ArticlePDF Available

Modelling the three-dimensional elastic constants of parallel-fibred and lamellar bone

Authors:

Abstract and Figures

The complex hierarchical structure of lamellar bone makes understanding structure–mechanical function relations, very difficult. We approach the problem by first using the relatively simple structure of parallel-fibred bone to construct a mathematical model for calculating Young's moduli in three-dimensions. Parallel-fibred bone is composed essentially of arrays of mineralized collagen fibrils, which are also the basic structural motif of the individual lamellae of lamellar bone. Parallel-fibred bone structure has orthotropic symmetry. As the sizes and shapes of crystals in bone are not well known, the model is also used to compare the cases of platelet-, ribbon- and sheet-reinforced composites. The far more complicated rotated plywood structure of lamellar bone results in the loss of the orthotropic symmetry of individual lamellae. The mathematical model used circumvents this problem by sub-dividing the lamellar unit into a thin lamella, thick lamella, transition zone between them, and the recently observed back-flip lamella. Each of these is regarded as having orthotropic symmetry. After the calculation of their Young's moduli they are rotated in space in accordance with the rotated plywood model, and then the segments are combined to present the overall modulus values in three-dimensions. The calculated trends compare well with the trends in microhardness values measured for circumferential lamellar bone. Microhardness values are, as yet, the only measurements available for direct comparison. Although the model is not directly applicable to osteonal bone, which is composed of many hollow cylinders of lamellar bone, the range of calculated modulus values and the trends observed for off-axis calculations, compare well with measured values. 1998 Chapman & Hall
Content may be subject to copyright.
00222461 ( 1998 Chapman & Hall 1497
JOURNAL OF MATERIALS SCIENCE: 33 (1998) 1497 1509
Modelling the three-dimensional elastic constants
of parallel-fibred and lamellar bone
U. AKIVA, H. D. WAGNER, S. WEINER*
Department of Materials and Interfaces, and
*
Department of Structural Biology, The
Weizmann Institute of Science, Rehovot 76100, Israel
E-mail. cpwagner@wis.weizmann.ac.il
The complex hierarchical structure of lamellar bone makes understanding
structuremechanical function relations, very difficult. We approach the problem by first
using the relatively simple structure of parallel-fibred bone to construct a mathematical
model for calculating Young’s moduli in three-dimensions. Parallel-fibred bone is composed
essentially of arrays of mineralized collagen fibrils, which are also the basic structural motif
of the individual lamellae of lamellar bone. Parallel-fibred bone structure has orthotropic
symmetry. As the sizes and shapes of crystals in bone are not well known, the model is also
used to compare the cases of platelet-, ribbon- and sheet-reinforced composites. The far
more complicated rotated plywood structure of lamellar bone results in the loss of the
orthotropic symmetry of individual lamellae. The mathematical model used circumvents
this problem by sub-dividing the lamellar unit into a thin lamella, thick lamella, transition
zone between them, and the recently observed ‘‘back-flip’’ lamella. Each of these is regarded
as having orthotropic symmetry. After the calculation of their Young’s moduli they are
rotated in space in accordance with the rotated plywood model, and then the segments are
combined to present the overall modulus values in three-dimensions. The calculated trends
compare well with the trends in microhardness values measured for circumferential lamellar
bone. Microhardness values are, as yet, the only measurements available for direct
comparison. Although the model is not directly applicable to osteonal bone, which is
composed of many hollow cylinders of lamellar bone, the range of calculated modulus
values and the trends observed for off-axis calculations, compare well with measured
values.
1998 Chapman & Hall
1. Introduction
Relating the mechanical properties of bone to its
structure is a very challenging task. Bone occurs in
several different forms or types and has a complex
hierarchical structure [1, 2]. Furthermore, bone struc-
ture, and, in particular, mineral content, changes with
increasing age of the bone, and this too has a direct
effect on mechanical properties [1, 3, 4]. Bone struc-
turemechanical relations have, therefore, to be re-
solved for well-defined bone types of a specific age.
Here we address this question first for the relatively
simple parallel-fibred bone, and then use this as
a basis for addressing the far more complex structure,
rotated plywood, that is present in lamellar bone [5].
We approach this by using our current knowledge of
the structure as a basis for constructing a mathematical
model of the material properties of the bone type, and
then compare the calculated results with the few appro-
priate data sets of experimentally measured values.
This approach has been used in several other
studied. Recent examples are the studies of Sasaki et
al. [6], Wagner and Weiner [7] and Currey et al. [8].
Sasaki et al. [6] used fibrolamellar bone which is
composed of both lamellar and parallel-fibred bone.
They measured the preferred orientations of the c-axes
of the crystals and introduced this information into
their model. They, like many earlier studies, assumed
that the crystals are needle-shaped. Direct observa-
tions [913] as well as the modelling studies of Wag-
ner and Weiner [7] and Currey et al. [8], both show
that most, if not all, the crystals are plate-shaped and
that bone should be regarded as a platelet-reinforced
composite material. The structure of the narwhal tusk
studied by Currey et al. [8] is very poorly known, and
this limits the extent to which its structure can be
related to measured mechanical properties.
Parallel-fibred bone is essentially an array of paral-
lel mineralized collagen fibrils [14]. The plate-shaped
crystals within each fibril are arranged in layers that
traverse the collagen fibril [15]. The structure of
lamellar bone has been the subject of many studies,
and was reviewed by Weiner and Traub [2]. Its
basic plywood motif has been well documented
[1619] and a well substantiated structural model for
collagen fibril organization was proposed by Giraud-
Guille [17]. In this model each lamella is composed of
arrays of parallel fibrils. The orientations of the fibril
arrays in adjacent lamellae are different, thus forming
Figure 1 Scanning electron micrograph of the fracture surface of rat
tibia lamellar bone (published previous [22]) showing the lamellar
units (between white bars), each of which is composed of a thin
lamella (1), a transition zone (2), a thick lamella (3) and a ‘‘back-flip’’
lamella’’ (4). We deliberately use a micrograph published previously
[22], to show that the ‘‘back-flip’’ lamella is present, but we did not,
at the time, note its significance. For more information, see [53].
a plywood-like motif [1619]. Note that not all
investigators concur with this view of the lamellar
structure [20]. Weiner et al. [5] observed that the
crystal layers are parallel to the lamellar boundary
plane in the thin lamellae, but are rotated relative to
the lamellar boundary plane in the thick lamella. They
therefore called this structure ‘‘rotated plywood’’. This
constituted the basis for a model that attempted to
relate structure to material properties by Wagner and
Weiner [7]. Our understanding of the rotated ply-
wood structure has subsequently been further refined
to include more information on the structure of the
transition zone between the thick and thin lamellae,
the presence of a so-called ‘‘back-flip’’ lamella and
semi-quantitative measurements of the plywood angle
[21, 22]. The basic features of the structure are shown
in Fig. 1.
The models proposed by Sasaki et al. [6], Wagner
and Weiner [7], Currey et al. [8] and indeed the one
proposed here, all predict the variations in elastic
modulus with changing orientations of the specimen
tested in relation to the long axis of the bone. This is
a fairly stringent test. A serious problem, however, is
the paucity of appropriate experimental measure-
ments of this type. Reilly and Burstein [23], Bonfield
and Grynpas [24] and recently Turner et al. [25]
collected such data for a variety of lamellar bones
and fibrolamellar bone. The lamellar bones used were
osteonal, which complicates the geometry enormously
due to the cylindrical nature of the osteon and its
central blood vessel. Fibrolamellar bone is a mixture
of two bone types [16], making its interpreta-
tions equivocal. Ziv et al. [15] performed a detailed
microhardness study on circumferential lamellar bone
(which is essentially composed of parallel lamellar
units) probed in many different orientations with re-
spect to the rotated plywood structure. Microhard-
ness is a convenient tool for doing this, but unfortu-
nately it is difficult to relate hardness values to most
other material properties of bone. The trends, how-
ever, should be the same. Evans et al. [26] made an
empirical study of microhardnessYoung’s modulus
relations using many different bones. Although a cor-
relation is observed, we hesitate to use this to convert
microhardness values to elastic moduli because the
samples were not oriented relative to their structures
and many different structures are all lumped together.
Here we present a more refined micromechanical
model for the three-dimensional Young’s modulus of
lamellar bone. The present model follows, at first, the
logic of our previous model [7], that is, the focus rests
first on the single lamella (parallel-fibred bone in
essence), then on bone viewed as an assembly of lamel-
lae. However, regarding the latter, the approach taken
here is now fully three-dimensional and eliminates
the need for the arbitrary factor, n, used in our pre-
vious model [7] to account for the weakening effect
induced by the rotation of the crystal layers in the
thick lamellae.
1.1. Bone structure: comments on the key
features incorporated into the model
The basic building block of bone is a mineralized
collagen fibril [27], which is composed of thin, plate-
shaped crystals of carbonated apatite (dahllite) ar-
ranged in parallel layers within the type I collagen
fibril [9]. The crystal thicknesses, as measured by
small-angle X-ray scattering are about 1.53.0 nm [12,
29]. The crystals themselves, once extracted, are very
irregular in shape, but are on average 50 nm long by
25 nm wide [11, 13]. Very little, however, is known
about their in vivo dimensions. As extraction almost
certainly causes breakage, they may be much larger, as
has been suggested by Boyde [30] who studied ion-
beam thinned dentine. We also know very little about
the nature of crystalcrystal contacts in vivo, and
about crystals that may be located between collagen
fibrils. The structure of the type I collagen fibril, to
a large extent, dictates the organization of the crystals
inside the fibril [31]. Although crystals initially form
in the gap zones [3133], they rapidly fill the gap
zones, and then spread into the overlap zones [34]
ultimately to form extended layers. Following the
model for the type I collagen fibril based on
stoichiometric measurements of cross-links [35, 36],
the layers of crystals are separated by four layers of
close-packed triple helical molecules. The distance
separating crystal layers is about 45 nm [2], based
on a fairly prominent 4 nm reflection present in X-ray
diffraction patterns of type I collagen [37] and from
direct measurements [38]. Note too that the propor-
tions of collagen, crystals and water vary in normal
bone from species to species, with the age of the animal,
and within the bone as a function of remodelling.
This has an important effect on almost all material
properties [3, 4]. The relative proportions of collagen,
1498
TABLE I Structural and material parameters of lamellar bone used for the calculations (refer to Figs 1 and 2 for notations)
Structural parameters Reference
Average crystal platelet dimensions 50]25]2nm3 [1, 3]
Distance between crystal layers: d"2 nm [2, 37]
corresponds to 4 nm spacing in structure (d"2nm
and 2]1/2 crystal thickness of 2 nm)
Lateral space between crystals u"0.1 nm (parallel to collagen fibril axis)
(assumed to be very small)
v"0.1 nm (perpendicular to collagen fibril axis)
Calculated (Equation 3) platelet volume fraction
(corresponds to 67% ash weight) »
1
"0.5 [15, 39]
Average lamellar unit 3.2 lm [21]
thickness (rat)
Average thick lamella 1.8 lm [21]
thickness (rat)
Average thin lamella thicknes 0.4 lm! [21]
Average transition zone 0.4 lm! [21]
(between thin and thick lamellae) thickness (rat)
Average thickness of the ‘‘back-flip’’ lamellae 0.6 lm [21]
w
1
and w
2
angles
Thin lamella w
1
"0 w
2
"0 [5]
Transition zone calculated
Thick lamella w
1
"90° w
2
"70" [21]
Back-flip lamella w
1
"120° w
2
"90° [21]
Material stiffness constants
Young’s modulus of mineral E
1
"114 GPa [40]
Shear modulus of mineral G
1
"44.5 GPa [40]
Poisson ratio of mineral m
1
"0.30
(calculated for isotropic platelets)
Young’s modulus of collagen E
.
"1.5 GPa [4]
Shear modulus of collagen G
.
"1 GPa [41]
Collagen’s Poisson’s ratio m
.
"0.38 [39]
(corresponds to m
12
"0.35 for the overall
Poisson’s ratio of the collagenmineral composite
calculated using the rule of mixtures)
! These two zones cannot be differentiated in scanning electron micrographs, so we arbitrarily divided them into two equal halves.
" No data are available for w
2
. Based only on impressions from scanning electron micrographs.
crystals and water more or less follow a regular pat-
tern [39].
Bundles of parallel mineralized collagen fibrils can
be arranged in a variety of ways to form different bone
types and associated tissues such as cementum, den-
tine and mineralized tendon. In this study we focus on
two such bone types: the parallel-fibred bone and
lamellar bone. In terms of the model, we regard paral-
lel-fibred bone as being composed of an array of
mineralized collagen fibrils, aligned not only along
their fibril axes, but also in terms of their crystal layers.
This highly ordered three-dimensional structure
is clearly idealized, and in vivo there is almost cer-
tainly less order particularly in the azimuthal orienta-
tions of the crystal layers [15]. We also use an
idealized rotated plywood structure for modelling
lamellar bone.
An analysis of the basic features of the rotated
plywood structure showed that much of the variation
within the structure can be described in terms of two
angles, the plywood angle, w
1
, and the rotation angle,
w
2
[7]. This relatively simple situation arises because
most of the mineralized collagen fibrils are in planes
parallel to the boundaries between lamellae. The ply-
wood angle, w
1
, describes the extent of offset of the
collagen fibril axes from one layer of collagen fibrils to
the next. The rotation angle, w
2
, is the extent of rota-
tion about the fibril axis from one layer to the next.
Wagner and Weiner [7] arbitrarily defined w
1
"0
when the fibrils are perpendicular to long axes of the
bone, and w
2
"0 when the crystal layers are parallel
to the lamellar boundary plane. Weiner et al. [21]
have measured w
1
angles between associated layers
using decalcified and vitrified thin sections of the rat
femur lamellar bone cut approximately parallel to the
lamellar boundary plane. They proposed a model in
which the w
1
values of a lamellar unit increase in
discrete steps of roughly 30°, such that the thick
lamella fibrils are roughly orthogonal to those of the
thin lamella; a result consistent with other observa-
tions [16, 17]. There also appears to be an additional
lamella in which the w
1
angle is 120°. This ‘‘back-flip’’
lamella borders on the lamellar boundary. In order to
simplify this situation, we propose to use the term
‘‘lamellar unit’’ to include the traditional thin and
thick lamellae, as well as the transition zone between
them and the 120° ‘‘back-flip’’ lamella. The structure
of an individual lamellar unit can be seen in Fig. 1.
Most of the features described above are incorporated
into the theoretical model. The specific values used
are, for the most part, for the midshaft of a rat tibia or
femur and are listed in Table I.
1499
2. Theoretical model
2.1. Elastic constants of parallel-fibred bone
Parallel-fibred bone and an individual lamella of
lamellar bone are here regarded as having the same
structures.
2.1.1. Platelet and ribbon reinforcement
The structure of the collagencrystal composite shows
that it should be modelled as a unidirectional plane-
parallel (UPP) platelet-reinforced composite, rather
than a fibre composite. UPP platelet ribbon- or sheet-
reinforced composites (where a ribbon has one of its
dimensions much larger than a platelet and a sheet has
both dimensions much larger than a platelet) have
three orthogonal planes of symmetry and, thus, their
elastic behaviour is characteristic of three-dimensional
orthotropic materials, which possess nine independent
elastic constants [42, 43].
Very little theoretical or experimental work has
been performed on the elastic properties of platelet- or
ribbon-reinforced composites, apart from some results
proposed by Halpin [44], a theoretical model by
Padawer and Beecher [45], and a model by Lusis et al.
[46]. Related work by Nielsen [47, 48] is also avail-
able. Some elastic constants (like Young’s moduli)
have been studied more than others, and therefore
theoretical models for some constants may not be
available. In the following it is assumed that the ‘‘1’’
direction is parallel to the long dimension of the plate-
let or the ribbon, the ‘‘2’’ direction is perpendicular to
the former, within the plane of the platelet or ribbon,
and the ‘‘3’’ direction is the out-of-plane direction. The
‘‘1’’ direction in vivo is also parallel to the axis of the
collagen fibril [10]. We now review the models avail-
able for each of the nine elastic constants for the UPP
platelet- and ribbon-reinforced composites.
2.1.1.1. ¸ongitudinal ½oungs modulus, E
1
. According
to the HalpinTsai model [44], the longitudinal (par-
allel to the long axis of the platelet or ribbon) modulus
of a UPP composite is given by
E
1
"E
.
1#AB»
1
1!B»
1
(1)
where
B"
E
1
/E
.
!1
E
1
/E
.
#A
(2)
and where E
.
is the matrix modulus, E
1
is the platelet
(or ribbon) modulus, »
1
is the volume fraction of
platelet reinforcement, given by
»
1
"
A
¸
¸#u
BA
¼
¼#v
BA
¹
¹#d
B
"
A
1#
u
¸
BA
1#
v
¼
BA
1#
d
¹
B
~1
(3)
where ¸ is the platelet length, ¼ is the platelet width
and ¹ is its thickness, and u, v and d are the longitudi-
nal, transverse in-plane and transverse out-of-plane
Figure 2 Schematic illustrations of the arrangements of highly sty-
lized platelet-shaped crystals in parallel-fibred bone introducing the
geometrical parameters and the defined (1, 2, 3) set of principal axes
constituting the model for this specific bone type.
distances between platelets, respectively (Fig. 2). In
particular, for ribbon-reinforced composites (uP0
and ¸PR)
»
1
"
A
1#
v
¼
BA
1#
d
¹
B
~1
The parameter A in Equations 1 and 2 is given by
A"2
A
¸
¹
B
#40»
10
1
(4)
For typical platelet dimensions and platelets volume
fraction in bone (Table I) 40»
10
1
;2(¸/¹). Thus, for
a UPP platelet-reinforced ‘‘boney’’ composite
A+2(¸/¹). In the case of ribbon-reinforced com-
posites, ¸PR, thus APR and (from Equation 2)
BP0 whereas ABP(E
1
/E
.
)!1. Equation 1, there-
fore, transforms into the rule of mixtures
E
1
"E
.
(1
1
)#E
1
»
1
(5)
Other models for Young’s modulus, E
1
, of platelet-
reinforced composites were proposed by Padawer and
Beecher [45] and by Lusis et al. [46]. Both have
similar expressions as follows
E
1
"E
.
(1
1
)#lE
1
»
1
(6)
in which, for the Padawer and Beecher (PB) model,
l"1![tanh(n)/n], whereas for the Lusis et al.
(LWX) model, l"1![ln(n#1)/n]. In both cases
n"
¸
¹
A
G
.
E
1
B
1@2
A
»
1
1
1
B
1@2
(7)
where G
.
is the matrix shear modulus, and the other
parameters were previously defined. Here also, as for
the Halpin and Tsai (HT) scheme (Equation 1), the
rule of mixtures is obtained for ribbon-reinforced
composites, because ¸PR, nPR and lP1in
both cases. The LWX scheme is essentially a modified
version of the PB model, as it takes into account
1500
plateletplatelet interactions at high platelet content.
The PB and LWX schemes are based on Cox’s early
model [9] but they take into account the platelet-like
geometry of the reinforcement. Available experimental
data (for glass, aluminium diboride and silicon carbide
platelets in PB [45] and for phlogopite and muscovite
mica flakes in LWX [46] fall short of all schemes due
to variability in the geometrical and the structural
parameters, but the general trends are quite satisfac-
tory. Thus, three different expressions are available for
the calculation of E
1
.
2.1.1.2. ¹ransverse in-plane ½oungs modulus E
2
.
Three different expressions are also available for the
transverse in-plane Young’s modulus, E
2
, based on the
HT, PB, and LWX schemes. The expressions are the
same as Equations 1 and 5, but the aspect ratio ¸/¹ is
now replaced by ¼/¹ in Equations 4 and 7. Regard-
ing Equation 4, one has again the simplification
40»
10
1
;2(¼/¹).
2.1.1.3. ¹ransverse out-of-plane ½oungs modulus, E
3
.
The HT model does not offer an expression for the
transverse out-of-plane Young’s modulus, E
3
. PB,
LWX models are not relevant here because they deal
with the modulus reduction factor, l, in the plane of
reinforcement only. We thus suggest the following
approximation for E
3
. By interchanging axes 2 and 3,
and 1 and 3, one obtains aspect ratios (¹/¼) and
(¹/¸), respectively. As the length and width of an
individual platelet are of the same order of magnitude,
and are both much larger than its thickness (Table I),
the aspect ratio for E
3
, to be used in Equations 4 and
7, is assumed to be A+2¹/(¼#¸). As ¹ gets small-
er with respect to ¼ and ¸, the value of A approaches
zero, as indeed was suggested by Nielsen [48] for
ribbons. For A"0, Equation 1 becomes the inverse
rule of mixtures
1
E
3
"
»
1
E
1
#
1
1
E
.
(8)
2.1.1.4. Shear moduli and Poissons ratios. For the
shear moduli, G
ij
, and the Poisson’s ratios m
ij
(i, j"1, 2, 3 and iOj) the HT equations may be used.
Table II lists the values of A used in the HT equations,
for the shear moduli and the Poisson’s ratios in the
cases of platelet or ribbon reinforcement, as suggested
by Halpin [44] and Nielsen [48]. Note from Table II
that no expressions for A exist for G
31
and m
31
in the
case of platelets, and for m
12
, m
31
in the case of ribbons.
For these missing values the following approaches will
be taken. Because for ribbon reinforcement
A
G
31
"A
G
23
(Table II), we will assume that the same
identity is valid for platelet reinforcement. Moreover,
because we have A
m
12
"R for platelet reinforcement,
the same will be assumed for ribbon reinforcement.
Finally, for both platelet and ribbon reinforcements
we will arbitrarily assume that A
m
31
"A
m
23
"0.
2.1.2. Sheet reinforcement
Sheet reinforcement is the limiting case of very wide
ribbons (large aspect ratio ¼/¹ ), as discussed by
TABLE II Literature values of the parameter A in Equations
1 and 2 for the principal elastic constants of platelet [44] and ribbon
[47, 48] reinforced composites
Elastic constant A (platelets) A (ribbons)
G
12
A
¸#¼
2¹ B
1.73
A
¼
¹ B
1.73
G
23
1
4!3m
.
0
G
31
0
m
12
R
m
23
00
m
31
——
Nielsen [47, 48]. Elastic constants for this case, as
suggested by Nielsen are
E
1
"E
2
"E
1
»
1
#E
.
(1
1
);
1
E
3
"
»
1
E
1
#
1
1
E
.
(9)
G
12
"G
1
»
1
#G
.
(1
1
);
1
G
23
"
1
G
31
"
»
1
G
1
#
1
1
G
.
(10)
m
12
"m
1
»
1
#m
.
(1
1
);
1
m
23
"
1
m
31
"
»
1
m
1
#
1
1
m
.
(11)
where the volume fraction, »
1
, now becomes, as uP0,
vP0 and ¸PR, ¼PR
»
1
"
A
1#
d
¹
B
~1
(12)
2.2. Elastic constants of lamellar bone
The calculation of Young’s modulus of lamellar bone
must take into account the fact that its basic repeating
structure, the lamellar unit, is composed of stacked
sets of ordered lamellae comprising a thin lamella,
a transition zone, a thick lamella and a fourth 120°
‘‘back-flip’’ lamella. Furthermore, each lamella has
a different orientation in space as defined by (
1
and
(
2
angles [7]. The relative orientation in space, and
thickness, of each lamella are important for accurate
numerical estimations of the overall Young’s modulus
of lamellar bone, using the following approach. We
assume [5, 7] that the principal axes of the thin
lamella are aligned with the principal axes of bone
(Fig. 3) and, thus, the stiffness constants of the thin
lamella in the major axes of bone (i.e. bone axis,
tangential axis, and radial axis) are merely those of an
individual lamella in its principal axes. We will first
transform the principal elastic moduli of each of the
remaining lamellae in a lamellar unit into their corres-
ponding values along the major axes of bone, then use
a rule of mixtures approach to calculate the overall
values of the moduli of bone along its principal axes.
For the purpose of transforming the elastic moduli
of the thick lamella and the ‘‘back-flip’’ lamella
1501
Figure 3 Schematic illustration (top) of the principal axes of a
long bone. BA, bone axis; TA, tangential axis; RA, radial axis. The
lower schematic illustration shows the orientations of the crystals
and the collagen fibrils in the thin lamella in relation to the 3 princi-
pal axes.
each along its principal axes (i.e. 1, 2 and 3 axes of
individual lamellae as defined above) into their elastic
moduli along the major axes of bone, we suggest the
following method.
We consider a rectangular system of coordinates
(x, y, z) with axes oriented arbitrarily with respect to
the principal directions of elasticity (1, 2, 3) of an or-
thotropic body. Following Lekhnitskii [42], we have
a general expression for Young’s moduli E
,,
, shear
moduli G
,-
and Poisson’s ratio m
,-
with k, l"x, y, z
along the arbitrary system of coordinates (x, y, z), in
terms of the cosine directors a
i
, b
i
,c
i
(i"1, 2, 3), the
corresponding Young’s moduli, E
ii
, and the other elas-
tic constants in the principal axes, m
ij
, (Poisson’s ra-
tios), and G
ij
(the shear moduli), as follows
(k, l"x, y, z, i, j"1, 2, 3 and k and l are linked to
i and j, respectively, such that k (or l)"x implies i (or
j)"1, k (or l)"y implies i (or j)"2, k (or l)"z
implies i (or j)"3)
E
kk
"
A
a
4
i
E
1
#
b
4
i
E
2
#
c
4
i
E
3
#a
2
i
b
2
i
I
12
#b
2
i
c
2
i
I
23
#c
2
i
a
2
i
I
31
B
~1
(13)
where
I
mn
"
1
G
mn
!2
m
mn
E
mm
(m, n"1, 2, 3 mOn) (14)
1
G
kl
"4
A
a
2
i
a
2
j
E
1
#
b
2
i
b
2
j
E
2
#
c
2
i
c
2
j
E
3
B
!8
A
m
12
E
1
a
i
a
j
b
i
b
j
#
m
23
E
2
b
i
b
j
c
i
c
j
#
m
31
E
3
c
i
c
j
a
i
a
j
B
#
(a
i
b
j
#a
j
b
i
)2
G
12
#
(b
i
c
j
#b
j
c
i
)2
G
23
#
(c
i
a
j
#c
j
a
i
)2
G
31
(kOl) (15)
m
kl
E
ii
"
C
m
12
E
1
(a
2
i
b
2
j
#a
2
j
b
2
i
)#
m
23
E
2
(b
2
i
c
2
j
#b
2
j
c
2
i
)
#
m
31
E
3
(c
2
i
a
2
j
#c
2
j
a
2
i
)
D
!
A
a
2
i
a
2
j
E
1
#
b
2
i
b
2
j
E
2
#
c
2
i
c
2
j
E
3
#
a
i
a
j
b
i
b
j
G
12
#
b
i
b
j
c
i
c
j
G
23
#
c
i
c
j
a
i
a
j
G
31
B
(kOl) (16)
The cosine directors a
i
, b
i
, c
i
, i"1, 2, 3 determine the
position and orientation of the arbitrary (x, y, z) sys-
tem with respect to the (1, 2, 3) system, via the trans-
formation
A
x
y
z
B
"
A
a
1
b
1
c
1
a
2
b
2
c
2
a
3
b
3
c
3
BA
1
2
3
B
or x"T·1 (17)
where T is the matrix of cosine directors. Thus,
a
1
"cos(x, 1), a
2
"cos(y, 1), etc. It results from
Equations 1317 that to calculate the nine elastic
moduli in the new coordinates system (x, y, z) one
needs to know: (a) the nine direction cosines a
i
, b
i
, c
i
,
(i"1, 2, 3) and (b) the nine elastic constants (Young’s
and shear moduli, and Poisson’s ratios) in the princi-
pal axes. Expressions for the nine elastic constants in
the principal axes were discussed in the section on
single lamellae. We now deal with the determination
of direction cosines.
For the calculation of the direction cosines the
following approach can be adopted. It is well known,
since Euler, that the relative orientations of two coor-
dinate systems may be specified by a set of three
independent angles [50]. There is a large number
of possible sequential choices for these three angles,
and we adopt the following transformations from
principal (or body) coordinates of the thin lamella
to spatial (or arbitrary) coordinates of thick or
‘‘back-flip’’ lamellae, based on the specific require-
ments in our studies of bone micromechanics [9].
Starting with the principal axes typical of a thin
lamella, and using the right-handed coordinate system
described in Fig. 4, one first rotates the initial (princi-
pal) system of axes 123 by an angle w
1
clockwise about
the 3 axis, and the new coordinate system is labelled
ngf. Next, the intermediate axes, ngf, are rotated
about the n axis counterclockwise by an angle w
2
to
produce another intermediate set, the n@g@f@ axes. Fi-
nally, the n@g@f@ axes are rotated counterclockwise by
an angle w
3
about the f@ axis to produce the desired
xyz system of axes. Fig. 4 illustrates the various stages
of the sequence. The elements of the complete trans-
formation T are obtained as the triple product of the
1502
T"T
3
T
2
T
1
"
A
cos w
1
cos w
3
#sin w
1
cos w
2
sin w
3
cos w
1
cos w
2
sin w
3
!sin w
1
cos w
3
sin w
2
sin w
3
sin w
1
cos w
2
cos w
3
!cos w
1
sin w
3
cos w
1
cos w
2
cos w
3
#sin w
1
sin w
3
sin w
2
cos w
3
!sin w
1
sin w
2
!cos w
1
sin w
2
cos w
2
B
(19)
Figure 4 Rotations defining the Eulerian angles. (a) First rotation,
w
1
, clockwise about axis 3, (b) second rotation, w
2
, counterclockwise
about axis n; (c) third rotation, w
3
, counterclockwise about axis f@.
individual rotations:
¹
1
"
A
cos w
1
!sin w
1
0
sin w
1
cos w
1
0
001
B
¹
2
"
A
10 0
0 cos w
2
sin w
2
0 !sin w
2
cos w
2
B
¹
3
"
A
cos w
3
sin w
3
0
!sin w
3
cos w
3
0
001
B
(18)
and thus, using Equation 17, one has x"T
·
1 where
By identification of the elements of matrix T
using Equations 17 and 19, the cosine directors a
i
, b
i
,
and c
i
are now defined in terms of the elementary
rotations, which are convenient from an experimental
viewpoint.
As discussed, much of the variation within the struc-
ture of lamellar bone can be described in terms of the
plywood angle w
1
which describes the extent of offset
of the collagen fibril axes from the thin lamella to
other lamella in a given lamellar unit and the rotation
angle w
2
which describes the extent of rotation about
the fibril axis relative to fibrils in the thin lamella.
These two elementary rotations are modelled exactly
by the first two rotations of the Eulerian transforma-
tion presented above. No third Eulerian angle w
3
is
needed in this particular case in order to complete the
transformation of the principal coordinates of the thin
lamella into those of the thick or the ‘‘back-flip’’
lamella. Thus, w
3
"0°. Therefore, the matrix for the
complete transformation from thin lamella to thick or
‘‘back-flip’’ lamellae is
¹"
A
cos w
1
!sin w
1
0
sin w
1
cos t
2
cos w
1
cos w
2
sin w
2
!sin w
1
sin w
2
!cos w
1
sin w
2
cos w
2
B
(20)
However, as the principal axes of the thin lamella are
those which correspond (Fig. 3) to the major axes of
bone, we are interested in the inverse transformation,
namely, transforming the principal elastic constants of
the thick and the ‘‘back-flip’’ lamellae into those in the
principal axes of the thin lamella. Thus, the desired
matrix of direction cosines should be that of the in-
verse transformation, which is actually the transposi-
tion of the matrix derived above, as we are dealing
with an orthogonal transformation [50]. Thus, the
matrix of direction cosines for the transformation of
principal axes from thick or ‘‘back-flip’’ lamellae to
thin lamella is
¹~1"
A
cos w
1
sin w
1
cos w
2
!sin w
1
sin w
2
!sin w
1
cos w
1
cos w
2
!cos w
1
sin w
2
0 sin w
2
cos w
2
B
(21)
To conclude, a method is available for calculating the
elastic moduli of the thick and ‘‘back-flip’’ lamellae in
the principal axes of the thin lamella which corres-
pond to the major axes of bone. For simplicity, the
elastic moduli of the intermediate lamellae (i.e. the
transition zone) in the major axes of bone are assumed
here to be the mathematical average of the corres-
ponding moduli of the thin and thick lamellae. More
accurate calculations should take into account the
exact contributions of the lamellae constituting the
transition zone but this awaits the determination of
the values of the rotational angle, w
2
, for each of these
lamellae which has not, as yet, been measured.
Assuming the thicknesses of the four types of lamel-
lae to be ¹ 5)*#,, ¹ 5)*/, ¹ 53!/4 and ¹ "& with a total
lamellar unit thickness of ¹"¹ 5)*#, 5)*/#
¹ 53!/4"&, the elastic moduli of bone in its major
axes may be assumed to be given by
D
B0/%
kl
"
1
¹
(¹ 5)*/D
5)*/
kl
5)*#,D
5)*#,
kl
53!/4D
53!/4
kl
"&D
"&
kl
(22)
1503
where D
B0/%
kl
stands for the overall elastic moduli E, G
and m of bone in its principal axes k, l(k, l"
BA, TA, RA), and D
5)*/
kl
, D
5)*#,
kl
, D
53!/4
kl
, D
"&
kl
stand for the
elastic constants of the thin lamellae, thick lamellae,
the transition zone and the ‘‘back-flip’’ lamellae, res-
pectively, in the same major axes of bone.
2.3. Young’s modulus of bone in an
arbitrary direction
Knowing the principal elastic constants of parallel
fibred bone (or lamellar bone), we are now also inter-
ested in calculating their Young’s modulus in any
direction in space. It is more convenient to do so by
expressing E along an arbitrary direction in space in
terms of spherical coordinates. One possibility is to
rotate the rigid framework of principal axes of parallel
fibred bone (or the rigid framework of principal axes
of lamellar bone) such that the ‘‘1’’ axis (or the bone
axis) would point into that arbitrary direction. This is
performed (Fig. 5) by first rotating the framework by
an angle, u, clockwise about the ‘‘1’’ axis (or the bone
axis) and then by an angle, h, clockwise about the new
‘‘3’’ axis (or the new radial axis). The matrix of trans-
formation is, thus
A
cos h !sin h 0
sin h cos h 0
001
BA
10 0
0 cos u !sin u
0 sin u cos u
B
"
A
cos h !sin h cos u sin h sin 6
sin h cos h cos u !cos h sin u
0 sin u cos u
B
(23)
Thus, the direction cosines a
1
, b
1
, c
1
needed for the
calculation of the Young’s modulus of parallel fibred
bone in an arbitrary direction in space defined by the
spherical coordinates h, u (Fig. 5), using Equation 13,
are now a
1
"cos h, b
1
"!sin h cos u, c
1
"sin h
sin u. Substituting a
1
, b
1
, c
1
into Equation 13 one gets
E(h, u)"
A
cos4 h
E
1
#
sin4 h cos4 u
E
2
#
sin4 h sin4 u
E
3
#cos2 h sin2 h cos2 uI
12
#sin4 h sin2 u cos2 uI
23
#sin2 h cos2 h sin2 uI
31
B
~1
(24)
where
I
ij
"
1
G
ij
!2
m
ij
E
ii
(i, j"1, 2, 3 iOj) (25)
Thus, Young’s modulus of parallel fibred bone in the
space defined by the 12, 23, and 31 planes
(0°)h)90°,0°)u)90°) is given as a function of
its elastic constants in its principal directions and the
spherical coordinates, h, u, as defined above. Equation
24 is valid for the calculation of the three-dimensional
Young’s modulus of lamellar bone in the space defined
by the BATA. TARA, and RABA, as well, with the
bone axis (BA), tangential axis (TA), and radial axis
Figure 5 Definition of direction in space for parallel-fibred bone (or
for lamellar bone) by spherical coordinates. (a) First rotation, u,
clockwise about the ‘‘1’’ axis (or the bone axis); (b) second rotation,
h, clockwise about the ‘‘3’’ axis (or the radial axis). The subsequent
rotations result in the ‘‘1’’ axis (or the bone axis) pointing to the
specific direction in space defined by h and u with regard to its
orientation in the original system of coordinates. Letting
0°)h)90° and 0°)u)90° the whole three-dimensional space
defined by the 12, 23, and 31 planes of parallel-fibred bone (or
the BATA, TARA, and RABA planes of lamellar bone) is
covered.
(RA) replacing ‘‘1’’, ‘‘2’’, and ‘‘3’’ axes of parallel fibred
bone, respectively (Fig. 5).
3. Strategy
The elastic constants of platelet-reinforced parallel-
fibred bone in its principal axes are first calculated
using the HT, PB and LWX models with the para-
meters listed in Table II. Next Young’s modulus of
platelet-reinforced parallel-fibred bone in three-di-
mensional space is calculated using Equation 24. For
this purpose, only the HT model is used as it contains
expressions for all nine elastic constants which are
necessary for three-dimensional calculations. For
comparison, Young’s moduli of platelet-reinforced
parallel-fibred bone in the 12 plane are calculated for
the HT, PB and LWX models, and the results for the
cases of platelet, ribbon and sheet reinforcement are
compared.
The calculated elastic constants in the principal
axes of a single UPP platelet-reinforced lamella are
used to calculate (using Equations 1316) the stiffness
constants of individual thin, thick, and ‘‘back-flip’’
lamellae in the major axes of bone. The orientations of
the latter two lamellae relative to the thin lamella in
terms of w
1
and w
2
are (Table I) w
1
"90° and
w
2
"70°, w
1
"120° and w
2
"90°, respectively. The
values of the stiffness constants of the transition zone
in the major axes of bone are taken as the average
1504
value of the respective thick and thin moduli. Using
the relative lamellae thicknesses listed in Table I, the
overall elastic constants of lamellar bone in its major
axes are calculated using Equation 22. The three-di-
mensional Young’s modulus of lamellar bone is cal-
culated by introducing the latter results into Equation
24. The model for rotated plywood lamellar bone
structure is then examined with respect to several
variables and is compared to existing experimental
data.
4. Results
4.1. Parallel-fibred bone
Fig. 6 shows the three-dimensional Young’s modulus
of parallel-fibred bone using the HT model for platelet
reinforcement. A highly anisotropic structure is re-
vealed with values of E
1
"33.6 GPa, E
2
"24.3 GPa
and E
3
"3.0 GPa for the Young’s moduli in the prin-
cipal axes of parallel-fibred bone. Young’s modulus of
parallel-fibred bone in the 12 plane (u"0), as cal-
culated using the three different models is reported in
Fig. 7. The HT, PB and LWX models all yield similar
trends for the relation between stiffness and angle h.
These similarities are expected because the assump-
tion of orthotropy is made in all models. The differ-
ences in three curves are due only to the different
expressions for E
1
and E
2
used in Equation 24 One
striking feature is the presence of a local maximum at
a value of hO0° or 90° for u"0° (12 plane). This
maximum progressively disappears (Fig. 6) as u in-
creases. As discussed by Wagner and Weiner [7], it
can be shown [51, 52] that for u"0°, E(h, u) has
a maximum different from both E
1
and E
2
for some
values of h other than 0° and 90° if
G
12
'
E
1
2(1#m
12
)
(26)
and, similarly, E(h, u) has a minimum different from
both E
1
and E
2
for some values of h other than 0° and
Figure 6 Three-dimensional Young’s moduli of parallel-fibred bone
(HT model) in the space defined by the 12 plane (u"0°), 13
plane (u"90°), nd 23 plane (h"90°).
Figure 7 Young’s modulus of parallel-fibred bone in the 12 plane
(u"0°) as calculated by the three different models (HT, PB, LWX).
90° if
G
12
(
E
1
2[(E
1
/E
2
)#m
12
]
(27)
The extremum (minimum or maximum) occurs at the
values of h which satisfy
tan2 h"
(2/E
1
)#(2m
12
/E
1
)!(1/G
12
)
(2/E
2
)#(2m
12
/E
1
)!(1/G
12
)
(28)
Thus, for example, using the HT model, the following
elastic constants are calculated for parallel-fibred
bone in the 12 plane (u"0°): E
1
"33.6 GPa,
E
2
"24.3 GPa, G
12
"19.8 GPa and m
12
"0.34. Sub-
stituting these values into Equations 26 and 27 a value
of 12.5 GPa for the right-hand side of Equation 26 and
a value of 9.8 GPa for the right-hand side of Equation
27 are obtained. This fulfils the conditions of Equation
26 but not of Equation 27, implying that E will have
a maximum between 0° and 90°. For Equation 28, this
maximum is calculated to occur at h+37°.
The cases of ribbon and sheet reinforcement as
compared to platelet reinforcement in parallel-fibred
bone are shown in Fig. 8. Ribbon reinforcement en-
hances the Young’s modulus of parallel-fibred bone in
the longitudinal direction of the reinforcement from
33.6 GPa (platelet reinforcement) to 57.5 GPa. The
Young’s modulus in the other two directions remain
the same as for platelets. For the case of sheet rein-
forcement E
1
"E
2
"57.5 GPa, whereas E
3
does not
change. Note, that in the mid-range of the h angle
(around 30°(h(60°) Young’s modulus shows an
intermediate value for ribbon reinforcement, whereas
for sheets, as for platelets, there is a local maximum of
the Young’s modulus in this range.
4.2. Lamellar bone
The three-dimensional Young’s modulus of lamellar
bone as calculated using the HT model for the structural
1505
Figure 8 Young’s modulus in the 12 plane (u"0°) of platelet,
ribbon- and sheet-reinforced parallel-fibred bone (HT model).
Figure 9 Three-dimensional Young’s moduli of lamellar bone as
calculated for the material and structural parameters listed in
Table I in the space defined by the bone axistangential axis plane
(u"0°), bone axisradial axis plane (u"90°), and tangential
axisradial axis plane (h"90°).
parameters listed in Table I versus the load
direction in space, as defined by the spherical coordi-
nates h, u, is shown in Fig. 9. The structure is clearly
anisotropic. The calculated Young’s modulus values
in the three orthogonal directions are shown in
Table III. Young’s modulus of bone in the bone
axistangential axis plane, as determined using the
HalpinTsai, Padawer and Beecher and the Lusis et
al. models, is shown in Fig. 10. All three models show
a smoothly decreasing trend in Young’s modulus from
the bone axis value to the tangential axis. Note that
the HT and LWX models give approximately the
Figure 10 Young’s modulus of lamellar bone in the bone
axistangential axis plane (u"0°) as calculated by the three differ-
ent models (HT, PB, LWX) for the material and structural para-
meters listed in Table I.
same values (2627 GPa and &9 GPa) for the
Young’s modulus in the bone axis and tangential axis,
respectively, whereas the LWX model gives lower re-
sults (&23 GPa and &8 GPa).
Table III also lists values for the Young’s modulus
in the bone axis and tangential axis (both calculated
by the LWX model), and in the radial axis (calculated
by the HT model) for various values of the rotational
angle, (
2
, of the thick and ‘‘back-flip’’ lamellae. For
each case, results for platelet, ribbon and sheet rein-
forcement are compared. Results for the specific case
in which the ‘‘back-flip’’ lamella in a lamellar unit is
absent, are also listed. Values of all other parameters
(Table I) were kept constant. Note that (
2
values of
the rotated plywood structure are not known, and
hence it is of interest to examine the sensitivity of the
model to variations of (
2
. The case in which indi-
vidual platelets adhere along their width dimension to
form long bands, the longitudinal direction of which is
the former width dimension of platelets (‘‘transversally
formed ribbons’’), which is also structurally possible, is
examined as well. Here, ¼PR (instead of ¸ in Equa-
tions 4 and 7) and, thus, the rule of mixtures is ob-
tained by all three models for E
2
but not for E
1
in an
individual lamella. In addition, A
G
12
"(¸/¹ )1.73
(Table II). All the other values of A for the remaining
elastic moduli for ribbons (Table II) do not change.
Table III shows that variations in (
2
affect the
Young’s modulus values in the radial axis direction in
particular. Furthermore, if the ‘‘back-flip’’ lamella is
removed, the modulus values are much less isotropic.
Table III also shows that the moduli for sheets and
longitudinally formed ribbons as compared to plate-
lets are extremely high in the bone axis (BA) direction,
whereas for transversally formed ribbons, values of the
Young’s modulus in the bone axis are much closer to
those of platelets.
1506
TABLE III Data for Young’s modulus of lamellar bone in the bone axis, tangential axis and radial axis directions for various values of the
rotational angle, (
2
, and for the case in which the ‘‘back-flip’’ lamella is absent: (a) using parameters defined in Table I; (b) as in (a), but
(
2
"50° for the thick lamella; (c) as in (a), but (
2
"30° for the ‘‘back-flip’’ lamella; (d) as in (a), but assuming no ‘‘back-flip’’ lamella present
(a) (b) (c) (d)
Lamella (
1
"90°, (
2
"70° (
1
"90°, (
2
"50° (
1
"90°, (
2
"70° (
1
"90°, (
2
‘‘Back-flip’’ (
1
"120°, (
2
"90° (
1
"120°, (
2
"90° (
1
"120°, (
2
"30° No ‘‘back-flip’’
lamella
PRR*SPRSPRSPR
E
1
(GPa) 22.9 41.2 24.6 48.4 22.9 41.2 48.4 25.3 44.1 51.3 26.2 48.8
E
2
(GPa) 8.3 13.6 6.5 13.6 9.5 14.6 14.8 10.0 15.1 15.4 9.2 15.9
E
3
(GPa) 13.6 12.8 24.4 24.5 9.2 8.6 16.1 10.7 9.8 14.4 12.2 11.1
P"platelet reinforcement; R"ribbon reinforcement; S"sheet reinforcement; R*"transversally formed ribbons (see text).
5. Discussion
The models presented here are based on our current
understanding of the parallel-fibred and rotated ply-
wood structures and their materials properties. The
calculated values reflect the accuracies and inaccur-
acies both in our understanding of the in vivo struc-
tures and in the construction of the mathematical
model, and of course do not take into account the
considerably natural variation that exists in all bones.
With so many unknowns, a fairly stringent way to test
the model is to compare it to measured values of
Young’s modulus in different directions via-a-vis the
bone axial directions, following the approach of Bon-
field and Grynpas [24] and Reilly and Burstein [23].
Unfortunately, here too difficulties are encountered.
No measured modulus values of pure parallel-fibred
bone exist. All the measurements made are of
fibrolamellar (also known as plexiform bone), which is
a mixture of both parallel-fibred and lamellar bone. In
the case of lamellar bone, our model is only relevant to
a planar array of lamellae, and not to the far more
complicated situation of osteonal bone, which is for
the most part, a series of cylinders made up of folded
lamellae [14]. No measurements of the Young’s
modulus of planar lamellar bone have been made to
date, because of the relatively large specimen size
needed. More or less planar arrays of circumferential
lamellar bone are present in certain bones, but these
are almost inevitably from small animals, and thus the
volumes available for modulus measurements, are too
small.
We have recently attempted partially to resolve this
problem by making microhardness measurements on
both pure parallel-fibred bone and planar lamellar
arrays at different angles relative to the bone axes. The
results are reported in Ziv et al. [15]. Although micro-
hardness values cannot be directly related to Young’s
modulus, an approximate linear relation for bone has
been observed26, 53, 54, 55. Ziv et al. [15] measured the
microhardness values of parallel-fibred bovine bone in
a zone close to the periosteum where lamellar bone is
absent. Measurements were made along the three
principal axes. It was found that each has a different
value and that the differences are large (compared to
lamellar bone), indicating a high degree of anisotropy.
These trends are well reproduced in the model for
platelets (cf. Figs 6 and 9 paying particular attention
to the scales). Ziv et al. [15] also measured intermedi-
ate off-axis angles in order to determine whether or
not the off-axis peak in Young’s modulus predicted by
Wagner and Weiner [7] is present. No such peak was
found. Interestingly, the intermediate values measured
are thus more consistent with the crystals being pres-
ent in the form of ribbons rather than platelets (follow-
ing Fig. 8) and it would be most interesting to try to
evaluate this possibility by obtaining better in vivo
information on crystal dimensions. Figs 11 and 12
compare two of the trends observed by Ziv et al. [15]
for lamellar bone with the calculated (LWX model,
platelet reinforcement, parameters as listed in Table I)
modulus values for the same axial directions. The
Figure 11 Trends in the microhardness measurements for the ‘‘a’’
set of planes [15] and in the Young’s modulus of the corresponding
directions in the bone axistangential axis plane, as calculated by
the LWX model for platelet reinforced lamellar bone (material and
structural parameters as listed in Table I).
1507
Figure 12 Trends in microhardness measurements for the ‘‘c’’ set of
planes [15] and in the Young’s modulus of the corresponding
directions in the bone axisradial axis plane, as calculated by the
LWX model for platelet-reinforced lamellar bone (material and
structural parameters as listed in Table I).
trends are remarkably similar, providing strong sup-
port for the validity of the model.
A comparison of the calculated values of Youngs’
modulus with the measured modulus values for bulk
fibrolamellar and osteonal bone are informative, des-
pite the fact that they are not directly comparable. In
general, the well-known measured anisotropy for
fibrolamellar and osteonal bone [1] (higher values
of Young’s modulus in the bone axis direction relative
to the radial and tangential directions) is consistently
reproduced in our model for platelets, ribbons and
sheets. Care must be taken when comparing the abso-
lute values and trends measured with the calculated
values for the reasons outlined above. We will restrict
the discussion to osteonal bone measurements, as at
least here we are dealing with only one structural type
(as opposed to fibrolamellar bone). The relevant data
sets of Reilly and Burstein [23] and Hasegawa et al.
[56] range in Young’s modulus values along the bone
axis to the tangential or radial directions from 35 GPa
to about 8 GPa. Our model using the structural para-
meters in Table I conforms well to this range for the
cases of platelets and ribbons, but not sheets
(Table III). The trends observed for off-axis measure-
ments of Young’s modulus of osteonal bone (Reilly
and Burstein [23], Pidaparti et al. [25]) show a
uniform decrease from the bone axis direction to the
tangential/radial directions. Our model reproduces
this trend well, but also predicts that the tangential
and radial directions of planar lamellar sheets will
have quite different absolute values (from Fig. 9). Such
measurements, when available, could also provide in-
direct information on whether or not the dominant in
vivo crystal shape of a given lamellar bone is in the
form of platelets or ribbons, as these have (Table III)
quite different tangential and radial values.
6. Conclusion
Relating bone structure, and, in particular, the lamel-
lar structure common in mammalian bones, to mech-
anical properties, is indeed a very challenging task.
The difficulties are not only due to the complex hier-
archical structure and the necessity for measuring
mechanical properties of very small specimens, but
also to the considerable structural variation that is an
inherent feature of biological materials. A flexible
mechanical model is, therefore, an important means of
overcoming some of these difficulties. A model of this
type may also be useful for better understanding com-
plex synthetic composite materials, and could, in par-
ticular, provide insights into the relative importance of
the various structural features in terms of bulk mater-
ial properties.
Acknowledgements
This study was funded by US Public Health Service
grant (DE 06954) from the NIDR to S. W. S. W. holds
the I.W. Abel Professorial Chair of Structural Biology.
References
1. J. D. CURREY, ‘‘The Mechanical Adapatations of Bones’’
(Princeton University Press, Princeton, NJ, 1984).
2. S. WEINER and W. TRAUB, FASEB J. 6 (1992) 879.
3. J. D. CURREY, J. Bomech 2 (1969) 1.
4. Idem, ibid. 2 (1969) 477.
5. S.WEINER,T.ARADand W. TRAUB, FEBS ¸ett. 285
(1991) 49.
6. N. SASAKI, T. IKAWA and A. FUKUDA, J. Biomech. 24
(1991) 57.
7. H. D. WAGNER and S. WEINER, ibid 25 (1992) 1311.
8. J. D. CURREY, K. BREAR and P. ZIOUPOS, ibid. 27 (1994)
885.
9. W. J. LANDIS, M. J. SONG, A. LEITH, L. McEWEN and
B. F. McEWEN, J. Struct. Biol. 110 (1993) 39.
10. R. STUHLER, Forscht. Geb. Ro¨ ntgenstrahlen 57 (1937) 231.
11. R. A. ROBINSON, J. Bone Surg. 34A (1952) 389.
12. E. WACHTEL and S. WEINER, J. Bone. Miner. Res. 9 (1994)
1651.
13. S. WEINER and P. A. PRICE, Calcif. ¹iss. Int. 39 (1976) 365.
14. H. FRANCILLON-VIE LLOT, V. DE BUFFERENI L, J.
CASTANET, J. GERAUDIE, F. J. MEUNIER, J. Y. SIRE,
L. ZYLBERBERG
and A. DE RICQLES,in ‘‘Skeletal Bio-
mineralization Patterns, Processes and Evolutionary Trends’’,
edited by J. G. Carter, Ch. 20, pp 471530.
15. V. ZIV, H. D. WAGNER and S. WEINER, Bone 18 (1996)
417.
16. W. GEBHARDT, Arch. Entw. Mech. Org. 20 (1906) 187.
17. M. M. GIRAUD-GUILLE, Calcif. ¹iss. Int. 42 (1988) 167.
18. S. A. REID, Anat. Embryol. 174 (1986) 329.
19. J. W. SMITH, J. Bone Joint Surg. 42B (1960) 588.
20. G. MARROTI, Calicif. ¹iss. Int. 53 (1993) 547.
1508
21. S. WEINER, T. ARAD, I. SABANAY and W. TRAUB, Bone,
in press.
22. V. ZIV, I. SABANAY, T. ARAD, W. TRAUB and S.
WEINER,
Microsc. Res. ¹ech. 33 (1996) 203.
23. D. T. REILLY and A. H. BURSTEIN, J. Biomec-. 8 (1975)
393.
24. W. BONFIELD and M. D. GRYNPAS, Nature 270 (1977)
453.
25. R. M. V. PIDAPARTI, A. CHANDRAN, Y. TAKANO and
C. H. TURNER, J. Biomech. 29 (1996) 909.
26. G. P. EVANS, J. C. BEHIRI, J. D. CURREY and W. BON-
FIELD,
J. Mater. Sci. Mater. Med. (1990) 38.
27. S.WEINER,T.ARADand W. TRAUB, Chem. Biol. Mineral.
¹iss. (1992) 93.
28. S. WEINER and W. TRAUB, FEBS ¸ett. 206 (1986) 262.
29. P. FRATZL, M. GROSCHNER, G. VOGEL, H. PLENK,
Jr, J. ESCHBERGER, N. FRATZL-ZELMAN, K. KOLLER
and K. KLAUSHOFER, Bone Miner. Res. 7 (1992) 329.
30. A. BOYDE, Cell ¹iss. Res. 152 (1974) 543.
31. M. J. GLIMCHER, Phil. ¹rans. R. Soc. ¸ond. B304 (1984)
479.
32. S. FITTON-JACKSON, Proc. R. Soc. ¸ond. Ser. B. 146 (1956)
270.
33. W. TRAUB, T. ARAD and S. WEINER, Conn. ¹iss. Res. 28
(1992) 99.
34. A. L. ARSENAULT, Calcif. ¹iss. Res. 6 (1991) 239.
35. E. P. KATZ, E. WACHTEL, M. YAMAUCHI and G. L.
MECHANIC,
Conn. ¹iss. Res. 21 (1989) 149.
36. M. YAMAUCHI, E. P. KATZ, O. KAZUNORI, K.
TERAOKA
and G. L. MECHANIC, ibid 18 (1989) 41.
37. M. W. K. CHEW and J. M. SQUIRE, Int. J. Biol. Macromol.
8 (1986) 27.
38. H. J. HOHLING,B.A.ASHTONand H. D. KOSTER, Cell.
¹iss. Res. 148 (1974) 11.
39. S. DOTY, R. A. ROBINSON and B. SCHOFIELD,
in ‘‘Handbook of Physiology,’’ (edited by G. D. Aurbach,
(American Physiology Society, Washington, DC, 1976) pp.
323.
40. R. S. GILMORE and J. L. KATZ, J. Mater. Sci. 17 (1982)
1131.
41. A. TANIOKA, T. TAZAWA, K. MIYASAKA and K.
ISHIKAWA,
Biopolymers 13 (1974) 735.
42. S. G. LEKHNITSKII, ‘‘Theory of Elasticity of an Anisotropic
Elastic Body’’ (Holden-Day, San Francisco, 1963) pp. 173.
43. U. AKIVA, E. ITZHAK and H. D. WAGNER, Compos. Sci.
¹ech., in press.
44. J. C. HALPIN, ‘‘Primer on Composite Materials: Analysis’’,
Revised Edition (Technomic Publishing, Lancaster, PA, 1984)
pp. 125137.
45. G. E. PADAWER and N. BEECHER, Polym. Engng Sci. 10
(1970) 185.
46. J. LUSIS, R. T. WOODHAMS and M. XHANTOS, ibid. 13
(1973) 139.
47. L. E. NIELSEN, J. Appl. Phys. 41 (1970) 4626.
48. L. E. NIELSEN, ‘‘Mechanical Properties of Polymers and
Composites’’, Vol. 2 (Marcel-Dekker, 1974) Ch. 8.
49. H. L. COX, J. Appl. Phys. 3 (1952) 72.
50. H. GOLDSTEIN, ‘‘Classical Mechanics’’ (Addison-Wesley,
1980) pp. 1438 and Appendix B.
51. B. D. AGARWAL and L. J. BROUTMAN, ‘‘Analysis and
Performance of Fiber Composites’’ (Wiley, New York, 1980).
52. B. GERSHON, D. COHN and G. MAROM, Biomaterials 11
(1990) 548.
53. J. D. CURREY and K. BREAR, J. Mater. Sci. Mater. Med.
1 (1990) 14.
54. G. P. EVANS, J. C. BEHIRI and W. BONFIELD, Adv. Bio-
mater. 8 (1988) 311.
55. R. HODGSKINSON, J. D. CURREY and G. P. EVANS, J.
Orthop. Res. 7 (1989) 754.
56. K. HASEGAWA, C. H. TURNER and D. B. BURR, Calcif.
¹iss. Int. 55 (1994) 381.
Received 19 August
and accepted 20 August 1997
1509
... In turn, each sublamella is made up by bundles of MCFs roughly parallel to each other, but having different orientations in the different sublayers, giving rise to the so-called rotated-twisted-plywood structure (Giraud-Guille, 1988;Weiner et al., 1999). The MCFs orientation in the th sublayer ( = 1, … , ) of the th lamella ( = 1, … , ), with respect to the osteon long axis , can be uniquely characterized by two angles and (Akiva et al., 1998;Weiner et al., 1999). In particular, as sketched in Figs. ...
... where the transformation matrices [T ] and [D ] depend on angles and and are computed as reported in Kollar and Springer (2003). Following well-established evidence (Weiner and Traub, 1992;Weiner et al., 1999;Akiva et al., 1998), sublamellae are assumed to be characterized by the same sequence of values for angles and , and for the thickness in each lamella, namely = , = , = , and = for each = 1, … , . A further numerical homogenization step allows to reduce the lamellar tissue to an equivalent locally-homogeneous linearlyelastic material, described by the fourth-order (undamaged) elastic stiffness tensor C . ...
... However, multiscale models that give a complete description of the hierarchical organization of cortical bone from the molecular scale to the macro-scale level are still lacking. Among the studies that have focused on a single microscopic scale level, some have described the mechanical behavior of lamellae (Akiva et al., 1998;Vercher et al., 2014;Yoon and Cowin, 2008b), while others have investigated osteons (Ascenzi et al., 2008;Pidaparti et al., 1996;Yoon and Cowin, 2008a). At the submicroscopic scale, just a few models have been developed to estimate the mechanical properties of the mineralized collagen fiber (Yoon and Cowin, 2008b), while most of the studies have been concerned with the mineralized collagen fibrils (Akkus, 2005;Buehler, 2008;Eppell et al., 2006;Ghanbari and Naghdabadi, 2009;Hang and Barber, 2011;Jäger and Fratzl, 2000;Sansalone et al., 2009;Shen et al., 2008;Van Der Rijt et al., 2006;Vercher-Martínez et al., 2015;Yuan et al., 2011). ...
Article
With osteoporosis and aging, structural changes occur at all hierarchical levels of bone from the molecular scale to the whole tissue, which requires multiscale modeling to analyze the effect of these modifications on the mechanical behavior of bone and its remodeling process. In this paper, a novel hybrid multiscale model for cortical bone incorporating the tropocollagen molecule based on the combination of finite element method and different homogenization techniques was developed. The objective was to investigate the influence of age-related structural alterations that occur at the molecular level, namely the decrease in both molecular diameter (due to the loss of hydration) and number of hydrogen bonds, on mechanical properties of the bone tissue. The proposed multiscale hierarchical approach is divided in two phases: (i) in Step 0, a realistic 3D finite element model for tropocollagen was used to estimate the effective elastic properties at the molecular scale as a function of the collagen molecule’s degree of hydration (represented by its external diameter) and the number of its intramolecular hydrogen bonds, and (ii) in Steps 1–10, the effective elastic constants at the higher scales from mineralized fibril to continuum cortical bone tissue were predicted analytically using homogenization equations. The results obtained in healthy mature cortical bone at different scales are in good agreement with the experimental data and multiscale models reported in the literature. Moreover, our model made it possible to visualize the influence of the two parameters (molecular diameter and number of hydrogen bonds) that represent the main age-related alterations at the molecular scale on the mechanical properties of cortical bone, at its different hierarchical levels. Keywords: Bone aging, multiscale model, tropocollagen, cortical bone, finite element modeling, homogenization method.
... The Young's moduli in all directions in space can be represented in radial coordinates, E(θ,ϕ), and be calculated as follows [50]: ...
Article
Full-text available
Additively manufactured (AM) materials and hot rolled materials are typically orthotropic, and exhibit anisotropic elastic properties. This paper elucidates the anisotropic elastic properties (Young’s modulus, shear modulus, and Poisson’s ratio) of Ti6Al4V alloy in four different conditions: three AM (by selective laser melting, SLM, electron beam melting, EBM, and directed energy deposition, DED, processes) and one wrought alloy (for comparison). A specially designed polygon sample allowed measurement of 12 sound wave velocities (SWVs), employing the dynamic pulse-echo ultrasonic technique. In conjunction with the measured density values, these SWVs enabled deriving of the tensor of elastic constants (Cij) and the three-dimensional (3D) Young’s moduli maps. Electron backscatter diffraction (EBSD) and micro-computed tomography (μCT) were employed to characterize the grain size and orientation as well as porosity and other defects which could explain the difference in the measured elastic constants of the four materials. All three types of AM materials showed only minor anisotropy. The wrought (hot rolled) alloy exhibited the highest density, virtually pore-free μCT images, and the highest ultrasonic anisotropy and polarity behavior. EBSD analysis revealed that a thin β-phase layer that formed along the elongated grain boundaries caused the ultrasonic polarity behavior. The finding that the elastic properties depend on the manufacturing process and on the angle relative to either the rolling direction or the AM build direction should be taken into account in the design of products. The data reported herein is valuable for materials selection and finite element analyses in mechanical design. The pulse-echo measurement procedure employed in this study may be further adapted and used for quality control of AM materials and parts.
Article
Directed energy deposition (DED) can potentially allow manufacturing of Nitinol (NiTi) parts that are difficult to fabricate using traditional methods. Here, large NiTi square bars were built vertically in the z-direction by Laser Engineered Net Shaping (LENS™), using a pre-alloyed Ni-Ti powder. Laser power, laser scan speed, and powder mass flow rate were optimized using the design of experiments (DOE) methodology. Each bar was cut at the top, middle, and bottom sections for microstructural and mechanical analyses. The middle section was found to be the densest, and with a complex grain structure slightly more equiaxed and finer than in the top and bottom sections. X-ray diffraction detected only the B2 phase in all three sections. A surprising mechanical behavior is reported, for the first time, where the middle section exhibits the lowest Young’s and shear moduli (based on an ultrasonic pulse-echo test) and the highest microhardness. Atom probe tomography (APT) allowed to rationalize the anomalous mechanical behavior based on the formation of a fine dispersion of Ni4Ti3 nanoprecipitates with the highest volume fraction in the middle section. These microstructural and mechanical heterogeneities have both fundamental and practical implications.
Article
Full-text available
The hierarchical design of bio-based nanostructured materials such as bone enables them to combine unique structure-mechanical properties. As one of its main components, water plays an important role in bone's material multiscale mechanical interplay. However, its influence has not been quantified at the length-scale of a mineralised collagen fibre. Here, we couple in situ micropillar compression, and simultaneous synchrotron small angle X-ray scattering (SAXS) and X-ray diffraction (XRD) with a statistical constitutive model. Since the synchrotron data contain statistical information on the nanostructure, we establish a direct connection between experiment and model to identify the rehydrated elasto-plastic micro- and nanomechanical fibre behaviour. Rehydration led to a decrease of 65%-75% in fibre yield stress and compressive strength, and 70% in stiffness with a 3x higher effect on stresses than strains. While in agreement with bone extracellular matrix, the decrease is 1.5-3x higher compared to micro-indentation and macro-compression. Hydration influences mineral more than fibril strain with the highest difference to the macroscale when comparing mineral and tissue levels. The effect of hydration seems to be strongly mediated by ultrastructural interfaces while results provide insights towards mechanical consequences of reported water-mediated structuring of bone apatite. The missing reinforcing capacity of surrounding tissue for an excised fibril array is more pronounced in wet than dry conditions, mainly related to fibril swelling. Differences leading to higher compressive strength between mineralised tissues seem not to depend on rehydration while the lack of kink bands supports the role of water as an elastic embedding influencing energy-absorption mechanisms.
Preprint
Full-text available
The multiscale architectural design of bio-based nanostructured materials such as bone enables them to combine unique structure-mechanical properties that surpass classical engineering materials. In biological tissues, water as one of the main components plays an important role in the mechanical interplay, but its influence has not been quantified at the length scale of a mineralised collagen fibre. Here, we combine in situ experiments and a statistical constitutive model to identify the elasto-plastic micro- and nanomechanical fibre behaviour under rehydrated conditions. Micropillar compression and simultaneous synchrotron small angle X-ray scattering (SAXS) and X-ray diffraction (XRD) were used to quantify the interplay between fibre, mineralised collagen fibrils and mineral nanocrystals. Rehydration led to a 65% to 75% decrease of fibre yield stress and compressive strength, and a 70% decrease of stiffness with a 3x higher effect on stress than strain values. While in good agreement with bone extracellular matrix, the decrease is 1.5-3x higher compared to micro-indentation and macro-compression. Hydration has a higher influence on mineral than fibril strain while the highest difference to the macroscale was observed comparing mineral and tissue levels. Results suggest that the effect of hydration is strongly mediated by ultrastructural interfaces while corroborating the previously reported water-mediated structuring of bone apatite providing insights towards the mechanical consequences. Results show that the missing reinforcing capacity of surrounding tissue is more pronounced in wet than dry conditions when testing an excised array of fibrils, mainly related to the swelling of fibrils in the matrix. Differences leading to higher compressive strength between mineralised tissues do not seem to depend on the rehydration state while fibril mobilisation follows a similar regime in wet and dry conditions. The lack of kink bands point towards the role of water as an elastic embedding, thus, adapting the way energy is absorbed. Statement of significance Characterising structure-property-function relationships of biomaterials helps us to elucidate the underlying mechanisms that enables the unique properties of these architectured materials. Experimental and computational methods can advance our understanding towards their complex behaviour providing invaluable insights towards bio-inspired material development. In our study, we present a novel method for biomaterials characterisation. We close a gap of knowledge at the micro- and nanometre length scale by combining synchrotron experiments and a statistical model to describe the behaviour of a rehydrated single mineralised collagen fibre. Results suggest a high influence of hydration on structural interfaces, and the role of water as an elastic embedding. Using a statistical model, we are able to deduce the differences in wet and dry elasto-plastic properties of fibrils and fibres close to their natural hydration state.
Preprint
In this paper, a multiscale rationale is applied to develop a bottom-up modelling strategy for analyzing the elasto-damage response of osteons, resulting in a first step towards a refined mechanical description of cortical bone tissue at the macroscale. Main structural features over multiple length scales are encompassed. A single osteon is described by considering a multi-layered arrangement of cylindrical lamellae and accounting for both lacunar micro-voids and thin interlamellar regions, these latter modelled as soft interfaces. A multi-step homogenization procedure has been conceived and numerically applied to describe the equivalent mechanical response of osteon constituents, upscaling dominant subscale mechanisms. A progressive stress-based damage approach has been implemented via a finite-element technique, allowing to describe interlaminar and/or intralaminar brittle failure modes. Proposed approach has been successfully validated by numerically reproducing available experimental tests of isolated osteons under different loading conditions. Present histologically-oriented multiscale model revealed to be sound and consistent, opening towards further insights about the influence on bone biomechanics of through-the-scales biophysical/biochemical alterations, possibly related to ageing or diseases.
Chapter
Bones represent about 15% of the body’s weight, a figure which doesn’t deserve any interpretation of consequence. Most importantly, ambulation, ventilation, and protection of the human body is bone dependent for a start, highlighting the mechanical function of the bones making them a structural material with mechanical characteristics like any other mineral building material even if the process of discovery and study of the linkage between micro-components and bulk material is opposite for the two types of material. After more than 2000 years of improvements, we know the right components to make a very good steel but we don’t know yet how to fight osteoporosis efficiently in its natural occurrence with age. Following the Wolff’s law: bone remodels stronger where external stress goes higher. Literature shows many studies try and give probable cause of this tenet without the actual key of the phenomenon, which might hopefully be tantamount to unlock the osteoporosis treatment stalemate. Bone cells are no different from the other body cells (at the notable exception of neurons) with constant renewing until the process dramatically seizes up. So, characterization of bone mechanical behavior is not dependent on the bone age, as bone is always “young,” but the age of the individual and biological disorders and mechanical disturbances fall upon it. Beside shape and appearance, bone is classified into two main types: cortical, hard and dense and trabecular, porous and where plentiful biological processes, with some not yet deciphered, are taking place. Following characterization concerns cortical and trabecular separately then the composite of both. The objective of this first chapter is to better understand all the complexity of human bones related to: its structure, its composition, its histology, and its mechanical behavior at different levels of scales.
Article
Full-text available
Bone is an intriguingly complex material. It combines high strength, toughness and lightweight via an elaborate hierarchical structure. This structure results from a biologically driven self-assembly and self-organisation, and leads to different deformation mechanisms along the length scales. Characterising multiscale bone mechanics is fundamental to better understand these mechanisms including changes due to bone-related diseases. It also guides us in the design of new bio-inspired materials. A key-gap in understanding bone’s behaviour exists for its fundamental mechanical unit, the mineralised collagen fibre, a composite of organic collagen molecules and inorganic mineral nanocrystals. Here, we report an experimentally informed statistical elasto-plastic model to explain the fibre behaviour including the nanoscale interplay and load transfer with its main mechanical components. We utilise data from synchrotron nanoscale imaging, and combined micropillar compression and synchrotron X-ray scattering to develop the model. We see that a 10-15% micro- and nanomechanical heterogeneity in mechanical properties is essential to promote the ductile microscale behaviour preventing an abrupt overall failure even when individual fibrils have failed. We see that mineral particles take up 45% of strain compared to collagen molecules while interfibrillar shearing seems to enable the ductile post-yield behaviour. Our results suggest that a change in mineralisation and fibril-to-matrix interaction leads to different mechanical properties among mineralised tissues. Our model operates at crystalline-, molecular- and continuum-levels and sheds light on the micro- and nanoscale deformation of fibril-matrix reinforced composites.
Article
Full-text available
Taking inspiration from natural materials, composite materials can be reinforced by creating matrix architectures that can better accommodate and control internal stresses. Despite the recent success in the synthesis of artificial assemblies for local reinforcement through the introduction of oriented fibers and plates into host multilayered composites, there is a lack of fundamental understanding of the factors that determine mechanical properties. Moreover, designing building blocks and interfaces that facilitate higher resistance and energy dissipation is highly challenging. When the intrinsic material is fixed, the mechanical and tribological properties can be further adjusted. In this study, europium oxide nanosheets are arranged in interlocked‐junction superstructures that resist sliding at junction points, thereby enhancing the mechanical properties of the nanosheet assemblies compared to those of the conventional face‐to‐face superstructures formed by parallel nanosheets. Furthermore, the crystalline origin of building blocks is revealed by demonstrating that faulty crystal nanosheets adopting an amorphous structure are different from single‐crystal nanosheets, with the former exhibiting superior mechanical reinforcement and improved abrasive resistance. Inspired by nacre and bones, an effective fabrication method for enhancing the mechanical and tribological performances of self‐assembled structures is reported via two strategies. First, a progressive interfacial sliding resistance is thought to be a source of strength and deformation hardening. Second, regulating the building blocks with an amorphous microstructure means they are harder and tougher than those that are single crystalline.
Article
Scitation is the online home of leading journals and conference proceedings from AIP Publishing and AIP Member Societies
Article
Using X-ray diffraction to monitor structural preservation, a technique has been developed for the processing of tendon collagen for cryo-ultramicrotomy which is largely free from artefacts. The resulting negatively stained cryosections provided electron micrographic images of fibrils showing regular lateral periodicities. Detailed analysis of the data provides convincing evidence that the structure observed in the fibrils, although showing spacings in the range 56.0–116.5 Å, is directly compatible with the Hulmes-Miller quasi-hexagonal packing model for collagen molecules, and not with earlier tetragonal models. There is circumstantial evidence that some form of higher molecular organization, possibly compressed microfibrils, may be present within this basic quasihexagonal lattice.
Article
A generalized equation is proposed for the relative elastic moduli of composite materials. By the introduction of a generalized Einstein coefficient and a function which considers the maximum volumetric peaking fraction of the filler phase, the moduli of many types of composite systems can be calculated.
Article
The hardness, Young's modulus of elasticity, tensile yield stress, ultimate stress and calcium content of 65 specimens from mammalian long bones and dental tissues were determined. The hardness was a very good predictor of the Young's modulus and yield stress, but a less good predictor of ultimate stress. The relationship in each case was nearly linear. All of these properties have the same type of non-linear relationship to the calcium content of the specimens. Over the range of calcium contents used here, hardness would be a useful guide to the mechanical properties of mineralized tissues in situations where conventional test specimens could not be produced, or where variations in mechanical properties over small distances are of interest.