Anda di halaman 1dari 8

Seminars in Cell & Developmental Biology 19 (2008) 459–466

Contents lists available at ScienceDirect

Seminars in Cell & Developmental Biology


journal homepage: www.elsevier.com/locate/semcdb

Review

Bone remodeling during fracture repair: The cellular picture


Aaron Schindeler a,b,∗ , Michelle M. McDonald a , Paul Bokko a , David G. Little a,b
a
Department of Orthopaedic Research & Biotechnology, The Children’s Hospital at Westmead, Sydney, Australia
b
Discipline of Paediatrics and Child Health, Faculty of Medicine, University of Sydney, Sydney, Australia

a r t i c l e i n f o a b s t r a c t

Article history: Fracture healing is a complex event that involves the coordination of a variety of different processes.
Available online 25 July 2008 Repair is typically characterized by four overlapping stages: the initial inflammatory response, soft callus
formation, hard callus formation, initial bony union and bone remodeling. However, repair can also be
Keywords: seen to represent a juxtaposition of two distinct forces: anabolism or tissue formation, and catabolism or
Bone remodeling remodeling. These anabolic/catabolic concepts are useful for understanding bone repair without giving
Fracture
the false impression of temporally distinct stages that operate independently. They are also relevant when
Fracture healing
considering intervention.
Bone repair
Bisphosphonates
In normal bone development, bone remodeling conventionally refers to the removal of calcified bone
tissue by osteoclasts. However, in the context of bone repair there are two phases of tissue catabolism:
the removal of the initial cartilaginous soft callus, followed by the eventual remodeling of the bony
hard callus. In this review, we have attempted to examine catabolism/remodeling in fractures in a sys-
tematic fashion. The first section briefly summarizes the traditional four-stage view of fracture repair
in a physiological manner. The second section highlights some of the limitations of using a temporal
rather than process-driven model and summarizes the anabolic/catabolic paradigm of fracture repair.
The third section examines the cellular participants in soft callus remodeling and in particular the role
of the osteoclast in endochondral ossification. Finally, the fourth section examines the effects of delaying
osteoclast-dependent hard callus remodeling and also poses questions regarding the crosstalk between
anabolism and catabolism in the latter stages of fracture repair.
© 2008 Elsevier Ltd. All rights reserved.

Contents

1. An overview of fracture repair: the four-stage model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 459


1.1. Stage 1: inflammation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 460
1.2. Stage 2: soft callus (fibrocartilage) formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 460
1.3. Stage 3: hard callus formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 461
1.4. Stage 4: bone remodeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 461
2. The anabolic/catabolic model of fracture repair . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 462
3. Soft callus remodeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 462
4. Hard callus remodeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 463
5. Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 464
Acknowledgment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 464
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 464

1. An overview of fracture repair: the four-stage model

∗ Corresponding author at: Orthopaedic Research & Biotechnology, Research


Fracture healing is a process that recapitulates certain aspects of
Building, The Children’s Hospital at Westmead, Locked Bag 4001, Westmead, NSW
skeletal development and growth involving a complex interplay of
2145, Australia. Tel.: +61 2 98450117; fax: +61 2 98453078. cells, growth factors, and extracellular matrix. Repair is convention-
E-mail address: AaronS@chw.edu.au (A. Schindeler). ally partitioned into four stages, each characterized by a specific set

1084-9521/$ – see front matter © 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.semcdb.2008.07.004
460 A. Schindeler et al. / Seminars in Cell & Developmental Biology 19 (2008) 459–466

of cellular and molecular events [1–3]. But in practical terms there activation of non-specific wound healing pathways that accom-
is rarely such a clear delineation of events; there is often signif- pany non-skeletal injuries. The extravasation (bleeding) within the
icant overlap between the different stages during fracture repair. fracture site is contained by the surrounding tissue and devel-
The four-stage model can still be a useful system for describing the ops into a hematoma. Degranulating platelets, macrophages, and
basic events that take place. other inflammatory cells (granulocytes, lymphocytes, and mono-
The four-stage model originated as the result of histological cytes) infiltrate the hematoma between the fractured fragments
observations of healing fractures in both human patients and ani- and combat infection, secrete cytokines and growth factors, and
mal models (Fig. 1A). However, research of the past several decades advance clotting into a fibrinous thrombus [1,2]. Over time, cap-
has explored both the cellular and molecular forces that drive the illaries grow into the clot, which is reorganized into granulation
underlying processes [4–6]. At the cellular level, inflammatory cells, tissue. Macrophages, giant cells and other phagocytic cells clear
vascular cells, osteochondral progenitors, and osteoclasts are key degenerated cells and other debris.
players in the repair process. At the molecular level, fracture repair This cellular response is coordinated by and involves the
is driven by the three main classes of factors: pro-inflammatory secretion of a range of cytokines and growth factors including
cytokines and growth factors, pro-osteogenic factors, and angio- transforming growth factor-␤ (TGF-␤), platelet-derived growth fac-
genic factors [7]. These help establish the relevant morphogenetic tor (PDGF), fibroblast growth factor-2 (FGF-2), vascular endothelial
fields by recruiting cells and stimulating growth and/or differ- growth factor (VEGF), macrophage colony stimulating factor (M-
entiation. Thereafter, the damaged soft tissues are repaired and CSF), interleukins-1 and -6 (IL-1 and -6), bone morphogenetic
the fracture is bridged by soft callus and later hard callus. The proteins (BMPs), and tumor necrosis factor-␣ (TNF-␣) [1–3]. This
bridging hard callus is eventually remodeled to re-establish the factors facilitate the recruitment of additional inflammatory cells
original geometry and function of the damaged tissue. In certain in a positive feedback loop, and also the migration and invasion
circumstances where there is absolute rigidity, or in well-apposed of multipotent mesenchymal stem cells [8,9]. Stem cells originat-
metaphyseal fractures, intramembranous hard callus formation ing from the periosteum [10], bone marrow [11,12], circulation [13],
may dominate without the clear presence of an intermediate car- and the surrounding soft tissues [14], have been implicated in bone
tilaginous soft callus. formation and repair.

1.1. Stage 1: inflammation 1.2. Stage 2: soft callus (fibrocartilage) formation

A fracture is typically associated with disruption of the local Most fractures possess some level of mechanical instability and
soft tissue integrity, interruption to normal vascular function, and heal by the process of endochondral ossification—bony callus for-
a distortion of the marrow architecture. This damage leads to mation is preceded by a cartilaginous template [1]. This stage is

Fig. 1. Models of fracture repair and the cellular participants. (A) A representative series of images of the four-stage model of fracture healing. Between the classical stages
2 and 3, the soft callus is systematically remodeled. (B) A more elaborate version of the anabolic/catabolic model of fracture repair [36] that incorporates the concepts of
non-specific anabolism (the early wound repair processes) and non-specific catabolism (soft callus remodeling). (C) A schematic showing the cellular contributors to the
fracture repair processes. The source(s) of mesenchymal progenitors that are able to differentiate into osteoblasts remain ambiguous.
A. Schindeler et al. / Seminars in Cell & Developmental Biology 19 (2008) 459–466 461

dominated on a cellular level by chondrocytes and fibroblasts, cells able to produce bone and bone healing is often delayed when
although the relative proportions of the different cell types can vary the periosteum is damaged or stripped [10,21]. The bone marrow
between fractures. These cells produce a semi-rigid soft callus that also contains mesenchymal stem cells capable of contributing to
is able to provide mechanical support to the fracture, as well as act bone formation during repair [11,12]. However, bones are able to
as a template for the bony callus that will later supersede it. Car- mount an effort at healing, including hard callus formation, even in
tilaginous callus is principally avascular, although its subsequent the absence of periosteal and/or marrow osteoprogenitors. Thus it is
replacement with woven bone involves vascular invasion. possible that osteoprogenitors may originate from multiple sources
Chondrocytes derived from mesenchymal progenitors prolif- including the circulation [13], the vasculature [22], and surrounding
erate and synthesize cartilaginous matrix until all the fibri- local tissues [14].
nous/granulation tissue is replaced by cartilage. Where cartilage The vasculature is known to be critical for formation of the
production is deficient, fibroblasts replace the region with gener- hard callus, with increased oxygen tension in the local region
alized fibrous tissue. Discrete cartilaginous regions progressively necessary for osteoblast differentiation. Stimulation of vessel for-
grow and merge to produce a central fibrocartilaginous plug mation using angiogenic factors can augment bone formation
between the fractured fragments that splints the fracture [7]. In and fracture healing in model systems [23,24]. However, recent
the final stages of soft callus production, the chondrocytes undergo studies have shown that delayed union can occur even after the
hypertrophy and mineralize the cartilaginous matrix before under- vasculature has been re-established, indicating that the blood
going apoptosis. supply alone is not the only determinant of fracture healing
Fibroblast proliferation and chondrocyte prolifera- success [25–27].
tion/differentiation are stimulated by the coordinated expression
of growth factors including TGF-␤2 and -␤3, PDGF, fibroblast 1.4. Stage 4: bone remodeling
growth factor-1 (FGF-1), and insulin-like growth factor (IGF)
[1,2]. Additionally, members of the BMP family (BMP-2, -4, -5 The final stage of fracture repair encompasses the remodeling
and -6) assist in promoting cell proliferation and chondrogenesis. of the woven bone hard callus into the original cortical and/or
In response to these factors, chondrocytes are able to generate trabecular bone configuration. This phase can also be referred to
considerable amounts of extracellular matrix proteins, particu- as secondary bone formation [1]. Initially, this involves converting
larly collagen II (or collagen X for hypertrophic chondrocytes) the irregular woven bone callus into lamellar bone, although the
[4]. standard cortical structure is eventually restored. The remodeling
Invasion of the soft callus by vascular endothelial cells, angio- process is driven by a coupled process of orderly bone resorption
genesis, and capillary in-growth is stimulated by pro-angiogenic followed by the formation of lamellar bone. Osteoblast-like cells
factors including VEGF, BMPs, FGF-1 and TGF-␤ [1,15,16]. VEGF may also play a minor role in proteolysis of osteoid elements, prior
expression by osteogenic cells is dependent on the master bone to new bone formation [28]. The process of neo-vascularization is
and cartilage regulatory factor Cbfa1/Runx2 [17], and is not highly maintained.
expressed by the early mesenchymal/inflammatory tissue at the The key cell type involved with the resorption of mineralized
fracture site [18]. In addition, angiopoietin I and II regulate vascu- bone is the osteoclast. The osteoclast is a large multinucleated cell
lar morphogenesis of larger vessels and development of collateral that is formed by the differentiation and fusion of haematopoietic
branches from existent vessels [1,4]. precursors [29]. To effect remodeling, osteoclasts become polarized
and adhere to a mineralized surface. They form a ruffled border,
1.3. Stage 3: hard callus formation which is sealed off and acid and proteinases are pumped into
the resorption domain. The acid environment demineralizes the
Also known as primary bone formation [1], this stage represents matrix, while proteinases degrade the organic components, such as
most active period of osteogenesis. It is characterized by high lev- collagen. The degradation products are removed through a vesic-
els of osteoblast activity and the formation of mineralized bone ular pathway from the ruffled border to the functional secretory
matrix, which arises directly in the peripheral callus in areas of sta- domain and the osteoclasts are either apoptosed or return to the
bility. In order for bridging new hard callus to form, the insecure non-resorbing form. Bone resorption by osteoclasts creates erosive
soft callus is gradually removed, concomitant with revasculariza- pits on the bone surface known as “Howship’s lacuna”. Once com-
tion. The new bone is known as the hard callus and it is typically pleted, osteoblasts are able to lay down new bone on the eroded
irregular and under-remodeled. Notably, hard callus can form in surface.
the absence of a cartilaginous template in intramembranous bone Two principal cytokines that are secreted by osteoblasts are
formation (during conditions of high mechanical stability) or in critical for the induction, survival and competency of osteoclasts:
appositional bone growth, where bone forms directly adjacent M-CSF and Receptor Activator of NF␬B Ligand (RANKL). M-CSF is
to an existing mineralized surface. However, in the majority of important for the primary induction of haematopoietic stem cell
orthopaedic instances, some level of endochondral ossification is differentiation towards an osteoclast lineage [30]. RANKL is a factor
present. produced by mature osteoblasts that is responsible for the coordi-
The initial woven bone matrix contains a combination of pro- nation of bone formation and bone resorption [31]. Osteoprotegrin
teinaceous and mineralized extracellular matrix tissue. This is (OPG) is a secreted decoy receptor that is an important regulator
synthesized by mature osteoblasts, which differentiate from osteo- of RANKL signaling and antagonizes osteoclast differentiation [32].
progenitors in the presence of osteogenic factors. Members of In addition to these specialized pro-osteoclastic factors, a range
the BMP family are critical mediators of this process, and have of growth factors and cytokines present in a healing fracture are
been shown to be sufficient for de novo bone formation [19,20]. known to promote osteoclastogenesis. These include interleukins,
Other growth factors are expressed during this time, although it TNF-␣, BMPs, and TGF-␤ [33–35].
is unclear whether their impact on bone healing predominantly The molecular pathways involved in osteoclast differentiation
affects osteoprogenitor migration/proliferation versus osteogenic and function are becoming increasingly well described. Many of
differentiation [1]. these molecules have been linked to genetic diseases affecting
The source(s) of osteoprogenitors responsible for fracture heal- osteoclasts, and/or identified as putative drug targets for osteoporo-
ing remains ambiguous. The periosteum is a rich source of stem sis therapies.
462 A. Schindeler et al. / Seminars in Cell & Developmental Biology 19 (2008) 459–466

2. The anabolic/catabolic model of fracture repair relevant as anti-resorptive drugs, including but not limited to
bisphosphonates (BPs), are widely used in the osteoporotic popula-
Although the sequential four-stage model describes the funda- tion. Should BPs or other interventions delay soft callus remodeling
mental events that occur over the timeline of a healing fracture, in osteoporotic fractures, this effect would be of considerable clin-
there are often significant overlaps between stages. For example, ical concern [42].
replacement of peripheral callus with lamellar bone remodeling One clinical case study suggested that the delayed or poor bone
occurs concurrently with cartilage mineralisation, vascular inva- healing observed in nine patients with osteoporosis related bone
sion and woven bone formation. Moreover, numerous fundamental loss was a result of their chronic treatment with BPs [43]. Although
events like cell proliferation, differentiation, matrix synthesis at the findings of this study showed extensive reductions in systemic
cellular level is not limited to any one stage [1]. bone formation in these patients, the direct link from the BP treat-
Applying the concepts of anabolism and catabolism may provide ment to the poor bone healing cannot be disproved, fuelling the
a useful alternative system for understanding the fracture repair concerns of clinicians treating osteoporotic patients with fractures.
process [36]. In this system, the outcome of the fracture repair pro- A clear and defined role for osteoclast activity during initial endo-
cess exists as a balance between the anabolic (bone forming) and chondral bone union after fracture therefore must be rigorously
catabolic (bone resorbing) responses. We have previously used this evaluated.
approach to describe hard callus remodeling [36]. However, as we With a focus on the clinical questions surrounding BP treatment
will go on to describe, anabolic/catabolic processes also occur dur- during periods of bone healing, our group has recently examined
ing the formation and removal of the fibrocartilage soft callus. Many the effect of the two BPs, zoledronic acid (ZA) and pamidronate
of the cell types recruited to the early stage fracture site and the (PAM), on endochondral union in a rat closed fracture model
associated regulatory growth factors and cytokines are seen in a [44,45]. In these studies, both therapeutic and extremely high doses
generalized response to tissue injury and are thus non-specific to of these agents failed to interfere with the achievement of endo-
bone. Consequently, we propose that these processes be referred to chondral union. These outcomes suggest that BPs may be safely
as non-specific anabolism and non-specific catabolism (Fig. 1B). We administered to patients undergoing periods of bone repair. How-
speculate that the speed of fracture healing may be determined by ever, as healthy growing rats were used, care needs to be made in
the processes of non-specific anabolism and catabolism (recruiting translating these results to the treatment of osteoporotic patients.
cells, revascularisation), while the strength of repair relates to the Osteoporotic animals show impaired fracture repair [46–48], and
mechanically driven balance between bone-specific anabolism and thus investigating the influence of BPs in osteoporotic rodent mod-
catabolism. els would be highly relevant.
Events in fracture healing are considerably influenced by numer- Based on this recent data, we propose a new model regarding
ous physiological and pharmacological factors in the prevailing osteoclast function and bone repair: that osteoclastic resorptive
environment that determine the state of the fracture and the pace function may be redundant during initial endochondral bone union.
of repair. Key features include the nature location and extent of In this model, soft callus remodeling is a non-specific catabolic
injury, the biomechanical forces, infection and/or disease, surgi- process that can involve the contribution of multiple cell types.
cal fixation, adjunctive drug therapies, nutrition and general bone In support of this hypothesis we have also examined endochon-
health, as well as any underlying genetic conditions. All of these fac- dral bone healing in the incisor absent (ia/ia) rat. These rats exhibit
tors can be described in terms of their effects on non-specific and extensive osteopetrosis due to harboring a mutant osteoclast pop-
specific anabolism and/or catabolism. The utility of this approach ulation with reduced functional activity [49]. Within the period
is that as pro-anabolic and anti-catabolic bone therapies become prior to recovery from osteopetrosis, fractures performed in these
more advanced, treatment approaches can be targeted to produce rats reached endochondral union within the same time frame as
the maximal benefit-dependent on the underlying bone healing their normal littermates [50].
deficiency [36]. This model is further supported by experiments using genet-
ically modified mice undertaken to examine the functional role
3. Soft callus remodeling of osteoclasts in endochondral bone repair. When fracture heal-
ing was examined in the osteoclast-deficient osteopetrotic (op/op)
Soft callus remodeling occurs between stages 2 and 3 of the tra- mouse, the removal of the cartilage callus was identical between
ditional four-stage model of fracture repair. This process involves mutant mice and normal littermates [51]. These investigators went
the gradual removal of the soft callus cartilage/fibrocartilage and its on to administer recombinant osteoprotegerin (OPG) as a pharma-
systematic replacement with woven bone [37]. Failure to remove cological means of inhibiting osteoclast formation. OPG acts as a
this tissue can lead to delay or arrest of the fracture repair process, decoy receptor for receptor activator of NF␬B (RANK) signaling, and
thus considerable research has been undertaken to understand the antagonizes the differentiation of osteoclast progenitors. Despite
cellular determinants of the process. It was traditionally believed very low osteoclast number, OPG-treated mice showed compara-
that osteoclasts were the key cell type involved with both soft callus ble union rates to untreated control mice [51]. Furthermore, neither
and hard callus remodeling [38]. However, recent evidence suggests of two recent studies examining fracture repair in rats using OPG
that osteoclasts may be somewhat redundant in the remodeling of treatment reported a delay in the achievement of initial endochon-
the soft callus. dral union [52,53].
The suggestion that osteoclasts may be involved with soft cal- One of the few examples of impaired initial union in the absence
lus remodeling is not without basis. There is histological evidence of osteoclasts involved Rank−/− knockout mice [51]. This mouse line
that large multinucleated cells are present within the soft cal- completely lacks osteoclasts, leading to a grossly osteopetrotic phe-
lus together with vascular endothelial cells at the invasion front notype such that tibial fractures could not be fixated as per the
[38–40]. These cells are also termed chondroclasts based on their control mice. Unstabilized Rank−/− mouse fractures formed abun-
physical location, although they stain for classic osteoclastic mark- dant soft callus, but hard callus formation was significantly delayed.
ers and exhibit a similar morphology [41]. While the authors suggested that this could be due to reduced bone
However, if osteoclasts are indeed critical for soft callus remod- formation and/or revascularization, it is unclear whether cartilage
eling, anti-osteoclastic agents could be expected to interfere with removal could have also been delayed. More likely, the magnitude of
this process and possibly delay fracture repair. This is clinically soft callus that formed in the absence of any fixation overwhelmed
A. Schindeler et al. / Seminars in Cell & Developmental Biology 19 (2008) 459–466 463

the normal soft callus remodeling process. Certainly, when osteo- can also affect a variety of cell signaling responses and cell fate
clast formation was blocked in wild type mice using an anti-RANK decisions [63]. Finally, the formation of new vessels may allow the
ligand antibody, this did not effect soft callus formation or remod- invasion of additional cells from the circulation to contribute to soft
eling, or the union rates of stabilized fractures [51]. callus remodeling.
If osteoclast function is redundant during endochondral fracture
repair then the role of other cell lineages must be considered. Such a
cell type must (i) be present in the soft callus during the remodeling 4. Hard callus remodeling
process and (ii) secrete factors able to break down collagen and
other extracellular matrix proteins found in soft callus cartilage. While osteoclasts may not be essential in the initial stages of
A group of proteases known as matrix metalloproteinases endochondral bone repair, remodeling of the hard callus is an
(MMPs) have been implicated as a fundamental class of col- osteoclast-dependent process. This is most clearly seen in fracture
lagenases and gelatinases responsible for extracellular matrix experiments where osteoclast activity is abolished following treat-
degradation. Secretion of MMPs has been demonstrated by a num- ment with a bisphosphonate. While there are conflicting reports
ber of cell types involved in or localized to the chondro-osseous regarding the overall benefit versus risk of BPs, all reports unequiv-
junction (the boundary between cartilage and bone tissue) dur- ocally show an inhibition of hard callus remodeling.
ing endochondral bone repair. MMP-9 co-localizes to cells of The literature contains numerous animal fracture models
the chondro-osseous junction, and has been shown to co-stain involving BP treatment that can vary in their selection of surgi-
with both inflammatory cells and osteoclasts [54,55]. MMP-13 co- cal site, fixation, and BP dosing regimen. Nevertheless, a number
expresses with markers for pre-osteoblasts, mature osteoblasts and of common themes emerge. Many of these studies found little or
hypertrophic chondrocytes. Unlike MMP-9 and MMP-13 has not no effect on the ultimate radiographic union rates or mechanical
been localized in osteoclasts. The absence of either MMP-9 or MMP- strength, but that remodeling of the hard callus to a normal lamel-
13 impedes soft callus removal, causing a delay in union during lar bone structure was delayed [64–66]. High dose local application
endochondral fracture repair [55,56]. MMP-14 is another abundant can ablate the bone formation response, but this is not through
metalloproteinase that plays a critical role in growth plate chondro- an osteoclastic mechanism, with decreased angiogenesis and pri-
cytes and may therefore also be relevant to endochondral fracture mary bone formation being observed [67]. Two clinical studies have
repair [57]. These studies support the concept that multiple MMP- indicated that BP treatment might inhibit initial union [43,68].
expressing cell types, not only osteoclasts, may be able to contribute However, these papers refer to examples of chronic and/or high
to soft callus remodeling. dose BP treatment that are likely to result in an overall decrease in
MMPs have also been strongly implicated in driving angiogene- bone turnover, rather than a specific inhibition of osteoclast func-
sis during tumor invasion and bone development [58]. The invasion tion.
of new blood vessels into the avascular hypertrophic cartilage of the One of the first studies to show a potential benefit for delay-
soft callus is a critical step in the process of endochondral bone ing hard callus remodeling used a single dose of PAM to treat
repair. Structural and immunohistochemical staining of healing a rat femoral open fracture model [69]. In this study, BP was
fractures showed capillaries lined with laminin positive basement administered via systemic injection (3 mg/kg s.c.) or locally by a
membranes containing endothelial cells adjacent to the hyper- poly-l-lactide-coated K-wire (0.1 or 1.0 mg) used for intramedullary
trophic chondrocyte matrix at the invasion front [59]. Perivascular fixation. Both delivery systems resulted in significant increases in
cells were commonly observed lining these vessels, where they callus volume and bone mineral density (BMD), as well as improve-
were suggested to participate in degrading the chondral matrix. ments in mechanical strength at early time points. A subsequent
Significantly, both vascular endothelial cells and perivascular cells study focusing on timing of anti-catabolic intervention was per-
are able to secrete MMPs. formed using ZA in a rat femoral-closed fracture model [70]. This
Angiogenesis has been deemed to be of critical importance study found that hard callus retention was maximized when ZA
to skeletal repair. TNP-470 is potent inhibitor of angiogenesis; was given at 2 weeks following fracture. When the drug was local-
it inhibits a specific isoform of methionine aminopeptidase and ized using 14C-labelled ZA, more was present in the callus of the
blocks the proliferation of vascular cells. Rat femoral fractures 2-week post-fracture dosed samples, supporting the concept that
treated with TNP-470 showed severely impaired healing [60]. BPs bind to areas of active bone formation/resorption [70].
Intramembranous and endochondral ossification were disturbed Although BPs can delay hard callus remodeling, this delay does
by the lack of angiogenesis. The early formation of fracture blastema not necessarily lead to negative functional outcomes. A remodeled
was also effected, with TNP-470-treated rats producing dense lamellar callus may have superior properties per unit area of bone,
fibrous tissue instead of a normal blastema. Similarly, inhibition but a larger woven bone callus may provide a similar mechanical
of VEGF signaling using a soluble, neutralizing VEGF receptor (Flt- advantage. In a rat closed femoral fracture study where rats were
IgG) resulted in a delay in the vascular invasion and replacement either pre-dosed or pre- and post-dosed with daily incadronate,
of cartilaginous callus with bone responsible for the transition two different results were observed. Pre-dosing alone led to a
from soft callus to hard callus [61]. This is highly relevant to clini- greater initial volume of hard callus, however this then remod-
cal situations where patients receiving drugs with anti-angiogenic eled to form a normal cortical shell over 6–16 weeks. In contrast,
effects sustain fractures. A recent bone healing study examined continuous dosing maintained a disorganized woven bone cal-
the effects of Rapamycin, immunosuppressive agent with known lus. While disordered, the large callus observed in the high-dose
anti-angiogenic properties. Rapamycin-treated mice exhibited a continuous group showed greater mechanical strength (ultimate
significant delay in endochondral repair. Calluses in treated mice load) in 3-point bending tests compared to any of the remodeled
were significantly smaller and exhibited increased fibrocartilagi- lamellar calluses [65]. In an analogous study using a similar frac-
nous tissue compared to untreated controls [62]. ture model, rats were dosed with a single bolus dose of ZA versus
Nevertheless, the precise role of angiogenic cells in soft callus multiple intermittent ZA doses, both led to an increase in bone
remodeling remains ambiguous. While vascular endothelial cells strength at 6 weeks [44]. Again, the callus properties of the two
and perivascular cells may secrete MMPs that degrade cartilage, dosing groups were divergent: the bolus dosing group had par-
like osteoclasts they may be redundant to the process. The process tially remodeled and possessed a cortex while the intermittent
of vascular invasion causes a change in oxygen tension that itself dosing group exhibited a substantial woven bone callus. Critically,
464 A. Schindeler et al. / Seminars in Cell & Developmental Biology 19 (2008) 459–466

the mechanical strength of both treatment groups did not differ sig- The secretion of sclerostin (the SOST gene product) by osteo-
nificantly, showing that delayed remodeling did not compromise cytes has emerged as an important regulator of bone anabolism
strength. [82,83]. SOST expression in osteocytes is upregulated in regions of
While BPs act to suppress the bone resorptive ability of osteo- disuse, which would be expected to reduce bone anabolism [84].
clasts, they do not prevent osteoclast formation or signaling While sclerostin is important for the systemic regulation of bone
crosstalk between osteoclasts and other cell types. However, in homeostasis it is unclear whether it plays a specific role in fracture
mice treated with 10 mg/kg human osteoprotegerin (hOPG), osteo- remodeling. SOST is expressed by cell types present in the remod-
clastogenesis was inhibited and osteoclast numbers were reduced eling callus (although not the early phases of bone repair) [85].
in healing tibial fractures by 93% at 3 weeks (P < 0.001) and 92% at Thus sclerostin may important for the normal downregulation of
8 weeks (P < 0.001) [52]. OPG treatment did not affect the strength specific anabolism in the callus, and may also represent a poten-
of the fractures, but decreased the material properties of the callus. tial pharmacological target for modulating the anabolic response
This reinforces the long-held concept that osteoclast-dependent during fracture remodeling in orthopaedic medicine.
bone remodeling is responsible for transforming a sizable woven
bone callus into a compact, mechanically efficient lamellar struc- 5. Concluding remarks
ture. As an alternative to pharmaceutical intervention, genetic
models of osteoclast dysfunction have been pursued. Bone repair Fracture repair is a complex and highly regulated process that
experiments in these mouse and rat models have yielded compara- is influenced by physiological, cellular, and molecular/genetic fac-
ble results to BP treatment. For example, following a closed fracture tors. Remodeling of the fracture callus is fundamental to the process
callus size was enhanced in the osteopetrotic ia/ia rats and hard of repair. This includes remodeling of the soft callus (non-specific
callus remodeling was also delayed [44,71,72]. catabolism), which is a largely osteoclast-independent process, and
Although hard callus remodeling is primarily driven by osteo- remodeling of the hard callus, which is an osteoclast-dependent
clasts, it is essential that a parallel anabolic activity is maintained process (specific catabolism). Future work is required to better
by osteoblasts. In the remodeling phase, the process of anabolism define the cell types capable of remodeling soft callus in the absence
is coupled to bone resorption. This is in contrast to the early of functional osteoclasts. The ability of anti-osteoclastic agents
anabolic response that forms the initial callus, which is primarily to modulate the catabolic response may be of medical utility in
cytokine/growth factor-driven and occurs in the absence of bone orthopaedics if a delay in remodeling may lead to stronger initial
resorptive forces. union. Finally, the coordinated regulation of hard callus remodeling
Coupling between anabolic and catabolic forces is intrinsic to by osteoblasts has been highlighted as an area for further study.
normal bone anabolism. When osteoclast formation is altered,
some compensatory anabolic change is typically observed. For
Acknowledgment
instance, Opg−/− mice develop excess osteoclasts but this is par-
tially offset by increased bone formation [73]. Conversely, bone
Our research into the mechanisms underlying fracture repair has
formation is reduced in c-Fos mutant mice that lack osteoclasts
been supported by the National Health & Medical Research Council
[74]. Due to the complexity of accurately measuring bone forma-
(NHMRC).
tion and resorption in a healing fracture, there are few examples
of anabolic/catabolic coupling in fracture repair. In osteogenesis
References
imperfecta patients on long-term PAM treatment, bone catabolism
is reduced and the anabolic response in bone healing is simi- [1] Gerstenfeld LC, Cullinane DM, Barnes GL, Graves DT, Einhorn TA. Fracture heal-
larly impaired [68]. In the osteoclast-deficient op/op mice, both ing as a post-natal developmental process: molecular, spatial, and temporal
anabolic and catabolic responses are reduced in the growth plate, aspects of its regulation. J Cell Biochem 2003;88:873–84.
[2] Einhorn TA. The cell and molecular biology of fracture healing. Clin Orthop Relat
a site of endochondral ossification analogous to a healing fracture
Res 1998;355:S7–21.
[75]. [3] Bolander ME. Regulation of fracture repair by growth factors. Proc Soc Exp Biol
Nevertheless, the precise mechanism underlying the coupling Med 1992;200:165–70.
[4] AI-Aql ZS, Alagl AS, Graves DT, Gerstenfeld LC, Einhorn TA. Molecular mech-
of anabolic and catabolic responses during remodeling remains
anisms controlling bone formation during fracture healing and distraction.
ambiguous [76]. Several models have been suggested to explain Osteogenesis. J Dent Res 2008;87:107–18.
the coupling process during remodeling. Active bone resorption is [5] Giannoudis PV, Einhorn TA, Marsh D. Fracture healing: the diamond concept.
capable of releasing numerous growth factors and osteoid break- Injury 2007;38:S3–6.
[6] Stone CA. A molecular approach to bone regeneration. Br J Plast Surg
down products including BMPs, TGF␤, and collagen peptides. These 1997;1997:369–73.
products may be responsible for stimulating local anabolism in [7] Barnes GL, Kostenuik PJ, Gerstenfeld LC, Einhorn TA. Growth factor regulation
regions of high bone resorption [77]. However, rodent models have of fracture repair. J Bone Miner Res 1999;14:1805–15.
[8] Iwaki A, Jingushi S, Oda Y, Izumi T, Shida JI, Tsuneyoshi M, et al. Localization
revealed that anabolism can remain coupled in the presence of non- and quantification of proliferating cells during rat fracture repair: detection of
functional osteoclasts, suggesting that the release of factors from proliferating cell nuclear antigen by immunohistochemistry. J Bone Miner Res
the bone matrix may not be critical [78,79]. Thus it is possible that 1997;12:96–102.
[9] Shapiro F. Bone development and its relation to fracture repair. The role of mes-
signaling molecules directly secreted by osteoclasts, such as osteo- enchymal osteoblasts and surface osteoblasts. Eur Cells Mater 2008;15:53–76.
clast inhibitory lectin (OCIL) [80], may be important for stimulating [10] Malizos KN, Papatheodorou LK. The healing potential of the periosteum. Molec-
osteoblast function. ular aspects. Injury 2005;36(Suppl 3):S13–9.
[11] Baksh D, Song L, Tuan RS. Adult mesenchymal stem cells: characterization,
In addition to anabolic/catabolic coupling, the process of remod-
differentiation, and application in cell and gene therapy. J Cell Mol Med
eling is modulated by mechanical forces. This tenet of orthopaedics 2004;8:301–16.
(Wolff’s law) stipulates that a healthy bone adapts to the load that it [12] Colnot C, Huang S, Helms J. Analyzing the cellular contribution of bone mar-
row to fracture healing using bone marrow transplantation in mice. Biochem
is placed under. This principle can manifest in many ways in fracture
Biophys Res Commun 2006;350:557–61.
repair. For instance, bone healing is impaired in unloaded situations [13] Eghbali-Fatourechi GZ, Lamsam J, Fraser D, Nagel D, Riggs BL, Khosla S. Circu-
such as disuse and/or rigid fixation. The osteocyte is a key cellu- lating osteoblast-lineage cells in humans. N Engl J Med 2005;352:1959–66.
lar player involved in mechanosensing in bone and signals from [14] Rumi MN, Deol GS, Singapuri KP, Pellegrini Jr VD. The origin of osteoprogenitor
cells responsible for heterotopic ossification following hip surgery: an animal
the osteocyte may be transduced to control both anabolism and model in the rabbit. J Orthop Res 2005;23:34–40.
catabolism [81]. [15] Kalfas IH. Principles of bone healing. Neurosurg Focus 2001;10:E1.
A. Schindeler et al. / Seminars in Cell & Developmental Biology 19 (2008) 459–466 465

[16] Deckers MML, van Bezooijen RL, van der Horst G, Hoogendam J, van der Bent [47] Meyer Jr RA, Tsahakis PJ, Martin DF. Age and ovariectomy impair both the nor-
C, Papapoulos SE, et al. Bone morphogenetic proteins stimulate angiogenesis malization of mechanical properties and the accretion of mineral by the fracture
through osteoblast-derived vascular endothelial growth factor A. Endocrinol- callus in rats. J Orthop Res 2001;19:428–35.
ogy 2002;143:1545–53. [48] Walsh WR, Sherman P, Howlett CR. Fracture healing in a rat osteopenia model.
[17] Zelzer E, Glotzer DJ, Hartmann C, Thomas D, Fukai N, Soker S, et al. Tissue specific Clin Orthop Relat Res 1997;342:218–27.
regulation of VEGF expression during bone development requires Cbfa1/Runx2. [49] Reinholt FP, Hultenby K, Heinegard D, Marks Jr SC, Norgard M, Anderson G.
Mech Dev 2001;106:97–106. Extensive clear zone and defective ruffled border formation in osteoclasts
[18] Pufe T, Wildemann B, Petersen W, Mentlein R, Raschke M, Schmidmaier G. of osteopetrotic (ia/ia) rats: implications for secretory function. Exp Cell Res
Quantitative measurement of the splice variants 120 and 164 of the angiogenic 1991;251:477–91.
peptide vascular endothelial growth factor in the time flow of fracture healing: [50] McDonald MM, Morse A, Mikulec K, Godfrey C, Sztynda T, Little DG. The
a study in the rat. Cell Tissue Res 2002;309:387–92. osteopetrotic Incisor absent rat exhibits reduced osteoclast activity, resulting
[19] Chen D, Zhao M, Mundy GR. Bone morphogenetic proteins. Growth Factors in normal endochondral fracture union but delayed hard callus remodelling. J
2004;22:233–41. Bone Miner Res 2007;22(Suppl 1):S253.
[20] Nakase T, Yoshikawa H. Potential roles of bone morphogenetic proteins (BMPs) [51] Flick LM, Weaver JM, Ulrich-Vinther M, Abuzzahab F, Zhang X, Dougall WC,
in skeletal repair and regeneration. J Bone Miner Metab 2006;24:425–33. et al. Effects of receptor activator of NFkappaB (RANK) signaling blockade on
[21] Hutmacher DW, Sittinger M. Periosteal cells in bone tissue engineering. Tissue fracture healing. J Orthop Res 2003;21:676–84.
Eng 2003;9(Suppl 1):S45–64. [52] Ulrich-Vinther M, Andreassen TT. Osteoprotegerin treatment impairs remod-
[22] Collett GD, Canfield AE. Angiogenesis and pericytes in the initiation of ectopic eling and apparent material properties of callus tissue without influencing
calcification. Circ Res 2005;96:930–8. structural fracture strength. Calcif Tissue Int 2005;76(4):280–6.
[23] Peng H, Usas A, Olshanski A, Ho AM, Gearhart B, Cooper GM, et al. [53] Ulrich-Vinther M, Schwarz EM, Pedersen FS, Soballe K, Andreassen TT. Gene
VEGF improves, whereas sFlt1 inhibits, BMP2-induced bone formation therapy with human osteoprotegerin decreases callus remodeling with limited
and bone healing through modulation of angiogenesis. J Bone Miner Res effects on biomechanical properties. Bone 2005;37:751–8.
2005;20:2017–27. [54] Vu TH, Shipley M, Bergers G, Berger JE, Helms JA, Hanahan D, et al. MMP-
[24] Tarkka T, Sipola A, Jämsä T, Soini Y, Ylä-Herttuala S, Tuukkanen J, et al. Adenovi- 9/Gelatinase B is a key regulator of growth plate angiogenesis and apoptosis of
ral VEGF-A gene transfer induces angiogenesis and promotes bone formation hypertrophic chondrocytes. Cell 1998;93:411–22.
in healing osseous tissues. J Gene Med 2003;5:560–6. [55] Colnot C, Thompson Z, Miclau T, Werb Z, Helms JA. Altered fracture repair in
[25] Brownlow HC, Reed A, Simpson AH. The vascularity of atrophic non-unions. the absence of MMP9. Development 2003;130:4123–33.
Injury 2002;33:145–50. [56] Behonick DJ, Xing Z, Lieu S, Buckley JM, Lotz C, Marcucio RS, et al. Role of Matrix
[26] Reed AA, Joyner CJ, Brownlow HC, Simpson AH. Human atrophic fracture non- metalloproteinase 13 in both endochondral and intramembranous ossification
unions are not avascular. J Orthop Res 2002;20:593–9. during skeletal regeneration. Plos One 2007;11:e1150.
[27] Harry LE, Sandison A, Paleolog EM, Hansen U, McCarthy ID, Pearse MF, et al. [57] Malemud CJ. Matrix metalloproteinases: role in skeletal development and
The effect of muscle and fasciocutaneous flaps on angiogenesis in open tibial growth plate disorders. Front Biosci 2006;11:1702–15.
fractures. J Orthop Trauma 2004;18(9 Suppl):S15. [58] Tang Y, Nakada MT, Kesavan P, McCabe F, Millar H, Rafferty P, et al. Extracellular
[28] Mulari MT, Qu Q, Härkönen PL, Väänänen HK. Osteoblast-like cells complete matrix metalloproteinase inducer stimulates tumor angiogenesis by elevating
osteoclastic bone resorption and form new mineralized bone matrix in vitro. vascular endothelial growth factor and matrix metalloproteinases. Cancer Res
Calcif Tissue Int 2004;75:253–61. 2005;65:3193–9.
[29] Teitelbaum SL. Bone resorption by osteoclasts. Science 2000;289:1504–8. [59] Mark H, Penington A, Nannmark U, Morrison W, Messina A. Microvascular inva-
[30] Fan X, Biskobing DM, Fan D, Hofstetter W, Rubin J. Macrophage colony stimulat- sion during endochondral ossification in experimental fractures in rats. Bone
ing factor down-regulates MCSF-receptor expression and entry of progenitors 2004;35:535–42.
into the osteoclast lineage. J Bone Miner Res 1997;12:1387–95. [60] Hausman MR, Schaffler MB, Majeska RJ. Prevention of fracture healing in rats
[31] Kong YY, Yoshida H, Sarosi I, Tan HL, Timms E, Capparelli C, et al. OPGL is a by an inhibitor of angiogenesis. Bone 2001;29:560–4.
key regulator of osteoclastogenesis, lymphocyte development and lymph-node [61] Street J, Bao M, Deguzman L, Bunting S, Peale Jr FV, Ferrara N, et al. Vascular
organogenesis. Nature 1999;397:315–23. endothelial growth factor stimulates bone repair by promoting angiogenesis
[32] Blair JM, Zhou H, Seibel MJ, Dunstan CR. Mechanisms of disease: roles of OPG, and bone turnover. Proc Natl Acad Sci USA 2002;99:9656–61.
RANKL and RANK in the pathophysiology of skeletal metastasis. Nat Clin Pract [62] Holsetin JH, Klein M, Garcia P, Histing T, Culemann U, Pizanis A, et al. Rapamycin
Oncol 2006;3:41–9. Affects early fracture healing in mice. Br J Pharmacol 2008 [ePub May 5, 2008].
[33] Quinn JM, Gillespie MT. Modulation of osteoclast formation. Biochem Biophys [63] Wan C, Gilbert SR, Wang Y, Cao X, Shen X, Ramaswamy G, et al. Activation of the
Res Commun 2005;328:739–45. hypoxia-inducible factor-1alpha pathway accelerates bone regeneration. Proc
[34] Lee SK, Lorenzo J. Cytokines regulating osteoclast formation and function. Curr Natl Acad Sci USA 2008;105:686–91.
Opin Rheumatol 2006;18:411–8. [64] Tarvainen R, Olkkonen H, Nevalainen T, Hyvönen P, Arnala I, Alhava E.
[35] Kaneko H, Arakawa T, Mano H, Kaneda T, Ogasawara A, Nakagawa M, et al. Effect of clodronate on fracture healing in denervated rats. Bone 1994;15:
Direct stimulation of osteoclastic bone resorption by bone morphogenetic pro- 701–5.
tein (BMP)-2 and expression of BMP receptors in mature osteoclasts. Bone [65] Li J, Mori S, Kaji Y, Mashiba T, Kawanishi J, Norimatsu H. Effect of bisphospho-
2000;27:479–86. nate (incadronate) on fracture healing of long bones in rats. J Bone Miner Res
[36] Little DG, Ramachandran M, Schindeler A. The anabolic and catabolic responses 1999;14:969–79.
in bone repair. J Bone Joint Surg [Br] 2007;89:425–33. [66] Li J, Mori S, Kaji Y, Kawanishi J, Akiyama T, Norimatsu H. Concentration of bis-
[37] Einhorn TA. The science of fracture healing. 1. J Orthop Trauma 2005;19(Suppl phosphonate (incadronate) in callus area and its effects on fracture healing in
10):S4–6. rats. J Bone Miner Res 2000;15:2042–51.
[38] Schell H, Lienau J, Epari DR, Seebeck P, Exner C, Muchow S, et al. Osteoclas- [67] Choi JY, Kim HJ, Lee YC, Cho BO, Seong HS, Cho M, et al. Inhibition of bone healing
tic activity begins early and increases over the course of bone healing. Bone by pamidronate in calvarial bony defects. Oral Surg Oral Med Oral Pathol Oral
2006;38:547–54. Radiol Endod 2007;103:321–8.
[39] Sawae Y, Sahara T, Sasaki T. Osteoclast differentiation at growth plate cartilage- [68] Munns CF, Rauch F, Travers R, Glorieux FH. Effects of intravenous pamidronate
trabecular bone junction in newborn rat femur. J Electron Microsc (Tokyo) treatment in infants with osteogenesis imperfecta: clinical and histomorpho-
2003;52:493–502. metric outcome. J Bone Miner Res 2005;20:1235–43.
[40] Takahara M, Naruse T, Takagi M, Orui H, Ogino T. Matrix metalloproteinase-9 [69] Amanat N, Brown R, Bilston LE, Little DG. A single systemic dose of pamidronate
expression, tartrate-resistant acid phosphatase activity, and DNA fragmen- improves bone mineral content and accelerates restoration of strength in a rat
tation in vascular and cellular invasion into cartilage preceding primary model of fracture repair. J Orthop Res 2005;23:1029–34.
endochondral ossification in long bones. J Orthop Res 2004;22:1050–7. [70] Amanat N, McDonald M, Godfrey C, Bilston L, Little D. Optimal timing of a single
[41] Cole AA, Walters LM. Tartrate-resistant acid phosphatase in bone and cartilage dose of zoledronic acid to increase strength in rat fracture repair. J Bone Miner
following decalcification and cold-embedding in plastic. J Histochem Cytochem Res 2007;22:867–76.
1987;35:203–6. [71] Schmidt CJ, Marks Jr SC, Jordan CA, Hawes LE. A radiographic and histo-
[42] McDonald MM, Little DG, Schindeler A. Bisphosphonate treatment and fracture logic study of fracture healing in osteopetrotic rats. Radiology 1977;122:
repair. BoneKey Osteovision 2007;9:236–51. 517–9.
[43] Odvina CV, Zerwekh JE, Rao DS, Maalouf N, Gottschalk FA, Pak CY. Severely [72] Marks Jr SC, Schmidt CJ. Bone remodeling as an expression of altered pheno-
suppressed bone turnover: a potential complication of alendronate therapy. J type: studies of fracture healing in untreated and cured osteopetrotic rats. Clin
Clin Endocrinol Metab 2005;90:1294–301. Orthop Relat Res 1978;137:259–64.
[44] McDonald MM, Dulai S, Godfrey C, Amanat N, Sztynda T, Little DG. Bolus or [73] Nakamura M, Udagawa N, Matsuura S, Mogi M, Nakamura H, Horiuchi H, et al.
weekly zoledronic acid administration does not delay endochondral fracture Osteoprotegerin regulates bone formation through a coupling mechanism with
repair but weekly dosing enhances delays in hard callus remodelling. Bone; bone resorption. Endocrinology 2003;144:5441–9.
2008 Jun 3. [Epub ahead of print]. [74] Grigoriadis AE, Wang ZQ, Cecchini MG, Hofstetter W, Felix R, Fleisch HA, et al. c-
[45] McDonald MM, Morse A, Mikulec K, Mai H, Munns C, Little DG. Normal endo- Fos: a key regulator of osteoclast-macrophage lineage determination and bone
chondral bone healing follows continuous bisphosphonate pre-treatment in a remodeling. Science 1994;266:443–8.
rat closed fracture model. J Bone Miner Res 2007;22(Suppl 1):S253. [75] Nishino I, Amizuka N, Ozawa H. Histochemical examination of osteoblastic
[46] Namkung-Matthai H, Appleyard R, Jansen J. Osteoporosis influences the early activity in op/op mice with or without injection of recombinant M-CSF. J Bone
period of fracture healing in a rat osteoporotic model. Bone 2001;28:80–6. Miner Metab 2001;19:267–76.
466 A. Schindeler et al. / Seminars in Cell & Developmental Biology 19 (2008) 459–466

[76] Martin TJ, Sims NA. Osteoclast-derived activity in the coupling of bone forma- [81] Rubin C, Judex S, Hadjiargyrou M. Skeletal adaptation to mechanical stimuli
tion to resorption. Trends Mol Med 2005;11:76–81. in the absence of formation or resorption of bone. J Musculoskelet Neuronal
[77] Mohan S, Baylink DJ. The role of IGF-II in the coupling of bone formation to Interact 2002;2:264–7.
resorption. In: Modern concepts of insulin-like growth factors (ed. Spencer, [82] van Bezooijen RL, ten Dijke P, Papapoulos SE, Löwik CW. SOST/sclerostin, an
EM) 1991; p. 19–174. osteocyte-derived negative regulator of bone formation. Cytokine Growth Fac-
[78] Marzia M, Sims NA, Voit S, Migliaccio S, Taranta A, Bernardini S, et al. Decreased tor Rev 2005;16:319–27.
c-Src expression enhances osteoblast differentiation and bone formation. J Cell [83] Winkler DG, Sutherland MK, Geoghegan JC, Yu C, Hayes T, Skonier JE, et al.
Biol 2000;151:311–20. Osteocyte control of bone formation via sclerostin, a novel BMP antagonist.
[79] Schaller S, Henriksen K, Sveigaard C, Heegaard AM, Hélix N, Stahlhut M, et al. EMBO J 2003;22:6267–76.
The chloride channel inhibitor NS3736 prevents bone resorption in ovariec- [84] Robling AG, Niziolek PJ, Baldridge LA, Condon KW, Allen MR, Alam I, et
tomized rats without changing bone formation. J Bone Miner Res 2004;19: al. Mechanical stimulation of bone in vivo reduces osteocyte expression of
1144–53. Sost/sclerostin. J Biol Chem 2008;283:5866–75.
[80] Nakamura A, Ly C, Cipetić M, Sims NA, Vieusseux J, Kartsogiannis V, et al. Osteo- [85] van Bezooijen RL, Roelen BA, Visser A, van der Wee-Pals L, de Wilt E, Karpe-
clast inhibitory lectin (OCIL) inhibits osteoblast differentiation and function in rien M, et al. Sclerostin is an osteocyte-expressed negative regulator of bone
vitro. Bone 2007;40:305–15. formation, but not a classical BMP antagonist. J Exp Med 2004;199:805–14.

Anda mungkin juga menyukai