Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/49626447

The effects of water and microstructure on the mechanical properties of


bighorn sheep (Ovis canadensis) horn keratin

Article  in  Acta biomaterialia · March 2011


DOI: 10.1016/j.actbio.2010.11.024 · Source: PubMed

CITATIONS READS

33 242

9 authors, including:

Michael Wesley Trim Mark F Horstemeyer


Engineer Research and Development Center - U.S. Ar... Liberty University
8 PUBLICATIONS   247 CITATIONS    530 PUBLICATIONS   10,627 CITATIONS   

SEE PROFILE SEE PROFILE

Hongjoo Rhee Lakiesha N Williams


Mississippi State University University of Florida
42 PUBLICATIONS   493 CITATIONS    57 PUBLICATIONS   371 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Rheology of Mucin-Nanoparticle Solutions View project

football helmet View project

All content following this page was uploaded by Mark F Horstemeyer on 24 February 2020.

The user has requested enhancement of the downloaded file.


Acta Biomaterialia 7 (2011) 1228–1240

Contents lists available at ScienceDirect

Acta Biomaterialia
journal homepage: www.elsevier.com/locate/actabiomat

The effects of water and microstructure on the mechanical properties


of bighorn sheep (Ovis canadensis) horn keratin
Michael W. Trim a,b,⇑, M.F. Horstemeyer a,b, Hongjoo Rhee b, Haitham El Kadiri a,b, Lakiesha N. Williams c,
Jun Liao c, Keisha B. Walters d, Joanna McKittrick e,f, Seong-Jin Park g
a
Department of Mechanical Engineering, Mississippi State University, Mississippi State, MS 39762, USA
b
Center for Advanced Vehicular Systems (CAVS), Mississippi State University, Mississippi State, MS 39762, USA
c
Department of Agriculture and Biological Engineering, Mississippi State University, Mississippi State, MS 39762, USA
d
Dave C. Swalm School of Chemical Engineering, Mississippi State University, Mississippi State, MS 39762, USA
e
Materials Science and Engineering Program, University of California, San Diego, La Jolla, CA 92037, USA
f
Department of Mechanical and Aerospace Engineering, University of California, San Diego, La Jolla, CA 92037, USA
g
Department of Mechanical Engineering, Pohang University of Science and Technology (POSTECH), Pohang, Gyeongbuk 790-784, Korea

a r t i c l e i n f o a b s t r a c t

Article history: The function of the bighorn sheep horn prompted quantification of the various parametric effects impor-
Received 20 May 2010 tant to the microstructure and mechanical property relationships of this horn. These parameters included
Received in revised form 16 November 2010 analysis of the stress-state dependence with the horn keratin tested under tension and compression, the
Accepted 17 November 2010
anisotropy of the material structure and mechanical behavior, the spatial location along the horn, and the
Available online 21 November 2010
wet–dry horn behavior. The mechanical properties of interest were the elastic moduli, yield strength,
ultimate strength, failure strain and hardness. The results showed that water has a more significant effect
Keywords:
on the mechanical behavior of ram horn more than the anisotropy, location along the horn and the type of
Horn
Keratin
loading state. All of these parametric effects showed that the horn microstructure and mechanical prop-
Stress-state dependence erties were similar to those of long-fiber composites. In the ambient dry condition (10 wt.% water), the
Mechanical properties longitudinal elastic modulus, yield strength and failure strain were measured to be 4.0 GPa, 62 MPa
Structure–property relations and 4%, respectively, and the transverse elastic modulus, yield strength and failure strain were 2.9 GPa,
37 MPa and 2%, respectively. In the wet condition (35 wt.% water), horn behaves more like an isotropic
material; the elastic modulus, yield strength and failure strain were determined to be 0.6 GPa, 10 MPa
and 60%, respectively.
Ó 2010 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

1. Introduction to be catastrophic failure at the maximum stresses developed dur-


ing fighting. The remarkable resilience of horns makes them an
The horns of male bovids (e.g. goats, sheep, cattle, buffalo, bison excellent study if one is attempting to understand damage-deter-
and antelope) are mainly used in combat with other males to gain ring mechanisms and impact-resistant materials in nature.
access to females for mating [1–5] and can also be used for visual The keratin sheath of horn has a higher energy-absorbing ability
display, defense against predators and thermoregulation. Bighorn than bone or antler [7]. This allows horn sheaths to localize defor-
sheep (Ovis canadensis) can exert up to an estimated 3400 N of mation away from the immediate area of the brain and other cra-
force during a fight [6], which is more than any other sheep spe- nial organs in the ram upon impact. Based on finite-element
cies. Unlike antlers, horns are permanent structures and are highly analyses of goat skull impacts, Farke [8] hypothesized that the
susceptible to damage accumulation over a lifetime [2]. Conse- keratinous horn sheaths are one of the most important features
quently, horns must be tough and resistant to flaw propagation, contributing to the shock-absorbing ability of ram skulls.
which could promote stress concentration and weaken the horns Horns are made up of a sheath of keratin and a core of cancel-
[5]. Kitchener [6] concluded that a crack must be more than 60% lous bone [9]. There are also areas of compact bone and air-filled
of the transverse basal dimension of the horn in order for there sinuses within the frontal bones and horn cores. The keratin sheath
is the primary impact load-bearing material of the horn and is the
focus of this investigation. The horn sheath is a composite material
⇑ Corresponding author at: Department of Mechanical Engineering, Mississippi
comprised of tough, crystalline fibers made of a-keratin set in a
State University, Mississippi State, MS 39762, USA. Tel.: +1 662 325 8718; fax: +1
662 325 5433. compliant, amorphous keratin matrix [10,11]. The keratin fibers
E-mail address: wtrim@cavs.msstate.edu (M.W. Trim). serve to strengthen and stiffen the structure by forming long,

1742-7061/$ - see front matter Ó 2010 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.actbio.2010.11.024
M.W. Trim et al. / Acta Biomaterialia 7 (2011) 1228–1240 1229

hollow, fiber-like tubules. This dispersed tubule microstructure has sion under quasistatic loads of homogeneous aluminum alloys, and
been observed in other tough biological materials such as hoof, Tucker et al. [23] showed that large differences exist under high
bone, antler and dentin [12]. Keratin is also found in many tough strain rates as well. As such, the tension–compression asymmetry
biological materials such as skin, hair, horns and hooves. In a horn, would induce a different kind of stress (and strain) response than if
the keratin fibers are parallel to the growth direction and are the structure had equal and opposite tension and compression
stacked in a lamellar fashion through the thickness of the horn. stresses. These stress-state differences have not been examined
Horn keratin has a lamellar structure (2–5 lm in thickness) for horn keratin.
stacked in the radial direction with tubules (40  100 lm in In a recent study, Tombolato et al. [13] examined the micro-
diameter) dispersed between the lamellae. The tubules extend structure and bending and compressive properties at different ori-
along the length of the horn in the growth direction [13]. The tu- entations in order to study the failure mechanisms of O. canadensis
bules are randomly spaced in the transverse and radial direction, horn keratin. Corresponding studies have also been performed on
which leads to the material behavior in these two directions being similar keratin-based materials such as rhinoceros horn [17], bo-
nearly identical. Therefore, horn keratin is a transversely isotropic vine hoof [24] and equid hoof [25]. However, the various stress-
material, i.e. isotropic in the transverse and radial directions. state properties, and in particular the tensile response and
The mechanical properties of keratin are also highly dependent mechanical property gradients, of horn have not been examined.
on moisture content [11,14–16]. On a living animal, ‘fresh’ horn In addition, microindentation has never been performed on ram
keratin contains around 20 wt.% water, but if left to soak, horn ker- horn keratin. Microindentation is an accurate method for
atin can absorb up to about 40 wt.% water, depending on the sample determining the hardness and modulus of a material [26]. Microin-
thickness [11]. In keratin, water interacts only with the amorphous dentation also provides a means to investigate the property gradi-
matrix and not with the crystalline fibers [17]. Wet horn keratin is ents through the cross-sectional area in order to give insight into
less susceptible to damage, because the more compliant matrix can the hardness levels that could in turn be used for bio-inspired
more readily yield and flow [18,19]. In completely dry horn keratin, designs.
the stiffness of the matrix and fiber are assumed to be equal [10], The primary purpose of this paper is to quantify the structure–
and this stiffness is higher compared to wet horn keratin. However, property relations of horn keratin for use in modeling and simula-
a balance between the stiffness of dry horn and the compliance of tion. The structure–property data presented can be used for consti-
wet horn must exist for optimal performance of the horn, i.e. to tutive modeling in finite-element analysis in order to solve
maximize energy absorption and minimize damage accumulation, boundary value problems related to rams striking one another with
the stiffness of the horn must be modulated. their horns. Previous data are not amenable to calculating material
The anisotropy of fiber composites is well known. The behavior constants since constitutive model calibration typically requires
of the fibers in a composite subjected to compressive loads is anal- homogeneous stress-states, i.e. uniaxial tension and compression.
ogous to the behavior of columns on an elastic foundation. Thus, The parameters considered in this study that affect the stress–
the response of a composite to a compressive load is strongly strain behavior are the following: material orientation (transverse
dependent on matrix properties such as the shear stiffness. This vs. longitudinal), spatial location within the horn (base vs. middle
observation is different from the response of the composite to lon- vs. tip of horn), stress-state (tension vs. compression), and hydra-
gitudinal tensile loads, which is governed primarily by the fibers. tion level (wet vs. dry).
Because of the curvature of a ram’s horn, when an impact oc- In this work, we investigate the stress-state-dependent struc-
curs, a multiaxial stress-state arises. Historically, a von Mises ture–property relations for different orientations and moisture
[21] assumption would be employed in a finite-element analysis contents at various locations throughout a horn. As such, the
of this type of structure and impact, but the von Mises stress as- mechanical properties obtained via microindentation, tensile and
serts that compression and tension would give equal and opposite compressive testing are compared. Fracture surfaces are analyzed
values of the stress tensor. Recently, Dighe et al. [22] showed that and mechanical property gradients throughout the horn keratin
fairly large differences exist between tension, compression and tor- sheath are investigated.

Fig. 1. Schematic illustration of the mechanical testing specimen locations, dimensions, and orientations.
1230 M.W. Trim et al. / Acta Biomaterialia 7 (2011) 1228–1240

2. Materials and methods The same was done for the transverse specimens. Similarly, 20
dog-bone-shaped tensile specimens were taken from each region
Two well-preserved bighorn sheep (O. canadensis) horn sheaths, of the horn, 10 of which were longitudinally oriented and 10 trans-
approximately 1 m in longitudinal length and 12 cm in diameter at versely oriented. Of the 10 longitudinal and the 10 transverse ten-
the base, were obtained from Montana Fish, Wildlife and Parks. The sion specimens from each region, five of each were tested in the
rams were killed for reasons unrelated to this study. The specimens ‘‘wet’’ condition and five were tested in the ‘‘dry’’ condition, i.e.
were stored in a controlled environment with a temperature of each uniaxial tension and compression test was repeated five
18 °C and a relative humidity of 30% until needed. times. The resulting stress–strain curves for the duplicate tests
Compressive and tensile testing was performed on a universal were averaged together and the standard deviation at various
testing machine (EM Model 5869, Instron, MA, USA) equipped with strain levels was calculated. In this parametric study, 60 compres-
a 50 kN load cell. Three sets of specimens used for tensile and com- sion tests and 60 tension tests were performed in total. No speci-
pression testing were cut from the base, middle and tip of the horn mens were harvested from the region of the horns where growth
using a water-jet cutting machine. Care was taken to cut the spec- lines were obvious, as the growth lines could potentially affect
imens such that the fiber orientation was aligned either parallel or the mechanical properties.
perpendicular to the long axis of the specimen. The dog-bone ten- To investigate the effects of moisture content of the horn kera-
sile specimens had a length of 37 mm, a width of 18 mm, a gage tin, specimens were tested in both wet and ambient dry conditions.
length of 12 mm, a gage width of 6 mm and a thickness of 3 mm. Prior to testing, the test-pieces for the wet condition were soaked
The cylindrical compression specimens had a diameter of 3 mm in deionized water for 3 days, which has been determined to be a
and thickness of 3 mm. A constant strain rate of 3.0  103 s1 sufficient period for complete swelling to occur [11]. Subsequent
was maintained for all testing. Toughness values were calculated to testing, the wet specimens were weighed and placed in a
as the area under the average tensile stress–strain curves. 130 °C oven for 24 h. Once dried, the specimens were reweighed
The 20 cylindrical specimens were harvested from each region to determine the weight per cent of water that was present in
of the horn (base, middle and tip) for compression testing, 10 in the specimen at time of testing. The specimens for the dry condi-
the longitudinal direction, and 10 in the transverse direction. Of tion were allowed to acclimate to ambient humidity and tempera-
the 10 longitudinal specimens from each region, five were tested ture (roughly 50% RH and 20 °C). Each mechanical test was
in the ‘‘wet’’ condition and five were tested in the ‘‘dry’’ condition. repeated five times and the results were averaged together.

Fig. 2. Longitudinal and transverse hardness profile of ram horn keratin in wet (35 wt.% water) and dry (10 wt.% water) conditions.
M.W. Trim et al. / Acta Biomaterialia 7 (2011) 1228–1240 1231

The density of the horn keratin taken from the base, middle and W 1 /w
q¼ ð1Þ
tip of the horn was determined using Archimedes’ principle. Cylin- W 2  ðW 2  W w Þ
drical samples, 3 mm in diameter and 3 mm thick, were harvested
A micromechanical testing machine (TI 900 Triboindentor, Hys-
from the three horn regions. The dry weight, W1, of each sample
itron Inc., Minneapolis, MN, USA) equipped with a Berkovich
was obtained using a digital scale. The samples were then impreg-
indentor tip was used to determine hardness and elastic modulus
nated with oil and reweighed to obtain W2. The oil-impregnated
of the horn sheath material. An indentation profile was made
sample was then immersed in water of known density, /w , via a
across a polished cross-section of the horn. Care was taken to not
suspension wire with known mass, Ww, to obtain W3. The Archime-
probe any voids within the material. Spacing between indentations
des density was then calculated as follows:

Fig. 3. Longitudinal and transverse elastic modulus profile of ram horn keratin in wet (35 wt.% water) and dry (10 wt.% water) conditions.

Fig. 4. (a) SEM micrograph of a polished transverse section of ram horn keratin. (b) Image-Analyzer output showing 6.3% porosity.
1232 M.W. Trim et al. / Acta Biomaterialia 7 (2011) 1228–1240

Table 1
Average (n = 15) tensile longitudinal and transverse tensile mechanical properties of bighorn sheep horn keratin.

Specimen Elastic modulus (GPa) Yield strength (MPa) Ultimate strength (MPa) Failure strain (%) Toughness (MJ m3)
Dry
Longitudinal 3.9 ± 0.2 62.0 ± 6.9 77.3 ± 7.2 3.5 ± 0.5 2.0 ± 0.3
Transverse 2.8 ± 0.5 37.4 ± 9.1 44.9 ± 9.8 2.1 ± 0.8 0.6 ± 0.2
Wet
Longitudinal 0.7 ± 0.1 13.3 ± 1.1 27.4 ± 4.5 61.2 ± 2.1 11.7 ± 1.2
Transverse 0.5 ± 0.2 7.6 ± 2.3 21.1 ± 4.9 59.3 ± 3.9 8.5 ± 2.5

Table 2 coater (Polaron SC7640, Quorum Technologies Ltd., CT, USA) prior
Average (n = 15) compressive longitudinal and transverse mechanical properties of to observation in secondary electron (SE) mode at 5 kV. SEM
bighorn sheep horn keratin. images were analyzed using the Image-Analyzer software package
Specimen Elastic modulus Yield strength Yield strain developed by the Center for Advanced Vehicular Systems (CAVS) at
(GPa) (MPa) (%) Mississippi State University to quantify the microstructural fea-
Dry tures of the ram horn keratin material.
Longitudinal 2.2 ± 0.1 72.1 ± 5.4 3.4 ± 1.0
Transverse 1.9 ± 0.2 60.6 ± 12.8 3.1 ± 1.5
Wet
Longitudinal 0.20 ± 0.1 3.7 ± 0.3 4.1 ± 0.2 3. Results and discussion
Transverse 0.10 ± 0.01 4.1 ± 0.7 4.0 ± 0.2

Similar to many other biological tissues [27], horn keratin is a


hierarchical material. At the molecular level, horn keratin com-
was approximately 500 lm to avoid any strain hardening or resid- prises helical, a-keratin protofibrils. These protofibrils assemble
ual stress effects. The mechanical testing specimen locations, ori- into rope-like structures called intermediate filaments [28]. The
entations and dimensions are schematically summarized in Fig. 1. crystalline intermediate filaments are oriented along the growth
Fracture surfaces were examined by field emission scanning direction and coil up into hollow, elliptically shaped tubules, with
electron microscopy (SEM) with a JSM-6500F microscope (JEOL average minor and major diameters of 29.6 and 93 lm, respec-
Ltd., Tokyo, Japan) equipped for energy-dispersive spectroscopy tively. These tubules, which resemble hollow reinforcing fibers,
(EDS). Specimens were mounted on aluminum sample holders are embedded in an amorphous keratin matrix. The matrix is akin
and all surfaces not being examined were coated with silver paint. to a randomly oriented, chopped fiber composite. There is also a
All specimens received a 12.5 nm platinum coating in a sputter porosity gradient through the thickness of the horn, with the high-

Fig. 5. Average (n = 15) tensile longitudinal and transverse stress–strain response for horn keratin in wet (35 wt.% water) and dry (10 wt.% water) conditions.
M.W. Trim et al. / Acta Biomaterialia 7 (2011) 1228–1240 1233

Table 3
Comparison of longitudinal and transverse elastic moduli of bighorn sheep horn keratin obtained via tension, compression, microindentation and three-point bending in wet
(35 wt.% water) and dry (10 wt.% water) conditions.

Tensile test: forward Tensile test: Compressive test: forward Microindentation test: Three-point bending test: Ref.
loading (GPa) unloading (GPa) loading (GPa) unloading (GPa) forward loading (GPa)
Dry
a
Longitudinal 3.9 ± 0.2 4.0 ± 0.2 2.2 ± 0.1 3.44 ± 0.41 –
– – 1.64 ± 0.3 – 2.20 ± 0.2 [13]
a
Transverse 2.8 ± 0.5 2.9 ± 0.3 1.9 ± 0.2 3.29 ± 0.28 –
– – 1.94 ± 0.3 – 1.69 ± 0.5 [13]
Wet
a
Longitudinal 0.7 ± 0.1 0.5 ± 0.2 0.20 ± 0.1 0.67 ± 0.03 –
– – 0.53 ± 0.2 – 0.81 ± 0.4 [13]
a
Transverse 0.5 ± 0.2 0.4 ± 0.3 0.10 ± 0.1 0.62 ± 0.03 –
– – 0.25 ± 0.1 – 0.63 ± 0.2 [13]
a
This work.

est porosity being at the outer surface [13]. At the macroscale, a Fig. 4a is a SEM micrograph of a polished, transverse section of
horn takes the shape of a logarithmic or growth spiral. horn keratin and Fig. 4b is the output from Image-Analyzer. The tu-
Hardness and elastic modulus profiles obtained via microinden- bules resemble elliptical pores when viewed at this orientation.
tation are shown in Figs. 2 and 3, respectively. The hardness and The average minor axis length of the elliptical tubules is 29.6 lm
modulus increase slightly radially when traversing from the out- and the average major axis length is 93.0 lm. The aspect ratio, de-
side surface of the horn to the core. There is about a 40% increase fined as the major axis length over the minor axis length, averaged
in the elastic modulus and hardness from the outer surface of 3.14; the average porosity is 6.3%. These values are consistent with
the horn to the core of the keratin sheath. This indicates that there the results of Tombolato et al. [13], who estimated the minor and
is a microstructure gradient from the core to the outer surface of major axis lengths to be 40 and 100 lm, respectively; and the
the horn. There is no significant difference between the longitudi- porosity to be 7%.
nal and transverse moduli of the keratin sheath. All indentations
were performed on the amorphous keratin matrix. This is the rea- 3.1. Anisotropy of horn keratin
son that there was no difference between longitudinal and trans-
verse properties on the indentation profiles. Wet horn keratin When comparing the anisotropic behavior of the horn keratin,
shows considerably more compliance than dry horn keratin. The one can observe from the data (Tables 1 and 2; Fig. 5) that the lon-
optical micrographs of the indentation profiles show significant gitudinal direction is stiffer (higher elastic modulus), stronger
swelling of the matrix in the wet keratin. In the longitudinal micro- (higher yield and ultimate stress) and more ductile (higher elonga-
graphs, the lamellae and porosity due to the tubules are clearly tions to failure) than the transverse direction, regardless of hydra-
observed. tion level (wet or dry) or loading state (tension or compression).

Fig. 6. SEM fractographs of ram horn keratin specimens tested in tension (a) longitudinal dry, (b) longitudinal wet, (c) transverse dry and (d) transverse wet conditions.
1234 M.W. Trim et al. / Acta Biomaterialia 7 (2011) 1228–1240

Fig. 7. SEM micrograph of longitudinal, dry horn keratin specimen fractured in tension. The growth direction is out of the page. Loading was applied parallel to the growth
direction. Very little fiber pullout occurred, demonstrating a high degree of fiber–matrix adhesion.

These mechanical property characteristics are similar to those of the forward loading moduli measurements when compared to
synthetic long-fiber-reinforced composite materials [29], although moduli measurements taken during unloading. The strong agree-
in horn keratin, the tubules act as the crystalline reinforcement ment between the elastic moduli measurements taken during for-
and the matrix comprises randomly oriented, chopped keratin fi- ward loading and during unloading indicates that machine
bers. When the tubules are oriented perpendicular to the loading compliance did not affect the tensile test results.
direction, they tend to produce stress concentrations at the inter- The longitudinal and transverse moduli values ranged from 0.3
face and within the matrix. As such, fiber composites subjected to 0.8 GPa for the wet horn keratin. The anisotropy of the elastic
to transverse tensile loads fail because of matrix cracking or inter- modulus decreased significantly with increasing moisture content
face debonding similar to their synthetic counterparts. Although of the horn keratin. The strong similarity between the ‘‘wet’’
the qualitative characteristics, and in particular the failure strain, mechanical properties in the longitudinal and transverse directions
of the dry horn keratin are similar to those of graphite–epoxy suggests that hydration severely degraded the matrix phase, which
long-fiber composites [29], the values for elastic modulus and ulti- led to a matrix-dominated deformation behavior. The hydrated,
mate strengths are approximately two orders of magnitude weaker amorphous matrix gives a much more isotropic response [14,19].
for the horn keratin. Table 3 also shows that in the dry condition, the tensile testing
When examining the different methods of quantifying the elas- resulted in a higher longitudinal elastic modulus than the microin-
tic moduli of the horn keratin, the microindentation and tensile dentation testing. This attests to a stiffening effect of the tubules,
loading and unloading moduli were generally larger than the for- since microindentation testing only probes the properties of the
ward compression and bending test results as summarized in Ta- randomly oriented keratin fiber matrix. Furthermore, the ambient
ble 3, which also includes data from Tombolato et al. [13]. For dry, transverse elastic modulus from microindentation is higher
example, the longitudinal, dry values ranged from 3.0 to 4.1 GPa than the elastic modulus obtained via tensile testing. In the horn
for the indentation and tensile unloading data, and from 1.3 to sheath, the tubules tend to debond from matrix under tensile load-
2.4 GPa for the compression and bending data. Similarly, for the ing. These trends, however, are much less noticeable in the wet
transverse dry horn keratin, the values ranged from 2.3 to horn keratin.
3.6 GPa for the indentation and tensile tests and from 1.6 to In terms of the anisotropic behavior of the failure strains, the
2.2 GPa for the compression and bending tests. Because of the com- dry, longitudinal compressive strength was only slightly higher
pliance in testing machines, lower values are usually expected for than the transverse compressive strength, indicating very good fi-

Fig. 8. SEM fractographs of transverse, wet horn keratin specimen fractured in tension. Loading was applied perpendicular to growth direction. Failure occurred
predominantly because of matrix failure, with some transverse fiber pullout.
M.W. Trim et al. / Acta Biomaterialia 7 (2011) 1228–1240 1235

Fig. 9. SEM fractographs of transverse, dry horn keratin specimen fractured in tension. Loading was applied perpendicular to growth direction. Failure exhibited delamination
and fiber fracture.

ber–matrix adhesion, as expected due to the strong chemical of synthetic composites, because of the greater failure strains and
bonds between fibers and matrix of identical composition. Water associated fairly large fracture toughness, as illustrated in Fig. 5.
apparently strengthens this bonding even more, which is evi- This flexibility of the horn moving back and forth between a stiffer
denced by the nearly identical longitudinal and transverse com- and a more flexible structure is not available for the synthetic long-
pressive strengths for wet horn keratin. fiber composites.
Related to the wet and dry specimen fracture surfaces, SEM
3.2. Wet vs. dry horn keratin images of tensile specimens revealed different failure phenomena
between the longitudinal and transverse conditions; examples
Although qualitative similarities exist between synthetic long- are illustrated in Fig. 6. The nearly flat fracture surface of the dry,
fiber composites and dry horn keratin in terms of anisotropy, the longitudinal specimen was characterized by brittle fracture. Very
qualitative character of wet horn keratin is very different to that little tubule pullout was observed, demonstrating a high degree

Fig. 10. Average (n = 5) longitudinal and transverse tensile stress–strain response for horn keratin. Samples located in the base, middle and tip regions of the horn in wet
(35 wt.% water) and ambient dry (10 wt.% water) conditions.
1236 M.W. Trim et al. / Acta Biomaterialia 7 (2011) 1228–1240

of tubule–matrix adhesion, as seen in Fig. 7. However, the wet, lon- 3.3. Location dependence within the horn
gitudinal fracture surface showed an extremely ductile fracture
mode, evidenced by a very deep, convoluted cup-and-cone type It may be expected that strength and stiffness differences exist
fracture. Significant necking was also observed on the wet, longitu- along the length of horn due to aging, with the material near the
dinal specimens, though the specimen had fully recovered to its tip of the horn being older than the material at the base of the horn
original shape and dimensions by the time of imaging. For the due to new horn being laid down as the animal grows larger.
transverse loading specimens, both the wet and dry specimens Figs. 10–12 show the stress–strain behavior of horn keratin at dif-
exhibited a shear-type failure mode. Tubules perpendicular to ferent locations in order to examine the anisotropy between the
the loading direction acted essentially to produce stress concentra- longitudinal and transverse directions, stress-state dependence be-
tions at the interface and in the matrix. When subjected to trans- tween tension and compression, and the wet–dry conditions. In
verse tensile loads, wet horn keratin failed predominantly particular, Fig. 10 shows the average (n = 5) tensile longitudinal
because of matrix failure (around the tubules), with some trans- and transverse tensile stress–strain behavior for horn keratin sam-
verse tubule pullout, as seen in Fig. 8. However, ambient dry horn ples located in the base, middle and tip regions in wet (35 wt.%
keratin exhibited delamination and tubule fracture, as seen in water) and ambient dry (10 wt.% water) conditions. Taking into ac-
Fig. 9. count the standard deviation associated with the five specimens
The larger failure strains and fracture toughness under the wet that were tested in each condition, it appears that the location
conditions indicate that wet horn keratin is more resilient than dry did not have a significant effect on the stress–strain behavior since
horn keratin, i.e. the wet horn keratin material can elastically store the response from the base, middle and tip of the horn were highly
more energy per unit volume than dry horn keratin. Under the wet consistent, with the average curve for each parametric test falling
conditions, the energy-absorbing capability is much greater than in within the bounds of deviation for the other tests. We suspect that
the dry conditions by approximately 5–10 times. Bovids display an with a larger sample population, these results would further con-
interesting behavior known as horning, where they frequently rub verge. As a note related to the anisotropic and wet–dry discussion
their horns in mud and against wet vegetation prior to fighting [5]. earlier, the mechanical properties were calculated by averaging the
The animals that keep the keratinous sheath of their horns ade- results from the base, middle and tip regions because it is not pos-
quately hydrated maintain toughness and notch-insensitivity in sible to distinguish the stress–strain behavior.
their horns against the desiccating environment. This lessens the Figs. 11 and 12 show the average (n = 5) longitudinal and trans-
probability of the animal sustaining an injury. verse compressive stress–strain behavior for horn keratin samples

Fig. 11. Average (n = 5) longitudinal and transverse compressive stress–strain response for horn keratin samples located in the base, middle and tip regions in the dry
(10 wt.% water) condition.
M.W. Trim et al. / Acta Biomaterialia 7 (2011) 1228–1240 1237

Fig. 12. Average (n = 5) longitudinal and transverse compressive stress–strain response for horn keratin samples located in the base, middle and tip regions in the wet
(35 wt.% water) condition.

The density of the horn keratin material as measured at the


base, middle and tip was determined to be 1.238, 1.237 and
1.237 g cm3, respectively. This indicates that similar to the tensile
and compressive stress–strain behavior, density also does not vary
significantly along the length of the horn. For comparison, KevlarÒ,
a lightweight, fiber reinforcement in composite materials that is
widely used for penetration resistance, has a density of
1.439 g cm3, and polycarbonate, a common polymer known for
its high impact resistance, has a density of about 1.21 g cm3 [30].

3.4. Stress-state dependence of horn keratin

When comparing the stress-state dependence of the horn kera-


tin, we see very different material responses, particularly when we
compare the tension and compression behaviors. Fig. 14a illus-
trates that for the dry horn keratin, the stress–strain behavior up
to approximately 1% strain, and the compression and tension
curves, are similar, but deviate beyond 3% strain. Another differ-
Fig. 13. Average (n = 15) compressive longitudinal and transverse stress–strain
response for horn keratin in wet (35 wt.% water) and dry (10 wt.% water)
ence between tension and compression for the dry horn keratin
conditions. is that the strain to failure is much greater for compression extend-
ing up to 20% before failure. Interestingly, the longitudinal stress–
strain behavior was greater than the transverse behavior under
located in the base, middle and tip regions in the wet and dry con- tension, but under compression the opposite behavior was realized
dition. Similar to the tension loading case, the compressive stress– later in strain (after 10% strain). A similar behavior is observed for
strain behavior did not vary greatly with location. Moreover, the wet horn keratin, as seen in Fig. 14b. In addition, for the wet
Fig. 13 shows the comprehensive average (n = 15) of the longitudi- horn keratin stress–strain behavior, one can observe that the ten-
nal and transverse compressive stress–strain behaviors of all three sion curves are higher than the compression curves for the both
locations in the wet and ambient dry conditions. the longitudinal and transverse directions.
1238 M.W. Trim et al. / Acta Biomaterialia 7 (2011) 1228–1240

When one compares the failure mechanisms under tension and


compression, different modes arise. Figs. 6 and 15 show the frac-
ture surfaces under tension and compression, respectively. As
shown in Fig. 15, the buckling of the lamellae under compression
effectively decreases the ‘‘work hardening’’ in the stress–strain
behavior when compared to the tensile loading. However, the ten-
sion failure strains are much lower than that of compression, even
when the tensile failure strains are increased because of water. The
damage progression under compression from the longitudinal frac-
ture specimens shown in Fig. 15a and b illustrate shear microbuck-
ling followed by delamination. This is quite different to the tensile
fracture behavior, which was observed to be fiber failure in the dry
specimen and ductile necking in the wet specimen. In the trans-
verse direction (Fig. 14c and d), both the tension and compression
loading fracture surfaces exhibited shear failure along the direction
of maximum shear stress (45° to loading axis) with some delami-
nation experienced in the compression specimens.
A blow to the horn during head-butting induces a local com-
pression loading condition, and hence extending the energy-
absorbing capability in compression is key. With the microbuck-
ling observed in compression, the energy absorption is extended
beyond what otherwise could be realized. This same microbuckling
of lamellae to extend the energy absorption was also observed in
turtle shells [31] and abalone nacre [32]. This microbuckling in
compression to increase energy absorption has also been observed
in synthetic long-fiber composites [33,34] and metal foams [35].
The compressive stress–strain behavior of the ram horn keratin,
as seen in Fig. 13, consists of three regimes: a linear elasticity re-
gion, a long collapse plateau, and finally densification. This type
of compressive stress–strain behavior is characteristic of synthetic
foams but has also been observed in biological materials such as
turtle shells [31] and bones [36].
Not only is the difference between tension and compression
important to the mechanical response of the spiraled ram horn,
but the unloading under each loading condition is also significant.
The unloading can reveal how much of the forward loading is elas-
tic, visocoelastic or viscoplastic. Furthermore, the unloading can
reveal more accurate elastic moduli measurements because of
Fig. 14. Stress-state and material orientation dependence of ram horn keratin for in the absence of machine compliance (the machine effect is very
(a) the ambient dry and (b) the wet condition. small, considering the elastic modulus of the horn keratin is

Fig. 15. SEM micrographs of compressive (a) longitudinal dry, (b) longitudinal wet, (c) transverse dry and (d) transverse wet horn keratin fracture specimens.
M.W. Trim et al. / Acta Biomaterialia 7 (2011) 1228–1240 1239

the stiffness and strength required for optimized impact resistance.


Given the function of ram horn, a structure–property parametric
study of bighorn sheep horn keratin was performed in order to
quantify the influence of several factors believed to potentially af-
fect the structure–property relations of the horn. These factors in-
cluded analysis of the stress-state dependence with the horn
keratin tested under tension and compression, the anisotropy of
the material structure and mechanical behavior, the spatial loca-
tion of the horn, and the wet–dry horn behavior.
The following conclusions can be made:

 Horn keratin behaves in an anisotropic manner similar to a


long-fiber composite with strengthening fibers in a matrix in
terms of the elastic moduli, strengths, and failure strains and
mechanisms. However, the anisotropy is reduced as water is
content is increased.
 The tubules serve to longitudinally stiffen the horn in tension
and absorb energy in transverse compression.
 Water dominates the horn keratin material behavior more than
the anisotropy, location in the horn and the type of loading
Fig. 16. Longitudinal and transverse tensile stress–strain response with unloading
state. This makes moisture content the most relevant parameter
for horn keratin in wet (35 wt.% water) and dry (10 wt.% water) conditions, showing
the inelastic strains for (a) dry horn keratin, (b) transverse–wet, and (c) longitu- in regards to influencing the mechanical behavior of horn
dinal–wet horn keratin. keratin.
 A clear tension–compression asymmetry exists within the horn
in which the tension stress–strain behavior exhibits a greater
2 GPa compared with that of steel, 200 GPa). Fig. 16 shows ten- initial modulus that is exacerbated in the wet state. This early
sion and compression forward loading and unloading stress–strain higher modulus in the tension curve leads to higher stress-
data to illustrate the non-monotonicity of the material behavior. states as a function of strain, and eventual fracture, sooner than
For both the wet and dry conditions, irreversible strains were evi- in compression.
denced with more exhibited for the wet condition. Because both  Tensile failure in the longitudinal direction occurred by matrix
reversible and irreversible strains arise from the loading–unload- separation followed by fracture of the reinforcing tubules and
ing sequence, both elasticity and inelasticity have occurred upon some tubule pull-out. The ambient dry horn keratin failed in a
the forward loading. From Fig. 16, one can observe that in the much more brittle manner, while wet horn keratin was much
dry condition, for a total tensile strain of 4%, 2% of the strain was more ductile.
elastic and 2% was inelastic. Hence, 50% of the total deformation  Tensile failure in the transverse direction occurred in wet horn
was inelastic. For the wet conditions up to a tensile strain of 20%, keratin primarily because of matrix failure, with some trans-
the elastic strain ranged from 5% to 7.5% and the inelastic strain verse fiber pullout. However, ambient dry horn keratin exhib-
ranged from 12.5% to 15% for the longitudinal and transverse direc- ited delamination and tubule fracture.
tions. Here, 60–75% of the deformation was inelastic. The inelastic-  Compressive failure in the longitudinal direction occurred by
ity exhibited within the horn keratin under tension arose from the shear microbuckling followed by delamination in both the
viscoplasticity of entangled molecular chains inducing friction wet and ambient dry conditions.
upon loading and also from the permanent damage that arose  Compressive failure in the transverse direction, for both the wet
within the material near the failure strain, which finally led to and dry specimens, exhibited a shear-type failure mode.
fracture.
We also note that Fig. 16 was used for the forward and unload-
ing determination of the elastic modulus according to our earlier Acknowledgements
discussion. As mentioned, the slope of the unloading curve gener-
ally provides a more accurate determination of elastic modulus This work was supported by the Center for Advanced Vehicular
since compliance of the testing machine is not a factor. The average Systems (CAVS) at Mississippi State University. This material is
longitudinal and transverse moduli obtained via unloading were based upon work supported by the Department of Energy, South-
measured as 4.0 and 2.9 GPa, respectively, for the dry horn keratin. ern Regional Center for Lightweight Innovative Design (SRCLID),
For the wet horn keratin, the longitudinal modulus was 0.5 GPa under award number DE-EE0002323. This report was prepared as
and the transverse modulus was 0.4 GPa. These values are in strong an account of work sponsored by an agency of the United States
agreement with the moduli obtained via forward loading, indicat- Government. Neither the United States Government nor any
ing that the compliance of the machine did not affect the agency thereof, nor any of their employees, makes any warranty,
measurements. expressed or implied, or assumes any legal liability or responsibil-
ity for the accuracy, completeness, or usefulness of any informa-
4. Conclusions tion, apparatus, product, or process disclosed, or represents that
its use would not infringe privately owned rights. Reference herein
The horn keratin of bighorn sheep (O. canadensis) is adapted for to any specific commercial product, process, or service by trade
excellent stiffness and strength under impact loading. When male name, trademark, manufacturer, or otherwise does not necessarily
bighorn sheep fight, their head-butts induce the largest impact constitute or imply its endorsement, recommendation, or favoring
forces of any ram. The combination of the horn spiral, the lessening by the United States Government or any agency thereof. The views
of the mean diameter as the horn length is extended from the skull, and opinions of authors expressed herein do not necessarily state
and the gradients of microstructure and associated mechanical or reflect those of the United States Government or any agency
properties from the horn’s center to the outer radius all point to thereof. We also appreciate the generosity of Mr. Neil Anderson
1240 M.W. Trim et al. / Acta Biomaterialia 7 (2011) 1228–1240

(Montana, Fish, Wildlife and Parks), who kindly provided our test [15] Bertram JEA, Gosline JM. Functional design of horse hoof keratin: the
modulation of mechanical properties through hydration effects. J Exp Biol
specimens. We also gratefully acknowledge American Eurocopter,
1987;130:121.
Columbus, MS for allowing us to use their water-jet cutter. Profes- [16] Fraser RDB, MacRae TP, Rogers GE. Keratins. Springfield, IL: Bannerstone
sor McKittrick gratefully acknowledges support from the National House; 1972.
Science Foundation, Biomaterials Program, Grant 0510138. [17] Druhala M, Feughelman M. Dynamic mechanical loss in keratin at low
temperatures. Colloid Polym Sci 1974;252:381.
[18] Vincent J. Structural biomaterials. Princeton, NJ: Princeton University Press;
Appendix A. Figures with essential colour discrimination 1990.
[19] Kitchener A. Effect of water on the linear viscoelasticity of horn sheath keratin.
J Mater Sci Lett 1987;6:321.
Certain figures in this article, particularly Figs. 1–5, 10–14 and [21] Von Mises R. Mechanik der festen Körper im plastisch deformablen Zustand.
16, are difficult to interpret in black and white. The full colour Gottingen Nachr Math Phys 1913;4:582.
[22] Dighe MD, Gokhale AM, Horstemeyer MF. Effect of loading condition and
images can be found in the on-line version, at doi:10.1016/
stress state on damage evolution of silicon particles in an Al–Si–Mg–base cast
j.actbio.2010.11.024. alloy. Metall Mater Trans A 2002;33:555.
[23] Tucker MT, Horstemeyer MF, Gullett PM, El Kadiri H, Whittington WR.
Anisotropic effects on the strain rate dependence of a wrought magnesium
References
alloy. Scr Mater 2009;60:182.
[24] Zhang D, Arola D, Reprogel R, Zhang W, Tasch U, Dyer R. A method for
[1] Geist V. Mountain sheep: a study in behavior and evolution. Chicago: Chicago characterizing the mechanical behaviour of hoof horn. J Mater Sci
University Press; 1971. 2007;42:1108.
[2] Goss RJ. Deer antlers—regeneration, function and evolution. New [25] Kasapi MA, Gosline JM. Micromechanics of the equine hoof wall: optimizing
York: Academic Press; 1983. crack control and material stiffness through modulation of the properties of
[3] Leuthold W. African ungulates: a comparative review of their ethology and keratin. J Exp Biol 1999;202:377.
behavioral ecology. Cambridge: Cambridge University Press; 1977. [26] Oliver WC, Pharr GM. An improved technique for determining hardness and
[4] Schaller GB. Mountain monarchs. Chicago: Chicago University Press; 1977. elastic modulus using load and displacement sensing indentation
[5] Kitchener A. Fracture toughness of horns and a reinterpretation of the horning experiments. J Mater Res 1992;7:1564.
behavior of bovids. J Zoolog Lond 1987;213:621. [27] Fratzl P, Weinkamer R. Nature’s hierarchical materials. Prog Mater Sci
[6] Kitchener A. An analysis of the forces of fighting of the blackbuck (Antilope 2007;52:1263.
cervicapra) and the bighorn sheep (Ovis canadensis) and the mechanical design [28] Fraser RD, MacRae TP, Parry DA, Suzuki E. Intermediate filaments in alpha-
of the horns of bovids. J Zoolog Lond 1988;214:1. keratins. Proc Natl Acad Sci 1986;83:1179.
[7] Kitchener A. The evolution and mechanical design of horns and antlers. In: [29] Johnston NJ. Toughened composites. Houston, TX: ASTM STP 937; 1987.
Rayner JMV, Wootton RJ, editors. Biomechanics in [30] Callister WD. Materials science and engineering: an introduction. New
evolution. Cambridge: Cambridge University Press; 1991. p. 229. York: John Wiley and Sons; 2007.
[8] Farke AA. Frontal sinuses and head-butting in goats: a finite element analysis. J [31] Rhee H, Horstemeyer MF, Hwang Y, Lim H, El Kadiri H, Trim W. A study on the
Exp Biol 2008;211:3085. structure and mechanical behavior of the Terrapene carolina carapace: a
[9] Packer C. Sexual dimorphism: the horns of African antelopes. Science pathway to design bio-inspired synthetic composites. Mater Sci Eng: C
1983;221:1191. 2009;29:2333.
[10] Fraser RDB, MacRae TP. Molecular structure and mechanical properties of [32] Menig R, Meyers MH, Meyers MA, Vecchio KS. Quasi-static and dynamic
keratins. In: Vincent JF, Currey JD, editors. The mechanical properties of mechanical response of Haliotis rufescens (abalone) shells. Acta Mater
biological materials. Cambridge: Cambridge University Press; 1980. p. 211. 2000;48:2383.
[11] Kitchener A, Vincent JFV. Composite theory and the effect of water on the [33] Sutcliffe MPF, Fleck NA. Microbuckle propagation in fibre composites. Acta
stiffness of horn keratin. J Mater Sci 1987;22:1385. Materialia 1997;45:921.
[12] McKittrick J, Chen PY, Tombolato L, Novitskaya EE, Trim MW, Hirata GA, et al. [34] Fleck NA, Sridhar I. End compression of sandwich columns. Compos Part A:
Energy absorbent natural materials and bioinspired design strategies: a Appl Sci Manuf 2002;33:353.
review. Mater Sci Eng: C 2010;30:331. [35] Gibson LJ, Ashby MF. Cellular solids: structure and
[13] Tombolato L, Novitskaya EE, Chen P-Y, Sheppard FA, McKittrick J. properties. Cambridge: Cambridge University Press; 1999.
Microstructure, elastic properties and deformation mechanisms of horn [36] Currey JD. Bones: structure and mechanics. Princeton, NJ: Princeton University
keratin. Acta Biomater 2010;6:319. Press; 2002.
[14] Feughelman M. Mechanical properties and structure of alpha-keratin
fibres. Sydney: University of New South Wales Press; 1997.

View publication stats

You might also like