Spider silk

A garden spider spinning its web
A female specimen of Argiope bruennichi wraps her prey in silk.
Indian summer by Józef Chełmoński (1875, National Museum in Warsaw) depicts a peasant woman with a thread of gossamer in her hand.
Spider cocoon

Spider silk is a protein fibre spun by spiders. Spiders use their silk to make webs or other structures, which function as sticky nets to catch other animals, or as nests or cocoons to protect their offspring, or to wrap up prey. They can also use their silk to suspend themselves, to float through the air, or to glide away from predators. Most spiders vary the thickness and stickiness of their silk for different uses.

In some cases, spiders may even use silk as a source of food.[1] While methods have been developed to collect silk from a spider by force,[2] it is difficult to gather silk from many spiders in a small space, in contrast to silkworm 'farms'.

Biodiversity

Uses

All spiders produce silks, and a single spider can produce up to seven different types of silk for different uses.[3] This is in contrast to insect silks, where an individual usually only produces one type of silk.[4] Spider silks may be used in many different ecological ways, each with properties to match the silk's function (see Properties section). As spiders have evolved, so has their silks' complexity and diverse uses, for example from primitive tube webs 300–400 million years ago to complex orb webs 110 million years ago.[5]

Use Example Reference
Prey capture The orb webs produced by the Araneidae (typical orb-weavers); tube webs; tangle webs; sheet webs; lace webs, dome webs; single thread used by the Bolas spiders for "fishing". [3][5]
Prey immobilization Silk used as "swathing bands" to wrap up prey. Often combined with immobilizing prey using a venom. In species of Scytodes the silk is combined with venom and squirted from the chelicerae. [3]
Reproduction Male spiders may produce sperm webs; spider eggs are covered in silk cocoons. [3][6]
Dispersal "Ballooning" or "kiting" used by smaller spiders to float through the air, for instance for dispersal. [7]
Source of food The kleptoparasitic Argyrodes eating the silk of host spider webs. Some daily weavers of temporary webs also eat their own unused silk daily, thus mitigating a heavy metabolic expense. [1][8]
Nest lining and nest construction Tube webs used by "primitive" spiders such as the European tube web spider (Segestria florentina). Threads radiate out of nest to provide a sensory link to the outside. Silk is a component of the lids of spiders that use "trapdoors", such as members of the family Ctenizidae, and the "water" or "diving bell" spider Argyroneta aquatica builds its diving bell of silk. [5]
Guide lines Some spiders that venture from shelter will leave a trail of silk by which to find their way home again. [8]
Drop lines and anchor lines Many spiders, such as the Salticidae, that venture from shelter and leave a trail of silk, use that as an emergency line in case of falling from inverted or vertical surfaces. Many others, even web dwellers, will deliberately drop from a web when alarmed, using a silken thread as a drop line by which they can return in due course. Some, such as species of Paramystaria, also will hang from a drop line when feeding. [8]
Alarm lines Some spiders that do not spin actual trap webs do lay out alarm webs that the feet of their prey (such as ants) can disturb, cuing the spider to rush out and secure the meal if it is small enough, or to avoid contact if the intruder seems too formidable. [8]
Pheromonal trails Some wandering spiders will leave a largely continuous trail of silk impregnated with pheromones that the opposite sex can follow to find a mate. [8]

Types

Meeting the specification for all these ecological uses requires different types of silk suited to different broad properties, as either a fiber, a structure of fibers, or a silk-globule. These types include glues and fibers. Some types of fibers are used for structural support, others for constructing protective structures. Some can absorb energy effectively, whereas others transmit vibration efficiently. In a spider, these silk types are produced in different glands; so the silk from a particular gland can be linked to its use by the spider. See the later section for details on the mechanical properties of silk and how the structure of silk can achieve these different properties.

Gland Silk Use
Ampullate (Major) Dragline silk—used for the web’s outer rim and spokes and the lifeline.
Ampullate (Minor) Used for temporary scaffolding during web construction.
Flagelliform Capture-spiral silk—used for the capturing lines of the web.
Tubuliform Egg cocoon silk—used for protective egg sacs.
Aciniform Used to wrap and secure freshly captured prey; used in the male sperm webs; used in stabilimenta.
Aggregate A silk glue of sticky globules.
Piriform Used to form bonds between separate threads for attachment points.

Properties

Mechanical properties

Each spider and each type of silk has a set of mechanical properties optimised for their biological function.

Most silks, in particular dragline silk, have exceptional mechanical properties. They exhibit a unique combination of high tensile strength and extensibility (ductility). This enables a silk fibre to absorb a lot of energy before breaking (toughness, the area under a stress-strain curve).

An illustration of the differences between toughness, stiffness and strength

A frequent mistake made in the mainstream media is to confuse strength and toughness, when comparing silk to other materials. As shown below in detail, weight for weight, silk is stronger than steel, but not as strong as Kevlar. Silk is, however, tougher than both.

It is important to note that the variability of mechanical properties of spider silk fibers may be important and it is related to their degree of molecular alignment.[9] Besides, mechanical properties depend strongly on the ambient conditions, i.e. humidity and temperature.[10]

Strength

In detail a dragline silk's tensile strength is comparable to that of high-grade alloy steel (450 - 2000 MPa),[11][12] and about half as strong as aramid filaments, such as Twaron or Kevlar (3000 MPa).[13]

Density

Consisting of mainly protein, silks are about a sixth of the density of steel (1.3 g/cm3). As a result, a strand long enough to circle the Earth would weigh less than 500 grams (18 oz). (Spider dragline silk has a tensile strength of roughly 1.3 GPa. The tensile strength listed for steel might be slightly higher—e.g. 1.65 GPa,[14][15] but spider silk is a much less dense material, so that a given weight of spider silk is five times as strong as the same weight of steel.)

Energy density

The energy density of dragline spider silk is roughly 1.2x108J/m3.[16]

Extensibility

Silks are also extremely ductile, with some able to stretch up to five times their relaxed length without breaking.

Toughness

The combination of strength and ductility gives dragline silks a very high toughness (or work to fracture), which "equals that of commercial polyaramid (aromatic nylon) filaments, which themselves are benchmarks of modern polymer fibre technology".[17][18]

Temperature

While unlikely to be relevant in nature, dragline silks can hold their strength below −40 °C (-40 °F) and up to 220 °C (428 °F).[19] As occurs in many materials, spider silk fibers undergo a glass transition. The glass-transition temperature depends on the humidity, as water is a plastiziser for the silk.[10]

Supercontraction

When exposed to water, dragline silks undergo supercontraction, shrinking up to 50% in length and behaving like a weak rubber under tension.[10] Many hypotheses have been suggested as to its use in nature, with the most popular being to automatically tension webs built in the night using the morning dew.

Highest-performance

The toughest known spider silk is produced by the species Darwin's bark spider (Caerostris darwini): "The toughness of forcibly silked fibers averages 350 MJ/m3, with some samples reaching 520 MJ/m3. Thus, C. darwini silk is more than twice as tough as any previously described silk, and over 10 times tougher than Kevlar".[20]

Types of silk

Many species of spider have different glands to produce silk with different properties for different purposes, including housing, web construction, defense, capturing and detaining prey, egg protection, and mobility (gossamer for ballooning, or for a strand allowing the spider to drop down as silk is extruded). Different specialized silks have evolved with properties suitable for different uses. For example, Argiope argentata has five different types of silk, each used for a different purpose:[21][22]

Silk Use
major-ampullate (dragline) silk Used for the web's outer rim and spokes and the lifeline. Can be as strong per unit weight as steel, but much tougher.
capture-spiral (flagelliform) silk Used for the capturing lines of the web. Sticky, extremely stretchy and tough. The capture spiral is sticky due to droplets of aggregate (a spider glue) that is placed on the spiral. The elasticity of flagelliform allows for enough time for the aggregate to adhere to the aerial prey flying into the web.
tubiliform (a.k.a. cylindriform) silk Used for protective egg sacs. Stiffest silk.
aciniform silk Used to wrap and secure freshly captured prey. Two to three times as tough as the other silks, including dragline.
minor-ampullate silk Used for temporary scaffolding during web construction.
Piriform (pyriform) Piriform serves as the attachment disk to dragline silk. Piriform is used in attaching spider silks together to construct a stable web.

Structural

Macroscopic structure down to protein hierarchy

Structure of spider silk. Inside a typical fiber there are crystalline regions separated by amorphous linkages. The crystals are beta-sheets that have assembled together.

Silks, as well as many other biomaterials, have a hierarchical structure (e.g., cellulose, hair). The primary structure is its amino acid sequence, mainly consisting of highly repetitive glycine and alanine blocks,[23][24] which is why silks are often referred to as a block co-polymer. On a secondary structure level, the short side chained alanine is mainly found in the crystalline domains (beta sheets) of the nanofibril, glycine is mostly found in the so-called amorphous matrix consisting of helical and beta turn structures.[24][25] It is the interplay between the hard crystalline segments, and the strained elastic semi-amorphous regions, that gives spider silk its extraordinary properties.[26][27] Various compounds other than protein are used to enhance the fiber's properties. Pyrrolidine has hygroscopic properties which keeps the silk moist while also warding off ant invasion. It occurs in especially high concentration in glue threads. Potassium hydrogen phosphate releases protons in aqueous solution, resulting in a pH of about 4, making the silk acidic and thus protecting it from fungi and bacteria that would otherwise digest the protein. Potassium nitrate is believed to prevent the protein from denaturing in the acidic milieu.[28]

This first very basic model of silk was introduced by Termonia in 1994[29] suggested crystallites embedded in an amorphous matrix interlinked with hydrogen bonds. This model has refined over the years: Semi-crystalline regions were found[24] as well as a fibrillar skin core model suggested for spider silk,[30] later visualized by AFM and TEM.[31] Sizes of the nanofibrillar structure and the crystalline and semi-crystalline regions were revealed by neutron scattering.[32]

It has been possible to relate microstructural information and macroscopic mechanical properties of the fibers.[33] The results show that ordered regions (i) mainly reorient by deformation for low-stretched fibers and (ii) the fraction of ordered regions increase progressively for higher stretching of the fibers.

Non-protein composition

Various compounds other than protein are found in spider silks, such as sugars, lipids, ions, and pigments that might affect the aggregation behaviour and act as a protection layer in the final fiber.[16]

Biosynthesis

The production of silks, including spider silk, differs in an important respect from the production of most other fibrous biological materials: rather than being continuously grown as keratin in hair, cellulose in the cell walls of plants, or even the fibers formed from the compacted faecal matter of beetles,[16] it is "spun" on demand from liquid silk precursor sometimes referred to as unspun silk dope, out of specialised glands.

The spinning process occurs when a fiber is pulled away from the body of a spider, be that by the spider’s legs, by the spider's falling and using its own weight, or by any other method including being pulled by humans. The name "spinning" is misleading as no rotation of any component occurs, but the name comes from when it was thought that spiders produced their thread in a similar manner to the spinning wheels of old. In fact the process is a pultrusion[34]—similar to extrusion, with the subtlety that the force is induced by pulling at the finished fiber rather than being squeezed out of a reservoir of some kind.

The unspun silk dope is pulled through silk glands, of which there may be both numerous duplicates and also different types on any one spider species.

Silk gland

The gland's visible, or external, part is termed the spinneret. Depending on the complexity of the species, spiders will have two to eight spinnerets, usually in pairs. There exist highly different specialised glands in different spiders, ranging from simply a sac with an opening at one end, to the complex, multiple-section Major Ampullate glands of the Nephila golden orb weaving spiders.[35]

Behind each spinneret visible on the surface of the spider lies a gland, a generalised form of which is shown in the figure to the right, "Schematic of a generalised gland".

Schematic of a generalised gland of a Golden silk orb-weaver. Each differently coloured section highlights a discrete section of the gland[36][37][38]

The gland described here will be based upon the major ampullate gland from a golden orb weaving spiders as they are the most-studied and presumed to be the most complex.

  1. The first section of the gland labelled 1 on Figure 1 is the secretory or tail section of the gland. The walls of this section are lined with cells that secrete proteins Spidroin I and Spidroin II, the main components of this spider’s dragline. These proteins are found in the form of droplets that gradually elongate to form long channels along the length of the final fiber, hypothesized to assist in preventing crack formation or even self-healing of the fiber.[39]
  2. The second section is the storage sac. This stores and maintains the gel-like unspun silk dope until it is required by the spider. In addition to storing the unspun silk gel, it secretes proteins that coat the surface of the final fiber.[17]
  3. The funnel rapidly reduces the large diameter of the storage sac to the small diameter of the tapering duct.
  4. The final length is the tapering duct, the site of most of the fiber formation. This consists of a tapering tube with several tight about turns, a valve almost at the end (mentioned in detail at point No. 5 below) ending in a spigot from which the silk fiber emerges. The tube here tapers hyperbolically, therefore the unspun silk is under constant shear stress, which is an important factor in fiber formation. This section of the duct is lined with cells that exchange ions and remove water from the fiber. The spigot at the end has lips that clamp around the fiber, controlling fiber diameter and further retaining water.
  5. Almost at the end of the tapering duct is a valve, approximate position marked "5" on figure 1. Though discovered some time ago, the precise purpose of this valve is still under discussion. It is believed to assist in restarting and rejoining broken fibers[40] acting much in the way of a helical pump, regulating the thickness of the fiber,[34] and/ or clamping the fiber as a spider falls upon it.[40][41] There is some discussion on the similarity of the silk worm’s silk press and the roles each of these valves play in the production of silk in these two organisms.

Throughout the process the unspun silk appears to have a nematic texture,[42] in a similar manner to a liquid crystal. This allows the unspun silk to flow through the duct as a liquid but maintain a molecular order.

As an example of a complex spinning field, the spinneret apparatus of an adult Araneus diadematus (garden cross spider) consists of the following glands:[28]

Artificial synthesis

Single strand of artificial spider silk produced under laboratory conditions.

In order to artificially synthesize spider silk into fibers, there are two broad areas that must be covered. These are synthesis of the feedstock (the unspun silk dope in spiders), and synthesis of the spinning conditions (the funnel, valve, tapering duct, and spigot). There have been a number of different approaches discussed below.

Feedstock

As discussed in the Structural section of the article, the molecular structure of unspun silk is both complex and extremely long. Though this endows the silk fibers with their desirable properties, it also makes replication of the fiber somewhat of a challenge. Various organisms have been used as a basis for attempts to replicate some components or all of some or all of the proteins involved. These proteins must then be extracted, purified and then spun before their properties can be tested. The table below shows the results including the true gold standard - actual stress and strain of the fibers as compared to the best spider dragline.

Organism Details Average Maximum breaking stress (MPa) Average Strain (%) Reference
Gold Standard: Darwin’s bark spider (Caerostris darwini) Malagasy spider famed for making webs with strands up to 25m long, across rivers. "C. darwini silk is more than twice as tough as any previously described silk" 1850 ±350 33 ±0.08 [20]
Gold Standard: Nephila clavipes Typical golden orb weaving spider 710–1200 18–27 [43][44]
Bombyx mori Silkworms Silkworms were genetically altered to express spider proteins and fibers measured.[45] 660 18.5 [46]
E. coli Synthesizing a large and repetitive molecule (~300 kDa) is complex, but required for the best silk. Here a 285 kDa protein was produced and spun. 508 ±108 15 ±5 [47]
Goats Goats were genetically modified to secrete silk proteins in their milk, which could then be purified. 285–250 30–40 [48]
Tobacco & potato plants Tobacco and potato plants were genetically modified to produce silk proteins. Patents were granted,[49] but no fibers have yet been described in the literature. n/a n/a [50]

Geometry

As was shown in the biosynthesis section, spider silks with comparatively simple molecular structure need complex ducts to be able to spin an effective fiber. There have been a number of methods used to produce fibers, of which the main types are briefly discussed below.

Syringe and needle

Feedstock is simply forced through a hollow needle using a syringe. This method has been shown to make fibers successfully on multiple occasions.[51][52]

Although very cheap and easy to assemble, the shape and conditions of the gland are very loosely approximated. Fibers created using this method may need encouragement to change from liquid to solid by removing the water from the fiber with such chemicals as the environmentally undesirable methanol[53] or acetone,[52] and also may require post-stretching of the fiber to attain fibers with desirable properties.[47][51]

Microfluidics

As the field of microfluidics matures, it is likely that more attempts to spin fibers will be made using microfluidics. These have the advantage of being very controllable and able to test spin very small volumes of unspun fiber[54][55] but setup and development costs are likely to be high. A patent has been granted in this area for spinning fibers in a method mimicking the process found in nature, and fibers are successfully being continuously spun by a commercial company.[56]

Electrospinning

Electrospinning is a very old technique whereby a fluid is held in a container in a manner such that it is able to flow out through capillary action. A conducting substrate is positioned below, and a large difference in electrical potential is applied between the fluid and the substrate. The fluid is attracted to the substrate, and tiny fibers jump almost instantly from their point of emission, the Taylor cone, to the substrate, drying as they travel. This method has been shown to create nano-scale fibers from both silk dissected from organisms and regenerated silk fibroin.

Other artificial shapes formed from silk

Silk can be formed into other shapes and sizes such as spherical capsules for drug delivery, cell scaffolds and wound healing, textiles, cosmetics, coatings, and many others.[57]

Research milestones

Due to spider silk being a scientific research field with a long and rich history, there can be unfortunate occurrences of researchers independently rediscovering previously published findings. What follows is a table of the discoveries made in each of the constituent areas, acknowledged by the scientific community as being relevant and significant by using the metric of scientific acceptance, citations. Thus, only papers with 50 or more citations are included.

Table of significant papers (50 or more citations)
Area of contribution Year Main researchers [Ref] Title of paper Contribution to the field
Chemical Basis 1960 Fischer, F. & Brander, J.[58] "Eine Analyse der Gespinste der Kreuzspinne" (Amino acid composition analysis of spider silk)
1960 Lucas, F. & et al.[59][60] "The Composition of Arthropod Silk Fibrons; Comparative studies of fibroins"
Gene Sequence 1990 Xu, M. & Lewis, R. V.[61] "Structure of a Protein Superfiber - Spider Dragline Silk"
Mechanical Properties 1964 Lucas, F.[62] "Spiders and their silks" First time compared mechanical properties of spider silk with other materials in a scientific paper.
1989 Vollrath, F. & Edmonds, D. T.[63] "Modulation of the Mechanical Properties of Spider Silk by Coating with Water" First important paper suggesting the water interplay with spider silk fibroin modulating the properties of silk.
2001 Vollrath, F. & Shao, Z.Z.[64] "The effect of spinning conditions on the mechanics of a spider's dragline silk"
2006 Plaza, G.R., Guinea, G.V., Pérez-Rigueiro, J. & Elices, M.[10] "Thermo-hygro-mechanical behavior of spider dragline silk: Glassy and rubbery states" Combined effect of humidity and temperature on the mechanical properties. Glass-transition temperature dependence on humidity.
Structural Characterization 1992 Hinman, M.B. & Lewis, R. V[23] "Isolation of a clone encoding a second dragline silk fibroin. Nephila clavipes dragline silk is a two-protein fiber"
1994 Simmons, A. & et al.[65] "Solid-State C-13 Nmr of Nephila-Clavipes Dragline Silk Establishes Structure and Identity of Crystalline Regions" First NMR study of spider silk.
1999 Shao, Z., Vollrath, F. & et al.[66] "Analysis of spider silk in native and supercontracted states using Raman spectroscopy" First Raman study of spider silk.
1999 Riekel, C., Muller, M.& et al.[67] "Aspects of X-ray diffraction on single spider fibers" First X-ray on single spider silk fibers.
2000 Knight, D.P., Vollrath, F. & et al.[68] "Beta transition and stress-induced phase separation in the spinning of spider dragline silk" Secondary structural transition confirmation during spinning.
2001 Riekel, C. & Vollrath, F.[69] "Spider silk fibre extrusion: combined wide- and small-angle X- ray microdiffraction experiments" First X-ray on spider silk dope.
2002 Van Beek, J. D. & et al.[25] "The molecular structure of spider dragline silk: Folding and orientation of the protein backbone"
Structure-Property Relationship 1986 Gosline, G.M. & et al.[70] "The structure and properties of spider silk" First attempt to link structure with properties of spider silk
1994 Termonia, Y[29] "Molecular Modeling of Spider Silk Elasticity" X-ray evidence presented in this paper; simple model of crystallites embedded in amorphous regions.
1996 Simmons, A. & et al.[24] "Molecular orientation and two-component nature of the crystalline fraction of spider dragline silk" Two types of alanine-rich crystalline regions were defined.
2006 Vollrath, F. & Porter, D.[71] "Spider silk as an archetypal protein elastomer" New insight and model to spider silk based on Group Interaction Modelling.
Native Spinning 1991 Kerkam, K., Kaplan, D. & et al.[72] "Liquid Crystallinity of Natural Silk Secretions"
1999 Knight, D.P. & Vollrath, F.[73] "Liquid crystals and flow elongation in a spider's silk production line"
2001 Vollrath, F. & Knight, D.P.[16] "Liquid crystalline spinning of spider silk" The most cited paper on spider silk
2005 Guinea, G.V., Elices, M., Pérez-Rigueiro, J. & Plaza, G.R.[9] "Stretching of supercontracted fibers: a link between spinning and the variability of spider silk" Explanation of the variability of mechanical properties.
Reconstituted /Synthetic Spider Silk and Artificial Spinning 1995 Prince, J. T., Kaplan, D. L. & et al.[74] "Construction, Cloning, and Expression of Synthetic Genes Encoding Spider Dragline Silk" First successful synthesis of Spider silk by E. coli.
1998 Arcidiacono, S., Kaplan, D.L. & et al.[75] "Purification and characterization of recombinant spider silk expressed in Escherichia coli"
1998 Seidel, A., Jelinski, L.W. & et al.[76] "Artificial Spinning of Spider Silk" First controlled wet-spinning of reconstituted spider silk.

Human uses

A cape made from Madagascar golden orb spider silk.[77]

Peasants in the southern Carpathian Mountains used to cut up tubes built by Atypus and cover wounds with the inner lining. It reportedly facilitated healing, and even connected with the skin. This is believed to be due to antiseptic properties of spider silk[78] and because the silk is rich in vitamin K, which can be effective in clotting blood.[79]

Some fishermen in the Indo-Pacific ocean use the web of Nephila to catch small fish.[28]

The silk of Nephila clavipes has recently been used to help in mammalian neuronal regeneration.[80]

At one time, it was common to use spider silk as a thread for crosshairs in optical instruments such as telescopes, microscopes,[81] and telescopic rifle sights.[82]

Due to the difficulties in extracting and processing substantial amounts of spider silk, the largest known piece of cloth made of spider silk is an 11-by-4-foot (3.4 by 1.2 m) textile with a golden tint made in Madagascar in 2009.[83] Eighty-two people worked for four years to collect over one million golden orb spiders and extract silk from them.[84]

In 2011, spider silk fibers were used in the field of optics to generate very fine diffraction patterns over N-slit interferometric signals utilized in optical communications.[85] In 2012, spider silk fibers were used to create a set of violin strings.[86]

Spider silk is used to suspend inertial confinement fusion targets during laser ignition, as it remains considerably elastic and has a high energy to break at temperatures as low as 10-20K. In addition, it is made from "light" atomic number elements that won't emit x-rays during irradiation that could preheat the target so that the pressure differential required for fusion is not achieved.[87]

Attempts at producing synthetic spider silk

Replicating the complex conditions required to produce fibers that are comparable to spider silk has proven difficult to accomplish in a laboratory environment. What follows is a miscellaneous list of attempts on this problem. However, in the absence of hard data accepted by the relevant scientific community, it is difficult to judge whether these attempts have been successful or constructive.

See also

References

  1. 1 2 Miyashita, Tadashi; Maezono, Yasunori; Shimazaki, Aya (2004). "Silk feeding as an alternative foraging tactic in a kleptoparasitic spider under seasonally changing environments". Journal of Zoology. 262 (3): 225–229. doi:10.1017/S0952836903004540.
  2. Work, Robert W.; Emerson, Paul D. (1982). "An Apparatus and Technique for the Forcible Silking of Spiders". Journal of Arachnology. 10 (1): 1–10. JSTOR 3705113.
  3. 1 2 3 4 Foelix, R. F. (1996). Series Editor, ed. Biology of Spiders. Oxford, N.Y: Oxford University Press. p. 330.
  4. Sutherland, TD; Young, JH; Weisman, S; Hayashi, CY; Merritt, DJ (2010). "Insect silk: One name, many materials". Annual Review of Entomology. 55: 171–88. doi:10.1146/annurev-ento-112408-085401. PMID 19728833.
  5. 1 2 3 Hillyard, P. (2007). Series Editor, ed. The Private Life of Spiders. London: New Holland. p. 160. ISBN 978-1-84537-690-1.
  6. Nentwig, W. & Heimer, S. (1987). Wolfgang Nentwig, ed. Ecological aspects of spider webs. Springer-Verlag. p. 211.
  7. Flying spiders over Texas! Coast to Coast. Chad B., Texas State University Undergrad Describes the mechanical kiting of Spider "ballooning".
  8. 1 2 3 4 5 Holm, Erik, Dippenaar-Schoeman, Ansie; Goggo Guide; LAPA publishers (URL: WWW.LAPA.co.za). 2010
  9. 1 2 Guinea, G.V.; Elices, M.; Pérez-Rigueiro, J. & Plaza, G.R. (2005). "Stretching of supercontracted fibers: a link between spinning and the variability of spider silk". Journal of Experimental Biology. 208 (1): 25–30. doi:10.1242/jeb.01344.
  10. 1 2 3 4 Plaza, Gustavo R.; Guinea, Gustavo V.; Pérez-Rigueiro, José; Elices, Manuel (2006). "Thermo-hygro-mechanical behavior of spider dragline silk: Glassy and rubbery states". Journal of Polymer Science Part B: Polymer Physics. 44 (6): 994–999. Bibcode:2006JPoSB..44..994P. doi:10.1002/polb.20751.
  11. Griffiths, J. R.; Salanitri, V. R. (1980). "The strength of spider silk". Journal of Materials Science. 15 (2): 491–6. Bibcode:1980JMatS..15..491G. doi:10.1007/BF00551703.
  12. "Overview of materials for AISI 4000 Series Steel". www.matweb.com. Retrieved 18 August 2010.
  13. "DuPont Kevlar 49 Aramid Fiber". www.matweb.com. Retrieved 18 August 2010.
  14. Ganio Mego, Paolo (c. 2002). "Material Tensile Strength Comparison". Archived from the original on 26 October 2009. Retrieved 3 January 2012.
  15. Shao, Zhengzhong; Vollrath, F (2002). "Materials: Surprising strength of silkworm silk". Nature. 418 (6899): 741. Bibcode:2002Natur.418..741S. doi:10.1038/418741a. PMID 12181556.
  16. 1 2 3 4 Porter, D.; Vollrath, F.; Shao, Z. (2005). "Predicting the mechanical properties of spider silk as a model nanostructured polymer". European Physical Journal E. 16 (2): 199–206. Bibcode:2005EPJE...16..199P. doi:10.1140/epje/e2005-00021-2.
  17. 1 2 Vollrath, F. & Knight, D. P. (2001). "Liquid crystalline spinning of spider silk". Nature. 410 (6828): 541–548. Bibcode:2001Natur.410..541V. doi:10.1038/35069000. PMID 11279484.
  18. "Spider Silk". www.chm.bris.ac.uk. Retrieved 18 August 2010.
  19. Yang, Y.; Chen, X.; Shao, Z.; Zhou, P.; Porter, D.; Knight, D. P.; Vollrath, F. (2005). "Toughness of Spider Silk at High and Low Temperatures". Advanced Materials. 17: 84–88. doi:10.1002/adma.200400344.
  20. 1 2 Agnarsson, Ingi; Kuntner, Matjaž; Blackledge, Todd A. (2010). Lalueza-Fox, Carles, ed. "Bioprospecting Finds the Toughest Biological Material: Extraordinary Silk from a Giant Riverine Orb Spider". PLoS ONE. 5 (9): 11234. Bibcode:2010PLoSO...511234A. doi:10.1371/journal.pone.0011234. PMC 2939878Freely accessible. PMID 20856804.
  21. Cunningham, Aimee (2009). "Taken for a spin: Scientists look to spiders for the goods on silk". Science News. 171 (15): 231–234. doi:10.1002/scin.2007.5591711509.
  22. Blackledge, TA; Hayashi, CY (2006). "Silken toolkits: Biomechanics of silk fibers spun by the orb web spider Argiope argentata (Fabricius 1775)". The Journal of Experimental Biology. 209 (Pt 13): 2452–61. doi:10.1242/jeb.02275. PMID 16788028.
  23. 1 2 Hinman, M. B. & Lewis, R. V. (1992). "Isolation of a clone encoding a second dragline silk fibroin. Nephila clavipes dragline silk is a two-protein fiber". J. Biol. Chem. 267 (27): 19320–19324. PMID 1527052.
  24. 1 2 3 4 Simmons, A. H.; Michal, C. A. & Jelinski, L. W. (1996). "Molecular orientation and two-component nature of the crystalline fraction of spider dragline silk". Science. 271 (5245): 84–87. Bibcode:1996Sci...271...84S. doi:10.1126/science.271.5245.84. PMID 8539605.
  25. 1 2 van Beek, J. D.; Hess, S.; Vollrath, F. & Meier, B. H. (2002). "The molecular structure of spider dragline silk: Folding and orientation of the protein backbone". Proc. Natl. Acad. Sci. U.S.A. 99 (16): 10266–10271. Bibcode:2002PNAS...9910266V. doi:10.1073/pnas.152162299. PMC 124902Freely accessible. PMID 12149440.
  26. Liu, Y.; Sponner, A.; Porter, D.; Vollrath, F. (2008). "Proline and Processing of Spider Silks". Biomacromolecules. 9 (1): 116–121. doi:10.1021/bm700877g. PMID 18052126.
  27. Papadopoulos, P.; Ene, R.; Weidner, I.; Kremer, F. (2009). "Similarities in the Structural Organization of Major and Minor Ampullate Spider Silk". Macromol. Rapid Commun. 30 (9–10): 851–857. doi:10.1002/marc.200900018.
  28. 1 2 3 Heimer, S. (1988). Wunderbare Welt der Spinnen. Urania. p.12
  29. 1 2 Termonia, Y. (1994). "Molecular Modeling of Spider Silk Elasticity". Macromolecules. 27 (25): 7378–7381. Bibcode:1994MaMol..27.7378T. doi:10.1021/ma00103a018.
  30. Vollrath, F.; Holtet, T.; Thogersen, H. C. & Frische, S. (1996). "Structural organization of spider silk". Proceedings of the Royal Society B. 263 (1367): 147–151. doi:10.1098/rspb.1996.0023.
  31. Sponner, A.; Vater, Wolfram, Wolfram; Monajembashi, Shamci, Shamci; Unger, Eberhard, Eberhard; Grosse, Frank, Frank; Weisshart, Klaus, Klaus (2007). Scheibel, Thomas, ed. "Composition and hierarchical organization of a spider silk". PLoS ONE. 2 (10): e998. Bibcode:2007PLoSO...2..998S. doi:10.1371/journal.pone.0000998. PMC 1994588Freely accessible. PMID 17912375.
  32. Sapede, D.; Seydel, T.; Forsyth, V. T.; Koza, M. M.; Schweins, R.; Vollrath, F.; Riekel, C. (2005). "Nanofibrillar structure and molecular mobility in spider dragline silk". European Biophysics Journal. 34 (20): 623. Bibcode:2005MaMol..38.8447S. doi:10.1021/ma0507995.
  33. Plaza, G.R.; Pérez-Rigueiro, J.; Riekel, C.; Perea, G.B.; Agulló-Rueda, F.; Burghammer, M.; Guinea, G.V.; Elices, M. (2012). "Relationship between microstructure and mechanical properties in spider silk fibers: identification of two regimes in the microstructural changes". Soft Matter. 8 (22): 6015–6026. doi:10.1039/C2SM25446H.
  34. 1 2 Wilson, R. S. (1969). "control of drag-line spinning in certain spiders". Am. Zool. 9: 103–&. doi:10.1093/icb/9.1.103.
  35. Dicko, C.; Porter, D.; Bond, J.; Kenney, J. M. & Vollratht, F. (2008). "Structural disorder in silk proteins reveals the emergence of elastomericity". Biomacromolecules. 9 (1): 216–221. doi:10.1021/bm701069y. PMID 18078324.
  36. Lefèvre, T.; Boudreault, S.; Cloutier, C. & Pézolet, M. (2008). "Conformational and orientational transformation of silk proteins in the major ampullate gland of Nephila clavipes spiders". Biomacromolecules. 9 (9): 2399–2407. doi:10.1021/bm800390j. PMID 18702545.
  37. Heim, M.; Keerl, D. & Scheibel, T. (2009). "Spider Silk: From Soluble Protein to Extraordinary Fiber". Angewandte Chemie International Edition. 48 (20): 3584–3596. doi:10.1002/anie.200803341.
  38. Knight, D. P. & Vollrath, F. (2001). "Comparison of the spinning of Selachian egg case ply sheets and orb web spider dragline filaments". Biomacromolecules. 2 (2): 323–334. doi:10.1021/bm0001446. PMID 11749189.
  39. Heinhorst, S.; Cannon, G. (2002). "Nature: Self-Healing Polymers and Other Improved Materials". J. Chem. Educ. 79: 10. Bibcode:2002JChEd..79...10H. doi:10.1021/ed079p10.
  40. 1 2 Vollrath, F. & Knight, D. P. (1998). "Structure and function of the silk production pathway in spider Nephila edulis". Int J Biol Macromol. 24 (2–3): 243–249. doi:10.1016/S0141-8130(98)00095-6. PMID 10342771.
  41. Wilson, R. S. (1962). "The Control of Dragline Spinning in the Garden Spider". Quarterly Journal of Microscopical Science. 103: 557–571.
  42. Magoshi, J.; Magoshi, Y. & Nakamura, S. (1985). "Physical properties and structure of silk: 9. Liquid crystal formation of silk fibroin". Polym. Commun. 26: 60–61.
  43. Elices, M.; Plaza, G.R.; Arnedo, M.A.; Perez-Rigueiro, J.; Torres, F.G. & Guinea, G. (2009). "Mechanical Behavior of Silk During the Evolution of Orb-Web Spinning Spiders". Biomacromolecules. 10 (7): 1904–1910. doi:10.1021/bm900312c. PMID 19505138.
  44. Swanson, B. O.; Blackledge, T. A.; Summers, A. P. & Hayashi, C. Y. (2006). "Spider dragline silk: Correlated and mosaic evolution in high-performance biological materials". Evolution. 60 (12): 2539–2551. doi:10.1554/06-267.1. PMID 17263115.
  45. Shao, Z. Z. & Vollrath, F. (2002). "Materials: Surprising strength of silkworm silk". Nature. 418 (6899): 741–741. Bibcode:2002Natur.418..741S. doi:10.1038/418741a. PMID 12181556.
  46. Wen, H. X.; et al. (2010). "Transgenic silkworms (Bombyx mori) produce recombinant spider dragline silk in cocoons". Molecular Biology Reports. 37 (4): 1815–1821. doi:10.1007/s11033-009-9615-2. PMID 19633923.
  47. 1 2 Xia, X. X.; et al. (2010). "Native-sized recombinant spider silk protein produced in metabolically engineered Escherichia coli results in a strong fiber". Proceedings of the National Academy of Sciences of the United States of America. 107 (32): 14,059–14,063. Bibcode:2010PNAS..10714059X. doi:10.1073/pnas.1003366107.
  48. Elices, M.; Guinea, G. V.; Plaza, G. R.; Karatzas, C.; Riekel, C.; Agulló-Rueda, F.; Daza, R.; Pérez-Rigueiro, J. (2011). "Bioinspired Fibers Follow the Track of Natural Spider Silk". Macromolecules. 44 (5): 1166–1176. Bibcode:2011MaMol..44.1166E. doi:10.1021/ma102291m.
  49. US patent 2008109923, Lewis, R. V., "Expression of spider silk proteins", published 2010-05-25, assigned to University of Wyoming
  50. Scheller, J. & Conrad, U. (2005). "Plant-based material, protein and biodegradable plastic". Current Opinion in Plant Biology. 8 (2): 188–196. doi:10.1016/j.pbi.2005.01.010. PMID 15753000.
  51. 1 2 Lazaris, A.; Arcidiacono,S, S; Huang,Y, Y; Zhou,J. F., JF; Duguay,F, F; Chretien,N, N; Welsh,E. A., EA; Soares,J. W., JW; Karatzas,C. N., CN (2002). "Spider silk fibers spun from soluble recombinant silk produced in mammalian cells". Science. 295 (5554): 472–476. Bibcode:2002Sci...295..472L. doi:10.1126/science.1065780. PMID 11799236.
  52. 1 2 Seidel, A.; Liivak, Oskar; Calve, Sarah; Adaska, Jason; Ji, Gending; Yang, Zhitong; Grubb, David; Zax, David B.; Jelinski, Lynn W. (2000). "Regenerated spider silk: Processing, properties, and structure". Macromolecules. 33 (3): 775–780. Bibcode:2000MaMol..33..775S. doi:10.1021/ma990893j.
  53. Arcidiacono, S.; Mello, Charlene M.; Butler, Michelle; Welsh, Elizabeth; Soares, Jason W.; Allen, Alfred; Ziegler, David; Laue, Thomas; Chase, Susan (2002). "Aqueous processing and fiber spinning of recombinant spider silks". Macromolecules. 35 (4): 1262–1266. Bibcode:2002MaMol..35.1262A. doi:10.1021/ma011471o.
  54. Kinahan, M. E.; et al. (2011). "Tunable Silk: Using Microfluidics to Fabricate Silk Fibers with Controllable Properties". Biomacromolecules. 12 (5): 1504–1511. doi:10.1021/bm1014624. PMC 3305786Freely accessible. PMID 21438624.
  55. Rammensee, S.; Slotta, U.; Scheibel, T. & Bausch, A. R. (2008). "Assembly mechanism of recombinant spider silk proteins (microfluidic)". Proceedings of the National Academy of Sciences of the United States of America. 105 (18): 6590–6595. Bibcode:2008PNAS..105.6590R. doi:10.1073/pnas.0709246105. PMC 2373321Freely accessible. PMID 18445655.
  56. Spintec Engineering GmbH (German)
  57. Eisoldt, L.; Smith, A. & Scheibel, T. (2011). "Decoding the secrets of spider silk". Mater. Today. 14 (3): 80–86. doi:10.1016/s1369-7021(11)70057-8.
  58. Fischer, F. & Brander, J. (1960). "Eine Analyse der Gespinste der Kreuzspinne". Hoppe-Seylers Zeitschrift für Physiologische Chemie. 320: 92–102. doi:10.1515/bchm2.1960.320.1.92.
  59. Lucas, F.; Shaw, J. T. B. & Smith, S. G. (1960). "The Composition of Arthropod Silk Fibrons". Insect Chemistry. Symp. 3: 208–214.
  60. Lucas, F.; Shaw, J. T. B. & Smith, S. G. (1960). "Comparative studies of fibroins.I. The amino acid composition of various fibroins and its significance in relation to their crystal structure and taxonomy". Journal of Molecular Biology. 2 (6): 339–349. doi:10.1016/S0022-2836(60)80045-9. PMID 13763962.
  61. Xu, M. & Lewis, R. V. (1990). "Structure of a Protein Superfiber - Spider Dragline Silk". Proceedings of the National Academy of Sciences of the United States of America. 87 (18): 7120–7124. Bibcode:1990PNAS...87.7120X. doi:10.1073/pnas.87.18.7120. PMC 54695Freely accessible. PMID 2402494.
  62. Lucas, F. (1964). "Spiders and their silks". Discovery. 25: 20–26.
  63. Vollrath, F. & Edmonds, D. T. (1989). "Modulation Of The Mechanical-Properties Of Spider Silk By Coating With Water". Nature. 340 (6231): 305–307. Bibcode:1989Natur.340..305V. doi:10.1038/340305a0.
  64. Vollrath, F.; Madsen, B. & Shao, Z. Z. (2001). "The effect of spinning conditions on the mechanics of a spider's dragline silk". Proceedings of the Royal Society B. 268 (1483): 2339–2346. doi:10.1098/rspb.2001.1590.
  65. Simmons, A.; Ray, E. & Jelinski, L. W. (1994). "Solid-State C-13 NMR of Nephila-Clavipes Dragline Silk Establishes Structure and Identity of Crystalline Regions". Macromolecules. 27 (18): 5235–5237. Bibcode:1994MaMol..27.5235S. doi:10.1021/ma00096a060.
  66. Shao, Z.; Vollrath, F.; Sirichaisit, J. & Young, R. J. (1999). "Analysis of spider silk in native and supercontracted states using Raman spectroscopy". Polymer. 40 (10): 2493–2500. doi:10.1016/S0032-3861(98)00475-3.
  67. Riekel, C.; Bränden, C; Craig, C; Ferrero, C; Heidelbach, F; Müller, M (1999). "Aspects of X-ray diffraction on single spider fibers". Int. J. Biol. Macromol. 24 (2–3): 179–186. doi:10.1016/S0141-8130(98)00084-1. PMID 10342763.
  68. Knight, D. P.; Knight, M. M. & Vollrath, F. (2000). "Beta transition and stress-induced phase separation in the spinning of spider dragline silk". Int. J. Biol. Macromol. 27 (3): 205–210. doi:10.1016/S0141-8130(00)00124-0. PMID 10828366.
  69. Riekel, C. & Vollrath, F. (2001). "Spider silk fibre extrusion: combined wide- and small-angle X- ray microdiffraction experiments". Int. J. Biol. Macromol. 29 (3): 203–210. doi:10.1016/S0141-8130(01)00166-0. PMID 11589973.
  70. Gosline, J. M.; DeMont, M. E. & Denny, M. W. (1986). "The structure and properties of spider silk". Endeavour. 10: 37–43. doi:10.1016/0160-9327(86)90049-9.
  71. Vollrath, F. & Porter, D. (2006). "Spider silk as an archetypal protein elastomer". Soft Matter. 2 (5): 377–385. Bibcode:2006SMat....2..377V. doi:10.1039/b600098n.
  72. Kerkam, K.; Viney, C.; Kaplan, D. & Lombardi, S. (1991). "Liquid Crystallinity of Natural Silk Secretions". Nature. 349 (6310): 596–598. Bibcode:1991Natur.349..596K. doi:10.1038/349596a0.
  73. Knight, D. P. & Vollrath, F. (1999). "Liquid crystals and flow elongation in a spider's silk production line". Proceedings of the Royal Society B. 266 (1418): 519–523. doi:10.1098/rspb.1999.0667.
  74. Prince, J. T.; McGrath, K. P.; Digirolamo, C. M. & Kaplan, D. L. (1995). "Construction, Cloning, and Expression of Synthetic Genes Encoding Spider Dragline Silk". Biochemistry. 34 (34): 10879–10885. doi:10.1021/bi00034a022. PMID 7662669.
  75. Arcidiacono, S.; Mello, C.; Kaplan, D.; Cheley, S. & Bayley, H. (1998). "Purification and characterization of recombinant spider silk expressed in Escherichia coli". Applied Microbiology and Biotechnology. 49 (1): 31–38. doi:10.1007/s002530051133. PMID 9487707.
  76. Seidel, A.; Liivak, O. & Jelinski, L. W. (1998). "Artificial Spinning of Spider Silk". Macromolecules. 31 (19): 6733–6736. Bibcode:1998MaMol..31.6733S. doi:10.1021/ma9808880.
  77. Maev Kennedy. "Spider silk cape goes on show at V&A". the Guardian.
  78. Heimer, S. (1988). Wunderbare Welt der Spinnen. Urania. p.14
  79. Jackson, Robert R. (1974). "Effects of D-Amphetamine Sulfate and Diazepam on Thread Connection Fine Structure in a Spider's Web". Journal of Arachnology. 2 (1): 37–41. JSTOR 3704994.
  80. Allmeling, Christina; Jokuszies, Andreas; Reimers, Kerstin; Kall, Susanne; Vogt, Peter M. (2006). "Use of spider silk fibres as an innovative material in a biocompatible artificial nerve conduit". Journal of cellular and molecular medicine. 10 (3): 770–7. doi:10.1111/j.1582-4934.2006.tb00436.x. PMID 16989736.
  81. Berenbaum, May R., Field Notes – Spin Control, The Sciences, The New York Academy Of Sciences, September/October 1995
  82. Example of use of spider silk for telescopic rifle sights. Books.google.co.uk. April 1955. Retrieved 24 August 2011.
  83. British art historian creates world's largest shawl woven from the silk of one million spiders Mail Online
  84. Leggett, Hadley (23 September 2009). "1 Million Spiders Make Golden Silk for Rare Cloth". Wired.
  85. Duarte F. J.; Taylor, T S; Black, A M; Davenport, W E; Varmette, P G (2011). "N-slit interferometer for secure free-space optical communications: 527 m intra interferometric path length". Journal of Optics. 13 (3): 5710. Bibcode:2011JOpt...13c5710D. doi:10.1088/2040-8978/13/3/035710.
  86. Osaki, Shigeyoshi (2012). "Spider Silk Violin Strings with a Unique Packing Structure Generate a Soft and Profound Timbre". Physical Review Letters. 108 (15): 154301. Bibcode:2012PhRvL.108o4301O. doi:10.1103/PhysRevLett.108.154301. PMID 22587257.
  87. http://www.lle.rochester.edu/media/publications/documents/theses/Bonino.pdf
  88. www.nexiabiotech.com Archived 1 September 2013 at the Wayback Machine.
  89. Kluge, Jonathan A.; Rabotyagova, Olena; Leisk, Gary G.; Kaplan, David L. (May 2008). "Spider silks and their applications". Trends in Biotechnology. 26 (5): 244–251. doi:10.1016/j.tibtech.2008.02.006. Retrieved 3 January 2012.
  90. Scheibel, Thomas (November 2004). "Spider silks: recombinant synthesis, assembly, spinning, and engineering of synthetic proteins". Microbial Cell Factories. 3 (1): 14. doi:10.1186/1475-2859-3-14. PMC 534800Freely accessible. PMID 15546497. Retrieved 3 January 2012.
  91. Xia, Xiao-Xia; Qian, Zhi-Gang; Ki, Chang Seok; Park, Young Hwan; Kaplan, David L.; Lee, Sang Yup (2010). "Native-sized recombinant spider silk protein produced in metabolically engineered Escherichia coli results in a strong fiber". Proceedings of the National Academy of Sciences. 107 (32): 14059–14063. Bibcode:2010PNAS..10714059X. doi:10.1073/pnas.1003366107. JSTOR 25708855.
  92. "University of Notre Dame and Kraig Biocraft Laboratories Create Artificial Spider Silk Breakthrough" (Press release). Kraig Biocraft Laboratories. 29 September 2010. Retrieved 3 January 2012.
  93. "Fraser Research Publicly Announced at Press Conference" (Press release). University of Notre Dame. 1 October 2010. Archived from the original on 10 October 2010. Retrieved 3 January 2012.
  94. "Draadkracht: spindoctors maken supersterk nepweb" [Wire strength: spin doctors make super strong fake cobweb] (in Dutch). KIJK. 21 April 2012. Retrieved 15 October 2014.
  • Forbes, Peter (4th Estate, London 2005). The Gecko's Foot – Bio Inspiration: Engineered from Nature, ISBN 0-00-717990-1 in H/B.
  • Graciela C. Candelas, José Cintron. "A spider fibroin and its synthesis", Journal of Experimental Zoology (1981), Department of Biology, University of Puerto Rico, Río Piedras, Puerto Rico 00931.
Wikinews has related news: Spiders' egg case silk gene found
Wikisource has the text of the 1905 New International Encyclopedia article Gossamer.
This article is issued from Wikipedia - version of the 11/14/2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.