Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Major ampullate silk gland transcriptomes and fibre proteomes of the golden orb-weavers, Nephila plumipes and Nephila pilipes (Araneae: Nephilidae)

  • Alessandra D. Whaite,

    Roles Conceptualization, Data curation, Investigation, Methodology, Validation, Writing – original draft, Writing – review & editing

    Affiliation GeneCology Research Centre and School of Science and Engineering, University of the Sunshine Coast, Sippy Downs, Queensland, Australia

  • Tianfang Wang,

    Roles Data curation, Methodology, Supervision, Writing – review & editing

    Affiliation GeneCology Research Centre and School of Science and Engineering, University of the Sunshine Coast, Sippy Downs, Queensland, Australia

  • Joanne Macdonald ,

    Roles Conceptualization, Funding acquisition, Methodology, Supervision, Writing – review & editing

    scummins@usc.edu.au (SFC); jmacdon1@usc.edu.au (JM)

    Affiliations GeneCology Research Centre and School of Science and Engineering, University of the Sunshine Coast, Sippy Downs, Queensland, Australia, Division of Experimental Therapeutics, Columbia University, New York City, New York, United States of America

  • Scott F. Cummins

    Roles Conceptualization, Data curation, Funding acquisition, Investigation, Methodology, Supervision, Writing – review & editing

    scummins@usc.edu.au (SFC); jmacdon1@usc.edu.au (JM)

    Affiliation GeneCology Research Centre and School of Science and Engineering, University of the Sunshine Coast, Sippy Downs, Queensland, Australia

Abstract

Natural spider silk is one of the world’s toughest proteinaceous materials, yet a truly biomimetic spider silk is elusive even after several decades of intense focus. In this study, Next-Generation Sequencing was utilised to produce transcriptomes of the major ampullate gland of two Australian golden orb-weavers, Nephila plumipes and Nephila pilipes, in order to identify highly expressed predicted proteins that may co-factor in the construction of the final polymer. Furthermore, proteomics was performed by liquid chromatography tandem-mass spectroscopy to analyse the natural solid silk fibre of each species to confirm highly expressed predicted proteins within the silk gland are present in the final silk product. We assembled the silk gland transcriptomes of N. plumipes and N. pilipes into 69,812 and 70,123 contigs, respectively. Gene expression analysis revealed that silk gene sequences were among the most highly expressed and we were able to procure silk sequences from both species in excess of 1,300 amino acids. However, some of the genes with the highest expression values were not able to be identified from our proteomic analysis. Proteome analysis of “reeled” silk fibres of N. plumipes and N. pilipes revealed 29 and 18 proteins, respectively, most of which were identified as silk fibre proteins. This study is the first silk gland specific transcriptome and proteome analysis for these species and will assist in the future development of a biomimetic spider silk.

Introduction

Spider silk is an outstanding proteinaceous fibre that outperforms other natural and synthetic fibres in tensile strength analyses. In addition to fibre strength, spider silks can also be tough, lightweight, highly extensible/flexible, biodegradable and stable across a broad temperature range [1]. Studies have also demonstrated the biocompatibility of spider silk: spider silk can be implanted into living tissue without eliciting an immune response [2, 3]. The potential of all these qualities present in one fibre type is unique and has led to intensive research interest in the molecular structure of the various silks spiders produce.

Orb-weaving spiders have up to seven different glands associated with producing silks and glues used for web architecture [major ampullate gland (MA), minor ampullate (Mi), flagelliform (Flag), aggregate (Ag), and pyriform (Py) glands], protection of eggs [tubuliform glands (Tu)], and prey wrapping [aciniform glands (Ac)] [4]. The molecular structure underpinning each of these silks and glues is the key to understanding the differing inherent mechanical properties of each type [5]. The major constituents in each silk gland are spider silk fibroins, called spidroins (Sp). While spidroins were named according to the gland from which they were primarily identified (for example the major ampullate spidroin (MaSp) from the MA gland), spidroins are not exclusively produced in their namesake gland, and proteomes have revealed silk to contain hundreds of proteins other than those derived from the spidroin gene family [611].

In the silk glands, silk proteins are secreted and then stored in the sac-like section of the gland in a water-soluble “molten fibril” or “aqua-melt” state, before processing occurs between the gland, duct and spinneret to produce the insoluble fibre we recognise within webs [12, 13]. Spidroin sequences are very large (approx. >10 kb) and code for a sequence that includes many repetitions of several repeat motifs, and is flanked by non-repetitive, highly conserved N- and C- termini [5,1416]. Typical repeat motifs in MA protein residues include stretches of poly-alanine (poly-A) and repeat glycine (G) motifs, such as GGX (where X could be A, glutamine, or tyrosine) [14]. An alternate MA spidroin contains stretches of poly-A and contains GPGGX motifs, where proline (P) is believed to contribute to the greater extensibility of this spidroin [10, 17].

Beyond spidroins, much remains to be understood about the tertiary and quaternary interactions of the various proteins and glycoproteins from storage in the soluble phase within the gland, to assembly via processing steps within the duct, and finally, the solid phase fibre as it leaves the spinnerets [1820]. Synthetic silks have not obtained the level of mechanical performance of their natural counterparts [2123]. The factors inhibiting successful biomimicry include the inability to mimic the natural production process in current expression systems, including the inability to express full-length recombinant spidroins, and, possibly, the lack of concurrent expression of a set of spidroins and other molecular components found within a silk thread. The non-repetitive N- and C- termini of ampullate silks have been found to play a role in fibre storage and assembly [24]. Further, the N- and C-termini are also thought to aid assembly of the secondary structures of the repeat regions within the processing duct. Indeed, thus far, the best example of a biomimetic spidroin has included these important highly conserved regions flanking a short repeat region [25].

The total primary MaSp sequence has been identified within 3 species; Latrodectus hesperus, also known as the Western Black Widow spider, the golden orb-weaving banana spider, Nephila clavipes, and Argiope bruennichi, the Wasp Spider [14, 26, 27]. Of more than 44,000 spider species, entire spider genomes have only been described for 6 species: L. hesperus (GenBank: JJRX00000000.1); the brown recluse, Loxosceles reclusa (JJRW000000000.1); the common house spider, Parasteatoda tepidariorum (AOMJ00000000.2); the African social velvet spider, Stegodyphus mimosarum (AZAQ00000000.1); the Brazilian white-knee tarantula, Acanthoscurria geniculata (GCA_000661875.1); and, most recently, N. clavipes (MWRG00000000.1)[2729]. In combination with their genome assembly for N. clavipes, Babb et al. characterised spidroin expression within the distinct silk glands [27]. Significant transcriptome coverage has also been achieved by Prosdocimi and colleagues in their infraorder comparative study of expressed RNAs of the spinning glands [30]; by Clark et al., based on multi-tissue de novo transcriptome assemblies of closely related cob-weavers [3133]; and Correa-Garhwal et al. examined silk expression in male spiders of the family Theridiidae [34].

In Queensland, Australia, Nephila plumipes (Latreille, 1804) and Nephila pilipes (Fabricius, 1793) are commonly encountered golden orb-weaving species [35]. Studies on N. plumipes have reported on the mechanical properties of their silk, the relationship between protein secondary structure and primary amino acid sequence, populations in the urban environment, and copulation behaviour, however, no transcriptome data is available on the MA gland of this species, and only a handful of silk sequences are available for N. pilipes in online databases [3642]. In this study, we report on a silk-gland specific transcriptome analysis for these golden orb-weaving species, N. plumipes and N. pilipes. Furthermore, proteomic analysis of silk fibres from these species was undertaken to compare predominant proteins within the silk fibre to predominant proteins expressed within the MA gland transcriptome. This study found that the silk gland transcriptome of N. plumipes and N. pilipes could be assembled into contiguous sequences, and proteome analysis of “reeled” silk fibres could confirm, and be used to mine for, spidroins within the transcriptomes. Novel proteins, which may be important constituents in the structure of spider silk, were also discovered in the silk proteome.

Methods

Animals and preparation of RNA

Golden orb spiders of the genus Nephila plumipes and Nephila pilipes were collected from the Sunshine Coast (26°41'43.1"S 153°05'56.7"E) and the Cooloola Coast (25°54'02.8"S 153°05'25.6"E) regions of Queensland, Australia, between the months of March and June 2013, and February and April 2016. Specimens were dissected immediately after sacrifice and each major ampullate gland were removed and either stored in RNAlater (Invitrogen) or immediately frozen in liquid nitrogen.

RNA was isolated from a pair of MA glands from an individual spider from each species using two different methods; the PicoPure RNA Isolation kit (Arcturus) and TRIzol Reagent (Invitrogen), according to manufacturer’s instructions, with the following changes:

For the PicoPure RNA extraction, an additional RNA purification step was performed using LiCl precipitation. Total RNA (30 μL) was divided into two aliquots for each species, and volume restored to 30 μL with 15 μL of RNase-free water so that samples were in duplicate. All RNA samples were incubated in 2.5 M LiCl (9.4 μL of 8 M LiCl) and 2.5 volumes (98.5 μL) of 100% EtOH for 1.5 h at -20°C. After incubation, the RNA was pelleted at 20,000 x g for 20 min at 4°C and the supernatant discarded. The pellet was washed with 70% EtOH and spun at 20,000 x g for 10 min at 4°C. The pellet was dried for 10 min at 37°C. The RNA concentration of a second aliquot from each species was estimated spectrophotometrically (NanoDrop 2000) after rehydration with 30 μL RNase-free water to ensure OD260/OD280 range was between 1.8 and 2.0. Dried samples of the first aliquot were stored at -80°C until shipment to BGI (China) for de novo RNA sequencing and bioinformatics.

For the TRIzol RNA extraction, all centrifugation steps were performed at 12,000 rpm at 4°C. RNA concentration and purity was estimated spectrophotometrically (NanoDrop 2000) to ensure OD260/OD280 range was between 1.8 and 2.0. Resuspended RNA was stored at -80°C until shipment to AGRF (Australia) for de novo RNA sequencing.

Next-generation sequencing, assembly and annotation

Total RNA from paired MA glands from individuals of both spider species were provided to the BGI for de novo RNA sequencing and bioinformatics using Illumina HiSeq 2000, and to the AGRF for library construction and paired-end sequencing using an Illumina HiSeq 2500 platform. Raw sequences were assembled into contigs using the Genomic CLC Workbench 9 software (default settings). Protein-coding regions were determined using the open reading frame (ORF) predictor [http://bioinformatics.ysu.edu/tools/OrfPredictor.html]. Blast2GO was utilised for functional annotation of protein-coding regions against the NCBI nr database [43]. Relative expression of genes in the transcriptome was determined based on reads per kilobase of transcript per million mapped reads (RPKM) values, utilizing the de novo RNA-seq CLC Genomic Workbench 9 software: transcripts per kilobase million mapped reads (TPM) are also reported.

Spidroin sequences of the genus Nephila were obtained from NCBI, compiled and used in a BLASTp search to identify homologous proteins derived from the N. plumipes and N. pilipes gland transcriptomes. Matches were manually assessed to determine conservation. Further, a “spidroin-like” database was created by examining the six translated nucleotide reading frames for the following spidroin-like amino acid motifs: AAAAA, GGYGG, GYGPG, GQQGP, and GAGAGG. Finally, CD-search [https://www.ncbi.nlm.nih.gov/Structure/bwrpsb/bwrpsb.cgi?] was utilised to identify specific hits to spidroin protein domains and superfamilies.

Spider silk preparation for proteomics

Spider silk threads were obtained by hand-reeling silk straight from live N. plumipes and N. pilipes spiders. Proteins were extracted from the silk by homogenisation in 100 μL protein extraction buffer (7 M urea, 2 M thiourea, 4% (w/v) CHAPS, 65 mM DTT), to which 100 μL of 30 mM Tris (pH 7.5) was added. Homogenate was vortexed and pulse-centrifuged several times to mix the dissolution buffer and Tris solution. The homogenate was vortexed for 15 min, pulse centrifuged, and then incubated in a sonicating water bath at ambient temperature for 20 min. Undissolved substances were pelleted by centrifugation at 12,000 x g for 8 min, and the supernatant containing the dissolved proteins were collected and stored at -80°C.

Sodium dodecyl sulphate—polyacrylamide gel electrophoresis (SDS-PAGE) and in-gel trypsin digestion

The spider silk proteins were separated by SDS-PAGE using a 4–20% polyacrylamide gradient gel (Amersham ECL Gel, GE Healthcare Life Sciences) according to manufacturer’s instructions. Samples were prepared 1:1 with 2x SDS sample buffer and incubated at 95°C for 5 min, then cooled to room temperature prior to pulse centrifugation. Samples were run for 1 h at 160 V (45 mA). Separated proteins were visualised by Coomassie Brilliant Blue (Bio-Rad) according to the staining process recommended by GE Healthcare. Upon completion of electrophoresis the proteins were precipitated with fixing solution [400 mL of EtOH, 100 mL of acetic acid, 500 mL of distilled water (DI)] for 30 min, followed by immersion in staining solution (1 tablet of PhastGel Blue-R 350, 400 mL of destaining solution; 250 mL of EtOH, 80 mL of acetic acid, 670 mL DI) for 10 min. The gel was subsequently preserved by immersion in preserving solution (25 mL of (87% v/v) glycerol with DI, 225 mL destaining solution) for 30 min. Protein sizes were estimated using a Pierce Blue Molecular Weight marker (Thermo Scientific).

Excised protein bands were washed in 50 mM NH4HCO3 at room temperature for 5 min, and then de-stained by incubating gel pieces for 30 min in 50 mM NH4HCO3 in 30% acetonitrile in a sonication water bath to remove Coomassie Brilliant Blue. Gel pieces were subsequently collected by in-gel trypsin digestion using the method described elsewhere [44]. The samples were reconstituted in 0.1% formic acid and stored at -20°C until mass spectroscopy.

NanoLC tandem TripleTof MS/MS analyses and protein identification

The spider silk extracts were analysed by LC-MS/MS on a Shimadzu Prominence Nano HPLC (Japan) coupled to a Triple ToF 5600 mass spectrometer (ABSCIEX, Canada) equipped with a nano electrospray ion source. Each extract (7 μL) was injected onto a 50 mm x 300 μm C18 trap column (Agilent Technologies, Australia) at 30 μL/min. The samples were de-salted on the trap column for 5 minutes using 0.1% formic acid (aq) at 30 μL/min. The trap column was then placed in-line with the analytical nano HPLC column, a 150 mm x 75 μm 300SBC18, 3.5 μm (Agilent Technologies, Australia) for mass spectrometry analysis. Linear gradients of 1–40% solvent B over 35 min at 300 nL/minute flow rate, followed by a steeper gradient from 40% to 80% solvent B in 5 min were used for peptide elution. Solvent B was held at 80% for 5 min for washing the column and returned to 1% solvent B for equilibration prior to the next sample injection. Solvent A consisted of 0.1% formic acid (aq) and solvent B contained 90/10 acetonitrile/ 0.1% formic acid (aq). The ionspray voltage was set to 2400V, declustering potential 100V, curtain gas flow 25, nebuliser gas 1 (GS1) 12 and interface heater at 150°C. The mass spectrometer acquired 500 ms full scan TOF-MS data followed by 20 by 50 ms full scan product ion data in an Information Dependant Acquisition mode. Full scan TOF-MS data was acquired over the mass range 350–1800 and for product ion ms/ms 100–1800. Ions observed in the TOF-MS scan exceeding a threshold of 100 counts and a charge state of +2 to +5 were set to trigger the acquisition of product ion, MS/MS spectra of the resultant 20 most intense ions. The data was acquired and processed using Analyst TF 1.5.1 software (AB SCIEX, Concord, Canada).

The LC-MS/MS data were imported to the PEAKS studio (Bioinformatics Solutions Inc., Waterloo, ON, Canada, version 7.0) with the assistance of MS Data Converter (Beta 1.3, http://sciex.com/software-downloads-x2110). The database search included our own Nephila sp. transcriptome-derived protein databases, our “spidroin-like” database made by motif-searching the six translated nucleotide reading frames, and non-redundant protein databases (GenBank and UniProt). De novo sequencing of peptides, database search and characterising specific PTMs were used to analyse the raw data; false discovery rate (FDR) was set to ≤ 1%, and [-10*log(p)] was calculated accordingly where p is the probability that an observed match is a random event. The PEAKS used the following parameters: (i) precursor ion mass tolerance, 0.1 Da; (ii) fragment ion mass tolerance, 0.1 Da (the error tolerance); (iii) tryptic enzyme specificity with two missed cleavages allowed; (iv) monoisotopic precursor mass and fragment ion mass; (v) a fixed modification of cysteine carbamidomethylation; and (vi) variable modifications including lysine acetylation, deamidation on asparagine and glutamine, oxidation of methionine and conversion of glutamic acid and glutamine to pyroglutamate.

Results and discussion

Nephila plumipes and N. pilipes (Fig 1A) were collected and the MA glands were removed (Fig 1B) for RNA isolation and sequencing. MA gland reference transcriptomes were constructed for each species by combining next-generation sequence (NGS) data from sequencing runs produced in 2013, with data produced in 2016 (GenBank Accession: SRR6747912, SRR6747911). The combined MA gland transcriptomes for N. plumipes and N. pilipes produced 42,351,802 and 46,060,170 total paired reads, respectively. Paired reads were assembled into 69,812 contiguous sequences (contigs) for N. plumipes with an average length of 685 nucleotides, and into 70,123 contigs for N. pilipes with an average nucleotide length of 672. The data returned from ORF prediction were 67,862 and 67,942 sequences, and Blast2Go annotation allowed for the annotation of approximately 25% and 29% of all transcripts plus identification of 48 and 35 spidroin contigs, for N. plumipes and N. pilipes, respectively. The total spidroin count for each species was increased to 73 and 60 spidroins for N. plumipes and N. pilipes (Table 1 and 2), upon mining for spidroin sequences identified within each corresponding transcriptome-derived silk proteome (S1 and S2 Tables), by analysing the six translated nucleotide reading frames for spidroin-like repeat motifs, and by BLAST searching unique silk sequences found in the related N. clavipes genome reported by Babb et al. [27].

thumbnail
Fig 1. The golden orb-weaving spiders and a representative MA gland.

(A) Nephila plumipes (photo by Alessandra Whaite) and Nephila pilipes (photo by Amos T Fairchild). (B) The major ampullate gland of Nephila pilipes, indicating the secretory section and the sac that stores silk proteins.

https://doi.org/10.1371/journal.pone.0204243.g001

thumbnail
Table 1. Total spidroins in each Nephila plumipes data set (2013 and 2016).

https://doi.org/10.1371/journal.pone.0204243.t001

thumbnail
Table 2. Total spidroins in each Nephila pilipes data set (2013 and 2016).

https://doi.org/10.1371/journal.pone.0204243.t002

A recent paper reporting on the genome and tissue transcriptomes of the golden-orb spider, N. clavipes, identified 28 spidroins [27]. Our transcriptomes included partial sequences with homology to several of the unique spidroins found in this closely related species, including AgSp-a, AgSp-c, Sp-5803, Sp8175, Sp74867 and, MaSp-c, -d -g and -h (Table 1 and 2). In N. plumipes, our study uncovered numerous matches to the N. clavipes spidroin Sp-907, potentially non-overlapping contigs aligning to different regions on the same gene. In N. pilipes, our study found sequences with homology to MiSp-a and -d, and Sp-1339. Our longest assembled spidroin contigs in both species matched to the N. clavipes the glue-like aggregate spidroin, AgSp-c. Aggregate spidroins, which form evenly-spaced droplets along flagelliform prey capture threads, have also been characterised in three other species from the family Araneidae, and three species from the family Theridiidae [45, 46], These spidroins vary greatly in length among species and it appears we have recovered a full-length aggregate spidroin (1609 aa) from N. pilipes. However, aggregate spidroins were not evident in the proteome of either species.

The N. pilipes and N. plumipes silk proteins were obtained after several predominant bands were excised from the Coomassie stained SDS-PAGE gel (Fig 2), followed by trypsin digestion and LC-MS/MS analysis. This analysis identified 29 and 18 proteins for N. plumipes and N. pilipes (Table 3 and 4), respectively (≥20.00 -10lgP, ≥ 2 peptide matches). For the silk proteins of N. plumipes mapped to the transcriptome, 24 were spidroins, mostly MaSp1-like spidroins, MiSp-like and novel N. clavipes spidroin, Sp-907, confirming the abundance of this spidroin in the transcriptome. Besides spidroins, a cuticle protein and a coiled-coil domain-containing protein were identified in the proteome along with two proteins which could not be identified. Cuticle proteins have been described previously in the MA gland [7, 47]. The coiled-coil domain protein found within the transcriptome and proteome of N. plumipes is potentially interesting as coiled-coil structures are also found to form in the silk of the Japanese yellow hornet, Vespa simillima [48]. N. pilipes silk proteins accounted for 13 of 18 proteins mapped back to the corresponding transcriptome. Again, most of these spidroins were MaSp-like, one MiSp and the remaining 5 proteins were not able to be annotated at this stage.

thumbnail
Fig 2. SDS-PAGE and Coomassie stain indicating prominent protein bands that were excised for LC-MS/MS analysis.

(A-F) Nephila plumipes, (G-K) Nephila pilipes. MWM, molecular weight marker.

https://doi.org/10.1371/journal.pone.0204243.g002

thumbnail
Table 3. Nephila plumipes proteins within the silk matching to the transcriptome.

https://doi.org/10.1371/journal.pone.0204243.t003

thumbnail
Table 4. Nephila pilipes proteins within the silk matching to the transcriptome.

https://doi.org/10.1371/journal.pone.0204243.t004

Beyond the matches made to the transcriptome, from the silk of N. plumipes and N. pilipes a further 2,420 and 2,658 de novo only peptides, respectively, were identified with high confidence (average local confidence above 70) that did not match any sequences within the transcriptome. This de novo dataset of unmatched potential proteins was BLASTed against NCBI protein databases, however the focus in this study is on those proteins relevant to their corresponding transcriptomes. Unlike a genome, a transcriptome can only provide us with genes transcribed at the time of RNA isolation and this may partly explain the discrepancy between matched and de novo peptides. A further explanation is that the hand-reeled silk also contains silk from different silk glands. Each silk gland ends at its own spigot on the surface of a spinneret. It is possible that as the silk thread passes past other spigots during collection, it also collects fibres from other silk glands. However, the spigots closest to the major ampullate spigot on the anterior lateral spinnerets produce pyriform spidroins and there was no evidence of these spidroins in the proteome [4]. Further, our gel-based extraction method might have missed proteins with relatively low molecular weight or low abundance, such as the cysteine-rich proteins identified by Pham et al. which were found to co-localise with spidroins in the MA silk of Latrodectus hesperus [8].

Quantitative analyses were undertaken and based on reads per kilobase of transcript per million mapped reads (RPKM) values by mapping the 2013 and 2016 data back to a combined de novo reference transcriptome. The 50 most highly expressed sequences of N. pilipes and N. plumipes were manually selected for further annotation (Tables 5 and 6). These abundant sequences were matched to sequences found in the NCBI or Uniprot public protein databases (accessed Oct-Dec 2017). Spidroins were, as expected, among the most highly expressed sequences of both datasets, numbering 23 and 26 spidroins for N. plumipes and N. pilipes, respectively. In both species, major and minor ampullate, and tubuliform spidroins were highly expressed in the MA gland. Interestingly, the N. plumipes sequence with the highest RPKM value could not be characterised based on BLAST protein prediction. Uncharacterised highly expressed sequences will be selected for functional annotation in future works.

thumbnail
Table 5. Nephila plumipes silk gland most abundantly expressed genes.

https://doi.org/10.1371/journal.pone.0204243.t005

thumbnail
Table 6. Nephila pilipes silk gland most abundantly expressed genes.

https://doi.org/10.1371/journal.pone.0204243.t006

This study found the MA gland alone produces six of the seven classes of silk products: MA, minor ampullate, flagelliform, tubuliform (also at times referred to as cylindriform silk), aciniform and aggregate silk products. Several other studies have also found multiple spidroin types expressed in a single gland [6, 7, 49, 50]. The only silk product not found to be produced by the MA gland of both N. pilipes and N. plumipes was pyriform adhesive silk, which is used to attach threads to objects and to each other [51]. The processing duct of the pyriform gland is shorter than most other ducts suggesting other silks require more extensive processing, which may explain why this silk is absent from the MA gland transcriptome. However, pyriform products are the least intensively studied of the spider silk repertoire, and the lack of pyriform annotation in our MA databases may be a reflection of poor representation in the public databases at the present time [51, 52].

Interestingly, in both N. plumipes and N. pilipes, tubuliform spidroins were found to be more highly expressed in the 2016 MA gland transcriptomes yet not expressed in the 2013 transcriptomes (see Table 1 and 2). Tubuliform silk is produced during reproduction for the formation of egg sacs [4, 53]. While no spiders were gravid at the time of dissection, it is possible they were collected and dissected just after the production of an egg sac, or just prior to vitellogenesis, and the 2016 transcriptomes reflect this in their relatively high expression of tubuliform silk transcripts. Expression of tubuliform spidroins in the MA gland has been previously noted in transcriptomic studies [50]. Vasanthavada et al. suggest that spiders can downregulate the production of various silks to maintain MA spidroin synthesis as an energetic trade-off, and Larracas et al. suggest that female spiders may shift synthesis of MA gland spidroins to tubuliform spidroins during the reproductive stage [6, 54, 55]. Our study did not find tubuliform spidroins in the silk proteome, however, the silk was collected and digested at the same time as the 2013 transcriptomes. It would be interesting to see if tubuliform spidroins could be found within the dragline silk of spiders prior to, during, or just post egg sac production.

This study is the first silk gland-specific transcriptome and proteome analysis in these Australian golden orb-weaving species. Major ampullate transcriptome analysis procured sequences for all silk types thus far known for golden orb spiders with the exception of pyriform adhesive silk. We found differential expression of tubuliform silk in the MA gland, suggesting a greater role for this gland producing tubuliform silks during spider reproduction. The silk proteome analysis resulted in 29 and 18 proteins for N. plumipes and N. pilipes that match to their corresponding MA gland transcriptomes.

Supporting information

S1 Table. Nephila plumipes proteins with homology to Nephila clavipes spidroins.

https://doi.org/10.1371/journal.pone.0204243.s001

(XLSX)

S2 Table. Nephila pilipes proteins with homology to Nephila clavipes spidroins.

https://doi.org/10.1371/journal.pone.0204243.s002

(XLSX)

S3 Table. List of peptides that were mapped back to the Nephila plumipes transcriptome.

https://doi.org/10.1371/journal.pone.0204243.s003

(XLSX)

S4 Table. List of peptides that were mapped back to the Nephila pilipes transcriptome.

https://doi.org/10.1371/journal.pone.0204243.s004

(XLSX)

Acknowledgments

We gratefully thank Dr Alun Jones (Institute for Molecular Bioscience, the University of Queensland) for advice and assistance with the LC-MS/MS. We also thank Dr Owen Seeman and Dr Robert Raven (Queensland Museum, Brisbane) for training in species identification and gland dissection. Thanks to Dr Min Zhao for assistance with bioinformatics.

References

  1. 1. Brown CP, Whaite AD, MacLeod JM, Macdonald J, Rosei F. With great structure comes great functionality: Understanding and emulating spider silk. J Mater Res. 2014;30(1):108–20.
  2. 2. Gellynck K, Verdonk P, Forsyth R, Almqvist KF, Van Nimmen E, Gheysens T, et al. Biocompatibility and biodegradability of spider egg sac silk. J Mater Sci Mater Med. 2008;19(8):2963–70. pmid:18360800
  3. 3. Kundu B, Kurland NE, Bano S, Patra C, Engel FB, Yadavalli VK, et al. Silk proteins for biomedical applications: Bioengineering perspectives. Prog Polym Sci. 2014;39(2):251–67.
  4. 4. Foelix RF. The biology of spiders. New York: Oxford University Press; 1996.
  5. 5. Gosline JM, Guerette PA, Ortlepp CS, Savage KN. The mechanical design of spider silks: From fibroin sequence to mechanical function. J Exp Biol. 1999;202(23):3295–303.
  6. 6. Larracas C, Hekman R, Dyrness S, Arata A, Williams C, Crawford T, et al. Comprehensive proteomic analysis of spider dragline silk from blackwidows: A recipe to build synthetic silk fibers. Int J Mol Sci. 2016;17(9).
  7. 7. Chaw RC, Correa-Garhwal SM, Clarke TH, Ayoub NA, Hayashi CY. Proteomic Evidence for Components of Spider Silk Synthesis from Black Widow Silk Glands and Fibers. J Proteome Res. 2015;14(10):4223–31. pmid:26302244
  8. 8. Pham T, Chuang T, Lin A, Joo H, Tsai J, Crawford T, et al. Dragline silk: A fiber assembled with low-molecular-weight cysteine-rich proteins. Biomacromolecules. 2014;15(11):4073–81. pmid:25259849
  9. 9. Dos Santos-Pinto JR, Garcia AM, Arcuri HA, Esteves FG, Salles HC, Lubec G, et al. Silkomics: Insight into the Silk Spinning Process of Spiders. J Proteome Res. 2016;15(4):1179–93. pmid:26923066
  10. 10. Hinman MB, Lewis RV. Isolation of a clone encoding a second dragline silk fibroin. Nephila clavipes dragline silk is a two-protein fiber. J Biol Chem. 1992;267(27):19320–4. pmid:1527052
  11. 11. Dos Santos-Pinto JRA, Arcuri HA, Priewalder H, Salles HC, Palma MS, Lubec G. Structural model for the spider silk protein spidroin-1. J Proteome Res. 2015;14(9):3859–70. pmid:26211688
  12. 12. Hijirida DH, Do KG, Michal C, Wong S, Zax D, Jelinski LW. 13C NMR of Nephila clavipes major ampullate silk gland. Biophys J. 1996;71(6):3442–7. pmid:8968613
  13. 13. Holland C, Vollrath F, Ryan AJ, Mykhaylyk OO. Silk and synthetic polymers: Reconciling 100 degrees of separation. Adv Mater. 2012;24(1):105–9. pmid:22109705
  14. 14. Ayoub NA, Garb JE, Tinghitella RM, Collin MA, Hayashi CY. Blueprint for a High-Performance Biomaterial: Full-Length Spider Dragline Silk Genes. PLoS ONE. 2007;2(6).
  15. 15. Rising A, Nimmervoll H, Grip S, Fernandez-Arias A, Storckenfeldt E, Knight DP, et al. Spider Silk Proteins–Mechanical Property and Gene Sequence. Zool Sci. 2005;22(3):273–81. pmid:15795489
  16. 16. Beckwitt R, Arcidiacono S. Sequence conservation in the C-terminal region of spider silk proteins (Spidroin) from Nephila clavipes (Tetragnathidae) and Araneus bicentenarius (Araneidae). J Biol Chem. 1994;269(9):6661–3. pmid:8120021
  17. 17. Tokareva O, Jacobsen M, Buehler M, Wong J, Kaplan DL. Structure-function-property-design interplay in biopolymers: Spider silk. Acta Biomater. 2013.
  18. 18. Augsten K, Muhlig P, Herrmann C. Glycoproteins and skin-core structure in Nephila clavipes spider silk observed by light and electron microscopy. Scanning. 2000;22(1):12–5. pmid:10768384
  19. 19. Vollrath F, Knight DP. Liquid crystalline spinning of spider silk. Nature. 2001;410(6828):541–8. pmid:11279484
  20. 20. Mello CM, Senecal K, Yeung B, Vouros P, Kaplan D. Initial Characterization of Nephila clavipes Dragline Protein. Silk Polymers. ACS Symposium Series. 544: J Am Chem Soc. 1993; p. 67–79.
  21. 21. Seidel A, Liivak O, Jelinski LW. Artificial spinning of spider silk. Macromolecules. 1998;31(19):6733–6.
  22. 22. Xia XX, Qian ZG, Ki CS, Park YH, Kaplan DL, Lee SY. Native-sized recombinant spider silk protein produced in metabolically engineered Escherichia coli results in a strong fiber. Proc Natl Acad Sci USA. 2010;107(32):14059–63. pmid:20660779
  23. 23. Teulé F, Miao YG, Sohn BH, Kim YS, Hull JJ, Fraser MJ Jr, et al. Silkworms transformed with chimeric silkworm/spider silk genes spin composite silk fibers with improved mechanical properties. Proc Natl Acad Sci USA. 2012;109(3):923–8. pmid:22215590
  24. 24. Hagn F, Eisoldt L, Hardy JG, Vendrely C, Coles M, Scheibel T, et al. A conserved spider silk domain acts as a molecular switch that controls fibre assembly. Nature. 2010;465(7295):239–42. pmid:20463741
  25. 25. Andersson M, Jia Q, Abella A, Lee XY, Landreh M, Purhonen P, et al. Biomimetic spinning of artificial spider silk from a chimeric minispidroin. Nat Chem Biol. 2017;13(3):262–4. pmid:28068309
  26. 26. Zhang Y, Zhao AC, Sima YH, Lu C, Xiang ZH, Nakagaki M. The molecular structures of major ampullate silk proteins of the wasp spider, Argiope bruennichi: A second blueprint for synthesizing de novo silk. Comp Biochem Physiol B Biochem Mol Biol. 2013;164(3):151–8. pmid:23262065
  27. 27. Babb PL, Lahens NF, Correa-Garhwal SM, Nicholson DN, Kim EJ, Hogenesch JB, et al. The Nephila clavipes genome highlights the diversity of spider silk genes and their complex expression. Nat Genet. 2017;49(6):895–903. pmid:28459453
  28. 28. Schwager EE, Sharma PP, Clarke T, Leite DJ, Wierschin T, Pechmann M, et al. The house spider genome reveals an ancient whole-genome duplication during arachnid evolution. BMC Biology. 2017;15(1).
  29. 29. Sanggaard KW, Bechsgaard JS, Fang X, Duan J, Dyrlund TF, Gupta V, et al. Spider genomes provide insight into composition and evolution of venom and silk. Nat Commun. 2014;5:3765. pmid:24801114
  30. 30. Prosdocimi F, Bittencourt D, da Silva FR, Kirst M, Motta PC, Rech EL. Spinning gland transcriptomics from two main clades of spiders (order: Araneae)—insights on their molecular, anatomical and behavioral evolution. PLoS ONE. 2011;6(6).
  31. 31. Clarke TH, Garb JE, Haney RA, Chaw RC, Hayashi CY, Ayoub NA. Evolutionary shifts in gene expression decoupled from gene duplication across functionally distinct spider silk glands. Scientific reports. 2017;7(1).
  32. 32. Clarke TH, Garb JE, Hayashi CY, Haney RA, Lancaster AK, Corbett S, et al. Multi-tissue transcriptomics of the black widow spider reveals expansions, co-options, and functional processes of the silk gland gene toolkit. BMC Genomics. 2014;15(1).
  33. 33. Clarke TH, Garb JE, Hayashi CY, Arensburger P, Ayoub NA. Spider Transcriptomes Identify Ancient Large-Scale Gene Duplication Event Potentially Important in Silk Gland Evolution. Genome Biology and Evolution. 2015;7(7):1856–70. pmid:26058392
  34. 34. Correa-Garhwal SM, Chaw RC, Clarke TH, Ayoub NA, Hayashi CY. Silk gene expression of theridiid spiders: Implications for male-specific silk use. Zoology. 2016.
  35. 35. Harvey MS, Austin AD, Adams M. The systematics and biology of the spider genus Nephila (Araneae:Nephilidae) in the Australasian region. Invertebr Syst. 2007;21(5):407–51.
  36. 36. Blamires SJ, Kasumovic MM, Tso IM, Martens PJ, Hook JM, Rawal A. Evidence of decoupling protein structure from spidroin expression in spider dragline silks. Int J Mol Sci. 2016;17(8).
  37. 37. Schneider JM, Herberstein ME, Bruce MJ, Kasumovic MM, Thomas ML, Elgar MA. Male copulation frequency, sperm competition and genital damage in the golden orb-web spider (Nephila plumipes). Aust. J. Zool. 2008;56(4):233–8.
  38. 38. Prenter J, Pérez-Staples D, Taylor PW. The effects of morphology and substrate diameter on climbing and locomotor performance in male spiders. Func Ecol. 2010;24(2):400–8.
  39. 39. Lowe EC, Wilder SM, Hochuli DF. Urbanisation at multiple scales is associated with larger size and higher fecundity of an orb-weaving spider. PLoS One. 2014;9(8):e105480. pmid:25140809
  40. 40. Griffiths JR, Salanitri VR. The strength of spider silk. J Mater Sci. 1980;15(2):491–6.
  41. 41. Tai PL, Hwang GY, Tso IM. Inter-specific sequence conservation and intra-individual sequence variation in a spider silk gene. Int J Biol Macromol. 2004;34(5):295–301. pmid:15556231
  42. 42. Blamires SJ, Liao CP, Chang CK, Chuang YC, Wu CL, Blackledge TA, et al. Mechanical Performance of Spider Silk Is Robust to Nutrient-Mediated Changes in Protein Composition. Biomacromolecules. 2015;16(4):1218–25. pmid:25764227
  43. 43. Conesa A, Gotz S, Garcia-Gomez JM, Terol J, Talon M, Robles M. Blast2GO: a universal tool for annotation, visualization and analysis in functional genomics research. Bioinformatics (Oxford, England). 2005;21(18):3674–6.
  44. 44. Wang T, Zhao M, Rotgans BA, Ni G, Dean JF, Nahrung HF, et al. Proteomic analysis of the venom and venom sac of the woodwasp, Sirex noctilio—Towards understanding its biological impact. J Proteom. 2016;146:195–206.
  45. 45. Collin MA, Clarke TH, Ayoub NA, Hayashi CY. Evidence from Multiple Species that Spider Silk Glue Component ASG2 is a Spidroin. Scientific reports. 2016;6.
  46. 46. Li X, Shi CH, Tang CL, Cai YM, Meng Q. The correlation between the length of repetitive domain and mechanical properties of the recombinant flagelliform spidroin. Biology Open. 2017;6(3):333–9. pmid:28126711
  47. 47. Davies GJG, Knight DP, Vollrath F. Chitin in the Silk Gland Ducts of the Spider Nephila edulis and the Silkworm Bombyx mori. PLoS ONE. 2013;8(8).
  48. 48. Kameda T, Kojima K, Togawa E, Sezutsu H, Zhang Q, Teramoto H, et al. Drawing-Induced Changes in Morphology and Mechanical Properties of Hornet Silk Gel Films. Biomacromolecules. 2010;11(4):1009–18. pmid:20178379
  49. 49. Lane AK, Hayashi CY, Whitworth GB, Ayoub NA. Complex gene expression in the dragline silk producing glands of the Western black widow (Latrodectus hesperus). BMC Genomics. 2013:846. pmid:24295234
  50. 50. Guerette PA, Ginzinger DG, Weber BHF, Gosline JM. Silk properties determined by gland-specific expression of a spider fibroin gene family. Science. 1996;272(5258):112–5. pmid:8600519
  51. 51. Blasingame E, Tuton-Blasingame T, Larkin L, Falick AM, Zhao L, Fong J, et al. Pyriform spidroin 1, a novel member of the silk gene family that anchors dragline silk fibers in attachment discs of the black widow spider, Latrodectus hesperus. J Biol Chem. 2009;284(42):29097–108. pmid:19666476
  52. 52. Chaw RC, Saski CA, Hayashi CY. Complete gene sequence of spider attachment silk protein (PySp1) reveals novel linker regions and extreme repeat homogenization. Insect Biochem Mol Biol. 2017.
  53. 53. Casem ML, Collin MA, Ayoub NA, Hayashi CY. Silk gene transcripts in the developing tubuliform glands of the Western black widow, Latrodectus hesperus. J Arachnol. 2010;38(1):99–103.
  54. 54. Vasanthavada K, Hu X, Tuton-Blasingame T, Hsia Y, Sampath S, Pacheco R, et al. Spider glue proteins have distinct architectures compared with traditional spidroin family members. The Journal of biological chemistry. 2012;287(43):35986–99. pmid:22927444
  55. 55. Herrera E, Nguyen LT, Escobar E, Ouriel W, Casem ML. Alteration of tubuliform silk gland cytoarchitecture with the reproductive cycle of the Western black widow spider, Latrodectus hesperus. Invertebr Biol. 2015;134(4):332–40.