mantis shrimp material

Upload: yap-yih-shan

Post on 12-Oct-2015

32 views

Category:

Documents


0 download

DESCRIPTION

A profound research done by UC.

TRANSCRIPT

  • q. Mailside

    e, CA, USliforn

    Keywords:BiomimeticCompositesBiomineralizationModeling

    (specically, Odontodactylus scyllarus). This crustaceans club has been designed to withstand the thou-

    ers over the past several decades have looked to the nacreous layerof mollusk shells as a model system for toughened biological com-posites [5,6]. Numerous efforts have been undertaken to mimic the

    s, Odontodeaching ac

    tions up to 104 m s2 and speeds of 23 m s1 [11,14]. Throuhammermechanism, shells of prey (e.g. mollusks), which cocharacteristically tough biological composites, are destroyedactyl club itself, however, is optimized to resist damage from adynamic striking event, and can withstand thousands of repeatedimpacts without experiencing catastrophic failure [1012]. Toaccomplish this, the club must be hard and stiff to deliver momen-tum to its prey, but must also be strong and tough to resist damagefrom an equal and opposite impact force. The composition and

    q Part of the Biomineralization Special Issue, organized by Professor HermannEhrlich. Corresponding author at: Department of Chemical and Environmental Engi-

    neering, Bourns Hall B357, University of California Riverside, Riverside, CA 92521,USA. Tel.: +1 (951) 827 2260; fax: +1 (951) 827 5696.

    E-mail address: [email protected] (D. Kisailus).

    Acta Biomaterialia 10 (2014) 39974008

    Contents lists availab

    Acta Biom

    journal homepage: www.elsevlogical materials, natural composites have received considerableattention in the eld of biomimetics [4]. As one example, research-

    [11,14]. The dactyl club of one smashing speciescyllarus (Fig. 1A), delivers high-velocity blows rhttp://dx.doi.org/10.1016/j.actbio.2014.03.0221742-7061/ 2014 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.econdsactyluscelera-gh thisnsist ofd. TheNatural materials are known for their efciency, performing arange of functions including structural support as well as mobilityand protection, using only the minimum quantities of a limitedselection of constituent materials [1]. These biological materialsare most often mineralized composites, consisting of an organicphase which templates the growth of mineral [2,3]. Complex hier-archical structures containing only a small amount of organic (typ-ically 510 wt.%) result in increased toughness and materialproperties exceeding those of the inorganic phase alone [1,2].Because of the performance and benign processing routes of bio-

    however, that the nacreous structure is vulnerable to attack froma smashing predator known as the stomatopod or mantis shrimp[10]. This aggressive marine crustacean uses a hammer-like striketo destroy the shells of mollusks to expose the soft body of theanimal.

    Stomatopods are divided into two classes, colloquially referredto as spearers and smashers, dependent upon the dactyl portion oftheir raptorial feeding appendage (slender barbed spears or bul-bous clubs, respectively) [1013]. Here, we focus on the smashingvariety, which employ a heavily mineralized club to impact theshells of prey animals in a dynamic strike lasting only millis1. Introductionsands of high-velocity blows that it delivers to its prey. The endocuticle of this multiregional structureis characterized by a helicoidal arrangement of mineralized ber layers, an architecture which resultsin impact resistance and energy absorbance. Here, we apply the helicoidal design strategy observed inthe stomatopod club to the fabrication of high-performance carbon berepoxy composites. Throughexperimental and computational methods, a helicoidal architecture is shown to reduce through-thickness damage propagation in a composite panel during an impact event and result in an increasein toughness. These ndings have implications in the design of composite parts for aerospace, automotiveand armor applications.

    2014 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

    brick-and-mortar structure of nacre [79]. It was recently shown,Article history:Available online 27 March 2014

    Through evolutionary processes, biological composites have been optimized to full specic functions.This optimization is exemplied in the mineralized dactyl club of the smashing predator stomatopodBio-inspired impact-resistant composites

    L.K. Grunenfelder a, N. Suksangpanya b, C. Salinas a, GK. Evans-Lutterodt d, S.R. Nutt e, P. Zavattieri b, D. KisaDepartment of Chemical and Environmental Engineering, University of California Riverb School of Civil Engineering, Purdue University, West Lafayette, IN 47907, USAcMaterials Science and Engineering Program, University of California Riverside, RiversiddNational Synchrotron Light Source, Brookhaven National Laboratory, Upton, NY 11973eDepartment of Chemical Engineering and Materials Science, University of Southern Ca

    a r t i c l e i n f o a b s t r a c tilliron a, N. Yaraghi c, S. Herrera a,us a,c,, Riverside, CA 92521, USA

    92521, USAAia, Los Angeles, CA 90089, USA

    le at ScienceDirect

    aterialia

    ier .com/locate /actabiomat

  • lubhe earc

    orig

    omaFig. 1. (A) Photograph of Odontodactylus scyllarus, along with an image of a dactyl cOptical micrograph of a transverse section of the stomatopod dactyl club revealing timage of a polished section through a single period, depicting a characteristic nestedhelicoidal organization of bers. (E) Schematic of Bouligand geometry, clarifying theber layers.

    3998 L.K. Grunenfelder et al. / Acta Biarchitecture of the club has been previously investigated, revealinga complex structure [10].

    The stomatopod dactyl club, as reported by Weaver et al., is amultiregional biological composite [10]. The club, like most crusta-cean exoskeletal elements, is formed through the templating ofmineral growth by an organic scaffold [15]. In the stomatopod, thisscaffold consists of bers of the long-chain polysaccharide a-chitin.The a-chitin template is mineralized in the exocuticle by orientedcrystalline hydroxyapatite [10]. The exocuticle region, which is themost heavily mineralized and hardest portion of the club, is termedthe impact region, as this is where the club impacts the prey. Theimpact region, outlined in blue in Fig. 1B, wraps around the entireclub, and is thickest at the impact surface. Underlying this hardouter layer is the endocuticle, which is mineralized with amor-phous calcium carbonate and calcium phosphate [10]. There aretwo regions of interest within the endocuticle, termed the striatedand periodic regions. The striated region of the club, located oneither side of the endocuticle (yellow in Fig. 1B), is characterizedby closely spaced striations. This region prevents lateral expansionof the club during a strike [10]. This is achieved through alignedmineralized ber layers that form a circumferential band aroundthe club. The remainder of the club is referred to as the periodicregion. While each region of the club plays a role in the dynamicresponse of the structure, the periodic region is key to energy dis-sipation and impact resistance [10].

    When viewed in transverse cross-section (Fig. 1B), the periodicregion exhibits a regular pattern of laminations, a theme com-monly observed in arthropod exoskeletons. It is this periodicitythat gives the region its name. Close observation by scanning elec-tron microscopy uncovers a characteristic nested arc pattern ineach laminated region (Fig. 1C). Bouligand has described this archi-tecture in detail, revealing that the nested arc appearance is anoptical illusion of sorts, resulting from a helicoidal stacking ofaligned ber layers [16,17]. This geometry is claried in a fracturedthat has been removed and sectioned (as indicated by the dashed line marks). (B)xocuticle (blue), striated (yellow) and periodic region (remainder of club). (C) SEMpattern. (D) SEM image of a fracture surface through multiple periods, revealing a

    ins of the nested arc pattern, a consequence of the helicoidal arrangement of aligned

    terialia 10 (2014) 39974008sample (Fig. 1D). The relationship between a rotated stacking ofber layers and the nested arc pattern observed in polished sec-tions is presented in Fig. 1E.

    The Bouligand structure in the crustacean cuticle is a complexhierarchical composite [15] consisting of acetyl glucosaminemonomers that polymerize to form long chains of the polysaccha-ride a-chitin, which assemble into chitin nanobrils that aggregateto form chitinprotein nanobers and organize into planar sheets[3,15,18,19]. The sheets then stack, with each layer rotated by asmall angle from the layer below, eventually completing a rotationof 180 [15,16,18,19]. The helicoidally arranged a-chitin bers aremineralized at the nanoscale with crystalline and/or amorphousmineral [19]. Thus, the cuticle is often viewed as a ber-reinforcedstructure, with the a-chitin bers embedded in a matrix of mineral[20,21]. The structurefunction relationships arising from thishierarchical design have been examined over a range of lengthscales through experiments and simulations on a variety of crusta-cean species [19,2126]. With a small rotation angle between con-secutive layers, a helicoidal composite displays an isotropicresponse to in-plane loading at the macroscale [22,27]. Addition-ally, the rotated architecture has been shown to increase thetoughness of biological composites when loaded in tension, withthe lamellae reorienting to adapt to the loading conditions andthus resist deformation [28]. The composite structure of the cuticleprovides the stiffness, strength and hardness required to protectthe animal and allow for movement and predation, with the spe-cic properties of each cuticle region dependent on the degree ofmineralization as well as the stacking the density of the Bouligandlayers [19,2225].

    Owing to its design, a natural bio-inspired corollary to theBouligand structure can be achieved using ber-reinforced com-posites. Unidirectional composite materials consist of aligned berlayers, similar to those observed in the crustacean exoskeleton.Through the stacking of unidirectional layers (plies), a helicoidal

  • showing the geometry to have enhanced properties when com-

    study, the impact damage to each sample was investigated through

    iomaexperimental observations as well as nite-element analysis. Fol-lowing impact testing, the residual strength of the samples wasdetermined through compression testing. This compression afterimpact protocol provides information on composite toughness, asit interrogates the ability of the sample to carry load followingthe onset of damage. Experimental and model results for ber-reinforced composites provide insight into toughening mecha-nisms at work in the stomatopod dactyl club, and reveal designguidelines for impact-resistant materials.

    2. Experimental

    2.1. Stomatopod specimen handling and sample preparation

    Live specimens of O. scyllarus were obtained from a commercialsupplier and housed in a recirculating articial seawater system.Specimens ranged in size from approximately 90 to 150 mm inlength. For analysis, dactyl clubs were removed, rinsed in deionizedwater, and air-dried at room temperature. To prepare polished sec-tions, clubs were embedded in epoxy (System 2000, Fiberglast,USA), sectioned with a diamond blade and polished with gradeddiamond abrasive solutions to a 50 nm grit size. Fractured sampleswere obtained along the transverse plane of the club using a razorblade. All samples were sputter coatedwith a thin layer of platinumand palladium to increase conductivity and eliminate charging andsubsequently imaged using a scanning electron microscope(XL30-FEG, Philips, USA) with 10 kV accelerating voltage.

    2.2. Composite fabrication

    The composite material used in this work was a carbon berepoxy prepreg with a unidirectional reinforcement (HexTow

    IM7, Hexcel, USA) specically formulated for out-of-autoclave, vac-uum bag only cure (CYCOM5320-1, Cytec Industries, USA). Fivesets of composite panels, each measuring 350 mm long 180 mmpared to a quasi-isotropic control sample [27]. Chen et al., drawinginspiration from a beetle exoskeleton, have reported an increase inthe fracture toughness of a glass ber-reinforced helicoidal com-posite architecture compared to that of a unidirectional control[30]. These studies have shown the benets of a helicoidal archi-tecture under static loading conditions. Based on the ability ofthe stomatopods club to withstand multiple high-energy impacts[10], the goal of this work was to examine the impact resistanceand energy absorption of a helicoidal structure subjected to adynamic strike, as well as to quantify the residual strength of thestructure following impact damage.

    Past studies examining the effect of impact on helicoidal com-posites have compared performance of helicoidal samples with asingle rotation angle to that of unidirectional or cross-ply controls[31,32]. Here, we examine three helicoidal rotation angles, andcompare results to unidirectional as well as quasi-isotropic con-trols. A quasi-isotropic layup was chosen as it is an aerospaceindustry standard design, and a robust baseline architecture. In thiscomposite can be readily obtained. Previous work has highlightedthe benets of a bio-inspired helicoidal architecture subjected to arange of loading conditions. Changes in ply orientation are knownto inuence the toughness and strength of composite materials byaltering damage mechanisms and propagation [29]. Cheng et al.published a comprehensive work on the exural stiffness andshear strength of glass ber-reinforced helicoidal composites,

    L.K. Grunenfelder et al. / Acta Bwide, were fabricated with 48 prepreg layers that were laid upwith different ply orientations. Two sets of standard controlcomposite panels were fabricated. The rst were unidirectionalsamples, with all plies oriented in the 0 direction. The secondwere quasi-isotropic panels consisting of prepreg layers orientedin the 0, 45 and 90 directions, with the layup symmetric aboutthe mid-plane. Mid-plane symmetry is critical in composite manu-facturing, to prevent the warping that can occur in non-symmetriclayups because of residual stresses that develop during elevatedtemperature cure [27,29]. In addition to the unidirectional andquasi-isotropic controls, three helicoidal (Fig. 2A) structures werefabricated with rotation angles of 7.8, 16.3 and 25.7. These rota-tion angles were chosen such that all panels had a consistent thick-ness, and mid-plane symmetry could be maintained. Details of thecomposite layup schemes are presented in Table 1.

    For all helicoidal samples, prepreg plies were cut to specicangles prior to layup. All samples were fabricated to be symmetricabout the mid-plane, meaning each right-handed rotation wasimmediately followed by a left-handed rotation (ber angle in eachsample is represented by the greyscale images in Fig. 2B).

    After layup, samples were vacuum bagged (>28 in. Hg) accordingto the layup schematic in Fig. 2C. Bagged samples were debulkedfor 4 h at room temperature to remove trapped air. After the roomtemperaturevacuumhold, sampleswerecuredaccording to theman-ufacturers recommended cure cycle (Fig. 2D). All samples were sub-jected to a freestanding post-cure cycle consisting of a ramp at2 C min1 to 121 C, immediately followedbya rampat0.6 C min1

    to 177 C. This lower ramp rate was employed to avoid crossing theinstantaneous glass transition temperature of the resin anddevitrify-ing the resin. The samples were held at 177 C for 2 h. Samples werenally cooled back to room temperature at 2 C min1.

    2.3. Analysis and mechanical testing of composites

    After curing, composite samples measuring 6.5 mm in thicknesswere machined to dimensions of 100 mm 150 mm using awater-fed tile saw. The dimensions of each sample were measuredwith calipers and recorded. Polished cross-sections of each samplewere prepared by embedding in epoxy (SamplKwick, Buehler, USA)and polishing with graded silicon carbide sandpapers down to agrit of 1200 (15 lm). Cross-sectional microstructures were exam-ined using a digital stereo microscope (Keyence VHX-600E).

    To determine the toughness of each composite, samples weresubjected to a standard compression after impact testing protocol.Three replicates were tested for each layup conguration. Impacttesting was performed in accordance with ASTM D7136 [33], usinga drop weight impact testing system equipped with a hemispher-ical tip, with a diameter of 16 mm (9250HV, Instron, USA), provid-ing an impact energy of 100 J. During impact, samples were held inplace by upper and lower clamp plates, with 76 mm diameter cir-cular holes at the center. After impact, photographs of each samplewere obtained to examine external damage, and dent depth wasmeasured using a depth gauge. Internal damage was probed usinga non-destructive ultrasound technique (Ultrasonic C-scans) usinga 10 MHz water-coupled transducer in reected mode (UPK-T36,NDT Automation, USA). Following ultrasound analysis, sampleswere end-loaded into a specialized xture and compressed to fail-ure, following ASTM D7173 [34]. The xture prevents global buck-ling of the sample during testing, leading to failure as a result ofpropagation of cracks and delaminations [35]. A displacement-con-trolled compressive loading of 1.25 mmmin1 was applied using aload frame with a 250 kN capacity (5585H, Instron, USA). Load anddisplacement were recorded during testing and used to calculateresidual strength.

    2.4. Modeling of impact damage

    terialia 10 (2014) 39974008 3999Experimental data from drop tower testing was used to validatemodels, allowing for the prediction of failure mechanisms and

  • oma4000 L.K. Grunenfelder et al. / Acta Biexamination of damage behavior in each specimen type underhigh-impact loading. Computational modeling of a drop tower testfollowing ASTM D7136 on IM7/5320-1 composite specimens wasperformed using the nite-element method with an explicit solver[36]. Most of the mechanical properties of the unidirectionalprepreg used in this work were obtained directly from manufac-turer data sheets (personal communication), except for the criticalenergy release rate values which were adopted from other works[37,38]. These properties are listed in Table 2.

    The quasi-isotropic control samples and three sets of helicoidalcomposite specimens examined in this study were modeled at theply level. Each unidirectional layer within the composite sampleswas modeled individually as continuum plane stress shellelements with the thickness being equal to the average laminathickness (0.135 mm) and the average element size near theimpact region being 0.08 mm. Fiber orientation was assigned toeach ply based on the layup details of each sample (see Table 1).

    Fig. 2. Details of composite processing. (A) Schematic representation of a helicoidal rotatrepresenting ber rotation angles in helicoidal composites, demonstrating the mid-planrotation. (C) Details of the composite layup and vacuum bagging scheme. (D) Composit

    Table 1Layup details for composite laminates.

    Sample Layup

    Unidirectional [0]48Quasi-isotropic [0/45/90]6sSmall angle [0/7.8/. . ./180]sMedium angle [0/16.3/. . ./180]2sLarge angle [0/25.7/. . ./180]3sterialia 10 (2014) 39974008The total model consisted of one layer with 144 integration points.In this layer, each ply of the composite is represented by threeadjacent points, thereby accounting for all 48 plies (Fig. 3). In total,the model consisted of 35,879 nodes and 35,853 large-strain shellelements (S4R). This modeling approach allows the simulation tocapture the variations of damage and stress states in the modelas a whole, as well as in each individual ply.

    The indentation head of the drop tower was considered as arigid body with a mass of 4.91 kg and providing an impact energyof 100 J (based on experimental conditions). In experimentalobservations, no sliding was observed between the compositeand the clamp plates. As such, the clamps were modeled with xedboundary conditions on the contact surfaces between the clampsand the top and bottom surfaces of the specimens.

    3. Results

    The helicoidal architecture found in the endocuticle of thestomatopod dactyl club was used as inspiration for the design ofimpact-resistant carbon berepoxy composites. Samples of thedactyl club were examined to determine the angle of rotationthroughout the entire endocuticle and ascertain details of the com-plex architecture of the club. Bio-inspired composites were thenfabricated and tested. The responses of the carbon ber sampleswere examined through experimental and analytical approaches,and related to the impact response in the organism.

    ion of ber layers (shown here for a medium-angle composite). (B) Greyscale imagese symmetry achieved by following each right-handed rotation with a left-handede cure cycle.

  • Despite the inability to directly measure or replicate the angles

    G

    5

    s

    1

    test

    ioma3.1. Analysis of a biological hammer

    The energy-absorbing periodic region of the stomatopod club isthe focus of the present study. Transverse polished sections of theclub provide information on fracture behavior in the sample.Cracks in the stomatopod club nest within the periodic region (asshown in Fig. 4A and in greater detail in Fig. 1C). No cracks areobserved to propagate through the entire sample. This is a conse-quence of the Bouligand architecture in the periodic region, andwill be discussed in more detail in Section 6.

    While a helicoidal architecture is a common feature of the crus-tacean cuticle, there are two unique aspects to the rotation in thestomatopod dactyl club. The rst observation is that the lengthof a 180 rotation (pitch) is larger than that observed in most crus-taceans. Secondly, the dactyl club displays a pitch gradient varyingfrom 115 lm near the impact surface to 30 lm at the interior ofthe club (Fig. 4B) as measured using image analysis software. Inter-

    Table 2Composite mechanical properties.

    E1 (GPa) E2 (GPa) m

    153.75 9.7 0.344

    rL (MPa) rL (MPa) rT (MPa) r

    T (MPa)

    2489.01 1985.69 81 312

    Fig. 3. Finite-element model of a drop tower

    L.K. Grunenfelder et al. / Acta Bestingly, there is an initial increase in pitch length, occurring overthe four periods closest to the impact region, followed by adecrease in the pitch length. This decrease is fairly linear untilreaching an asymptotic threshold of 3035 lm. The pitch changeobserved in the club indicates that either the angle of rotation isincreasing moving from the impact region inward, or the spacingbetween ber layers is decreasing.

    High-resolution scanning electron microscopy (SEM) of a trans-verse fractured section in the periodic region reveals the presenceof mineralized ber bundles (Fig. 4C). These bundles measure1.13 0.02 lm in thickness (with individual bers ranging from20 to 50 nm in diameter). In addition, round or elliptical featuresare observed in polished sections within the periodic region(Fig. 4D and E) at the boundary of each pitch, where bers are ori-ented perpendicular to the eld of view. We use these features todene the thickness of each mineralized layer. Subsequent layersare oriented at an angle to the plane of view until a rotation of180 is completed. Using image analysis, these features are mea-sured to be 1.14 0.17 lm in diameter, and are consistent in sizethroughout the stomatopod club, regardless of pitch length (seeFig. 4D and E). This measurement is consistent with the thicknessof a bundle of mineralized bers (Fig. 4C), indicating that each bun-dle denes a unidirectional sheet in the Bouligand architecture. Bydening each layer as having a thickness of 1 lm, the angle of rota-tion throughout the periodic region was calculated to range from1.6 in the outermost region to 6.2 in the inner region of the club.present in the club, the use of multiple angles provides a meansof determining the inuence of a small vs. large angle of rotationon material response.While this calculation of rotation angle is an approximation, itgives a rst estimate to the range of angles present in the club.

    4. Biologically inspired composites

    The aim of this work was to apply the helicoidal design strategyobserved in the stomatopod dactyl club to the fabrication of high-performance composite panels. The rotational angles examined inour biomimetic composites, 7.8, 16.3 and 25.7, were constrainedby the need to maintain a consistent number of layers and imposemid-plane symmetry. These angles were, in most cases, larger thanthose calculated for the stomatopod club, with the smallest angle(7.8) providing a close match to the innermost region of the club.

    and details of the helicoidal composite panel.12 (GPa) G23 (GPa) q (Vf = 0.67) (kg m3)

    .79 2.9 1624.9

    L (MPa) sT (MPa) Gcr;L (kN m1) Gcr;T (kN m1)

    25.48 111.7 40 0.3

    terialia 10 (2014) 39974008 4001The mechanical properties of helicoidal composites were com-pared to those of control layups (unidirectional, quasi-isotropic)to examine the inuence of the rotated design on impact resis-tance. With mid-plane symmetry imposed, all samples remainedat after curing, and no warping was observed. Polished sectionsof each sample conrmed that panels were void-free. Microstruc-tural analysis of each sample type (Fig. 5) demonstrates a changein grey-scale intensity through the thickness the sample (withthe exception of the unidirectional panel), which is caused bychanges in ber rotation angle. Samples were sectioned such thatthe 0 direction is perpendicular to the polished face.

    4.1. Impact response

    Samples measuring 100 mm 150 mmwere subjected to high-energy (100 J) impact using a drop tower. Following the impactevent, the extent of damage to the panels was assessed by examin-ing visible external damage, measuring the depth of the indent,and performing ultrasound scans of the internal damage. Threesamples were tested for each layup condition, yielding consistentresults. The images in Fig. 6AE are representative of the damagetype for each layup conguration. Initially, external damage wasexamined on the backside of the panel, opposite the impact event.The high-energy impact produced catastrophic failure in the unidi-rectional control samples, resulting in a splitting of the panels

  • oma4002 L.K. Grunenfelder et al. / Acta Bi(Fig. 6A). All other samples, however, showed varying degrees ofimpact damage. For the quasi-isotropic controls (Fig. 6B) theimpact event resulted in puncture through the backside of thepanels, accompanied by ber breakage. The helicoidal samples, incontrast, did not show indentation puncture, but rather splits onthe back surface. It is of note that these splits did not, in all cases,occur immediately opposite to the impact location, but were ratherdisplaced from the central impact point. For the small-angle com-posite panels, large cracks were observed along the long edge ofthe sample (Fig. 6C).

    As a second indication of impact damage, the dent depth pro-duced by the impact (on the impact side of the panel) was mea-sured (Fig. 6F). Because the unidirectional panels were split bythe impact, it was not possible to measure dent depths for thosepanels. In all cases, the helicoidal samples showed a reduced dent

    Fig. 4. (A) Overview of the periodic region, showing change in pitch length from the extefrom the impact region inward. (C) Fractured surface highlighting out-of-plane mineralizsmall pitch region. Yellow lines used in (D) and (E) as guides to show arching in sample

    Fig. 5. Micrographs of polished sections from the (A) unidirectional, (B) quasi-isotrterialia 10 (2014) 39974008depth when compared to the quasi-isotropic controls. The percent-age reduction in depth is reported on the plot in Fig. 6F. From dentdepth measurements, the medium angle composites showed thegreatest reduction in external indent damage, with an averagedepth reduction of 49%.

    The internal damage eld caused by impact was investigatedusing a non-destructive pulseecho ultrasound technique. Scanswere performed in a water tank using a 10 MHz transducer inreected mode. Representative ultrasonic C-scan images are pre-sented in Fig. 6GI. The color scale on the images represents thereected signal intensity, with green indicating a reected intensityof approximately 90%, yellow 50% and red 10%. White areas areempty pixels, resulting from regions where no signal was detected(indicating a high degree of internal damage). The extent of theinternal damage eld in the region scanned is marked with dashed

    rior inward. (B) Pitch length as a function of distance through the center of the clubed ber bundles. (D) Polished section of a large pitch region. (E) Polished section ofs.

    opic, (C) small-angle, (D) medium-angle and (E) large-angle composite panels.

  • iomaL.K. Grunenfelder et al. / Acta Blines in Fig. 6, and average area of the damaged region for eachlayup conguration is reported. The ultrasound data reveals thatthe quasi-isotropic control samples (Fig. 6G) have the smallestinternal damage eld. The damage spread in the control samplesis also symmetric. The helicoidal samples, however, show a morewidespread and asymmetric damage eld. The extent of the lateralspread of damage is most pronounced in the small-angle compos-ites (Fig. 6H), and decreases with increasing ply rotation angle.

    4.2. Residual strength

    Following analysis of impact damage, samples were end-loadedand compressed to failure. This compression after impact test proto-col is a composite industry standard test for toughness, as it interro-gates the ability of the sample to support a load after the onset ofdamage. Load vs. displacement data was recorded for each paneltype (Fig. 7A). Compressive residual strength was calculated fromthe peak load achieved during the test (Pmax) divided by the cross-sectional area of the panel (A), as detailed in the following equation:

    FCAI PmaxA

    1

    Details of the residual strength in each panel are presented(Fig. 7B) with the percentage change in residual strength vs. thequasi-isotropic controls indicated for helicoidal samples. Thesmall-angle panels reached the lowest loads during compression

    Fig. 6. Impact damage. Photographs showing damage in the (A) unidirectional, (B) quaDent depth measurements for each sample. Percentage reduction in dent depth listed for(G) quasi-isotropic, (H) small-angle, (I) medium-angle and (J) large-angle composites. Thdamage area for each layup type is reported.terialia 10 (2014) 39974008 4003and thus have the lowest residual strength. The medium- andlarge-angle panels, however, show a marked increase in residualstrength when compared to the quasi-isotropic controls (16% and18%, respectively, Fig. 7B).

    After compressive failure, the damage behavior in each compos-ite was assessed (Fig. 7CF). Dotted black lines mark the lateralextent of visible cracking. For the quasi-isotropic controls(Fig. 7C), the damage is localized to a narrow region directly in linewith the impact event. However, the damage in the helicoidal sam-ples is spread laterally to varying degrees. The small-angle compos-ites, which showed edge cracking following impact, exhibit damageacross the entire length of the test specimen, and show consider-able cracking (Fig. 7D). In the medium-angle composites, cracksare distributed laterally, with damage propagating farther outwardin lower plies. The large-angle composites show the narrowestdamage window for the helicoidal samples, though the damage isstill more disbursed than in the case of the quasi-isotropic controls.To obtainmore information about the damagemechanisms at workin helicoidal composites, computational models were employed.

    5. Modeling of a dynamic strike

    For the computational modeling, the composite material isassumed to have a transversely isotropic elastic response until fail-ure. Using a nite-element framework, damage can therefore bemodeled with Hashin progressive failure criteria [39]. This method

    si-isotropic, (C) small-angle, (D) medium-angle and (E) large-angle composites. (F)helicoidal samples. Ultrasonic C-scan images showing internal damage elds in thee extent of the internal damage eld is indicated by dashed lines, and the average

  • oma4004 L.K. Grunenfelder et al. / Acta Bihas been veried by Icten and Karakuzu through nite-elementanalysis of pinned-joint carbonepoxy composite panels and com-parison with experimental data [40]. The Hashin damage modelallows damage to take place in the ber and matrix. This enablesthe prediction of different failure modes, considering failure as aresult of ber/matrix tension or compression, in addition to speci-fying the damage spatial distribution. The criteria of failure in bermode under tension and compression can be expressed accordingto plane-stress conditions as:

    r11rL

    2 r12

    sL

    2 1 when r11 0

    r11rL

    2 1 when r11 < 0

    2

    While the failure under tensile and compressive state under matrixmode is:

    r22rT

    2 r12

    sL

    2 1 when r22 0

    r222sT

    2 r

    T

    2sT

    2 1

    " #r22rT

    r12

    sL

    2 1 when r22 < 0

    3

    Fig. 7. Composite residual strength and compressive damage. (A) Loaddisplacement cucompressive data, with percentage change listed for helicoidal samples. Photographs ofangle composites after compression tests.terialia 10 (2014) 39974008where E1 is Youngs modulus in the longitudinal direction (berdirection), E2 is the Youngs modulus in the transverse direction, mis the Poissons ratio, G12 is the shear modulus in the longitudinaldirection, G23 is the shear modulus in the transverse direction, qis the density based on 67% ber volume fraction, rL is the tensilestrength in the longitudinal direction, rL is the compressivestrength in the longitudinal direction, sL is the shear strength inthe longitudinal direction, rT is the tensile strength in transversedirection rT is the compressive strength in transverse direction,sT is the shear strength in the transverse direction, and rij are thestress tensor components.

    The nite-element model is set up such that three integrationpoints independently represent the nonlinear mechanical behaviorof each ply across its thickness. In this way this computationalmodel can qualitatively capture the stress components and dam-age mechanisms of every ply at any given time. To effectively visu-alize such failure modes in the simulation results of the 3-Dmodels, a visualization method has been implemented by display-ing only the nodal points that meet Hashin failure criteria, Eqs. (2),(3). This allows the visualization of the 3-D failure patterns in theinterior of the sample. Fig. 8 shows a qualitative description of thedamage distribution in the interior of the plate at a representativetime during the impact event. Additionally, the rotations of

    rves from compression after impact testing. (B) Residual strength calculated fromdamage in the (C) quasi-isotropic, (D) small-angle, (E) medium-angle and (F) large-

  • iomaFig. 8. Damaged nodal points within the composite specimens based on Hashinfailure criteria for the (A) quasi-isotropic control, (B) small-angle composite, (C)medium-angle composite and (D) large-angle composite (see Supplementary videofor more information). Color code indicates the dominant damage mechanisms.

    L.K. Grunenfelder et al. / Acta Bdamage resulting from different ber orientations are clearlyvisible for each (see Supplementary video for more information).

    The dominant failure mechanism near the impact surface of thepanel is observed to consist mainly of ber compression (Fig. 8),indicating that bers near the impact surface undergo bucklingand matrix damage [39]. In the bottom section of the panel, thedominant failure mechanism is matrix cracking. This results fromthe bending mode of the central region of the panel during theimpact process, which leads to tensile forces in the bottom section.Damage in the middle region is not signicant, mainly because thestress values in the neutral plane are not high enough during theimpact process. The model results show a wide in-plane spreadof damage in the small angle composites, which is reduced witha reduction in angle, and smallest in the quasi-isotropic controls.It should be pointed out that the present model does not take intoaccount delamination between plies. However, it clearly capturesthe in-plane damage modes which are consistent with the experi-mental observations.

    6. Discussion

    The stomatopod dactyl club has evolved to withstand thou-sands of high-velocity impacts. The helicoidal architecture in theperiodic region of this robust biological composite provides ameans of dissipating energy from such high-velocity impactevents. This is showcased in the nesting of cracks within the peri-odic region (Fig. 4A). This fracture pattern indicates that cracks fol-low the path of least resistance, propagating between mineralizedbers, rather than severing them. This results in a rotating crackfront, which yields a large surface area per unit crack length inthe direction of crack propagation [10]. The Bouligand structurein the periodic region increases the toughness and energy absorp-tion of the cuticle by redirecting cracks and preventing propagation

    Each layer of colored points represents individual plies.of damage through to the surface of the club, thereby preventingcatastrophic failure [41]. This same strategy is applied in this workto the production of impact-resistant high-performancecomposites.

    In a ber-reinforced structure, a helicoidal layup eliminates thelarge angle mismatch between adjacent layers. As elastic proper-ties are a function of ber angle, this graded design results in asmooth change in in-plane stiffness, reducing interlaminar shearstresses (a key source of delamination) [31,32]. This reduction instress results in the ability to withstand greater deections priorto the onset of delamination [31]. The gradual rotation angle alsoleads to greater in-plane isotropy [42]. Additionally, as is observedin the stomatopod club, a helicoidal organization of bers results ina complex pattern of crack propagation. Here we have shown,through experimental results and computational models, that ahelicoidal layup results in a wide in-plane spread of damage,accompanied by a non-symmetrical rotating damage eld. Cracksin a carbon berepoxy composite will move preferentiallythrough the matrix, rather than severing the stiff bers. With aconstant and gradual change in ber angle in each layer, continu-ous matrix cracking is prevented [32] and cracks are redirectedin every ply, leading to a more tortuous crack path and thus greaterenergy dissipation. Through in-plane spreading of cracks and crackredirection, catastrophic propagation of damage through the thick-ness of the sample is prevented.

    Three helicoidal rotation angles were investigated in this work.Following a high-energy impact event, all three showed reducedthrough-thickness damage in terms of dent depth and visible dam-age when compared to unidirectional and quasi-isotropic controls.Not surprisingly, the unidirectional composites failed catastrophi-cally when subjected to impact. In a unidirectional sample, a crackcan propagate through the entire thickness of the panel withoutthe need to sever bers. Thus, the unidirectional controls were splitdown the center, with the crack running parallel to the alignedber layers. The quasi-isotropic controls, with bers oriented inmultiple orientations, did not fail catastrophically. The impactevent did, however, result in puncture through the back side ofthe panels, indicating that a high degree of damage was transferredthrough the thickness of the samples. Consistent with this hypoth-esis is the measurement of dent depth, which was highest for thequasi-isotropic controls. In the helicoidal samples, dent depthwas reduced, and no puncture was observed. This reduction inexternal damage is attributed to the in-plane spread of damage,observed in ultrasound C-scans (Fig. 6GJ).

    Computational modeling was used to provide additional quali-tative insight into crack behavior in helicoidal composites. Modelresults reveal that the tensile damage in the bottom of each sampleis mainly dominated by matrix fracture (in green, Fig. 8) withoutshowing signs of ber fracture. This is an indication that cracksare aligned to the ber direction in the matrix of the bottom layers.While the quasi-isotropic model (Fig. 8A) shows a distinct damagepattern in each layer, the helicoidal samples (Fig. 8BD) exhibit asmooth transition between consecutive layers. In other words, thehelicoidal samples show a nonsymmetrical damage pattern thatrotates from ply to ply. This smooth rotation does not necessarilyfollow the ber rotation. However, we surmise that the initial dam-age and subsequent stress concentration in the bottom layer isresponsible for guiding the damage that occurs in the upper layers.For example, if a crack in the matrix occurs in the 0 direction in thebottom layer (mostly following the direction of the bers), thestress concentration will affect the directionality of the damagezone in the second and subsequent layers. Although the model isnot capable of indicating the true orientation of the cracks, it clearly

    terialia 10 (2014) 39974008 4005shows the preferential orientation of the damage zone, and thedominant failuremode in each layer. The top layers of the helicoidalsamples, near the impact surface, also exhibit different damage

  • c con w

    omapatterns when compared to the quasi-isotropic control. A smoothrotation between the compression damage is observed in the heli-coidal samples. Simulation results clearly show the highest degreeof in-plane damage in the small-angle helicoidal sample. This par-ticular sample also shows extensive compressive damage in thematrix (which is not shown in the other samples). The extensivedamage in the small-angle simulation is consistent with experi-mental observations. Simulation results (Fig. 9) are consistent withexperimental C-scans of impact damage in (Fig. 6GJ), validatingthe model results, and substantiating the fact that a smaller anglechange provides wider in-plane damage dissipation.

    All helicoidal samples showed a reduction in through-thicknessimpact damage when compared to the controls. However, thesmall-angle composite showed the lowest reduction, which islikely a result of experimental factors, the most critical being sam-ple dimensions. Observations of small-angle samples followingimpact testing, as well as ultrasound C-scans, reveal that the in-plane spread of damage resulting from a 100 J impact in a samplewith a ply rotation angle of 7.8 is sufcient to reach the edges of asample 100 mm in width. The damage, therefore, is not containedwithin the sample, but rather damage progresses to the edge of thepanel, large-scale deformations occur, and the sample surfacebecomes uneven (Fig. 6C). A at surface is necessary to obtainaccurate measurements using a depth gauge, and thus this damagemode inuences the experimental results. The ability to observecracks at the edge of the sample, however, provides further insightinto damage propagation. The cracks observed along the edges ofthe small-angle composite are at an angle that is not parallel tothe panel surface. This indicates that delamination is not the mainfailure mode, but rather that crack jumping is occurring betweenangled layers. Similar behavior has been reported in previous stud-ies, as off-axis plies are more susceptible to matrix cracking thandelamination [43].

    After impact, the residual strength of the samples, or their abil-ity to carry load following the onset of damage, was measuredusing an in-plane compression test. Residual strength is a criticalmeasure of toughness in ber-reinforced composites. Composite

    Fig. 9. Model results (top view of damaged nodal points) for the (A) quasi-isotropicomposite. The zero degree direction runs left to right to allow for direct compariso

    4006 L.K. Grunenfelder et al. / Acta Bipanels are often damaged in service (e.g. bird strikes on aircraft[44]), and detection of damage in composite structures is a difculttask [45]. Therefore, it is essential that composite designs maintainstrength after damage is initiated. The compression after impacttesting of helicoidal composites revealed an increase in residualstrength for the medium and large-angle samples. Again, the per-formance of the small-angle composites were adversely inuencedby edge cracking. An increase in residual strength for a helicoidallayup has been suggested in the past via three-point bend test data[27], but has not previously been shown for samples with impactdamage, or reported using a standard test method. The observationof an increase in residual strength as a function of ber angle hasbroad impact. All samples examined in this study were fabricatedwith the same number of layers, using the same constituent mate-rials. Introducing a helicoidal layup scheme, however, results inincreased toughness.All plies for the carbon berepoxy samples fabricated in thiswork were measured, cut and laid up by hand. This introducesthe possibility of human error, specically in the form of ply mis-alignment, which will directly inuence sample performance. Thepotential for misalignment is greatest for smaller rotation angles,another factor which may have inuenced the response of thesmall-angle composite. While hand layup of complex laminategeometries is tedious and prone to error, helicoidal compositescould be readily produced in industry using automated layup ofunidirectional tapes, a common practice. Thus, the ndings in thisstudy are of practical importance for aerospace, automotive andarmor applications, where impact resistance and weight savingsare critical.

    Through the examination of multiple helicoidal rotation angles,conclusions can be drawn regarding the inuence of angle onmate-rial response. As mentioned, results for the small-angle compositeswere inuenced by experimental constraints. The medium- andlarge-angle data, however, reveal an increase in through-thicknessimpact resistance for a smaller angle of rotation. This observation isconsistent with previous studies, which have shown an increase inother properties (exural strength and stiffness) for a smaller berrotation angle [27]. These ndings are of interest considering thestructural arrangement of the stomatopod dactyl club. From SEManalysis, we hypothesize that the rotation angle in the club is small-est close to the impact surface. Therefore, we can expect optimumin-plane energy dissipation in the region most inuenced by thehigh-energy strike. Additionally, a smaller rotation angle equatesto a larger pitch length and a higher number of mineralized berlayers. More ber layers will result in increased crack redirection,as well as crack arrest, as cracks propagating through the mineralphase are known to stop when an organic ber is encountered[20]. This tougheningmechanism also sheds light on the large pitchlength observed in the club.

    The pitch length in the stomatopod dactyl club, measured in thesample investigated here to reach up to 115 lm, is considerablylarger than the pitch lengths reported for other crustaceans. Inthe American lobster, Homarus americanus, for example, examina-

    ntrol, (B) small-angle composite, (C) medium-angle composite and (D) large-angleith the ultrasound scans in Fig. 6.

    terialia 10 (2014) 39974008tion of both claw and carapace samples show average pitch lengthsof 610 lm in the exocuticle and 2535 lm in the endocuticle[19,24,26,42], with the pitch length remaining relatively constantthroughout each region [19]. Although the stomatopod is a smalleranimal than the lobster, it employs a different predation strategy,implementing a hammer-like strike (high peak force, short time),rather than a crushing mechanism (low force, long time), todestroy prey [11]. This high-energy strike results in a need forimpact resistance not required in the conventional arthropod exo-skeletal structure. With a larger pitch length indicative of a smallerrotation angle, and a larger number of mineralized ber layers,both factors which result in increased energy dissipation, it is notsurprising that the dactyl club displays a larger pitch than cuticlesamples from other crustaceans.

    Through examination of the dactyl club, we have revealed anexpanded pitch length in the Bouligand structure of the periodic

  • helicoidal samples, with a lower degree of damage propagation

    iomaCertain gures in this article, particularly Figs. 14 and 69 aredifcult to interpret in black and white. The full colour images canbe found in the on-line version, at http://dx.doi.org/10.1016/j.actbio.2014.03.022

    References

    [1] Wegst UGK, Ashby MF. The mechanical efciency of natural materials. PhilMag 2007;84:216786.

    [2] Dunlop JWC, Fratzl P. Biological composites. Annu Rev Mater Res2010;40:124.

    [3] Giraud-Guille M-M. Plywood structures in nature. Curr Opin Solid State MaterSci 1998;3:2217.

    [4] Sarikaya M. An introduction to biomimetics: a structural viewpoint. MicroscRes Tech 1994;27:36075.

    [5] Luz GM, Mano JF. Biomimetic design of materials and biomaterials inspired bythrough the thickness of the panel. Three ber rotation angles wereinvestigated in this work. Because of sample dimensions, the resultsfor the small-angle panels were skewed. Examination of medium-and large-angle samples, however, showed a reduction inthrough-thickness impact damage, and an increase in residualstrength over control samples, and indicated that a smaller rotationangle results in a reduction in impact damage.

    Acknowledgments

    The authors gratefully acknowledge nancial support from theAir Force Ofce of Scientic Research (AFOSR-FA9550-12-1-0245)and the National Science Foundation (DMR-0906770). Assistancefrom Dr. Timotei Centea at the University of Southern Californiaand Mark Ostermeier at Cytec Engineered Materials is alsoacknowledged.

    Appendix A. Supplementary data

    Supplementary data associated with this article can be found, inthe online version, at http://dx.doi.org/10.1016/j.actbio.2014.03.022.

    Appendix B. Figures with essential colour discriminationIn this work, we have examined the impact resistance of a heli-coidal architecture in both a biological composite (the stomatopoddactyl club) and ber-reinforced samples. Carbon berepoxy com-posites with a helicoidal architecture were shown to exhibit achange in impact response when compared to unidirectional andquasi-isotropic controls. Damage was shown to spread in-plane inregion, as well as a distinct pitch gradient. These ndings indicate asmaller rotation angle at the outer region of the periodic and thusgreater energy absorption (in-plane damage spread) in the regionnearest the impact. If the hypothesis of a change in ber angle asa function of pitch length is inaccurate, the remaining possibilityis that the rotation angle remains consistent through the club,and the spacing between mineralized ber layers increases as afunction of pitch length. This scenario would imply that the outerregion of the periodic is more heavily mineralized, with the innerregion having a denser packing of a-chitin layers. Evidence of thishas not been seen in past elemental analysis or micromechanicaltesting of the club [10], though higher-resolution data would berequired to observe a mineral gradient.

    7. Conclusions

    L.K. Grunenfelder et al. / Acta Bthe structure of nacre. Philos Trans R Soc A 2009;367:1587605.[6] Barhelat F. Nacre frommollusk shells: a model for high-performance structural

    materials. Bioinspir Biomim 2010;5.[7] Finnemore A, Cunha P, Shean T, Vignolini S, Guldin S, Oyen M, et al. Biomimeticlayer-by-layer assembly of articial nacre. Nat Commun 2012;3.

    [8] Espinosa HD, Juster AL, Latourte FJ, Loh OY, Gregoire D, Zavattieri PD. Tablet-level origin of toughening in abalone shells and translation to syntheticcomposite materials. Nat Commun 2011;2.

    [9] Wang J, Cheng Q, Tang Z. Layered nanocomposites inspired by the structureand mechanical properites of nacre. Chem Soc Rev 2012;41:9451404.

    [10] Weaver JC, Milliron GW, Miserez A, Evans-Lutterodt K, Herrera S, Gallana I,et al. The stomatopod dactyl club: a formidable damage-tolerant biologicalhammer. Science 2012;336:127580.

    [11] Patek SN, Caldwell RL. Extreme impact and cavitation forces of a biologicalhammer: strike forces of the peacock mantis shrimp Odontodactylus scyllarus. JExp Biol 2005;208:365564.

    [12] Full RJ, Caldwell RL, Chow SW. Smashing energies: prey selection and feedingefciency of the stomatopod, Gonodactylus bredini. Ethology 1989;81.

    [13] Currey JD, Nash A, Boneld W. Calcied cuticle in the stomatopod smashinglimb. J Mater Sci 1982;17:193944.

    [14] Patek SN, Korff WL, Caldwell RL. Biomechanics: deadly strike mechanism of amantis shrimp. Nat Commun 2004;428:81920.

    [15] Fabritius H, Sachs C, Raabe D, Nikolov S, Friak M, Neugebauer J. Chitin in theexoskeletons of arthropoda: from ancient design to novel materials science. In:Gupta NS, editor. Chitin. Netherlands: Springer; 2011. p. 3560.

    [16] Bouligand Y. Twisted brous arrangements in biological materials andcholesteric mesophases. Tissue Cell 1972;4:189217.

    [17] Bouligand Y. Sur une architecture torsadee repandue dans de nombreusescuticles dArthropodes. C R Acad Sci Paris 1965;261:36658.

    [18] Fabritius H-O, Karsten ES, Balasundaram K, Hild S, Huemer K, Raabe D.Correlation of structure, composition and local mechanical properties in thedorsal carapace of the edible crab Cancer pagurus. Z Kristallogr2012;227:76676.

    [19] Raabe D, Sachs C, Romano P. The crustacean exoskeleton as an example of astructurally and mechanically graded biological nanocomposite material. ActaMater 2005;53:428192.

    [20] Bouligand Y. The renewal of ideas about biomineralisations. C R Palevol2004;3:61728.

    [21] Sachs C, Fabritius H, Raabe D. Inuence of microstructure on deformationanisotropy of mineralized cuticle from the lobster Homarus americanus. JStruct Biol 2008;161:12032.

    [22] Nikolov S, Petrov M, Lymperakis L, Friak M, Sachs C, Fabritius H-O, et al.Revealing the design principles of high-performance biological compositesusing ab initio and multiscale simulations: the example of lobster cuticle. AdvMater 2010;22:51926.

    [23] Bobelmann F, Romano P, Fabritius H, Raabe D, Epple M. The composition of theexoskeleton of two crustacea: the American lobster Homarus americanus andthe edible crab Cancer pagurus. Thermochim Acta 2007;463:658.

    [24] Fabritius H-O, Sachs C, Triguero PR, Raabe D. Inuence of structural principleson the mechanics of a biological ber-based composite material withhierarchical organization: the exoskeleton of the lobster Homarusamericanus. Adv Mater 2009;21:391400.

    [25] Sachs C, Fabritius H, Raabe D. Experimental investigation of the elastic-plasticdeformation of mineralized lobster cuticle by digital image correlation. J StructBiol 2006;155:40925.

    [26] Sachs C, Fabritius H, Raabe D. Hardness and elastic properties of dehydratedcuticle from the lobster Homarus americanus obtained by nanoindentation. JMater Res 2006;21:198795.

    [27] Cheng L, Thomas A, Glancey JL, Karlsson AM. Mechanical behavior of bio-inspired laminated composites. Compos A 2011;42:21120.

    [28] Zimmermann EA, Gludovatz B, Schaible E, Dave NKN, Yang W, Meyers MA,et al. Mechanical adaptability of the Bouligand-type structure in naturaldermal armour. Nat Commun 2013;4:17.

    [29] Andersons J, Konig M. Dependence of fracture toughness of compositelaminates on interface ply orientations and delamination growth direction.Compost Sci Technol 2004;64:213952.

    [30] Chen B, Peng X, Cai C, Niu H, Wu X. Helicoidal microstructure of Scarabaeicuticle and biomimetic research. Mater Sci Eng A 2006;423:23742.

    [31] Ravi-Chandar K. Design optimization and characterization of helicoidalcomposites with enhanced impact resistance. Army Research Ofce 2011.

    [32] Apichattrabrut T, Ravi-Chandar K. Helicoidal composites. Mech Adv MaterStruct 2006;13:6176.

    [33] ASTM. D7136/D7136M-12. Standard test method for measuring the damageresistance of a ber-reinforced polymer matrix composite to a drop-weightimpact.

    [34] ASTM. D7137/D7137M-12. Standard test method for compressive residualstrength properties of damaged polymer matrix composite plates.

    [35] Sanchez-Saez S, Barbero E, Zaera R, Navarro C. Compression after impact ofthin composite laminates. Compos Sci Technol 2005;65:19119.

    [36] ABAQUS, version 6.1. Providence, RI.[37] Hansen P, Martin R. DCB, 4ENF and MMB delamination characterization of S2/

    8552 and IM7/8552. London: European Research Ofce of the US Army; 1999.[38] Li S, Thouless MD, Wass AM, Schoroeder JA, Zavattieri PD. Use of mode-I

    cohesive-zone models to describe the fracture of an adhesivley-bondedpolymer-matrix composite. Compos Sci Technol 2005;65:28193.

    [39] Hashin Z. Failure criteria of unidirectional ber composites. J Appl Mech

    terialia 10 (2014) 39974008 40071980;47:32934.[40] Icten BM, Karakuzu R. Progressive failure analysis of pin-loaded carbon-epoxy

    woven composite plates. Compos Sci Technol 2002;62:125971.

  • [41] Meyers MA, McKittrick J, Chen P-Y. Structural biological materials: criticalmehcanics-materials connections. Science 2013;339:7739.

    [42] Cheng L, Wang L, Karlsson AM. Image analysis of two crustacean exoskeletonsand implications of the exoskeletal microstructure on the mechanicalbehavior. J Mater Res 2008;23:285472.

    [43] Tao J, Sun CT. Inuence of ply orientation on delamination in compositelaminates. J Compos Mater 1998;32:193347.

    [44] Georgiadis S, Gunnion AJ, Thomson RS, Cartwright BK. Bird-strike simulationfor certication of the Boeing 787 composite moveable trailing edge. ComposStruct 2008;86:25868.

    [45] Zhou G, Sim LM. Damage detection and assessment in ber-reinforcedcomposite structures with embedded bre optic sensorsreview. SmartMater Struct 2002;11:92539.

    4008 L.K. Grunenfelder et al. / Acta Biomaterialia 10 (2014) 39974008

    Bio-inspired impact-resistant composites1 Introduction2 Experimental2.1 Stomatopod specimen handling and sample preparation2.2 Composite fabrication2.3 Analysis and mechanical testing of composites2.4 Modeling of impact damage

    3 Results3.1 Analysis of a biological hammer

    4 Biologically inspired composites4.1 Impact response4.2 Residual strength

    5 Modeling of a dynamic strike6 Discussion7 ConclusionsAcknowledgmentsAppendix A Supplementary dataAppendix B Figures with essential colour discriminationReferences