Anda di halaman 1dari 9

ARTICLE IN PRESS

Journal of Biomechanics 38 (2005) 1517–1525


www.elsevier.com/locate/jbiomech
www.JBiomech.com

Fracture in human cortical bone: local fracture criteria and


toughening mechanisms
R.K. Nallaa, J.S. Stölkenb, J.H. Kinneyb, R.O. Ritchiea,
a
Materials Sciences Division, Lawrence Berkeley National Laboratory, and Department of Materials Science and Engineering,
University of California, Berkeley, CA 94720 USA
b
Lawrence Livermore National Laboratory, Livermore, CA 94550
Accepted 19 July 2004

Abstract

Micromechanical models for fracture initiation that incorporate local failure criteria have been widely developed for metallic and
ceramic materials; however, few such micromechanical models have been developed for the fracture of bone. In fact, although the
fracture event in ‘‘hard’’ mineralized tissues such as bone is commonly believed to be locally strain-controlled, only recently has
there been experimental evidence (using double-notched four-point bend testing) to support this widely held belief. In the present
study, we seek to shed further light on the nature of the local cracking events that precede catastrophic fracture in human cortical
bone, and to define their relationship to the microstructure. Specifically, numerical computations are reported that demonstrate that
the stress and strain states ahead of such a notch are qualitatively similar irrespective of the deformation mechanism (pressure-
insensitive plasticity vs. pressure-sensitive microcracking). Furthermore, we use the double-notched test to examine
crack–microstructure interactions from a perspective of determining the salient toughening mechanisms in bone and to characterize
how these may affect the anisotropy in fracture properties. Based on preliminary micromechanical models of these processes, the
relative contributions of various toughening mechanisms are established. In particular, crack deflection and uncracked-ligament
bridging are identified as the major mechanisms of toughening in cortical bone.
r 2004 Elsevier Ltd. All rights reserved.

Keywords: Bone; Fracture; Toughening; Microstructure

1. Introduction et al., 1994; Ford and Keaveny, 1996; Yeh and Keaveny,
2001), until recently (Nalla et al., 2003a) there has been
Recent interest in bone and its mechanical properties no experimental evidence to verify this hypothesis.
has led to extensive research into how it fractures (e.g., Specifically, the experimental verification of strain-
Lindahl and Lindgren, 1967; Burnstein et al., 1976; controlled fracture initiation in bone was achieved using
Wright and Hayes, 1976; Katz, 1980; Ashman and Rho, a double-notched four-point bend test. The basis of this
1988; Park and Lakes, 1992; Keaveny et al., 1994; test is two-fold: (i) since both notches experience the
Vashishth et al., 1997; Zioupos and Currey, 1998; Burr, same bending moment, when one notch breaks, the
2002; Nalla et al., 2003a), although several questions other is ‘‘frozen’’ at the point of fracture, and (ii) in
remain unaddressed. For example, although models for the presence of any degree of inelasticity, the maximum
bone fracture are invariably based on the concept of a stresses peak ahead of the notch whereas the maximum
strain-controlled critical fracture event (e.g., Keaveny strains are at the notch. Consequently, microstructural
examination of the unbroken notch can give insight into
Corresponding author. Tel.: +1-510+486-5798; fax: +1-510-486- cracking events immediately preceding fracture; in
4881. particular, the location of the microstructural cracking
E-mail address: roritchie@lbl.gov (R.O. Ritchie). event that precedes macroscopic fracture at a notch

0021-9290/$ - see front matter r 2004 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jbiomech.2004.07.010
ARTICLE IN PRESS
1518 R.K. Nalla et al. / Journal of Biomechanics 38 (2005) 1517–1525

gives a definitive indicator of whether fracture initiation The region simulated extended over a distance equiva-
is locally stress- or strain-controlled. Strain-controlled lent to 100 times the notch radius in both directions.
fracture will initiate at the notch whereas stress- Uniaxial displacement was applied along the upper
controlled will initiate ahead of the notch. Preliminary boundary. Infinite-body conditions were approximated
experiments on human cortical bone (Nalla et al., by constraining the edge of the region to move with
2003a) (and also on human dentin (Nalla et al., 2003b) parallel motion.
showed that the initial local fracture events do indeed To simulate deformation by pressure-insensitive
initiate at the notch in both materials, which is plasticity and pressure-sensitive microcracking, two
consistent with a strain-based failure mechanism. nonlinear materials models were used. The first ‘‘plas-
The notch-field stress and strain distribution used in tic-damage’’ model (PD) used a Mises yield criterion,
these studies (Hill, 1950; Griffiths and Owen, 1971), based on tissue material properties provided by Niebur
however, pertained to a Mises solid, i.e., based on et al. (2000). At stresses beyond the yield stress, the
pressure-insensitive plasticity as in metals, whereas material becomes perfectly plastic; plastic strain was
inelasticity in bone may involve pressure-insensitive accumulated without change in unloading modulus. An
mechanisms such as plasticity in the collagen fibers and isotropic Young’s modulus of 18.7 GPa was assumed.
pressure-sensitive mechanisms such as microcracking The second model, utilized to better simulate deforma-
(Currey, 2001; Taylor, 2003), where the corresponding tion by microcracking, was based on the oriented
notch solutions are unknown. Consequently, in this ‘‘brittle-damage’’ model (BD) of Govindjee et al.
work, we determine the stress distributions using finite- (1995). This model treats the compliance tensor as an
element analysis for both pressure-insensitive plasticity internal variable, and adopts the maximum damage-
and pressure-sensitive microcracking, with the objective dissipation principle (Simo and Ju, 1987) (equivalent to
of further developing the double-notch test to probe the the maximum plastic-work principle in models of
local fracture events in bone. We use the technique to associative-plasticity). In compression, the constituent
seek further understanding of (i) the criteria for crack response was taken to be plastic yielding (identical to
initiation, and (ii) the salient toughening mechanisms PD model); in tension, a smeared-crack model was used
associated with subsequent crack growth and how they to simulate damage evolution. This has the effect of
are affected by the hierarchical microstructure of bone. reducing the element stiffness in response to an
To date, although various toughening mechanisms have increased microcrack density, the growth of which is
been proposed, including microcracking, crack deflec- controlled by the toughness. A fracture toughness,
tion and crack bridging (e.g., Vashishth et al., 1997, K c ¼ 3:2 MPaOm (strain-energy release rate 500 J/
2000; Yeni and Norman, 2000; Parsamian and Norman, m2) was selected from the middle of the range of
2001; Thompson et al., 2001; Yeni and Fyhrie, 2001; toughnesses for bone (Zioupos and Currey, 1998;
Wang et al., 2001, 2002; Nalla et al., 2003a, 2004, 2005), Lucksanambool et al., 2001; Phelps et al., 2000). As
little definitive information exists for their relative microcracking damage was considered to be unrecover-
potency. Moreover, as they undoubtedly operate in able, this reduced the unloading modulus. The damage
concert, this almost certainly controls the orientation- threshold for microcrack nucleation was taken to be
dependence of the toughness with respect to the 0.6% strain. Both models are fully described by Stölken
microstructure. Consequently, we also seek to investi- and Kinney (2003).
gate how microstructure affects this anisotropy of
fracture toughness by quantifying the effect of these 2.2. Experimental fracture testing
mechanisms by simple theoretical modeling.
Human cortical bone from the mid-diaphyses of three
freshly frozen cadaveric humeri (from 34-year old
2. Methods and materials females) was used; three orientations were studied
(Fig. 1):
2.1. Numerical computations
 anti-plane longitudinal (medial–lateral)–– long axes of
Events that result in macroscopic fracture can be the osteons in the plane of the notch/crack, but
described as either locally stress- or strain-controlled. To perpendicular to the (nominal) crack-propagation
derive the stress/strain distributions required to achieve direction (N ¼ 9),
this distinction, numerical analysis using the nonlinear,  in-plane longitudinal (proximal–distal)–– long axes of the
implicit, three-dimensional finite-element code NIKE3D osteons in the plane of the notch/crack, but perpendi-
(Maker and Hallquist, 1995) was used to simulate cular to the crack-propagation direction (N ¼ 7),
inelastic deformation ahead of a notch in plane-strain.  transverse–– long axes of the osteons perpendicular to
Using symmetry, only one quadrant of the problem was the plane of the notch/crack and to the crack-
modeled with a 4000 linear finite element graded mesh. propagation direction (N ¼ 6).
ARTICLE IN PRESS
R.K. Nalla et al. / Journal of Biomechanics 38 (2005) 1517–1525 1519

Samples were loaded to failure under displacement


In-plane control in 25 1C HBSS at a cross-head displacement rate
Anti-plane longitudinal of 0.01 mm/s, with applied loads and load-line displace-
longitudinal ments simultaneously monitored in situ. Stress-intensity
factors, K, were computed from handbook solutions
(ASTM Standard E-399-90, 2001). Due to substantial
crack deflections, mode-I stress intensities for the
transverse orientation may be somewhat questionable.
Consequently, additional measurements were made of
the work-of-fracture, Wf, an alternative measure of
toughness obtained by dividing the area under the
Transverse Transverse load–displacement curve by twice the nominal crack-
surface area; however, this latter parameter is both size-
and geometry-dependent.

Fig. 1. Schematic illustration of the various specimen orientations,


namely transverse, proximal–distal (in-plane longitudinal) and med- 3. Results
ial–lateral (anti-plane longitudinal), taken from the humeri (with
respect to the direction of the osteons, indicated in gray). 3.1. Numerical computations

Normalized notch-field stress/strain distributions


from the finite-element simulations are shown in Fig. 2.
Experiments on N ¼ 22 samples (thickness For both models of inelasticity, i.e., the plasticity (PD)
B1.1–2.6 mm, width W1.7–2.7 mm, spacing between
inner loading points S10–15 mm, notch root-radius
r200–300 mm, notch depth a/W0.3–0.4) were per-
formed using the double-notched four-point bending
geometry, as this enabled a determination of whether
crack initiation was stress- or strain-controlled and
further permitted the development of stable cracks.
Samples were kept hydrated during fabrication; prior to
testing, they were soaked in 25 1C Hanks’ balanced salt
solution (HBSS) in air-tight containers for 40 h1.
All testing was conducted in 25 1C HBSS using an
ELFs 3200-series mechanical-testing machine; bend
bars were loaded under displacement control at a
constant displacement rate of 0.01 mm/s. The area
around the unfractured notch in failed double-notched
samples was examined using optical microscopy and,
after coating with gold-palladium, in a scanning electron
microscope (SEM) operating in the back-scattered
electron mode.
Additional fracture-toughness tests were conducted
using the three-point-bending geometry (B1.1–2.6 mm,
W1.7–2.7 mm, S ¼ 5:025:5 W ) in the three orienta-
tions (N ¼ 3 for each orientation). A fatigue precrack
was grown from the notch at a load ratio (ratio of
minimum-to-maximum loads) of R ¼ 0:1 at 2 Hz
frequency; the final maximum stress intensity was
1–2 MPaOm with final precrack length of 0.4–0.6 W.
Fig. 2. Non-linear, finite-element computations of the distributions of
(a) tensile stress, s11, and (b) strain, e11, normalized by the yield stress,
1
It is possible that degradation of the collagen and/or leaching of the sy, and yield strain, ey, respectively, as a function of distance, x, ahead
mineral into solution could alter properties with time. However, no of a round hole, normalized by the radius of the notch, r. Calculations
such changes could be detected in parallel investigations in dentin are shown for inelastic deformation based on classical pressure-
following short-time storage under identical conditions (Habelitz et al., insensitive, shear-driven plastic deformation (PD) model and pressure-
2002). sensitive microcracking (brittle damage—BD) model.
ARTICLE IN PRESS
1520 R.K. Nalla et al. / Journal of Biomechanics 38 (2005) 1517–1525

and brittle-damage (BD) models, the stresses peak ahead 5 µm

of the notch whereas the strains peak directly at the root


of the notch. Thus, both distributions are qualitatively
similar for materials undergoing either shear-driven
(pressure-insensitivity) plasticity, e.g., as in metals, or
pressure-sensitive microcracking. These results are con-
sistent with those for J2 plasticity (Hill, 1950; Griffiths
and Owen, 1971) and with experimentally measured 1 µm
crack-tip strain distributions reported by Nicolella et al.
(2001) for bone. Thus, the double-notched-bend techni-
1 mm
que can be used to distinguish between stress- and
strain-controlled fracture in terms of where cracking
first initiates in materials undergoing both forms of
(a)
‘‘yielding’’, namely shear-driven plastic-flow and pres-
sure-sensitive microcracking. This result validates the
use of the double-notch test for this purpose in 1 µm
mineralized tissue materials such as bone.

3.2. Fracture testing

Crack initiation: Consistent with the preliminary data


of Nalla et al. (2003a), results from the double-notched 1 mm
four-point-bend tests showed that without exception, all
(b)
precursor cracks formed directly at the notch root; there
was no evidence of precursor cracking at the site of
10 µm
maximum stress (Fig. 3). This provides experimental
verification to endorse the concept that crack initiation
in human cortical bone is locally strain-controlled. In
the transverse orientation, i.e., how bone typically fails
in vivo, in addition to incipient cracking at the notch
root, there are instances of crack formation at micro-
structural features, such as the Haversian canals, near 1 mm
the notch (Fig. 3d), consistent with the higher strains (c)
measured at such features (Nicolella et al., 2001).
Crack growth: With respect to crack growth, it is
evident from Fig. 3 that the underlying microstructure
has a marked influence on the nature of the crack path,
particularly in the transverse orientation. Whereas
fracture in the longitudinal orientations (Figs. 3a,b)
nominally follows an expected trajectory dictated by the 100 µm
path of maximum tensile stresses, cracks in the 1 mm
transverse orientation extend initially in an unlikely (d)
direction perpendicular to the notch (Figs. 3c, d).
Fig. 3. Optical micrographs (left) of typical double-notch specimens
SEM micrographs of the subsequent propagation of after fracture, together with scanning electron micrographs (right) of
such cracks are shown in Fig. 4. Fig. 4a shows the the area of interest for the purpose of determining the failure criterion.
relatively deflection-free trajectory of a 1 mm long Note the absence of crack initiation at the location of maximum stress
crack in the medial–lateral orientation. However, crack/ as illustrated for the in-plane longitudinal (proximal–distal) orienta-
microstructural interactions are evident in other orien- tion (a: top, right). A much stronger influence of the underlying
microstructure can be seen for the transverse orientations (c and d) as
tations, which leads to toughening. Figs. 4b,c show compared to the longitudinal ones (a and b). Also, for the transverse
evidence of so-called uncracked-ligament bridging, an orientation, in some instances, cracking was observed ahead of the
extrinsic-toughening mechanism involving two-dimen- notch at a Haversian canal (d). The schematic insets in each optical
sional uncracked regions along the crack path that can micrograph show the orientation of the specimen and the black dashed
bridge the crack on opening. This mechanism, which is lines (in a–c) indicate the notch surface.
commonly seen in metal–matrix composites (Shang and
Ritchie, 1989) and intermetallics (Campbell et al., 1999),
is often the result of either the non-uniform advance of
ARTICLE IN PRESS
R.K. Nalla et al. / Journal of Biomechanics 38 (2005) 1517–1525 1521

6
HUMAN CORTICAL BONE
25°C, Wet

FRACTURE TOUGHNESS, Kc (MPa m)


4
(a) (b) 100 µm
100 µm

3.5 MPa√m

5.3 MPa√m
2.2 MPa√m
(c) 10 µm (d) 1 µm 1

Fig. 4. Scanning electron micrographs illustrating crack–microstruc-


ture interactions. For the anti-plane longitudinal (medial–lateral) 0
orientation, (a) the secondary osteons (indicated by black arrows) have (a)
a weak influence on crack path (after Nalla et al., 2003a), (b) evidence 3000
of bridging due to a secondary lamella, also shown at higher HUMAN CORTICAL BONE
magnification in (c), and (d) possible collagen fibril-based bridging. 25°C, Wet
The insets show the specimen orientation used and the white arrows in 2500
(a–d) indicate the direction of nominal crack growth.
WORK OF FRACTURE, Wf (J/m2 )

2000

the crack front and/or the imperfect link-up of micro-


cracks initiated ahead of the crack tip with the main 1500
crack. Uncracked-ligament bridging and microcracking
in the vicinity of the crack are also seen in the
proximal–distal orientation (Nalla et al., 2003a). The 1000

microcracking can in principle lead to extrinsic toughen-


ing since it can cause dilation and increase the
500
compliance of the region surrounding the crack. In
addition, crack bridging by individual collagen fibrils
may provide further toughening (Fig. 4d). 0
However, the largest microstructural influence on the (b)
crack path can be seen in the transverse orientation, Fig. 5. The measured toughness of bone, in terms of (a) fracture
where secondary osteons run along the specimen length. toughness and (b) work of fracture results, obtained in this study for
As mentioned previously, crack initiation and initial different orientations (schematically shown) with respect to the
crack growth out of the notch did not occur on a plane osteons. The half-error bars indicate one standard deviation; the
numbers in (a) are the average toughness values.
normal to the maximum tensile stress, but rather
perpendicular to this in the nominal direction of the
osteon system (Fig. 3c). As suggested by Yeni and The measured fracture-toughness and work-of-frac-
Norman (2000), it appears that the initiated crack can be ture values, given in Fig. 5 show Kc to be 2–6 MPaOm,
deflected from the nominal direction of propagation i.e., consistent with the previous results of Zioupos and
along the cement lines that are the interfaces between Currey (1998), Lucksanambool et al. (2001), and Phelps
the secondary osteons and the surrounding matrix as et al. (2000). In terms of Kc values, the transverse
this provides a weaker path away from the path of orientation is between 51–140% tougher than the
maximum tensile stress. Indeed, although there is longitudinal orientations. However, since the Kc values
significant cracking ahead of the notch at a Haversian in the transverse orientation may be somewhat ques-
canal (indicated by white arrow) in Fig. 3d, this seems to tionable, work-of-fracture measurements are also shown
be ignored by the main crack at the notch. The marked and indicate a toughness in the transverse orientation
(901) deflections in crack path seen in this case can lead that is again far in excess of the corresponding values for
to substantial toughening and may be the major factor either of the longitudinal orientations; similar qualita-
contributing to the anisotropy in fracture properties in tive trends have been reported by Lucksanambool et al.
bone. (2001). Such measured anisotropy in toughness is
ARTICLE IN PRESS
1522 R.K. Nalla et al. / Journal of Biomechanics 38 (2005) 1517–1525

believed to be a direct result of the various active  macroscopic crack deflection


toughening mechanisms, namely crack bridging, crack  microcracking.
deflection and microcracking. While such mechanisms
have long been hypothesized in bone (e.g., microcrack- Since the potency of these mechanisms depends on
ing by Vashishth et al. (1997, 2000), Parsamian and specific microstructural features, which in turn vary
Norman (2001); crack deflection by Yeni and Norman with orientation, the marked anisotropy in the
(2000); collagen fiber bridging by Thompson et al. fracture toughness can be considered to result
(2001), Wang et al. (2001, 2002), Yeni and Fyhrie directly from the relative contributions of these
(2001); and uncracked-ligament bridging by Nalla et al. mechanisms.
(2003a, 2004, 2005)), evidence of their quantitative The influence of the microstructure is strongest for the
relevance has largely been lacking; this issue is addressed cracks growing transversely. This is believed to be the
below. result of crack deflection along the cement lines, which
offers a path of lower resistance. For cracks growing
longitudinally, the osteons do not seem to influence the
4. Discussion macroscopic crack growth substantially (Fig. 4a) (see
also Nalla et al., 2004). This is reflected in the measured
The microstructure of a material can influence the fracture-toughness results, which also display a marked
fracture toughness in two primary ways (Ritchie, 1988, orientation-dependence.
1999): The highest toughness was observed in the transverse
orientation, consistent with previous results (Lucksa-
 it can affect the inherent resistance to microstructural
nambool et al., 2001; Phelps et al., 2000; Behiri and
damage and fracture ahead of the crack tip, which is
Bonfield, 1989); here the crack path deflects at 901 to
termed intrinsic toughening, and/or
the plane of maximum tensile stress (Figs. 3c,d). The
 it can promote crack-tip shielding, i.e., act to reduce
effect of this deflection is to increase the toughness
the local stress intensity actually experienced at or
substantially, as shown by the following analysis.
behind the crack tip, which is termed extrinsic
Assuming, for the sake of simplicity, that these
toughening.
deflections/kinks represent in-plane tilts through an
angle, a, to the crack plane, then the local mode-I and
Crack propagation can be considered as a mutual
mode-II stress intensities, k1 and k2, at the deflected
competition between these two classes of mechanisms,
crack tip are given by (Bilby et al., 1978; Cotterell and
i.e., microstructural damage in the process zone ahead
Rice, 1980)
of the crack tip, which acts to promote crack extension,
and extrinsic crack-tip shielding behind the tip, which k1 ðaÞ ¼ c11 ðaÞK I þ c12 ðaÞK II ;
acts to impede it. Whereas intrinsic toughening tends to
dominate in ductile materials and to affect crack- k2 ðaÞ ¼ c21 ðaÞK I þ c22 ðaÞK II ; (1)
initiation resistance, extrinsic mechanisms are invariably where KI(=5.33 MPaOm) and KII(=0) are, respectively,
the main source of toughening in brittle materials, and the mode-I and mode-II far-field stress intensities for a
govern crack-growth resistance. Given the relatively main crack, and the coefficients, cij(a), are mathematical
brittle nature of bone and the nature of the crack–mi- functions of the deflection angle, a(901) (Bilby et al.,
crostructure interactions (see Section 3.2), extrinsic 1978; Cotterell and Rice, 1980). The effective stress
mechanisms appear to provide the principal contribu- intensity at the tip of the deflected crack tip, Kd, can
tions to the toughness of bone, akin to other mineralized then be calculated by summing the mode-I and mode-II
tissues such as dentin in vitro (Nalla et al., 2003c). contributions in terms of the strain-energy release rate,
In the present work, firstly in terms of intrinsic viz.:
damage, we have provided sound evidence that the onset
of fracture in human cortical bone is consistent with a K d ¼ ðk21 þ k22 Þ1=2 ; (2)
strain-based criterion, which has been so widely used in
which suggests that the value of the stress intensity at the
theoretical models of the mechanical behavior of bone
crack tip is reduced locally by 50% due to such
(Keaveny et al., 1994; Ford and Keaveny, 1996; Yeh and
deflection to 2.7 MPaOm, compared to that for an
Keaveny, 2001). Secondly, in terms of the salient
undeflected crack. This calculation is consistent with the
toughening mechanisms, it is apparent from the current
toughness being approximately twice as high in this
observations of the microstructural crack paths that this
orientation as compared to the longitudinal orienta-
toughening arises extrinsically from several distinct
tions.
sources:
In the proximal–distal and medial–lateral orienta-
 crack bridging by uncracked ligaments tions, conversely, crack bridging appears to be the
 crack bridging by intact collagen fibrils prominent source of toughening. Theoretical estimates
ARTICLE IN PRESS
R.K. Nalla et al. / Journal of Biomechanics 38 (2005) 1517–1525 1523

of bridging due to uncracked ligaments in the latter Using a volume-fraction, fm3% from present
orientation can be made based on a limiting crack- observations and those of Vashishth et al. (1997) for
opening approach (Shang and Ritchie, 1989) : the in-plane longitudinal orientation, the contribution to
the toughness from microcracking would be
f ul K I ½ð1 þ l ul =rbÞ1=2  1
K br ¼ ; (3) 0.1 MPaOm; similar values are obtained for the other
1  f ul þ f ul ð1 þ l ul =rbÞ1=2 orientations. As impractically high microcrack volume-
where ful is the area-fraction of bridging ligaments on fractions of 60–90% would be needed for this
the crack plane (0.2–0.4, from crack-path observa- mechanism to account for the observed toughness of
tions), KI is the applied stress intensity (2.4 MPaOm), lul bone (Nalla et al., 2004), we do not believe that
is the bridging-zone size (50–300 mm, from crack-path constrained microcracking is a significant toughening
observations), r is a rotational factor (0.20–0.47) and mechanism.
b ¼ W  a: Again, substituting typical values for these In summary, direct experimental evidence has been
parameters, a contribution to the toughness on the order presented in support of a strain-controlled mechanism
of Kbr0.3 MPaOm was obtained for the medial–lateral for the onset of fracture in bone. In addition, a series
(anti-plane longitudinal) orientation. Similar analyses of extrinsic toughening mechanisms have been
for the proximal–distal (in-plane longitudinal) orienta- identified, specifically crack bridging by uncracked
tion yielded even higher values of 1–1.6 MPaOm due ligaments and collagen fibrils, crack deflection, and
to the larger (mm) bridging zones in this orientation microcracking, based on microscopic observations
(Nalla et al., 2004) of the interaction of the crack path with the underlying
For toughening associated with bridging by collagen microstructure. Using simple theoretical modeling
fibrils, the uniform-traction Dugdale-zone model (Evans for quantification, the dominant toughening mechan-
and McMeeking, 1986) can be employed to estimate the isms are deemed to be (i) crack deflection along
resulting decrease in the stress intensity, Kfb, due to the cement lines in the transverse orientation, and
‘‘fiber-bridging’’, viz: (ii) crack bridging by uncracked ligaments in the
longitudinal orientations. The roles of collagen
K fb ¼ 2sb f f ð2l f =pÞ1=2 ; (4) fibril bridging and microcracking are relatively insignif-
icant. Based on this appreciation of the toughening
where sb is the normal bridging stress on the fibrils
mechanisms in bone, the anisotropy in the fracture
(100 MPa), ff is the effective area-fraction of the
toughness of bone with orientation can be understood in
collagen fibrils on the crack plane (0.15, from crack-
terms of a large contribution from crack deflection in
path observations), and lf is the bridging-zone length
the transverse orientation as compared to a smaller
(10 mm, from crack-path observations). Using these
contribution from crack bridging in the longitudinal
estimates, a value of Kfb0.08 MPaOm can be obtained
orientations.
for both longitudinal orientations where such bridging
was observed, suggesting that bridging by individual
collagen fibers contributes little to the toughness of
bone. However, this mechanism might be expected to be
Acknowledgments
significant for individual microcracks, where due to the
substantially smaller size-scales collagen bridges would
This work was supported in part by the National
be able to span a larger percentage of the crack length.
Institutes of Health under Grant No. 5R01 DE015633
Finally, based on accepted models for microcrack
(for RKN), by the Office of Science, Office of Basic
toughening due to dilation and modulus reduction
Energy Science of the Department of Energy under
(Evans and Faber, 1984; Hutchinson, 1987; Sigl,
Contract No. DE-AC03-76SF00098 (for ROR), and by
1996), the increase in toughness due to microcracks
the Laboratory Science and Technology Office, LLNL,
can be expressed as
pffiffiffiffiffi under the auspices of the US Department of Energy
K m ¼ 0:22m E 0 f m l m þ bf m K c ; (5) W-7405-ENG48 (for JSS and JHK). The authors
wish to thank Drs. C. M. Puttlitz and Z. Xu (Depart-
where em is the residual volumetric strain (=0.002, as ment of Orthopedic Surgery, University of California,
calculated from data in Sigl, 1996), E0 is the plane-strain San Francisco, CA, USA) for supply of the human
elastic modulus, fm is the volume-fraction of micro- cortical bone used in this study, Dr. J. J. Kruzic
cracks, lm is the height of the microcrack zone, b is a (Materials Sciences Division, Lawrence Berkeley
factor dependent on Poisson’s ratio (1.2 (Sigl, 1996)2. National Laboratory, Berkeley, CA, USA) for many
2 useful discussions, and Dr. A. P. Tomsia (Materials
Note that the value of em used will tend to overestimate the
toughening since em is directly proportional to the residual stresses in
Sciences Division, Lawrence Berkeley National Labora-
the material, which can be hundreds of MPa in ceramics, much higher tory, Berkeley, CA, USA) for his continued support and
than could be sustained in bone. encouragement.
ARTICLE IN PRESS
1524 R.K. Nalla et al. / Journal of Biomechanics 38 (2005) 1517–1525

References Nalla, R.K., Kinney, J.H., Ritchie, R.O., 2003b. On the fracture of
human dentin: Is it stress- or strain-controlled? Journal of
Ashman, R.B., Rho, J.Y., 1988. Elastic modulus of trabecular bone Biomedical Materials Research 67A, 484–495.
material. Journal of Biomechanics 21, 177–181. Nalla, R.K., Kinney, J.H., Ritchie, R.O., 2003c. Effect of orientation
ASTM Standard E 399-90 (Reapproved 1997), 2001. Annual Book of on the in vitro fracture toughness of dentin: The role of toughening
ASTM Standards, vol. 03.01: Metals–– Mechanical testing; mechanisms. Biomaterials 24, 3955–3968.
Elevated and low-temperature tests; Metallography. ASTM, West Nalla, R.K., Kruzic, J.J., Kinney, J.H., Ritchie, R.O., 2005.
Conshohocken. Mechanistic aspects of fracture and R-curve behavior in human
Behiri, J.C., Bonfield, W., 1989. Orientation dependence on fracture cortical bone. Biomaterials 26, 217–231.
mechanics of bone. Journal of Biomechanics 22, 863–872. Nalla, R.K., Kruzic, J.J., Ritchie, R.O., 2004. On the origin of the
Bilby, B.A., Cardew, G.E., Howard, I.C., 1978. Stress intensity factors toughness of mineralized tissue: Microcracking or crack bridging.
at the tips of kinked and forked cracks. In: Fracture 1977, vol. 3. Bone 34, 790–798.
Pergamon Press, Oxford, pp. 197–200. Nicolella, D.P., Nicholls, A.E., Lankford, J., Davy, D.T., 2001.
Burnstein, A.H., Reilley, D.T., Martens, M., 1976. Ageing of bone Machine vision photogrammetry: a technique for measurement of
tissue: Mechanical properties. Journal of Bone and Joint Surgery microstructural strain in cortical bone. Journal of Biomechanics 34,
58A, 82–86. 135–139.
Burr, D.B., 2002. The contribution of the organic matrix to bone’s Niebur, G.L., Feldstein, M.J., Yuen, J.C., Chen, T.J., Keaveny, T.M.,
material properties. Bone 31, 8–11. 2000. High-resolution finite element models with tissue strength
Campbell, J.P., Venkateswara Rao, K.T., Ritchie, R.O., 1999. The asymmetry accurately predict failure of trabecular bone. Journal of
effect of microstructure on fracture toughness and fatigue crack Biomechanics 33, 1575–1583.
growth behaviour in g-titanium aluminide based intermetallics. Park, J.B., Lakes, R.S., 1992. Biomaterials: An Introduction. Plenum
Metallurgical and Materials Transactions A 30A, 563–577. Press, New York.
Cotterell, B., Rice, J.R., 1980. Slightly curved or kinked cracks. Parsamian, G.P., Norman, T.L., 2001. Diffuse damage accumulation
International Journal of Fracture 16, 155–169. in the fracture process zone of human cortical bone specimens and
Currey, J.D., 2001. Sacrificial bonds heal bone. Nature 414, 699. its influence on fracture toughness. Journal of Materials Science 12,
Evans, A.G., Faber, K.T., 1984. Crack-growth resistance of micro- 779–783.
cracking brittle materials. Journal of the American Ceramic Society Phelps, J.B., Hubbard, G.B., Wang, X., Agrawal, C.M., 2000.
67, 255–260. Microstructural heterogeneity and the fracture toughness of bone.
Evans, A.G., McMeeking, R.M., 1986. On the toughening of ceramics Journal of Biomedical Materials Research 51, 735–741.
by strong reinforcements. Acta Metallurgica 34, 2435–2441. Ritchie, R.O., 1988. Mechanisms of fatigue crack propagation in
Ford, C.M., Keaveny, T.M., 1996. The dependence of shear failure metals, ceramics and composites: role of crack-tip shielding.
properties of trabecular bone on apparent density and trabecular Materials Science and Engineering 103, 15–28.
orientation. Journal of Biomechanics 29, 1309–1317. Ritchie, R.O., 1999. Mechanisms of fatigue-crack propagation in
Govindjee, S., Kay, G.J., Simo, J.C., 1995. Anisotropic modelling ductile and brittle solids. International Journal of Fracture 100,
and numerical simulation of brittle damage in concrete. Interna- 55–83.
tional Journal for Numerical Methods in Engineering 38, Shang, J.-K., Ritchie, R.O., 1989. Crack bridging by uncracked
3611–3633. ligaments during fatigue-crack growth in SiC-reinforced alumi-
Griffiths, J.R., Owen, D.R.J., 1971. An elastic–plastic stress analysis num-alloy composites. Metallurgical Transactions A 20A, 897–908.
for a notched bar in plane strain bending. Journal of the Mechanics Sigl, L.S., 1996. Microcrack toughening in brittle materials containing
and Physics of Solids 19, 419–431. weak and strong interfaces. Acta Materialia 44, 3599–3609.
Habelitz, S., Marshall, G.W., Balooch, M., Marshall, S.J., 2002. Simo, J.C., Ju, J.W., 1987. Strain- and stress-based continuum damage
Nanoindentation and the storage of teeth. Journal of Biomechanics models-I. Formulation. International Journal of Solids and
35, 995–998. Structures 23, 821–840.
Hill, R., 1950. The Mathematical Theory of Plasticity. Clarendon Stölken, J.S., Kinney, J.H., 2003. On the importance of geometric
Press, Oxford. nonlinearity in finite-element simulations of trabecular bone
Hutchinson, J.W., 1987. Crack tip shielding by micro-cracking in failure. Bone 33, 494–504.
brittle solids. Acta Metallurgica 35, 1605–1619. Taylor, D., 2003. How does bone break? Nature Materials 2, 133–134.
Katz, J.L., 1980. The structure and biomechanics of bone. Symposium Thompson, J.B., Kindt, J.H., Drake, B., Hansma, H.G., Morse, D.E.,
of the Society for Experimental Biology 34, 137–168. Hansma, P.K., 2001. Bone indentation recovery time correlates
Keaveny, T.M., Wachtel, E.F., Ford, C.M., Hayes, W.C., 1994. with bond reforming time. Nature 414, 773–776.
Differences between the tensile and compressive strengths of bovine Vashishth, D., Behiri, J.C., Bonfield, W., 1997. Crack growth
tibial trabecular bone depends on modulus. Journal of Biomecha- resistance in cortical bone: Concept of microcrack toughening.
nics 27, 1137–1146. Journal of Biomechanics 30, 763–769.
Lindahl, O., Lindgren, A.G., 1967. Cortical bone in man I. Variation Vashishth, D., 2000. In vivo diffuse damage in human vertebral
of the amount and density with age and sex. Acta Orthopaedica trabecular bone. Bone 26, 147–152.
Scandinavica 38, 133–140. Vashishth, D., Tanner, K.E., Bonfield, W., 2000. Contribution,
Lucksanambool, P., Higgs, W.A.J., Higgs, R.J.E.D., Swain, M.W., development and morphology of microcracking in cortical bone
2001. Fracture toughness of bovine bone, influence of orientation during crack propagation. Journal of Biomechanics 33, 1169–1174.
and storage media. Biomaterials 22, 3127–3132. Wang, X., Bank, R.A., Tekoppele, J.M., Agrawal, C.M., 2001. The
Maker, B.N., Hallquist, J.O., 1995. NIKE-3D—a nonlinear, implicit, role of collagen in determining bone mechanical properties. Journal
three-dimensional finite element code for solid and structural of Orthopaedic Research 19, 1021–1026.
mechanics––user’s manual, UCRL-MA-105268-REV-1. Lawrence Wang, X., Shen, X., Li, X., Agrawal, C.M., 2002. Age-related changes
Livermore National Laboratory, Livermore. in the collagen network and the toughness of bone. Bone 31, 1–7.
Nalla, R.K., Kinney, J.H., Ritchie, R.O., 2003a. Mechanistic fracture Wright, T.M., Hayes, W.C., 1976. The fracture mechanics of fatigue
criteria for the failure of human cortical bone. Nature Materials 2, crack propagation in compact bone. Journal of Biomedical
164–168. Materials Research 10, 637–648.
ARTICLE IN PRESS
R.K. Nalla et al. / Journal of Biomechanics 38 (2005) 1517–1525 1525

Yeh, O.C., Keaveny, T.M., 2001. Relative roles of microdamage and Yeni, Y.N., Fyhrie, D.P., 2001. Collagen-bridged microcrack model
microfracture in the mechanical behavior of trabecular bone. for cortical bone tensile strength. In: Proceedings of the 2001
Journal of Orthopaedic Research 19, 1001–1007. Summer Bioengineering Conference, BED-vol. 50. ASME, New
Yeni, Y.N., Norman, T.L., 2000. Calculation of porosity and osteonal York, pp 293–294.
cement line effects on the effective fracture toughness of cortical Zioupos, P., Currey, J.D., 1998. Changes in the stiffness, strength,
bone in longitudinal crack growth. Journal of Biomedical Materials and toughness of human cortical bone with age. Bone 22,
Research 51, 504–509. 57–66.

Anda mungkin juga menyukai