You are on page 1of 11

Analysis of an ultra hard

magnetic biomineral in
chiton radular teeth
Recent analyses of the ultrastructural and mechanical properties of
mineralized biological materials have demonstrated some common
architectural features that can help explain their observed damage
tolerance. Nature has accomplished this feat through the precise control
of anisotropic crystal nucleation and growth processes in conjunction
with nanoscale control over the self-assembly of spatially distinct
organic and inorganic phases, resulting in effective inhibition of crack
propagation through these materials. One such example is found in the
hyper-mineralized and abrasion resistant radular teeth of the chitons, a
group of herbivorous marine mollusks who have the surprising capacity
to erode away the rocky substrates on which they graze1-4. Through
the use of modern microscopy and nanomechanical characterization
techniques, we describe the architectural and mechanical properties of
the radular teeth from Cryptochiton stelleri. Chiton teeth are shown to
exhibit the largest hardness and stiffness of any biominerals reported
to date, being notably as much as three-fold harder than human enamel
and the calcium carbonate-based shells of mollusks. We explain how the
unique multi-phasic design of these materials contributes not only to
their functionality, but also highlights some interesting design principles
that might be applied to the fabrication of synthetic composites.
James C. Weavera, Qianqian Wanga, Ali Miserezb, Anthony Tantuccioc, Ryan Strombergd, Krassimir N. Bozhilove, Peter Maxwellg,
Richard Nayd, Shinobu T. Heierf, Elaine DiMasih, David Kisailusa,*
aDepartment of Chemical and Environmental Engineering, University of California, Riverside, CA 92521 (USA)
bMaterials Department and Marine Science Institute, University of California, Santa Barbara, Santa Barbara, CA-93106-5050 (USA)

Now at: School of Materials Science and Engineering, Nanyang Technological University, 50 Nanyang Avenue, Singapore 639798
cDepartment of Chemical Engineering, The Cooper Union for the Advancement of Science and Art, New York, NY 10003-7120, USA
dNanomechanics Research Laboratory, Hysitron, Inc., 10025 Valley View Road, Minneapolis, MN 55344 (USA)
eCentral Facility for Advanced Microscopy and Microanalysis, University of California, Riverside, CA 92521
fThermo Fisher Scientific, 1400 Northpoint Parkway, Suite 10, West Palm Beach, Florida 33407
gMaterials Department, University of California, Santa Barbara, Santa Barbara, CA-93106-5050 (USA)
hNational Synchrotron Light Source, Bldg 725D, Brookhaven National Laboratory, Upton NY 11973

*E-mail: david@engr.ucr.edu

42 JAN–FEB 2010 | VOLUME 13 | NUMBER 1-2 ISSN:1369 7021 © Elsevier Ltd 2010
Analysis of an ultra hard magnetic biomineral in chiton radular teeth INSIGHT

Ever since early humans first used antler to shape stone tools These biominerals display similar, and frequently superior,
via lithic reduction, we have been fascinated by the structural mechanical robustness to those exhibited by many engineering
complexity and mechanical robustness of biominerals. The materials with similar chemical composition, such as structural
19th century saw an explosion in the interest to catalog global ceramics. Biological systems have accomplished this feat through the
biodiversity and through the course of these expeditions, many demonstrated ability to control size, morphology, crystallinity, phase,
new species were discovered and the details of their skeletal and orientation of the mineral under benign processing conditions
systems were documented in exquisite detail (Fig. 1). It was these (i.e., near-neutral pH, room temperature, etc.). They utilize organic-
detailed ultrastructural studies5-7 that helped lay the groundwork inorganic interactions and carefully controlled microenvironments that
for the modern field of biomineralization2. In more recent times, enable kinetic control during the synthesis of inorganic structures4.
the availability of advanced instrumentation such as synchrotron In these biomineralized systems, minerals and organic
X-ray diffraction and high resolution transmission electron and macromolecules exist in close proximity and at nanoscale dimensions.
scanning probe microscopy in conjunction with the latest advances Interactions at these interfaces are vital to the functions of structural
in genomics and proteomics have permitted significant progress materials found in nature such as shells, teeth, and bone4. Although
into not only the nanoscale characterization of these materials, the organic constituents of these biological composite materials
but also the factors regulating their controlled fabrication. are present in relatively small quantities 4, they significantly alter

(a) (b) (c)

(d) (e) (f)

(g) (h) (i)

Fig. 1 Evolution of the field of biomineralization through time.19th Century: Gross morphological skeletal anatomy of Caryophyllia profunda (a), Cassis sp. (b),
and Euplectella aspergillum (c) collected during the Challenger expedition between 1873 and 1876. 20th Century: Electron microscopy studies of the aragonitic
spherulites from Aphrocallistes vastus (d), mineralized tablets of nacre from Haliotis rufescens (e), and fused hexactine siliceous spicules from Aphrocallistes vastus
(f). 21st Century: TEM image of a focused-ion beam milled sclerite from Corallium rubrum (g), High resolution TEM and associated electron diffraction pattern of a
single nacre tablet from Haliotis rufescens (h), and atomic force micrograph of the concentric lamellae of consolidated silica nanoparticles of a spicule from Tethya
aurantia (i). (a-c) adapted from5-7 and (g) adapted from8.

JAN–FEB 2010 | VOLUME 13 | NUMBER 1-2 43


INSIGHT Analysis of an ultra hard magnetic biomineral in chiton radular teeth

the mechanical behavior of the bulk structures. In organic-inorganic by Ricketts and Calvin3 who wrote, “After one experiment the writers
composites, the existence of the organic phase leads to significant decided to reserve the animals for times of famine; one tough, paper-
energy dissipation at the interfaces during loading, resulting in a thin steak was all that could be obtained from a large Cryptochiton,
combination of properties that may well improve the abrasion and and it radiated such a penetrating fishy odor that it was discarded
wear resistance of the structure in comparison to monolithic materials before it reached the frying pan.”
with equivalent chemical composition. One such example are the New radular teeth are continually secreted and shaped inside a
hypermineralized teeth of chitons, which are shown to be harder and specialized secretory organ, the radular sack, where they are fused
stiffer than any other known biomineral as discussed in detail below. with the underlying radular cuticle that maintains tooth alignment
The chitons (Mollusca, Polyplacophora; Fig. 2a) are an ancient with respect to one another. Slow migration of the cuticle strip
group of mollusks with a fossil record dating back nearly half a billion brings the newly formed teeth forward until they are in a functional
years. Despite their long and successful history and their ecological position, at the anterior-most region of the radula2. Because every
importance in rocky coastal habitats1 they are a comparatively small stage of tooth development is contained on a single radula, it
group with about 650 modern species. Chitons are flattened and represents an ideal model system for investigating the dynamic
usually elongated mollusks that are protected dorsally by a shell processes of biomineralization2, 9-12. Building on the pioneering
consisting of eight overlapping plates. The foot is broad and powerful, work of Lowenstam and colleagues in the 1960s4 our present
well adapted for clinging tightly to the hard surfaces on which the study reports a detailed investigation of the unique properties of
animal grazes for algae. Like most other groups of mollusks, the the radular teeth of C. stelleri. Lowenstam’s initial investigations
chitons have a radula (Fig. 2b, c), a rasping, toothed conveyor belt-like into the architecture of the radular teeth from C. stelleri revealed
structure, which is used for feeding. The composition and morphology that the tricuspid mineralized cap is a multi-component system
of the radular teeth vary from group to group and depend to a large consisting predominantly of an amorphous phosphatic core covered
extent on the dietary specifics and the mechanical properties of the with a hard veneer of magnetite. Although mineralogical analyses
substrates on which they feed1-3. by Raman and EDS have been performed previously at various
Cryptochiton stelleri is native to the temperate Northern Pacific locations on radular tooth cross-sections12, detailed elemental and
and is the world’s largest species of chiton, reaching a maximum mechanical maps during the tooth maturation process have not been
length of up to 33 cm1-3, thus making it an ideal research system for reported.
investigating radular tooth architecture and mechanics. Despite its In this work, we present a more detailed understanding of the
large size, the meat of this mollusk has no commercial value and is mineralized radular teeth via nano-mechanical and elemental mapping
generally not considered palatable. This is exemplified in an account in order to provide insight into the important structural gradients that

(a) (c) (e)

(f)

(b) (d)

Fig. 2 Morphological features of the chiton radula. External (a) and internal (b) anatomy of a representative chiton showing the location of the radula, a rasping,
toothed conveyor belt-like structure used for feeding. Details of the anterior region of the radula from C. stelleri (c-f). Optical (c) and backscattered SEM (d)
imaging and X-ray transmission studies (e) reveal the nature of the electron density distribution of the tricuspid tooth caps. Cross-sectional studies through the
mature teeth from C. stelleri (f1,2) reveal a concentric biphasic structure. (a) adapted from14.

44 JAN–FEB 2010 | VOLUME 13 | NUMBER 1-2


Analysis of an ultra hard magnetic biomineral in chiton radular teeth INSIGHT

contribute to the unique impact and abrasion-resistant properties of 125 GPa and a corresponding hardness ranging from 9 to 12 GPa. To
this material. the best of our knowledge, these values represent the highest modulus
yet reported for a biomineral22. The hardness is notably about 3
Results and discussion: times higher than that of enamel and nacre, which exhibit indentation
The mature radular teeth are each composed of a mineralized tricuspid hardness and modulus of 3 – 4 GPa and 65 – 75 Gpa22, respectively,
cap with a stalk-like flexible attachment to the base of the radula making this material exceptionally well suited for the continuous
(Fig. 2d-f). When examined by optical microscopy, the second row scraping activity of the radular teeth. In contrast, the weakly
of lateral teeth appears black and lustrous (Fig. 2c), and synchrotron crystalline core region has a modulus of ca. 25 GPa and a hardness
X-ray transmission studies reveal the nature of the electron density of ca. 2 GPa. Mechanical mapping of cross-sections through these
distribution of this material (Fig. 2e). These observations are also two regions of the teeth reveals a distinct gradient in mechanical
supported by backscattered electron microscopy studies13 where the properties with the modulus of the leading edge of the tooth ca. 15%
tooth caps are significantly brighter compared to the background higher than that on the trailing edge. This design strategy results
material of the stalk and basal supporting membrane (Fig. 2d,f). in an uneven wear pattern along the scrapping edge of the tooth
Cross-sectional examinations by scanning electron microscopy and establishes a self-sharpening condition (Fig. 4b), an observation
(Fig. 2f1, 2) and energy dispersive spectroscopy (EDS) of the radular consistent with radula structural studies on other species23.
teeth of C. stelleri reveal that they are composed of two distinct mineral Indentation fracture studies of the magnetite veneer reveal that
phases2,4 (Fig. 3a). The core region of the teeth is enriched in iron cracks propagating through this material usually travel parallel to
phosphate and the exterior of the teeth in iron oxide (Fig. 3a,b). EDS the long axis of the tooth rather than perpendicular to it (Fig. 5b).
analysis also reveals a significantly higher C content in the tooth core In contrast, cracks propagating through the tooth core are largely
as well as small amounts of Ca, K, Na, Mg, and Si (Fig. 3b). X-ray and isotropic and lack any defined directionality (Fig. 5a). In addition, as
electron diffraction (Figs. 3c; 5d) of the tooth exterior region reveal a propagating crack passes across the boundary between the core
that it is composed of ferromagnetic iron oxide (magnetite)2,4,15, material and the outer magnetite shell, significant crack deflection
while EDS, Raman spectroscopy11,12,16 and dark-field transmission occurs at this interface (Fig. 5a) because of the stress redistribution
electron microscopy of the tooth core material suggest the presence occurring either from small-scale yielding of the thin organic layers,
of weakly crystalline hydrated iron phosphate (Figs. 3b,d; 5e)2,4. The or from debonding at the organic/mineral interface, with the latter
iron phosphate rich core region of the radular teeth exhibits two mechanism being more efficient in protecting the uncracked material
intense Raman bands at 1050 and 1010 cm-1 along with weak bands at across the interface24 (Fig. 5a). This strategy of crack deflection is very
1110 cm-1 and in the range of 400 – 700 cm-1 (Fig. 3d). It is generally effective at maintaining the tooth structural integrity and preventing
accepted that the stretching and bending vibrations of phosphate catastrophic failure of the material. Examination of fractured tooth
groups occur at 1000 – 1200 and 400 – 700 cm-1, respectively. The cross-sections reveals how the organization of the magnetite
Raman peaks of crystalline materials are typically sharp and well crystallites facilitates this mechanical response (Fig. 5c).
separated in the regions of 900 – 1200 cm-1 and 100 – 400 cm-1.17,18 Scanning electron microscopy of these fractured surfaces reveals
In contrast, the Raman peaks of amorphous/nanocrystalline materials that the magnetite is organized into ca. 250 nm wide bundles of
are wider, weaker and typically not well resolved for the vibrations from crystallites that are oriented parallel to the long axis of the teeth, each
the phosphate or iron-oxygen moieties, which is consistent with our of which is surrounded by a thin organic layer4. Transmission electron
analyses of the core region of the radular teeth (Fig. 3d). The raman microscopy analysis of focused ion beam-milled samples from both
spectra also suggest a large α-chitin component in the core material the core region (Fig. 5e) and the outer magnetite shell region (Fig. 5d)
(Fig. 3d) and TEM studies reveal the presence of abundant chitin fibers confirm both the rod-like orientation of the magnetite crystallites and
throughout this region (Fig. 5e)19,20. In addition, EDS point scans across the weakly crystalline, organic rich nature of the core material. Selected
focused ion beam milled samples of the tooth core material reveal area electron diffraction and dark-field imaging of the outer magnetite
the presence of isolated domains of silica. These siliceous domains shell also reveal that although the magnetite crystallites (each of which
appear more electron transparent in TEM studies compared to the measure ca. 30 nm in diameter) are organized into aligned rod-like
comparatively electron dense surrounding hydrated iron phosphate columns, there is no preferred orientation of the constituent crystallites
(Supplemental Fig. 2)2,4. (Fig. 5d). In contrast, the diffuse electron diffraction pattern of the
Nanomechanical analyses of polished cross-sections through both tooth core material clearly demonstrates the weakly crystalline form of
the tooth tip and mid-region reveal that the two mineral phases the iron phosphate phase.
(the magnetite veneer and the core of weakly crystalline hydrated While the hardness and modulus of the teeth are directly related to
iron phosphate – see Fig. 2f) exhibit distinct mechanical properties the intrinsic mechanical properties of the constituent mineral phases,
(Fig. 4a). The magnetite veneer has a modulus ranging from 90 to indentation fracture measurements of sodium hypochlorite-treated

JAN–FEB 2010 | VOLUME 13 | NUMBER 1-2 45


INSIGHT Analysis of an ultra hard magnetic biomineral in chiton radular teeth

(a)
(b)

(c)

(d)

Fig. 3 Elemental and phase analysis of the radular teeth from C. stelleri. EDS analysis of the mature radular teeth (a,b) reveal an iron phosphate core surrounded
by a thick veneer of iron oxide. Powder XRD (c) suggests the presence of α-chitin and magnetite as the dominant crystalline phases. A chitin reference pattern21
is shown in orange and the crystal structure of magnetite is shown to the right of the diffraction pattern. Raman Spectroscopy (d) identifies the two likely mineral
phases in the shell and core regions as magnetite and hydrated iron phosphate, respectively. A Raman linescan (d-left) through the region highlighted in the
backscattered SEM image illustrates the transition zone between the two mineral phases. The dominant peak in the blue spectrum at 670 cm-1 corresponds to the
νFe-O vibration of magnetite while the dominant peak in the pink spectrum at 1010 cm-1 corresponds to the νP-O vibration of iron phosphate. A higher resolution
raman spectrum of the radular tooth core material with the labeled peaks corresponding to α-chitin, phosphate and water is shown in 2D-right. See Supplemental
Fig.3 for more details.

46 JAN–FEB 2010 | VOLUME 13 | NUMBER 1-2


Analysis of an ultra hard magnetic biomineral in chiton radular teeth INSIGHT

teeth reveal that following destruction of the tooth organic matrix, the equivalent to the hardest structural ceramics such as silicon carbide (SiC)
fracture toughness drops precipitously, while the modulus and hardness and boron carbide (B4C). For reference, a geological magnetite mineral
remain largely unaffected (<15% reduction, unpublished work). standard is also shown on the plot. Interestingly, the latter exhibits
This illustrates the important point that structural integrity can be hardness values (10.5 ± 0.2 GPa) equivalent to the radular teeth of C.
attained only in the presence of the organic matrix that facilitates the stelleri, but is significantly stiffer (175 GPa), i.e. 1.6 to 1.9 more stiff than
anisotropic organization of the magnetite crystallites, binds them into the radular teeth. Small-scale sliding at the mineral/organic interfaces
a composite structure, and plays a critical role in crack blunting and during external loading is a plausible cause to explain the larger
deflection at interfaces24. The highly mineralized nature of the external compliance of radular tooth magnetite. From the materials selection plot,
magnetite region of the radular teeth is exemplified by its low organic one expects the radular tooth magnetite to exhibit a higher tolerance
content (ca. 2%) compared to the core iron phosphate region (ca. 8%). to abrasion, here expressed in terms of residual indentation depth after
To gain additional insight into the abrasion resistance of the radular contact loading, i.e. the smaller the residual depth, the better abrasion
teeth from C. stelleri, we plotted the E/H combination on an Ashby- resistance. We performed indentation curves with a spherical tip
type plot that compares the abrasion tolerance of a wide range of (nominal tip radius 1 μm) on the two materials to verify this assertion
materials (Fig. 6a). Here the guidelines correspond to abrasion damage up to normal peak load of 1 mN. Representative curves are presented
by localized plastic deformation against a blunt and rigid contact in Fig. 6c, together with the predicted curves from the elastic Hertzian
(Fig. 6b), for which the critical normal load for yielding Py (according to solution (with the actual radius tip calibrated from indentation curves
Hertzian contact mechanics) is proportional to the materials property on fused quartz in the elastic domain). Since deviation from the elastic
group25, Py ∝ H3/E2. Hence by plotting E vs. H values on a log-log plot, solution corresponds to the onset of yielding, comparison between the
materials lying on lines with a slope of 2/3 are predicted to perform elastic solution and the obtained experimental data allows detection of
equivalently against yield damage at the contact. the critical load at yielding27. The data and analysis confirm that (1) the
Based on these criteria, we predict that the radular teeth of critical load at yielding is indeed higher for the radular teeth as predicted
C. stelleri perform better against wear damage than the hardest known from the index Py ∝ H3/E2, and (2) the residual indentation depth hr
biominerals: enamel and nacre. Perhaps even more remarkable, the of the radular teeth after incursion at a peak load of 1 mN is smaller,
predicted abrasion resistance against a blunt contact exceeds that of resulting in a better tolerance to abrasion than geological magnetite (in
zirconia (a common ceramic used in load-bearing applications)26 and is the absence of friction).

(a) (b)

Fig. 4 Nanomechanical testing of radular tooth cross-sections from C. stelleri. (a) Indentation (left) and the corresponding gradient (right) maps of modulus
(upper) and hardness (lower) through a tooth tip and mid-region reveal that the leading edge of the tooth has a higher modulus and hardness than the trailing
edge, thus establishing the self-sharpening condition illustrated in (b). The direction of tooth movement against the substrate is indicated by the blue arrow.
(c) adapted from23.

JAN–FEB 2010 | VOLUME 13 | NUMBER 1-2 47


INSIGHT Analysis of an ultra hard magnetic biomineral in chiton radular teeth

When abrasion is produced by sharp stiff particles, plastic deformation and microcracking becomes the dominant wear mechanism. It can
initially occurs at the contact and the abrasion resistance approximately be shown that abrasion damage is then proportional to the material
4
scales with the hardness25. In such a case, the radular teeth from C. property group K Ic / H·E2 where KIc is the fracture toughness31. Hence,
stelleri again outperform all other known biominerals. But yielding can a moderate increment of fracture toughness significantly improves the
be quickly followed by cracking at the contact in brittle materials25,28-30, abrasion resistance (power-law exponent of 4). While KIc of radular tooth

(a) (b)

(c)

(d) (e)

Fig. 5 SEM and TEM analysis of failure modes in the radular teeth of C. stelleri. While crack propagation through the tooth core is isotropic (a), those propagating
through the outer magnetite shell travel parallel to the long axis of the teeth (b). SEM analysis of a fractured tooth (c) reveals its closely packed rod-like
ultrastructure. TEM bright-field, dark-field, and selected area electron diffraction studies of focused ion beam milled tooth sections through the magnetite shell
region (d) and the core region (e) show the highly parallel nature of the magnetite crystallite bundles (d), the organic-rich fibrous core material, and the weakly
crystalline nature of the associated mineral phase (e).

48 JAN–FEB 2010 | VOLUME 13 | NUMBER 1-2


Analysis of an ultra hard magnetic biomineral in chiton radular teeth INSIGHT

magnetite still awaits experimental determination, it is anticipated that of nanotechnology and nanomanufacturing, exploiting control
its complex hierarchical microstructure toughens the structure against mechanisms established by nature to make novel materials and
catastrophic fracture in comparison to monolithic magnetite, much devices exhibiting unique properties. For example, the graded modulus
to what has been established for the hierarchical calcium carbonate design of the radular teeth and their anisotropic rod-like substructure
microstructures of nacre and conch shells32,33. Additionally, tolerance provides design cues that could be used in the fabrication of ultrahard
to abrasion is proportional to the factor 1/H.E2. Supplemental Table 1 materials for precision cutting or wear-resistant components. The
compares this index (as well as those discussed earlier) with literature graded modulus design would be particularly useful for creating a self-
values for various biominerals and engineering ceramics. These data sharpening condition in instances where location or situation would
suggest that the radular teeth from C. stelleri perform as well as hard not permit regular replacement or sharpening of the cutting blades. In
ceramics, but appear less tolerant to microcracking from a sharp abrasive addition to understanding the structure function relationships in these
than enamel or mollusk shells. Knowledge of the radular tooth’s fracture unique abrasion- and impact-tolerant materials, the high abundance
toughness, however, will be necessary to confirm this prediction. These of crystalline magnetite also make them an ideal model system for
results emphasize the importance of the geometry of the contact elucidating the synthesis mechanisms of magnetic materials at ambient
in defining the abrasion tolerance of biominerals. A comprehensive temperatures and pressures and near-neutral pH. Toward this endeavor,
experimental study of wear at the micro-scale is the topic of an on-going future investigation of the structure and composition of the tooth
investigation. organic matrix could ultimately lead to the identification of specific
The results obtained from the investigations outlined above chemical functionalities that play a critical role in controlling magnetite
have the potential to enable progress in the emerging fields nucleation and growth.

(a) (c)

(b)

Fig. 6 Abrasion performance of the radular teeth from C. stelleri against a blunt contact. (a) Property map resistance of yield damage (stiff abrasive) for various
materials (adapted from25) predicts that the radular teeth perform better than known biominerals against a spherical contact and approaches that of the hardest
engineering ceramics. (b) Schematic description of yield initiation beneath the contact during a normal loading/unloading cycle. (c) Representative indentation
curves with a spherical tip on the radular teeth and a geological magnetite control: deviation from the elastic Hertzian solution corresponds to incipient yielding and
confirms that the radular teeth have a higher critical load (in agreement with Py α H3/E2), while its residual impression, hr, is smaller at that peak load.

JAN–FEB 2010 | VOLUME 13 | NUMBER 1-2 49


INSIGHT Analysis of an ultra hard magnetic biomineral in chiton radular teeth

Materials and methods geological magnetite control at prescribed peak loads of 1 mN (0.1 mN/
Research specimens sec. loading rate) with a 5 second hold. Calibration of the tip radius was
Live specimens of Cryptochiton stelleri were collected from Vancouver, done by performing a series of indents on fused quartz (reduced modulus
British Columbia between October 2008 and March 2009 and 69 GPa) in the elastic domain and fitting the loading curves with the
maintained live in a recirculating seawater system until ready for Hertzian’s solution. Best fit gave an actual radius of curvature of 400 nm.
use. The radulae were dissected from these specimens and rinsed in
fresh seawater to remove any loose organic debris and immediately Scanning electron microscopy and
dehydrated through a graded series of ethanol treatments. Following Energy dispersive spectroscopy
complete dehydration to 100% ethanol, the anterior-most region Polished tooth samples were gold coated and examined with a FEI
of each radula (containing the fully mature teeth) was removed and XL-40 or a Tescan VEGA TS-5130MM scanning electron microscope
embedded in Spurrs resin. The resin blocks obtained were polished equipped with an IXRF systems energy dispersive spectrometer. In
to P1200 with progressively finer grades of silicon carbide paper and addition to the polished sections, whole fractured teeth were also
then with diamond lapping films of particles sizes down to 0.25 μm to analyzed by SEM to identify their failure modes. Additional teeth
obtain a smooth finish. were soaked overnight in a 5.25% sodium hypochlorite (NaOCl)
The resulting samples were examined either by nanoindentation, solution to remove the organic phase. The resulting samples were
Raman spectroscopy, back-scattered and secondary scanning electron then soaked overnight in DI water, mechanically fractured, rinsed
microscopy, or focused ion beam milled for transmission electron briefly with ethanol and air dried for examination by SEM. Additional
microscopy as described below: NaOCl-treated teeth were prepared, embedded and polished for
nanomechanical testing as described above.
Nanoindentation
Full-map indentation tests were performed on cross-sections through Transmission electron microscopy
the tip and mid-region of the mature radular teeth from C. stelleri Polished tooth cross-sections were prepared for TEM using a focused
in ambient air using a Triboindenter nanomechanical testing system ion beam milling machine (Leo Gemini, 1540XB) to obtain thin
(Hysitron, Minneapolis, MN, USA) with either a berkovich or a cube sections measuring ca 100nm thick. The resulting foils were examined
corner tip at a peak force of 5mN. The load function consisted of using a FEI-PHILIPS CM300 transmission electron microscope at 300 kV
a 5-second loading to 5 mN, followed by a 5-second hold at that accelerating voltage.
force, and then a 5-second unloading. The resulting indents in a grid-
array measured ca. 12.5 μm apart and the hardness and reduced Raman spectroscopy
modulus were calculated from the unloading curve of each, using Polished tooth samples prepared as described above were investigated
the Oliver-Pharr method34. Similar indentation measurements were using Raman spectroscopy (λ0= 780 nm) under low laser power
also performed on the same tooth samples that had been hydrated conditions (6 mW) to minimize sample heating. Raman linescans
with deionized water using rows of intents that crossed from the were performed with mineral standards to help identify the various
outer magnetite shell region into the iron phosphate core region. The inorganic phases present in the different regions of the teeth using a
measurements were performed to investigate the possible influence that ThermoFisher DXR dispersive Raman microscope.
sample dehydration might have on the measured mechanical properties.
These results revealed that due to the highly mineralized nature of the X-ray transmission and diffraction
teeth, there was on average only a ca. 15% reduction in both modulus For X-ray diffraction, the mineralized caps from the mature teeth of C.
and hardness within both mineral phases of the teeth in the hydrated stelleri were individually mechanically separated from the supporting
vs. the dry state. For this reason, and to standardize measurements from chitinous stalk, ground into a fine powder, and examined using a Philips
sample to sample, all subsequent measurements were performed with X’PERT Powder Diffractometer with Cu Kα radiation.
dry samples. Synchrotron X-ray transmission imaging was performed at beamline
The resulting modulus and hardness data were plotted using XYZPlot X13B at the National Synchrotron Light Source, Brookhaven National
(Hysitron) and the corresponding gradient maps were created by Laboratory, using 19 keV X-rays (λ = 0.65 Å) and a beam spot focused
applying a Gaussian blur function to the resulting indentation maps to ≈ 10 μm ×10 μm. Specimens consisting of several radular teeth
which were superimposed on the corresponding back-scattered electron still attached to their basal membrane were mounted in the beam
micrographs. For comparative purposes, a profile of indents was also and scanned in 25 μm steps both parallel and perpendicular to the
performed on polished sections of conch shells (Strombus gigas) using the basal plane. Transmitted intensity was recorded using a photodiode
procedures described above. Indents were also performed with a spherical detector fixed beyond the sample and normalized by incident intensity
tip (nominal radius 1 μm) on both the polished radula sections and a measured with an ion chamber.

50 JAN–FEB 2010 | VOLUME 13 | NUMBER 1-2


Analysis of an ultra hard magnetic biomineral in chiton radular teeth INSIGHT

Acknowledgements: (a)
We thank Philip Bruecker and Shane Anderson for help with specimen
acquisition, Sara Krause for the illustration in Fig. 1b, and Dr. Kenneth
Evans-Lutterodt of the National Synchrotron Light Source for contributing
his expertise at the microdiffraction endstation X13B. The NSLS is
supported under USDOE Contract DE-AC02-98CH10886. This work made
use of the Scanning Electron Microscopy facilities in the laboratory of D.
Morse at UCSB. Professor Kisailus acknowledges support from his initial
complement at U. C. Riverside.
(b)
Author Contributions:
D.K. and J.W. designed the study. D.K., J.W. and A.M. wrote the paper.
J.W., Q.W. and A.T. prepared specimens. D.K., J.W. and K.B. collected and
interpreted microscopy data. D.K., J.W., Q.W., and S.H. gathered Raman
data. D.K. and S.H. interpreted the Raman data. J.W., A.M., P.M., R.N. and
R.S. performed mechanics experiments. A.M., R.N. and R.S. interpreted
mechanics data. J.W. and E.D. collected and interpreted synchrotron data.

Supplemental Information (c)

Fig. S2 In focused ion beam milled samples isolated from the radular tooth cores,
randomly dispersed domains of electron transparent material are routinely
observed (Fig. S2a). To investigate the nature of these domains (some of
which are indicated by dotted yellow circles in Fig. S2a), EDS point scans were
performed to look for variability in elemental composition (Fig. S2b). The results
from one of these measurements are shown in Fig. S2c. EDS elemental point
spectra reveal that these electron transparent regions contain elevated quantities
of silicon compared to levels observed in the background material. These results
Fig. S1 Electron diffraction pattern of the outer magnetite shell of the radular indicate that while the radular teeth of C. stelleri do in fact contain silica as has
teeth of C. stelleri. Miller indices corresponding to the first 4 rings of the been reported previously2,4, the siliceous material is phase segregated from the
diffraction pattern are indicated. bulk core material of weakly crystalline hydrated iron phosphate.

Table S1 Materials property group of various biominerals and structural ceramics. Yield damage from a blunt contact
is proportional to H3/E2; yield from a sharp contact mostly depends on H, while microcraking from a sharp abrasive
4
depends on the index KIc /H E2. KIc of the radular teeth is unknown, hence only the factor 1/H E2 is presented and used
for comparison.
H [GPa] E [GPa] H3/E2, [GPa] 1/H E2 [GPa-3] References
Radular teeth from C. stelleri 9 – 12 90 - 125 0.10 – 0.14 7·10-6 – 1.4·10-5 This study
Magnetite 10.5 175 0.03 7·10-6 This study
Enamel 3.5 – 4.5 65 – 75 0.01 – 0.02 4·10-5 – 6.8·10-5 35-37
Aragonite (Abalone and Conch shells) 3–4 70 – 75 0.01 6.3·10-5 This study, 38,39
Guanoine 4.5 62 0.02 5.8·10-5 40
Zirconia 10 – 15 140 – 240 0.05 – 0.06 1.2·10-6 – 5.1·10-6 26,41,42
SiC 20 – 30 410 – 480 0.05 – 0.12 1.5·10-7 – 3·10-7 42,43
B4C 30 430 – 480 0.13 1.6·10-7 42,44

JAN–FEB 2010 | VOLUME 13 | NUMBER 1-2 51


INSIGHT Analysis of an ultra hard magnetic biomineral in chiton radular teeth

255 Lattice modes


288 Lattice modes
344 Lattice modes
452 ν2: symmetric bending mode of the O-P-O bonds
477 ν2: symmetric bending mode of the O-P-O bonds
550 ν4: antisymmetric bending mode of the O-P-O bonds
570 ν4: antisymmetric bending mode of the O-P-O bonds
600 ν4: antisymmetric bending mode of the O-P-O bonds
615 ν4: antisymmetric bending mode of the O-P-O bonds
834 ν: symmetric stretching mode of the C-O-C bonds in alpha-chitin
899 ν: stretching mode of the C-C bonds from chitin
1010 ν1: symmetric stretching mode of the P-O bonds
1030 ν3: antisymmetric stretching mode of the P-O bonds
1110 ν3: antisymmetric stretching mode of the P-O bonds
1280 Amide III bands of chitin alpha-helix
1330 Amide III bands of chitin alpha-helix
1380 Amide III bands of chitin alpha-helix
1460 δ: bending (CH2) - in alpha-chitin
1550 δ: bending (CH2) - in alpha-chitin
1620 δ: bending (OH) - this is hydrogen bonded to C=O in alpha-chitin
1660 ν: stretching mode of the C=O bonds in random coil / helix in
alpha-chitin
Fig. S3 Detailed analysis of the Raman data obtained from the core region of the 2860 ν: stretching mode C-H bonds
radular teeth of C. stelleri. The specific vibrations corresponding to specific peak 2930 ν: stretching mode C-H bonds
locations are listed to the right of the figure: 3360 ν: stretching (OH)

REFERENCES 21. Cardenas, G., et al., J. Appl. Polym. Sci. (2004) 93, 1876-1885
1. Morris, R. H., et al., Intertidal Invertebrates of California, Stanford Press, Stanford, 22. Ashby, M. F., et al., Proceedings of the Royal Society of London Series
Calif., USA, (1980) A-Mathematical and Physical Sciences (1995) 450, 123-140
2. Lowenstam, H. A., and Weiner, S., On biomineralization, Oxford University Press, 23. van der Wal, P., et al., Mater. Sci. & Eng. C-Biomimetic and Supramolecular
Oxford, England, (1989) Systems (2000) 7, 129-142
3. Ricketts, E. F., et al., Between Pacific Tides fifth edition, Stanford Univ. Press, 24. Chan, K. S., et al., Mater. Sci. Eng. A- Structural Materials Properties
Stanford, Calif., USA, (1985) Microstructure and Processing (1993) 167, 57-64
4. Towe, K. M., and Lowenstam, H. A., J. Ultrastructure Res (1967) 17, 1-13 25. Zok, F. W., and Miserez, A., Acta Mater. (2007) 55, 6365-6371
5. Moseley, H. N., and Madreporarian Corals, Report of the scientific results of the 26. Medvedovski, E., Wear (2005) 249, 821-828
voyage of H.M.S. Challenger during the years 1873-76: Zoology Part VII. Report 27. Schuh, C.A., and Lund, A., J. Mat. Res. (2004) 19, 2152
on certain Hydroid, Alcyonarian, and Madreporarian Corals, Neill and Company,
Edinburgh, UK, (1881) 28. Williams, J. A., Tribolog. Int. (2005) 38, 863-870

6. Watson, R. B., Report of the scientific results of the voyage of H.M.S. Challenger 29. Hsu, S. M., and Shen, M., Wear (2004) 256, 867-878
during the years 1873-76: Zoology Part XLII. Report on the Scaphopoda and 30. Sharp, S. J., et al., Acta Metallurgica et Materialia. (1993) 41, 685-692
Gasteropoda, Neill and Company, Edinburgh, UK, (1886) 31. Miserez, A., et al. Adv. Funct. Mater. (2008) 18, 1241-1248
7. Schulze, F. E., Report of the scientific results of the voyage of H.M.S. Challenger 32. Kamat, S., et al., Nature (2000) 405, 1036-1040
during the years 1873-76: Zoology Part LIII. Report on the Hexactinellida, Neill
and Company, Edinburgh, UK, (1887) 33. Barthelat, F., and Espinosa, H. D., Experimental Mechanics (2007) 47, 311-324

8. Vielzeuf, D., et al., American Mineralogist (2008) 93, 1799–1815 34. Oliver, W. C., and Pharr, G. M., J. Mater. Res. (2004) 19, 3-20

9. Weiner, S., and Addadi, L., Science (2002) 298, 375-376 35. Marshall, G. W., et al., J. Biomed. Mater. Res. (2001) 54, 87.

10. Macey, D. J., and Brooker, L. R., J. Morphology (1996) 230, 33-42 36. Habelitz, S., et al., J. Biomechanics (2002) 35, 995.

11. Lee, A. P., et al., J. Biol. Inorganic Chem. (2003) 8, 256-262 (2003) 37. Jackson, A. P., et al., Proceedings of the Royal Society of London Series B-Biological
Sciences (1988) 234, 415.
12. Brooker, L. R., et al., Mar. Biol. (2003) 142, 447-454
38. Imbeni, V., et al., Nat. Mater. (2005) 4, 229.
13. Wealthall, R. J., et al., J. Morphology (2005) 265, 165-175
39. Menig, R., et al., Acta Mater. (2000) 48, 2383.
14. Lydekker, R., The Royal Natural History, London Warne F. & Co., New York, (1896)
40. Bruet, B. J. F., et al., Nat. Mater. (2008) 7, 748.
15. Carefoot, T. H., Proc. Malacol. Soc. London (1965) 36, 203-212
41. Cawley, J. D., The Encyclopedia of Materials Science and Technology
16. Brooker, L. R., et al., Venus (Tokyo) (2006) 65, 71-80 (Buschow, K. H. et al.,), Pergamon, (2001).
17. Zaghib, K., and Julien, C. M., Journal of Power Sources (2005) 142, 279-284 42. Tressler, R. E., The Encyclopedia of Materials Science and Technology
18. Gremlich, H.U., and Yan, B., Infrared and Raman Spectroscopy of Biological (Buschow, K. H., et al.,), Pergamon, (2001).
Materials, Marcel Dekker, New York, (2001) 43. Srinavassan, M., and Rafaneillo, W., Carbide, Nitride and Boride Materials Synthesis
19. Macey, D. J., et al., Acta Zoologica (1996) 77, 287-294 and Processing (Weimler, A.), Springer, (1996).
20. Evans, L. A., et al., Philosophical Transactions of the Royal Society of London 44. Dewith, G., J. Mater. Sci. (1984) 19, 457.
Series B-Biological Sciences (1990) 329, 87-96

52 JAN–FEB 2010 | VOLUME 13 | NUMBER 1-2

You might also like