Abstract

Background

Venom glands play a key role in the predation and defense strategies of almost all spider groups. However, the spider family Uloboridae lacks venom glands and has evolved an adaptive strategy: they excessively wrap their prey directly with spider silk instead of paralyzing it first with toxins. This shift in survival strategy is very fascinating, but the genetic underpinnings behind it are poorly understood.

Results

Spanning multiple spider groups, we conducted multiomics analyses on Octonoba sinensis and described the adaptive evolution of the Uloboridae family at the genome level. We observed the coding genes of myosin and twitchin in muscles are under positive selection, energy metabolism functions are enhanced, and gene families related to tracheal development and tissue mechanical strength are expanded or emerged, all of which are related to the unique anatomical structure and predatory behavior of spiders in the family Uloboridae. In addition, we also scanned the elements that are absent or under relaxed purifying selection, as well as toxin gene homologs in the genomes of 2 species in this family. The results show that the absence of regions and regions under relaxed selection in these spiders’ genomes are concentrated in areas related to development and neurosystem. The search for toxin homologs reveals possible gene function shift between toxins and nontoxins and confirms that there are no reliable toxin genes in the genome of this group.

Conclusions

This study demonstrates the trade-off between different predation strategies in spiders, using either chemical or physical strategy, and provides insights into the possible mechanism underlying this trade-off. Venomless spiders need to mobilize multiple developmental and metabolic pathways related to motor function and limb mechanical strength to cover the decline in adaptability caused by the absence of venom glands.

Introduction

“Venomous” is a common way that people perceive spiders (Araneae). The toxic, painful, and even fatal bite is always frightening. Indeed, almost all spiders are venomous. In the earliest diverging suborder Mesothelae, fangs that could inject venom are already present [1]. Some highly toxic species make spiders even more notorious. As an important means of hunting and defense, the toxin system gives spiders an outstanding advantage in environmental suitability and has allowed them to spread throughout the world. Of course, there are always exceptions, and some outliers are believed to be lacking venom glands. Currently, known spiders without venom grands include Holarchaea (2 species) [2–4] and the entire family Uloboridae, with the latter being the most prosperous group [5]. Compared to Holarchaea, which has a smaller body size, fewer species, and limited distribution, the family Uloboridae provides a satisfactory model for us to study the evolution of an important synapomorphy of spiders and the adaptation strategy brought about by the loss of important functional traits.

Due to the absence of venom glands, the predation methods of uloborids are also relatively specialized and excessive. Generally speaking, using venom to paralyze prey is an effective chemical attack strategy. However, many kinds of spiders integrate both chemical attack and physical attack strategies. Certain species within the Araneoidea family utilize entanglement initially to restrain larger prey before a venomous final strike [6]. Although this predatory tactic may alleviate the selective pressure associated with venom usage, entanglement appears rudimentary in comparison to the prey-wrapping behavior exhibited by uloborids. In Uloboridae, this physical attack as the sole means of attack can span from a few minutes to nearly an hour, with the spider silk utilized sometimes exceeding a hundred meters in length [7–9]. Therefore, considerable physical endurance is indispensable for the successful execution of this predatory tactic. Previous anatomical records indicate uloborids have well-developed trachea [10], and many branches of the trachea extend into the prosoma and appendages [11]. However, there is no relevant research that can link these adaptive characteristics to the venom gland deficiency of this group.

The spider Octonoba sinensis (NCBI:txid198412) belongs to the family Uloboridae. Their body size is relatively larger than other members of the family [12], and this species is widely distributed in East Asia, Southeast Asia, and North America [2]. Their habitat is close to human buildings, and the populations are large, so they can be easily collected in cities (Fig. 1A, B). These aforementioned characteristics make it an ideal model species of the Uloboridae to study the biological characteristics of this family.

Observations of Octonoba sinensis. (A) Two specimens of O. sinensis on their orb webs in their natural environment. (B) Adult female of O. sinensis. (C) The perspective diagram in (D), with arrows numbered 1–4 indicating the shooting angles of the 4 images in (D). (D) Scanning electron microscope image of O. sinensis fangs.
Figure 1:

Observations of Octonoba sinensis. (A) Two specimens of O. sinensis on their orb webs in their natural environment. (B) Adult female of O. sinensis. (C) The perspective diagram in (D), with arrows numbered 1–4 indicating the shooting angles of the 4 images in (D). (D) Scanning electron microscope image of O. sinensis fangs.

In this study, we generated a chromosome-level genome assembly of O. sinensis. High-quality annotation was performed on the genomes of Pardosa pseudoannulata and Dolomedes plantarius, as well as optimization of annotation for the genomes of Uloborus diversus and Latrodectus elegans. Furthermore, we assembled a full-protein reference for Deinopis sp. using transcriptome data. By leveraging multiomics datasets from various spider species, we explored the molecular basis of the unique adaptive strategies in the family Uloboridae. In the selection pressure analysis, we found that some important genes related to muscle movement have undergone significant positive selection. In gene family analysis, we observed the absence or relaxation in selection pressure of toxins and developmental related genes. Gene families related to tracheal development, skeletal development, and tissue force-bearing structures have significantly expanded or emerged in the genomes of this group. Moreover, energy metabolism–related genes exhibit high expression, and enzyme activity in related pathways is significantly increased. These all provide a plausible explanation for the improvement in respiratory efficiency and the well-developed tracheal system observed.

Results

Prey-wrapping behavior observations and fang morphology

Through our observations, like other species in the family Uloboridae, O. sinensis only arrest their prey through extensive silk wrapping. The time usually exceeds 3 minutes (sometimes even 8–9 minutes), during which there may be several brief breaks (Additional Files 13). This time is much higher than the previously recorded wrapping time of other spider species (Araneoidea, 10.1 ± 1.2 seconds for flies, 43.2 ± 6.7 seconds for dragonflies) [6].

In addition, for the first time, we examined the fangs of O. sinensis from multiple angles with a scanning electron microscope (SEM). Generally speaking, if a species has venom glands, a channel opening for injecting venom should be found on the fangs [1, 13]. We did not observe this in O. sinensis, which further provides evidence that uloborids are not equipped to deliver venom (Fig. 1C, D).

Genome assembly and annotation

We assembled an O. sinensis genome of 1.34 Gb, which is slightly smaller than the prediction of 1.47 Gb based on Illumina data (Additional File 4: Supplementary Table  S1). The average GC content is 32.57%, and N50 value is 139.92 Mb. A total of 20 scaffolds were obtained, of which more than 99.9% of the sequences were loaded onto 9 scaffolds that reached the chromosome level (Fig. 2A), which was consistent with the previous karyotype analysis of O. sinensis [14]. The BUSCO [15, 16] complete analysis was 95.3% (arachnida_odb10, Additional File 4: Supplementary Table  S2).

Genome and comparative genomic analysis results from Octonoba sinensis. (A) Assembly and structural annotation of the O. sinensis genome. The distribution of different elements is marked with lowercase letters from the outside to the inside in the circle diagram, a: gene; b: SINE, c: LINE, d: LTR; e: DNA transposon; f: highly conserved elements (HCEs); g: GC content. The inner lines of the circle graph are collineation, dark red lines are collineation between chromosomes, and pink lines represent collineation within the chromosome. (B) Phylogeny of multiple spider species. All nodes received 100% bootstrap support. New annotated genomes provided in this study are marked in light blue, and new genome assemblies and annotations are marked in light red. The 95% confidence interval of the divergence time is represented by the gray bars on the phylogeny, and the number of expanded and contracted gene families at each node is represented by light blue and light green, respectively. The asterisk indicates the foreground branch in the selection pressure analysis. The branch with only transcriptome data cannot deploy CAFE analysis and is highlighted by a red branch. The collinearity between O. sinensis and other spider species is visualized through circle diagrams positioned adjacent to each respective species. The proportion of collinearity segments within the O. sinensis genome is emphasized in bold percentages.
Figure 2:

Genome and comparative genomic analysis results from Octonoba sinensis. (A) Assembly and structural annotation of the O. sinensis genome. The distribution of different elements is marked with lowercase letters from the outside to the inside in the circle diagram, a: gene; b: SINE, c: LINE, d: LTR; e: DNA transposon; f: highly conserved elements (HCEs); g: GC content. The inner lines of the circle graph are collineation, dark red lines are collineation between chromosomes, and pink lines represent collineation within the chromosome. (B) Phylogeny of multiple spider species. All nodes received 100% bootstrap support. New annotated genomes provided in this study are marked in light blue, and new genome assemblies and annotations are marked in light red. The 95% confidence interval of the divergence time is represented by the gray bars on the phylogeny, and the number of expanded and contracted gene families at each node is represented by light blue and light green, respectively. The asterisk indicates the foreground branch in the selection pressure analysis. The branch with only transcriptome data cannot deploy CAFE analysis and is highlighted by a red branch. The collinearity between O. sinensis and other spider species is visualized through circle diagrams positioned adjacent to each respective species. The proportion of collinearity segments within the O. sinensis genome is emphasized in bold percentages.

In the annotation of repetitive sequences of the genome, we found that the proportion of repetitive regions was 55.08%, and the most recognizable element was DNA transposons, which accounted for 18.5% of the genome (Additional File 4: Supplementary Table  S3, Additional File 5: Supplementary Fig. S1), similar to other spider groups with genomic data [17–19]. In our assembly, 24,579 coding genes were annotated, and 24,563 genes have obtained effective functional annotation in at least one of the following databases: NCBI-Nr, Swiss-Prot, or EggNOG v5.0 [20]. The chromosome loading rate of coding genes was 99.09%. The BUSCO assessment of protein level was 94.8% (Additional File 4: Supplementary Table  S2).

Divergence time estimation and synteny analysis

Genomes of O. sinensis, U. diversus [22], and the model species—the house spider (Parasteatoda tepidariorum) [21]—and 7 other representative species (Additional File 4: Supplementary Table  S4) were selected for orthologous gene identification (see Methods). A total of 1,560 single-copy gene families shared by all species were identified and used to construct the phylogenetic tree (Fig. 2B). All nodes have 100% ultrafast bootstrap support, and the topolog and the estimated divergence time of each node are similar to those of previous studies [23–25].

Among the species involved in the above analysis, genomes with chromosome-level assemblies were selected for synteny analysis with O. sinensis (Additional File 4: Supplementary Table  S4). There is a trend that the closer the relationship, the stronger the collinearity (Fig. 2B). Because Deinopis sp. lack comprehensive genomic data, this species was not included in the synteny analysis (red branch, Fig. 2B).

Genes under positive selection and energy metabolism in muscle

In the selection pressure analysis, at the node of the Uloboridae (asterisk, Fig. 2B), 401 genes are under positive selection (Additional File 4: Supplementary Table  S5). Although these genes did not achieve effective Gene Ontology (GO) enrichment (P-adjust < 0.05), we found that there is tissue preference in the expression of some positive selection genes (PSGs), and in O. sinensis, these genes have the highest enrichment in embryos and muscles (Fig. 3A). It is worth noting that these PSGs in muscle tissue include myosin (gene ID: g27351, P = 2.59e-04), an important molecular motor [26, 27], and twitchin (gene ID: g8872, P = 7.55e-05), a key regulator of muscle movement [28].

Positive selection genes and energy metabolism. (A) Tissue enrichment of genes under positive selection. (B) Differential expression analysis of homologous genes in the legs of Parasteatoda tepidariorum and Octonoba sinensis (only reciprocal best hit genes were considered, fold-change > 1.5, P < 0.05). Genes exhibiting higher expression levels in O. sinensis legs are designated with red upper triangles, whereas those with higher expression in P. tepidariorum legs are designated with green lower triangles. (C) GO terms enrichment analysis. Compared to the legs of P. tepidariorum, genes with higher expression (fold-change > 1.5, P < 0.05) in the legs of O. sinensis were analyzed. GO terms related to the mitochondrion are highlighted in red. Metabolic rate measurement (D) and enzyme activity (E) of hexokinase (HK), pyruvate dehydrogenase (PDH), alpha-ketoglutarate dehydrogenase complex (α-KGDHC), NADH dehydrogenase (ND), and ATP synthase (ATPase). Os: O. sinensis; Pt: P. tepidariorum; -t: under fatigue treatment. Significant differences are denoted by *P < 0.05 and **P < 0.01; n.s., not significant.
Figure 3:

Positive selection genes and energy metabolism. (A) Tissue enrichment of genes under positive selection. (B) Differential expression analysis of homologous genes in the legs of Parasteatoda tepidariorum and Octonoba sinensis (only reciprocal best hit genes were considered, fold-change > 1.5, P < 0.05). Genes exhibiting higher expression levels in O. sinensis legs are designated with red upper triangles, whereas those with higher expression in P. tepidariorum legs are designated with green lower triangles. (C) GO terms enrichment analysis. Compared to the legs of P. tepidariorum, genes with higher expression (fold-change > 1.5, P < 0.05) in the legs of O. sinensis were analyzed. GO terms related to the mitochondrion are highlighted in red. Metabolic rate measurement (D) and enzyme activity (E) of hexokinase (HK), pyruvate dehydrogenase (PDH), alpha-ketoglutarate dehydrogenase complex (α-KGDHC), NADH dehydrogenase (ND), and ATP synthase (ATPase). Os: O. sinensis; Pt: P. tepidariorum; -t: under fatigue treatment. Significant differences are denoted by *P < 0.05 and **P < 0.01; n.s., not significant.

To further explore the evolution of the motor function of Uloboridae, we compared the transcriptome data of the legs between O. sinensis and other species. We used the model species P. tepidariorum, which is also a web-building spider, as a control. Under consistent standardized conditions, the results showed that a large number of genes were differentially expressed (Fig. 3B, Additional File 4: Supplementary Table  S6). Compared to the legs of P. tepidariorum, genes with higher expression in the legs of O. sinensis were most enriched in the mitochondrial matrix; meanwhile, significant enrichment was also observed in other GO terms about mitochondria (Fig. 3C, Additional File 5: Supplementary Fig. S2). This result suggests that at least in the legs, O. sinensis requires greater energy consumption compared to typical web-building spiders. On this basis, we examined the activities of several key enzymes in the mitochondria involved in energy metabolism, including hexokinase, pyruvate dehydrogenase, α-ketoglutarate dehydrogenase, NADH dehydrogenase, and ATP synthase. Our results revealed that the activity levels of these 5 enzymes in the body of O. sinensis were higher compared to those in P. tepidariorum (Fig. 3E). Furthermore, while there was no significant difference in CO2 production rates between O. sinensis and P. tepidariorum in a resting state, O. sinensis exhibited significantly higher rates under fatigue treatment (Fig. 3D). These findings suggest that the evolution related to energy metabolism in the motor organs may be a key factor in the sustained output power of species in the Uloboridae family.

Expanded and new emergent gene families

To compare the genomic differences between Uloboridae and other spiders, we used CAFE v4.2 [29] to analyze the gene family expansions and contractions. Results indicate that 123 gene families have undergone significant expansion at the ancestral node of the Uloboridae (Fig. 4A). In O. sinensis, we found 4 of these families to have a high number of annotated genes that corresponded to FH2 domain containing 1 (FHDC1), FBN1, WD40 repeat proteins (WD40), and 7-(pass)-transmembrane domain receptors 1 (7tm_1) (Fig. 4B, Additional File 4: Supplementary Table  S7). Interestingly, studies have shown that FHDC1 proteins not only play a crucial role in the development of the tracheal system in fruit flies [30] but also have significant implications in muscle movement [31, 32]. In addition, the preproprotein of FBN1 is proteolytically processed to generate 2 proteins, including the extracellular matrix component fibrillin 1 and the protein hormone asprosin. Fibrillin 1 is an extracellular matrix glycoprotein that serves as a structural component of calcium-binding microfibrils. These microfibrils provide force-bearing structural support in elastic and nonelastic connective tissue throughout the body. Asprosin has been shown to regulate glucose homeostasis [33]. Apart from the 4 superfamilies mentioned above, we have deployed GO enrichment analysis of the other expanded gene families. REVIGO [34] results show that these annotated genes were mainly enriched in the transport of carbohydrates and organic acids, the immune system, and the functions related to transposable elements (Fig. 4C).

Expanded and new emergent gene families. (A) Species phylogeny for calculating the expansion and contraction of gene families. Uloboridae branches are in red. (B) The 4 superfamilies with the highest number of clearly annotated genes. (C) REVIGO plot of GO enrichment results for the other expanded gene families, excluding 4 superfamilies. The numbers in the figure refer to different GO terms: 1, response to molecules of bacterial origin; 2, DNA recombination; 3, DNA transposition; 4, receptor-mediated endocytosis; 5, carbohydrate transport; 6, DNA integration; 7, organic anion transport; 8, sialic acid transport; 9, organic acid transport; 10, transposition; 11, carbohydrate transmembrane transport; 12, response to type II interferon; 13, carboxylic acid transport; 14, response to interleukin 4; 15, cellular response to biotic stimulus; 16, cellular response to molecule of bacterial origin; 17, cellular response to type II interferon; 18, cellular response to interleukin 4; 19, monoatomic anion transmembrane transport; 20, inorganic cation transmembrane transport; 21, carbohydrate derivative transport; 22, proton transmembrane transport; 23, organic acid transmembrane transport; 24, carboxylic acid transmembrane transport. (D) GO enrichment of new emergent gene families in Uloboridae.
Figure 4:

Expanded and new emergent gene families. (A) Species phylogeny for calculating the expansion and contraction of gene families. Uloboridae branches are in red. (B) The 4 superfamilies with the highest number of clearly annotated genes. (C) REVIGO plot of GO enrichment results for the other expanded gene families, excluding 4 superfamilies. The numbers in the figure refer to different GO terms: 1, response to molecules of bacterial origin; 2, DNA recombination; 3, DNA transposition; 4, receptor-mediated endocytosis; 5, carbohydrate transport; 6, DNA integration; 7, organic anion transport; 8, sialic acid transport; 9, organic acid transport; 10, transposition; 11, carbohydrate transmembrane transport; 12, response to type II interferon; 13, carboxylic acid transport; 14, response to interleukin 4; 15, cellular response to biotic stimulus; 16, cellular response to molecule of bacterial origin; 17, cellular response to type II interferon; 18, cellular response to interleukin 4; 19, monoatomic anion transmembrane transport; 20, inorganic cation transmembrane transport; 21, carbohydrate derivative transport; 22, proton transmembrane transport; 23, organic acid transmembrane transport; 24, carboxylic acid transmembrane transport. (D) GO enrichment of new emergent gene families in Uloboridae.

Based on the results of orthologous gene identification, we screened for gene families that are exclusively shared in Uloboridae and not found in any other spider species (species in Fig. 2B), designating them as new emergent gene families. In total, 269 such gene families have been identified in both O. sinensis and U. diversus, with a total of 658 members in the O. sinensis genome (Additional File 4: Supplementary Table  S8). GO enrichment results showed that these genes are more enriched in GO terms that are related to bone trabecular development (Fig. 4D).

Generally speaking, spiders lack endurance, necessitating the prompt subduing of their prey within a brief timeframe during hunting [13]. However, uloborids can exercise intensely for nearly an hour [9]. Our results indicate that genes related to tracheal development, skeletal development, tissue force-bearing structures, and energy metabolism have significantly expanded or emerged in the genome of Uloboridae. We believe that the evolution of these aspects is highly likely to be related to their increased endurance.

Absent regions and genes under relaxed purifying selection

Due to the absence of venom glands, the genes or functional regions specifically involved in the venom gland system in Uloboridae may be subjected to relaxed purifying selection or gradually lost from the genome. To obtain this information, highly conserved elements (HCEs) were searched spanning 9 spider genomes (Fig. 4A), and the sites with Uloboridae-specific deletions among them were identified (Additional File 5: Supplementary Fig. S3, Additional File 4: Supplementary Tables S9S10). In addition, absent genes and genes under selective relaxation in the O. sinensis and U. diversus genomes were analyzed against the background of species with venom glands (Fig. 2B, Additional File 5: Supplementary Fig. S3, Additional File 4: Supplementary Tables S11–S12). We conducted functional enrichment on the above results and found that the biggest difference between Uloboridae and background species comes from the development, especially the neurodevelopment, related to gene family (Fig. 5A, B).

Absent regions and regions under relaxed selection. (A) GO enrichment of genes under relaxed selection in Octonoba sinensis. (B) GO enrichment of genes related to missing HCEs in O. sinensis. (C) The expression patterns of the genes specifically absent in Uloboridae were examined in the model species Parasteatoda tepidariorum. The heatmap is plotted based on the z-score transformed from transcripts per million (TPM) values. Different-colored markers are used to distinguish between distinct gene types, while an asterisk identifies the gene that is notably absent from the venom gland–specific expression module of P. tepidariorum.
Figure 5:

Absent regions and regions under relaxed selection. (A) GO enrichment of genes under relaxed selection in Octonoba sinensis. (B) GO enrichment of genes related to missing HCEs in O. sinensis. (C) The expression patterns of the genes specifically absent in Uloboridae were examined in the model species Parasteatoda tepidariorum. The heatmap is plotted based on the z-score transformed from transcripts per million (TPM) values. Different-colored markers are used to distinguish between distinct gene types, while an asterisk identifies the gene that is notably absent from the venom gland–specific expression module of P. tepidariorum.

In the homologs missing from Uloboridae, we found that 5 genes belong to 2 toxin-related gene families (LRR and CRISP), as well as 3 transcription factors. It is worth noting that the expression patterns of the homologs of 2 LRR genes (LOC122270931, LOC107442855) in P. tepidariorum indicate their highest expression in venom glands, although neither of them has been annotated as homologs to known toxin genes (Fig. 5C). Three transcription factors belong to important components that activate transcription: “protein c-ets-2,” the “cooled coil and C2 domain containing protein (CC2D)” that regulates neurotransmitter expression, and the component of the STAGA complex: “ataxin-7.”

In addition, we referred to the venom gland–specific expression module of P. tepidariorum obtained in a previous study (n = 1,088, Additional File 4: Supplementary Table  S13) [21] and found that only 1 gene (LOC107440400), which is specifically lost in Uloboridae, intersects with this module. This gene is a SOBP (Sine Oculis-Binding Protein Homolog), and the protein encoded by this gene is involved in the development of the cochlea, and genetic defects are also related to intellectual disability (Fig. 5C).

Deficiency of toxin genes in O. sinensis

To search for toxin genes in O. sinensis, we integrated the results of previous studies and established a comprehensive toxin protein database (Additional File 6) and screened toxin gene homologs with the same threshold in different species. In the O. sinensis and U. diversus genomes, we identified 12 and 11 homologs, respectively, that have similar structures to members of 6 major toxin or venom component gene families (latrotoxin, latrodectin, CRISP, ICK, TCTP, EF-hand, and ctenitoxin). Our findings indicate that although the Uloboridae family has a relatively low number of toxin-related homologs, their count still exceeds that of some venomous spiders, such as Ectatosticta davidi [35–38] (Additional File 4: Supplementary Table  S14). Nevertheless, we discovered that none of these O. sinensis genes’ orthologs in P. tepidariorum exhibit high expression levels in venom glands (Fig. 6). Given that O. sinensis had 3 latrotoxin homologs out of 12 toxin homologs (highest category, Additional File 4: Supplementary Table S14) and latrotoxins are not known outside of Theridiidae, a phylogenetic analysis of this gene family has been conducted. In this analysis, we found that the 3 homologs in O. sinensis are not clustered on the same branch as the reported latrotoxin (Additional File 5: Supplementary Fig. S4) [39]. There is a hypothesis regarding the evolution of venom components that ancestors of toxin proteins were originally proteins with normal physiological functions that were recruited in venom glands to play the role of venom components [40], and our results also support this viewpoint.

The expression patterns of toxin homolog coding genes in Octonoba sinensis and Parasteatoda tepidariorum. The heatmap is plotted based on the z-score transformed from TPM values. The ribbon connects the homolog genes. The black asterisks indicate the best match of possible O. sinenesis pseudogenes in P. tepidariorum. The black squares indicate homologs matching the O. sinenesis existence position in the collinearity (see Fig. 7B: red ribbons).
Figure 6:

The expression patterns of toxin homolog coding genes in Octonoba sinensis and Parasteatoda tepidariorum. The heatmap is plotted based on the z-score transformed from TPM values. The ribbon connects the homolog genes. The black asterisks indicate the best match of possible O. sinenesis pseudogenes in P. tepidariorum. The black squares indicate homologs matching the O. sinenesis existence position in the collinearity (see Fig. 7B: red ribbons).

To further explore the evolutionary processes of O. sinensis toxin genes in the absence of venom glands, we conducted a pseudogene search on noncoding regions of the O. sinensis genome, but traditional search methods did not identify pseudogenes (blastn, E-value 1e-5, matching length 50 bp) [41]. However, we did find more traces of toxin gene homologs in the blastx search through the protein sequence in the toxin genes database. These results include 48 different genomic regions, but only 1 of which had an effective hit with the relatively reliable toxin gene of P. tepidariorum (LOC107440051) (Additional File 4: Supplementary Table  S15, Fig. 6).

By searching for toxin gene homologs in collinearity fragments of O. sinensis and P. tepidariorum, a particular class of genes was found in O. sinensis. These genes are located in the same place as the P. tepidariorum toxin gene homologs in the collinearity segment, but they can no longer be identified as toxin genes (below the minimum recognition threshold; see Methods) (Fig. 7B, red ribbon). There are 3 pairs of such genes, including 1 pair of CRISP genes (g31478∼LOC107437238) and 2 pairs of ICK genes (g6736∼LOC107446942 and g6736∼LOC107446932). Their expression patterns in P. tepidariorum indicate that only the CRISP gene is a reliable toxin gene (Fig. 6, black squares). In O. sinensis, this CRISP gene (g31478) cannot be unambiguously classified as a homolog of toxin genes due to changes in protein structure (Additional File 4: Supplementary Table  S16). However, compared with other toxin genes, for which it is difficult to find pseudogenes, the ortholog of this CRISP gene in O. sinensis has complete gene structures and can be expressed in multiple tissues of a venomless spider (Additional File 4: Supplementary Table  S17). These all indicate that this CRISP gene (g31478) must play the role of a nontoxic gene. We believe that this observation suggests a potential functional shift between toxic and nontoxic genes in spiders.

Collinearity containing toxin homologs between Octonoba sinensis and Parasteatoda tepidariorum. (A) Collinearity of all toxin homologs. If the toxin homologs were lost in O. sinensis, the collinearity fragments are represented by a pink ribbon and otherwise by a dark red band. (B) Gene correspondence in each collinearity fragment where toxin gene deletion occurs in (A). Genes located on the positive strand of the genome are represented in blue, while those located on the negative strand are represented in green. Each lowercase letter corresponds to the collinearity fragments represented by the same letter in (A); the toxin gene names in P. tepidariorum are listed, and the red gene names highlight the toxin genes that are highly expressed in the venom glands. The toxin genes in the gray background belong to the latrotoxin gene family, and different silhouettes indicate the type of toxin and the target group of toxicity. Mouse = vertebrates (alpha-latrotoxin), fly = insects (delta-latrotoxin); the red ribbons indicate that the collinearity of the linked genes is extremely weak, below the threshold (see Methods).
Figure 7:

Collinearity containing toxin homologs between Octonoba sinensis and Parasteatoda tepidariorum. (A) Collinearity of all toxin homologs. If the toxin homologs were lost in O. sinensis, the collinearity fragments are represented by a pink ribbon and otherwise by a dark red band. (B) Gene correspondence in each collinearity fragment where toxin gene deletion occurs in (A). Genes located on the positive strand of the genome are represented in blue, while those located on the negative strand are represented in green. Each lowercase letter corresponds to the collinearity fragments represented by the same letter in (A); the toxin gene names in P. tepidariorum are listed, and the red gene names highlight the toxin genes that are highly expressed in the venom glands. The toxin genes in the gray background belong to the latrotoxin gene family, and different silhouettes indicate the type of toxin and the target group of toxicity. Mouse = vertebrates (alpha-latrotoxin), fly = insects (delta-latrotoxin); the red ribbons indicate that the collinearity of the linked genes is extremely weak, below the threshold (see Methods).

Discussion

During evolution, the loss of important organs is accompanied by a series of adaptive evolution, such as the enhancement of nonvisual senses of some eyeless organisms [42–46]. In Uloboridae, due to the absence of venom glands, prey-wrapping as the sole attack strategy is excessive. Multiomics analysis revealed the absence or possible functional shift of toxin genes in O. sinensis. At the same time, compared to venomous spiders, a series of genomic changes related to muscle, aerobic respiration, and metabolism of energy substances were observed in the genomes. These findings imply an evolutionary trade-off between the loss of venom glands and the enhancement of physical attack strategies.

Both gene family evolution and selection pressure shaped the genomic change related to the specialized physical attack of Uloboridae. For example, myosin and twitchin were under positive selection. Myosin provides the driving force for muscle contraction as a molecular motor [26, 27], while twitchin is located at the junction of muscle fibers and regulates the speed and force of muscle contraction [28]. The positive selection of these 2 genes suggests that they may play a role in the strengthening of muscle energy efficiency and contraction strength. Additionally, the gene family expansions of key proteins involved in muscle fiber connection, FHDC1 [31, 32], and the structural component of connective tissue, FBN1 [33], likely provided genetic resources for the optimization of muscle and connective tissue.

Aerobic respiration is the process that directly supplies energy to muscles. We found a higher expression of mitochondria-related genes in the legs of O. sinensis compared to P. tepidariorum (a venomous spider) (Fig. 3C). In addition, physiological and biochemical measurements also indicate an advantage of uloborids in respiratory efficiency and related enzyme activity (Fig. 3D, E). Moreover, the expansion of gene families involved in tracheal development [30], coupled with previous records that uloborids possess a more complex tracheal system compared to other web-building spiders [10, 11], all indicates that uloborids have the ability to provide sufficient oxygen supply for enhanced aerobic respiration.

The supply of energy substances is also crucial for sustained physical output. We found that the gene families involved in the transport of carbohydrates and organic acids have undergone expansion (Fig. 4C), which may help uloborids’ enhanced utilization and metabolism of substrates (or intermediate products) in aerobic respiration. Additionally, asprosin, encoded by an expanded gene family [33], acts as hormone-regulating glucose homeostasis, potentially mobilizing energy reserves to further provide energy for long-term physical activity.

Simultaneously, through analyzing expression patterns and synteny relationships, we discovered that toxin genes in venomous spiders are expressed in other tissues of uloborids and may perform nontoxic functions. This observation suggests a potential functional shift between toxic and nontoxic genes in spiders. In addition to toxin genes, the absence of certain genes exclusively in uloborids raises concerns. These include protein c-ets-2, CC2D, and ataxin-7, which are transcription factors involved in various metabolic and developmental processes. Additionally, notable missing genes include the SOBP gene and 2 LRR genes. The former is involved in nervous and organ development, while the latter has a similar structural domain to LRR toxin genes, all of which are highly expressed in venom glands of model species.

For predators, there exists a pervasive evolutionary trade-off between chemical and physical attack strategies. Previous research has frequently favored the exploration of chemical strategies, particularly venom. However, against the backdrop of nearly all spiders being toxic predators, our study delves into the genetic basis underlying the alternative choice in this trade-off. Unsurprisingly, reliable toxin gene was not identified in Uloboridae, but these adaptive evolutions ranging from muscle to aerobic respiration and then to the supply of energy substances provide strong support for the exceptional physical endurance demands of this group and compensate for their decreased adaptability due to the absence of venom glands. Furthermore, some development-related gene and element deletions were observed in uloborids. Although the association between these deletions and the absence of spider venom glands remains unclear, they still hold potential for exploring the evolutionary mechanisms underlying this phenomenon.

Methods

Sample collection and DNA extraction

Live O. sinensis were field-captured from Olympic Park, Chaoyang District, Beijing, China. To minimize the contamination of impurities in the digestive tract as much as possible, all samples were starvation-reared for more than 1 week at room temperature. Genome DNA for both short- and long-read sequencing were isolated from the cephalothoraxes of adult female spiders using the Blood & Cell Culture DNA Kit (QIAGEN).

Observation of predation behavior and examination of fangs

We recorded a series of videos to observe the predation behavior, using a Logitech StreamCam 960–001,282. Wrapping duration was timed. If the spider stopped for more than 3 seconds, the timing also stopped but would continue if it began again. If the spider began to soak (O. sinensis will emit liquid onto their prey before eating) or left the prey, the video was terminated.

The dissected spider fangs were pasted onto a copper substrate at different angles, dried with a CO2 critical point drying method, coated with gold, and then observed with the SEM (model: FEI Quanta 450).

Genome sequencing and genome size estimation

Short-insert libraries of O. sinensis were sequenced with the DNBSEQ-G400 (RRID:SCR_017980) using paired-end (PE) reads of 150 bp (BGI). To remove low-quality reads and adapters, raw reads were trimmed by Trimmomatic (RRID:SCR_011848) v0.39 [47]. A total of 259.82 Gb of clean data were obtained for O. sinensis for survey analysis and assembly correction.

For long-read sequencing, “SMRTbell” (double-stranded DNA template capped by hairpin loops at both ends) libraries were constructed according to the standard protocol of PacBio using 15 kb of preparation solution (PacBio). The high-fidelity (HiFi) libraries were sequenced on 3 SMRT cells on the PacBio Sequel II system (RRID:SCR_017990) in Circular Consensus Sequencing mode at Novogene Technology and generated 67.18 Gb HiFi data (3,967,026 reads) total [48].

To further improve the continuity of the assembled genomes, chromosome conformation capture (Hi-C) experiments were conducted [49]. Hi-C libraries were prepared following a published protocol with minor modifications [50]. For cross-linking, samples were fixed with 1% formaldehyde. The cross-linked DNA was digested with MboI restriction endonuclease and marked with biotin-14-dCTP to remove nonligated DNA fragments. The ligated DNA was extracted with a QIAamp DNA Mini Kit (Qiagen). The purified DNA was then sheared to ∼350-bp fragments and followed by a standard Illumina library preparation protocol [51]. Hi-C sequencing of O. sinensis was conducted on the DNBSEQ-G400 platform with PE 150 bp (BGI). We then filtered the raw reads using Juicer v1.6.2 [52] to remove low-quality reads and adapters, yielding 201.99 Gb of clean data.

Before de novo assembly, we estimated the genome size of this species. Using the Illumina data, Jellyfish (RRID:SCR_005491) (v2.1.3) [53] was employed to calculate the frequency of each k-mer (k = 17–31). Then, the genome size of O. sinensis was estimated using a method based on k-mer distribution.

De novo genome assembly and quality assessment

PacBio reads were first assembled using 2 de novo assemblers: hifiasm (RRID:SCR_021069) v0.15.2 [54] and wtdbg2 (RRID:SCR_017225) v2.5 [55]. The best assembly was selected according to the optimal continuity and completeness. The final version of contigs was polished with Racon (RRID:SCR_017642) v1.4.17 for 3 rounds based on long reads and NextPolish (RRID:SCR_025232) v1.4.0 [56] using short reads. Contig-level genome completeness assessment was performed using BUSCO (RRID:SCR_015008) v5.2.2 [15, 16]. Genome consistency assessment was evaluated by mapping the short reads to the genome with Minimap2 (RRID:SCR_018550) v2.24-r1122 [57] and SAMtools/BCFtools (RRID:SCR_005227) v1.10 [58].

We used Hi-C–based proximity-guided assembly to generate chromosomal-level genome assemblies for O. sinensis. Hi-C library sequencing data were mapped to the contig-level genome using Juicer (RRID:SCR_017226). The 3D-DNA (RRID:SCR_017227) v180922 [59] pipeline was executed to construct the chromosomes and correct the errors. We further performed correction with Juicebox Assembly Tools (RRID:SCR_021172) v1.11.08 [60]. The completeness of the chromosomal level assembly was assessed by BUSCO.

RNA extraction, sequencing, and expression analysis

Different tissue samples (spinneret, leg, brain, gut, silk gland, pedipalp, chelicera, and abdomen) from adult female O. sinensis and eggs (development stage undetermined) were dissected for total RNA extraction using an RNAsimple Total RNA kit (TIANGEN). The RNA sequencing libraries were constructed with insert sizes of ∼150 bp and sequenced on the NovaSeq 6000 platform (RRID:SCR_016387). We produced about 6 Gb of data per sample. Low-quality reads, reads with adapters, and unknown bases were filtered using Trimmomatic (RRID:SCR_011848).

In addition, legs from adult female P. tepidariorum were subjected to transcriptome analysis according to the above process (see Data Availability), and transcriptome data from 4 tissue transcriptomes (PRJNA934108, including brain, ovary, silk gland, and venom gland) were downloaded from the Sequence Read Archive (SRA) database.

Clean reads were aligned to the genome using Hisat2 (RRID:SCR_015530) [61], followed by the quantification of all samples with HTSeq (RRID:SCR_005514) [62] to determine the count value. Subsequently, TPMs were derived through automated scripts.

In the differential expression analysis across 2 species, only reciprocal best hit (RBH) genes were extracted for quantification. To facilitate comparison, ortholog (RBH) gene IDs of O. sinensis were replaced by P. tepidariorum gene IDs for comparison and figure illustration. Finally, differential expression analysis was conducted by R package LIMMA (RRID:SCR_010943) [63].

Genome annotation

The RepeatModeler (RRID:SCR_015027) v2.0.2 [64] and RepeatMasker (RRID:SCR_012954) v4.1.2-p1 [65] pipelines were used to annotate repetitive sequences in the genome.

Gene annotation was based on the braker (RRID:SCR_018964) v2.1.6 [66] pipeline, which combines the whole protein sequences of the 9 other species in this study (Additional File 4: Supplementary Table  S4) and more than 320 Gb of multitissue transcriptome data for comprehensive annotation.

Gene function annotation is based on NCBI-Nr, Swiss-Prot (RRID:SCR_021164), and EggNOG (RRID:SCR_002456) v5.0 databases. The transfer RNA was predicted using the program tRNAscan-SE (RRID:SCR_008637) v2.09 [67]. Other noncoding RNAs were annotated with the Rfam (RRID:SCR_007891) v14.8 [68] database through infernal (RRID:SCR_011809) v1.1.4 [69].

Among the other species involved in this study, P. pseudoannulata [70] and D. plantarius only have assembled sequence data currently. The genome of L. elegans has high assembly quality [37], but the protein BUSCO score is only 63.7%; U. diversus only has 15,750 annotated protein-coding genes, significantly less than other spider species. To obtain more reliable results for downstream analysis, we annotated the genomes of the aforementioned species based on transcriptome data. For P. pseudoannulata, D. plantarius, and L. elegans, we used our annotations, and for U. diversus, we used our annotation to supplement the original one. All annotation strategies are based on transcriptome data from the SRA (Additional File 4: Supplementary Table  S18) database and carried out through the TransDecoder (RRID:SCR_017647) pipelines. We also added Deinopis sp. of Deinopidae and assembled all of its proteins sequence using transcriptome data (DRR297048), and quality control was done using fastp (RRID:SCR_016962) version 0.21.0 [71]. De novo assemblies were done using Trinity (RRID:SCR_013048) v2.11.0 [72] under default settings. Open Reading Frame (ORF) prediction was done using TransDecoder. Redundancy reduction was done with CD-HIT (RRID:SCR_007105) version 4.8.1 (-c 0.98 -n 10) [73]. All the newly provided annotations mentioned above have a protein BUSCO score of over 90% (Additional File 4: Supplementary Table  S2).

Orthologous gene identification, phylogenetic and synteny analysis

We used OrthoFinder (RRID:SCR_017118) v2.5.4 [74] to analyze the annotation information of species in this study (Additional File 4: Supplementary Table  S5). In the pipeline, Mafft (RRID:SCR_011811) v7.453 [75] was used to perform multiple sequence alignment, blastp (RRID:SCR_001010) v2.9.0+ [76] was used to perform sequence searches, and the phylogenetic tree was constructed using IQ-TREE (RRID:SCR_017254) v2.2.0 [77]. The calibration points are from fossil specimens [23]. The analysis was run twice, once to calculate the divergence time of different species and the second to calculate the expansion and contraction of gene families. The latter did not include Deinopis sp., as the only transcriptome data cannot determine the number of gene copies (Fig. 4A).

The phylogenetic tree of latrotoxin homologs (Additional File 5: Supplementary Fig. S4) was reconstructed (1,000 bootstrap) using the neighbor-joining method, following alignment of the full-length protein sequences via Mafft v7.453.

Collinearity analysis was conducted between O. sinensis and other species with chromosome-level genomes (T. antipodiana, A. bruennichi, L. elegans, U. diversus, P. pseudoannulata, D. plantarius, and E. davidi [78]) using the MCscan (RRID:SCR_017650) pipeline (Python version) in the jcvi toolkit (RRID:SCR_021641) [79].

Test for selection pressure and gene family expansions/contractions

One-to-one ortholog identification among 10 species (Fig. 2B) was performed using the RBH method by blastp v2.9.0+. O. sinensis was used as a reference species. Finally, 5,848 RBH clusters (Additional File 4: Supplementary Table  S19) were retained for analysis.

To scan for genes under positive selection in O. sinensis, RBH clusters were used for the branch site model analysis using CODEML in the PAML v4.9j toolkit [80]. Each gene family sets O. sinensis and U. diversus as foreground branches (Fig. 2B). “Model A” and “Model A-null” models were compared. “Model A” assumes that the selection pressure of foreground branches is greater than that of background branches, and “Model A-null” is an alternative hypothesis.

To statistically test which genes of Uloboridae are under relaxed purifying selection, we used RELAX in the HYPHY (RRID:SCR_016162) v2.5.2 [81] toolkit to infer the free relaxation parameter k at the node of O. sinensis and U. diversus branches for genes shared by all species in the phylogeny of Fig. 2B (Additional File 4: Supplementary Table  S4). The relaxation parameter k is an exponent for selection parameters between the foreground and the background branches. A k > 1 suggests selection is more intensified in the foreground branch vs. the background branch and vice versa.

For gene family expansions/contractions, mcmctree in the PAML (RRID:SCR_014932) v4.9j toolkit was used to estimate the divergence time of each node in the phylogenetic tree of OrthoFinder pipeline results (without Deinopis sp.) (Fig. 4A). Next, we used CAFE (RRID:SCR_005983) v4.2 [29] under the default parameters to analyze the gene family expansions and contractions of 9 spider species.

Identification of HCEs

To identify the HCEs, we initially generated pairwise sequence alignments across all 9 spider genomes (Fig. 4A) with LASTZ (RRID:SCR_018556) v1.04.15 [82] and chainNet [83], using the P. tepidariorum genome as the reference. We then used MULTIZ v11.2 [84] to combine the pairwise alignments into multiple sequence alignments. Subsequently, we ran phyloFit in the PHAST package (RRID:SCR_003204) [85] with the topology from OrthoFinder to estimate the neutral (“nonconserved”) model based on 4-fold degenerate sites. With the nonconserved model as input, we ran phastCons [86] to estimate conserved models with its intrinsic function and predicted the HCEs.

The distribution of HCEs in exons, introns, 2,000 bp upstream and 2,000 bp downstream of genes, and intergenic regions was summarized with Annovar (RRID:SCR_012821) [87, 88] based on the genome annotation information of O. sinensis or P. tepidariorum.

Homologs of toxin gene family identification and analysis

Based on ArachnoServer 3.0 [89], a specialized spider venom database, and integrating toxin protein sequences obtained from other toxin research of spiders [37, 38, 90], we have compiled a new reference dataset. This dataset was used to conduct a blastp search for candidate toxin genes (E-value less than 1e-10, matching length greater than 70% of the reference sequence, and hit area mismatch less than 30%). We also established hidden Markov models for different types of toxin proteins based on the database and further confirmed the results obtained from blastp using HMMER (RRID:SCR_005305) v 3.3 [91]. To search for hidden toxin-related pseudogenes in O. sinensis, we referred to the identification criteria of human genome pseudogenes [41] and used blastn v2.9.0+ and blastx v2.9.0+ to search for candidates.

Physiological index measurement

Assays were performed as described previously [92]. CO2 production rate was used as a proxy for metabolic rate (MR). The assays were conducted in a closed-circuit system with a volume of 73.3 mL at a temperature of 25°C, a pressure of 100.6 kPa, and a flow rate of 110 mL/min. MR was calculated as the amount of CO2 produced per gram of body mass per second, using the equation MR = MCO2/T/body mass, where MCO2 represents the amount of oxygen substance (in mol). To put spiders into a state of fatigue, we stimulated the spider’s legs with a dissecting needle and kept it in a high-intensity state of exercise for 10 minutes.

Enzyme activity was measured using the corresponding reagent kit (Wuhan Mosak Biotechnology Co., Ltd. KT50129, KT50577, KT42310, KT41589, KT87867). Female spider individuals in a resting state were fixed in liquid nitrogen after weighing and stored at −80°C. Prior to testing, the samples were homogenized and diluted to 500 μL as the test solution following the supplier’s protocol.

Additional Files

Additional Files 1–3. Hunting video.

Additional File 4:

Supplementary Table S1. Genome survey prediction of Octonoba sinensis.

Supplementary Table S2. Genome assembly.

Supplementary Table S3. Repeat sequences of Octonoba sinensis genome.

Supplementary Table S4. Transcriptome assembly and genomes used in this study.

Supplementary Table S5. Genes under positive selection pressure on the node of the family Uloboridae.

Supplementary Table S6. Transcriptome differential expression analysis of the legs of Parasteatoda tepidariorum and Octonoba sinensis.

Supplementary Table S7. Genes list of significantly expanded gene families in Octonoba sinensis.

Supplementary Table S8. New emergent gene family members in the Octonoba sinensis genome.

Supplementary Table S9. All HCEs, using the Parasteatoda tepidariorum genome as a reference.

Supplementary Table S10. Missing HCEs in Octonoba sinensis and Uloborus diversus.

Supplementary Table S11. Specific missing orthologous groups in Octonoba sinensis and Uloborus diversus.

Supplementary Table S12. Genes under relaxed selection pressure in Uloboridae.

Supplementary Table S13. Parasteatoda tepidariorum venom gland-specific expression module.

Supplementary Table S14. Quantitative distribution of different toxin homologs across species.

Supplementary Table S15. “Pseudogene” blastx.

Supplementary Table S16. Protein domain search.

Supplementary Table S17. Expression levels (TPM) in different organizations of the genes connected by the red band to the Parasteatoda tepidariorum toxin gene in Fig. 7H.

Supplementary Table S18. The SRA data used for genome annotation.

Supplementary Table S19. Clusters of Reciprocal Best Hits (RBH) and RBH between Octonoba sinensis and Parasteatoda tepidariorum.

Additional File 5:

Supplementary Fig. S1. Recognizable elements of the Octonoba sinensis genome.

Supplementary Fig. S2. GO enrichment analysis of genes expressed at higher levels (fold-change > 4, P < 0.05) in O. sinensis legs relative to those in P. tepidariorum legs.

Supplementary Fig. S3. Distribution of all highly conserved elements (HCEs) in P. tepidariorum and missing HCEs and genes in O. sinensis.

Supplementary Fig. S4. Phylogenetic tree of homologs of neurotoxin genes (Latrotoxin).

Additional File 6. Dataset of toxin reference genes.

Abbreviations

BUSCO: Benchmarking Universal Single-Copy Orthologs; GO: Gene Ontology; HCE: highly conserved element; HiFi: high-fidelity; MR: metabolic rate; ORF: open reading frame; PE: paired end; PSG: positive selection gene; RBH: reciprocal best hit; SEM: scanning electron microscope; SRA: Sequence Read Archive; TPM: transcripts per million.

Acknowledgement

We thank Prof. Zhonge Hou and Assis. Prof. Fengyuan Li for academic suggestions, Wei Wang at Guangxi Normal University for her suggestions on anatomical techniques, and Dr. Nadia Ayoub and Dr. Sandra Correa-Garhwal for their careful review and valuable suggestions on this article. Yiming Zhang thanks his wife Lingling Liu, in particular, for the invaluable support over the years.

Author Contributions

Yiming Zhang (Data analysis [equal], Methodology [lead], Investigation [supporting], Photo/Video shooting [equal], Visualization [lead], Writing—original draft [lead], Writing—review & editing [lead]), Yunxiao Shen (Data analysis [equal], Methodology [supporting], Visualization [supporting], Writing—original draft [supporting]), Pengyu Jin (Investigation [supporting], Writing—review & editing [supporting]), Bingyue Zhu (Investigation [supporting], Methodology [supporting]), Yejie Lin (Species identification [lead], Photo shooting [equal]), Tongyao Jiang (Methodology [supporting]), Xianting Huang (Methodology [supporting]), Yang Wang (Video shooting [equal]), Zhe Zhao (Funding acquisition [supporting], Species identification [supporting]), Shuqiang Li (Funding acquisition [lead], Investigation [lead], Writing—original draft [supporting], Writing—review & editing [supporting]).

Funding

This study was supported by the Science and Technology Fundamental Resources Investigation Program of China (2023FY100200) and the Program of National Natural Sciences Foundation of China (NSFC–32,170,447, NSFC–32,370,490).

Data Availability

The sequencing data are deposited in NCBI BioProjects PRJNA1018860 and PRJNA1019401, as well as ScienceDB [93]. All additional supporting data are available in the GigaScience repository, GigaDB [94].

Competing Interests

The authors declare that they have no competing interests.

References

1.

Foelix
 
RF
,
Erb
 
B
.
Mesothelae have venom glands
.
J Arachnol
.
2010
;
38
(
3
):
596
98
. .

2.

NMBE—World Spider Catalog
. https://wsc.nmbe.ch/.
Accessed 9 September 2023
.

3.

Forster
 
RR
,
Platnick
 
NI
.
A review of the archaeid spiders and their relatives, with notes on the limits of the superfamily Palpimanoidea (Arachnida, Araneae)
.
Bull AMNH 1984;178:Article 1
.

4.

Rix
 
MG
.
A review of the Tasmanian species of Pararchaeidae and Holarchaeidae (Arachnida, Araneae)
.
J Arachnol
.
2005
;
33
(
1
):
135
52
. .

5.

Opell
 
BD
.
Revision of the genera and tropical American species of the spider family Uloboridae
.
Bull Museum Comparative Zool Harvard College
.
1979
;
148
:
443
549
.. https://biostor.org/reference/646.

6.

Robinson
 
MH
,
Olazarri
 
J
.
Units of behavior and complex sequences in the predatory behavior of Argiope argentata (Fabricius): (Araneae: araneidae)
. In:
Smithsonian Contributions to Zoology
.
Washington
, USA:
Smithsonian Institution Press
;
1971
;
65
. .

7.

Weng
 
JL
,
Barrantes
 
G
,
Eberhard
 
WG
.
Feeding by Philoponella vicina (Araneae, Uloboridae) and how uloborid spiders lost their venom glands
.
Can J Zool
.
2006
;
84
(
12
):
1752
62
. .

8.

Eberhard
 
WG
,
Barrantes
 
G
,
Weng
 
JL
.
Tie them up tight: wrapping by Philoponella vicina spiders breaks, compresses and sometimes kills their prey
.
Sci Nat Heidelberg
.
2006
;
93
(
5
):
251
54
. .

9.

Lubin
 
YD
.
Web buiding and prey capture in the Uloboridae
. In:
Spiders: Webs, Behavior, and Evolution
.
Stanford, USA
:
Stanford University Press
;
1986
:
132
71
.

10.

Opell
 
BD
.
The relationship of book lung and tracheal systems in the spider family uloboridae
.
J Morphol
.
1990
;
206
(
2
):
211
16
. .

11.

Opell
 
BD
.
The respiratory complementarity of spider book lung and tracheal systems
.
J Morphol
.
1998
;
236
(
1
):
57
64
. .

12.

de Plancy
 
VC
.
Arachnides Recueillis Aux Environs De Pékin
. In:
Annales de la Société Entomologique de France
.
Paris, FR
:
La Société
;
1880
.

13.

Rainer
 
F
.
Biology of Spiders
.
Oxford, UK: Oxford University Press
;
2011.
;

14.

Wang
 
X
,
Wang
 
Y
,
Yang
 
Z
, et al.  
On the karyotype of Octonnba sinensis
.
J Hebei Normal Univ
.
1997
;
21
(
4
):
423
26
.

15.

Seppey
 
M
,
Manni
 
M
,
Zdobnov
 
EM
.
BUSCO: assessing genome assembly and annotation completeness
.
Methods Mol Biol
.
2019
;
1962
:
227
45
. .

16.

Simao
 
FA
,
Waterhouse
 
RM
,
Ioannidis
 
P
, et al.  
BUSCO: assessing genome assembly and annotation completeness with single-copy orthologs
.
Bioinformatics
.
2015
;
31
(
19
):
3210
12
. .

17.

Hu
 
W
,
Jia
 
A
,
Ma
 
S
, et al.  
A molecular atlas reveals the tri-sectional spinning mechanism of spider dragline silk
.
Nat Commun
.
2023
;
14
(
1
):
837
. .

18.

Fan
 
Z
,
Yuan
 
T
,
Liu
 
P
, et al.  
A chromosome-level genome of the spider Trichonephila antipodiana reveals the genetic basis of its polyphagy and evidence of an ancient whole-genome duplication event
.
Gigascience
.
2021
;
10
(
3
):giab016. .

19.

Liu
 
S
,
Aagaard
 
A
,
Bechsgaard
 
J
, et al.  
DNA methylation patterns in the social spider, Stegodyphus dumicola
.
Genes
.
2019
;
10
(
2
):
137
. .

20.

Huerta-Cepas
 
J
,
Szklarczyk
 
D
,
Heller
 
D
, et al.  
eggNOG 5.0: a hierarchical, functionally and phylogenetically annotated orthology resource based on 5090 organisms and 2502 viruses
.
Nucleic Acids Res
.
2019
;
47
(
D1
):
D309
14
. .

21.

Zhu
 
B
,
Jin
 
P
,
Zhang
 
Y
, et al.  
Genomic and transcriptomic analyses support a silk gland origin of spider venom glands
.
BMC Biol
.
2023
;
21
(
1
):
82
. .

22.

Miller
 
J
,
Zimin
 
AV
,
Gordus
 
A
.
Chromosome-level genome and the identification of sex chromosomes in Uloborus diversus
.
Gigascience
.
2022
;
12:giad002
:, .

23.

Magalhaes
 
I
,
Azevedo
 
G
,
Michalik
 
P
, et al.  
The fossil record of spiders revisited: implications for calibrating trees and evidence for a major faunal turnover since the Mesozoic
.
Biol Rev Camb Philos Soc
.
2020
;
95
:
184
217
. .

24.

Shao
 
L
,
Zhao
 
Z
,
Li
 
S
.
Is phenotypic evolution affected by spiders' construction behaviors?
.
Syst Biol
.
2023
;
72
:
319
40
. .

25.

Wheeler
 
WC
,
Coddington
 
JA
,
Crowley
 
LM
, et al.  
The spider tree of life: phylogeny of Araneae based on target-gene analyses from an extensive taxon sampling
.
Cladistics
.
2017
;
33
(
6
):
574
616
. .

26.

Foth
 
BJ
,
Goedecke
 
MC
,
Soldati
 
D
.
New insights into myosin evolution and classification
.
Proc Natl Acad Sci USA
.
2006
;
103
(
10
):
3681
86
. .

27.

Squire
 
J
.
Special issue: the actin-myosin interaction in muscle: background and overview
.
Int J Mol Sci
.
2019
;
20
(
22
):
5715
. .

28.

Ayme-Southgate
 
A
,
Vigoreaux
 
J
,
Benian
 
G
, et al.  
Drosophila has a twitchin/titin-related gene that appears to encode projectin
.
Proc Natl Acad Sci USA
.
1991
;
88
(
18
):
7973
77
. .

29.

De Bie
 
T
,
Cristianini
 
N
,
Demuth
 
JP
, et al.  
CAFE: a computational tool for the study of gene family evolution
.
Bioinformatics
.
2006
;
22
(
10
):
1269
71
. .

30.

Matusek
 
T
,
Djiane
 
A
,
Jankovics
 
F
, et al.  
The Drosophila formin DAAM regulates the tracheal cuticle pattern through organizing the actin cytoskeleton
.
Development
.
2006
;
133
(
5
):
957
66
. .

31.

Iskratsch
 
T
,
Ehler
 
E
.
Formin-g muscle cytoarchitecture
.
Bioarchitecture
.
2011
;
1
(
2
):
66
68
. .

32.

Valencia
 
DA
,
Quinlan
 
ME
.
Formins
.
Curr Biol
.
2021
;
31
(
10
):
R517
22
. .

33.

Summers
 
KM
,
Bush
 
SJ
,
Davis
 
MR
, et al.  
Fibrillin-1 and asprosin, novel players in metabolic syndrome
.
Mol Genet Metab
.
2023
;
138
(
1
):
106979
. .

34.

Supek
 
F
,
Bosnjak
 
M
,
Skunca
 
N
, et al.  
REVIGO summarizes and visualizes long lists of gene ontology terms
.
PLoS One
.
2011
;
6
(
7
):
e21800
. .

35.

Escuer
 
P
,
Pisarenco
 
VA
,
Fernandez-Ruiz
 
AA
, et al.  
The chromosome-scale assembly of the Canary Islands endemic spider Dysdera silvatica (Arachnida, Araneae) sheds light on the origin and genome structure of chemoreceptor gene families in chelicerates
.
Mol Ecol Resour
.
2022
;
22
(
1
):
375
90
. .

36.

Luo
 
J
,
Ding
 
Y
,
Peng
 
Z
, et al.  
Molecular diversity and evolutionary trends of cysteine-rich peptides from the venom glands of Chinese spider Heteropoda venatoria
.
Sci Rep
.
2021
;
11
(
1
):
3211
. .

37.

Wang
 
Z
,
Zhu
 
K
,
Li
 
H
, et al.  
Chromosome-level genome assembly of the black widow spider Latrodectus elegans illuminates composition and evolution of venom and silk proteins
.
Gigascience
.
2022
;
11
:
giac049
. .

38.

Zhu
 
B
,
Jin
 
P
,
Hou
 
Z
, et al.  
Chromosomal-level genome of a sheet-web spider provides insight into the composition and evolution of venom
.
Mol Ecol Resour
.
2022
;
22
(
6
):
2333
48
. .

39.

Chen
 
M
,
Blum
 
D
,
Engelhard
 
L
, et al.  
Molecular architecture of black widow spider neurotoxins
.
Nat Commun
.
2021
;
12
(
1
):
6956
. .

40.

Luddecke
 
T
,
Herzig
 
V
,
von Reumont
 
BM
, et al.  
The biology and evolution of spider venoms
.
Biol Rev Camb Philos Soc
.
2022
;
97
(
1
):
163
78
. .

41.

Zhang
 
Z
,
Carriero
 
N
,
Zheng
 
D
, et al.  
PseudoPipe: an automated pseudogene identification pipeline
.
Bioinformatics
.
2006
;
22
(
12
):
1437
39
. .

42.

Protas
 
ME
,
Trontelj
 
P
,
Patel
 
NH
.
Genetic basis of eye and pigment loss in the cave crustacean, Asellus aquaticus
.
Proc Natl Acad Sci
.
2011
;
108
(
14
):
5702
7
. .

43.

Gore
 
AV
,
Tomins
 
KA
,
Iben
 
J
, et al.  
An epigenetic mechanism for cavefish eye degeneration
.
Nat Ecol Evol
.
2018
;
2
(
7
):
1155
60
.

44.

Mojaddidi
 
H
,
Fernandez
 
FE
,
Erickson
 
PA
, et al.  
Embryonic origin and genetic basis of cave associated phenotypes in the isopod crustacean Asellus aquaticus
.
Sci Rep
.
2018
;
8
(
1
):
16589
. .

45.

Piatigorsky
 
J
.
A genetic perspective on eye evolution: gene sharing, convergence and parallelism
.
Evol Educ Outreach
.
2008
;
1
(
4
):
403
14
. .

46.

Krishnan
 
J
,
Rohner
 
N
.
Cavefish and the basis for eye loss
.
Philos Trans R Soc Lond B Biol Sci
.
2017
;
372
(
1713
):
20150487
. .

47.

Bolger
 
AM
,
Lohse
 
M
,
Usadel
 
B
.
Trimmomatic: a flexible trimmer for Illumina sequence data
.
Bioinformatics
.
2014
;
30
(
15
):
2114
20
. .

48.

Wenger
 
AM
,
Peluso
 
P
,
Rowell
 
WJ
, et al.  
Accurate circular consensus long-read sequencing improves variant detection and assembly of a human genome
.
Nat Biotechnol
.
2019
;
37
(
10
):
1155
62
. .

49.

Lu
 
L
,
Liu
 
X
,
Huang
 
W
, et al.  
Robust hi-C maps of enhancer-promoter interactions reveal the function of non-coding genome in neural development and diseases
.
Mol Cell
.
2020
;
79
(
3
):
521
34
...

50.

Rao
 
SS
,
Huntley
 
MH
,
Durand
 
NC
, et al.  
A 3D map of the human genome at kilobase resolution reveals principles of chromatin looping
.
Cell
.
2014
;
159
(
7
):
1665
80
. .

51.

Meyer
 
M
,
Kircher
 
M
.
Illumina sequencing library preparation for highly multiplexed target capture and sequencing
.
Cold Spring Harb Protoc
.
2010
;
2010
(
6
):
pdb.prot5448
. .

52.

Durand
 
NC
,
Shamim
 
MS
,
Machol
 
I
, et al.  
Juicer provides a one-click system for analyzing loop-resolution hi-C experiments
.
Cell Syst
.
2016
;
3
(
1
):
95
98
. .

53.

Marcais
 
G
,
Kingsford
 
C
.
A fast, lock-free approach for efficient parallel counting of occurrences of k-mers
.
Bioinformatics
.
2011
;
27
(
6
):
764
70
. .

54.

Cheng
 
H
,
Concepcion
 
GT
,
Feng
 
X
, et al.  
Haplotype-resolved de novo assembly using phased assembly graphs with hifiasm
.
Nat Methods
.
2021
;
18
(
2
):
170
75
. .

55.

Ruan
 
J
,
Li
 
H
.
Fast and accurate long-read assembly with wtdbg2
.
Nat Methods
.
2020
;
17
(
2
):
155
58
. .

56.

Hu
 
J
,
Fan
 
J
,
Sun
 
Z
, et al.  
NextPolish: a fast and efficient genome polishing tool for long-read assembly
.
Bioinformatics
.
2020
;
36
(
7
):
2253
55
. .

57.

Li
 
H
.
Minimap2: pairwise alignment for nucleotide sequences
.
Bioinformatics
.
2018
;
34
(
18
):
3094
100
. .

58.

Danecek
 
P
,
Bonfield
 
JK
,
Liddle
 
J
, et al.  
Twelve years of SAMtools and BCFtools
.
Gigascience
.
2021
;
10
(
2
):
giab008
. .

59.

Dudchenko
 
O
,
Batra
 
SS
,
Omer
 
AD
, et al.  
De novo assembly of the Aedes aegypti genome using Hi-C yields chromosome-length scaffolds
.
Science
.
2017
;
356
(
6333
):
92
95
. .

60.

Dudchenko
 
O
,
Shamim
 
MS
,
Batra
 
S
, et al.  
The Juicebox Assembly Tools Module Facilitates De Novo Assembly of Mammalian Genomes with Chromosome-length Scaffolds for under $1000
.
Cold Spring Harbor, USA
:
Cold Spring Harbor Laboratory Press
;
2018.
;

61.

Kim
 
D
,
Paggi
 
JM
,
Park
 
C
, et al.  
Graph-based genome alignment and genotyping with HISAT2 and HISAT-genotype
.
Nat Biotechnol
.
2019
;
37
(
8
):
907
15
. .

62.

Anders
 
S
,
Pyl
 
PT
,
Huber
 
W
.
HTSeq—a Python framework to work with high-throughput sequencing data
.
Bioinformatics
.
2015
;
31
(
2
):
166
69
. .

63.

Ritchie
 
ME
,
Phipson
 
B
,
Wu
 
D
, et al.  
limma powers differential expression analyses for RNA-sequencing and microarray studies
.
Nucleic Acids Res
.
2015
;
43
(
7
):
e47
. .

64.

Flynn
 
JM
,
Hubley
 
R
,
Goubert
 
C
, et al.  
RepeatModeler2 for automated genomic discovery of transposable element families
.
Proc Natl Acad Sci USA
.
2020
;
117
(
17
):
9451
57
. .

65.

Tarailo-Graovac
 
M
,
Chen
 
N
.
Using RepeatMasker to identify repetitive elements in genomic sequences
.
Curr Protoc Bioinformatics
.
2009
;.

66.

Bruna
 
T
,
Hoff
 
KJ
,
Lomsadze
 
A
, et al.  
BRAKER2: automatic eukaryotic genome annotation with GeneMark-EP+ and AUGUSTUS supported by a protein database
.
NAR Genom Bioinform
.
2021
;
3
(
1
):
lqaa108
. .

67.

Chan
 
PP
,
Lin
 
BY
,
Mak
 
AJ
, et al.  
tRNAscan-SE 2.0: improved detection and functional classification of transfer RNA genes
.
Nucleic Acids Res
.
2021
;
49
(
16
):
9077
96
. .

68.

Kalvari
 
I
,
Nawrocki
 
EP
,
Ontiveros-Palacios
 
N
, et al.  
Rfam 14: expanded coverage of metagenomic, viral and microRNA families
.
Nucleic Acids Res
.
2021
;
49
(
D1
):
D192
D200
. .

69.

Nawrocki
 
EP
,
Eddy
 
SR
.
Infernal 1.1: 100-fold faster RNA homology searches
.
Bioinformatics
.
2013
;
29
(
22
):
2933
35
. .

70.

Yu
 
N
,
Li
 
J
,
Bao
 
H
, et al.  
Chromosome-level genome of spider pardosa pseudoannulata and cuticle protein genes in environmental stresses
.
Sci Data
.
2024
;
11
(
1
):
121
. .

71.

Chen
 
S
,
Zhou
 
Y
,
Chen
 
Y
, et al.  
fastp: an ultra-fast all-in-one FASTQ preprocessor
.
Bioinformatics
.
2018
;
34
(
17
):
i884
90
. .

72.

Grabherr
 
MG
,
Haas
 
BJ
,
Yassour
 
M
, et al.  
Full-length transcriptome assembly from RNA-seq data without a reference genome
.
Nat Biotechnol
.
2011
;
29
(
7
):
644
52
. .

73.

Li
 
W
,
Godzik
 
A
.
Cd-hit: a fast program for clustering and comparing large sets of protein or nucleotide sequences
.
Bioinformatics
.
2006
;
22
(
13
):
1658
59
. .

74.

Emms
 
DM
,
Kelly
 
S
.
OrthoFinder: phylogenetic orthology inference for comparative genomics
.
Genome Biol
.
2019
;
20
(
1
):
238
. .

75.

Katoh
 
K
,
Standley
 
DM
.
MAFFT multiple sequence alignment software version 7: improvements in performance and usability
.
Mol Biol Evol
.
2013
;
30
(
4
):
772
80
. .

76.

Camacho
 
C
,
Coulouris
 
G
,
Avagyan
 
V
, et al.  
BLAST+: architecture and applications
.
BMC Bioinf
.
2009
;
10
:
421
. .

77.

Minh
 
BQ
,
Schmidt
 
HA
,
Chernomor
 
O
, et al.  
IQ-TREE 2: new models and efficient methods for phylogenetic inference in the genomic era
.
Mol Biol Evol
.
2020
;
37
(
5
):
1530
34
. .

78.

Fan
 
Z
,
Wang
 
LY
,
Xiao
 
L
, et al.  
Lampshade web spider ectatosticta davidi chromosome-level genome assembly provides evidence for its phylogenetic position
.
Commun Biol
.
2023
;
6
(
1
):
748
. .

79.

Tang
 
H
,
Bowers
 
JE
,
Wang
 
X
, et al.  
Synteny and collinearity in plant genomes
.
Science
.
2008
;
320
(
5875
):
486
88
. .

80.

Yang
 
Z
.
PAML 4: phylogenetic analysis by maximum likelihood
.
Mol Biol Evol
.
2007
;
24
(
8
):
1586
91
. .

81.

Kosakovsky
 
PS
,
Poon
 
A
,
Velazquez
 
R
, et al.  
HyPhy 2.5-A customizable platform for evolutionary hypothesis testing using phylogenies
.
Mol Biol Evol
.
2020
;
37
(
1
):
295
99
. .

82.

Harris
 
RS
.
Improved Pairwise Alignment of Genomic DNA
.
Philadelphia
:
Pennsylvania State University
;
2007
.

83.

Kent
 
WJ
,
Baertsch
 
R
,
Hinrichs
 
A
, et al.  
Evolution's cauldron: duplication, deletion, and rearrangement in the mouse and human genomes
.
Proc Natl Acad Sci USA
.
2003
;
100
(
20
):
11484
89
. .

84.

Blanchette
 
M
,
Kent
 
WJ
,
Riemer
 
C
, et al.  
Aligning multiple genomic sequences with the threaded blockset aligner
.
Genome Res
.
2004
;
14
(
4
):
708
15
. .

85.

Hubisz
 
MJ
,
Pollard
 
KS
,
Siepel
 
A
.
PHAST and RPHAST: phylogenetic analysis with space/time models
.
Brief Bioinform
.
2011
;
12
(
1
):
41
51
. .

86.

Siepel
 
A
,
Bejerano
 
G
,
Pedersen
 
JS
, et al.  
Evolutionarily conserved elements in vertebrate, insect, worm, and yeast genomes
.
Genome Res
.
2005
;
15
(
8
):
1034
50
. .

87.

Yang
 
H
,
Wang
 
K
.
Genomic variant annotation and prioritization with ANNOVAR and wANNOVAR
.
Nat Protoc
.
2015
;
10
(
10
):
1556
66
. .

88.

Wang
 
K
,
Li
 
M
,
Hakonarson
 
H
.
ANNOVAR: functional annotation of genetic variants from high-throughput sequencing data
.
Nucleic Acids Res
.
2010
;
38
(
16
):
e164
. .

89.

Pineda
 
SS
,
Chaumeil
 
PA
,
Kunert
 
A
, et al.  
ArachnoServer 3.0: an online resource for automated discovery, analysis and annotation of spider toxins
.
Bioinformatics
.
2018
;
34
(
6
):
1074
76
. .

90.

Haney
 
RA
,
Matte
 
T
,
Forsyth
 
FS
, et al.  
Alternative transcription at venom genes and its role as a complementary mechanism for the generation of venom complexity in the common house spider
.
Front Ecol Evol
.
2019
;
7
:
85
. .

91.

Mistry
 
J
,
Finn
 
RD
,
Eddy
 
SR
, et al.  
Challenges in homology search: HMMER3 and convergent evolution of coiled-coil regions
.
Nucleic Acids Res
.
2013
;
41
(
12
):
e121
. .

92.

Roberts
 
SP
,
Harrison
 
JF
,
Dudley
 
R
.
Allometry of kinematics and energetics in carpenter bees (Xylocopa varipuncta) hovering in variable-density gases
.
J Exp Biol
.
2004
;
207
(
Pt 6
):
993
1004
. .

93.

Zhang
 
Y
,
Li
 
S
.
Genome of Octonoba sinensis
. V3.
Science Data Bank
.
2024
. .

94.

Zhang
 
Y
,
Shen
 
Y
,
Jin
 
P
, et al.  Supporting data for “A Trade-Off in Evolution: The Adaptive Landscape of Spiders without Venom Glands.”.
GigaScience Database.
 
2024
. .

Author notes

Yiming Zhang and Yunxiao Shen contributed equally to this work.

This is an Open Access article distributed under the terms of the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted reuse, distribution, and reproduction in any medium, provided the original work is properly cited.