You are on page 1of 11

1983

The Journal of Experimental Biology 215, 1983-1993


© 2012. Published by The Company of Biologists Ltd
doi:10.1242/jeb.064790

RESEARCH ARTICLE
Comparison of structural, architectural and mechanical aspects of cellular and
acellular bone in two teleost fish
Liat Cohen1, Mason Dean3, Anna Shipov1, Ayelet Atkins1, Efrat Monsonego-Ornan2 and Ron Shahar1,*
1
Koret School of Veterinary Medicine and 2School of Biochemistry and Nutrition, The Robert H. Smith Faculty of Agriculture, Food
and Environment, The Hebrew University of Jerusalem, Israel and 3Max Planck Institute of Colloids and Interfaces, Department of
Biomaterials, Am Mühlenberg 1, 14424 Potsdam, Germany
*Author for correspondence (rons@savion.huji.ac.il)

Accepted 13 February 2012

SUMMARY
The histological diversity of the skeletal tissues of fishes is impressive compared with that of other vertebrate groups, yet our
understanding of the functional consequences of this diversity is limited. In particular, although it has been known since the mid-
1800s that a large number of fish species possess acellular bones, the mechanical advantages and consequences of this
structural characteristic – and therefore the nature of the evolution of this feature – remain unclear. Although several studies have
examined the material properties of fish bone, these have used a variety of techniques and there have been no direct contrasts of
acellular and cellular bone. We report on a comparison of the structural and mechanical properties of the ribs and opercula
between two freshwater fish – the common carp Cyprinus carpio (a fish with cellular bone) and the tilapia Oreochromis aureus (a
fish with acellular bone). We used light microscopy to show that the bones in both fish species exhibit poor blood supply and
possess discrete tissue zones, with visible layering suggesting differences in the underlying collagen architecture. We performed
identical micromechanical testing protocols on samples of the two bone types to determine the mechanical properties of the bone
material of opercula and ribs. Our data support the consensus of literature values, indicating that Youngʼs moduli of cellular and
acellular bones are in the same range, and lower than Youngʼs moduli of the bones of mammals and birds. Despite these
similarities in mechanical properties between the bone tissues of the fish species tested here, cellular bone had significantly
lower mineral content than acellular bone; furthermore, the percentage ash content and bone mineral density values (derived from
micro-CT scans) show that the bone of these fishes is less mineralized than amniote bone. Although we cannot generalize from
our data to the numerous remaining teleost species, the results presented here suggest that while cellular and acellular fish bone
may perform similarly from a mechanical standpoint, there are previously unappreciated differences in the structure and
composition of these bone types.
Key words: acellular bone, teleost fish, mechanical properties.

INTRODUCTION The bone matrix is permeated by an interconnected network of


Bone forms a tough and protective load-bearing framework for the microscopic channels (canaliculi), which contain cellular processes
bodies of vertebrates. The magnitude and direction of the principal of osteocytes, and thus embedded osteocytes are able to
loads to which bones are exposed can vary. Furthermore, bone is communicate with their neighbors.
loaded repeatedly and cyclically during the animal’s life; therefore, In the middle of the 19th century it was observed that in the great
fatigue damage in the form of microcracks is bound to accumulate majority of teleost fish species (and thus a large proportion of
within the bone material. Most vertebrate bones are able to cope vertebrates), bone is acellular and does not contain osteocytes. The
with a changing mechanical environment and damage accumulation ubiquity of osteocytes throughout vertebrate animals, including
through adaptation of their morphology and self-repair (termed mammals, birds, amphibians and reptiles, and their proposed
modeling and remodeling, respectively). Both processes depend on mechanosensory function make this observation surprising.
the ability of bone to sense local changes in its mechanical Interestingly, the bone of basal teleosts is cellular; therefore, the
environment, such as increasing strain or strain energy density levels, loss of osteocytes is a derived property (Parenti, 1986; Kranenbarg
which result from the accumulation of microcracks or changing et al., 2005a); as the main function of bone is mechanical, this
loads. This mechanosensory ability is widely thought to be provided suggests that the evolution of acellularity confers a mechanical
by bone cells – the osteocytes (Adachi et al., 2009; Bonewald, 2011). advantage to the skeleton. However, based on current knowledge
Osteocytes arise from the bone-forming cells (osteoblasts) (Franz- of mammalian bone biology, lack of cells would hinder the
Odendaal et al., 2006). In the bones of mammals, some osteoblasts functionality of fish bone, restricting its ability to respond to
become embedded within the extracellular matrix they deposit changing loading conditions and to repair microdamage. And yet,
(osteoid), and this matrix then becomes mineralized. Once embedded various reports suggest that acellular fish bone functions
and trapped within the matrix, the osteoblasts become known as mechanically in a manner similar to its cellular counterpart. For
osteocytes, residing within small spaces named osteocytic lacunae. example, several authors have reported that the acellular bone of

THE JOURNAL OF EXPERIMENTAL BIOLOGY


1984 L. Cohen and others

neoteleosts – the eight most-derived superorders of bony fishes, samples were wrapped in saline-soaked gauze and stored at –20°C
which include the Percomorpha, the largest and most diverse group until further testing.
of fishes – responds to mechanical loads by the process of modeling,
as a cellular bone would (Huysseune et al., 1994; Hegrenes, 2001; Histology and light microscopy
Kilhara et al., 2002; Kranenbarg et al., 2005a; Kranenbarg et al., Rib and operculum bone samples, taken from four carp and four
2005b; Nemoto et al., 2007; Kitamura et al., 2010; Totland et al., tilapia, were fixed overnight in 4% paraformaldehyde in PBS at
2011). Furthermore, although osteocytic lacunae are believed to act 4°C. They were then decalcified in 0.5moll–1 EDTA, pH7.4. Once
as stress concentrators (e.g. Nicolella et al., 2006), and therefore decalcified, they were dehydrated in graded ethanol solutions and
their absence could conceivably decrease the rate of microdamage Histo-clear© (National Diagnostics, Atlanta, GA, USA). They were
accumulation, it is reasonable to assume that acellular bone still then embedded in Paraplast (SPI supplies, West Chester, PA, USA),
accumulates some fatigue damage from repeated cyclic mechanical cut by microtome into 5m sections and mounted on glass slides
loads, and therefore would benefit from the ability to replace (Superfrost, Thermo Scientific, Barrington, IL, USA). The
damaged areas. However, the manner in which acellular fish bone embedded sections were deparaffinized in xylene, washed twice with
is able to undergo modeling and remodeling without the strain- ethanol, rehydrated through a graded series of ethanol solutions and
sensing osteocytes is currently unknown, although osteoblasts and stained with hematoxylin and eosin (H&E).
bone lining cells are reasonable candidates for mediators of these Five histological sections from each of the four carp opercula
processes (Witten and Huysseune, 2009). were used to obtain estimates of osteocyte density within the bone.
The structure and mechanical properties of bones from a large To determine osteocyte density, a large rectangle was outlined
variety of vertebrates from different taxonomic levels (mammals, digitally over a high-resolution image of each sample such that it
amphibians, reptiles and birds) have been extensively studied. It is covered as much of the section as possible, the area of the rectangle
thus somewhat surprising that so little is known about the structure was determined and all osteocytes within this region were counted
and mechanical properties of the bones of modern teleost fish, as manually. Osteocyte density was determined as the number of
the number of teleost fish species likely represents at least 50% of osteocytes within the rectangle divided by its area.
the rest of the vertebrate species combined (Currey, 2010). Very In addition, several unstained, fresh samples of ribs and opercula
few studies have examined the material properties of fish bone, with from four carp and four tilapia were prepared for light microscopy.
cellular fish bone being the most studied (Roy et al., 2000; Rho et Transverse and axial rib slices were sectioned from the proximal
al., 2001; Erickson et al., 2002; Wang et al., 2002; Zhang et al., region of the fifth rib, and both axial and dorso-ventral strips were
2002; Paxton et al., 2006) and, to the best of our knowledge, with prepared from the operculum. The regions from which the samples
only one study providing values for acellular fish bone (Horton and in the different bones were obtained are schematically shown in
Summers, 2009). However, none of these previous studies tested Fig.1. Bone samples were prepared using a slow-speed water-cooled
and compared the two bone types, and valid comparisons among diamond-blade saw (Isomet® low speed saw, Buehler, Lake Bluff,
the studies are difficult because they employed a wide diversity of IL, USA), and their surfaces ground and polished using a
methods and focused on disparate species and different skeletal grinding–polishing device (Minimet®1000, Buehler). The prepared
elements. The dearth of data regarding the mechanics of fish bone surfaces were examined by a reflected-light microscope (Olympus®
tissue most likely stems from the fact that fish bones are typically BX 51 microscope, Olympus, Tokyo, Japan) at different
small and their shape is extremely irregular, making sample magnifications, and images were captured using a high-resolution
preparation and testing very difficult (Currey, 2010). In order to camera (Olympus® DP 71 digital camera).
properly compare the mechanical properties of cellular and acellular
bone, it is imperative that samples of the two bone types be tested Water, organic and ash content
by exactly the same method using multiple replicates. Opercula from eight mature carp and eight mature tilapia were used
In the study reported here, we examined in detail the structure for determination of ash content. Small pieces (800–1200mg) of
and mechanics of two bones (operculum and rib) obtained from fish each operculum were carefully dissected and cleaned of all soft
with acellular bones [the tilapia Oreochromis aureus (Steindachner, tissue, then placed in acetone solution for 12h to remove all lipids.
1864); Acanthopterygii] and from fish with cellular bones (the Samples were weighed with a semi-analytical scale (±0.05mg) to
common carp, Cyprinus carpio L.; Ostariophysi). We used various determine lipid-free wet mass, then placed in a ceramic cup and
techniques to describe their morphological features, mean mineral heated in an oven (Tuttnauer, Jerusalem, Israel) to 100°C for 3h to
density and microstructural features. These analyses are paired with remove all unbound water. After heating, the samples were
standardized mechanical testing methods to determine and compare immediately weighed to determine their dry mass (organic and
the mechanical properties of cellular and acellular fish bone tissue. mineral content) and placed (within their ceramic cups) in an oven
at 500°C for 10h, so that all organic material was burned. The
MATERIALS AND METHODS remaining ash was weighed immediately, so that the dry bone ash
Specimens and sample preparation content (%) of each sample could be determined. Water (wet mass
Sixteen freshly killed, healthy 18month old carp weighing – dry mass), organic (dry mass – ash mass) and ash content were
672–1281g, and 15 freshly killed, healthy 14month old tilapia, also calculated by dividing by wet mass to determine a percentage.
weighing 356–576g, were obtained from a commercial fishery
(Madan, Maagan Michael, Israel). Mechanical testing
The operculum and fifth rib were then harvested from all fish by Mechanical testing was performed with the samples immersed in
manual dissection within 1h. Samples of all bones were collected saline, using a custom-built micromechanical testing device (see
for histological study and immediately fixed in 4% Fig.2). All samples were thawed immediately before testing for 24h
paraformaldehyde solution (Sigma, St Louis, MO, USA). Samples at 4°C, then for 1h at room temperature. The operculum beams
were also collected for light microscopy, micro-computed were tested in three-point bending; rib samples were tested in axial
tomography (micro-CT) scanning and mechanical testing. These compression because three-point bending tests require long, straight

THE JOURNAL OF EXPERIMENTAL BIOLOGY


Structure and mechanics of acellular bone 1985

A
Dorso-ventral

Axial

Fig.2. Schematic drawing of the micromechanical testing device with close-


Fig.1. Schematic drawing of the samples obtained for mechanical testing. up photographs demonstrating the loading scenario of (A) 3-point bending
Two kinds of rectangular beam were excised from the operculum. From of operculum beams and (B) compression testing of rib sections. Note in B
each fish, one beam was cut such that its long axis was in the axial the platen is in contact with the rounded end of the anvil, allowing the
(cranio-caudal) direction and a second beam (from the contralateral platen to compensate for slight deviations from parallelism that may occur
operculum) was cut such that its long axis was in the dorso-ventral between the two cut edges of the rib section.
direction. The rib sample was removed as a short (~2.5mm long) section
obtained from the proximal, straight and uniform part of the fifth rib.

samples with a consistent cross-section, and these conditions could for each sample (mm2) as an average of the CSA of the proximal
not be adequately met for ribs. and distal ends, as determined from digital images (ImageJ, NIH,
Bethesda, MD, USA, version 1.44I).
Operculum samples
Two beams were prepared from opercula from each of eight carp Three-point bending of operculum beams
and seven tilapia: beams from the left operculum were oriented along The operculum beams were tested by three-point bending. The
the axial orientation of the operculum, while beams prepared from beams were placed in the custom-built micromechanical testing
the right operculum were oriented in a perpendicular (dorso-ventral) device within a saline-containing test chamber (see Fig.2A). A
orientation (see Fig.1). Thirty beams of roughly similar size were stationary anvil was attached to the wall of the test chamber. This
prepared in all (14 from tilapia opercula and 16 from carp opercula). anvil consisted of two supports, which were 8mm apart. A
The length of the axial beams, for both types of fish, ranged between moving anvil with a single prong was attached to a load cell (model
11.4 and 21.93mm, and that of the dorso-ventral beams ranged 31, Honeywell, Sensotec, Columbus, OH, USA), which was
between 12.31 and 23.73mm. The width of the dorso-ventral beams attached in turn to a high-precision linear motor [Physik
was 4.94±0.42mm (mean ± s.d., here and for all subsequent values Instrumente (PI) GmbH, Karlsruhe, Germany]. The tip of the
unless otherwise stated) for the carp and 5.03±0.20mm for the moving prong was centered between the bottom supports (see
tilapia. The width of the axial beams was 5.25±0.30mm for the carp Fig.2A). All parts of the testing system (including the chamber
and 5.11±0.07mm for the tilapia. The thickness of the beams was itself) were made of stainless steel. The parts contacting the sample
the native thickness of the operculum, and was measured for each (bottom supports and upper prong) had rounded profiles (1mm
beam at several locations. Thickness ranged between 0.60 and diameter) to reduce stress concentration. The upper prong was
1.04mm in the carp samples and between 0.48 and 0.84mm in the brought into contact with the tested beams at a predetermined
tilapia samples; however, there was never more than 5% variation preload (2N), the chamber was filled with physiological saline
in depth across the length of any given beam. solution at room temperature until the sample was fully immersed,
and bending tests were conducted under displacement control at
Rib samples a rate of 500mm–1 up to a final displacement of 1500m (at this
Eight samples of the left fifth rib of carp and seven samples of the displacement the beams were in the plastic, post-yield zone, but
left fifth rib of tilapia were prepared. A tubular segment (carp had not fractured yet). Force–displacement data were collected by
segment length 2.66±0.04mm; tilapia segment length custom-written software (LabView, National Instruments, Austin,
2.65±0.19mm) was obtained from the proximal, reasonably straight TX, USA) at 50Hz. The resulting load–displacement curves were
part of each rib, using the Isomet® slow-speed water-cooled used to estimate the Young’s modulus of the beam material. It
diamond-blade saw (see Fig.1). In compression testing (see below), should be noted that as shear deformation was not taken into
stress calculation requires knowledge of the cross-sectional area of consideration, estimated values of the Young’s modulus are
the sample; as biological samples do not have a perfectly uniform slightly underestimated. However, as the mean span-to-depth ratio
cross-sectional area (CSA), we chose samples with relatively small of all tested beams was 12.06±3.52, this effect is relatively minor
CSA variation (<10%) and set the effective transverse bone CSA (Spatz et al., 1996; Draper and Goodship, 2003). Young’s modulus

THE JOURNAL OF EXPERIMENTAL BIOLOGY


1986 L. Cohen and others

was calculated based on the following relationship obtained from of the load cell, making the correction valid (if the load cell was
beam theory: much more compliant than the sample, such a correction would be
inaccurate).
S × L3
E= , (1) In order to validate the results obtained by this experimental setup,
48 I first a hollow circular tube was prepared from Delrin®
where E is the Young’s modulus of the bone material in the direction (polyoxymethylene). This material has known mechanical properties
coinciding with the long axis of the beam (Nmm–2), S is the slope (the Young’s modulus of Delrin® is 3.2GPa, quite similar to that
of the linear part of the force–displacement curve (Nmm–1), L is of the bones tested here). The tube was 2.5mm long, with a 1.3mm
the distance between the stationary supports (mm) and I is the o.d. and 0.95mm i.d., and was thus similar in dimensions to the
moment of inertia (mm4) of each operculum beam. The moment of tested rib sections. The tube was compressed 4 times and the
inertia I of a rectangular beam is calculated as follows: Young’s modulus for Delrin® determined by these experiments was
2.92±0.096GPa, within an acceptable margin of error of the
b × d3
I= , (2) material’s actual stiffness.
12
where b is the beam width (mm) and d is the beam depth (mm). Micro-CT scanning
The tubular rib sections and opercular beams were scanned using
Compression testing of rib samples a micro-CT scanner (1174 scanner, SkyScan®, Kontich, Belgium).
All rib samples were tested by axial compression under displacement The X-ray source was set at 50kVp (peak kilovoltage) and 800A.
control in the same micromechanical testing device (Fig.2B). Each In total, 450 projections were acquired over an angular range of
sample was placed between two anvils, one of which was stationary 180deg. The operculum samples of both carp and tilapia were
and attached to the wall of the test chamber, while the other was scanned with an isotropic voxel size of 27.6m and an integration
movable and attached to a load cell, which was attached in turn to time of 3900ms, while the carp and tilapia rib samples were scanned
a high-precision linear motor. The moving anvil had a well- at an isotropic voxel size of 11.1 and 8.6m, with integration times
lubricated ball-like edge of 10mm diameter upon which a stainless of 3600 and 3500ms, respectively. All scans were performed with
steel platen with a corresponding lubricated half-sphere cavity was a 0.25mm aluminium filter to decrease beam-hardening effects.
fitted, such that the platen could rotate to adjust for slight Scans were reconstructed and analysis was performed using
misalignment of the sample’s edges (see Fig.2B). One end of the commercial software (NRecon® SkyScan® software, version 1.6.1.2
sample was placed against the platen and the movable anvil was and CT analyser® SkyScan® software, version 1.9.3.2, respectively).
moved carefully until it contacted the opposite end of the sample, Bone mineral density (BMD) was determined based on calibration
and a desired preload was obtained (2N). The sample was then fully with two phantoms of known mineral density (0.25 and
immersed in saline and loading proceeded at a constant rate of 0.75mgcm–3) supplied by SkyScan®, which were scanned under
60mmin–1, up to a total displacement of 60m. exactly the same conditions as the operculum and rib specimens.
Force–displacement data were collected by custom-written software
(LabView). The resulting load–displacement curves were analyzed Statistics
to calculate the Young’s modulus of the rib sections tested. Statistical analyses of quantitative data (BMD, ash content and
The modulus was calculated as the ratio of stress to strain, where mechanical testing results) were made to find differences between
the stress was determined by: carp and tilapia groups, as well as between ribs and opercula within
F each species of fish. The quantitative data were compared between
σ= , (3) and within groups by the non-parametric Mann–Whitney test. The
A level of significance was set at P≤0.05.
where  is stress (Nmm–2), F is the applied force (N) and A is the
average bony cross-sectional area of the rib section (mm2) RESULTS
determined as described above. The strain was determined by: Histology and light microscopy
The current study focused primarily on the comparison of the
ΔL
ε= , (4) material properties of the bones of these two teleost species.
L However, as the mechanical behavior of bone is strongly linked to
where  is the strain (dimensionless), L is the shortening of the its structure, we also provide a brief description of some features
compressed rib segment under load (mm) and L is its original length discerned from a light microscopy study of stained and unstained
(mm). Thus the modulus could be extracted from the results of the bone samples.
experiment as: Fish opercula are bony plates, thin and slightly curved in the
medio-lateral direction. Both tilapia and carp opercula exhibited
FL
E= . (5) three distinct tissue zones, observable under reflected light
AΔL microscopy in either coronal sections (i.e. opercula divided into
As the rib segments were quite stiff, part of the movement of the rostral and caudal sections) or transverse sections (i.e. divided into
anvil was due to compression of the load cell. Load cell compliance dorsal and ventral sections). The lateral (outer) and medial (inner)
was determined by compressing the two stainless anvils against each thirds of the thickness of the bone showed clear lamellar organization
other, and evaluating the resulting load–deformation curve. Thus (see area bracketed in Fig.3B), sandwiching a middle, less-ordered
for a given load, sample compression could be determined by (isotropic) tissue (see area bracketed in Fig.3A), similar in
subtracting the load cell compression displacement from the total appearance to ‘woven’ bone (Currey, 2008). The blood supply to
sample-load cell compression displacement. Because for the same all bones in both types of fish is quite sparse, but blood vessels
load the deflection of the load cell was found to be about 50% of tended to be localized in the disordered woven zone. The lateral
that of the sample-load cell unit, sample stiffness was similar to that and medial lamellar zones appeared thicker and more compact in

THE JOURNAL OF EXPERIMENTAL BIOLOGY


Structure and mechanics of acellular bone 1987

Fig.3. Light microscopy and histology images of operculum


A B and rib samples from carp (left column) and tilapia (right
column). (A)Low magnification cross-section of carp
operculum. White bracket indicates the extent of the
disorganized (ʻwovenʼ bone) mid-region. (B)Low magnification
cross-section of tilapia operculum. Black bracket indicates the
extent of the layered region. (C)High magnification cross-
section of carp operculum. Black arrows point to osteocyte
lacunae. Note the woven bone region on the right of the
image. (D)High magnification cross-section of tilapia
100µm 100µm operculum. (E)High magnification flat medial surface of the
carp operculum, showing osteocyte lacunae (black arrow) and
C D their network of canaliculi (black arrowheads). (F)Axial cross-
section of a rib from a tilapia, demonstrating outer layered
zones, sandwiching a woven interior zone. (G)Cross-section
of carp operculum stained with H&E, with osteocyte nuclei
visible within lacunae (black arrows). (H)Cross-section of
tilapia operculum stained with H&E. Note the absence of cells.
The structural layering arrangement can be clearly seen.
(I)Cross-section of carp rib stained with H&E. Osteocyte
nuclei can be clearly seen within lacunae (black arrows).
25µm 25µm
(J)Cross-section of tilapia rib stained with H&E. Note the
layered structure and absence of cells.
E F

25µm 100µm

G H

50µm 50µm

I J

50µm 50µm

tilapia than in carp. A similar layering of tissue zones was also and with a much higher density than the osteocytic lacunae in the
observed in frontal (i.e. perpendicular to the long axis) and cellular carp bone. The voids appeared as small ‘pores’ (2–4m
longitudinal cross-sections of ribs of both species, where a less- wide) and/or wavy and occasionally branched ‘lines’ (~15–40m
organized inner zone (sometimes perforated by the central lumina long) (Fig.4); we took these to be the same structural features, but
of the ribs) was surrounded by an outer zone of concentrically viewed in different orientations, and therefore interpret the structures
ordered layers (Fig.3F). as tapered or blind-ended tubules. As tubules were only visible in
When samples of opercula of both species were polished on their median (and not coronal or transverse) sections of opercula, we
medial or lateral faces (i.e. perpendicular to the sections mentioned verified them to be consistent structural features and not preparation
above), no distinct tissue layering or zonal arrangement was visible artifacts by examining multiple additional tissue sections from
(Fig.4). However, on these polished surfaces, we consistently saw individuals of both species, prepared using a variety of methods
dense collections of what appeared to be small voids in the tissue, (embedded vs non-embedded samples, polished vs unpolished, etc.),
similar in size and density in the two bone types and typically smaller but the appearance described above remained consistent. Although

THE JOURNAL OF EXPERIMENTAL BIOLOGY


1988 L. Cohen and others

Fig.4. Tubules seen in the paramedian plane of


A B the opercular bone of carp (A,B) and tilapia (C,D),
viewed under reflected light microscopy
(unstained). Arrows in each image indicate
individual tubules; however, there are many
tubules visible in each panel. All images show
predominantly end-on views of tubules except D,
where some tubules are visible in a more
longitudinal view (the sample is somewhat
transparent and so structures are visible beneath
the surface). Higher magnification views of
morphologies shown in C and D are presented in
200µm 20µm the insets in those images. All of these
morphologies were seen on opercular samples
C D polished from both lateral (A,D) and medial (B,C)
sides and also in ribs. Note the different zones of
tissue orientation/organization in A, and also how
tubule openings follow the tissue contours around
the vessel in the center of B, supporting our
conclusion that these are structural features of
both bone types and not artifacts.

10µm 200µm 10µm 20µm

opercula polished in the median plane showed no discrete tissue between the BMD of the operculum and rib of carp (Fig.5), although
zones, tubules appeared to clump spatially according to their there was a tendency towards a lower BMD in the operculum
orientation, with clusters of pores and clusters of wavy lines (P0.09).
appearing to localize separately (Fig.4A). These clusters did not
seem to be linked to the medio-lateral position in the operculum Youngʼs modulus
(e.g. ‘pores’ did not appear to be localized only in the outer, lamellar The computed Young’s moduli of the opercula and ribs of both
layer). Similar tubules were observed in rib samples as well but types of fish (tilapia and carp) are presented in Fig.6. The mean
with a lower density than seen in opercula. value of Young’s moduli obtained from the operculum samples
reveals that the stiffness of the acellular bones of tilapia is similar
Osteocyte distribution and density to that of the cellular bones of the carp, with values ranging between
Osteocytes can be seen within lacunae scattered within the sections 3.5 and 10.87GPa. In the carp, the Young’s moduli of the operculum
obtained from the carp operculum and rib (black arrows in Fig.3G,I). bone in the axial and dorso-ventral directions were not statistically
Osteocyte lacunae showing clearly stained nuclei were counted in different (medians of 4.1 and 5.7GPa, respectively, and ranges of
several randomly selected cross-sections of four different carp 3.6–10.0 and 3.9–6.8GPa, respectively, P0.29), suggesting that this
and yielded a mean (±standard deviation) density of bone is transversely isotropic. In the tilapia opercula, the median
424±33osteocytesmm–2. There are no adequate data for lacunar Young’s modulus in the axial direction (7.7GPa, range
shapes and sizes for fishes, but if we assume that they are similar 5.3–10.9GPa) was approximately 40% higher than in the dorso-
to those seen in mammals [i.e. elliptical in shape with major and ventral direction (5.8 GPa, range 5.1–6.0GPa).
minor diameters of 10 and 5m, respectively (McCreadie et al., The rib sections were loaded in compression along the long axis
2004; Hannah et al., 2010)] the area fraction taken by the osteocytic of the rib. The median values of Young’s modulus obtained
lacunae in carp bones is 1.6%. The osteocytic lacunae (black arrows (8.6GPa for tilapia, range 4.1–11.0GPa, and 7.5GPa for carp, range
in Fig.3C,E) in the bones of the carp show distinct canalicular 3.6–14.5GPa) were approximately 40% higher than those found for
networks interconnecting adjacent lacunae (black arrowhead in the dorso-ventral orientation of the operculum.
Fig.3E); however, their distribution is less dense than in mammals
(R.S., personal observation) (Skedros, 2005; Li et al., 1991). Water, organic and ash content
The dry bone ash content of opercula in tilapia (64.49%) is higher
BMD than that of carp (57.58%), in agreement with the values reported
The BMD of carp opercula, determined by micro-CT, was above for bone mineral density (Fig.7A). Water, ash and organic
significantly (P0.001) lower than the BMD of the tilapia opercula content relative to lipid-free wet mass were 14.2%, 49.4% and
(median 766mgcm–3, range 735–824mgcm–3 and median 36.4%, respectively, for carp, and 14.0%, 55.5% and 30.5%,
912mgcm–3, range 883–963mgcm–3, respectively, see Fig.5). respectively, for tilapia (Fig.7B).
Similarly, the BMD of carp rib was significantly (P0.001)
lower than the BMD of tilapia rib (median 785mgcm–3, DISCUSSION
range 737–837mgcm–3 and median 1066mgcm–3, range In this study we compared the material composition (BMD; ash,
1038–1102mgcm–3, respectively, see Fig.5). In tilapia, the BMD water and organic content), structure and mechanical properties
of the operculum was significantly (P0.002) lower than the BMD between the acellular opercula and ribs of a neoteleost fish (O.
of the rib (Fig.5), while no significant difference was observed aureus) and those of a more basal teleost (the carp C. carpio) whose

THE JOURNAL OF EXPERIMENTAL BIOLOGY


Structure and mechanics of acellular bone 1989

1200 12
A
Rib
1100 10

8
BMD (mg cm–3)

1000 Operculum
6
900
Rib Operculum 4
800 2

Young’s modulus (GPa)


700 0
Carp Carp Tilapia Tilapia
axial DV axial DV
600
Carp Tilapia Fish species and beam orientation
Fish species

Fig.5. Box plot of bone mineral density (BMD) values in carp and tilapia
16
B
14
opercula and ribs.
12
10
bones are cellular. Whereas material composition was shown to vary 8
significantly between cellular and acellular bone in the species tested 6
here, and gross structure to vary to some degree, the mechanical 4
properties were shown to be largely similar (with the exception of 2
the orientation-dependent moduli of tilapia opercula).
0
Structurally, the tissues are easily visibly distinguished from one Carp Tilapia
another by the presence and absence of osteocytes, but are more Fish species
similar to each other in gross, histological morphology than they
are to the more-studied cellular bone types of amniotes. In particular, Fig.6. Youngʼs moduli of (A) carp and tilapia operculum bone in the dorso-
differences in osteocyte presence and associated canalicular density ventral (DV) and axial directions (from bending tests), and of (B) carp and
tilapia fifth rib bone (from compression tests).
are striking, even for the cellular boned carp where osteocytes are
much less densely packed than those of birds and mammals (424
vs 700–1800cellsmm–2) (Li et al., 1991; Skedros, 2005). The low
osteocyte density, coupled with the apparent relative scarcity of bone in carp (Bigi et al., 2000) and studies of intramuscular fish
blood vessels in both tilapia and carp bone, make both fish bone bone (mineralized myoseptal tendons) (e.g. Lee and Glimcher, 1991;
types observed here appear, at lower magnifications, as more solid Zhou et al., 2007) and scales (made of acellular bony tissue in both
(less porous) blocks of tissue compared with amniote bone, also cellular and acellular boned species, yet perhaps dental in
suggesting they may have a lower metabolic rate. Our data and evolutionary origin) (e.g. Meunier, 1987; Bigi et al., 2000; Sire and
published work on medaka (Ekanayake and Hall, 1988) point to a Huysseune, 2003) point to structural systems in fish bone types that
relative lack of blood vessels in teleost bone; however, Moss reported are similar to but simpler than amniote bone, where tissue layers
extremely vascular regions in both acellular and cellular bone (Moss, differ in their collagen orientation (Lee and Glimcher, 1991; Currey,
1963) and so it remains unclear whether fish bone as a rule is less 2008). The visible evidence of tissue organization in the ribs and
vascular than amniote bone. opercula of the two species in our study provides a roadmap for
Although lacking the rich cellular patterning of most amniote future biological and material science examinations into fundamental
bones, carp and tilapia bones still exhibited pronounced structural questions of growth, mineralization and organization in fish bones.
features. The obvious tissue zones, viewed in coronal/transverse The observed interspecific differences in tissue layers (e.g. the
cross-sections of opercula and ribs, suggest differences in underlying thicker woven region in carp opercula relative to that of tilapia) also
collagen organization. In particular, the disorganized central (woven) offer a valuable opportunity for testing and modeling basic tissue-
region may be the remains of the earliest primary bone formed during specific structure–function relationships.
development, whereas the laminated lateral and medial layers point The dense network of narrow tubules (2–4m wide, ~15–40m
to a later process of layered tissue deposition in both types of bone. long), a major structural feature observed under reflected light
This finding agrees with those of Moss (Moss, 1960; Moss, 1961), microscopy in ground and polished sections, remains more difficult
who described fish bone (both cellular and acellular) as a mixture to interpret (Fig.4). These structures do not appear to be artifacts,
of lamellar and woven bone, and is further supported by the common as we observed them in rib and opercular bones of both species
use of opercular annuli to age post-hatch individuals of many species (albeit with a higher density in opercular bone) and under multiple
of bony fishes (e.g. Frost and Kipling, 1959; Vilizzi and Walker, forms of sample preparation. However, their structural interpretation
1999; Khan and Khan, 2009). The observed histological differences and the extent to which they perforate the thickness of the skeletal
therefore likely reflect underlying structural patterning, as collagen element are not clear. The tubules are apparently not associated with
organization has been shown to play a role in mineral platelet osteocytes or vascular tissue, and bear strong resemblance to the
packing, size and distribution in vertebrate bones (Weiner et al., canaliculi of Williamson [~2–4m wide canals with an apparent
1999; Currey, 2008; Bigi et al., 2000). It is unclear to what degree nutritive function in the cellular dermal skeletons and endoskeletons
fish endoskeletal bones exhibit the hierarchical structural of basal bony fishes (Sire and Meunier, 1994)] and endoskeletal
organization of amniote bony tissues, but one study of endoskeletal tubules (~2–8m wide) observed in the bones of several teleost

THE JOURNAL OF EXPERIMENTAL BIOLOGY


1990 L. Cohen and others

68 A B
66

% Ash content of dry bone


64

62

60

58

56
Carp Tilapia

Fig.7. (A)Ash content (percentage by mass) of the operculum of carp and tilapia. (B)Ternary diagram of the constituents of various vertebrate skeletal
elements, compiled and modified from Currey (Currey, 2002) by Macesic and Summers (Macesic and Summers, 2012), and including data from our study.
The diagram shows bone from terrestrial vertebrates (gray circles), teleost fish (black circles) and calcified cartilage (white circles) from the propterygia
(pelvic skeletal elements composed of an unmineralized cartilage core covered by a mineralized rind) and vertebral column of elasmobranch fishes (sharks,
rays and relatives). Relative proportions were determined by percentage wet mass. Comparative data were obtained from previous publications (Currey,
2002; Macesic and Summers, 2012; Porter et al., 2006; Toppe et al., 2007; Zioupos et al., 2000).

species and containing thin Sharpey’s-like fibers, probably periosteal Pellegrino, 1969; Lees, 1987a; Szpak, 2011; Toppe et al., 2007)
in origin (Kölliker, 1857) [figs11,12 in Moss (Moss, 1961); figs4–6 (Fig.7B). This supports theoretical conjectures that mineral is less
in Hughes et al. (Hughes et al., 1994)]. It is possible also that the tightly packed among collagen fibrils in fish bone relative to
features we observed are the bundles of ‘uncalcified collagen fibers’ mammalian bone (Lees et al., 1984; Lees, 1987a; Lees, 1987b), but
described by Moss (Moss, 1960; Moss, 1963) [which may also be the reasons for the differences among fish species are unclear; a
the ‘T-fibers’ of Hughes et al. (Hughes et al., 1994)], anchored much wider array of species must be studied to understand this
directly in the matrix but not housed in tubes; however, the pored feature in a phylogenetic context and studies correlating bulk tissue
end-on perspective of these structures argues they are tubular in mineral density and nanoscale mineral packing would be very
nature (e.g. Fig.4B,C). We could not verify the contents or detail instructive.
the structure and distribution of these tubules; however, their As the BMD of cellular fish bone is significantly lower than that
presence in both an acanthopterygian fish with acellular bones of acellular fish bone, we expected Young’s modulus to also be
(tilapia) and an ostariophysian fish with cellular bones (carp), and much lower. The Young’s modulus of bone is determined by its
the reported presence of similar structures in bones/scales of basal composition (mostly mineral density and porosity) and by its
cellular bony fishes and several other species of acellular microarchitecture. However, counter to our predictions, values of
acanthopterygian fishes, suggests these are extremely widespread Young’s moduli of bone material obtained from samples of ribs
structural features. A three-dimensional analysis of tissue patterning and opercula from the two fish types tested here were quite similar:
would provide fundamental insight into the material organization for both cellular and acellular fish bone, median Young’s modulus
of fish bone, and the simple shape of the opercula and ribs make values were found to be in the range 4.1–8.5GPa. The reason for
them ideal skeletal elements for examination. Studies are currently the similarity of material stiffness in bones with such a large
underway to examine these features in three dimensions using high- difference in mineral density is not clear, but surely relates to some
resolution X-ray imaging and scanning electron microscopy. degree to the organization and interactions of mineral and organic
Despite gross structural similarities between the bone types here components in the tissues (Fig.7B).
(aside from cell presence or absence), the acellular operculum and There are very few studies reporting the material properties of
ribs of tilapia have a much higher mineral density (by as much as fish bones; however, the values for Young’s moduli observed in
25%) and ash content (by about 10%) than the cellular bones of the this study are within the range reported (3.7–8.4GPa) recently for
carp. The lower mineral density of cellular bone could be attributed the acellular rib bone of another fish species, the great sculpin,
to its higher porosity due to the presence of lacunae (Currey, 2008); Myoxocephalus polyacanthocephalus (Horton and Summers, 2009)
however, the low percentage volume occupied by the osteocytic (Fig.8A). Such values are lower than those reported for the cortical
lacunae in the carp [estimated in our study to be around 1.6%, bones of birds and mammals. For instance, Young’s modulus of
compared with ~5–9% in amniotes (Currey, 1988)] makes such an equine cortical bone was reported to be between 17.4 and 23.5GPa
explanation unlikely. A literature review of percentage ash values in the axial direction, and between 9.3 and 13.9GPa in the transverse
for acellular and cellular fish bone (N>20 species for each bone direction (Shahar et al., 2007). Currey reported on the Young’s
type; M.D., J. W. C. Dunlop and R.S., in preparation) suggests that modulus of a wide range of amniote bones (but no fish), which were
a lower mineral content is a consistent feature of cellular bone (~50% all tested by a similar protocol, and with the exception of antlers
ash) relative to acellular bone (~58% ash); these differences are also (which are particularly compliant because of their exceptionally low
evident in Toppe et al.’s (Toppe et al., 2007) comparison of several mineral content), the range of values reported fell between 14.2 and
acellular and cellular species. Furthermore, compared with amniote 28.2GPa (Currey, 1999). Even in 4week old mice, which could be
bones, which have ~61–74% dry ash content (Currey, 2008; Szpak, expected to have low moduli as a result of their young age and
2011), fish bones are in general less well mineralized (Biltz and consequent low mineral content, the axial Young’s modulus of

THE JOURNAL OF EXPERIMENTAL BIOLOGY


Structure and mechanics of acellular bone 1991

Kranenbarg et al., 2005a


A Cellular Acellular
Species with material data

Polypteriformes
17.6 GPaB 1
P. senegalensis
Acipenseriformes
15.5 GPaI 2
Lepisosteiformes C. harengus

Amiiformes 8.51 GPaB,C I 3,4


Actinopterygii
C. carpio
Osteoglossomorpha
8.50 GPaI 5,6
Elopomorpha
D. rerio

Teleostei 16.68 GPaH 7


Clupeomorpha
O. mykiss
B Unpublished
tilapia data Ostariophysi 14.30 GPaH 7
14

S. trutta
12 Protacanthopterygii
Modulus (GPa)

10
7.20 GPaB,C 3
Inc.
bichir Basal neoteleosts
data
8 O. aureus

6
Paracanthopterygii
Without Euteleostei 6.48 GPaB 8
bichir data
4
Acanthopterygii
Comp./ Hard./
bend. indent. Neoteleostei M. polyacanthocephalus

Fig.8. (A)Phylogeny of actinopterygian fishes detailing the current state of knowledge of the phylogenetic distribution and material properties of cellular (red)
and acellular (green) bone. Pie charts at the tips of the phylogeny represent the relative proportions of cellular and acellular boned species sampled for each
group, as reported from a literature consensus by Kranenbarg et al. (Kranenbarg et al., 2005a). Pie chart size is scaled according to the number of species
with available data, not the overall number of species in each group (e.g. Kranenbarg et al. collected data for 281 acanthopterygian fishes but that group
contains >10,000 species); however, the differences in pie chart size among groups are roughly representative of group size (e.g. there are many more
acanthopterygian species than polypteriform species). The eight species for which material properties data have been reported in the literature are listed in
the right-hand column, colored according to bone type and with blue dashed arrows indicating the phylogenetic groups to which they belong. Mean moduli
are presented inside each fish icon with superscript letters representing test method (B, bending; C, compression; I, indentation; H, hardness) and numbers
representing literature sources (see below). Mean moduli per species have been, in many cases, averaged from a variety of skeletal elements (e.g. ribs,
opercula, vertebrae, intramuscular bones) and tissue orientations, and from data collected by non-equivalent methods (e.g. nanoindentation and bending);
values are therefore provided only to show the approximate range and order of magnitude of existing data for fish bone. We strongly urge readers to consult
the original works and not cite these values directly. (B)Means and standard deviations for bone moduli versus testing method (comp./bend.,
compression/bending; hard./indent., hardness/indentation) for the eight examined cellular and acellular boned species. Many of the cellular bone samples
were tested dry and by hardness/indentation, therefore likely over-estimating the average modulus for cellular bone relative to acellular bone. This inflation
of modulus is emphasized by nanoindentation data for tilapia opercula (shown; A.S., unpublished data), which are approximately two times larger than the
modulus we report in three-point bending. Three-point bending data for bichir (Polypteriformes) pectoral girdle (cellular) report a particularly high Youngʼs
modulus (17.6GPa) and standard deviation (7.8GPa), suggesting either abnormal properties relative to cellular bone from other fishes or that three-point
bending was an improper test for a skeletal element with such a non-uniform shape; we therefore show the effect of removing this value from the cellular
bone data for bending/compression. Pie charts and phylogeny modified from fig.1 of Kranenbarg et al. (Kranenbarg et al., 2005a), with tree topology
adjusted to reflect the current hypothesis of sister group relationship between Clupeomorpha and Ostariophysi (e.g. Zaragüeta-Bagils et al., 2002).
References for material property data: 1(Erickson et al., 2002); 2(Rho et al., 2001); 3this study; 4(Roy et al., 2000); 5(Zhang et al., 2002); 6(Wang et al.,
2002); 7(Paxton et al., 2006) – we estimate Youngʼs modulus from their microhardness data using an equation from Rapoff et al. (Rapoff et al., 2008);
8
(Horton and Summers, 2009).

cortical bone was reported to be 8.9±1.1GPa (Lev-Tov Chattah et measured in the rib. In tilapia (acellular bone), the axial modulus
al., 2009). in the rib and the modulus in the antero-posterior orientation in
The relatively low stiffness of fish bone is likely related to its the transverse plane of the operculum were similar, and much
comparatively low mineral content, but whether the lower stiffness higher (by about 50%) than the modulus in the dorso-ventral
in fact confers a functional advantage on fishes (e.g. whether more orientation in the transverse plane of the operculum. These
compliant bone is better suited to fish ecologies) is unknown. findings suggest that there may be an inherent difference between
Fish bone stiffness is probably also related to some unique features the hierarchical architecture of acellular and cellular bone, at least
of its microarchitecture. In the carp (cellular bone), the moduli in the bones of these two fishes, beyond the presence/absence of
in orthogonal directions within the median plane of the operculum osteocytes and their lacunae. We believe this will be clarified
were similar, and substantially lower than the axial modulus through analyses of the ontogeny of tissue mineralization and of

THE JOURNAL OF EXPERIMENTAL BIOLOGY


1992 L. Cohen and others

crystal orientation in species with cellular and acellular bone Our data suggest that fish bone has fundamental structural,
types. It would be of particular interest to examine how tissue compositional and material performance differences relative to other
material properties and mineral microstructure change in fish bony tissues; yet it remains unclear what selective pressures might
bones that show large-scale ontogenetic changes or spatial grades have led to the loss of osteocytes in some fish groups. We are
in the degree of cellularity (Moss, 1961; Moss, 1963; Witten et currently exploring the microstructure of cellular and acellular fish
al., 2001; Witten and Huysseune, 2009). bone in multiple species. These analyses will allow us to test the
It is important to note that material property data for bone are hypothesis that differences in fibril structure and fibril-array
highly dependent on sample preparation and testing techniques, and arrangement are responsible for the differences found between fish
therefore comparisons among studies that used different bone types (i.e. mineral density and isotropy), and between fish bone
methodologies may not be valid. To the best of our knowledge, and that of other vertebrates (e.g. low stiffness and cellularity). To
only eight studies (including ours) have performed mechanical tests provide a useful phylogenetic perspective and allow accurate
aimed at determining material properties of fish bone, and only two comparisons of mechanical property data for acellular and cellular
of these [ours and that of Horton and Summers (Horton and fish bone, it would be valuable for future work on fish bone to
Summers, 2009)] have focused on acellular bone (see Fig.8). standardize material testing methodology and, if possible, to choose
Ignoring for a moment the differences in sample orientation (i.e. test species according to those phylogenetic groups for which there
axial vs transverse tests), the skeletal element examined and are existing data. It would be particularly useful to see comparisons
phylogeny (Fig.8A), the combined data from these studies suggest within the Perciformes, which have many species of convenient size
that cellular bone, on average, is stiffer than acellular bone (11.4±2.6 for testing and, although largely acellular boned, also contain several
vs 7.0±1.3GPa) (Fig.8B). However, the majority of studies reportedly cellular-boned species within the Scombrinae (Kölliker,
examining cellular bone tested dried samples using nanoindentation, 1857; Moss, 1961). The Protocanthopterygii also represent a
a technique for measuring nanoscale material moduli. Not only is particularly interesting group for study as they possess nearly equal
the indentation modulus often higher for dry than for hydrated numbers of cellular- and acellular-boned species (Moss, 1961;
biological samples (Guidoni et al., 2006; Enders et al., 2004) but Kranenbarg et al., 2005a; Kranenbarg et al., 2005b).
also it is unclear how it can be compared with Young’s moduli In the comparison of carp and tilapia bone, the lack of cells does
obtained by common techniques used to test materials at the not seem to confer an obvious material performance advantage
micrometer scale, such as bending, tension and compression tests relative to cellular fish bones; as such, the cause of the evolutionary
(Hengsberger et al., 2003). When all available fish material property development towards acellularity in neoteleost fish bone remains
data are plotted according to testing method, the values generated unclear. An explanation previously offered was that acellularity
by nanoindentation (i.e. the majority of data for fish cellular bone) could increase the stiffness of fish bone, which is low relative to
are higher and show a wider range of variation (Fig.8B). This effect that of mammalian bone (Horton and Summers, 2009). Our data
of testing method is underlined by nanoindentation data of carp (Roy support Horton and Summers’ rejection of this hypothesis because
et al., 2000), which provided an indentation modulus more than twice the bones of the more basal teleost, which are cellular, were shown
our reported Young’s modulus in compression from axial tests, and here to have a similar stiffness to that of acellular bones from a
by nanoindentation data of the same tilapia opercula used in this neoteleost, despite the presence of osteocytic lacunae. Alternatively,
study but tested dry (A.S., unpublished data), which provided an the primary function of osteocytes may be osteocytic osteolysis
indentation modulus of 14.5±1.5GPa (Fig.8B), approximately twice (Clark et al., 2005; Bonucci, 2009; Witten and Huysseune, 2009;
as high as the Young’s modulus we report from three-point bending Bonewald, 2011). As fish are unlikely to face calcium deficiency
of the same tissue. Based on the likelihood that existing data for given the high availability of minerals in their natural habitat (both
fish bone material properties are test dependent and on our marine and fresh water), selective pressure would have favored the
comparison of tilapia and carp bone, we posit that cellular and loss of osteocytes, as the high metabolic cost of their maintenance
acellular bone are, in general, equivalent from the standpoint of would outweigh the benefit. The loss of osteocytes in fishes then
material stiffness. implies that the other functions typically performed by these cells
Our study presents the first comparison of acellular and cellular are either of lesser importance to fish bone or are addressed through
fish bone materials, performed by uniform methodology on samples as yet unidentified mechanisms.
of standardized size. Yet, although these species provide valuable
reference points for acellular and cellular bone, they are quite ACKNOWLEDGEMENTS
phylogenetically disparate (Fig.8A), with one a derived perciform The authors wish to thank Prof. John Currey for critically reviewing the manuscript
neoteleost (tilapia: Acanthopterygii) and the other a more basal and making many useful comments, Dr Ofer Ashulin and Prof. Berta Sivan for
help with obtaining samples, Dr Joshua Milgram for help with sample preparation
teleost (carp: Ostariophysi). In fact, the scant data on fish bone and Dr Gilad Segev for help with statistical analysis.
material properties have been drawn from a very wide phylogenetic
sampling, made up of one polypteriform fish [Polypterus FUNDING
senegalensis (Erickson et al., 2002)] and one to two representatives This work was supported in part by an Alexander von Humboldt Post-doctoral
each from four superorders of Teleostei: Clupeomorpha [Clupea Fellowship to M.D.
harengus (Rho et al., 2001)], Ostariophysi [Danio rerio (Wang et
al., 2002; Zhang et al., 2002); Cyprinus carpio (this study) (Roy et REFERENCES
al., 2000)], Protacanthopterygii [Oncorhynchus mykiss, Salmo trutta Adachi, T., Aonuma, Y., Ito, S.-I., Tanaka, M., Hojo, M., Takano-Yamamoto, T. and
Kamioka, H. (2009). Osteocyte calcium signaling response to bone matrix
(Paxton et al., 2006)] and Acanthopterygii [Myoxocephalus deformation. J. Biomech. 42, 2507-2512.
polyacanthocephalus (Horton and Summers, 2009); Oreochromis Bigi, A., Koch, M., Panzavolta, S., Roveri, N. and Rubini, K. (2000). Structural
aureus (this study)]. All of the studied species are freshwater fishes aspects of the calcification process of lower vertebrate collagen. Conn. Tiss. Res.
41, 37-43.
except M. polyacanthocephalus and C. harengus, and the two Biltz, R. M. and Pellegrino, E. D. (1969). The chemical anatomy of bone: I. A
acanthoptherygian species are the only acellular boned fish that have comparative study of bone composition in sixteen vertebrates. J. Bone Joint Surg.
51, 456-466.
been examined (Fig.8). Bonewald, L. (2011). The amazing osteocyte. J. Bone. Miner. Res. 26, 229-238.

THE JOURNAL OF EXPERIMENTAL BIOLOGY


Structure and mechanics of acellular bone 1993
Bonucci, E. (2009). The osteocyte: the underestimated conductor of the bone Macesic, L. J. and Summers, A. P. (2012). Flexural stiffness and composition of the
orchestra. Rend. Lincei Sci. Fis. Nat. 20, 237-254. batoid propterygium as predictors of punting ability. J. Exp. Biol. 215 (in press).
Clark, W. D., Smith. E. L., Linn. K. A., Paul-Murphy, J. R., Muir, P. and Cook, M. E. McCreadie, B. R., Hollister, S. J., Schaffler, M. B. and Goldstein, S. A. (2004).
(2005). Osteocyte apoptosis and osteoclast presence in chicken radii 0-4 days Osteocyte lacuna size and shape in women with and without osteoporotic fracture. J.
following osteotomy. Calcif. Tissue Int. 77, 327-336. Biomech. 37, 563-572.
Currey, J. (1988). The effect of porosity and mineral content on the Youngʼs modulus Meunier, F. J. (1987). Nouvelles données sur lʼorganisation spatiale des fibres de
of elasticity of compact bone. J. Biomech. 21, 131-139. collagène de la plaque basale des écailles des Téléostéens. Ann. Sci. Nat. Zool. 9,
Currey, J. D. (1999). The design of mineralized hard tissues for their mechanical 113-121.
functions. J. Exp. Biol. 202, 3285-3294. Moss, M. L. (1960). Osteogenesis and repair of acellular teleost bone. Anat. Rec. 136,
Currey, J. D. (2002). Bones, pp. 436. Princeton: Princeton University Press. 246-247.
Currey, J. D. (2008). Collagen and the mechanical properties of bone and calcified Moss, M. L. (1961). Studies of the acellular bone of teleost fish: 1. Morphological and
cartilage. In Collagen: Structure and Mechanics (ed. P. Fratzl), pp. 397-420. New systematic variations. Acta Anat. 46, 343-462.
York: Springer. Moss, M. L. (1963). The biology of acellular teleost bone. Ann. N. Y. Acad. Sci. 109,
Currey, J. D. (2010). Mechanical properties and adaptations of some less familiar 337-350.
bony tissues. J. Mech. Behav. Biomed. Mater. 3, 357-372. Nemoto, Y., Higuchi, K., Baba, O., Kudo, A. and Takano, Y. (2007). Multinucleate
Draper, E. R. C. and Goodship, A. E. (2003). A novel technique for four-point bending osteoclasts in medaka as evidence of active bone remodeling. Bone 40, 399-408.
of small bone samples with semi-automatic analysis. J. Biomech. 36, 1497-1502. Nicolella, D. P., Moravits, D. E., Gale, A. M., Bonewald, L. F. and Lankford, J.
Ekanayake, S. and Hall, B. K. (1988). Ultrastructure of the osteogenesis of acellular (2006). Osteocyte lacunae tissue strain in cortical bone. J. Biomech. 39, 1735-1743.
vertebral bone in the Japanese Medaka, Oryzias latipes (TeIeostei, Cyprinidontidae). Parenti, L. R. (1986). The phylogenetic significance of bone types in euteleost fishes.
Am. J. Anat. 182, 241-249. Zool. J. Linn. Soc. 87, 37-51.
Enders, S., Barbakadse, N., Gorb, S. and Arzt, E. (2004). Exploring biological Paxton, H., Bonser, R. H. C. and Winwood, K. (2006). Regional variation in the
surfaces by nanoindentation. J. Mater. Res. 19, 880-887. microhardness and mineralization of vertebrae from brown and rainbow trout. J. Fish
Erickson, G. M., Catanese, J., 3rd and Keaveny, T. M. (2002). Evolution of the Biol. 68, 481-487.
biomechanical properties of the femur. Anat. Rec. 268, 115-124. Porter, M. E., Beltrán, J. L., Koob, T. J. and Summers, A. P. (2006). Material
Franz-Odendaal, T. A., Hall, B. K. and Witten, P. E. (2006). Buried alive: how properties and biochemical composition of mineralized vertebral cartilage in seven
osteoblasts become osteocytes. Dev. Dyn. 235, 176-190. elasmobranch species (Chondrichthyes). J. Exp. Biol. 209, 2920-2928.
Frost, W. E. and Kipling, C. (1959). The determination of the age and growth of pike Rapoff, A. J., Rinaldi, R. G., Hotzman, J. L. and Daegling, D. J. (2008). Elastic
(Esox lucius L.) from scales and opercular bones. ICES J. Mar. Sci. 24, 314-341. modulus variation in mandibular bone: a microindentation study of Macaca
Guidoni, G., Denkmayr, J., Schöberl, T. and Jager, I. (2006). Nanoindentation in fascicularis. Am. J. Phys. Anthropol. 135, 100-109.
teeth: influence of experimental conditions on local mechanical properties. Philos. Rho, J. Y., Mishra, S. R., Chung, K., Bai, J. and Pharr, G. M. (2001). Relationship
Mag. 86, 5705-5714. between ultrastructure and nanoindentation properties of intramuscular herring
Hannah, K. M., Thomas, C. D. L., Clement, J. G., De Carlo, F. and Peele, A. G. bones. Ann. Biomed. Eng. 29, 1082-1088.
(2010). Bimodal distribution of osteocyte lacunar size in the human femoral cortex as Roy, M. E., Nishimoto, S. K., Rho, J. Y., Bhattacharya, S. K., Lin, J. S. and Pharr,
revealed by micro-CT. Bone 47, 866-871. G. M. (2000). Correlations between osteocalcin content, degree of mineralization and
Hegrenes, S. (2001). Diet-induced phenotypic plasticity of feeding morphology in the mechanical properties of C. carpio rib bone. J. Biomed. Mater. Res. 54, 547-553.
orangespotted sunfish, Lepomis humilis. Ecol. Freshwater Fish. 10, 35-42. Shahar, R., Zaslansky, P., Barak. M., Friesem, A. A., Currey, J. D. and Weiner, S.
Hengsberger, S., Enstroem, J., Peyrin, F. and Zysset, P. (2003). How is the
(2007). Anisotropic Poissonʼs ratio and compression modulus of cortical bone
indentation modulus of bone tissue related to its macroscopic elastic response? A
determined by speckle interferometry. J. Biomech. 40, 252-264.
validation study. J. Biomech. 36, 1503-1509.
Sire, J.-Y. and Huysseune, A. (2003). Formation of dermal skeletal and dental tissues
Horton, J. M. and Summers, A. P. (2009). The material properties of acellular bone in
in fish: a comparative and evolutionary approach. Biol. Rev. 78, 219-249.
a teleost fish. J. Exp. Biol. 212, 1413-1420.
Sire, J.-Y. and Meunier, F. J. (1994). The canaliculi of Williamson in holostean bone
Hughes, D. R., Bassett, J. R. and Moffat, L. A. (1994). Histological identification of
(Osteichthyes, Actinopterygii): a structural and ultrastructural study. Acta Zool. 75,
osteocytes in the allegedly acellular bone of the sea breams Acanthopagrus
235-247.
australis, Pagrus auratus and Rhabdosargus sarba (Sparidae, Perciformes,
Skedros, J. G. (2005). Osteocyte lacuna population densities in sheep, elk and horse
Teleostei). Anat. Embryol. 190, 163-179.
calcanei. Cells Tissues Organs 181, 23-37.
Huysseune, A., Sire, J. Y. and Meunier, F. J. (1994). Comparative study of lower
Spatz, H. C., OʼLeary, E. J. and Vincent, F. V. (1996). Youngʼs moduli and shear
pharyngeal jaw structure in two phenotypes of Astatoreochromis alluaudi (Teleostei:
moduli in cortical bone. Proc. R. Soc. Lond. B 263, 287-294.
Cichlidae). J. Morphol. 221, 25-43.
Khan, M. A. and Khan, S. (2009). Comparison of age estimates from scale, opercular Szpak, P. (2011). Fish bone chemistry and ultrastructure: implications for taphonomy
bone, otolith, vertebrae and dorsal fin ray in Labeo rohita (Hamilton), Catla catla and stable isotope analysis. J. Archaeol. Sci. 38, 3358-3372.
(Hamilton) and Channa marulius (Hamilton). Fish. Res. 100, 255-259. Toppe, J., Albrektsen, S., Hope, B. and Aksnes, A. (2007). Chemical composition,
Kilhara, M., Ogata, S., Kawano, N., Kubota, I. and Yamaguchi, R. (2002). Lordosis mineral content and amino acid and lipid profiles in bones from various fish species.
induction in juvenile red sea bream, Pagrus major, by high swimming activity. Comp. Biochem. Physiol. 146B, 395-401.
Aquaculture 212, 149-158. Totland, G. K., Fjelldal, P. G., Kryvi, H., Lokka, G., Wargelius, A., Sagstad, A.,
Kitamura, K., Suzuki, N., Sato, Y., Nemoto, T., Ikegame, M., Shimizu, N., Kondo, Hansen, T. and Grotmol, S. (2011). Sustained swimming increases the mineral
T., Furusawa, Y., Shigehito, W. and Hattori, A. (2010). Osteoblast activity in the content and osteocyte density of salmon vertebral bone. J. Anat. 219, 490-501.
goldfish scale responds sensitively to mechanical stress. Comp. Biochem. Physiol. Vilizzi, L. and Walker, K. F. (1999). Age and growth of the common carp, Cyprinus
156A, 357-363. carpio, in the River Murray, Australia: validation, consistency of age interpretation,
Kölliker, A. (1857). On the different types in the microscopic structure of the skeleton and growth models. Environ. Biol. Fish. 54, 77-106.
of osseous fishes. Proc. R. Soc. Lond. B 9, 656-668. Wang, X. M., Cui, F. Z., Ge, J., Zhang, Y. and Ma, C. (2002). Variation of
Kranenbarg, S., van Cleynenbreugel, T., Schipper, H. and van Leeuwen, J. L. nanomechanical properties of bone gene mutation in the zebrafish. Biomat. 23,
(2005a). Adaptive bone formation in acellular vertebrae of sea bass (Dicentrarchus 4557-4563.
labrax L.). J. Exp. Biol. 208, 3493-3502. Weiner, S., Traub, W. and Wagner, H. D. (1999). Lamellar bone: structure-function
Kranenbarg, S., Waarsing, J. H., Muller, M., Weinans, H. and van Leeuwen, J. L. relations. J. Struct. Biol. 126, 241-255.
(2005b). Lordotic vertebrae in sea bass (Dicentrarchus labrax L.) are adapted to Witten, P. E and Huysseune, A. (2009). A comparative view on mechanisms and
increased loads. J. Biomech. 38, 1239-1246. functions of skeletal remodeling in teleost fish, with special emphasis on osteoclasts
Lee, D. D. and Glimcher, M. J. (1991). Three-dimensional spatial relationship and their function. Biol. Rev. 84, 315-346.
between the collagen fibrils and the inorganic calcium phosphate crystals of pickerel Witten, P. E, Hansen, A. and Hall, B. K. (2001). Features of mono- and
(Americanus americanus) and herring (Clupea harengus) bone. J. Mol. Biol. 217, multinucleated bone resorbing cells of the zebrafish Danio rerio and their contribution
487-501. to skeletal development, remodeling, and growth. J. Morphol. 250, 197-207.
Lees, S. (1987a). Considerations regarding the structure of the mammalian Zaragüeta-Bagils, R., Lavoué, S., Tillier, A., Bonillo, C. and Lecointre, G. (2002).
mineralized osteoid from viewpoint of the generalized packing model. Conn. Tiss. Assessment of otocephalan and protacanthopterygian concepts in the light of
Res. 16, 281-303. multiple molecular phylogenies. C. R. Biol. 325, 1191-1207.
Lees, S. (1987b). Possible effect between the molecular packing of collagen and the Zhang, Y., Cui, F. Z., Wang, X. M., Feng, Q. L. and Zhu, X. D. (2002). Mechanical
composition of bony tissues. Int. J. Biol. Macromol. 9, 321-326. properties of skeletal bone in gene-mutated Stöpseldtl28d and wild-type zebrafish
Lees, S., Bonar, L. and Mook, H. (1984). A study of dense mineralized tissue by (Danio rerio) measured by atomic force microscopy-based nanoindentation. Bone
neutron-diffraction. Int. J. Biol. Macromol. 6, 321-326. 30, 541-546.
Lev-Tov Chattah, N., Sharir, A., Weiner, S. and Shahar, R. (2009). Determining the Zhou, H., Burger, C., Sics, I., Hsiao, B. S., Chu, B., Graham, L. and Glimcher, M.
elastic modulus of mouse cortical bone using electronic speckle pattern J. (2007). Small-angle X-ray study of the three-dimensional collagen mineral
interferometry (ESPI) and micro computed tomography: a new approach for superstructure in intramuscular fish bone. J. Appl. Crystallogr. 40, S666-S668.
characterizing small-bone material properties. J. Biomech. 45, 84-90. Zioupos, P., Currey, J. D. and Casinos, A. (2000). Exploring the effects of
Li, K., Zernicker, R. F., Barnard, R. J. and Li, A. F. Y. (1991). Differential response hypermineralisation in bone tissue by using an extreme biological example. Conn.
of rat limb bones to strenuous exercise. J. Appl. Phys. 70, 554-560. Tiss. Res. 41, 229-248.

THE JOURNAL OF EXPERIMENTAL BIOLOGY

You might also like