#PAGE_PARAMS# #ADS_HEAD_SCRIPTS# #MICRODATA#

The beta-1, 4-N-acetylglucosaminidase 1 gene, selected by domestication and breeding, is involved in cocoon construction of Bombyx mori


Authors: Chunlin Li aff001;  Xiaoling Tong aff001;  Weidong Zuo aff001;  Hai Hu aff001;  Gao Xiong aff001;  Minjin Han aff001;  Rui Gao aff001;  Yue Luan aff001;  Kunpeng Lu aff001;  Tingting Gai aff001;  Zhonghuai Xiang aff001;  Cheng Lu aff001;  Fangyin Dai aff001
Authors place of work: State Key Laboratory of Silkworm Genome Biology, Key Laboratory of Sericultural Biology and Genetic Breeding, Ministry of Agriculture and Rural Affairs, College of Biotechnology, Southwest University, Chongqing, China aff001
Published in the journal: The beta-1, 4-N-acetylglucosaminidase 1 gene, selected by domestication and breeding, is involved in cocoon construction of Bombyx mori. PLoS Genet 16(7): e32767. doi:10.1371/journal.pgen.1008907
Category: Research Article
doi: https://doi.org/10.1371/journal.pgen.1008907

Summary

Holometabolous insects have distinct larval, pupal, and adult stages. The pupal stage is typically immobile and can be subject to predation, but cocoon offers pupal protection for many insect species. The cocoon provides a space in which the pupa to adult metamorphosis occurs. It also protects the pupa from weather, predators and parasitoids. Silk protein is a precursor of the silk used in cocoon construction. We used the silkworm as a model species to identify genes affecting silk protein synthesis and cocoon construction. We used quantitative genetic analysis to demonstrate that β-1,4-N-acetylglucosaminidase 1 (BmGlcNase1) is associated with synthesis of sericin, the main composite of cocoon. BmGlcNase1 has an expression pattern coupled with silk gland development and cocoon shell weight (CSW) variation, and CSW is an index of the ability to synthesize silk protein. Up-regulated expression of BmGlcNase1 increased sericin content by 13.9% and 22.5% while down-regulation reduced sericin content by 41.2% and 27.3% in the cocoons of females and males, respectively. Genomic sequencing revealed that sequence variation upstream of the BmGlcNase1 transcriptional start site (TSS) is associated with the expression of BmGlcNase1 and CSW. Selective pressure analysis showed that GlcNase1 was differentially selected in insects with and without cocoons (ω1 = 0.044 vs. ω2 = 0.154). This indicates that this gene has a conserved function in the cocooning process of insects. BmGlcNase1 appears to be involved in sericin synthesis and silkworm cocooning.

Keywords:

Insects – Evolutionary genetics – Domestic animals – Larvae – Genotyping – Protein synthesis – Silkworms – Silk

Introduction

Insects have evolved adaptations for each developmental stage and this has helped them become a successful group. Cocoons are an adaptation that protects insect pupae. For example, the wild silkworm (Bombyx mandarina) camouflages its pupa by using its silk to wrap mulberry leaves around the cocoon. Antherea pernyi and Antheraea yamamai spin cocoons with natural colors adapted to their environments for concealment. Cocoon adaptions help protect insects from predators and parasites. At least 17 insect orders have species with the ability to spin silk. Many insects construct cocoons during a specific developmental stage and this is usually the final instar [1]. Deciphering the genetic and evolutionary basis underlying cocoon construction will advance our understanding of the adaptations of these insects.

The silkworm (Bombyx mori) is the best example of an insect that spins cocoons. The silkworm cocoon is a capsule structure enclosed by silk. It provides a strong, breathable space that is waterproof. An intact cocoon is spun with a single continuous silk fiber that ranges between 500 m to 1500 m in length [2]. This silk fiber is mainly composed of fibroin and sericin proteins. Fibroin serves as the scaffold of the cocoon, while sericin acts as an adhesive to bind the silk fibroin. Cocoon spinning is a complex physiological behavior, involving synthesis, secretion, assembly, and spinning of the silk proteins. Silk research has mainly focused on silk assembly and spinning behavior [3, 4]. It is noteworthy that silk proteins, including sericin and fibroin, are the material basis of cocoons. An appropriate amount of silk protein is needed for cocoon construction. Studying the molecular basis of silk protein synthesis provides insight into the genetic and evolutionary basis of cocoon construction.

The silkworm is the only domesticated invertebrate producing economic quantities of silk. The domestication process began > 7500 years ago [5, 6]. Long-term domestication and breeding greatly improved the silk protein synthesis of the domestic silkworm compared to its wild counterpart (Bombyx mandarina). The cocoon shell weight (CSW), an important index for silk protein content, increased from 0.05 g to 0.5 g. Domestication and breeding also produced several silkworm strains with varying ability to synthesize silk protein. These strains can be used to study the genetic basis underlying the synthesis of silk protein. We used the silkworm as a model to study the genetic basis underlying silk protein synthesis and cocoon construction.

The ability to synthesize silk protein, measured by CSW, is a quantitative trait [2]. Several studies have explored the genetic basis of this trait using quantitative trait loci (QTL) mapping [711] and selection analysis [12, 13]. However, fine mapping and cloning of genes that affect silk protein synthesis has not been done. We previously identified a major locus on the 11th chromosome that controls CSW [11, 14]. In this study, we performed a genetic analysis based on selective genotyping and association analysis. We identified a gene (BmGlcNase1) that affects the synthesis of sericin and is involved in the formation of silkworm cocoons.

Results

Genetic analysis associates BmGlcNase1 with variation in CSW

Preliminary QTL analysis was made using IS-Dazao and 872B, a pair of silkworm strains with varied CSW, as parents. We mapped csw2, a locus that explained 15.07% of the variation of CSW in a segregating population, to the region bracketed by markers at 9.2 and 17.8 mega base pair (Mb) on the 11th chromosome (Fig 1A) [11, 14]. To precisely identify the mapping region, we performed a selective genotyping experiment. IS-Dazao and 872B were also used as parents. Two subgroups included 200 males with low CSW values and the another 200 males with high CSW values were selected from a backcross population containing 4325 individuals [11, 15]. Via chromosome walking, we identified four sequential indels with the highest linkage significance in the region between 10.74 and 11.01 Mb (Fig 1B). The genotyping result of these indels in the 400 individuals suggests that they were tightly linked.

Genetic analysis of <i>csw2</i>.
Fig. 1. Genetic analysis of csw2.
(A) Preliminary mapping based on selective genotyping. The genomic position of csw2 was first confirmed by genotyping the markers used in preliminary QTL mapping (Part 1 in S1 Table) [11]. The linkage significance (p) between each marker and the CSW was calculated by one-way ANOVA. The line with the ruler above represents the 11th chromosome; the grey color bar shows the negative logarithm of linkage significance of the corresponding chromosome section. Chr, chromosome; Mb, mega base pair. (B) Fine mapping of csw2 by selective genotyping. Each point in the frame represents an Indel marker (Part 2 in S1 Table) developed in csw2 and red points are the Indels with the highest linkage significance. The linkage significance (p) between each Indel and CSW was calculated by one-way ANOVA. (C) Association analysis based on germlines. One-way ANOVA was used to calculate the association significance between each Indel and CSW. The Bonferroni adjustment was used to obtain the threshold in association analysis. Each point represents an Indel marker and red points are the Indels with significance lower than the threshold (associated Indel). Numbers below indicate the genomic location of Indels. (D) Diagram of the association region with red highlights. Line with the ruler above represents the genomic region containing BmGlcNase1. The black frame represents the BmGlcNase1 gene and the arrow shows the gene direction. The solid red line shows the genomic location of the associated Indels.

We conducted an association analysis using the germplasm strains conserved in the silkworm gene bank. Ninety-nine Indel markers were developed in this region and 95 germlines with normal viability and distant kinship were selected as the association analysis population. Selection was based on germplasm conservation and breeding experience (S8 Table). This population exhibited substantial CSW variation. The XiaFang strain had the highest CSW (0.383 g) while strain 19–460 had the lowest CSW (0.078 g). By genotyping the indels in the association analysis population, we identified a cluster of indels associated with CSW (Fig 1C). Among these, Indel-10.762 had the highest association significance. This indel is located 1.2 kilo base pairs (kb) upstream of the transcription start site (TSS) of gene BMgn011646. This suggests that BMgn011646 may be involved in the control of CSW (Fig 1D). Functional annotation showed that BMgn011646 encodes β-1,4-N-acetylglucosaminidase 1 (BmGlcNase1).

Expression of BmGlcNase1 is correlated with silk gland development and CSW

The associated Indel is located upstream of the TSS of BmGlcNase1, indicating that the transcription level of this gene may be associated with CSW. Therefore, we investigated the expression pattern of BmGlcNase1. The spatial expression pattern shows that it has a relatively specific high expression in the middle silk gland (S1A Fig), but no expression was detected in the posterior silk gland. Silk glands are silkworm tissues that synthesize silk protein. They consist of anterior, middle, and posterior parts. The middle and posterior silk glands synthesize sericin and fibroin proteins, respectively. The volume study of parental silk glands showed that 872B, the parent with higher CSW, had a larger silk gland compared to IS-Dazao, the parent with lower CSW (S1B Fig). We determined the temporal expression profile of BmGlcNase1. BmGlcNase1 was mainly expressed in the middle and late stages of embryonic development and in the fifth instar which are key stages for silk gland development (Fig 2A). Silk gland cell division and active endomitosis during these stages determines the number and size of silk gland cells and thus the final silk gland volume. The expression pattern of BmGlcNase1 was highly associated with silk gland development. In the fifth instar, the daily transcription level was highly associated with the silk gland volume (Fig 2A). These results indicate that BmGlcNase1 is involved in silk gland development.

Expression pattern of <i>BmGlcNase1</i>.
Fig. 2. Expression pattern of BmGlcNase1.
(A) Correlation between silk gland development and temporal expression of BmGlcNase1. Dot plot shows the relative expression level of BmGlcNase1 in each sample. E3-E216, the embryo at 3–216 hours after incubation; Hatch, the new hatched larvae; I_F, feeding phase in the 1st instar; I_M, molting phase in the 1st instar; I-V, the 1st to the 5th instar; V0-V7, start to the 7th day of the 5th instar; the height of the green bar represent the volume of silk gland. Each point represents the average of 3 replicate and the error bar shows the standard deviation (SD.). (B) Association analysis between the expression level of BmGlcNase1 and CSW of varied silkworm strains. Each point represents the average expression level of BmGlcNase1 in the silk gland of one silkworm strain. The grey points represent strains with relatively lower CSW, while the red points represent these with higher CSW. Three replicates were performed in this experiment with silk glands from three larvae in each replicate. Linear regression analysis was used to determine the correlation between the expression level of BmGlcNase1 and CSW. Two tailed Student’s t-test was used to compare the expression difference between the silkworm with lower and higher CSW. ** represents a significant difference at the 0.01 level.

We surveyed the expression of this gene in the middle silk gland of additional strains with varied CSW. Thirty-six strains were allowed to feed up to the last day of the fifth instar (S8 Table). The middle silk gland of each was then dissected at this time because the gene exhibited high expression on this day. Transcript level analysis showed a significant association between gene expression and CSW (r = 0.7051) (Fig 2B; p<0.0001). Based on the CSW performance of each strain, we divided the strains into two groups with low and high CSW. We then compared the expression level of each group. There was significantly higher expression in the group with higher CSW than in the low CSW group (Fig 2B). This suggests that BmGlcNase1 may play an important role in silk gland development and may be associated with the CSW of silkworms.

A cluster of noncoding variants was associated with BmGlcNase1 expression and CSW

We used rapid amplification of cDNA ends (RACE) to obtain the full cDNA of parents, IS-Dazao and 872B, and analyzed their gene structure. BmGlcNase1 contains 11 exons; the open reading frame (ORF) starts at the 3rd exon and ends at the 11th exon (Fig 3A). Two types of transcript were detected in both strains, IS-Dazao and 872B. Compared with the type I transcript, there was one 106 bp fragment insertion in the second exon of the type II transcript. However, there was no variation in the ORF region between them (Fig 3A). The sequence of the type I transcript in Is-Dazao and 872B was then compared. There were 33 SNPs in the full cDNA and, among them, three were located in the 5’ untranslated region (UTR), and another 30 SNPs were in the ORF (Fig 3A). Amino acid (AA) analysis showed that only the first SNP in the ORF led to the substitution (Gly-Arg). Protein structure analysis suggested that this substitution was in the signal peptide, but no signal peptide characteristic difference was detected between the coding products of IS-Dazao and 872B. These results indicated that variation in the BmGlcNase1 transcription level might be associated with the variation in CSW.

Sequence variation of <i>BmGlcNase1</i>.
Fig. 3. Sequence variation of BmGlcNase1.
(A) cDNA sequence and gene structure of BmGlcNase1. The wireframes represent the structure of the two BmGlcNase1 transcripts, of which the blue frames indicate the exons in open reading frame (ORF) and the blacks indicate the untranslated exons. The horizontal lines indicate the introns. The grey frame in transcript 2 shows the inserted fragment in this transcript. Red vertical bars indicate the synonymous SNPs and the yellow bar shows the synonymous SNP. The ruler below shows the position of each part of BmGlcNase1 on the 11th chromosome. The 107- (63591–86430) indicate the 10,763,591 to 10,786,430 nt of the 11th chromosome. (B) Genomic sequence comparison of BmGlcNase1 between parents; columns represent the number of variants in this region. The combination of wireframes and ruler showed the gene structure of BmGlcNase1. The grey histogram shows the number of SNP/Indels in each 100 bp bin along the genomic region covering BmGlcNase1. (C) Association analysis between upstream indels and CSW. Each point indicates the association significance of one Indel upstream of BmGlcNase1, of which the red points show those with the highest significance. The paired linkage disequilibrium (r2) values are shown at the bottom. (D-E) Comparison of BmGlcNase1 expression and CSW between strains with L and H haplotypes. Two tailed Student’s t-test was used to compare the two indexes between the two groups. Error bar, SD. ** represents a significant difference at the 0.01 level.

We determined the genomic sequence of both parents. In total, 30 kb of the genomic region covering BmGlcNase1 was sequenced. A large number of variations were detected in this region (2,242), including 1932 SNPs and 309 indels. These variations were mainly enriched in the upstream and intron regions, rather than distributed evenly (Fig 3B). The association analysis showed that indels in the upstream region were associated with CSW. We thus genotyped indels in the 1.5 kb upstream region of TSS in the association analysis population and calculated their association significance with CSW. Indels in the region between 1631 bp and 1264 bp exhibited the highest association with CSW (p = 1.59e-08) (Fig 3C) (S5 Table). Genotyping results of multiple strains showed that these associated indels were grouped into two haplotypes (L/H). CSW significantly varied between the two haplotypes (Fig 3D). The BmGlcNase1 gene exhibited significantly higher expression in H haplotype strains than that in L haplotypes (Fig 3E). These results suggest that variants in this region may contribute to the expression variation of BmGlcNase1 and the CSW differences among silkworm strains.

Functional validation showed BmGlcNase1 has a significant effect on sericin synthesis

To study the effect of BmGlcNase1 on CSW, we used transgenic technology to elevate and knockdown its expression in the middle silk gland (S2 Fig) by transgenic technology. Reverse PCR indicated that the insertion site of the overexpression and interference constructs is located in the intergenic region, which rules out the disruption of other silkworm genes (S2B and S2E Fig). To avoid interference with the feeding environment, 150 normal and 150 transgenic newly hatched larvae were mixed for feeding until the wandering stage, during which the larvae search for places to spin their cocoons. During the wandering stage, the larvae were differentiated for cocoon construction by fluorescence screening. The transcript content in the middle silk gland of the over-expressed (OE) group was twice that of the control (Fig 4A). In the RNAi group, the transcription level of BmGlcNase1 was significantly reduced (Fig 4B). The CSW of the OE group was significantly higher than that of the control, with 10.9% increase in females (p = 0.0005) and 8.5% in males (p = 0.0038) (Fig 4A). The CSW was significantly decreased in the RNAi group, with 7% reduction in both males (p = 0.0056) and females (p = 0.0017) (Fig 4B). We compared the volume of silk glands between the transgenic strain and the control group at the beginning of the wandering stage. The results showed that compared with the control, the silk gland volume of the overexpressed strain was slightly but significantly increased (p = 0.47). Although the silk gland volume of the knock-down strain is smaller than the control group, it is not significant (0.053) (S2C and S2F Fig). Spatial expression pattern analysis showed that BmGlcNase1 exhibited relatively high expression in the middle silk gland, while the expression level in the posterior silk gland was below the level of detection. This suggests that BmGlcNase1 can specifically affect the synthesis of sericin. To determine whether its ectopic expression can affect fibroin synthesis, we constructed an ORF-containing vector driven by the promoter of the fibroin heavy chain gene and obtained the positive strain (S3 Fig). The transcription level was significantly elevated (9×) in the posterior silk gland (p<0.01) while the CSW showed no difference between the ectopic expression and the control group (S3 Fig).

Functional analysis of <i>BmGlcNase1</i>.
Fig. 4. Functional analysis of BmGlcNase1.
(A-B) Overexpression and knockdown of BmGlcNase1 and the effect on CSW. The above histograms in both (A) and (B) show the expression level of BmGlcNase1 in control and over-expression groups and the lows show the CSW of the two groups. OE, over-expression. For expression level determination, three replicates were performed with silk glands from three larvae in each sample. (C-D) Effect of BmGlcNase1 overexpression and knockdown on the synthesis of sericin and fibroin. Two tailed Student’s t-test was used for comparison. Error bar, SD. N.S., no significant; * and ** represented significant differences at the 0.01 level.

These results suggest that BmGlcNase1 can specifically affect sericin synthesis, while it does not influence fibroin synthesis. To confirm this, we surveyed the sericin content in the cocoons of the OE, RNAi, and control groups. Sericin was significantly elevated in OE cocoons, with a 13.9% and 22.5% increase in females (p<0.01) and males (p<0.01), respectively. However, the fibroin content in the two groups was similar (Fig 4C). Compared to the control, sericin in the RNAi group was significantly lower (p<0.01), with 41.2% and 27.3% reductions in females and males, respectively (Fig 4D). The fibroin content also exhibited a significant reduction (p = 0.038) in the females of the RNAi group, while the fibroin reduction in males (p = 0.067) was not significant. These data indicate that sericin is an essential component of cocoons and indispensable for normal cocoon formation. A reduction of sericin in cocoons will reduce the production of fibroin and limit silk secretion. In summary, the expression of BmGlcNase1 affects the synthesis of sericin and cocoon formation.

Associated variants selected during domestication and breeding

CSW is the most important index for silk yield and the main selection target during silkworm breeding. Genes or regions associated with CSW may also be targets of selection. Therefore, we studied the selection signals in the genomic region covering BmGlcNase1, and we focused on associated regions. For this study, 20 wild silkworms (Bombyx mandarina), 30 local varieties, and 30 breeding strains were collected (S8 Table) to detect the selection signals from domestication and breeding. Using long fragment amplification, we determined the genomic sequences covering BmGlcNase1 from these strains. We created a balanced mix of these strains to construct three DNA pools for NGS. We obtained 1.85, 1.89, and 2.31 -giga base pair (Gb) of clean sequencing data for the wild, local, and breeding pools respectively, which resulted in a depth of more than 60,000×. Based on this, the pooled heterozygosity (Hp) was calculated. Hp in the wild silkworm pool was significantly higher than in the other two silkworm pools, with an average of 0.256 for wild strains. Hp was 0.113 and 0.120 for the local varieties and breeding strains, respectively (Fig 5A). This indicates that strong selection occurred in this region during domestication. Five regions, including the associated region upstream of TSS, were identified with greatly reduced heterozygosity. Hp of this region was 0.056 and 0.043 in the local variety pool and breeding pool, respectively. These values are much lower than values for the total region, suggesting strong selection. The Hp of the flanking region was much higher than that of the associated site in all three of the pools. This suggests that the reduction of Hp in the local and breeding pools is not the result of a hitchhiking effect.

Selection analysis of <i>BmGlcNase1</i>.
Fig. 5. Selection analysis of BmGlcNase1.
(A) Pooled heterozygosity of the genomic region of BmGlcNase1. The black, green and red lines represent the heterozygosity of wild silkworm (Bombyx mandarina), local varieties and breeding strains of Domestic silkworm (Bombyx mori), respectively. Numbers under the horizontal axis indicate the position relative to TSS. The thick lines below highlight the region with heterozygosity reduction in domesticated silkworms compared with the wild silkworms and the red shows the associated region upstream of BmGlcNase1. The wireframes indicate the structure of BmGlcNase1. (B-C) Sliding window analysis of nucleotide diversity (π) and of the sequenced region in wild, local, and breeding silkworm strains. The lines with circles, squares, and triangles represent wild silkworms, local varieties, and breeding strains, respectively. The black and red asterisks indicate the significance level that Tajima’D deviates from 0 is lower than 0.05 and 0.01, respectively. (D) Frequency of the haplotypes at the associated region in wild, local, and breeding strains. The yellow and green region indicate the frequency of H and L haplotype. (E) Ratio of sericin to CW in wild, local, and breeding strains. Red, orange, and green represent the ratio of sericin to cocoon weight of wild silkworm, local and breeding silkworm strains. For wild silkworm, six individuals, including three males and three females, were surveyed. For domesticated silkworms, 12 individuals, including 6 males and 6 females were surveyed. Error bar is one SD.

We sequenced the associated region in wild and domesticated silkworm strain via PCR amplification. We obtained the sequence of a 750 bp region around the associated site in 12 wild, 22 local, and 26 breeding strains (S1 File). For the entire sequence region, the nucleotide diversity (π) was 0.16419 in the wild silkworm strains, 0.09161 in local strains, and 0.04462 in breeding strains. This illustrates a gradual reduction of nucleotide diversity from wild to local and from local to breeding silkworm strains (Table 1). The sliding window analysis of the whole sequence suggests that the biggest difference among the three populations is within the associated region (Fig 5B). A neutral test based on sliding window analysis showed that Tajima’s D and Fu and Li’s F values are both lower than 0 in domesticated silkworm strains, and values were particularly low in breeding strains (Fig 5C and Table 1). These results indicate that the associated region was subject to strong selection in domestication, especially in the breeding strains.

Tab. 1. Nucleotide diversity and neutral tests.
Nucleotide diversity and neutral tests.

The local varieties and breeding strains exhibited opposite haplotype patterns for the associated region (S5 Table). The ratio of haplotype-L in 30 local varieties was 78.33%, while the ratio of haplotype-H was 86.67% in the breeding populations (Fig 5D). This suggests that the two haplotypes were successively purified during domestication and breeding. The association between BmGlcNase1 and sericin synthesis suggests that sericin synthesis was reduced during silkworm domestication but elevated during the improvement stage. To test this inference, we studied the ratio of sericin to cocoon weight (summation weight of cocoon shell and pupae, CW) of several wild silkworms, local varieties, and breeds. Compared to the wild silkworm, the relative sericin content in local varieties was slightly decreased (3.13% vs. 3.24%). However, this ratio was significantly elevated in breeding strains during the improvement stage compared to local strains (6.67% vs. 3.13%, p<0.01) (Fig 5E). These results suggest that the associated region upstream of BmGlcNase1 was subjected to selection during domestication and breeding. In the domestication process, haplotype-L, which is related to lower CSW, was purified. This resulted in lower sericin synthesis in local varieties. In the breeding process, haplotype-H, related to higher CSW, was selected. This led to higher sericin content in breeding silkworm strains.

GlcNase1 was differentially selected between insects with, and without cocoons

We demonstrated that BmGlcNase1 influences silkworm cocoons by affecting sericin synthesis. Orthologues of BmGlcNase1 are also expressed in the silk gland of wild silkworms and other cocoon spinning insects of Bombycoidea [1617]. To determine the relationships between GlcNase1 and cocoon formation in other insects, we studied the presence of differential selection on GlcNase1 in insects with and without cocoons by estimating the ratio (ω) of non-synonymous to synonymous substitution rates using a likelihood method. We selected 3 Lepidoptera and 3 Hymenoptera species that form cocoons and six species, lacking cocoons, that also belong to Lepidoptera and Hymenoptera (S6 Fig). Assuming that the GlcNase1 sequences from all of the branches in the tree of the 12 taxa have the same ω (model A in Table 2), we estimated that ω = 0.06912. Next, we tested whether a model that allows different ω values for insects, with and without cocoons (model B in Table 2), fits the data significantly better than model A. In model B, the ω for insects with cocoons was 0.04843, while the ω for insects lacking cocoons was 0.15417. Model A and model B were significantly different (p value = 0.0398), suggesting that the selection pressure on GlcNase1 is different between the selected insects with cocoons and those lacking a cocoon. These results suggest that GlcNase1 might have been subject to strong purifying selection during the evolution of the cocoon trait.

Tab. 2. Likelihood ratio tests of selective pressures on Glcnase1 in insects.
Likelihood ratio tests of selective pressures on <i>Glcnase1</i> in insects.

Discussion

Potential roles of BmGlcNase1 and hexosamine flux in silk protein synthesis

BmGlcNase1 encodes the β-1,4-N-acetylglucosaminidase 1 protein, which is responsible for hydrolyzing N-linked hexosamines on some proteins. Hexosamino-glycosylation is a type of protein with post-translational modification. Hexosamino-glycosylation includes N-linked and O-linked types and both types are key steps of the hexosamine signaling pathway [18]. This pathway is assumed to be a nutrient sensor [19]; the concentrations of UDP-GlcNAc are sensitive to glucose titer. UDP-GlcNAc content could influence the hexosamine modification of proteins and then regulate cell proliferation [20, 21] and oncogenesis [22, 23]. Intracellular hexosamine flux can affect the N-glycan number and the degree of branching in surface glycoprotein via cooperation with the Golgi pathway [24]. Interestingly, the receptors promoting growth, such as the EGF receptor (EGFR), harbor more N-glycan sites, while the receptors inhibiting growth possess fewer N-glycan sites. The former shows a hyperbolic response when intracellular UDP-GlcNAc content increases, while the latter exhibits a switch-like response. Thus, elevated intracellular UDP-GlcNAc content will first activate the growth-promoting receptors and facilitate the growth of cells and organs. BmGlcNase1 has no transmembrane structure, which indicates that it may be a nucleocytoplasmic protein that hydrolyzes GlcNAc from nucleocytoplasmic glycoproteins. This may elevate the hexosamine flux, which would activate growth-promoting receptors to facilitate the development of the silk glands. However, N-hexosamine studies in insects have focused on chitin metabolism. For example, genes in the acetylglucosaminidase family have been reported to degrade chitin and were therefore termed chitinases. We genome widely identified the acetylglucosaminidase genes in silkworm. In addition to BmGlcNase1, the results revealed nine additional genes that were annotated as acetylglucosaminidase. These genes are expressed in various tissues where chitin is absent, such as fat body, testes, ovaries, and hemolymph (S4 Fig). This suggests that they have other functions besides chitin metabolism. Among them, BmGlcNase1 was highly expressed in the middle silk glands. Considering the functions of N-glycan related to cell proliferation, oncogenesis, and protein synthesis in mammals, BmGlcNase1 and hexosamine flux might also be involved in regulating the proliferation of silk gland cells, the development of the middle silk gland, and sericin protein synthesis.

Selection pattern of BmGlcNase1 and its effect on sericin content in cocoons

Selection analysis showed that haplotype-L and haplotype-H were purified during domestication and breeding, respectively. Domestication resulted in a decrease in the ratio of sericin to CW, and breeding resulted in an increase in this ratio. During the domestication and breeding process, easy degumming and high reel-ability (% of silk that can be reeled) of cocoons were breeder targets. The sericin content of cocoons affects the silk reeling index. Excessive sericin in cocoons will increase energy consumption and decrease the reel-ability because of the increased adhesion among the silk fibers [25]. Thus, selection has led to a reduction of sericin in the cocoon. We calculated the ratio of sericin in the cocoon shell of wild, local, and breeding strains. The ratio was much lower in local varieties than in wild silkworms (33.7–26.8% decrease), but the reduction was somewhat less when comparing local strains to breeding strains (26.8–23.9%) (S5 Fig). Early domestication may have rapidly reduced the ratio of sericin to the cocoon shell. When this ratio was reduced to a certain level, breeding selection was unable to further decrease the sericin content. During the latter process, the cocoon shell ratio (CSR) (ratio of cocoon shell weight to cocoon weight), greatly increased (0.3–0.1). This led to increased silk protein synthesis efficiency, including sericin. These data may explain why breeding silkworm strains have a higher sericin synthesis efficiency than local strains. They also indicate that a certain amount of sericin is necessary for silk fiber spinning and cocoon construction.

Contribution of sericin to silkworm cocoon construction

Functional validation of BmGlcNase1 demonstrated that increases in sericin have no effect on fibroin synthesis, while decreases in sericin may slightly reduce the content of fibroin in cocoons. This shows that sericin is an essential component of silkworm cocoons and the fibrosis of silk proteins is a critical process for cocoon spinning. This process is accompanied by the dehydration of fibroin and the exchange of ions. Sericin plays a critical role in this process. It absorbs water lost during the dehydration of fibroin and it also mediates the ion exchange during the fibrosis of fibroin. Sericin promotes exocrine processes by avoiding premature fibrosis of silk fibroin and protecting the wall of spinnerets [26]. When wild silkworm cocoons are spun, sericin-coated silk fibers are used to wrap a layer of mulberry leaves around the cocoon for camouflage. Thus, sericin is needed for the fibrosis of silk fiber and, the formation and concealment of silkworm cocoons.

The evolution of Glcnase1 and cocoon construction in insects

Insects that form cocoons occur in many evolutionarily distant orders suggesting that this trait may be the result of convergent evolution. The selective pressure on GlcNase1 in insects with cocoons and insects without cocoons would be significantly different, suggesting that GlcNase1 may be involved in the cocoon formation of many insect species. The ω (dN/dS) for both types of insects is less than 0.25. This indicates that it was subjected to purifying selection in all of the surveyed insects. Homologs of BmGlcNase1 have multiple functions. Many members of this gene family are involved in chitin hydrolysis [27, 28]. Chitin is the main component of the exoskeleton, and the coding sequence conservation of GlcNase1 indicates that it may be involved in chitin metabolism in insects not forming cocoons. GlcNase1 may have evolved new functions related to cocoon formation in insects with cocoons. In addition, the phylogeny constructed by the coding sequence of GlcNase1 showed similarity to the surveyed species, rather than a clustered distribution based on species with, or without, cocoons (S7 Fig). This suggests that GlcNase1 independently evolved cocoon-related functions in different insects. Convergent evolution can emerge in two ways: parallel genetic evolution and collateral genetic evolution. Parallel genetic evolution refers to mutations that arise and spread in independent lineages, and collateral genetic evolution refers to mutations that either arise from a common ancestor or result from hybridization [29]. The phylogeny of the focal genes will generally be congruent with that of the species in parallel evolution [30, 31], while the phylogeny of the focal genes will be inconsistent with the phylogenetic history of species in collateral evolution [3234]. Thus, the phylogeny of GlcNase1 and its role in insect cocooning may be an example of parallel genetic evolution.

We investigated whether the potential neofunctionalization of GlcNase1 results from convergent/parallel sites. Sequence alignment showed that the Glycohydro_20b2 domain at the N’ end was poorly conserved and no specific conserved amino acid sequence was detected. The Glycohydro_20 domain at the C’ end shows a higher degree of conservation in all insects (S8 Fig). The amino acids at 215, 255, 484, and 609 sites were highly conserved in insects with cocoons (S7 Fig), while they showed a high diversity in insects without cocoons. This was especially true for the amino acid variation at sites 255 and 484 located in the Glycohydro_20 domain. This suggests that the conservative amino acid sites in insects with cocoons may be involved in the neofunctionalization of GlcNase1.

Conclusion

We demonstrated that the gene BmGlcNase1 affects the synthesis of sericin. The gene was selected by domestication and breeding. Selection significantly reduced the sericin content in domestic silkworm cocoons compared to the content in its wild ancestor. These data, combined with the functional conservation of GlcNase1 in insects with cocoons, suggest that BmGlcNase1 is involved in the cocoon process of silkworms and in other insects with cocoons.

Materials and Methods

Silkworm samples

We used two populations of silkworms for genetic analysis. An extreme phenotype population for selective genotyping and germlines for association analysis. To construct the former population, two silkworm strains (IS-Dazao and 872B), which were used in preliminary QTL analysis, were selected as parents to produce a backcross population, IS-Dazao♀×(IS-Dazao×872B)F1♂. The individuals in this population were reared on fresh mulberry leaves to the cocoon stage at 25°C with a 12:12 h (L:D) photoperiod. At the eye coloring stage, the cocoons of each individual were opened to distinguish the sexes based on features of the pupal abdomen, and the cocoon shell weight (CSW) was measured for the two sexes. Based on the phenotype data, male individuals with extreme CSW values were selected from the progeny of a single female moth. The number of individuals with extremely high and extremely low CSW (H sub-population and L sub-population, respectively) made up approximately 10% of the total males in the mapping population. The two sub-populations were used as the extreme phenotype populations for selective genotyping. For association analysis, 95 germlines were chosen from the silkworm gene bank of Southwest University (Chongqing, China) with three selection criteria: the strains had normal viability, geographical or systematical source silkworm strains should be as diverse as possible to avoid the possible kinship and they should exhibited substantial CSW variation (S8 Table). These silkworm strains were reared on fresh mulberry leaves at 25°C with a 12:12 h (L:D) photoperiod. At the pupal stage, 100 males and 100 females were selected from each germline for CSW determination using the method detailed above. Genomic DNA of individuals from the two populations was obtained via a standard phenol-chloroform extraction method [35].

Genotyping and statistics for genetic analysis

To narrow the mapping region by selective genotyping, we developed 18 indel markers in the preliminary mapping region (S1 Table). These were genotyped via PCR amplification and polyacrylamide gel electrophoresis (PAGE). The linkage significance between these indels and CSW was calculated by one-way ANOVA. After genotyping analysis, the mapping region was narrowed to a 270 kb region. We then developed 99 indel markers in this region for association analysis. These were also genotyped by PCR amplification and PAGE. The association significance between indels and CSW was calculated using one-way ANOVA. Primers used to amplify these indels are listed in S2 Table.

Sampling and expression pattern analysis

To investigate the expression level of BmGlcNase1, various silkworm tissues were dissected from IS-Dazao larvae on the third day of the 5th instar and placed into cold normal saline (NS). The dissected tissues included muscle, testis, wing disc, midgut, malpighian tubules, anterior-middle silk gland, fat body, ovary, head, posterior silk gland, integument, and hemolymph. For temporal expression pattern analysis, 17 samples were selected beginning at 3 h after incubation and ending at day seven of the 5th instar. At the embryonic stage, eggs laid by one female moth were divided into three groups and preserved in three tubes with TRIzol (Invitrogen, USA). From the first day of the first instar to the molting stage of the 4th instar, whole larvae were preserved. At the 5th instar, the anterior-middle silk glands were dissected from larvae each day. For the association analysis between expression and CSW variation, 36 germlines (S8 Table) with varied CSW values were selected and reared to the final day of the fifth instar. Middle silk glands were dissected on this day from each germline. All of these samples were preserved in cold TRIzol. Total RNA was extracted and purified using TRIzol (Invitrogen, USA) according to the manufacturer instructions. Then, the first strand cDNA was generated via reverse transcription using a PrimeScript RT Reagent Kit (Takara, Japan) according to the supplier instructions. The correlation between the expression level of BmGlcNase1 in the silk glands of each strain and CSW was analyzed by linear regression. Analysis of the difference in expression of BmGlcNase1 between different groups was conducted using Student’s t-test. For each strain, three replicates were performed and each replicate contained middle silk glands from three larvae. Primers used for real-time PCR and RT-PCR are listed in S6 Table.

Amplification and sequencing

To study the sequence variation between parents for the full-length cDNA of BmGlcNase1, we performed cDNA cloning and rapid amplification of cDNA ends (RACE). The total RNA was extracted from the middle silk gland of parents on the final day of the 5th instar according to the procedures detailed above. For cDNA cloning, four pairs of primers were designed (S3A Table). Two strands of primer for double-end RACE amplification were also designed based on the inner cDNA sequence. The RACE amplification was performed using the instructions of the Marathon cDNA Amplification Kit (Takara). All of these products were then inserted into a pM19-T vector (Takara, Japan) and sequenced. To obtain the genomic sequence of BmGlcNase1, we designed 21 pairs of primers to successively amplify the 30 kb genomic region. These products were also inserted into a pM19-T vector (Takara, Japan) and sequenced. All of the primers used for sequencing were listed in S3B Table. Primers for association analysis of the Indels in upstream region of BmGlcNase1 were listed in S4 Table.

Transgenic function analysis

Transgenic-based over-expression and knock-down were used to validate the function of BmGlcNase1. The Ser1 and FibH promoters used to drive the expression of BmGlcNase1 in middle and posterior silk glands were synthesized based on the sequences in Wang et al. [36] and Zhao et al. [37]. The primers, ORF/RNAi fragment, and SV40 termination sequence were then sequentially linked together with the restriction sites NotI and kpnI. The expression core was then inserted into the PiggyBac transposon with dsRed and EGFP, both of which were driven by 3 × p3 promoter. Primers used for vector construction were list in S7 Table. The over-expression and knock-down constructs, accompanied with a helper vector, were then, respectively, injected into the newly laid non-diapause eggs of IS-Dazao (incubated at 16°C) and 305, a multivoltine silkworm strain with a relatively high CSW. The new G1 larvae were then subjected to fluorescence screening to identify the positive G1 individuals. Through back-crosses and self-crosses with control lines, we established seven overexpression, two knock down and two posterior silk gland ectopic expression lines. By detecting the expression level of the BmGlcNase1 gene of each line, we selected the line with the highest overexpression and interference effects for the following experiment. Inverse PCR was then performed to identify the insertion sites of transgenic silkworms according to the method of Wang et al. [36]. To accurately evaluate the CSW of transgenic strains and the corresponding control, we reared 150 positive and 150 control individuals together to remove interference from environmental differences. On the last day of the 5th instar, fluorescence screening was used to distinguish the positive individuals from controls. The CSW values of these individuals were then measured using the method detailed above. The fibroin content of cocoons of various strains was determined by degumming based on the sodium carbonate solution reported by Dai et al. [38]. Cocoons were heat-dried at 70°C for 7 h and weighted. Then the dried cocoons were boiled in 150 mL 0.5% sodium–carbonate solution at 100°C for 2 h, followed by washing 45 min with deionized water. Degummed cocoons were dried at 65°C for 24 h and weighted. The corresponding sericin content was calculated by the following formula: Weight(Sericin) = CSW-Weight(Fibrion).

Selection analysis in domestication and breeding

A total of 20 wild (B. mandarina), 30 local, and 30 breeding silkworm strains (S8 Table) were selected for pooled heterozygosity investigation. To capture the genomic fragment in these strains, four pairs of primers with an average product length of 8327 bp were designed. These were amplified via LA-Taq (Takara, Japan) under the conditions specified in the instructions. This product was purified by the Gel Extraction Kit (Omega, USA) using manufacturer instructions. Then, the products were equally mixed to produce the wild, local, and breeding pools. The Illumina sequencing libraries were constructed and then subjected to sequencing under the paired-end mode on Illumina HiSeq 2500 (Novogene, Beijing, China). We generated approximately 1.85, 1.89, and 2.31 Gb raw reads for wild, local, and breeding pool, respectively. The genomic sequence of BmGlcNase1 in IS-Dazao obtained in this research was used as a reference. The raw data were then subjected to joint trimming, low-quality data filtering, comparison with the reference genome, and variant calling. Based on this, the Hp of each sample was calculated based on the method of Erik [39]. Multi-sequence alignment of nucleic acid sequences was conducted by using mega software (muscle algorithm). Through DNAsp v6 software [40], the nucleotide polymorphism (pi), Tajima ‘s D value and Fu and Li ‘s F value were calculated by sliding window (window size: 100 bp, sliding step size: 25 bp). Sericin content in the cocoons of wild, local, and breeding strains was also determined using the method above. The weight ratios of sericin to cocoon and to cocoon shell were calculated using the following equations:



Selection pressure analysis

To estimate selection pressures on the GlcNase1 gene in cocooning insects and non-cocooning insects, we selected six cocooning insect species and six non-cocooning insect species (S9A Fig). The GlcNase1 of Bombyx mandarina was from Silkbase (http://silkbase.ab.a.u-tokyo.ac.jp/cgi-bin/index.cgi). Transcript sequence of Antheraea pernyi GlcNase1 was from the transcriptome Shotgun Assemblies in Dong et al. [17]. Genome and proteome data of the other species were downloaded from NCBI by the accessions as follows: GCA_000572035.2 (Microplitis demolitor), GCA_000806365.1 (Fopius arisanus), GCA_001412515.3 (Diachasma alloeum), GCA_002156985.1 (Helicoverpa armigera), GCA_000836235.1 (Papilio xuthus), GCA_009193385.2 (Nasonia vitripennis), GCA_000599845.3 (Trichogramma pretiosum), GCA_001465965.1 (Polistes dominula) and GCA_001266575.1 (Operophtera brumata). The protein sequence in each species was searched by pblast, and then the corresponding transcript sequence was obtained by tblastn. Each transcript was mapped to the corresponding genome by blastn to obtain its genomic location information and determine the copy number of GlcNase1 in each species. Real-time PCR was used to compare the DNA content of the Antheraea pernyi GlcNase1 relative to the housekeeping genes, eukaryotic translation initiation factor 4A (ETI-4) and Actin3 (S6 Table), to determine the genome copy number (S9B Fig). After confirmation of the GlcNase1 copy number of each species, the transcript sequences were aligned using ClustalW (codon) followed by manual adjustments (S6 Fig). A maximum-likelihood method was used to estimate the selective pressures using branch models in codeml of the EasyCodeML package [41]. The ratio of non-synonymous to synonymous substitution rates, termed ω, was used to estimate the mean selection pressures on different branches of the tree. We first estimated ω across the tree under a one-ratio model. Then, we estimated an independent ω value for cocoon forming and non-cocoon forming branches under the free-ratio model. The maximum-likelihood tree of GlcNase1 was constructed by MEGA7.1.0 [42] with 1000 bootstrap runs.

Supporting information

S1 Fig [a]
Expression pattern of in various tissues at the third day of the 5 instar larva.

S2 Fig [pdf]
Screening of transgenic positive lines and detection of insertion sites.

S3 Fig [a]
Ectopic expression of in the posterior silk gland.

S4 Fig [pdf]
Expression pattern of genes of family of silkworm.

S5 Fig [pdf]
Ratio of sericin to cocoon shell weight in wild silkworm, local and breeding domesticated silkworm strains.

S6 Fig [pdf]
Nucleotide sequences of insects used for selection pressure analysis.

S7 Fig [pdf]
Phylogenetic analysis of selected insects.

S8 Fig [pdf]
Protein sequence comparison of in insects with cocoons and without cocoons.

S9 Fig [a]
Information about the orthologues in other insects.

S1 Table [xlsx]
Primers used for selective genotyping.

S2 Table [xlsx]
Primers used for association analysis.

S3 Table [xlsx]
Primers used for sequencing.

S4 Table [xlsx]
Primers used for association analysis of the Indels in the upstream region of .

S5 Table [xlsx]
Genotype of variants at the upstream of .

S6 Table [xlsx]
Primers for real time PCR and RT-PCR.

S7 Table [xlsx]
Primers used for plasmid construction.

S8 Table [xlsx]
Information on the silkworm samples used.

S1 File [txt]
Sequence information.


Zdroje

1. Sutherland TD, Young JH, Weisman S, Hayashi CY, Merritt DJ. Insect silk: one name, many materials. Annu Rev Entomol. 2010; 55: 171–88. doi: 10.1146/annurev-ento-112408-085401 19728833

2. Xiang ZH. Quantitative trait genetics of silkworm. In: Xiang ZH, editor. Biology of Sericulture. Forestry Press of China, Beijing; 2005. pp. 242–53.

3. Andersson M, Johansson J, Rising A. Silk Spinning in Silkworms and Spiders. Int J Mol Sci. 2016; 17: 1290.

4. Sparkes J, Holland C. Analysis of the pressure requirements for silk spinning reveals a pultrusion dominated process. Nat Commun. 2017; 8: 594. doi: 10.1038/s41467-017-00409-7 28928362

5. Gong Y, Li L, Gong D, Yin H, Zhang J. Biomolecular Evidence of Silk from 8,500 Years Ago. PLoS One. 2016; 11(12):e0168042. doi: 10.1371/journal.pone.0168042 27941996

6. Yang SY, Han MJ, Kang LF, Li ZW, Shen YH, Zhang Z. Demographic history and gene flow during silkworm domestication. BMC Evol Biol. 2014; 14: 185. doi: 10.1186/s12862-014-0185-0 25123546

7. Lu C, Li B, Zhao AC, Xiang ZH. QTL mapping of economically important traits in Silkworm (Bombyx mori). Science in China Ser. C Life Sciences. 2004; 47: 477–484. doi: 10.1360/03yc0260 15623161

8. Mirhoseini SZ, Rabiei B, Potki P, Dalirsefat SB. Amplified fragment length polymorphism mapping of quantitative trait loci for economically important traits in the silkworm, Bombyx mori. J Insect Sci. 2010; 10: 153. doi: 10.1673/031.010.14113 21070171

9. Lie Z, Cheng L, Dai FY, Fang SM. Mapping of major quantitative trait loci for economic traits of silkworm cocoon. Genet Mol Res. 2010; 9: 78–88. doi: 10.4238/vol9-1gmr676 20092037

10. Zhan S, Huang J, Guo Q, Zhao Y, Li W, Miao X, et al. An integrated genetic linkage map for silkworms with three parental combinations and its application to the mapping of single genes and QTL. BMC Genomics. 2009; 10: 389. doi: 10.1186/1471-2164-10-389 19698097

11. Li C, Zuo WD, Tong XL, Hu H, Qiao L, Song JB, et al. A composite method for mapping quantitative trait loci without interference of female achiasmatic and gender effects in silkworm, Bombyx mori. Anim Genet. 2015; 46: 426–32. doi: 10.1111/age.12311 26059330

12. Xia Q, Guo Y, Zhang Z, Li D, Xuan Z, Li Z, et al. Complete resequencing of 40 genomes reveals domestication events and genes in silkworm (Bombyx mori). Science. 2009; 326: 433–6. doi: 10.1126/science.1176620 19713493

13. Xiang H. Liu X, Li M, Zhu Y, Wang L, Cui Y, et al. The evolutionary road from wild moth to domestic silkworm. Nat Ecol Evol. 2018; 2: 1268–1279. doi: 10.1038/s41559-018-0593-4 29967484

14. Li C, Tong XL, Zuo WD, Luan Y, Gao R, Han MJ, et al. QTL analysis of cocoon shell weight identifies BmRPL18 associated with silk protein synthesis in silkworm by pooling sequencing. Sci Rep. 2017; 7: 17985. doi: 10.1038/s41598-017-18277-y 29269837

15. Darvasi A, Soller M. Selective genotyping for determination of linkage between a marker locus and a quantitative trait locus. Theor Appl Genet. 1992; 85: 353–359. doi: 10.1007/BF00222881 24197326

16. Fang SM, Hu BL, Zhou QZ, Yu QY, Zhang Z. Comparative analysis of the silk gland transcriptomes between the domestic and wild silkworms. BMC Genomics. 2015; 16:60. doi: 10.1186/s12864-015-1287-9 25887670

17. Dong Y, Dai FY, Ren YD, Liu H, Chen L, Yang PC, et al. Comparative transcriptome analyses on silk glands of six silkmoths imply the genetic basis of silk structure and coloration. BMC Genomics. 2015; 16:203. doi: 10.1186/s12864-015-1420-9 25886738

18. Pinho SS, Reis CA. Glycosylation in cancer: mechanisms and clinical implications. Nat Rev Cancer. 2015; 15: 540–55. doi: 10.1038/nrc3982 26289314

19. Love DC, Hanover JA. The hexosamine signaling pathway: deciphering the "O-GlcNAc code". Sci STKE. 2005; 312: re13.

20. Varki A. Biological roles of oligosaccharides: all of the theories are correct. Glycobiology. 1993; 3: 97–130. doi: 10.1093/glycob/3.2.97 8490246

21. Lau KS, Partridge EA, Grigorian A, Silvescu CI, Reinhold VN, Demetriou M, et al. Complex N-glycan number and degree of branching cooperate to regulate cell proliferation and differentiation. Cell. 2007; 129: 123–34. doi: 10.1016/j.cell.2007.01.049 17418791

22. Huang B, Wu Q, Ge Y, Zhang J, Sun L, Zhang Y, et al. Expression of N-acetylglucosaminyltransferase V in gastric cancer correlates with metastasis and prognosis. Int J Oncol. 2014; 44: 849–57. doi: 10.3892/ijo.2014.2248 24399258

23. Takahashi N, Yamamoto E, Ino K, Miyoshi E, Nagasaka T, Kajiyama H, et al. High expression of N-acetylglucosaminyltransferase V in mucinous tumors of the ovary. Oncol Rep. 2009; 22: 1027–32. doi: 10.3892/or_00000531 19787216

24. Lau KS, Dennis JW. N-Glycans in cancer progression. Glycobiology. 2008; 18: 750–60. doi: 10.1093/glycob/cwn071 18701722

25. Zhou SB, Dong W, Zhao LJ. Discussion on Incorporating Sericin Content into Cocoon Quality Assessment System. Management and Depth. 2008; 11: 34–36.

26. Xiang ZH. Quantitative trait genetics of silkworm. In: Xiang ZH, editor. Biology of Sericulture. Forestry Press of China, Beijing; 2005. pp. 305–06.

27. Liu T, Zhang H, Liu F, Wu Q, Shen X, Yang Q. Structural determinants of an insect beta-N-Acetyl-D-hexosaminidase specialized as a chitinolytic enzyme. J Biol Chem. 2011; 286: 4049–58. doi: 10.1074/jbc.M110.184796 21106526

28. Hogenkamp DG, Arakane Y, Kramer KJ, Muthukrishnan S, Beeman RW. Characterization and expression of the beta-N-acetylhexosaminidase gene family of Tribolium castaneum. Insect Biochem Mol Biol. 2008; 38: 478–89. doi: 10.1016/j.ibmb.2007.08.002 18342252

29. Stern DL. The genetic causes of convergent evolution. Nat Rev Genet. 2013; 14: 751–64. doi: 10.1038/nrg3483 24105273

30. Dobler S, Dalla S, Wagschal V, Agrawal A. A. Community-wide convergent evolution in insect adaptation to toxic cardenolides by substitutions in the Na, K-ATPase. Proc Natl Acad Sci U S A. 2012; 109: 13040–5. doi: 10.1073/pnas.1202111109 22826239

31. Zhen Y, Aardema ML, Medina EM, Schumer M, Andolfatto P. Parallel molecular evolution in an herbivore community. Science. 2012; 337: 1634–7. doi: 10.1126/science.1226630 23019645

32. Colosimo PF, Hosemann KE, Balabhadra S, Villarreal GJ, Dickson M, Grimwood J, et al. Widespread parallel evolution in sticklebacks by repeated fixation of Ectodysplasin alleles. Science. 2005; 307: 1928–33. doi: 10.1126/science.1107239 15790847

33. Olofsson JK. Bianconi M, Besnard, Dunning LT, Lundgren MR, Holota H, et al. Genome biogeography reveals the intraspecific spread of adaptive mutations for a complex trait. Mol Ecol. 2016; 25: 6107–6123. doi: 10.1111/mec.13914 27862505

34. Song Y. Endepols S, Klemann N, Richter D, Matuschka FR, Shih CH, et al. Adaptive introgression of anticoagulant rodent poison resistance by hybridization between old world mice. Curr Biol. 2011; 21: 1296–301. doi: 10.1016/j.cub.2011.06.043 21782438

35. Nagaraja GM, Nagaraju J. Genome fingerprinting of the silkworm, Bombyx mori, using random arbitrary primers. Electrophoresis. 1995;16(9):1633–8. doi: 10.1002/elps.11501601270 8582347

36. Wang F, Xu H, Yuan L, Ma S, Wang Y, Duan X, et al. An optimized sericin-1 expression system for mass-producing recombinant proteins in the middle silk glands of transgenic silkworms. Transgenic Res. 2013; 22: 925–38. doi: 10.1007/s11248-013-9695-6 23435751

37. Zhao AC, Zhao TF, Zhang YS, Xia QY, Lu C, Zhou ZY, et al. New and highly efficient expression systems for expressing selectively foreign protein in the silk glands of transgenic silkworm. Transgenic Res, 2010; 19: 29–44. doi: 10.1007/s11248-009-9295-7 19533404

38. Dai ZJ, Sun W, Zhang Z. Comparative analysis of iTRAQ-based proteomes for cocoons between the domestic silkworm (Bombyx mori) and wild silkworm (Bombyx mandarina). J Proteomics. 2019; 192: 366–373. doi: 10.1016/j.jprot.2018.09.017 30287406

39. Axelsson E, Ratnakumar A, Arendt ML, Maqbool K, Webster MT, Perloski M, et al. The genomic signature of dog domestication reveals adaptation to a starch-rich diet. Nature. 2013; 495(7441): 360–4. doi: 10.1038/nature11837 23354050

40. Rozas J, Ferrer-Mata A, Sánchez-DelBarrio JC, Guirao-Rico S, Librado P, Ramos-Onsins SE, et al. DnaSP 6: DNA Sequence Polymorphism Analysis of Large Datasets. Mol. Biol. Evol. 2017; 34: 3299–3302. doi: 10.1093/molbev/msx248 29029172

41. Gao F, Chen C, Arab DA, Du Z, He Y, Ho SYW. EasyCodeML: A visual tool for analysis of selection using CodeML. Ecology and Evolution. 2019; 9(7):3891–3898. doi: 10.1002/ece3.5015 31015974

42. Kumar S, Stecher G, Tamura K. MEGA7: Molecular Evolutionary Genetics Analysis Version 7.0 for Bigger Datasets. Mol Biol Evol. 2016; 33(7):1870–4. doi: 10.1093/molbev/msw054 27004904


Článek vyšel v časopise

PLOS Genetics


2020 Číslo 7
Nejčtenější tento týden
Nejčtenější v tomto čísle
Kurzy Podcasty Doporučená témata Časopisy
Přihlášení
Zapomenuté heslo

Zadejte e-mailovou adresu, se kterou jste vytvářel(a) účet, budou Vám na ni zaslány informace k nastavení nového hesla.

Přihlášení

Nemáte účet?  Registrujte se

#ADS_BOTTOM_SCRIPTS#