Topical Review

Nanomechanics of silk: the fundamentals of a strong, tough and versatile material

and

Published 16 June 2016 © 2016 IOP Publishing Ltd
, , Citation Isabelle Su and Markus J Buehler 2016 Nanotechnology 27 302001 DOI 10.1088/0957-4484/27/30/302001

0957-4484/27/30/302001

Abstract

Spider silk is a remarkable material that provides a template for upscaling molecular properties to the macroscale. In this article we review fundamental aspects of the mechanisms behind these behaviors, discuss the molecular makeup, chemical designs, and how these integrate in a complex arrangement to form webs, cocoons and other material architectures. Moreover, this review paper explores the unique ability of silk to tolerate various kinds of defects, in a way enabling this material platform to serve as one of the most resilient materials in nature. We conclude the discussion with a summary of key scaling laws, an attempt model and define hierarchical length-scales, and the translation to synthetic materials.

Export citation and abstract BibTeX RIS

1. Introduction

Biological materials are known for their highly optimized structure and functions. They often surpass synthetic materials in strength and toughness. They have become a source of inspiration for high-performance material design [15]. One of the most fascinating natural materials is silk, as it combines outstanding mechanical properties: strength, toughness and robustness. In particular, spiders are able to produce up to seven types of silks, each optimized to fulfill different specific functions [69]. For instance, spider webs are composed of structural radial dragline silk threads and spiral viscid and more extensible silk for the capture of prey [6, 10, 11]. Combining silk properties and spider web architectures, the webs are strong enough to withstand the weight of the spider and its prey and extensible to absorb energy from impact and environmental threats [10, 12]. With a maximum strength of up to 1.7 GPa spider dragline silk (e.g. Caerostris darwini) [13] is comparable in strength to artificial materials such as synthetic polyamide (PA66 and PET, 1 GPa) [14], steel (1.5 GPa) and Kevlar (3.6 GPa) [15]. However, most silks, spun by caterpillars, silkworms and spiders, show a large variability in their mechanical properties, with a Weibull statistical modulus ranging from 0.6 to 7, depending of their type, treatment and age, regardless of the specie [16]. Moreover, previous studies have shown that along the length of a single silk fiber, the stress–strain behavior type can be variable [1618]. The stress–strain curves are differentiated into five types, where they display a linear elastic behavior, in one (type I, III) or two steps (type II), followed by a quasi plateau with a kink (type I) or smooth transition (type III). On the other hand, type IV follows a nonlinear stiffening behavior and type V silk fails during the linear elastic stage [16, 17]. However, many silks displays remarkable extensibility with 50%–60% strain at failure, which makes silk tougher than most industrial materials, for an equal weight [6, 19, 20]. Confronted with natural threats, silk and silk structures, cocoons and webs, are extremely resilient despite being lightweight. Spider webs are designed to function even with defects and cocoon structure itself is an armor protecting the pupae inside [10, 21, 22]. Silk has the ability to adapt and tune its structure and properties, making silk a remarkably versatile material [5, 23].

The impressive tensile strength and toughness of silk is due to its hierarchical structure, ranging from protein sequence to complex silk architectures at the macroscale. Silk is strong because of the stiff β-sheet nanocrystals resulting from protein folding, and extensible thanks to the hidden length of a disorganized amorphous phase, where the core molecular structure is controlled by weak hydrogen bonds [2426].

In addition to having superior tensile properties, silk displays supercontraction in presence of water. This phenomena refers to the fact that silk fibers shrink up to 50% of their length [27, 28], when exposed to water. Moreover, silk is biodegradable and biocompatible, which offers potential for biomedical applications such as tissue regeneration, drug delivery, wound suture [2933].

Indeed, silk's remarkable mechanical and biological properties have led to intense research on its structure and production process, in order to create new innovative and high-performance materials with tailored properties. However, the production of silk and silk-inspired materials remains a challenge, because silk structure is controlled by self-assembly and the specific spinning process which are not yet fully understood [6, 3436].

Adopting a bottom-up approach, spider silk is a remarkable material that provides a template for upscaling nanoscale and molecular properties to the macroscale. In this article, we review the fundamentals of the mechanics of silk across scales from the molecular scale to the macroscale. In particular, we discuss the resilience of silk and silk architectures. Finally, we conclude in a summary of key scaling laws and the translation to synthetic silks and materials.

2. Mechanical properties of silk across scales

2.1. A hierarchical structure

Silk is a protein material and has a hierarchical organization that reaches from the atomic to the macroscale. Figure 1 describes the hierarchical organization of spider silk. At the molecular scale, silk is a sequence of amino acids which dictates the folding mechanism and the final molecular structure. Polyalalnine-rich regions fold into β-strands bonded by hydrogen bonds which form β-sheets. They are stacked into stiff β-sheet nanocrystals which are embedded in a disordered semi-amorphous phase derived from the folding of glycine-rich protein [6, 2426]. This semi-amorphous phase constitutes the majority of silk and consists of disordered secondary structures. Along the fiber axis, the distribution of crystalline regions is heterogeneous [37]. Up a level of scale, the silk nanocomposites form a network that is nanoconfined into fibrils. The fibrils are bundled together into a fiber of a few micrometers which are bundled into fibers to finally form webs [6, 2426]. For silkworm silk, the protein chain is organized into two fibroins which are bonded to each other. The pair of fibroins is coated in sericin glue to form the fiber [38, 39].

2.2. Molecular structure and spinning process

Spider silk's mechanical properties originate from the complex molecular structure of the amino acid sequence. The different types of silk have different sequences which result in different molecular structures and different mechanical properties to fulfill specific functions [40, 41]. However, the pattern between the different silks are similar. The proteins are composed mainly of repetitive amino acid motifs and bounded by non-repetitive N-termini and C-termini [42, 43]. The sequences of the termini are highly conserved, they remain the same for different silk types and spider species [8, 44]. They play a fundamental role in the storage of the spider protein in the glands and the self-assembly process for the formation of fibers [8, 43]. The common repeats are: (i) GPGGX/GPGQQ (where X is Q, L, Y or R [45]), (ii) GGX, (iii) poly-Ala/poly-Gly-Ala, and (iv) spacers [44, 46]. In particular, dragline silk from the spider N. Clavipes is known to be one of the strongest biomaterials, and is produced to build structural radial threads of webs. Dragline silk is produced by the major ampullate gland of the spider and is composed of two types of protein: major ampullate dragline silk protein 1 (MaSp1) and major ampullate dragline silk protein 2 (MaSp2) [6, 47, 48]. They represent 81% and 19% of dragline silk respectively [49]. The majority of the fiber is composed of MaSp1, while MaSp2 is found in clusters within the fiber [50, 51]. The major difference between MaSp1 and MaSp2 is the quantity of proline (P) residues which is 15% of MaSp2. There is no proline in MaSp1. The three principal repeats of MaSp1 are poly-Ala and poly-Gly-Ala and GGX. MaSp2 has the same repeats except that it has GPGGX domains instead of glycine-rich domains (GGX)[44]. The polyalanine-rich domains form stiff β-sheet nanocrystals responsible for the silk's strength. GGX motifs fold into 310-helix and GPGXX into β-spirals forming the semi-amorphous and disordered domain responsible for silk's elasticity [6, 50, 52]. The primary sequence controls the folding of the protein into its secondary structure which influences the mechanical and biological properties of silk. The behavior of silk can be predicted by understanding the role of each group of amino acids in the formation of the secondary structure (β-sheet, 310-helix, β-spirals) and their structural role (strength and elasticity). Thus, it is possible to control and improve silk's properties by modifying the protein sequence. For example, an increase of polyalanine domains would result in the formation of more β-sheet crystals, and consequently a stronger fiber. With only a few different amino acids, spiders can produce complex and versatile silks [16].

Although the primary structure is essential for the formation of strong, tough and versatile silk, it has to be associated with a particular spinning process during which certain structural features are formed. The silk produced from recombinant protein silk has lower mechanical properties than naturally spun spider silk: the spinning process is fundamental [8, 35, 36]. Dragline silk spinning process occurs in the major ampulate glands. First, in the tail, specialized cells secrete spidroins or spider protein which are then stored in the sac or ampulla. The stored spidroin is a highly concentrated aqueous crystalline solution [5356], called spinning dope (up to 50% w/v) [6, 8, 57]. Next, the dope undergoes elongation and shear flow in the S-shape tapered spinning duct [6, 8]. The elongational flow forces the spidroin to align and engenders the formation of β-sheet nanocrystals highly oriented along the fiber axis, where β-sheets are tightly stacked and bonded by hydrogen bonds. Throughout the spinning duct, the dope is subjected to a decrease in pH, from moderately basic to moderately acidic pH. Simultaneously, phosphate and potassium ions are added to the dope, while water, sodium and chloride ions are removed from the dope. Phase separation occurs and water is extracted. After dehydration the fiber is no longer soluble, from a chemical point of view [8, 5860]. Finally, the already aligned spidroin fiber is further extended and aligned after exiting from the spigot. The natural spinning process is illustrated in figure 2, during which the emerging silk fiber is pulled and stretched. A valve located before the spigot helps restart the spinning process when fibers internally break [8]. In the spinning duct, the transition from spidroin to silk fibers is explained by two theories. The first theory considers the spinning dope to possess a liquid-crystalline phase characteristics. The second theory considers the formation of micelles due to the hydrophobic and hydrophilic domains of the spidroin. Respectively, they are inside of the micelles and outside in direct contact of the water solvent [57]. In both cases, shear flow is essential for the formation of aligned fibers. The N-termini and C-termini are crucial to avoid premature folding of the stored spidroin. At a low pH, C-termini have a molten globule state which has no tertiary structure but conserves its secondary structure elements. This suggests that C-termini might be at the origin of the ordered aggregation into fibrils [61]. As for the N-termini, transition from monomer to homodimer are more likely to occur in a low salt concentration environment [62].

Figure 1. Refer to the following caption and surrounding text.

Figure 1. Schematic depiction of the hierarchical structure of spider silk, from hydrogen bonded chains to spider web structures, featuring multiple length-scales in which structural organization is found (from protein secondary structures to macroscopic geometries) [24, 65, 90]. The three images in the top panel were reprinted (adapted) from [24] with permission from Springer.

Standard image High-resolution image
Figure 2. Refer to the following caption and surrounding text.

Figure 2. Schematic of natural spinning process during silk production. A highly unfolded silk protein (dope) undergoes shear and elongation flow in the spinning duct. Associated with a decrease in pH and ion exchange, the produced silk fiber is rich in highly aligned β-sheet crystals. Adapted and reprinted with permission from [64]. Copyright 2016 American Chemical Society.

Standard image High-resolution image

Recently, Lin et al [63] have investigated the crucial spinning processing conditions and design parameters of spidroin self-assembly, through mesoscopic modeling and experiments. The synthetic silks used for the study are composed of three main motifs: polyalanine including hydrophobic domains (at the origin of stiff β-sheet crystals), GGX including hydrophilic domains (at the origin of the semi-amorphous phase) and hydrophilic hexastidine fusion tag to facilitate purification. These micelles merge together into larger micelles after shear flow. Similarly, the β-sheet crystals are larger. The protein is elongated and aligned. Furthermore, homogeneous and continuous network along the fiber axis (or shear flow direction) is only obtained for longer chains. The optimal ratio is obtained for an intermediate ratio of hydrophobic and hydrophilic domains. Shear flow triggers the transition from isotropic to anisotropic cylindrical structures and leads to stronger and more robust fibers [4, 63].

The folding self-assembly mechanism has not been fully understood up until now. Giesa et al [64] have investigated the crucial role of shear flow on the transition from the protein sequence to secondary structures. Molecular dynamics simulations were applied to model the transition of Nephila clavipes MaSp1 protein into secondary structures experimentally determined for silk fibers. The shear stress needed to transform unfolding protein into β-sheets is between 300 and 700 MPa, which also corresponds to transition shear stress for protein self-assembly. The shear stress was estimated at the same time period taken for the assessment of the maximum amount of β-sheets figure 3(A(i)). The transition shear stress is lower than the failure shear stress (indicated by dashed lines; note that the shear stress is derived from the pulling force divided by the effective area). The time period corresponds to the maximum content of β-sheets secondary structures (figure 3(A(ii))). The following drop in shear stress indicates that the molecule enters a low energy state after β-sheets crystals formation. It is observed that the transition into β-sheets occurs at around 20% lower stress than the failure stress. The transition shear stress is independent of the pulling speed and the number of chain repeats [64]. The different stages of protein assembly are illustrated in figure 3(B). Silk dope is first stored in an unfolded state (Stage 1), shear flow then aids to fold the protein and creates aligned β-sheet crystals (Stage 2). The aligned protein was subjected to the presence of a water/ion solvent and vacuum environment, to determine the stability of the post-shear flow protein. Post-spun fibers' natural environment is the air and is considered to be an intermediate between vacuum and water/ion solvent environment. Indeed, as discussed before, ion exchange occurs in the spinning duct and the exited fiber is water-insoluble and stable [8, 58]. In vacuum (Stage 3a), the protein conserves its β-sheets structures independently of chain repeats, and all the proteins are stable. In solvent and in the presence of ions (Stage 3b), a disordered and unfolded spidroin-like state is recovered for short chain protein. The protein is water soluble. The smallest stable molecule is composed of a minimum of six polyalanine repeats, the β-sheets structures remain stable and the β-sheets content is relatively constant (figure 3(B))[64]. Recently, Giesa et al [64] suggested that the stability of β-sheets structures originates from the helices close in space in the spidroin disordered state.

Figure 3. Refer to the following caption and surrounding text.

Figure 3. (A) Shear stress and secondary structure of silk during self-assembly. Reprinted (adapted) with permission from [64]. Copyright 2016 American Chemical Society. (i) Estimation of the transition shear stress, which corresponds to the maximum content in β-sheet structures. (ii) Content in secondary structures during pulling simulation. Transition is located at highest content in β-sheets. (B) Illustrations of different stages of protein self-assembly process. Reprinted (adapted) with permission from [64]. Copyright 2016 American Chemical Society. Stage 1: silk dope is stored in an unfolded state. Stage 2: shear flow folds the protein and creates aligned β-sheet crystals. Stage 3a: in vacuum environment. Stage 3b: in water and in the presence of ions, short chain protein recover their unfolded state. Only the larger structure (a = 5) remains stable. The parameter 'a' is the number of repetitions and represents chain length. For example, a = 5 means the protein contains six polyalanine stretches.

Standard image High-resolution image

Both Giesa et al [64] and Lin et al [63] showed the importance of shear flow and molecular makeup in the spinning process. The results of these studies can contribute to the improvement of microfluidic devices and the design and production of silk-inspired protein materials. Moreover, silk and spinning-inspired protein show attractive advantages: the processing occurs in a mild environment (room temperature and aqueous solvent) and using just a few building blocks. A better understanding could lead to new bio-inspired materials using limited resources and low energy processing.

2.3. Nanomechanics

As shown in the preceding section, the primary structure, associated with a specific spinning process, determines the folding mechanism and the characteristics of the secondary structures. The main secondary structures are: β-sheets nanocrystals controlling the stiffness of the fiber and embedded in a semi-amorphous phase composed of 310-helix, β-spirals and β-turns [6568]. Specifically, β-sheet nanocrystals are composed of tightly stacked and hydrogen-bonded β-sheets, formed by anti-parallel β-strands (these β-sheet crystals represent about 10%–15% of silk volume [15, 69],). Their dimensions are typically on the order of a few nanometers [6971]. Under tension, the nanocrystals are under lateral loading and they form interlocking regions to transfer the load between chains, and β-sheet nanocrystals reinforce the fiber akin to cross-links in a polymer network. The structures ultimately fail at large loads and deformation [69].

The high strength of the nanocrystals is due to the cooperation of the hydrogen bonds between the β-strands and β-sheets. In earlier work [72] we studied the deformation under shear of the hydrogen bond of the β-strands and β-sheets, by applying a structural elastic model. Hydrogen bonds are considerably deformed within a small region of a size of a few bonds, in which the deformation and rupture of hydrogen bonds are delocalized, acting in concert [73]. When the cluster of hydrogen bonds are confined in small amounts so that they correspond to the size of cooperative deformation, the hydrogen bonds are used fully, at maximum utilization of the associated material volume [72]. Henceforth, the force applied on β-strands or β-sheets, bonded by hydrogen bonds, of characteristic length ${L}_{0}$ will remain the same even for larger length. The characteristic length corresponds to the cooperative length, within which the hydrogen bonds deform cooperatively [71, 74]

Equation (1)

where b (0.35 nm [71]) is the distance between two hydrogen bonds, ${K}_{1}$ and ${K}_{2}$ are the stiffness of the β-sheet nanocrystals (41.5 N m−1 [72]) and hydrogen bonds (2.66–8.43 N m−1 [72]) respectively [71, 74]. Although weak and non-covalent, hydrogen bonds contribute to the strength of the crystal, by resisting to shear failure through cooperation. In addition, they can reform after breaking.

To explore further mechanisms of deformation of β-sheets nanocrystals, large-scale molecular dynamic pull-out and bending simulations were reported that were aimed to determine the influence of crystal size under lateral load [69]. It has been established that for small β-sheets nanocrystals, deformations are controlled by shear and for large crystals, by bending. It is important to note that hydrogen bonds are in tension when the crystal is under bending (large size) which does not allow cooperation of hydrogen bonds. However for smaller crystals, hydrogen bonds work cooperatively to resist shear load. Bending to shear transition length is about 2.5 nm. In addition, because small crystals are stiff, under pull-out simulations crystals undergoes a stick-slip mechanism after the initial fracture. The stick-slip mechanism is as follows: the loaded β-sheet is pulled out of the crystal, breaking and reforming its hydrogen bonds (figure 4(B)). The peaks in the small crystal force-displacement plot show hydrogens bonds breaking and reforming and a substantial increase the total dissipated energy (figure 4(A(i))). The dependence of pull-out strength with crystal sizes has shown that small crystals show a strong and stiff response, which implies greater pull-out force and strength (figure 4(A(ii))). The variations of resilience and toughness with size show that smaller crystals dissipate more energy due to the stick-slip failure (figure 4(A(ii))). Critical nanoconfinement size is determined where strength (figure 4(A(ii))), resilience and toughness (figure 4(A(iii))) are at a maximum. The initial stiffness, breaking strength and toughness are maximized for a crystal of length of about 3 nm, in agreement with the previous result (figure 4(A)). The elastic energy storage before initial failure or resilience also increases with the decrease in crystal size. The estimated critical β-strand length (crystal width) is about 1–2 nm. Larger crystals are brittle and fail at a lower load resulting in a crack-like flaw, because of local failure of hydrogen bonds under tension (figure 4(C)). The study on larger crystals is relevant because under thermal excitation or other types of loading, the crystal may bend [24, 65, 69]. Nanoconfinement transforms weak hydrogen bonds into strong, tough and resilient β-sheets crystals.

Figure 4. Refer to the following caption and surrounding text.

Figure 4. Nanoconfinement of β-sheet nanocrystals and results from pull-out simulations carried out using molecular dynamics. Reprinted by permission from Macmillan Publishers Ltd: Nature Materials [69], copyright 2010. (A) Force–displacement, strength and toughness depend on crystal size β-sheet crystals. (i) Force–displacement plot. (ii) Variation of pull-out strength with crystal size. (iii) Variation of toughness and resilience with crystal size. (B) Stick-slip fracture mechanism, observed in small crystals. (C) Brittle fracture of silk nanocrystals, triggered once the crystal size exceeds a critical dimension.

Standard image High-resolution image

At the nanoscale, silk represents a two-phase composite organized in a semi-amorphous phase reinforced with stiff β-sheets nanocrystals. The phases provide extensibility and fracture strength respectively [68]. Stretching simulations on silk nanocomposites of MaSp1 and MaSp2 demonstrated the nonlinear behavior of spider silk. The characteristic force–displacement curve obtained from stretching of the nanocomposite follow four regimes (figure 5). First (i), the composite undergoes an initial stiff behavior [68]. The semi-amorphous phase stretch homogeneously until a yield point. At yield, the hydrogen bonds break resulting in the failure of the 310-helix, β-spirals and β-turns disordered secondary conformations [24, 65, 75]. (ii), The composite enters a softening regime [68]. The semi-amorphous phase unfolds along the pulling direction. The softening mechanism arises from the unraveling of the hidden length, unlocked from the rupture of the hydrogen bonds [24, 65, 75]. The decrease in β-turns content illustrates the unfolding of the secondary structures of the semi-amorphous phase. Simultaneously, the content in β-strand increases, which means that new hydrogen bonds are formed, leading to the creation of small β-sheet crystals in the semi-amorphous region (figure 6(B)) [24, 68]. In particular, the transitions from α-helices to β-sheets, have been observed at the chemical bond scale [16]. Colomban and Dinh [16] carried out a Raman analysis to determine the nanomechanics of silk fibers. As silk protein is a polyamide chain, they applied a controlled strain on the νN–H mode and measured the Raman shift generated. A perfect correlation was observed between the wavenumber shift-strain behavior of the hydrogen bonds at the nanoscale and stress–strain behavior of silk fibers at macroscale [1618]. The softening regime (ii) is followed by (iii) a much stiffer regime [68]. The load is transferred from the fully extended semi-amorphous phase to the β-sheet nanocrystals [75]. Finally (iv), in the short softening regime preceding failure, the nanocomposite fails through a stick-slip mechanism: hydrogen bonds in the crystalline region break and lead to strands sliding, resulting in the separation of smaller intact nanocrystals [68, 75]. This specific mechanism is at the origin of silk's strength, extensibility and toughness.

Figure 5. Refer to the following caption and surrounding text.

Figure 5. Nonlinear constitutive behavior of spider dragline silk. (A) Stress–strain behavior for defect-free spider dragline silk. Reprinted (adapted) with permission from [90]. Copyright 2011 American Chemical Society. (B) Molecular nanostructure deformations and mechanism. Reprinted (adapted) with permission from [68]. Copyright 2010 The Royal Society. Regime (i): linear elastic regime. Regime (ii): unfolding of the semi-amorphous phase. Regime (iii): stiffer regime. Regime (iv): short softening regime, immediately prior to failure. Note that the constitutive behavior shown in panel A is derived from molecular dynamic simulations [75, 90], and that regime (iv) has not been observed experimentally [1618].

Standard image High-resolution image
Figure 6. Refer to the following caption and surrounding text.

Figure 6. (A) Stress–strain response for silk fibrils with different size of β-sheet crystals. Reprinted (adapted) with permission from [75]. Copyright 2010 American Chemical Society. (B) Strain-softening behavior (regime (ii)). Reprinted (adapted) with permission from [68]. Copyright 2010 The Royal Society. (i) Force–displacement plot describing the softening behavior. (ii) Secondary structure content as a function of displacement.

Standard image High-resolution image

The importance of nanoconfinement of β-sheet crystals is identified again when associated with a semi-amorphous phase. In the stress–strain plots of different crystal size nanocomposites, the composite with small crystals (3 nm) shows large strength, toughness and a stick-slip mechanism (figure 6(A)). Only for small β-sheet crystals can the composite fully take advantage of the extensibility of the semi-amorphous phase through unfolding. Thus, nanoconfinement contributes vastly to the extensibility and energy dissipation capacity of silk [75].

In addition to stick-slip (shear) failure and brittle (bending) failure [69], it has been hypothesized that the strain stiffening behavior (stiffer regime (iii) described above) may be caused by β-sheet crystals unfolding [76]. Using molecular dynamics calculations, a recent study [70] proposed a mechanism for silk's strain stiffening behavior following failure of β-sheet crystals. Residual strengths resulting from two types of crystal failure mechanisms were compared: lateral separation of two β-sheets (corresponding to a stick-slip mechanism) and unfolding of the β-sheet. The residual strength was calculated from electrostatic and van der Waals force. As the residual strength of the unfolding of the β-sheet is significantly higher for any deformation state, the study proposed the following deformation mechanism: first, the β-sheet crystals fail by stick-slip giving rise to smaller crystals or just the separation of one β-sheet. If the failure strength of the smaller crystals is greater than that of the original crystal, they participate in the stiffening mechanism. Second, the β-sheets unfold. When completely unfolded and stable, they participate in a secondary stiffening regime. The straightened β-sheets perform as a fiber reinforcing the nanocomposite (figure 7) [70].

Figure 7. Refer to the following caption and surrounding text.

Figure 7. Schematic of a hypothetic failure mechanism of β-sheet crystals, within the context of the nanocomposite. They fail first through stick-slip, then unfold. Reprinted and adapted with permission from [70].

Standard image High-resolution image

In conclusion, at the nanoscale, silk is a two-phase nanocomposite that consists of the nanoconfined β-sheet and semi-amorphous phase, which combined, provide strength, extensibility and toughness. Cooperative deformation of hydrogen bonds transform weakness into strength within the crystalline regions.

2.4. Macroscale silk architectures and mechanics

Silk's upper-scale hierarchical level is reflected in macroscopic silk architectures, such as spider webs (in 2D or 3D), or silkworm cocoons. Silk macroscale structures were likely adapted through evolution to ensure characteristic functions. For instance, viscid spider silk is used to trap prey whereas dragline silk is used for framework of the web. Out of the numerous types of spider webs (cob webs, brushed sheet webs...), orb web is the most familiar and studied [77, 78]. As web construction is energy costly, spiders minimize material use while exploiting fully the unique properties of the different silks and securing durability of the web [79, 80]. Orb web architectures are organized in radial dragline silk (which properties are discussed previously) and spiral viscid or sticky silk [10, 15].

Recently, Qin et al [79] have investigated the mechanical performance of a 3D-printed synthetic elastomeric spider web to determine the influence of structural design and material distribution in web mechanics and design. Spider web experiments are difficult to repeat and compare, because webs are unique and testing requires destruction of webs. Consequently, testing on a 3D-printed web-inspired structure and simulations on its model were carried out. Under point loading, e.g. prey impact or spider weight, it is preferable to distribute uniformly silk in all the fibers in order to optimize strength. Radial and spiral threads as well should have the same diameter/thickness. On the other hand, for a distributed loading, e.g. wind loading, the web is the strongest when radial threads are thicker permitting sacrificial failure of thinner spiral threads. Defects are easily repairable. Furthermore, these different geometries are observed in nature as spiders optimize their use of silk through web architecture. For example, centimeter-scale spider webs (garden spider webs) are optimized for local load: the spiral and radial threads have similar diameter. Meter-scale spider webs (Caerostris darwini webs) are optimized for wind load as they cover a larger surface: radial threads are thicker [4, 79].

Although silk's properties are usually obtained through tensile testing of the fibers, Koski et al [81] determined the stiffness tensors of the fibers of an orb web using non-invasive, non-destructive Brillouin light scattering method. The stiffness of the dragline radial, viscid spiral silk threads and the silk junctions were mapped on the orb web. When hydrated, spider silk is subjected to supercontraction; it shortens its length by 50% and increases by more than 40% the stiffness of the fibers. Even though water droplets apply a gravity load, they stiffen the web considerably and the deflection of the web can decrease by 17% [77, 81].

While a spider web's robustness rests in part on its discrete architecture, a silkworm cocoon's remarkable properties are hypothesized to rely on its multi-layered nonwoven structure. Cocoons are spun from a single and continuous fiber that can reach 1 km in length. Layers and intersecting intra-layer fibers are bonded by sericin glue, which can be considered as the matrix of the composite. A cocoon is a three-dimensional laminated composite composed of a porous matrix reinforced by randomly arranged and oriented fibers. Cocoon performance highly depends on its porosity and bonding connectivity [8286]. Understanding cocoon geometry and mechanics can result in new bio-inspired designs for materials, for instance, in random fiber and particulate composites [82].

Further understanding of the interplay between silk properties and macroscale structural optimization of webs could provide structural solutions for engineering, especially for resilient and lightweight design. In particular, robustness of webs and their underlying mechanics will be discussed in the following section.

3. Resilience of spider silk

3.1. Case study 1: resilience of fibers against defects

Silk fibers not only have excellent mechanical strength and toughness but also are extremely robust against defects. Indeed, they perform remarkably well even though it has been observed that silk fibers often have defects such as cavities cracks, surfaces, and tears [8789]. Defects can be very large (several hundreds of nanometers) for a micrometer size fiber and they create stress concentrations which could trigger fiber failure [90].

Spider silk fibers are composed of a bundle of fibrils defined by a three-dimensional two-phase composite. To understand how interplay between β-sheet crystals and semi-amorphous phase at the nanoscale influences and helps to resist fracturing at the macroscale, Giesa et al [90] applied a mesoscale spring-model to connect length-scales (figure 8(C)). The springs follow the same four-regime nonlinear behavior of the two-phase composite derived from nanomechanics (figure 5(A)). Various loading cases were investigated to describe the mechanical response of fibrils under different loadings (tension and shear) and defects. The crack size is 50% of the length (L) or width (H) of the fibril sample figure 8(A)). Fracture strengths were measured for the different cases and different fibril widths H, and were compared to fracture strengths of defect-free silk fibers obtained experimentally (figure 8(D)). Fracture stresses and strains increase considerably as the width H decreases. When fibril size H converges toward the critical length-scale H* = 50 ± 30 nm, failure stresses and strains also converge toward the experimental data of defect-free fibers. The large crack does not impact the mechanical response of fibrils, if H = H*. When H reaches H* = 50 ± 30 nm, there is a drastic increase of material entering regime (iv) in which the β-sheet crystals fail through stick-slip. After this threshold, 100% of the material contribute to resisting failure through the stick-slip mechanism, as seen in figure 8(B)). As a result, when fibrils are nanoconfined at H = H*, they behave as defect-free fibers by distributing stresses homogeneously. The entire fibril works at resisting fracture. Through nanoconfinement of fibrils, macroscale properties can be traced back to molecular scale properties (stick-slip and unfolding mechanisms).

Figure 8. Refer to the following caption and surrounding text.

Figure 8. Determination of a fibril's critical size. Reprinted (adapted) with permission from [90]. Copyright 2011 American Chemical Society. (A) Flawed fibrils under tension (mode I, (1) and (2)) and shear (mode II, (3) and (4)) loading. (B) Percentage of fibril that has reached stiffening regime (regime (iii)) (blue) and stick-slip failure (regime (iv)) (red). When H is decreased until it is equal to H* = 50 + −30 nm, there is a drastic increase of material entering regime (iv). After this threshold, 100% of the material contribute to resisting failure through the stick-slip mechanism. H* is indeed the critical size. (C) Setup of a simple spring model to assess the length-scale effects studied here. (D) Influence of fibril size H on failure strain and failure strength, for different loading cases and defects. Nanoconfined fibrils at H = H* and below reach high strength and extensibility, as seen also in experimental data and defect-free simulation results.

Standard image High-resolution image

3.2. Case study 2: resilience of spider webs

In nature, spider webs are rarely intact as they are subjected to numerous threats, such as wind load, impact from debris and prey, resulting in failure of threads. Despite having defects, they remain functional for hosting their spiders and capturing prey [91].

Indeed, simulations on orb webs with different defects (removal of threads) not adjacent to the loaded thread have shown that defects do not impact web behavior and failure mechanisms [10]. For all webs, failure is only localized near the loaded thread. When loading is applied on radial threads, web deflection and damage are larger than in the spiral thread loading case. Moreover, webs fail at a considerably larger load when load is applied to radial threads. This suggests that spiral threads are non-structural; made out of viscid/sticky silk, their role is to catch prey. The localized damage of webs under local loading can be traced back to their molecular structure and nanomechanics in particular, the underlying nonlinear behavior. When a local load is applied to a radial thread (and prior to failure), all the radial threads reach the onset of yield stress. At this point, only the loaded thread enters the unfolding of the semi-amorphous phase regime (regime (ii)) at a low stress. As shows the bar plot figure 9(A(i)) under local loading on a radial thread, nonlinear behavior leads to stress concentration in the loaded thread only, and the other threads stay at the yield stress. Then the silk enters the third regime: load is transferred to β-sheet crystals and the silk stiffens. As soon as the failure stress is reached, β-sheet crystals, located in the loaded region of the fiber, fail. Only the loaded thread fails (figure 9(A(i))) [10, 92]. To show that a spider web's robustness results from its nonlinear stiffening behavior, Cranford et al [10] also compared the previous case to the response of webs with the same geometry but composed of silks with different stress–strain properties: linear elastic and elastic-perfectly plastic (softening behavior). For both new cases, the damage area increased compared with the nonlinear behavior case, especially for the web made of softening behavior silk. This is due to the transfer of load to the adjacent radial fibers (linear elastic case) (figure 9(A(ii))) or even to the entire web (softening case) (figure 9(A(iii))). Although they fail at a higher loading force thanks to the cooperation of all the threads in resisting load and deflect less, they lose in robustness when compared to a web made of nonlinear stiffening silk (figure 9(A(ii)(iii))) [10, 92].

Figure 9. Refer to the following caption and surrounding text.

Figure 9. Influence of nonlinear material behavior of silk and loading type on web performance. Reprinted by permission from Macmillan Publishers Ltd: Nature [10], copyright 2012. (A) Web performance and failure under local loading for different material behavior: (i) atomistically derived dragline silk (nonlinear stiffening behavior), (ii) linear elastic and (iii) elastic-perfectly plastic (softening material behavior). The respective stress–strain curves are represented in the left panel. Comparison of failures (center panel) shows that webs made of dragline silk perform better by sacrificial failure of one thread. The bar plots (right panel) describe the stress transfer to the radial threads depending on material behavior. (B) Stress–strain plots describing different types of material behavior with parameter α. (C) Structural robustness versus α. (D) Web performance under global (wind) loading.

Standard image High-resolution image

Under wind loading the three types of webs fail at 60 m s−1 wind speed. For the natural silk web, the deflection is controlled by dragline radial silk threads, because they are stiffer than viscid spiral threads. Webs deflect similarly for wind speed lower than 10 m s−1. However, for higher wind speed, the nonlinear stiffening behavior is a disadvantage for web performance. Their webs have a considerably larger deflection because of the softening regime (unfolding) of dragline silk, making the radial threads deflect more (figure 9(D)) [10, 92]. Larger deflection means more risk for the web to be caught in vegetation or walls to which they are fixed.

In the same way that nonlinear stiffening increases the flaw-tolerance of fibers [93], nonlinear stiffening of silk also increases the robustness of spider webs. Similarly to the previous case study, a power-law with a controlling parameter N (N > 1 for nonlinear stiffening) is used, so that $\sigma \propto {\varepsilon }^{N}$ ($\sigma $ and $\varepsilon $ are the stress and strain, respectively). As α increases, there is a transition from softening to stiffening material (figure 9(B)). For spider webs, quantized fracture mechanics can be applied instead to describe their failure behavior. The scaling law is derived [10, 92], as

Equation (2)

with ${\rm{\Omega }}$ the structural robustness defined by fraction of intact web after failure, S is a system-dependent constant and is the fraction of damaged web associated with linear elastic behavior. Here, $\alpha =\tfrac{1}{(\mathrm{1+1}/N)}$ and $\alpha $ is a parameter describing the behavior of silk: linear elastic ($\alpha =\mathrm{1/2}),$ nonlinear stiffening ($\alpha \gt \mathrm{1/2}),$ or nonlinear softening ($\alpha \lt \mathrm{1/2}).$ According to the scaling law, structural robustness increases with $\alpha ,$ which means that a nonlinear stiffening behavior increases spider web robustness (figure 9(C)) [10, 92].

The macroscale spider web robustness originates from failure of sacrificial elements (threads) due to highly localized failure, which is possible because of nonlinear stiffening. Similarly, silk threads are also extremely resilient because of this particular behavior. As for nonlinear stiffening, this characteristic is due to the interplay between stiff nanocrystals and semi-amorphous phase, which comes from the very protein sequence. Silk's resilience originates from its hierarchical organization.

4. Scaling laws: a model to define hierarchical length-scales

The resilience of silk can generally be explained throughout the different levels of its hierarchical organization, starting from protein sequence, nanoconfined β-sheet crystals, two-phase composite, its underlying nonlinear behavior, and nanoconfined fibrils up until the macroscale structural design of webs. The same applies to fracture toughness of silk fibers, and other biological fibers: it heavily relies on the length-scale of interest [94]. Recently, Giesa et al [94] established an analytical model using scaling laws to define the important role of length-scale. They derived the macroscale fracture characteristics of hierarchical fibers, in particular, spider silk, from the interatomic potential.

In fracture mechanics, the process zone or cavitation box bounds the zone around the crack in which the material resists crack propagation and decreases the risk of fracture[94, 95]. The fracture toughness K depends on the characteristic length-scale ${l}_{0}$ of the process zone ($K\sim \sqrt{{l}_{0}})$ [96]. To avoid catastrophic failure and become flaw-tolerant, it is preferable to have a large ${l}_{0}$ (up to the fiber diameter): the stresses are less concentrated at the crack tip and the damage does not spread only through crack surface formation. The process zone size can be increased by other toughening mechanisms through the nanostructure of the silk fiber [97]. For instance, dissipation can occur through unfolding of the semi-amorphous phase or through the stick-slip mechanism of the β-sheet crystals [94].

Instead of using continuum mechanics (e.g. Griffith's theory of fracture mechanics [98, 99],), Giesa et al [94] derived the length-scale ${l}_{0}$ directly from the interatomic potential. They consider that atomic bonds failure implies fiber fracture. From the generalized Lennard-Jones (n, m) potential and the Dugdale–Barenblatt yield-strip model [100], the ${l}_{0}/{r}_{0}$ ratio is estimated to be between 0.5 and 16 for different sets of (n, m) parameters. The parameter ${r}_{0}$ denotes the equilibrium lattice spacing of a cubic lattice. The process zone size is approximated to be of the order of a few nanometers [94]. However, the fracture toughness K derived from the approximations is about 1000 times smaller than measured fracture toughness. The reason for this mismatch is spider silk's heterogeneity through hierarchy. Thus, r0 is the characteristic length-scale and depends on the hierarchical level of interest [94].

For spider silk, the length-scale of observation ${r}_{0}$ is atomic bonds (e.g. hydrogen bonds) and the process zone size ${l}_{0}$ is of a few nanometers, as previously discussed. Thus, to get a large process zone size, the dimensions of the substructure has to be of the same order of magnitude. At this scale, the nanostructure is nanoconfined β-sheet crystals (2–4 nm [69]). Larger in scale, silk is a two-phase composite described by a network of β-sheet crystals linked by a semi-amorphous phase. In this case, the length-scale of observation is ${r}_{0}\simeq 10\;{\rm{nm}}$ (distance between β-sheet crystals). The process zone size is calculated to be 20–150 nm and agrees with the ${l}_{0}/{r}_{0}$ ratio. The substructure to be considered is nanoconfined fibrils (40–80 nm in diameter [101]); fracture toughness increases with unfolding of the semi-amorphous phase. Silk's final level of hierarchy is silk fiber which is composed of an ensemble of weakly bonded flaw-tolerant fibrils. The fracture toughness increases by sliding and delocalization of fibrils [94]. The toughness mechanism across scales is described above are illustrated in figure 10(A).

Figure 10. Refer to the following caption and surrounding text.

Figure 10. Fracture toughness and confinement of length-scales. Copyright 2013 Wiley. Used with permission from [94]. (A) Toughness mechanism across scales in spider silk. (B) Model for transferring stress concentrations to the upper scale and making extremely resilient materials.

Standard image High-resolution image

In conclusion, the resilience of silk fibers is due to transfer of stress concentration from one scale to another, all the way to the upper scale, through particular toughness mechanisms, as described in figure 10(B). This model shows that in hierarchical materials, each substructure is composed of building blocks, of size r, repeated n times, making a new motif. The motif needs to be confined to r* which is of the same order as the process zone size of the substructure which follows this scaling law: ${l}_{0}/{r}_{0}=[0.5,\;16]$ (figure 10(B)). Each length-scale nanostructure (β-sheet crystals, fibril, and fiber) is confined to a large process zone (${l}_{0}\simeq {\rm{nanostructure}}$ size). Hierarchy and nano-substructures explain the large process zone of the overall fiber (300–1000 nm) and their remarkable fracture toughness (figure 10(B)) [94].

5. Summary and future directions

In this review article we discussed how hierarchical organization makes silk, in particular spider dragline silk, extremely strong, tough, robust, and versatile. At the lowest level of the hierarchy, silk is a sequence of ordered amino acids. Combined with a specific spinning process, the protein folds into a two-phase composite organized into highly aligned β-sheet nanocrystals and unordered semi-amorphous phase. The strength of β-sheet crystals arises from its nanoconfined size, leading to cooperation of weak hydrogen bonds against shear. Interplay between crystals and semi-amorphous phase results in a nonlinear stiffening behavior of silk. Strength is due to β-sheet nanocrystals and extensibility due to the hidden length in the semi-amorphous phase. Up a scale, silk nanoconfined fibrils are bundled into fibers which are produced to make macroscale spider webs. In addition, silk is extremely resilient and robust against defects. Fibers' robustness comes from its nanoconfined fibrils and nonlinear stiffening behavior, which make them behave as defect-free fibers by homogeneous distribution of stress. To face wind load and prey impacts or just their own weight, spiders have designed and optimized extremely resilient and functional webs. Their robustness comes from failure of sacrificial threads, which is also due to the nonlinear behavior of silk, preventing catastrophic failure of webs. Webs remain functional and can be repaired. Finally, using scaling laws, we discussed how fibers' nanostructure is essential for transferring stress from one scale to a higher scale which increase considerably fracture toughness of the overall fiber.

Understanding silk's hierarchical structure and its underlying mechanisms provides essential insights for silk-inspired material design and production. Silk has become a template for upscaling molecular properties to the macroscale. For example, it is theoretically possible to modify a protein sequence to tune synthetic silk properties. If motifs responsible for β-sheet crystals are added, the fiber is likely to be stiffer but less extensible [6]. Indeed, simulations and experiments have shown that increasing the length of hydrophobic motifs (responsible for β-sheet crystal formation) and increasing the overall length of the protein chain resulted in stronger fibers [63, 66, 102, 103]. Environment (e.g., ions and pH) also influences properties of recombinant silk [58]. Silk protein is primarily obtained through reconditioning of spun silk or bio-synthesis of recombinant silk protein [104]. Contrarily to silkworm silk, which is easily harvested from cocoons, mass production of spider silk remains a challenge. Recombinant silk production seems to be an interesting alternative. One successful attempt at producing recombinant spider silk was realized by using metabolically engineered E. coli to synthesize native-sized recombinant silk protein, which resulted in fibers with comparable mechanical properties [103]. Another successful alternative is to use transgenic silkworms to produce spider/silkworm composite silk in cocoons [105, 106]. In addition to spinning fibers, other products can be processed from recombinant spider silk such as: particles, films, hydrogels, or foams [107, 108]. Silk-based particles could be used for drug delivery to carry substances as they are biodegradable and biocompatible. Silk films could be used as cell-supporting scaffolds. Silk hydrogels could be applied to functional coating, tissue engineering, and drug delivery as well. Silk sponges and foams are able to transport nutrients and waste, which makes them suitable for cell culture. Recently, carbon-reinforced silk reinforced was spun by spiders, after being exposed to carbon nanotube or graphene dispersions, and showed improved mechanical properties [109].

In conclusion, silk's remarkable properties are due to its hierarchical structure. Knowing how scales are linked and how they influence the uppermost scale will broaden our understanding of silk's properties and mechanisms at each level, and likely do the same for other hierarchical materials. It will consequently improve our knowledge of how to tune and manipulate material properties for various applications, ranging from biomedical to structural optimization. Silk is a blueprint for material design and processing.

Acknowledgments

We acknowledge support from NIH (U01 EB01497), ONR-PECASE (N00014-10-1-0562), and AFOSR (FA9550-11-1-0199).

Please wait… references are loading.