You are on page 1of 9

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

net/publication/6819048

The role of osteocytes and bone microstructure in preventing osteoporotic


fractures

Article  in  Osteoporosis International · February 2007


DOI: 10.1007/s00198-006-0222-y · Source: PubMed

CITATIONS READS

45 191

3 authors:

Jan Geert Hazenberg David Taylor


Teva Pharmaceutical Industries Ltd. Trinity College Dublin
19 PUBLICATIONS   737 CITATIONS    233 PUBLICATIONS   5,495 CITATIONS   

SEE PROFILE SEE PROFILE

Clive Lee
Chinese Academy of Sciences
33 PUBLICATIONS   1,278 CITATIONS   

SEE PROFILE

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

FINITE FRACTURE MECHANICS View project

Investigations into the mechanical properties of insect cuticle (PhD) View project

All content following this page was uploaded by Jan Geert Hazenberg on 31 May 2014.

The user has requested enhancement of the downloaded file.


Osteoporos Int (2007) 18:1–8
DOI 10.1007/s00198-006-0222-y

REVIEW

The role of osteocytes and bone microstructure in preventing


osteoporotic fractures
Jan G. Hazenberg & David Taylor & T. Clive Lee

Received: 30 March 2006 / Accepted: 8 August 2006 / Published online: 14 September 2006
# International Osteoporosis Foundation and National Osteoporosis Foundation 2006

Abstract The skeleton alters its geometry following trauma, repair defects. Unfortunately, as we grow older it tends to
the introduction of artificial defects and of fatigue-induced work against us, especially in woman after the age of 40,
microcracks. The precise mechanism by which the skele- resulting in osteoporosis. Osteoporosis is a progressive
ton adapts remains unclear. Microcracks might directly bone disease in which bone tissue is normally mineralized,
affect the cell by damaging the osteocyte cell network or but the amount of bone is decreased and the structural
causing apoptosis. Bone microstructure may play an integrity is impaired, becoming more brittle, thinner and
important role in these processes by diverting and more prone to fracture [3]. Annually, 1.5-million fragility
arresting propagating microcracks and so prevent fracture fractures occur in the USA, including 700,000 vertebral
failure. This paper discusses the effects of microstructure fractures, 300,000 hip fractures, 250,000 Colles’ fractures,
on propagating cracks, how microdamage may act as a and 250,000 at other skeletal sites [4]. The annual cost
stimulus for bone adaptation and its potential effects on associated with this has been estimated at $17 billion. Hip
bone biochemistry. fractures are associated with the greatest morbidity, with
50% of patients unable to walk without assistance and 25%
Keywords Bone adaptation . Microdamage . Osteocytes . requiring long-term domiciliary care, and mortality, with
Osteoporotic fracture . RANKL 10–20% dying within 6 months [5]. In this paper, we
review how bone morphology affects damage accumulation
and the role that osteocytes play in preventing fracture in
Introduction bone, from microstructure to cell signalling.

Bones are subjected each day to cyclic tensile, compressive


and torsional loading, making them very susceptible to Bone microstructure and crack arrest
microdamage accumulation. The amount of microdamage
increases as our bones get older and as the number of bone Bone consists of a hard mineral phase, mainly comprised of
cells decrease. Caucasian bones are more susceptible to hydroxyapatite, and a fibrous scaffold, mainly composed of
damage accumulation than those of African Americans [1, type I collagen. The stiffness caused by the mineral phase
2]. Fortunately, bone contains a mechanism which is able to provides resistance to compressive stresses, while the
collagen provides bone with toughness and resistance to
tensile stresses. Throughout the matrix are vascular canals
J. G. Hazenberg (*) : T. C. Lee
that supply the bone cells, embedded in the hard matrix,
Department of Anatomy, Royal College of Surgeons in Ireland,
St. Stephen’s Green, with nutrients. The main vascular canals (Haversian canals)
Dublin 2, Ireland run approximately parallel to the long axis of the bone,
e-mail: jhazenberg@rcsi.ie whereas the smaller Volkmann’s canals radiate tangentially.
D. Taylor : T. C. Lee
Is it the sole function of these vascular canals to provide
Trinity Centre for Bioengineering, Trinity College, nutrients to the bone or might they have a secondary
Dublin 2, Ireland function in maintaining mechanical integrity?
2 Osteoporos Int (2007) 18:1–8

It has been known for over 40 years that bone in vivo propagation requires strain energy, which causes the crack
contains small cracks [6, 7]. These cracks are the visual to increase in length. It seems that shorter cracks tend to
manifestation of fatigue damage, caused by the cycles of stabilise and become dormant after reaching some critical
stress that occur during daily activities. To date, most length, whereas those longer than the critical length are
studies have focused either on detecting cracks formed in self-propagating and may cause fracture [21]. A number of
vivo in humans and animals [8, 9] or on experiments to researchers have noted that the vast majority of the cracks
study the initiation and propagation of cracks in specimens can be found in older bone, located between the osteons.
of dead bone during either brittle fracture [10–12] or fatigue The percentages of cracks found in interstitial bone by
[13–15]. Experimental measurements of cracks tend to be various researchers [22–24] are, respectively, 87%, 62%
made from histological sections oriented transversely to the and 85%. According to Martin and Burr [25], microcracks
bone’s long axis. Individual cracks have an average length formed in interstitial bone propagate to the cement line or
(measured tip-to-tip) of 60–100 μm. In longitudinal concentric lamellae and debond or separate the osteon from
sections, which have rarely been used, crack lengths appear the surrounding bone. Furthermore, propagating micro-
significantly longer than those in transverse sections [9, cracks under uniaxial tension exhibit an intermitted crack
16]. This implies that they are taking advantage of a growth pattern [26, 27]. Using optical microscopy to
relatively easy growth direction afforded by the anisotropic monitor propagating cracks, it was shown that this
‘grain’ of the material. Observations have been made of propagation behaviour was related to the cracks encounter-
different microcrack appearances in tension and compres- ing vascular canals, which either caused crack growth rates
sion. Cracks induced by compression are usually relatively to decrease or arrested the crack completely [27]. The
long, straight and start at obvious stress concentrators, such available data support the theory that cement lines,
as channels and lacunae [17, 18]. Figure 1 shows a surrounding the osteons and other vascular canals act as
scanning electron microscope (SEM) image of this phe- microcrack-arresting features or decrease their propagation
nomenon. A microcrack has propagated towards and then rate. These microscopic observations have shown that the
through an osteocyte lacuna in a bone subjected to a micro-mechanisms of crack propagation and toughening are
uniform load, while a second microcrack has initiated from complex and involve contributions from several different
a neighbouring lacuna. It shows that once cracks accumu- factors, including microdamage zones ahead of the crack tip
late, secondary damage formation occurs, which weakens and bridging actions across the crack faces [28–30].
the bone structure and increases fracture risk. Tensile Experiments conducted on bone demonstrate that fracture
cracks, on the other hand, are often orientated with the resistance is not constant but increases with crack length
grain of the bone and are associated with planes of [29, 31, 32]. Others have argued that microdamage, in the
weakness, such as cement lines and inter-lamellar debond- form of small cracks, or diffuse damage, in front of larger
ing [19, 20]. This is illustrated in Fig. 2. In general, crack microcracks acts as an energy-dissipating mechanism which
may cause crack arrest [33–35]. Observations of collagen
fibres crossing the crack faces could be another mechanism
by which crack propagation is controlled. This extrinsic
mechanism [36], could explain the observed increase in
toughness with crack length [37]. As the crack propagates,
more ligaments will be left in its wake, limiting the opening
and shear displacements between the crack faces [38].
Furthermore, microdamage accumulation and propagation
in cortical bone and trabecular bone might be completely
different in nature. Microdamage in cortical bone involves
cracks that are embedded in a solid matrix, whereas cracks
in trabecular bone are typically surface cracks. A surface
crack of the same length, under similar loading conditions,
will have more strain energy to propagate than those in
cortical bone. The consequences of a single fractured
trabecula might not pose a threat to the overall mechanical
integrity, while a propagating crack in cortical bone could
Fig. 1 A scanning electron microscope image of a microcrack (A),
result in fracture of the complete structure. It is evident
which has propagated towards and through an osteocyte lacuna. A from these studies that microcrack propagation in bone is a
second microcrack has initiated from a neighbouring osteocyte lacuna multi-scale problem.
Osteoporos Int (2007) 18:1–8 3

Fig. 2 Three examples (a, b, c) of microcracks propagating from a notch (N). They preferred to propagate in a direction which is parallel to the
osteonal direction instead of taking a direction perpendicular to the maximum principal stresses

Bone adaptation the surface bone lining cells. Although the functional role
of the osteocyte remains partly unknown, a key role in the
Bone adaptation is a dynamic process which involves both regulation of remodelling the skeletal architecture in
modelling and remodelling. Modelling is a process by response to mechanical load has been suggested. Marotti
which bone changes in length and diameter during growth et al. [48] suggested that an inhibitory signal is generated
and development. Unlike modelling, which involves either by the osteocytes, which is passed through their cell
resorption or formation, bone remodelling follows an processes to osteoclasts to prevent bone resorption. Martin
activation, resorption, formation sequence [39]. Remodel- [49] extended this concept by developing a hypothetical
ling of bone requires a complex arrangement and interac- concept for the down-regulation of remodelling by osteo-
tion of cells, collectively called basic multicellular units cytes. He proposed that the strength of the inhibitory signal
(BMUs). Both processes are tightly controlled by a number perceived by each osteoclast is proportional to the number
of cells. The four main types of cells which can be found in of osteocytes connected to lining cells located on quiescent
bone are bone lining cells, osteoblasts, osteocytes and bone surfaces, and inversely proportional to the distance
osteoclasts. Bone lining cells, osteoblasts and osteoclasts from the surface. Regions of bone exhibiting high rates of
are located on the bone surfaces, whereas the osteocytes are remodelling were shown to correlate with relatively low
in the bone matrix. Palumbo et al. [40] proposed that pre- numbers of viable osteocytes, as low osteocyte density
osteoblasts mature into osteoblasts, some of which are reduced this inhibition of remodelling [50]. This is
committed to differentiate further. These committed osteo- interesting since it has been reported that osteocytes in
blasts start to lose their activity and are buried inside the woven bone appear at an early stage of bone repair and
bone matrix by adjacent osteoblasts. Consequently, pre- develop shorter processes which are irregularly distributed
osteocytes start the formation of cellular processes and in the osteoid matrix. Osteocytes in lamellar bone form
become osteoid osteocytes. The mature osteocyte radiates long cell processes that are regularly distributed in mature
processes in all directions, forming a network inside the bone matrix. Based on the pattern of development of bone
hard mineralized matrix. There is great interest in osteo- processes by the osteocytes [51], woven bone osteocytes
cytes as potential regulators of damage and repair [41–45]. may be necessary for induction of lamellar bone osteocytes
followed by active appositional growth of lamellar bone at
an early stage of bone repair. It has been estimated that the
Osteocytes osteocyte population in woven bone is approximately four
to eight times as large as that in lamellar bone [52], which
Human cortical bone contains between 13,900 and 19,400 might indicate that woven bone with increased lacunar
osteocytes per mm3 and trabecular bone roughly 13,000 density may undergo remodelling at an accelerated rate
osteocytes per mm3 [46, 47]. These osteocytes form an [53]. Given the potentially important role of osteocytes,
interconnected network through their dendritic processes, immuno-fluorescence microscopy has been used to study
allowing communication between individual osteocytes and the morphology of these cells [54–56]. Osteocytes contain
4 Osteoporos Int (2007) 18:1–8

between 40 and 60 cell processes per osteocyte with a cell- showed microcracks in the bone cortex. After 10 days of
to-cell distance between 20 and 30 μm [56] and a diameter fatigue loading, resorption cavities were observed. Two rats
varying from 50 to 410 nm [57]. However, there are still had no microdamage following fatigue loading and no
questions which remain to be answered. For example, resorption cavities were observed. Hsieh et al. [63] and Lee
African Americans have higher BMD indexes [2], lower at al. [64] also used the isolated ulna loading model and
osteon densities [58], higher osteocyte densities [59] and reported bone deposition, but no evidence of microdamage
are less prone to stress fractures [60] compared with white formation was found. This is rather surprising, given the
Americans. Do these various parameters contribute to a loading amplitude and frequency used for the experiments.
reduction in the accumulation and initiation of micro- However, what these studies have shown is that it is
damage and therefore of stress and fragility fractures? irrelevant how many load cycles to which bones are
subjected, the stimulus caused by the induced strain
saturates. In these studies, there was an asymptotic
Microdamage-stimulated bone adaptation approach to saturation as the duration of loading was
increased from 36 to 720 cycles per day. Increasing the
Other researchers have developed theoretical models based duration of a loading bout therefore results in diminishing
on the idea that adaptation is controlled by the level of returns in bone formation, again suggesting that cells may
fatigue damage, i.e. by the number and length of cracks become non-responsive to repeated mechanical stimuli
[42, 61]. Martin [61], for example, used the amount of [65]. Even if these studies show that the effective altered
damage (mean crack length times crack density) as the strain would diminish, microdamage might indirectly cause
stimulus for bone remodelling and the activation frequency a sustained local stimulus. Microdamage causes a local
of BMUs is directly related to the amount of damage stress concentrator around its perimeter. If strain levels do
present in a local area. As a consequence of damage not decrease, cracks will continue to propagate and
formation, BMUs are activated, which increases the therefore the local strains, caused by these cracks, will
porosity resulting in increased strain. This promotes continue to increase [66]. It has been suggested that
damage accumulation, which is not dangerous while the increased strain levels cause osteocytic apoptosis due to
load remains below a critical level. The increasing damage the presence of microcracks which affect osteocyte homeo-
removal overtakes the rising damage formation rate, stasis [22, 67, 68]. Recent experimental work which has
resulting in a new state of equilibrium. On the other hand, tried to investigate this phenomenon has indicated that
when the effective strain levels are below a critical value, osteocyte apoptosis increases osteoclastic activity [69].
BMUs are activated, causing resorption of bone. This idea Various in vitro studies have focused on the role of
is an attractive one because the level of damage provides a osteocytes in bone regulation, especially regarding inhibi-
direct measure of the potential risk of failure. If the damage, tion and activation of osteoclastic resorption [70–74]. As
and particularly its rate of increase, is greater than can be soon as an osteocyte network is damaged and cells undergo
repaired, then failure will occur unless adaptation is apoptosis, the inhibitory signal is abolished and osteoclasts
initiated to reduce the stress level. An advantage of this are released for resorption [75].
approach is that it automatically accounts for the dynamic If bone is the activation signal to trigger BMU activity,
loading history as the driving force for the remodelling how can bone cells detect the presence of cracks? If all
process. However, there remains a lack of knowledge of microcracks were removed from the matrix, the mechanical
how bone detects damage and ‘decides’ to initiate repair or integrity of bone might be compromised since tunnelling
adaptation. BMUs would weaken the structure. On the other hand, if
microdamage remained undetected, various microcracks
might propagate to form macrocracks leading to fracture.
Detection of microdamage by osteocytes So how do the osteocytes decide whether a crack is
dangerous enough to require repair? If this principle were
Recently, Qui et al. [62] found that regions with osteocyte extended, and the rate of microdamage accumulation were
densities <728/mm2, such as interstitial bone, are 3.8-times so high that the remodelling process could not keep up,
more likely to contain microdamage than regions with then the only way to reduce propagation would be to
higher osteocyte densities. The questions remain: if micro- deposit new bone at surfaces in order to reduce strain
damage can trigger bone adaptation and what may be the levels. But how do the osteocytes decide whether the level
role of osteocytes in this process? Using the isolated ulna of damage is high enough to require surface adaptation? A
loading model, intracortical bone remodelling in rats was possible answer to these questions might lie in damage to
shown by Bentolila et al. [45] to be triggered by fatigue the cell processes. Recently, a theoretical model was
loading. In this experiment, 14 out of a total of 16 rats developed based on this principle.
Osteoporos Int (2007) 18:1–8 5

A typical microcrack is elliptical in shape with the minor


axis having a length of 100 μm and the major axis
measuring lengths up to 700 μm, orientated at approxi-
mately 20° to the longitudinal axis of the bone [9]. Given
that human cortical bone contains between 13,900 and
19,400 osteocytes per mm3 [46], with 40–60 cell processes
each, it is likely that the microcrack crosses this network.
Since microcracks tend to propagate in a direction similar
to the local lamellar orientation, it would mean that cracks Fig. 3 A new schematic proposal for bone cell signalling in damage-
stimulated bone remodelling
subjected to tensile loading open up and produce a shear
displacement. Cracks subjected to compressive loading
would result in compression of the fracture surfaces and
an additional shear displacement, resulting in the rupture of are embedded in the hard mineralised matrix, limiting our
cell processes. Using linear elastic fracture mechanics and knowledge regarding bone regulation. The development of
various osteocyte densities, it was predicted that small the MLO-Y4 osteocyte cell line has provided more data for
microcracks (<30 μm) do not produce enough shear this particular. For example, Kuratu et al. [86] used a novel
displacement to rupture cell processes, independent of the approach to investigate the effect of microdamage on the
stress level to which these cracks were subjected. Micro- cellular network. In their study, damage was introduced to a
cracks with a typical length of 100 μm show ruptured cell three-dimensional osteocyte cell culture with a 21-gauge
processes at stress levels exceeding 28 MPa. If the stress needle. Microdamage was found to have no significant
levels are increased for cracks of this length, several effect on the number of dead cells compared with the
hundred cell processes will rupture. Extremely large cracks control. However, it was found that the bone marrow cells,
(>300 μm) tend to rupture several thousand cell processes which were seeded on top of the gel, were TRACP-positive
even at extremely low stress levels [66]. Given that hardly in a region close to the damage site. Furthermore, it was
any crack lengths in excess of 100 μm are reported in the shown, using an ELISA assay, that the gel-embedded
literature, it might indicate that cell process rupture plays an osteocytes secreted significant amounts of both M-CSF
important role in monitoring local microdamage accumula- and RANKL. In other words, if damage to the osteocyte
tion. Experimental evidence, in which propagating cracks processes causes up-regulation of RANKL, then this would
were monitored under a UV-epiflourescence microscope provide a new pathway to activate osteoclastic activity (see
with additional staining of the cytoskeleton, showed that Fig. 3) and so link microdamage, resorption and formation
the majority of cell processes rupture between the two crack of BMUs.
faces. Only near the crack tip, a region where crack-face In summary, it has been know for half a century that
displacements are negligible, did cell processes remain bones contain microcracks [6]. Their initiation, propagation
intact [76]. and arrest are affected by microstructural features [20, 87]
and the nano-composition of bone [37]. There is evidence
to suggest that microcracks are repaired by a targeted
remodelling process [43, 45, 88]. How this process is
Biochemistry of bone cell signalling regulated at a cellular level still remains unknown. One
proposed mechanism has focused on ruptured osteocyte cell
The response of osteocytes to damaged cell processes processes affecting signalling pathways [66, 89] and
influences the cell-to-cell interaction and, therefore, the causing up-regulation of RANKL [86]. The second mech-
cellular network biochemical homeostasis in bone. The anism proposed is that microdamage accumulation causes
bone remodelling cycle is controlled by various systemic osteocyte apoptosis, which triggers osteoclastic activity [22,
factors (see Fig. 3). TGF-β, for example, is released and 67, 75]. Although these two mechanisms differ, a common
activated by resorbing osteoclasts [77] and promotes pathway might be present where ruptured cell processes
different stages of maturation of osteoblasts [78], which is trigger osteocyte apoptosis, causing an up-regulation of
also controlled by IGF [79]. Osteoblasts produce osteopro- RANKL which stimulates osteoclastic activity and the
tegerin (OPG) and express RANKL on their surfaces. The formation of BMUs.
secretion of OPG and the binding of RANK-RANKL
stimulate the differentiation and activation of osteoclasts
[80]. While osteoclasts, osteoblasts and the signalling
Acknowledgement This work was financially supported through the
pathways between them have received much attention EMBARK Postdoctoral Fellowship by the Irish Research Council for
[81–85], osteocytes are less well understood since they Science.
6 Osteoporos Int (2007) 18:1–8

References 23. Norman TL, Wang Z (1997) Microdamage of human cortical


bone: incidence and morphology in long bones. Bone 20:375–
379
1. Wenzel TE, Schaffler MB, Fyhrie DP (1996) In vivo trabecular 24. O’Brien FJ, Taylor D, Lee TC (2002) An improved labelling
microcracks in human vertebral bone. Bone 19:89–95 technique for monitoring microcrack growth in compact bone.
2. Aloia JF, Vaswani A, Delerme-Pagan C, Flaster E (1998) J Biomech 35:523–526
Discordance between ultrasound of the calcaneus and bone 25. Martin RB, Burr DB (1982) A hypothetical mechanism for the
mineral density in black and white women. Calcif Tissue Int stimulation of osteonal remodelling by fatigue damage. J Biomech
62:481–485 15:137–139
3. Seeman E, Bianchi G, Adami S, Kanis J, Khosla S, Orwoll E 26. Akkus O, Rimnac CM (2001) Cortical bone tissue resists fatigue
(2004) Osteoporosis in men–consensus is premature. Calcif Tissue fracture by deceleration and arrest of microcrack growth. J Biomech
Int 75:120–122 34:757–764
4. Lewiecki EM (2004) Management of osteoporosis. Clin Mol 27. Hazenberg JG, Taylor D, Lee TC (2006) Mechanisms of short
Allergy 2:9 crack growth at constant stress in bone. Biomaterials 27:2114–
5. Riggs BL, Melton LJ 3rd (1995) The worldwide problem of 2122
osteoporosis: insights afforded by epidemiology. Bone 17:505S– 28. Nalla RK, Kinney JH, Ritchie RO (2003) Effect of orientation on
511S the in vitro fracture toughness of dentin: the role of toughening
6. Frost HM (1960) Presence of microscopic cracks in vivo in bone. mechanisms. Biomaterials 24:3955–3968
Henry Ford Hosp Bull 8:25–35 29. Nalla RK, Kinney JH, Ritchie RO (2003) Mechanistic fracture
7. Burr DB, Stafford T (1990) Validity of the bulk-staining technique criteria for the failure of human cortical bone. Nat Mater 2:164–
to separate artifactual from in vivo bone microdamage. Clin 168
Orthop 305–308 30. Nalla RK, Kruzic JJ, Ritchie RO (2004) On the origin of the
8. Lee TC, Myers ER, Hayes WC (1998) Fluorescence-aided toughness of mineralized tissue: microcracking or crack bridging?
detection of microdamage in compact bone. J Anat 193(Pt 2):179– Bone 34:790–798
184 31. Malik CL, Stover SM, Martin RB, Gibeling JC (2003) Equine
9. O’Brien FJ, Taylor D, Dickson GR, Lee TC (2000) Visualisation cortical bone exhibits rising R-curve fracture mechanics. J Biomech
of three-dimensional microcracks in compact bone. J Anat 197 36:191–198
(Pt 3):413–420 32. Nalla RK, Kruzic JJ, Kinney JH, Ritchie RO (2005) Mechanistic
10. Norman TL, Vashishth D, Burr DB (1995) Fracture toughness of aspects of fracture and R-curve behavior in human cortical bone.
human bone under tension. J Biomech 28:309–320 Biomaterials 26:217–231
11. Yeni YN, Brown CU, Wang Z, Norman TL (1997) The influence 33. Vashishth D, Koontz J, Qiu SJ, Lundin-Cannon D, Yeni YN,
of bone morphology on fracture toughness of the human femur Schaffler MB, Fyhrie DP (2000) In vivo diffuse damage in human
and tibia. Bone 21:453–459 vertebral trabecular bone. Bone 26:147–152
12. Feng Z, Rou J, Han S, Ziv I (2000) Orientation and loading 34. Vashishth D, Tanner KE, Bonfield W (2003) Experimental
condition dependence of fracture toughness in cortical bone. validation of a microcracking-based toughening mechanism for
Mater Sci Eng C 11:41–46 cortical bone. J Biomech 36:121–124
13. Zioupos P, X TW, Currey JD (1996) The accumulation of fatigue 35. Vashishth D (2004) Rising crack-growth-resistance behavior in
microdamage in human cortical bone of two different ages in cortical bone: implications for toughness measurements. J Bio-
vitro. Clin Biomech (Bristol, Avon) 11:365–375 mech 37:943–946
14. Lee TC, Arthur TL, Gibson LJ, Hayes WC (2000) Sequential 36. Ritchie RO (1988) Mechanisms of fatigue crack propagation in
labelling of microdamage in bone using chelating agents. J Orthop metals, ceramics and composites: role of crack-tip shielding.
Res 18:322–325 Mater Sci Eng A103:15–28
15. O’Brien FJ, Taylor D, Lee TC (2003) Microcrack accumulation at 37. Yeni YN, Fyhrie DP (2003) A rate-dependent microcrack-
different intervals during fatigue testing of compact bone. J Biomech bridging model that can explain the strain rate dependency of
36:973–980 cortical bone apparent yield strength. J Biomech 36:1343–
16. Burr DB, Martin RB (1993) Calculating the probability that 1353
microcracks initiate resorption spaces. J Biomech 26:613– 38. Hazenberg JG, Taylor D, Clive Lee T (2006) Mechanisms of short
616 crack growth at constant stress in bone. Biomaterials 27:2114–
17. Reilly GC, Currey JD (1999) The development of microcracking 2122
and failure in bone depends on the loading mode to which it is 39. Frost HM (1987) Bone “mass” and the “mechanostat”: a proposal.
adapted. J Exp Biol 202(Pt 5):543–552 Anat Rec 219:1–9
18. Reilly GC (2000) Observations of microdamage around osteocyte 40. Palumbo C, Palazzini S, Marotti G (1990) Morphological study of
lacunae in bone. J Biomech 33:1131–1134 intercellular junctions during osteocyte differentiation. Bone
19. Carter DR, Hayes WC (1977) Compact bone fatigue damage: a 11:401–406
microscopic examination. Clin Orthop 265–274 41. Viceconti M, Seireg A (1990) A generalized procedure for
20. Vashishth D, Behiri JC, Bonfield W (1997) Crack growth predicting bone mass regulation by mechanical strain. Calcif
resistance in cortical bone: concept of microcrack toughening. Tissue Int 47:296–301
J Biomech 30:763–769 42. Prendergast PJ, Taylor D (1994) Prediction of bone adaptation
21. Mohsin S, O’Brien FJ, Lee TC (2006) Osteonal crack barriers in using damage accumulation. J Biomech 27:1067–1076
ovine compact bone. J Anat 208:81–89 43. Martin RB, Stover SM, Gibson VA, Gibeling JC, Griffin LV
22. Schaffler MB, Choi K, Milgrom C (1995) Aging and matrix (1996) In vitro fatigue behaviour of the equine third metacarpus:
microdamage accumulation in human compact bone. Bone remodeling and microcrack damage analysis. J Orthop Res
17:521–525 14:794–801
Osteoporos Int (2007) 18:1–8 7

44. Burr DB, Forwood MR, Fyhrie DP, Martin RB, Schaffler MB, 65. Turner CH, Owan I, Takano Y (1995) Mechanotransduction in
Turner CH (1997) Bone microdamage and skeletal fragility in bone: role of strain rate. Am J Physiol 269:E438–E442
osteoporotic and stress fractures. J Bone Miner Res 12:6–15 66. Taylor D, Hazenberg JG, Lee TC (2003) The cellular transducer in
45. Bentolila V, Boyce TM, Fyhrie DP, Drumb R, Skerry TM, damage-stimulated bone remodelling: a theoretical investigation
Schaffler MB (1998) Intracortical remodeling in adult rat long using fracture mechanics. J Theor Biol 225:65–75
bones after fatigue loading. Bone 23:275–281 67. Noble BS, Reeve J (2000) Osteocyte function, osteocyte death
46. Sissons HA, O’Connor P (1977) Quantitative histology of and bone fracture resistance. Mol Cell Endocrinol 159:7–13
osteocyte lacunae in normal human cortical bone. Calcif Tissue 68. Verborgt O, Gibson GJ, Schaffler MB (2000) Loss of osteocyte
Res 22 (Suppl):530–533 integrity in association with microdamage and bone remodeling
47. Qin L, Mak AT, Cheng CW, Hung LK, Chan KM (1999) after fatigue in vivo. J Bone Miner Res 15:60–67
Histomorphological study on pattern of fluid movement in cortical 69. Noble BS, Peet N, Stevens HY, Brabbs A, Mosley JR, Reilly GC,
bone in goats. Anat Rec 255:380–387 Reeve J, Skerry TM, Lanyon LE (2003) Mechanical loading:
48. Marotti G, Ferretti M, Muglia MA, Palumbo C, Palazzini S (1992) biphasic osteocyte survival and targeting of osteoclasts for bone
A quantitative evaluation of osteoblast-osteocyte relationships on destruction in rat cortical bone. Am J Physiol Cell Physiol 284:
growing endosteal surface of rabbit tibiae. Bone 13:363–368 C934–C943
49. Martin RB (2000) Does osteocyte formation cause the nonlinear 70. Shimizu H, Sakamoto M, Sakamoto S (1990) Bone resorption
refilling of osteons? Bone 26:71–78 by isolated osteoclasts in living versus devitalized bone: differ-
50. Metz LN, Martin RB, Turner AS (2003) Histomorphometric ences in mode and extent and the effects of human recombinant
analysis of the effects of osteocyte density on osteonal morphol- tissue inhibitor of metalloproteinases. J Bone Miner Res 5:
ogy and remodeling. Bone 33:753–759 411–418
51. Kusuzaki K, Kageyama N, Shinjo H, Takeshita H, Murata H, 71. Maejima-Ikeda A, Aoki M, Tsuritani K, Kamioka K, Hiura K,
Hashiguchi S, Ashihara T, Hirasawa Y (2000) Development of Miyoshi T, Hara H, Takano-Yamamoto T, Kumegawa M (1997)
bone canaliculi during bone repair. Bone 27:655–659 Chick osteocyte-derived protein inhibits osteoclastic bone resorp-
52. Parfitt AM (1984) The cellular basis of bone remodeling: the tion. Biochem J 322(Pt 1):245–250
quantum concept reexamined in light of recent advances in the 72. Plotkin LI, Weinstein RS, Parfitt AM, Roberson PK, Manolagas
cell biology of bone. Calcif Tissue Int 36(Suppl 1):S37–S45 SC, Bellido T (1999) Prevention of osteocyte and osteoblast
53. Hernandez CJ, Majeska RJ, Schaffler MB (2004) Osteocyte apoptosis by bisphosphonates and calcitonin. J Clin Invest
density in woven bone. Bone 35:1095–1099 104:1363–1374
54. Tanaka-Kamioka K, Kamioka H, Ris H, Lim SS (1998) Osteocyte 73. Kogianni G, Mann V, Ebetino F, Nuttall M, Nijweide P, Simpson
shape is dependent on actin filaments and osteocyte processes are H, Noble B (2004) Fas/CD95 is associated with glucocorticoid-
unique actin-rich projections. J Bone Miner Res 13:1555–1568 induced osteocyte apoptosis. Life Sci 75:2879–2895
55. Kamioka H, Honjo T, Takano-Yamamoto T (2001) A three- 74. Liu Y, Porta A, Peng X, Gengaro K, Cunningham EB, Li H,
dimensional distribution of osteocyte processes revealed by the Dominguez LA, Bellido T, Christakos S (2004) Prevention of
combination of confocal laser scanning microscopy and differen- glucocorticoid-induced apoptosis in osteocytes and osteoblasts by
tial interference contrast microscopy. Bone 28:145–149 calbindin-D28k. J Bone Miner Res 19:479–490
56. Sugawara Y, Kamioka H, Honjo T, Tezuka K, Takano-Yamamoto 75. Gu G, Mulari M, Peng Z, Hentunen TA, Vaananen HK (2005)
T (2005) Three-dimensional reconstruction of chick calvarial Death of osteocytes turns off the inhibition of osteoclasts and
osteocytes and their cell processes using confocal microscopy. triggers local bone resorption. Biochem Biophys Res Commun
Bone 36:877–883 335:1095–1101
57. You LD, Weinbaum S, Cowin SC, Schaffler MB (2004) 76. Hazenberg JG, Freeley M, Foran E, Lee TC, Taylor D (2005)
Ultrastructure of the osteocyte process and its pericellular matrix. Microdamage: A cell transducing mechanism based on ruptured
Anat Rec 278A:505–513 osteocyte processes. J Biomech
58. Cho H, Stout SD, Madsen RW, Streeter MA (2002) Population- 77. Bonewald LF (1999) Establishment and characterization of an
specific histological age-estimating method: a model for known osteocyte-like cell line, MLO-Y4. J Bone Miner Metab 17:61–65
African-American and European-American skeletal remains. 78. Simmons ED Jr, Pritzker KP, Grynpas MD (1991) Age-related
J Forensic Sci 47:12–18 changes in the human femoral cortex. J Orthop Res 9:155–167
59. Qiu S, Rao DS, Palnitkar S, Parfitt AM (2005) Differences in 79. Machwate M, Zerath E, Holy X, Pastoureau P, Marie PJ (1994)
osteocyte and lacunar density between Black and White American Insulin-like growth factor-I increases trabecular bone formation
women. Bone and osteoblastic cell proliferation in unloaded rats. Endocrinology
60. Beck TJ, Ruff CB, Shaffer RA, Betsinger K, Trone DW, Brodine 134:1031–1038
SK (2000) Stress fracture in military recruits: gender differences 80. Burger EH, Klein-Nulend J (1999) Mechanotransduction in bone:
in muscle and bone susceptibility factors. Bone 27:437–444 role of the lacuno-canalicular network. FASEB J 13(Suppl):S101–
61. Martin RB (1995) Mathematical model for repair of fatigue S112
damage and stress fracture in osteonal bone. J Orthop Res 81. Pensler JM, Patel PK, Langman CB (1997) Osteoblast-directed
13:309–316 osteoclast metabolism from patients with premature coronal
62. Qiu S, Sudhaker Rao D, Fyhrie DP, Palnitkar S, Parfitt AM (2005) synostosis. Plast Reconstr Surg 99:1518–1521
The morphological association between microcracks and osteo- 82. Gay CV, Gilman VR, Sugiyama T (2000) Perspectives on
cyte lacunae in human cortical bone. Bone 37:10–15 osteoblast and osteoclast function. Poult Sci 79:1005–1008
63. Hsieh YF, Turner CH (2001) Effects of loading frequency on 83. Shimoaka T, Ogasawara T, Yonamine A, Chikazu D, Kawano H,
mechanically induced bone formation. J Bone Miner Res 16:918–924 Nakamura K, Itoh N, Kawaguchi H (2002) Regulation of
64. Lee KC, Maxwell A, LE Lanyon (2002) Validation of a technique osteoblast, chondrocyte, and osteoclast functions by fibroblast
for studying functional adaptation of the mouse ulna in response growth factor (FGF)-18 in comparison with FGF-2 and FGF-10.
to mechanical loading. Bone 31:407–412 J Biol Chem 277:7493–7500
8 Osteoporos Int (2007) 18:1–8

84. Phan TC, Xu J, Zheng MH (2004) Interaction between osteoblast 87. Currey JD, Brear K, Zioupos P (1996) The effects of ageing and
and osteoclast: impact in bone disease. Histol Histopathol changes in mineral content in degrading the toughness of human
19:1325–1344 femora. J Biomech 29:257–260
85. Horowitz MC, Bothwell AL, Hesslein DG, Pflugh DL, Schatz DG 88. Burr DB, Milgrom C, Boyd RD, Higgins WL, Robin G, Radin
(2005) B cells and osteoblast and osteoclast development. EL (1990) Experimental stress fractures of the tibia. Biological
Immunol Rev 208:141–153 and mechanical aetiology in rabbits. J Bone Joint Surg Br
86. Kurata K, Heino HJ, Higaki H, Vaananen HK (2006) Bone 72:370–375
marrow cell differentiation induced by mechanically damaged 89. Hazenberg JG, Freeley M, Foran E, Lee TC, Taylor D (2006)
osteocytes in 3D gel-embedded culture. J Biomed Mater Res Microdamage: a cell transducing mechanism based on ruptured
21:616–625 osteocyte processes. J Biomech 39:2096–2103

View publication stats

You might also like