbp, and the second part derived from gut microbes had an average of 8.39 Gbp. Among 40 samples, 35 (N = 19 for B. friseanus; N = 16 for B. prshewalskyi) had over 15 Gbp corresponding to the host. The remaining five had 3.45–11.96 Gbp corresponding to the host. All 40 samples had over 1.2 Gbp corresponding to the gut microbiome.
Fig 1.
Sample sites of the two bumblebee species. Workers of B. friseanus were collected from the yellow dot sites on Yulong Snow Mountain and Shika Snow Mountain (3,200 m). Collection sites of B. prshewalskyi are marked with blue dots on Baima Snow Mountain above 4,200 m.
Host population structure and adaptation
Genome alignment indicated an average of 112.5× sequencing coverage for each individual relative to a 225-Mb reference genome of Bombus pyrosoma. The average nucleotide identity (ANI) between B. friseanus and B. prshewalskyi genomes was 99.03%. A total of 112,132 single nucleotide polymorphisms (SNPs) were identified. Reduced genome-wide linkage disequilibrium and genetic diversity (π) were observed in the B. prshewalskyi populations at extremely high elevations (Fig. S1A and B). Based on linkage disequilibrium, the estimated effective population sizes of B. friseanus and B. prshewalskyi populations were 1,065 and 69, respectively. Population genetic analyses showed that B. friseanus and B. prshewalskyi populations formed two branches and divided distinctly in the maximum likelihood tree. This clear division was demonstrated further by genetic structure analysis at K = 2 (Fig. 2A). Principal component analysis (PCA) also supported the clear division of these two species. In PCA, the first principal component (PC) axes explained 11.18% of the total variation separating both bumblebee populations (Fig. S1C).
Fig 2.
Host population differentiation and adaptation to higher (4,200 m) elevations. (A) Maximum likelihood tree and population structure of two bumblebee species. (B) The result of the McDonald–Kreitman test on coding genes of bumblebees. Orange diamonds indicate those genes under positive selection, while gray circles indicate other genes. FN / FS: the fixed nonsynonymous/synonymous ratio between species; PN / PS: the polymorphism nonsynonymous/synonymous ratio within species. (C) Candidate genes under positive selection are involved in diverse gene ontology (GO) processes. Lines indicate that genes are associated with the GO processes.
The McDonald–Kreitman (MK) test identified 23 candidate genes under positive selection from 2,009 genes with sufficient polymorphisms (Fig. 2B; Table S3). Eighteen of them were involved in multiple GO processes (Fig. 2C), ranging from basic to more complicated cellular processes. The basic cellular biological processes included the regulation of DNA transcription (GO:0006351, BRM, ZN236), RNA metabolic process [GO:0006396 (RNA), PNPT1; GO:0010586 (miRNA), LIN41], protein processing (GO:0015031, CUBN; GO:0006457, TRAP1; GO:0018126, OGFD1), actin cytoskeleton organization (GO:0030036, SPTN5), inner dynein arm assembly (GO:0036159, DYH1B), and hydrogen peroxide catabolic process (GO:0042744, PERO). The more complicated processes included response to stress/oxidative stress (GO:0006950, TRAP1; GO:0006979, PERO; GO:0034599, PNPT1), regulation of cellular respiration (GO:0043457, PNPT1; GO:1901856, TRAP1), developmental processes (GO:0035329, GO:0045742, BRM; GO:0008543, LIN41; GO:0045926, PNPT1), immune system processes (GO:0045088, BRM; GO:0008063, SPZ4), cell redox homeostasis (GO:0045454, NHLC2), cilium movement involved in cell motility (GO:0060294, DYH1B), locomotory behavior (GO:0007626, CAD23), oxidation–reduction processes (GO:0055114, FAS, FASC, OGFD1, LAC1), cellular response to glucose stimulus (GO:0071333, Zn236), and the insulin signaling pathway (GO: 0008286, INSR; GO: 0046627, GRB14). In general, all these functions were grouped within six physiological processes, namely, cellular respiration, stress response, growth and development, locomotory behavior, immune response, and energy metabolism.
Microbiota composition and functions
Taxonomic profiles of bumblebee gut microbiomes estimated using the Kraken2 software showed that the core gut bacteria of bumblebees (i.e., Gilliamella, Snodgrassella, Lactobacillus, Bifidobacterium, and Apibacter) dominated most samples, although seven samples were dominated exclusively by pathogens. Specifically, four samples from B. friseanus were dominated by eukaryotic Crithidia sp., and three samples of B. prshewalskyi were dominated by potentially pathogenic bacteria including Pseudomonas sp., Serratia sp., and Rahnella sp. (Fig. 3A).
Fig 3.
Gut communities of two bumblebee species differing in composition and function. (A) Taxonomic profiles of bumblebee gut microbiota estimated using the Kraken2 software. (B) Principal coordinate analysis (PCoA) based on the distance calculated by KmerFreqCalc. The clusters of bumblebee gut microbiotas were significantly associated with host specificity [permutational analysis of variance (PERMANOVA) F1, 29= 14.30, r2 = 0.33, P < 0.01] and related weakly with plant species foraged on by worker bees (PERMANOVA F8, 22= 3.07, r2 = 0.53, P < 0.05). (C) Kyoto Encyclopedia of Genes and Genomes pathways with differential relative abundances between gut microbiomes of B. friseanus and B. prshewalskyi.
PCoA based on the distance calculated by KmerFreqCalc showed distinct clusters of the gut microbiomes from two bumblebee species. The first PC explained 43.5% and the second 34.1% (Fig. 3B). The two clusters were significantly associated with host specificity [permutational analysis of variance (PERMANOVA) F1, 29= 14.30, r2 = 0.33, P < 0.01). In addition, the difference between gut microbiomes showed a weak relation to the plant species foraged on by the worker castes of both species (PERMANOVA F8, 22= 3.07, r2 = 0.53, P < 0.05).
The Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway enrichment analysis identified a total of 72 pathways in all samples. The LEfSe analysis showed pathways with significantly different abundances between the two bumblebee species involved in cell motility, cell community, membrane transport, signal transduction, amino acid metabolism, carbohydrate metabolism, and energy metabolism (Fig. 3C; Table S4). The relatively enriched KEGG pathways in the B. prshewalskyi gut microbiota included the flagellar assembly (ko02040), biofilm formation (ko02026), quorum sensing (ko02024), ATP-binding cassette (ABC) transporters (ko02010), two-component systems (ko02020), butanoate metabolism (ko00650), sulfur relay system (ko04122), and valine/leucine/isoleucine degradation (ko00280). In comparison, the more enriched pathways in the gut of B. friseanus included the bacteria secretion system (ko03070), methane metabolism (ko00680), lysine biosynthesis (ko00300), cysteine/methionine metabolism (ko00270), glycine/serine/threonine metabolism (ko00260), streptomycin biosynthesis (ko00521), aminoacyl-tRNA biosynthesis (ko00970), and base excision repair (ko03410).
Metagenomic binning
Metagenomic binning yielded 23 bins showing >85% completeness and <5% contamination (Table S5). Taxonomic identification classified eight bins representing genera of eusocial bee gut core bacteria. This represented one bin from Gilliamella, two from Snodgrassella, two from Lactobacillus, two from Bifidobacterium, and one from Apibacter. Phylogenetic analysis showed that most of the core bacterial bins clustered together within typical strains of bumblebee gut bacteria, forming distinct clusters with their relatives in the honeybee gut, excluding the bin of Snodgrassella_1 located at the base of the Snodgrassella clade (Fig. S2). Quantitative analysis showed that the relative abundance patterns of four bins (Snodgrassella_1, Snodgrassella_2, Bifidobacterium_2, and Lactobacillus_2) differed significantly in the two bumblebee species (Wilcoxon rank-sum test, P < 0.01) (Fig. 4A; Fig. S3). Notably, Snodgrassella bins showed clear host specificity. Snodgrassella_1 dominated B. friseanus, while Snodgrassella_2 dominated B. prshewalskyi.
Fig 4.
Metagenome binning revealed a pair of vicariously distributed species in Snodgrassella bacteria and the adaptation of Snodgrassella bacteria to the gut environment in a host restricted to a higher elevation. (A) The abundance of metagenome assembly bins in each healthy sample of bumblebees. The bins with significantly different abundances in two bumblebee species are marked with ** (P < 0.01, rank-sum test) or *** (P < 0.001, rank-sum test). (B) Comparison of gene compositions between two Snodgrassella bins using the genome of Snodgrassella alvi_wkB2 as the reference. (C) KEGG pathway enrichment analysis in the unique genes of Snodgrassella bins. (D) Maximum likelihood tree and population structure of two Snodg rassella bins. (E) The result of the McDonald–Kreitman test on genes in Snodgrassella bins. FN / FS: fixed nonsynonymous/synonymous ratio between species; PN / PS: polymorphism nonsynonymous/synonymous ratio within species. Colored diamonds indicate the genes under positive selection ((FN / FS) / (PN / PS) >1 and Fisher’s exact test P < 0.05) in diverse KEGG pathways, while gray circles indicate other genes.
Two Snodgrassella bins
The completeness of Snodgrassella_1 and Snodgrassella_2 bins was 94.77% and 88.82%, respectively. The ANI between the two bins was 77.65%. Phylogenomic analysis indicated that the two Snodgrassella bins differed significantly. Snodgrassella_1 clustered on the base of the Snodgrassella clade, while Snodgrassella_2 clustered with other strains from bumblebees (Fig. S2). Alignment with the genome of Snodgrassella alvi_wkB2 showed that each bin had unique genes (59 for Snodgrassella_1 and 336 for Snodgrassella_2), with 1,377 genes shared (Fig. 4B; Table S6). Five pathways showed enrichment in the 336 Snodgrassella_2 unique genes, including metabolic pathways, microbial metabolism in a diverse environment, butanoate metabolism, oxidative phosphorylation, and sulfur metabolism. No pathways were enriched in the 59 genes unique to Snodgrassella_1 (Fig. 4C). Further analysis indicated that genes involved in butanoate metabolism showed significant differences in the two bins (Fig. S4A). Snodgrassella_2 had twice as many genes involved in the butanoate synthesis pathway as Snodgrassella_1 (Fig. S4B).
To identify the adaptation signals of the two Snodgrassella bins, 15 sets of sequencing reads for each of the Snodgrassella bins were derived from raw data, identifying a total of 95,215 SNPs. Population genetic analyses, including population genetic structure (K = 2) and the maximum likelihood tree, showed a clear division between Snodgrassella_1 and Snodgrassella_2 (Fig. 4D). The MK test identified 22 candidate genes under positive selection from 496 with sufficient polymorphisms (Table S7; Fig. 4E). These genes involved in multiple KEGG pathways, including beta-lactam resistance (ko01501: two genes), two-component systems (ko02020: four genes), ABC transporters (ko02010: two genes), purine metabolism (ko00230: two genes), amino acid metabolism (ko00250: one gene, ko00260: one gene, and ko00400: one gene), riboflavin metabolism (ko00740: one gene), pentose phosphate pathway (ko00030: one gene), and homologous recombination (ko03440: one gene).
DISCUSSION
We focused on two bumblebee species from the same subgenus as well as their gut microbiota, allowing us to hypothesize that genetic diversity in these two prospective host insects occurs along distinct zoogeographic ranges as shaped by adaptive selection. We also expected to detect differences between their gut microbiotas via metagenomic approaches, which may be involved in host genome differences.
We identified 23 genes under positive selection in B. prshewalskyi living at extremely high elevations. Consequently, it is not surprising that the gut microbiotas of both bumblebee species are significantly different in composition and function. Under positive selection, these 23 genes in B. prshewalskyi underlie diverse biological processes, implying that bumblebees adapt to extremely high elevations via complex and systematic mechanisms. Three genes associated with the regulation of cellular respiration (TRAP1 and PNPT1) and response to stress/oxidative stress (TRAP1, PNPT1, and PERO) significantly reflect the adaptation to hypoxia at greater heights. In particular, the positively selected BRM gene indicates that transcriptional regulation plays an important role in maintaining normal development and growth of bumblebees living under harsh conditions (i.e., lower ambient temperatures and hypoxia). A recent study on males of B. terrestris in Britain also showed that environmental pressures most likely contributed to recent changes in genes underlying physiology, neurology, and wing development (9).
Furthermore, BRM together with Spätzle (Spz) suggests ongoing stress on the bumblebees’ immune systems, with BRM regulating the innate immune response and SPZ in the Toll-like signaling pathway. The Toll-like signaling pathway is one of the most important intracellular signaling pathways in innate immunity systems in bees (24, 25) and other insects including Drosophila (26). In humans, acute high-elevation exposure upregulated the inflammatory signaling pathways and may have sensitized the toll-like receptor 4 (TLR4) signaling pathway to subsequent inflammatory stimuli (27). Further evaluation of the expression level and function of BRM or Spz and other immunological tests may provide additional support via future experimentation. Seven genes (FAS, FASC, OGFD1, LAC1, INSR, GRB14, and Zn236) involved in metabolic processes imply strong pressures on energy metabolism in montane bumblebees. Among them, INSR (insulin-like peptide receptor) and GRB14 are mentioned because they are involved in the insulin or insulin-like growth peptide signaling (IIS) pathway, an evolutionarily conservative nutrient-sensing pathway that modulates energy metabolism and development in metazoans (28, 29). In honeybees, the IIS pathway participates in the regulation of caste development (30–32) and the response to cold stress (33). The IIS pathway is also reported to contribute to the adaptive response to hypoxia in other insects including Drosophila (34, 35) and Tibetan locusts (6). Consequently, we propose here that the genetic variants of INSR and GRB14 in B. prshewalskyi play important roles in adapting bumblebee workers to hypoxia at extremely high elevations by regulating energy metabolism.
Additionally, the results suggest a potential correlation between the gut microbiota and those bumblebee adaptations, focusing on butanoate metabolism in the microbiota of B. prshewalskyi and the bins of Snodgrassella_2. Butanoate (or butyrate) is one of the short-chain fatty acids (SCFAs) generated by gut microbes. It can activate G-coupled-receptors directly, inhibit histone deacetylases, and serve as an energy substrate, regulating an animal’s physiology and health (36–38). In humans, butyrate is the primary energy source for colonocytes and also protects against colorectal cancer and inflammation (39, 40). It can improve the host’s metabolism of carbohydrates and lipids by serving as a signal molecule (41). Mice fed with a butyrate-enriched high-fat diet showed increased thermogenesis and energy expenditure and appeared resistant to obesity (42). Though simpler than the human or mice gut microbiota, the honeybee gut microbiota exhibits a variety of beneficial effects, from gut-centric functions like digestion (43), detoxification (44), and defense from pathogens (45), to more peripheral processes like behavior (46). The bumblebee gut microbiota has also been shown to provide some resistance to a trypanosomatid parasite (Crithidia bombi) (47). These effects are thought to be mediated, in part, by SCFAs (43). Considered synthetically, we propose that butyrate concentrations in the guts of bumblebees affect their energy metabolism and gut immunity. As both processes appear to be under stress at extremely high elevations, the results suggest that the gut microbiota may influence their hosts’ adaptations by adjusting the concentration of butyrate. Biochemical details of the impact effect of the gut microbiota on the bumblebee host should be elucidated further by experimentation on the gut butyrate and host physiology of these two bumblebee species or on related model species.
The different gut communities in the two bumblebee species reflect the hosts’ distinctive influence on their own gut microbiota. It is well known that host genetic background and diet are major factors influencing the gut microbiota (48–51). As plant species foraged upon by the bumblebee, samples were shown to correlate weakly with differences in gut microbiotas, and these differences may be mainly due to the hosts’ genetic background changes. The 22 genes involved in multiple KEGG pathways under positive selection in Snodgrassella_2 suggest strong host stress on their signaling pathway and antibiotic resistance process. Snodgrassella bacteria, forming a biofilm on the bumblebees’ ileum wall (15), directly encounter their host’s immune system. A recent study reported that the host specificity of honeybee gut bacteria was determined through reactive oxygen species that are regulated by immune deficiency and Toll pathways (52). Therefore, those host genetic variants involved in the immune system probably contribute to the species replacement of Snodgrassella.
In conclusion, we have helped to show that at extremely high elevations, B. prshewalskyi shows signs of positive selection in diverse biological processes, including cellular respiration, stress response, growth and development, locomotory behavior, immune system, and energy metabolism. We also found differences in both the composition and function of the bumblebee gut microbiome along an elevation gradient. Differences between the host bee species and their gut microbe species, as driven by adaptive selection, are probably functional and suggest a closer association between the genomes of hosts and their microbiomes. As adaptations shared by these bumblebee species with their microbiota indicate some degree of specialization, it may also reflect some degree of coevolution, but this will require further study.
MATERIALS AND METHODS
Bumblebee sampling
We collected workers of B. friseanus and B. prshewalskyi for gut extraction from 11 August to 15 August 2020. Workers of B. friseanus were collected between 2,700 and 3,200 m on Yulong Snow Mountain and the Shika Snow Mountain, northwest Yunnan. Collections of B. prshewalskyi came from 4,200 to 4,700 m on Baima Snow Mountain (Fig. 1, and Table S1 for the detailed information of sampling sites). As bumblebees are eusocial insects, we tried to avoid collecting individuals from the same colony by sampling more than one site on the three mountains for both species. The subsequent population genetic analyses of both species showed a high divergence among individual workers, suggesting that we did collect specimens from more than one colony (Fig. 2A).
We recorded the observation date, time, and plant species on which each foraging worker was collected. Each specimen was netted and stored in a separate 50 mL Eppendorf tube and euthanized at 4°C prior to gut dissection. Specimens were identified as species using morphological characters (18) under field conditions. Gut dissections and their preservation were performed in the field on the day of collection using a Stereo Microscope (Zeiss Stemi 508) as follows.
The surface of each bumblebee was first washed in droplets of 70% ethanol for 30 s and then rinsed with sterile water three times to remove external contaminants. We then pulled out the entire gut from the abdomen terminus using sterilized forceps. Each gut specimen was rinsed twice immediately with 0.9% sterile NaCl solution to maintain cell pressure. Each sample was stored separately in a 2mL cryotube placed in liquid nitrogen. After returning to the laboratory, gut specimens were frozen at –80°C until DNA extraction. In total, we collected 20 gut samples for each Bombus species.
Metagenomic shotgun sequencing and processing
Total genomic DNA was extracted from gut materials using the E.Z.N.A. Soil DNA Kit (Omega Bio-Tek, Norcross, GA, USA). The concentration and purity of extracted DNA were determined with TBS-380 and NanoDrop2000, respectively. DNA extract quality was checked on a 1% agarose gel. The DNA extract was fragmented to an average size of about 400 bp using Covaris M220 (Gene Company Limited, China) for paired-end library construction. The paired-end library was constructed using NEXTFLEX Rapid DNA-Seq (Bioo Scientific, Austin, TX, USA). Adapters containing the full complement of sequencing primer hybridization sites were ligated to the blunt end of fragments. Paired-end sequencing was performed on Illumina NovaSeq 6000 (Illumina Inc., San Diego, CA, USA) at Majorbio Bio-Pharm Technology Co., Ltd. (Shanghai, China) using NovaSeq Reagent Kits, generating an average of 36 gigabyte bases.
Reads were first trimmed off with Trimmomatic v0.39 (parameters used: LEADING:3 TRAILING:3 SLIDINGWINDOW:4:15 MINLEN:40) (53) and inspected with FastQC v0.11.8 (54) to validate the trimming quality. Clean reads of each metagenomic sample were mapped against the reference genome of B. pyrosoma (GCA_14825855.1) using BWA v0.7.17-r1188 (55) to separate reads from the host and its microbes. B. pyrosoma belongs to the same subgenus as the two species, and its genome was annotated using Trinotate (http://trinotate.sourceforge.net/). About 80% of the total reads were derived from the host, and most of the remaining reads belonged to gut microbes. These two datasets were used in the subsequent bumblebee genomic and gut microbiome analyses.
Population genetics and evolutionary analyses
The genome sequences of the two bumblebee species were generated based on the reference genome of B. pyrosoma and the alignment BAM files. The ANI between genomes was calculated using FastANI v.1.1 (56). To obtain high-quality SNPs, the BAM files were further processed as follows. PCR duplicates were removed by SAMtools v1.9 (57). The UnifiedGenotyper method in GATK v4.1.4.0 software (58) was used for SNP calling with default parameters across 40 individuals. To obtain a reliable SNP, we performed a filtering step with the following set of parameters: QD <2.0 || MQ <40.0 || FS >60.0 || SOR > 3.0 || MQRankSum <−12.5 || ReadPosRankSum <−8.0. The resulting VCF file was further filtered using VCFtools v 0.1.17 (59) (parameters: -remove-indels --max-alleles 2 --minDP 4 --minQ 70). We combined these methods to generate final SNPs. The effective population size (Ne) was estimated by NeEstimator (60) using the linkage disequilibrium method. Genome nucleotide diversity (π) was calculated using VCFtools v 0.1.17 (59). The significant difference between the two species was evaluated using the Wilcoxon rank-sum test. The linkage disequilibrium was calculated by PopLDdecay (61).
PCA for the first two components was performed based on the SNP site of the bumblebee in Plink (62). The resulting graph was plotted using the ggplot2 package in R (63).
The population structure was determined using ADMIXTURE v 1.3.0 (64). The estimation was K = 2 to K = 4, and the cross-validation (CV) error estimation was minimized at K = 2. A maximum-likelihood phylogenetic tree was constructed by the RAxML software (65), with the GTRGAMMA model and 100 bootstraps. An ascertainment bias correction was performed to correct for the impact of invariable sites in the data.
The MK test (8) was used to make inferences regarding historical selection by comparing fixed divergence and polymorphism at synonymous and nonsynonymous sites using a 2 × 2 contingency table. Significance was tested using Fisher’s exact test. To correct the false positives, the proportion of amino acid fixations driven by positive selection (α) was calculated following the procedure proposed by Shapiro et al. (66).
Gut microbiome analyses
For each sample, the shotgun reads of the gut microbiome were randomly subsampled to reach the same sequencing depth using Seqtk v1.3-r106 (67) with parameters: -s 100, number = 1687456. Taxonomic profiles were estimated using the Kraken2 software (68) based on a custom comprehensive database, which was built with reference genomes of representatives of the Archaea, Eeubacteria, fungi, and protozoans. Relative abundance was estimated by Bracken (69). If Crithidia or pathogenic bacteria reads ever accounted for >50%, that sample was classified as “unhealthy individuals” and removed from downstream analyses. Similarity between samples was compared using KmerFreqCalc (k = 21) (70). PCoA was performed to illustrate the results. Significant differences between sample clusters across host specificity and diet type (the flower of the plant species on which a bumblebee foraged for nectar and/or pollen) were tested using PERMANOVA (adonis function in the vegan package v2.6-2, R software).
All host-filtered reads were used to build de novo assemblies separately using metaSPAdes v3.13.1 (71) with default parameters. The MetaGeneMark v3.38 (72) gene prediction tool was employed to predict genes, and functional profiling was determined by emapper.py from eggNOG mapper v2.1.3 (73). Gene abundance was quantified with Salmon v0.14.1 (74) using subsampled reads. KEGG pathway enrichment analyses were performed by ClusterProfiler v4.0 (75). The variance between annotated pathways was analyzed using the LEfSe software (76).
Metagenomic binning was performed following the metaWRAP pipeline (77). MetaSPAdes v3.13.1 (71) was used to produce co-assemblies for healthy samples from the same species. The completeness and contamination of each bin were estimated using CheckM v1.1.3 (78). Each bin was assigned taxonomically against the Genome Taxonomy Database with GTDB-tk v1.4.1 (79). Only bins meeting both criteria (completeness ≥80%; contamination ≤5%) were retained to calculate the ANI using FastANI v.1.1 (56). A 99% ANI threshold was set to distinguish strains from the same genus or family. For sibling bins, the one showing higher completeness and lower contamination was selected as one of the core bacterial strains in the bumblebee’s gut. Assemblies of core bacteria from the guts of other species of eusocial bees were downloaded from the NCBI to construct a phylogenetic tree with the bacterial bins using PhyloPhlAn v3.0.60 (80). Relative bin abundance was estimated by Salmon v0.14.1 (74). Rank-sum tests were used to determine whether the differences between bacteria abundance were significant.
Comparative analysis of two Snodgrassella bins
Protein-coding regions were predicted using Prokka v1.14.6 (81). Predicted protein sequences were aligned to the genome of Snodgrassella alvi wkB2 (KEGG: T03063). Pathway enrichment analyses were performed with ClusterProfiler v4.0 (75). Genes involved in the butanoate metabolism pathway (ko00650) were compared between two Snodgrassella bins and highlighted in the KEGG pathway map.
Sequencing reads belonging to each bin were derived from metagenomic shotgun sequencing reads using BWA v0.7.17-r1188 (55) and SAMtools v1.9 (57). We used one of the Snodgrassella bins (Snodgrassella_2) as the reference genome to perform the subsequent admixture analysis and MK test as described previously.
ACKNOWLEDGMENTS
We thank Dr. Peter Bernhardt of Missouri Botanical Garden, USA, for improving an earlier version of this manuscript. We also thank Xiao-Qing Fan of the Kunming Institute of Botany (KIB), CAS, for making field collections.
This work was supported by the Strategic Priority Research Program of the Chinese Academy of Sciences (XDB31000000) and the Yunlin Scholarship of Yunnan Province (YLXL20170001).
Contributor Information
Qi Wu, Email: wuq@im.ac.cn.
Hong Wang, Email: wanghong@mail.kib.ac.cn.
Janet K. Jansson, Pacific Northwest National Laboratory Biological Sciences Division, Richland, USA
DATA AVAILABILITY
Whole metagenome shotgun sequencing datasets are deposited in the National Genomics Data Center (NGDC database) with the BioProject accession number PRJCA012598.
SUPPLEMENTAL MATERIAL
The following material is available online at https://doi.org/10.1128/msystems.01219-23.
Genomic differences of two bumblebee species and the metagenomic analyses results.
Sample and sequencing dataset information, the results of population genetics and evolutionary analyses, and the results of metagenomic analyses.
ASM does not own the copyrights to Supplemental Material that may be linked to, or accessed through, an article. The authors have granted ASM a non-exclusive, world-wide license to publish the Supplemental Material files. Please contact the corresponding author directly for reuse.
REFERENCES
- 1. Cheviron ZA, Brumfield RT. 2012. Genomic insights into adaptation to high-altitude environments. Heredity (Edinb) 108:354–361. doi: 10.1038/hdy.2011.85 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2. Simonson TS, Yang Y, Huff CD, Yun H, Qin G, Witherspoon DJ, Bai Z, Lorenzo FR, Xing J, Jorde LB, Prchal JT, Ge R. 2010. Genetic evidence for high-altitude adaptation in Tibet. Science 329:72–75. doi: 10.1126/science.1189406 [DOI] [PubMed] [Google Scholar]
- 3. Qiu Q, Zhang G, Ma T, Qian W, Wang J, Ye Z, Cao C, Hu Q, Kim J, Larkin DM, et al. 2012. The yak genome and adaptation to life at high altitude. Nat Genet 44:946–949. doi: 10.1038/ng.2343 [DOI] [PubMed] [Google Scholar]
- 4. Li M, Tian S, Jin L, Zhou G, Li Y, Zhang Y, Wang T, Yeung CKL, Chen L, Ma J, et al. 2013. Genomic analyses identify distinct patterns of selection in domesticated pigs and Tibetan wild boars. Nat Genet 45:1431–1438. doi: 10.1038/ng.2811 [DOI] [PubMed] [Google Scholar]
- 5. Montero-Mendieta S, Tan K, Christmas MJ, Olsson A, Vilà C, Wallberg A, Webster MT. 2019. The genomic basis of adaptation to high-altitude habitats in the eastern honey bee (Apis cerana). Mol Ecol 28:746–760. doi: 10.1111/mec.14986 [DOI] [PubMed] [Google Scholar]
- 6. Ding D, Liu G, Hou L, Gui W, Chen B, Kang L. 2018. Genetic variation in PTPN1 contributes to metabolic adaptation to high-altitude hypoxia in Tibetan migratory locusts. Nat Commun 9:4991. doi: 10.1038/s41467-018-07529-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7. Sun C, Huang J, Wang Y, Zhao X, Su L, Thomas GWC, Zhao M, Zhang X, Jungreis I, Kellis M, et al. 2021. Genus-wide characterization of bumblebee genomes provides insights into their evolution and variation in ecological and behavioral traits. Mol Biol Evol 38:486–501. doi: 10.1093/molbev/msaa240 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8. McDonald JH, Kreitman M. 1991. Adaptive protein evolution at the Adh locus in Drosophila. Nature 351:652–654. doi: 10.1038/351652a0 [DOI] [PubMed] [Google Scholar]
- 9. Colgan TJ, Arce AN, Gill RJ, Ramos Rodrigues A, Kanteh A, Duncan EJ, Li L, Chittka L, Wurm Y. 2022. Genomic signatures of recent adaptation in a wild bumblebee. Mol Biol Evol 39:msab366. doi: 10.1093/molbev/msab366 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10. Zeng B, Zhang S, Xu H, Kong F, Yu X, Wang P, Yang M, Li D, Zhang M, Ni Q, Li Y, Fan X, Yang D, Ning R, Zhao J, Li Y. 2020. Gut microbiota of Tibetans and Tibetan pigs varies between high and low altitude environments. Microbiol Res 235:126447. doi: 10.1016/j.micres.2020.126447 [DOI] [PubMed] [Google Scholar]
- 11. Karl JP, Berryman CE, Young AJ, Radcliffe PN, Branck TA, Pantoja-Feliciano IG, Rood JC, Pasiakos SM. 2018. Associations between the gut microbiota and host responses to high altitude. Am J Physiol Gastrointest Liver Physiol 315:G1003–G1015. doi: 10.1152/ajpgi.00253.2018 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12. Zhao F, Yang L, Zhang T, Zhuang D, Wu Q, Yu J, Tian C, Zhang Z. 2023. Gut microbiome signatures of extreme environment adaption in Tibetan pig. NPJ Biofilms Microbiomes 9:27. doi: 10.1038/s41522-023-00395-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13. Zhang Z, Xu D, Wang L, Hao J, Wang J, Zhou X, Wang W, Qiu Q, Huang X, Zhou J, Long R, Zhao F, Shi P. 2016. Convergent evolution of rumen microbiomes in high-altitude mammals. Curr Biol 26:1873–1879. doi: 10.1016/j.cub.2016.05.012 [DOI] [PubMed] [Google Scholar]
- 14. Kwong WK, Moran NA. 2016. Gut microbial communities of social bees. Nat Rev Microbiol 14:374–384. doi: 10.1038/nrmicro.2016.43 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15. Hammer TJ, Le E, Martin AN, Moran NA. 2021. The gut microbiota of bumblebees. Insectes Soc 68:287–301. doi: 10.1007/s00040-021-00837-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16. Kwong WK, Engel P, Koch H, Moran NA. 2014. Genomics and host specialization of honey bee and bumble bee gut symbionts. Proc Natl Acad Sci U S A 111:11509–11514. doi: 10.1073/pnas.1405838111 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17. Li Y, Leonard SP, Powell JE, Moran NA. 2022. Species divergence in gut-restricted bacteria of social bees. Proc Natl Acad Sci U S A 119:e2115013119. doi: 10.1073/pnas.2115013119 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18. Williams PH, Altanchimeg D, Byvaltsev A, De Jonghe R, Jaffar S, Japoshvili G, Kahono S, Liang H, Mei M, Monfared A, Nidup T, Raina R, Ren Z, Thanoosing C, Zhao Y, Orr MC. 2020. Widespread polytypic species or complexes of local species? Revising bumblebees of the subgenus Melanobombus world-wide (Hymenoptera, Apidae, Bombus). Eur J Taxon 719:1–120. doi: 10.5852/ejt.2020.719.1107 [DOI] [Google Scholar]
- 19. Liang H, Zhao Y-H, Rafferty NE, Ren Z-X, Zhong L, Li H-D, Li D-Z, Wang H. 2021. Evolutionary and ecological factors structure a plant–bumblebee network in a biodiversity hotspot, the Himalaya–Hengduan mountains. Funct Ecol 35:2523–2535. doi: 10.1111/1365-2435.13886 [DOI] [Google Scholar]
- 20. Zhao Y-H, Lázaro A, Ren Z-X, Zhou W, Li H-D, Tao Z-B, Xu K, Wu Z-K, Wolfe LM, Li D-Z, Wang H. 2019. The topological differences between visitation and pollen transport networks: a comparison in species rich communities of the Himalaya–Hengduan mountains. Oikos 128:551–562. doi: 10.1111/oik.05262 [DOI] [Google Scholar]
- 21. Freeman BG, Strimas-Mackey M, Miller ET. 2022. Interspecific competition limits bird species' ranges in tropical mountains. Science 377:416–420. doi: 10.1126/science.abl7242 [DOI] [PubMed] [Google Scholar]
- 22. Hedberg O. 1969. Evolution and speciation in a tropical high mountain flora. Biol J Linn Soc 1:135–148. doi: 10.1111/j.1095-8312.1969.tb01816.x [DOI] [Google Scholar]
- 23. Poisot T, Bever JD, Nemri A, Thrall PH, Hochberg ME. 2011. A conceptual framework for the evolution of ecological specialisation. Ecol Lett 14:841–851. doi: 10.1111/j.1461-0248.2011.01645.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24. Evans JD, Aronstein K, Chen YP, Hetru C, Imler JL, Jiang H, Kanost M, Thompson GJ, Zou Z, Hultmark D. 2006. Immune pathways and defence mechanisms in honey bees Apis mellifera. Insect Mol Biol 15:645–656. doi: 10.1111/j.1365-2583.2006.00682.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25. Larsen A, Reynaldi FJ, Guzman-Novoa E. 2019. Fundaments of the immune system of honey bee (Apis mellifera). Revision. Rev Mex Cienc Pecu 10:705–728. doi: 10.22319/rmcp.v10i3.4785 [DOI] [Google Scholar]
- 26. Valanne S, Wang J-H, Rämet M. 2011. The Drosophila toll signaling pathway. J Immunol 186:649–656. doi: 10.4049/jimmunol.1002302 [DOI] [PubMed] [Google Scholar]
- 27. Pham K, Frost S, Parikh K, Puvvula N, Oeung B, Heinrich EC. 2022. Inflammatory gene expression during acute high-altitude exposure. J Physiol 600:4169–4186. doi: 10.1113/JP282772 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28. Li D, Chen X, Zhu F, Chen K. 2020. Insulin-like peptides in model insects. Invert Surviv J 17:186–195. doi: 10.25431/1824-307X/isj.v0i0.186-195 [DOI] [Google Scholar]
- 29. Antonova Y, Arik AJ, Moore W, Riehle MA, Brown MR. 2012. Insulin-like peptides: structure, signaling, and function, p 63–92. In Insect endocrinology. Elsevier. [Google Scholar]
- 30. de Azevedo SV, Hartfelder K. 2008. The insulin signaling pathway in honey bee (Apis mellifera) caste development - differential expression of insulin-like peptides and insulin receptors in queen and worker larvae. J Insect Physiol 54:1064–1071. doi: 10.1016/j.jinsphys.2008.04.009 [DOI] [PubMed] [Google Scholar]
- 31. Wang Y, Azevedo SV, Hartfelder K, Amdam GV. 2013. Insulin-like peptides (AmILP1 and AmILP2) differentially affect female caste development in the honey bee (Apis mellifera L.). J Exp Biol 216:4347–4357. doi: 10.1242/jeb.085779 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32. Wheeler DE, Buck NA, Evans JD. 2014. Expression of insulin/insulin-like signalling and TOR pathway genes in honey bee caste determination. Insect Mol Biol 23:113–121. doi: 10.1111/imb.12065 [DOI] [PubMed] [Google Scholar]
- 33. Ramirez L, Luna F, Mucci CA, Lamattina L. 2021. Fast weight recovery, metabolic rate adjustment and gene-expression regulation define responses of cold-stressed honey bee brood. J Insect Physiol 128:104178. doi: 10.1016/j.jinsphys.2020.104178 [DOI] [PubMed] [Google Scholar]
- 34. Wong DM, Shen Z, Owyang KE, Martinez-Agosto JA. 2014. Insulin-and warts-dependent regulation of tracheal plasticity modulates systemic larval growth during hypoxia in Drosophila melanogaster. PLoS One 9:e115297. doi: 10.1371/journal.pone.0115297 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35. Harrison JF, Haddad GG. 2011. Effects of oxygen on growth and size: synthesis of molecular, organismal, and evolutionary studies with Drosophila melanogaster. Annu Rev Physiol 73:95–113. doi: 10.1146/annurev-physiol-012110-142155 [DOI] [PubMed] [Google Scholar]
- 36. Koh A, De Vadder F, Kovatcheva-Datchary P, Bäckhed F. 2016. From dietary fiber to host physiology: short-chain fatty acids as key bacterial metabolites. Cell 165:1332–1345. doi: 10.1016/j.cell.2016.05.041 [DOI] [PubMed] [Google Scholar]
- 37. Makki K, Deehan EC, Walter J, Bäckhed F. 2018. The impact of dietary fiber on gut microbiota in host health and disease. Cell Host Microbe 23:705–715. doi: 10.1016/j.chom.2018.05.012 [DOI] [PubMed] [Google Scholar]
- 38. van der Hee B, Wells JM. 2021. Microbial regulation of host physiology by short-chain fatty acids. Trends Microbiol 29:700–712. doi: 10.1016/j.tim.2021.02.001 [DOI] [PubMed] [Google Scholar]
- 39. Valdes AM, Walter J, Segal E, Spector TD. 2018. Role of the gut microbiota in nutrition and health. BMJ 361:k2179. doi: 10.1136/bmj.k2179 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40. Flint HJ, Scott KP, Louis P, Duncan SH. 2012. The role of the gut microbiota in nutrition and health. Nat Rev Gastroenterol Hepatol 9:577–589. doi: 10.1038/nrgastro.2012.156 [DOI] [PubMed] [Google Scholar]
- 41. De Vadder F, Kovatcheva-Datchary P, Goncalves D, Vinera J, Zitoun C, Duchampt A, Bäckhed F, Mithieux G. 2014. Microbiota-generated metabolites promote metabolic benefits via gut-brain neural circuits. Cell 156:84–96. doi: 10.1016/j.cell.2013.12.016 [DOI] [PubMed] [Google Scholar]
- 42. Gao Z, Yin J, Zhang J, Ward RE, Martin RJ, Lefevre M, Cefalu WT, Ye J. 2009. Butyrate improves insulin sensitivity and increases energy expenditure in mice. Diabetes 58:1509–1517. doi: 10.2337/db08-1637 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43. Zheng H, Powell JE, Steele MI, Dietrich C, Moran NA. 2017. Honeybee gut microbiota promotes host weight gain via bacterial metabolism and hormonal signaling. Proc Natl Acad Sci U S A 114:4775–4780. doi: 10.1073/pnas.1701819114 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44. Zheng H, Nishida A, Kwong WK, Koch H, Engel P, Steele MI, Moran NA, Maleszka R, McFall-Ngai MJ. 2016. Metabolism of toxic sugars by strains of the bee gut symbiont Gilliamella apicola. mBio 7:e01326-16. doi: 10.1128/mBio.01326-16 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45. Kwong WK, Mancenido AL, Moran NA. 2017. Immune system stimulation by the native gut microbiota of honey bees. R Soc Open Sci 4:170003. doi: 10.1098/rsos.170003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46. Liberti J, Kay T, Quinn A, Kesner L, Frank ET, Cabirol A, Richardson TO, Engel P, Keller L. 2022. The gut microbiota affects the social network of honeybees. Nat Ecol Evol 6:1471–1479. doi: 10.1038/s41559-022-01840-w [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47. Koch H, Schmid-Hempel P. 2011. Socially transmitted gut microbiota protect bumble bees against an intestinal parasite. Proc Natl Acad Sci U S A 108:19288–19292. doi: 10.1073/pnas.1110474108 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48. Muegge BD, Kuczynski J, Knights D, Clemente JC, González A, Fontana L, Henrissat B, Knight R, Gordon JI. 2011. Diet drives convergence in gut microbiome functions across mammalian phylogeny and within humans. Science 332:970–974. doi: 10.1126/science.1198719 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49. Lynch JB, Hsiao EY. 2019. Microbiomes as sources of emergent host phenotypes. Science 365:1405–1409. doi: 10.1126/science.aay0240 [DOI] [PubMed] [Google Scholar]
- 50. Groussin M, Mazel F, Sanders JG, Smillie CS, Lavergne S, Thuiller W, Alm EJ. 2017. Unraveling the processes shaping mammalian gut microbiomes over evolutionary time. Nat Commun 8:14319. doi: 10.1038/ncomms14319 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51. Davenport ER. 2016. Elucidating the role of the host genome in shaping microbiome composition. Gut Microbes 7:178–184. doi: 10.1080/19490976.2016.1155022 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52. Guo L, Tang J, Tang M, Luo S, Zhou X. 2023. Reactive oxygen species are regulated by immune deficiency and toll pathways in determining the host specificity of honeybee gut bacteria. Proc Natl Acad Sci U S A 120:e2219634120. doi: 10.1073/pnas.2219634120 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53. Bolger AM, Lohse M, Usadel B. 2014. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics 30:2114–2120. doi: 10.1093/bioinformatics/btu170 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54. Andrews S. 2010. FastQC: a quality control tool for high throughput sequence data. Babraham Bioinformatics, Babraham Institute, Cambridge, United Kingdom. [Google Scholar]
- 55. Li H, Durbin R. 2010. Fast and accurate long-read alignment with Burrows–Wheeler transform. Bioinformatics 26:589–595. doi: 10.1093/bioinformatics/btp698 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56. Jain C, Rodriguez-R LM, Phillippy AM, Konstantinidis KT, Aluru S. 2018. High throughput ANI analysis of 90K prokaryotic genomes reveals clear species boundaries. Nat Commun 9:5114. doi: 10.1038/s41467-018-07641-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57. Li H, Handsaker B, Wysoker A, Fennell T, Ruan J, Homer N, Marth G, Abecasis G, Durbin R, 1000 Genome Project Data Processing Subgroup . 2009. The sequence alignment/map format and SAMtools. Bioinformatics 25:2078–2079. doi: 10.1093/bioinformatics/btp352 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58. DePristo MA, Banks E, Poplin R, Garimella KV, Maguire JR, Hartl C, Philippakis AA, del Angel G, Rivas MA, Hanna M, McKenna A, Fennell TJ, Kernytsky AM, Sivachenko AY, Cibulskis K, Gabriel SB, Altshuler D, Daly MJ. 2011. A framework for variation discovery and genotyping using next-generation DNA sequencing data. Nat Genet 43:491–498. doi: 10.1038/ng.806 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59. Danecek P, Auton A, Abecasis G, Albers CA, Banks E, DePristo MA, Handsaker RE, Lunter G, Marth GT, Sherry ST, McVean G, Durbin R, 1000 Genomes Project Analysis Group . 2011. The variant call format and VCFtools. Bioinformatics 27:2156–2158. doi: 10.1093/bioinformatics/btr330 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60. Do C, Waples RS, Peel D, Macbeth GM, Tillett BJ, Ovenden JR. 2014. NeEstimator v2: re-implementation of software for the estimation of contemporary effective population size (Ne) from genetic data. Mol Ecol Resour 14:209–214. doi: 10.1111/1755-0998.12157 [DOI] [PubMed] [Google Scholar]
- 61. Zhang C, Dong SS, Xu JY, He WM, Yang TL. 2019. PopLDdecay: a fast and effective tool for linkage disequilibrium decay analysis based on variant call format files. Bioinformatics 35:1786–1788. doi: 10.1093/bioinformatics/bty875 [DOI] [PubMed] [Google Scholar]
- 62. Purcell S, Neale B, Todd-Brown K, Thomas L, Ferreira MAR, Bender D, Maller J, Sklar P, de Bakker PIW, Daly MJ, Sham PC. 2007. PLINK: a tool set for whole-genome association and population-based linkage analyses. Am J Hum Genet 81:559–575. doi: 10.1086/519795 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63. Wickham H. 2016. ggplot2: elegant graphics for data analysis. Springer. [Google Scholar]
- 64. Alexander DH, Novembre J, Lange K. 2009. Fast model-based estimation of ancestry in unrelated individuals. Genome Res 19:1655–1664. doi: 10.1101/gr.094052.109 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65. Silvestro D, Michalak I. 2012. raxmlGUI: a graphical front-end for RAxML. Org Divers Evol 12:335–337. doi: 10.1007/s13127-011-0056-0 [DOI] [Google Scholar]
- 66. Shapiro JA, Huang W, Zhang C, Hubisz MJ, Lu J, Turissini DA, Fang S, Wang HY, Hudson RR, Nielsen R, Chen Z, Wu CI. 2007. Adaptive genic evolution in the Drosophila genomes. Proc Natl Acad Sci U S A 104:2271–2276. doi: 10.1073/pnas.0610385104 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67. Li H. 2013. Seqtk: a fast and lightweight tool for processing FASTA or FASTQ sequences. Github. [Google Scholar]
- 68. Wood DE, Lu J, Langmead B. 2019. Improved metagenomic analysis with Kraken 2. Genome Biol 20:257. doi: 10.1186/s13059-019-1891-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69. Lu J, Breitwieser FP, Thielen P, Salzberg SL. 2017. Bracken: estimating species abundance in metagenomics data. PeerJ Comput Sci 3:e104. doi: 10.7717/peerj-cs.104 [DOI] [Google Scholar]
- 70. Dong J, Liu S, Zhang Y, Dai Y, Wu Q. 2020. A new alignment-free whole metagenome comparison tool and its application on gut microbiomes of wild giant pandas. Front Microbiol 11:1061. doi: 10.3389/fmicb.2020.01061 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71. Nurk S, Meleshko D, Korobeynikov A, Pevzner PA. 2017. metaSPAdes: a new versatile metagenomic assembler. Genome Res 27:824–834. doi: 10.1101/gr.213959.116 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72. Zhu W, Lomsadze A, Borodovsky M. 2010. Ab initio gene identification in metagenomic sequences. Nucleic Acids Res 38:e132–e132. doi: 10.1093/nar/gkq275 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73. Huerta-Cepas J, Forslund K, Coelho LP, Szklarczyk D, Jensen LJ, von Mering C, Bork P. 2017. Fast genome-wide functional annotation through orthology assignment by eggNOG-Mapper. Mol Biol Evol 34:2115–2122. doi: 10.1093/molbev/msx148 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74. Patro R, Duggal G, Love MI, Irizarry RA, Kingsford C. 2017. Salmon provides fast and bias-aware quantification of transcript expression. Nat Methods 14:417–419. doi: 10.1038/nmeth.4197 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75. Wu T, Hu E, Xu S, Chen M, Guo P, Dai Z, Feng T, Zhou L, Tang W, Zhan L, Fu X, Liu S, Bo X, Yu G. 2021. clusterProfiler 4.0: a universal enrichment tool for interpreting omics data. Innovation (Camb) 2:100141. doi: 10.1016/j.xinn.2021.100141 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76. Segata N, Izard J, Waldron L, Gevers D, Miropolsky L, Garrett WS, Huttenhower C. 2011. Metagenomic biomarker discovery and explanation. Genome Biol 12:1–18. doi: 10.1186/gb-2011-12-6-r60 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77. Uritskiy GV, DiRuggiero J, Taylor J. 2018. MetaWRAP-a flexible pipeline for genome-resolved metagenomic data analysis. Microbiome 6:158. doi: 10.1186/s40168-018-0541-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78. Parks DH, Imelfort M, Skennerton CT, Hugenholtz P, Tyson GW. 2015. CheckM: assessing the quality of microbial genomes recovered from isolates, single cells, and metagenomes. Genome Res 25:1043–1055. doi: 10.1101/gr.186072.114 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79. Chaumeil P-A, Mussig AJ, Hugenholtz P, Parks DH, Hancock J. 2020. GTDB-Tk: a toolkit to classify genomes with the genome taxonomy database. Bioinformatics 36:1925–1927. doi: 10.1093/bioinformatics/btz848 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80. Segata N, Börnigen D, Morgan XC, Huttenhower C. 2013. PhyloPhlAn is a new method for improved phylogenetic and taxonomic placement of microbes. Nat Commun 4:1–11. doi: 10.1038/ncomms3304 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81. Seemann T. 2014. Prokka: rapid prokaryotic genome annotation. Bioinformatics 30:2068–2069. doi: 10.1093/bioinformatics/btu153 [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Genomic differences of two bumblebee species and the metagenomic analyses results.
Sample and sequencing dataset information, the results of population genetics and evolutionary analyses, and the results of metagenomic analyses.
Data Availability Statement
Whole metagenome shotgun sequencing datasets are deposited in the National Genomics Data Center (NGDC database) with the BioProject accession number PRJCA012598.




