You are on page 1of 5

journal of the mechanical behavior of biomedical materials 56 (2016) 1 5

Available online at www.sciencedirect.com

www.elsevier.com/locate/jmbbm

Research paper

Exploring the shock response of spider webs


V. Tietscha, J. Alencastreb,n, H. Wittea, F.G. Torresb,n
a

FachgebietBiomechatronik, Technische Universitt Ilmenau, Max-Planck-Ring 12, 98693 Ilmenau, Germany


Department of Mechanical Engineering, Ponticia Universidad Catolica del Peru, Av. Universitaria 1801, Lima 32, Peru

art i cle i nfo

ab st rac t

Article history:

Spider orb-webs are designed to allow for quick energy absorption as well as the constraint

Received 6 July 2015

of drastic oscillations occurring upon prey impact. Studies on spider silk illustrate its

Received in revised form

impressive mechanical properties and its capacity to be used as technical bers in

2 November 2015

composite materials. Models have previously been used to study the mechanical properties

Accepted 9 November 2015

of different silk bers, but not the behavior of the spider web as a whole. Full spider webs

Available online 19 November 2015

have been impacted by a projectile and the transverse displacement was measured by

Keywords:

means of a laser interferometer. The damping and stiffness of the entire webs were

Spider silk

quantied considering the orb-web as a single-degree-of-freedom (SDOF) system. The

Spider orb-web

amplitude, the period duration, and the energy dissipation of the oscillations have also

Vibration

been reported from the experiments. The analysis of the energy dissipation conrmed that

Dynamic behavior

the webs of orb-web spiders are optimized for the capture of a single or few large prey,
rather than several small prey. The experiments also conrmed that the overall stiffness of
the web displayed a non-linear behavior. Such non-linearity was also observed in the
damping characteristics of the webs studied.
& 2015 Elsevier Ltd. All rights reserved.

1.

Introduction

made from dragline silk (Prez-Rigueiro et al., 2010), whereas


the spiral threads are made from viscid silk. The spiral

The mechanical performance of spiders webs allows them to

threads are glue-coated and capture the prey (Torres et al.,

absorb the energy of ying prey (Blackledge and Hayashi,

2014), and it has been hypothesized that these spiral threads

2006; Elices et al., 2009). Spider silks must have enough

function as shock absorbers through viscous dissipation as

strength to absorb the kinetic energy of ying insect prey,


while also minimizing the return of that energy to the prey in

kinetic energy is converted to heat (Denny, 1976).


Spider orb-webs are designed to allow for quick energy

order to prevent it from bouncing off the web (Kelly et al.,

absorption as well as the constraint of drastic oscillations,

2011). Spiders use different types of silks such as viscid,

which occurs upon prey impact (Du et al., 2011). These

dragline, spiral, and cocoon silk (Gosline et al., 1986), each

vibrations are used by spiders to locate and attack prey.

of which has different mechanical properties and functions.

Barrows (1915) reported that Epeira sclopetaria, an orb-weaving

The primary silks composing the web are the dragline and

spider, orients itself in the center of the web in order to

viscid silks; the structural radial threads of the orb webs are

charge and seize ies that impact the web (Parry, 1965).

Corresponding authors.
E-mail addresses: jalenca@pucp.edu.pe (J. Alencastre), fgtorres@pucp.pe (F.G. Torres).

http://dx.doi.org/10.1016/j.jmbbm.2015.11.007
1751-6161/& 2015 Elsevier Ltd. All rights reserved.

journal of the mechanical behavior of biomedical materials 56 (2016) 1 5

Moreover, Craig et al. (1985) reported that small spiders build


oscillating webs, and those webs characterized by large
amplitude oscillations are able to intercept small, slow ying
prey more efciently than webs characterized by small
amplitude oscillations.
Sensenig et al. (2012) reported that in a web of Araneus
trifolium impacted by a projectile of approximately 180 mJ of
kinetic energy, most of the energy transfer occurs in the rst
100 ms, while the remainder occurs in subsequent oscillations. The study also found that the web could oscillate more
than six times; however, only negligible energy remains after
the rst two oscillations. This is supported by a study by Alam
et al. (2007), which also reports that the maximum dynamic
response is found in the rst natural frequency of the web.
Dynamic mechanical analysis (DMA) tests of individual
spider threads have been reported in the literature
(Blackledge et al., 2005; Torres et al., 2013). Blackledge et al.
(2005) found that the loss tangent of spider silk rapidly
increases during the rst 23% of strain, and reaches a
maximum at the yield point. They also reported an initial
storage modulus of 911 GPa for the rst 2% of strain for the
bers of different types of spider silk. At these small strains,
low damping levels can be expressed. This, according to Kelly
et al. (2011), allows orb webs to retain full functionality of
their silk until the impact of a larger, more energetically
valuable prey. By contrast, at relatively high strains under
repetitive loading conditions, orb webs from species such as
Argiope aurantia or Argiope trifasciata exhibit damping capacities of 4050% (Kelly et al., 2011).
Studies on spider silk illustrate its impressive mechanical
properties and its capacity to be used as technical bers in
composite materials. Bai et al. (2006) reported that silk can be
modied for direct use in micro-electromechanical systems
or as a part of a special composite material that enhances
mechanical properties. Moreover, Allmeling et al. (2006)
reported that spider silk is resistant to fungal and bacterial
decomposition and that its molecular structures promote
cellular adhesion and migration, making it an ideal biomaterial candidate.
Models have previously been used to study the mechanical properties of different silk bers, but not the behavior of
the spider web as a whole. Porter et al. (2005) developed one
such model, taking into account the chemical composition
and semi-crystalline morphology of the bers. A molecular
model of dragline silk elasticity has also been developed to
reproduce the complex stressstrain curves for the dragline
in wet and dry states (Termonia, 1994).
Additionally, studies have used projectiles simulating
ying prey to study the energy absorbed and dissipated by
spider webs (Kelly et al., 2011; Sensenig et al., 2012, 2013,
2010). Sensenig et al. (2012) used video-based imaging methods to measure the length change of the silk threads of orb
webs during the projectile impact. By using the damping
capacity measured for individual threads, the energy
absorbed and dissipated by the entire orb web was estimated
for different species of spiders.
In this study, we used an experimental setup that allowed
us to directly measure the vibration of a spider orb-web as a
whole. The aim of this work is to quantify the characteristics
of the vibrations caused by prey impact in order to be able to

Fig. 1 Variation of displacement with time during a


projectile impact.

establish the mechanical properties of orb-webs, under the


assumption that such vibrations are forced vibrations of a
damped, single degree of a freedom, spring mass system.

2.

Results and discussion

For the experiments reported here a disk was attached to the


center of the web and the position, and velocity of that disk
was measured while being impacted by a projectile. Fig. 1
shows the horizontal displacement of the disk during the
impact. A decaying amplitude of the oscillations can be
observed as a function of time. The rst deection represents
the displacement during the rst impact of the prey. This
effect allows the web to absorb high kinetic impacts that
would otherwise cause an extension above the elastic range
and therefore destroy the web. A similar response has been
reported for a single spider silk thread by Du et al. (2011).
These results are also in agreement with the results reported
by Sensenig et al. (2012). In those experiments, the energy
dissipation in orb webs spun by diverse species of spiders
was estimated using data derived from high-speed videos of
web deformation under prey impact together with the damping properties of individual threads. They estimated that
most of the energy damping occurred within 0.1 s and that
approximately 50% of the remaining prey energy was dissipated at each subsequent oscillation.
For our analysis, the dynamic behavior of the web system
studied here was represented by a single-degree-of-freedom
(SDOF) system. Linear SDOF systems are commonly used to
describe the vibration of simple systems. In this type of
systems, stiffness and damping are constants. A linear
model, however, was not able to represent the behavior of
the system during the rst damped vibration in the experiments reported here. Non-linear SDOF systems have been
used in cases with dry friction, Coulomb damping and nonlinear stiffness (Chen and Tomlinson, 1996). We have also
used a dynamic non-linear model to t our experimental data
considering that stiffness and damping were not constant.
Fig. 2 shows the horizontal displacement of the disk

journal of the mechanical behavior of biomedical materials 56 (2016) 1 5

Fig. 2 Comparison between the experimental


measurements and the dynamic model for the displacement
of the web.

!
4 2
c2

m
T2web 4m2

Fig. 4 shows the dependence of the stiffness as a function


of the maximum displacement per period. An exponential
function was used to plot an interpolation. We considered the
stiffness to be function only of the displacement. The stiffness of the web at displacements and velocities near zero
approaches a constant value. In contrast, the value of the
stiffness in the rst cycles is not constant. The variation of
the stiffness values can be accounted for geometric and
material properties, pre-stressing forces on the web, as well
as the construction parameters of an individual spider web.
The non-linear behavior of the system stiffness could also
be attributed to a stiffening caused by the deformation of
individual silk threads. However, we have estimated the
average strain of an individual thread during impact at
around 0.05%. Previous work on individual spider silk bers
showed that considerable stiffening of spider silk occurs at
strain values well over 10%. We can therefore conclude that
single ber stiffening shall not be the main source of the
dynamic non-linearity of the system (Du et al., 2011).
A comparison between the energy dissipation from the
dynamic non-linear model, the experimental measurements,
and a linear model (c and k Econstants) is depicted in Fig. 5. It
is clear that the linear model differs from the measured energy dissipation of the web, while the dynamic nonlinear model more accurately describes such behavior. The

Fig. 3 Damping as a function of velocity of vibration.

measured during impact together with the non-linear calculated points. This dynamic non-linear model adequately
describes the decaying amplitude of displacement during
the impact of a projectile. This suggests a more accurate
representation of stiffness and damping with regard to the
linear model.
In fact, damping and stiffness are functions of different
parameters such as displacement, velocity, and acceleration.
The damping of the system was calculated as a function of
the amplitude during each vibration cycle. Fig. 3 shows the
damping of the system plotted as a function of velocity. The
graph shows a non-linear damping behavior in the form of an
increasing e-function. For the sake of simplicity, damping
was considered to be a function only of velocity. This is in
agreement with the results found by Lin et al. (1995). They
showed that the aerodynamic damping (which depends on
the velocity) of the spider web, produced by the movement of
the web against the air, plays a key role on the dissipation of
energy.
In order to calculate the stiffness an empirical expression
that relates the stiffness, damping and mass of the system
with the period duration (Tweb) was used:

  0:5
k
c2

1
Tweb 2
2
m 4m
where m, c and k stand for mass, damping and stiffness,
respectively.
The period duration and the mass were directly measured.
The stiffness was calculated by solving Eq. (2) for k:

Fig. 4 Stiffness as a function of web displacement.

Fig. 5 Energy dissipation of the web: comparison between


the experimental measurements, the dynamic model and a
linear model.

journal of the mechanical behavior of biomedical materials 56 (2016) 1 5

apparent differences between the measurements and the


dynamic model are assumed to result from the viscoelastic
properties of the spider silk, the composite character of the
orb-web consisting of different types of silks, and the special
geometric assembly of the orb-web in radii and helices.
The energy dissipated when a real prey impacts the web
would be higher than the energy measured here due to the
fact that the plastic spheres used in this study bounced off
the web after impact. A real prey would stick to the web and
therefore more energy would be dissipated. Further studies
would be necessary to measure the energy dissipated in a real
prey impact.
Fig. 5 also shows that in the rst periods the observed
capacity for energy dissipation is higher than predicted by the
linear model. The energy dissipation of the web follows an
exponential trend and it reaches much higher values, compared to the conventional linear system. This is in part due to
the shock absorption and higher stiffness in the rst periods
(Fig. 1), and to some extent a consequence of the higher
damping at higher velocities (Fig. 3).
The initial non-linearity of the system would provide the
orb-web with the capacity of dissipating the impact energy of
large prey and preventing damages for such an impact. In
fact, it has been shown that the bulk of biomass captured
over a spider's lifetime comes from large prey (Venner and
Casas, 2005). Sensenig et al. (2012) suggested that while
spiders capture numerous small prey, their webs are optimized to target large, high-value prey due to the large
energetic return from such prey.

3.

Experimental

3.1.

Materials

Fig. 6 Orb web of Argiope argentata in the PMMA frame


prepared for the experiments.
custom-made leaf spring at the contact instant with the
projectiles. Projectiles weight was also necessary for kinetic
energy calculation.
The projectiles were red at a distance of 10 cm from the
webs. The web motion after impact was assessed by means of
a Laser-Interferometer (BREL & KJR TYPE 3544), which
tracked the position of the polystyrene disks over time. The
tests were carried out at 23 1C and a relative humidity of 85%.

3.3.
Orb-webs from Argiope argentata (Fig. 6), an American species
of the Araneidae family, were used in this study. Female
adult spiders of 3.55 mm in prosoma width were collected in
the vicinity of the outskirts of Lima, Peru. The weight of the
spiders ranged from 200 to 350 mg. Each spider was housed in
a 400 mm  400 mm PMMA cage. Orb-webs were constructed
by the spiders on rectangular PMMA frames placed inside
the cages.

3.2.

Experimental methods

The experimental rig used is shown in Fig. 7. The spiders


were removed from the constructed web and were replaced
by foamed polystyrene disks of 4 mm in thickness with
approximately the same weight of the spiders. One side of
the disk had a small reector to allow for the no-contact
interferometer measurement. The other side was used as the
contact area for the projectiles that simulate the ying prey.
Eleven plastic spheres were used as projectiles to represent the ying prey (See Table 1 in the Supplementary
information section). The projectiles were shot at the web
using a custom-made leaf spring, which had an accelerometer (BREL & KJR TYPE 4383) attached to allow for the
projectile velocity measure. Kinetic energy was calculated for
each experiment by measuring the maximum velocity of the

The single-degree-of-freedom (SDOF) model

In order to understand the web response during the impact, a


simplied, damped, single degree of freedom model was used
(See Supplementary information). Amplitudes and period
times of each single vibration cycle were recorded. The
recorded values were used to obtain the stiffness and damping of the system as a function of the vibration parameters
such as the period duration, amplitude, velocity and acceleration. The threads were modeled as KelvinVoigt viscoelastic materials. All measurement data and model results
were processed using the same algorithms to ensure a valid
comparison. The models were simulated using Simulinks.

4.

Conclusions

We have performed dynamic experiments on full spider webs


in order to characterize their shock response. Our experiments conrmed that a spider web can be described as a
dynamic system, which has time-dependent properties as
the impact-caused vibrations decay. Considering the orbweb as a single-degree-of-freedom system, allowed for the
description of the amplitude, the period duration, and the
energy dissipation during impact. The stiffness plots obtained as function of displacement conrmed that stiffness is

journal of the mechanical behavior of biomedical materials 56 (2016) 1 5

Fig. 7 The experimental rig used. The PMMA frame with the spider's web was clamped to a wooden holding structure. The
wooden structure was covered with polystyrene to isolate the experimental rig from air ow and other disturbances during
the measurement. After the PMMA frame was installed, the interferometer and the projectile were aligned and the
measurements were carried out.

not constant and its value increases steadily during the rst
few vibration cycles after impact, allowing for increased
energy dissipation. This energy dissipation exceeds the
values predicted by a linear model, conrming that the webs
of orb-web spiders are optimized to dissipate the kinetic
energy of large ying prey.

Acknowledgment
F.T. and J.A. would like to acknowledge the Vice-Rectorate of
Research of the Ponticia Universidad Catolica del Peru for
nancial support (VRI-165).

Appendix A.

Supplementary material

Supplementary data associated with this article can be found


in the online version at http://dx.doi.org/10.1016/j.jmbbm.
2015.11.007.

r e f e r e n c e s

Blackledge, T.A., Hayashi, C.Y., 2006. J. Exp. Biol. 209, 31313140.


Elices, M., Plaza, G.R., Arnedo, M.A., Perez-Rigueiro, J., Torres, F.G.,
Guinea, G.V., 2009. Biomacromolecules 10, 19041910.
Kelly, S.P., Sensenig, A., Lorentz, K.A., Blackledge, T.A., 2011.
Zoology 114, 233238.
Gosline, J.M., DeMont, M.E., Denny, M.W., 1986. Endeavour 10,
3743.
Perez-Rigueiro, J., Plaza, G.R., Torres, F.G., Hijar, A., Hayashi, C.,
Perea, G.B., Elices, M., Guinea, G.V., 2010. Int. J. Biol. Macromol.
46, 555557.

Torres, F.G., Troncoso, O.P., Cavalie, F., 2014. Mater. Sci. Eng. C 34,
341344.
Denny, M., 1976. J. Exp. Biol. 65, 483506.
Du, N., Yang, Z., Liu, X.Y., Li, Y., Xu, H.Y., 2011. Adv. Funct. Mater.
21, 772778.
Barrows, W.M., 1915. Biol. Bull. 29, 316332.
Parry, D.A., 1965. J. Exp. Biol. 43, 185192.
Craig, C.L., Okubo, A., Andreasen, V., 1985. J. Theor. Biol. 115,
201211.
Sensenig, A.T., Lorentz, K.A., Kelly, S.P., Blackledge, T.A., 2012. J. R.
Soc. Interface 9, 18801891.
Alam, M.S., Wahab, M.A., Jenkins, C.H., 2007. Mech. Mater. 39,
145160.
Blackledge, T.A., Swindeman, J.E., Hayashi, C.Y., 2005. J. Exp. Biol.
208, 19371949.
Torres, F.G., Troncoso, O.P., Torres, C., Cabrejos, W., 2013. Mater.
Sci. Eng. C 33, 14321437.
Bai, J., Ma, T., Chu, W., Wang, R., Silva, L., Michal, C., Chiao, J.C.,
Chiao, M., 2006. Biomed. Microdevices 8, 317323.
Allmeling, C., Jokuszies, A., Reimers, K., Kall, S., Vogt, P.M., 2006.
J. Cell. Mol. Med. 10, 770777.
Porter, D., Vollrath, F., Shao, Z., 2005. Eur. Phys. J. E: Soft Matter
Biol. Phys. 16, 199206.
Termonia, Y., 1994. Macromolecules 27, 73787381.
Sensenig, A.T., Kelly, S.P., Lorentz, K.A., Lesher, B., Blackledge,
T.A., 2013. J. Exp. Biol. 216, 33883394.
Sensenig, A.T., Agnarsson, I., Blackledge, T.A., 2010. J. Evol. Biol.
23, 18391856.
Chen, Q., Tomlinson, G.R., 1996. J. Vib. Acoust. 118, 252263.
Lin, L., Edmonds, D., Vollrath, F., 1995. Nature 373, 146148.
Du, N., Yang, Z., Yang Liu, X., Li, Y., Yao Xu, H., 2011. Adv. Funct.
Mater. 21, 772778.
Venner, S., Casas, J., 2005. Proc. R. Soc. B 272, 15871592.

You might also like