You are on page 1of 13

Biomaterials 21 (2000) 2347}2359

Biomaterial developments for bone tissue engineering


Karen J.L. Burg *, Scott Porter, James F. Kellam
Department of Bioengineering, Clemson University, 501 Rhodes Engineering Research Building, Clemson, SC 29634-0905, USA
Department of Orthopaedic Surgery, Carolinas Medical Center, Charlotte, NC 28232-2861, USA

Abstract
The development of bone tissue engineering is directly related to changes in materials technology. While the inclusion of materials
requirements is standard in the design process of engineered bone substitutes, it is also critical to incorporate clinical requirements in
order to engineer a clinically relevant device. This review presents the clinical need for bone tissue-engineered alternatives to the
present materials used in bone grafting techniques, a status report on clinically available bone tissue-engineering devices, and recent
advances in biomaterials research. The discussion of ongoing research includes the current state of osseoactive factors and the delivery
of these factors using bioceramics and absorbable biopolymers. Suggestions are also presented as to the desirable design features that
would make an engineered device clinically e!ective.  2000 Elsevier Science Ltd. All rights reserved.
Keywords: Absorbable; Bone morphogenetic protein; Bone tissue engineering; Ceramic; Demineralized bone matrix; Polymer

1. Introduction
1.1. Rationale for bone tissue-engineering
There are multiple clinical reasons to develop bone
tissue-engineering alternatives, including the need for
better "ller materials that can be used in the reconstruction of large orthopaedic defects and the need for orthopaedic implants that are mechanically more suitable to
their biological environment. The traditional biological
methods of bone-defect management include autografting and allografting cancellous bone, applying vascularized grafts of the "bula and iliac crest, and using
other bone transport techniques. Although these are the
standard treatments, shortcomings are encountered with
their usage. Since bone grafts are avascular and dependent on di!usion, the size of the defect and the viability of
the host bed can limit their application. Furthermore, the
new bone volume maintenance can be problematic due
to unpredictable bone resorption. In large defects, the
grafts can be resorbed by the body before osteogenesis is
complete [1,2]. Not only is the operating time required
for harvesting autografts expensive, but often the donor
tissue is scarce, and there can be signi"cant donor site
morbidity associated with infection, pain, and hematoma

* Corresponding author.

[3}7]. Allografting introduces the risk of disease and/or


infection; it may cause a lessening or complete loss of the
bone inductive factors [8]. Vascularized grafts require
a major microsurgical operative procedure requiring
a sophisticated infrastructure. Distraction osteogenesis
techniques are often laborious and lengthy processes that
are reserved for the most motivated patients [9,10]. Another method of bone defect repair is via bone cement
"llers. Bone cements are prepared in the operating room
and therefore can be susceptible to infection.
Bone marrow replacement is another possible tissueengineering application for the treatment of patients
following high-dose chemotherapy and/or radiation
treatment [11]. The acquisition of bone marrow requires
the sterile aspiration of the marrow from the posterior
iliac crest. Marrow may be used in tissue-engineering
culture and speci"cally as a basis for bone marrow expansion. The progenitor cells can be cultivated and selected as needed.
Bone tissue engineering may potentially provide alternative solutions that possess better mechanical properties than those used currently. This may decrease the
vascular insult of the implant to the bone and cause
less-stress shielding, perhaps decreasing the incidence of
implant-related osteopenia and subsequent refracture.
The mechanical properties of a bone tissue-engineered
construct could be modulated to avoid such sequelae.
A related application is the use of a tissue-engineered

0142-9612/00/$ - see front matter  2000 Elsevier Science Ltd. All rights reserved.
PII: S 0 1 4 2 - 9 6 1 2 ( 0 0 ) 0 0 1 0 2 - 2

2348

K.J.L. Burg et al. / Biomaterials 21 (2000) 2347}2359

implant surface to permanently stabilize implants by


coating the prosthesis with cells or tissue before implantation. This could be extremely useful in reconstructive orthopaedic surgeries that potentially have high
incidences of failure secondary to large bone defects [12].
Bone regeneration requires four components: a morphogenetic signal, responsive host cells that will respond
to the signal, a suitable carrier of this signal that can
deliver it to speci"c sites then serve as a sca!olding for
the growth of the responsive host cells, and a viable, well
vascularized host bed [13,14]. Bone tissue engineering,
for the purpose of this review, is the use of a sca!olding
material to either induce formation of bone from the
surrounding tissue or to act as a carrier or template for
implanted bone cells or other agents. Materials used as
bone tissue-engineered sca!olds may be injectable or
rigid, the latter requiring an operative implantation procedure. To this end, the areas of materials research can be
generally divided into acellular and cellular, with drug
delivery overlapping in both areas.
The "rst part of the research discussion in this review
will focus on acellular systems. Acellular will be classi"ed
as materials on or in which no additional cellular component is cultured. The discussion will include materials
that are clinically available as well as those materials that
have potential clinical use and are in the early stages of
research. Acellular materials will be regarded as solid,
absorbable "llers that will disappear over time, or porous
sca!olding that immediately allows room for bone
growth into the construct. The second part of the research discussion will focus on cellular systems. The
cellular materials are classi"ed, for this review, as scaffolds to which a cellular component is added prior to
implantation.
The third part of the research discussion will focus on
materials designed for drug delivery. The use of growth
factors has a potential to markedly increase sca!old
e!ectiveness. Transforming growth factor b and the
more recently discovered bone morphogenetic proteins
(BMPs) are integral in the regulation of embryologic
bone formation and post-traumatic bone healing/formation [13,15}17]. Researchers have demonstrated that
BMP can cause bone formation de novo if present in
su$cient quantity [18]. BMP has been delivered to localized sites in order to repair bone defects and nonunions
in several di!erent experimental models [19}33].
The materials commonly used in all three approaches
are ceramics, polymers or composites. The ceramics and
polymers are either absorbable or nonabsorbable, and
the polymers can be naturally derived materials or
synthesized materials. Bone tissue-engineering systems
have included demineralized bone matrix, collagen
composites, "brin, calcium phosphate, polylactide,
poly(lactide-co-glycolide), polylactide-polyethylene glycol,
hydroxyapatite, dental plaster, and titanium [13,
34}38].

2. Approaches toward bone tissue engineering


2.1. Desired features of a bone tissue-engineering material
Before considering the desired features of potential
tissue-engineering materials, it is useful to understand
two concepts of bone regeneration for tissue-engineering
constructs, speci"cally osteoconduction and osteoinduction. Osteoinduction is de"ned as the ability to cause
pluripotential cells, from a nonosseous environment to
di!erentiate into chondroctyes and osteoblasts, culminating in bone formation [39}41]. An osteoinductive material allows repair in a location that would normally not
heal if left untreated [8]. Osteoconduction supports ingrowth of capillaries and cells from the host into a threedimensional structure to form bone. An osteoconductive
material guides repair in a location where normal healing
will occur if left untreated.
Several years ago, researchers became aware that the
osteoconductive properties of the synthetic absorbable
polymers were dependent on their location and the structure of the polymer [42]. The tubular absorbable polymers used in long bone defects, for example, are believed
to promote bone growth by excluding the surrounding
soft tissue and its undesired cellular elements from the
defect, by maintaining an osteogenic-rich medullary environment within the defect, and by allowing direct bone
growth onto the polymer skeleton [42}45]. Polymers can
di!er in their molecular weight, polydispersity, crystallinity, and thermal transitions, allowing di!erent absorption rates. Their relative hydrophobicity and percent
crystallinity can a!ect cellular phenotype. Variations in
surface charge will a!ect cellular spreading or a$nity for
the surface, which can also cause changes in phenotypic
expression [46].
Local tissue responses to polymers in vivo depend on
the biocompatibility of the polymer as well as its degradative by-products [47]. The mechanism of erosion can
also a!ect the pH of the surrounding environment and
subsequent response. The absorbable polyesters, for
example, are largely hydrolysed through bulk erosion
[48}50]. Although early absorption times may demonstrate a stable pH [51], there is the potential for a sudden
decrease in local pH after a prolonged absorption time in
slower absorbing systems. Oxygen tension will also a!ect
the type of cell that proliferates; therefore, the correct
environment is very important [52].
Poly(lactide-co-glycolide) matrices are often used as
construct materials. They can be custom synthesized to
meet an absorption time requirement, and they are also
clinically familiar. There are several methods of processing these porous, synthetic matrices. The most common
method is solution cast, particulate-leached as developed
by Mikos [53]. In order to re"ne this method, it is quite
possible to modulate the pore topography and size to
suit a particular cell type, e.g., osteoblasts. Work in our

K.J.L. Burg et al. / Biomaterials 21 (2000) 2347}2359

2349

Fig. 1. (a,b) Bone topography (40;magni"cation) varies according to function and location. Photographs courtesy of L. Jenkins, HT (Clemson
University Department of Bioengineering). (c,d) Polylactide sca!old topography (20;magni"cation) may be varied by processing technique.

laboratory [54] as well as by others [55] has shown that


the pore shape can have a profound e!ect on the attachment and long-term survival of cells on a surface. Our
work has shown that, for a speci"c cell type, there is an
optimal pore topography that can be readily modulated
by careful selection of porogen. Since bone has very
di!erent structures depending on its function and location, it stands to reason that the same pore shape may
not be ideal for all potential uses. Fig. 1 reinforces this
point by demonstrating di!erent appearances of bone
and di!erent polymer sca!old topographies.
Pore size and tortuosity can be carefully modulated to
control the release of a material complexed to the polymer [56,57]. Pore size is also very cell-type speci"c.
Gogolewski's laboratory reported that, although polyester membranes with pore sizes up to 200 lm diameter
promoted bone growth within a 1-cm defect of the radii
of rabbits, smaller pore sizes promoted the most growth
[43]. Tsurga and coworkers have suggested that the
optimal pore size of ceramics that supports ectopic bone
formation is 300}400 lm [38]. Holmes similarly suggested that the optimal pore range is 200}400 lm with the
average human osteon size of approximately 223 lm
[58].

Porosity can, however, adversely a!ect important


mechanical characteristics of a polymer, requiring more
complex material designs. Increased hydroxyapatite
(HA) porosity, for example, decreases its malleability and
reduces its ability to conform to the irregular surfaces
that may be present in host bone [26,59,60]. Continuity
of the host bone/polymer interface is essential for osseointegration [26]. Koempel and coworkers determined
that re-engineering a porous HA to deliver rhBMP 2 resulted in nearly uniform "xation. The authors postulated
that the porous HA acted as a conduit for the BMPinduced growth of new bone allowing better "xation of
the HA composite despite its porosity [28].
Osteoblast proliferation is sensitive to surface topography [52], strain or other mechanical stimuli. The aspect
ratio of a material will in#uence the bony or "broblastic
di!erentiation of a tissue. Particle size, shape, and surface
roughness a!ect cellular adhesion, proliferation, and
phenotype. Cells can discriminate even the subtlest changes in topography, and they are most obviously sensitive to chemistry, topography, and surface energy. Such
surface features are particularly interesting when considering an absorbable material, since this is a dynamic
material, always presenting a new surface. Additionally,

2350

K.J.L. Burg et al. / Biomaterials 21 (2000) 2347}2359

Table 1
Select critical consideration in bone tissue-engineering sca!old design
[27,58,62]
Desirable qualities of a bone tissue-engineering sca!old
Available to surgeon on short
notice
Absorbs in predictable manner
in concert with bone growth
Adaptable to irregular wound
site, malleable
Maximal bone growth through
osteoinduction and/or
osteoconduction
Correct mechanical and
physical properties for
application
Good bony apposition

Promotes bone ingrowth


Does not induce soft tissue growth
at bone/implant interface
Average pore sizes approximately
200}400 lm
No detrimental e!ects to surrounding tissue due to processing
Sterilizable without loss of
properties
Absorbable with biocompatible
components

the surface energy may play a role in attracting particular


proteins to the surface of the material and, in turn, this
will a!ect the a$nity of the cells to the material. Table 1
is a summary of several critical considerations in the
design of a bone tissue-engineering material. Brekke [61]
provides a detailed list of desirable device characteristics
of a manufactured bone graft substitute.

3. Bone tissue-engineering materials research


3.1. Acellular systems
3.1.1. Acellular systems: naturally derived polymers
In the early 1960s, Urist and several co-workers realized that demineralized bone consistently induced bone
formation in ectopic tissues of experimental animals
[63}65]. A hydrochloric acid extraction process decalci"ed the bone matrix [64,66,67], producing demineralized
bone matrix (DBM), a compound that was later shown to
possess inherent osteoconductive and osteoinductive
properties [66]. Recent research has proven that cortical
bone is the preferred choice for DBM synthesis as it is
more osteoinductive with a lower antigenic potential
than cancellous bone [68]. Urist described the histological process by which new bone is formed under the
in#uence of DBM in ectopic sites [66]. The tissue is
initially in"ltrated by in#ammatory and mesenchymal
cells. Early angiogenesis, progenitor cells, and chondrocytes can be appreciated by 3 weeks. Shortly thereafter, osteoblasts, osteocytes, and chondrocytes appear to
lead to the synthesis of cartilage which is transformed
into woven bone over the next several weeks. By 4 weeks
post-implantation, osteoclasts and bone remodeling cells
are present, and the bone marrow is formed at about 4}6
weeks [14,66]. This process recapitulates the process of
endochondral ossi"cation.

Russell and Block point out that the processing of the


DBM can greatly in#uence the "nal osteoinductive ability [69]. Ethylene oxide, the sterilization agent for many
synthetic, absorbable devices [70], can render the DBM
completely devoid of osteoinductive potential, although
the osteoconductive potential would remain [69]. In
addition, they assert that ethanol is e!ective in reducing
the bacterial load without a!ecting the osteoinductive
potential of the matrix.
There has been a persistent orthopaedic interest in
DBM because of its therapeutic potential in the treatment of bony defects, nonunion, and its application in
joint fusion procedures. Russell and Block evaluated 21
studies that used DBM in the treatment of a variety of
clinical and experimental orthopaedic situations that
would have normally warranted treatment with standard
bone graft [69]. Even though the sterilization procedures
varied, they determined that more than 80% of the
authors reported favorable results with the use of DBM
regardless of the level of di$culty of the cases [69].
Tiedeman and coworkers were able to clinically demonstrate a 77% (30 of 39 patients) union rate for a series of
patients with non-unions, arthrodeses, acute fractures
with bone loss or comminution, and osseous or cavitary
defects that were treated with DBM plus autogenous
bone marrow aspirate [3]. Excluding the patients who
had a documented non-union before the study's inception, DBM increased the union rate for the remaining
patients to 90% (26 of 28 patients). These rates were
obtained despite the sterilization of DBM with ethylene
oxide [3].
Gepstein and coworkers examined the ability of DBM
to heal large long bone defects in rats [4]. In their study,
defects that were greater than 50% of the length of the
radial diaphysis were created in the bilateral front limbs
of 33 rats. At 21 days postoperatively, densitometry
measurements of the defect treated with DBM were within 91% of the native ulna for the 27 surviving rats. By 35
days, 71% of the rats had radiographic union, and the
other 29% had at least union of one end of the defect [4].
Urist and Daws obtained a low pseudarthrosis rate of
12% in their series of 40 patients with intertransverse
process fusion using autologous bone graft from the
spinous processes. The graft was harvested during the
surgical approach and supplemented with DBM. This
rate compared favorably to the reported rate in the best
series using iliac crest bone graft at that time. Furthermore, Urist and Daws were able to perform multilevel
fusions without the increase in morbidity common in
these procedures when the bone graft source is the iliac
crest [68].
MaK rtson's investigations of the use of viscose cellulose
sponge determined that it too is useful for bone tissue
engineering. Sponges placed in the femoral shafts of rats
showed that primarily osteoconduction occurred, since
bone formation was predominantly at the periphery of

K.J.L. Burg et al. / Biomaterials 21 (2000) 2347}2359

the sponges. No microstructural or physicochemical


characteristics of the sponges were given.
3.1.2. Acellular systems: synthetic polymers
Synthetic injectable materials are "nding appeal as
bone tissue-engineered sca!olds, due in part to their
minimally invasive implantation. Elissee! and coworkers
developed photopolymerizable materials that can be injected as liquids and photopolymerized to localize the
material [71]. Although they have tested this for transdermal application, with the increased interest in minimally invasive procedures, this may have orthopaedic
application as well.
Synthetic matrices have also been assessed as acellular
bone tissue-engineering materials. Gogolewski's group
used a poly-L-lactide (PLLA) membrane to cover 1-cm
defects that they created in the radii of skeletally mature
rabbits [45]. They chose a polymer with pore sizes of
5}15 lm diameter, a thickness of 250 lm, and an in vivo
life span of 18}24 months. At the conclusion of the study,
they had grossly and histologically demonstrated that
cortical bone had regenerated to span the defect [45]. In
another study by Gogolewski's group, they used the
Yucatan pig model to create a critical defect of 25% of
the length of the radius [72]. The defects were covered
with a PLLA or a poly-L-co-D,L-lactide (PLDL) membrane or with a PLLA or PLDL membrane that had
been synthesized with calcium carbonate in order to
decrease the relative amount of polymer in the membrane. They demonstrated that the membrane facilitated
the rapid formation of new bone growth without adverse
reactions. Importantly, the addition of calcium carbonate
did not change the dynamics of the membranes even
though the amount of actual polymer within them was
reduced by 50% [72]. Gugala and Gogolewski failed to
induce signi"cant bone formation when they used membranes with a pore size of 10}20 lm to cover 4-cm defects
in the tibiae of sheep [44]. They concluded that bone
defects that are greater than a critical size will not heal,
even if the synthetic membranes are used to facilitate
bone regeneration. When they used cancellous bone graft
in conjunction with synthetic membranes in the same
sized defect, they were able to induce signi"cant bone
healing. They postulated that, in addition to the
aforementioned roles synthetic membranes might play,
they could also optimize the contact between the soft
tissues and the bone graft and avoid the graft's excessive
resorption [44].
Another possible "ller material is poly-e-caprolactone-co-lactide. This has been observed in nonosseous
applications [73,74] but is newly applied to bone tissue
engineering as a paste or wax. Ekholm and coworkers
evaluated the absorption and biocompatibility of this
copolymer using a rat femoral defect model [75]. The
copolymer was speci"cally a paste of 40 : 60 poly-e-caprolactone-co-D,L-co-L-lactide with a 50 : 50 D-lactide to

2351

L-lactide

ratio. The material elicited a moderate in#ammatory response in bone and was still present after 1 yr in
this preliminary study.
Photocrosslinkable polyanhydrides are new materials
that present certain advantages in orthopaedic applications [76]. They absorb via surface erosion and therefore
are not susceptible to sudden losses in mass or load
dumping in delivery applications. The photopolymerizable element adds the potential for microfabrication of
porous sca!olds but also could allow an injectable material that can be subsequently crosslinked. Initial mechanical studies show that these polymers demonstrate
enhanced mechanical integrity [76].

3.1.3. Acellular systems: composites


Polymer composites with ceramic "llers have also been
investigated. Peter and coworkers reported a method in
which poly(lactide-co-glycolide) constructs formed by
solvent casting/particulate leaching were crushed and
then compression molded with hydroxyapatite (HA) in
order to improve compressive yield strength [77]. Mikos
reports a poly(propylene fumarate) (PPF) biodegradable
bone cement that can be combined with a leachable
component and injected into osseous defects [78,79]. The
polymerization of the material itself can be adjusted to
cause foaming through the release of carbon dioxide,
thereby producing a porous sca!old [80]. The injectable
nature of the material allows it to "ll irregular osseous
defects, and the leachable component allows room for
bone ingrowth (Fig. 2). This particular material also has
potential as a drug delivery system. Composite systems
can be formed with tricalcium phosphate (TCP) to improve mechanical integrity [81]. Similarly, Bennett
showed that a poly-dioxanone-co-glycolide based composite reinforced with HA or TCP can be used as an
injectable or moldable putty [82]. During the crosslinking reaction following injection, carbon dioxide is released allowing the formation of interconnected pores.
The carbon dioxide also causes expansion of the material
and essentially a press-"t, seamless interface.
Zhang and Ma prepared PLLA/apatite composites by
soaking porous PLLA constructs in simulated body #uid
to allow development of apatite throughout the sca!old
[83]. Apatite formation was enhanced by hydrolysis and
nucleation, and both size and frequency was correlated to
amount of exposed surface area. This may be of direct
interest as a bone tissue-engineering sca!old or indirectly
in assessing currently existing sca!olds. Zhang and Ma
have also examined PLLA/HA constructs formed
through a standard polymer processing technique, thermally induced phase separation (TIPS) [84]. As with
other absorbable systems formed in this manner [85],
the polymer microstructure can be readily controlled
through manipulation of the component concentrations
as well as the processing temperature and cooling rate.

2352

K.J.L. Burg et al. / Biomaterials 21 (2000) 2347}2359

Friedman and coworkers have created a new tetracalcium phosphate that addresses the di$culties with malleability that can be encountered with the high-porosity
ceramics [59]. BoneSource2+ is a hydroxyapatite that is
supplied in a powder. When it is mixed with sterile water,
it is changed into a conformable, paste-like consistency.
When it sets, it does so with a microporous structure of
8}12 lm. Despite its microporous structure, the authors
note that, unlike older ceramic HA implants, BoneSource2+ is rapidly adherent to bone and possesses the
unique quality of direct conversion to new bone without
loss of implant volume. They have termed this process
osteoconversion [59]. In a study of 103 patients with
cranial defects in which BoneSource2+ was used, the
success rate, which was based on maintenance of the
implant and implant volume at 24 months, was approximately 97% [59]. Interpore 200 is a coralline-derived
porous HA trabecular-like structure with average pore
sizes of 200 lm. Hydroxyapatite is a very slowly degraded material and therefore can be fashioned into an
appropriate shape and prefabricated as a vascularized
bone #ap. Levine successfully showed the potential of
this system in rabbits [27].
Ceramic processing is also advancing with the developments of photopolymerizable biopolymers. Garg and
coworkers leveraged stereolithography technology to
fabricate ceramic constructs using a concentrated colloidal dispersion in an aqueous photocurable polymer
solution [87]. This has potential for controlling pore size
and porosity and therefore precision fabrication of porous templates.
Fig. 2. (a) Scanning electron micrograph of PPF sca!old: 10 wt% PPF,
90 wt% porogen; porogen removed with water leaching after PPF
crosslinking. (b) Scanning electron micrograph of PPF sca!old: 30 wt%
PPF, 70 wt% porogen; porogen removed with water leaching after
PPF crosslinking. Micrographs courtesy of Dr. A.G. Mikos (Rice
University Department of Bioengineering).

3.1.4. Acellular systems: ceramics


The absorbable, inorganic materials that have been
investigated include CaCO (argonite), CaSO 2H O

\ 
(plaster of Paris), and Ca (PO ) (beta-whitlockite,


a form of TCP) [47]. The most widely studied calcium
phosphate ceramics are TCP, HA (Ca (PO ) (OH) ),



and the newest tetracalcium phosphate [47,59]. The appeal of the calcium phosphates rests largely with their
biocompatibility. Since they are protein free, minimal
immunologic reactions, foreign body reactions, or systemic toxicity have been reported with their use [47,86].
Although the inorganic ceramics have not shown osteoinductive ability, they certainly possess osteoconductive abilities as well as a remarkable ability to bind
directly to bone [47,86].

3.2. Role of bone morphogenetic protein (BMP) systems


in bone tissue engineering
Near the time that Urist and coworkers "rst generated
demineralized bone matrix, they proposed that its osteoinductive properties were caused by a novel factor
[39]. Several years later, Urist and Strates named this
factor bone morphogenetic protein (BMP) and, in conjunction with several other coworkers, puri"ed rat and
rabbit BMP [67,88,89]. Since then, several groups of
investigators have shown that BMP is actually a group of
proteins responsible for a variety of events in embryogenesis and in the postnatal skeleton [13,17].
BMPs are of particular interest to the "eld of orthopaedics because of the critical role that they have been
shown to play in embryological bone formation, osteoinduction, and bone repair as well as the possibility that
they may assist in the replacement of the standard autogenous bone graft [15,17,25,33,35,90,91]. To date, 15
BMPs have been characterized and cloned [13]. Wang
and Sampath demonstrated that recombinant human
BMP (rhBMP) 2 and 7, respectively, are capable of
inducing bone formation in a process that mimics
endochondral ossi"cation at ectopic sites in a rat model

K.J.L. Burg et al. / Biomaterials 21 (2000) 2347}2359

[90,92]. Using rats, Einhorn demonstrated that the normal pathway of healing for fractures could be accelerated
with the percutaneous injection of rhBMP 2 [35,93].
Bostrom and coworkers were able to demonstrate
intense staining of fracture calluses with the use of
anti-BMP 2 and four antibodies. They were able to
demonstrate an almost biphasic intensity of staining that
related to a primary intensity within the primitive mesenchymal and chondrocytic cells, and a second period of
intensity within osteoblasts as they invaded the cartilaginous callus [15].
Though the use of BMPs may prove to be important
to orthopaedic surgery as a whole, it is evident that
BMPs could be invaluable in orthopaedic reconstructive
surgery. BMP has been delivered to localized sites in
order to repair bone defects and nonunions in several
di!erent experimental models [19}33]. Delivery systems
have included demineralized bone matrix, collagen
composites, "brin, calcium phosphate, polylactide,
poly(lactide-co-glycolide), polylactide-polyethylene glycol,
hydroxyapatite, dental plaster, and titanium [13,34}38,
94]. An ideal delivery system would allow a slow release
of the BMPs, be biologically and immunologically inert,
quickly absorbed, and supportive of cell proliferation
and angiogenesis. It would also possess enough
rigidity to withstand deforming forces until absorbed.
Lastly, it would be easily stored, handled, and sterilized
[13,14].
3.2.1. BMP systems: naturally derived polymers
In long bones such as the tibia, the existing treatment
options for large segmental defects are usually limited to
multistaged reconstruction and/or amputation [95]. Several authors have demonstrated that BMP 2 and 7,
delivered to clinical and experimental osseous defects up
to 17 cm in length, have signi"cantly and favorably
a!ected the ability of these defects to heal [13,18,24,
25,33,35,96}99]. Yasko and coworkers created 5-mm defects in the femora of 45 adult rats [33]. By showing
a 100% union rate using a combination of rhBMP 2
and DBM as a carrier, they concluded that BMP might
prove to be a bone graft substitute of unlimited quantity
[33].
Reddi and Levine both cite insoluble collagen as a
potential carrier for BMP, but point out that data are
limited for this matrix [27,56]. This may be, in part, due
to low compression strengths of constructs of this type.
In an isolated study comparing delivery from collagen
matrices versus HA, TCP, glass beads, and polymethylmethacrylate, the collagen was superior as a drug-releasing matrix [56].
In a series of six patients who had undergone several
lower extremity procedures to gain soft tissue control
following an acute traumatic event, Johnson and
coworkers achieved union in "ve of six patients with
tibial defects that spanned 3}17 cm [25]. The authors

2353

used autologous bone graft supplemented with hBMP


and induced union with a single operative procedure. In
another series, these same authors gathered 25 patients
with resistant nonunions that included partial or complete segmental bone defects [99]. Twenty-three of these
patients had an average of three prior surgical procedures that failed to promote union of the bony ends.
Within 3}7 months of treatment with hBMP and a DBM
derivative as an onlay or inlay graft, 20 of the 25 patients
achieved union. Four of the remaining "ve patients
obtained union after a second procedure [99].
Although BMPs may represent an integral component
in the future of orthopaedic reconstructive surgery, problems potentially exist with the types of carriers currently
in use. The crude extract of BMP is hydrophobic, but as
the protein is puri"ed, it becomes more hydrophilic. The
implantation of puri"ed, hydrophilic BMP promotes
their dispersion shortly after implantation and before
osteoinduction can occur [100}102]. To prevent the dispersion and to deliver the BMPs to the desired sites, they
are normally complexed to carriers that allow a slow
release. Most of the protein-derived carriers such as
DBM and the collagen composites have the potential of
antigenicity and exposure of the recipient to viruses and
other infectious agents [3,13,17,25,68,96,97,99].
3.2.2. BMP systems: synthetic polymers
Although select synthetic polymers such as PLLA can
also pose design issues due to in#ammatory responses or
extended absorption times [21,43,44,51,94,100], investigators have examined the possibility of using synthetic
polymers as vehicles to allow the targeted delivery of
BMPs [21,47,94,100,103]. Ferguson and coworkers
showed that experimentally created trephine defects
could be successfully treated with puri"ed bovine BMP
[21]. They complexed the bBMP with a poly-lactide-coglycolide (PLGA) in one animal. This yielded unabsorbed polymer surrounded by new bone as late as 16 weeks
into the experiment. The authors concluded that longlasting copolymers might provide a barrier to complete
bone formation. Moreover, they postulated that more
quickly degrading polymers could be used to aid in the
controlled delivery of BMP to defects without interfering
with bone formation [21]. Kirker-Head and coworkers
used rhBMP 2 combined with poly-D,L-lactide-co-glycolide (PDLGA) particles of 150}500 lm and autogenous
blood to heal 2.5-cm defects in the femora of "ve of 10
sheep. In the sheep that demonstrated union, the polymer
was completely absorbed and woven and lamellar bone
bridged the defect [103]. Miyamoto evaluated several
polylactic acids (PLAs) and polylactides of varying molecular weights for their potential use as a carrier [100].
These polymers were synthesized using polycondensation and ring-opening techniques, respectively. The differing polymerizations, and therefore molecular weights,
allowed a variety of appearances of the "nal PLA/BMP

2354

K.J.L. Burg et al. / Biomaterials 21 (2000) 2347}2359

or PLLA/BMP complex that ranged from viscous liquid


to pellets. It also allowed for di!ering lengths of time
necessary for complete degradation. The BMP used in
the experiment was puri"ed from rat osteosarcomas.
Among PLAs or PLLAs of di!ering molecular weights,
they determined that only PLA with a molecular weight
of 650 daltons allowed BMP to reproducibly induce bone
formation at ectopic sites (i.e., rat muscle belly). Furthermore, by 2 weeks, the polymer was completely degraded
[100]. In a follow-up study, the authors demonstrated
that when they complexed the 650-dalton PLA/BMP
composite with a 200-dalton polyethylene glycol (PEG),
they induced bone formation and the formation of
haematopoietic marrow 3 weeks after implantation of
the complex into rat muscle bellies [94]. The authors
also noted twice as much bone formation in the rats
that had been given the PLA/PEG/BMP composite.
With the addition of hydroxyapatite, this composite
was converted from a viscous liquid to a doughy
paste. The authors postulate that the viscous composite
could be used as an injectable osteoinductive material
while the doughy composite could be used as a matrix
[94].
Not only does the molecular weight a!ect the BMP
delivery, so does the shape of the biomaterial delivering
BMP [56]. Reddi compared BMP loaded beads and
discs of HA and observed that the beads were inactive,
highlighting the need to customize the structure to suit
the speci"c application.
3.2.3. BMP systems: ceramics
Koempel and coworkers suggest that the addition of
BMP to a porous HA, of 200}500 lm pore size, would
augment the ingrowth of host bone [26]. Tsuruga and
coworkers [38] made similar observations by systematically comparing a series of "ve BMP-releasing HAs, each
with a di!erent pore range. They found that 300}400 lm
yielded the highest bone formation. Takahashi and
coworkers demonstrated 100% fusion in 14 goats that
underwent anterior multilevel cervical spine arthrodesis
when they used synthetic porous HA and rhBMP 2 [32].
Despite the lack of postoperative immobilization and the
shortcomings of porous HA (the brittleness, the slow
degradation, the lack of osteoinduction, and the decreased mechanical strength), the authors still demonstrated a 100% solid fusion rate with high-dose BMP
and HA [32]. Asahina and coworkers demonstrated superior bone induction in comparison to controls when an
HA/collagen/bovine BMP composite was placed into
surgically created mandibular defects in a primate model
[19]. Gao and coworkers showed that a sheep
BMP/collagen composite demonstrated osteoconductive
and osteoinductive properties when applied to an HA
tubular construct with a pore size of 200}400 lm [22]. In
their model of tibial defects in sheep, they reconstructed
the defects with the BMP composite, and the new bone

was characterized by more intense callus formation and


increased mechanical strength as compared with their
control groups [22].

4. Cellular systems
4.1. Cellular systems: naturally derived polymers
The collagen materials have also been applied as cellular sca!olding systems. As collagen possesses no inherent
structural mechanical properties, engineering modi"cations may be useful to provide a sti!er polymer to assist
in force transmission of bone during the regenerative
phase of healing. Yaylaoglu and coworkers demonstrated that porous collagen foams could be treated with
calcium solution to allow the deposition of calcium phosphate and improvement of mechanical integrity [104].
This technique shows promise in chondrocyte culture
and has great potential for bone application as well. It is
not known whether the collagen-based system has longterm stability in culture. Du and coworkers have demonstrated that collagen sheets can be used as the basis for
composite bone tissue-engineering sca!olds [105]. Du
obtained commercially available collagen sheets, precipitated HA onto the surface, then placed bone fragments
along the surface, rolling the composite into a tube. The
pore sizes in this material range from tens to hundreds of
microns; the material is absorbable and #exible. Cells
migrated from the bone fragments into the matrix, suggesting that the material is bioactive.
4.2. Cellular systems: synthetic polymers
Polyglycolide (PG) "brous, nonwoven mesh is another
tissue-engineering candidate. Clinically well known and
having the advantage of fast absorption, this material has
been applied to almost every area of tissue-engineering
research, including bone [106]. Polyglycolide mesh demonstrates relatively low mechanical integrity in vitro
[107] and, for this reason, would be inappropriate as
a bone tissue-engineering construct. By combining this
material with a second, reinforcing material, a stable
construct can be formed. This has been accomplished in
the past using a PL solution to bond the mesh [108].
More recently a speci"c bone application has been accomplished by coating a PG-based tube with PL [106].
Puelacher and coworkers applied PG mesh seeded with
osteoprogenitor cells to the hollow portion of the stabilized tubes. These constructs showed promise as long
bone defect replacements in a rat femoral defect model.
Since joints are the con#uence of several tissue types,
including bone, researchers have recently undertaken the
complicated task of reconstructing a whole joint. This
requires coculture, with each polymer section custom
designed to the speci"c cell type. Preliminary studies used

K.J.L. Burg et al. / Biomaterials 21 (2000) 2347}2359

2355

PL-stabilized PG mesh cocultured with chondrocytes,


tenocytes, and periosteal osteoblasts in an attempt to
form a "nger joint [109]. The separate parts were sewn
together with absorbable suture, and implanted subcutaneously in nude mice. Although the eventual goal is
complex, the basic concepts can be immediately applied
to osteochondral defect repair.
New synthetic polymers are also of great interest as
potential cellular sca!olds. Attawia and coworkers have
focused on poly-anhydride-co-imides as bone tissueengineering sca!old alternatives [110]. These materials
absorb by surface erosion, thus having the advantage of
a more predictable mass loss. Furthermore, they have
high mechanical strength and rigidity, their compressive
moduli ranging from 10 to 60 MPa.

premarket noti"cations (510(k)) and premarket approvals. Premarket noti"cations are from the manufacturer to the FDA that state the intent of marketing
a device for the "rst time or reintroduce a device to the
market that has been signi"cantly adapted. Premarket
approval is the most strict type of device application,
requesting to take to market or continue marketing
a Class III medical device. Included within the Class III
FDA classi"cation are devices that `present a potential,
unreasonable risk of illness or injurya and devices that
`are of substantial importance in preventing impairment
of human healtha.

4.3. Cellular systems: ceramics

Much work remains to be undertaken in the analysis


of progenitor cells and their di!erentiation, particularly
with regard to their interactions with biomaterials. Biomaterial development and "nal design will be essential to
the appropriate stimulation and di!erentiation of bone
cells. The environment in which these polymer}tissue
systems are cultivated will greatly a!ect the long-term
tissue viability. Furthermore, it will also be necessary to
focus on creating the optimal micromechanical environment. Bioreactor studies examining the interaction of
cellular concentration with material type and the in#uence of residence time on bone development will also be
of interest.
Clinically, it is always of interest to improve methods
of bone regeneration in order to reduce costs and the
surgical trauma to the patient. The ability to use autogenous bone forming cells attached to a mechanically
sound, biologically active BMP impregnated, replaceable
sca!olding would be ideal. This would allow implantation of the cellular construct without associated surgeries
to harvest a graft or implant a structural device. As
a compromise, a similar cellular construct could be used
in conjunction with conventional "xation techniques.
Both techniques would rely on a viable bed of host tissue,
so that there is no local tissue necrosis or dense avascular
scar tissue.
The ideal replacement material for osseous defects
associated with orthopaedic reconstructive surgery
would be one that is initially pliable and readily molded
but which hardens quickly once implanted. The photopolymerizing polymers may "nd application in open
surgeries; they could also potentially be used in arthroscopic applications or minimally invasive procedures.
Polymeric foams may be good candidates as delivery
vehicles for cells or drugs such as antibiotics, since they
would conform to the defect.
Bone tissue engineering is one area that will bene"t
greatly from the current e!orts by the American Society
for Testing and Materials to construct tissue-engineering
guidelines. It is very di$cult to correlate studies from

Solchaga and coworkers compared two commercially


available HA-based constructs with a well-characterized
porous calcium phosphate ceramic [111]. Speci"cally,
Hya! 11 (Fidia Advanced Biopolymers; Abano Terme,
Italy) constructs (100}400 lm size pores, porosity of
80%) were compared with ACP (Fidia Advanced Biopolymers; Abano Terme, Italy) constructs (10}300 lm
size pores, 85% porosity). Porous calcium phosphate
ceramic with 60% HA, 40% TCP (200}400 lm pore
sizes, porosity of 60%) were also examined in this study.
All materials were seeded with marrow progenitor cells
before subcutaneous implantation in rats. The ACP constructs and TCP ceramics bound identical numbers of
cells, signi"cantly lower than those bound to the Hya!
11. The ACP constructs absorbed relatively quickly without a protective cell layer, disappearing in the rat within
3 weeks in subcutaneous location. The structure of Hya!
11 is much more open with larger pores, allowing good
distribution of cells throughout. The ACP constructs
tripled in volume when hydrated, perhaps diminishing
the available cell space within the pores. This demonstrated the advantage of high porosities in achieving
good cellular distribution and the importance of correct
pore size selection [111].

5. Clinically available tissue-engineering options and


speci5cations
According to the food and drug administration (FDA),
at this time there are no approved orthopaedic devices
that incorporate tissue-derived components such as cells
and/or growth factors. Additionally, there is currently no
conclusive data to support indications for these devices.
There are, however, "llers without tissue-derived components that may be classi"ed as tissue-engineering materials. They are listed in Table 2 along with manufacturer
name and FDA "ling classi"cation. The table lists both

6. Future areas of development

2356

K.J.L. Burg et al. / Biomaterials 21 (2000) 2347}2359

Table 2
Current FDA-listed bone void "llers
Device type

Product name

Product description

Regulation number

Applicant

Calcium sulfate preformed


pellets

K990131, cleared 3/2/99

Interpore Cross International

Stimulan-calcium sulfate
bone void "ller

Calcium sulfate preformed


pellets

K982663, cleared 2/26/99

Encore Orthopaedics
International

Pro Osteon implant 500R


resorbable bone graft
substitute

Calcium sulfate preformed


pellets

K980817, cleared 9/25/98

Interpore International

Profusion bone void "ller

Calcium sulfate preformed


pellets

K973704, cleared 4/3/98

Biogeneration

Wright plaster of paris pellets Calcium sulfate preformed


pellets

K963562, cleared 5/7/97

Wright Medical Technology, Inc.

Wright plaster of paris


bone void "ller kit

Calcium sulfate preformed


pellets

K963587, cleared 3/24/97

Wright Medical Technology, Inc.

Wright plaster of paris


pellets (subject to revision)

Calcium sulfate preformed


pellets

K960978, cleared 6/21/96

Wright Medical Technology, Inc.

Methyl methacrylate for


cranioplasty

K983009, cleared 11/25/98

Etex Corporation

MEDPOR surgical granule


implants

Methyl methacrylate for


cranioplasty

K982040, cleared 9/8/98

Porex Surgical, Inc.

Norian cranial repair


system (CRS) bone cement

Methyl methacrylate for


cranioplasty

K973789, cleared 5/18/98

Norian Corporation

Bonesource hydroxyapatite
cement (HAC)

Methyl methacrylate for


cranioplasty

K964537, cleared 1/24/97

Osteogenics, Inc.

Bonesource hydroxyapatite
cement (HAC)

Methyl methacrylate for


cranioplasty

K953339, cleared 6/27/96

Osteogenics, Inc.

Cranioplastic, acrylic
cranioplasty material

Methyl methacrylate for


cranioplasty

K873689, cleared 10/26/87

Dentsply International

Collagraft

Collagen matrix for bone


repair

P900039

Collagen Corporation

Pro Osteon 500

Porous hydroxyapatite
bone graft substitute
blocks and granules

P860005

Interpore Cross
International

501(k) Devices, calcium


Pro Osteon implant 500R
sulfate bone void "llers resorbable bone graft
or equivalent
substitute

510(k) Devices, CranioBSM * bone substitute


facial Bone Void Fillers material

Premarket Approval
Devices, Bone Fillers
or equivalent

di!erent laboratories when each has an independent set


of processing parameters or uses a speci"c polymer of
often unreported physical parameters. Standardization
will hopefully expedite the fabrication of successful
tissue-engineered alternatives.

Acknowledgements
The authors wish to acknowledge the input of N.Y.
Sloan and A. Torres-Cabassa from the United States
Food and Drug Administration. The authors would also
like to thank L. Jenkins of the Clemson University Department of Bioengineering as well as A.G. Mikos and
J.P. Fisher of the Rice University Department of Bioengineering for their contributions.

References
[1] Brown KLB, Cruess RL. Bone and cartilage transplantation
surgery. J Bone Jt Surg Am 1982;64-A:270}9.
[2] Enneking WF, Eady JL, Burchardt H. Autogenous cortical bone
grafts in the reconstruction of segmental skeletal defects. J Bone
Jt Surg Am 1980;62-A:1039}58.
[3] Tiedeman JJ, Garvin KL, Kile TA, Connolly JF. The role
of a composite, demineralized bone matrix and bone marrow
in the treatment of osseous defects. Orthopaedics 1995;
18:1153}8.
[4] Gepstein R, Weiss RE, Hallel T. Bridging large defects in bone
by demineralized bone matrix in the form of a powder. A radiographic, histological, and radioisotope-uptake study in rats.
J Bone Jt Surg Am 1987;69:984}92.
[5] Summers BN, Eisenstein SM. Donor site pain from the ilium.
J Bone Jt Surg Br 1989;71-B:677}80.
[6] Coventry MB, Tapper EM. Pelvic instability. J Bone Jt Surg Am
1972;54-A:83}101.

K.J.L. Burg et al. / Biomaterials 21 (2000) 2347}2359


[7] Younger EM, Chapman MW. Morbidity at bone graft donor
site. J Orthop Trauma 1989;3:192}5.
[8] Bostrom RD, Mikos AG. Tissue engineering of bone. In: Atala
A, Mooney D, Vacanti JP, Langer R, editors. Synthetic biodegradable polymer sca!olds. Boston: BirkhaK user, 1997. p. 215}34.
[9] Ilizarov GA. The tension-stress e!ect on the genesis and growth
of tissues: part I. Clin Orthop Rel Res 1989;238:249}90.
[10] Ilizarov GA. The tension-stress e!ect on the genesis and growth
of tissues: part II. Clin Orthop Rel Res 1989;239:263}85.
[11] Koller MR, Palsson MA, Manchel I, Maher RJ, Palsson B".
Tissue culture surface characteristics in#uence the expansion of
human bone marrow cells. Biomaterials 1998;19:1963}72.
[12] Dekker RJ, De Bruijn JD, Van den Brink I, Bovell YP, Layrolle P,
Van Blitterswijk CA. Bone tissue engineering on calcium phosphate-coated titanium plates utilizing cultured rat bone marrow
cells: a preliminary study. J Mater Sci Mater Med 1998;9:859}63.
[13] Croteau S, Rauch F, Silvestri A, Hamdy RC. Bone morphogenetic proteins in orthopaedics: from basic science to clinical practice. Orthopaedics 1999;22:686}95.
[14] Harakas NK. Demineralized bone-matrix-induced osteogenesis.
Clin Orthop Rel Res 1984:239}51.
[15] Bostrom MPG, Lane JM, Berberian WS, Missri AAE, Tomin E,
Weiland A. Immunolocalization and expression of bone morphogenetic proteins 2 and 4 in fracture healing. J Orthop Res
1995;13:357}67.
[16] Reddi AH. Regulation of cartilage and bone di!erentiation by
bone morphogenetic proteins. Curr Opin Cell Biol 1992;4:850}5.
[17] Ripamonti U, Duneas N. Tissue morphogenesis and regeneration by bone morphogenetic proteins. Plast Reconstr Surg
1998;101:227}39.
[18] Wozney JM, Rosen V, Byrne M, Celeste AJ, Moutsatsos I, Wang
EA. Growth factors in#uencing bone development. J Cell Sci
(Suppl) 1990;13:149}56.
[19] Asahina I, Watanabe M, Sakurai N, Mori M, Enomoto S.
Repair of bone defect in primate mandible using a bone morphogenetic protein (BMP)}hydroxyapatite}collagen composite.
J Med Dent Sci 1997;44:63}70.
[20] Cunningham BW, Kanayama M, Parker LM, Weis JC, Sefter
JC, Fedder IL. Osteogenic protein versus autologous interbody
arthrodesis in the sheep thoracic spine. A comparative endoscopic study using the Bagby and Kuslich interbody fusion
device. Spine 1999;24:509}18.
[21] Ferguson D, Davis WL, Urist MR, Hurt WC, Allen EP. Bovine
bone morphogenetic protein (BMP) fraction-induced repair of
craniotomy defects in the rhesus monkey (Macaca speciosa).
Clin Orthop Rel Res 1987;219:251}8.
[22] Gao TJ, Lindholm TS, Kommonen B, Ragni P, Paronzini A,
Lindholm TC. Enhanced healing of segmental tibial defects in
sheep by a composite bone substitute composed of tricalcium
phosphate cylinder, bone morphogenetic protein, and type IV
collagen. J Biomed Mater Res 1996;32:505}12.
[23] Gao T, Lindholm TS, Marttinen A, Urist MR. Composites of
bone morphogenetic protein (BMP) and type IV collagen, coralderived coral hydroxyapatite, and tricalcium phosphate ceramics. Int Orthop 1996;20:321}5.
[24] Gerhart TN, Kirker-Head CA, Kriz MJ, Holtrop ME, Hennig
GE, Hipp J. Healing segmental femoral defects in sheep using.
recombinant human bone morphogenetic protein. Clin Orthop
Rel Res 1993;293:317}26.
[25] Johnson EE, Urist MR, Finerman GA. Repair of segmental
defects of the tibia with cancellous bone grafts augmented with
human bone morphogenetic protein. A preliminary report. Clin
Orthop Relat Res 1988:249}57.
[26] Koempel JA, Patt BS, O'Grady K, Wozney J, Toriumi DM. The
e!ect of recombinant human bone morphogenetic protein-2 on
the integration of porous hydroxyapatite implants with bone.
J Biomed Mater Res 1998;41:359}63.

2357

[27] Levine JP, Bradley J, Turk AE, Ricci JL, Benedict JJ, Steiner G.
Bone morphogenetic protein promotes vascularization and osteoinduction in preformed hydroxyapatite in the rabbit. Ann
Plastic Surg 1997;39:158}68.
[28] Linde A, Hedner E. Recombinant bone morphogenetic protein-2
enhances bone healing, guided by osteopromotive e-PTFE
membranes: an experimental study in rats. Calcif Tissue Int
1995;56:549}53.
[29] Miyamoto S, Takaoka K, Ono K. Bone induction in monkeys
by bone morphogenetic protein. A trans-"lter technique. J Bone
Jt Surg Br 1993;75:107}10.
[30] Sandhu HS, Kanim LE, Toth JM, Kabo JM, Liu D, Delamarter
RB. Experimental spinal fusion with recombinant human bone
morphogenetic protein-2 without decortication of osseous
elements (published erratum appears in Spine 1997 Oct
15;22(20):2463). Spine 1997;22:1171}80.
[31] Suzawa M, Takeuchi Y, Fukumoto S, Kato S, Ueno N,
Miyazono K. Extracellular matrix-associated bone morphogenetic proteins are essential for di!erentiation of murine
osteoblastic cells in vitro. Endocrinology 1999;140:2125}33.
[32] Takahashi T, Tominaga T, Watabe N, Yokobori Jr. AT, Sasada
H, Yoshimoto T. Use of porous hydroxyapatite graft containing
recombinant human bone morphogenetic protein-2 for cervical
fusion in a caprine model. J Neurosurg 1999;90:224}30.
[33] Yasko AW, Lane JM, Fellinger EJ, Rosen V, Wozney JM, Wang
EA. The healing of segmental bone defects, induced by recombinant human bone morphogenetic protein (rhBMP-2). A radiographic, histological, and biomechanical study in rats (published
erratum appears in J Bone Jt Surg Am 1992;74(7):1111). J Bone
Jt Surg Am 1992;74:659}70.
[34] DeGroot J. Carriers that concentrate native bone morphogenetic protein in vivo. Tissue Eng 1998;4:337}41.
[35] Bostrom MP, Camacho NP. Potential role of bone morphogenetic proteins in fracture healing. Clin Orthop Rel Res
1998:S274}82.
[36] Whang K, Tsai DC, Nam EK, Aitken M, Sprague SM, Patel PK.
Ectopic bone formation via rhBMP-2 delivery from porous
bioabsorbable polymer sca!olds. J Biomed Mater Res
1998;42:491}9.
[37] Sakou T. Bone morphogenetic proteins: from basic studies to
clinical approaches. Bone 1998;22:591}603.
[38] Tsuruga E, Takita H, Itoh H, Wakisaka Y, Kuboki Y. Pore size
of porous hydroxyapatite as the cell-substratum controls BMPinduced osteogenesis. J Biochem 1997;121:317}24.
[39] Urist MR, Silverman MF, Buring K, Dubuc FL, Rosenburg JM.
The bone induction principle. Clin Orthop Rel Res
1967;53:243}83.
[40] Urist MR, Strates BS. Bone morphogenetic protein. J Dent Res
1971;50:1392}406.
[41] Urist MR. The search for and discovery of bone morphogenetic
protein. In: Urist MR, O'Connor BT, Burwell RG, editors. Bone
grafts, derivatives and substitutes. London: Butterworth, 1994.
p. 315}62.
[42] Kulkarni RK, Moore EG, Hegyeli AF, Leonard F. Biodegradable poly(lactic acid) polymers. J Biomed Mater Res
1971;5:169}81.
[43] Pineda LM, Busing M, Meinig RP, Gogolewski S. Bone regeneration with resorbable polymeric membranes. III. E!ect of
poly(L-lactide) membrane pore size on the bone healing process
in large defects. J Biomed Mater Res 1996;31:385}94.
[44] Gugala Z, Gogolewski S. Regeneration of segmental diaphyseal
defects in sheep tibiae using resorbable polymeric membranes:
a preliminary study. J Orthop Trauma 1999;13:187}95.
[45] Meinig RP, Rahn B, Perren SM, Gogolewski S. Bone regeneration with resorbable polymeric membranes: treatment of diaphyseal bone defects in the rabbit radius with poly(L-lactide)
membrane. A pilot study. J Orthop Trauma 1996;10:178}90.

2358

K.J.L. Burg et al. / Biomaterials 21 (2000) 2347}2359

[46] Hollinger JO, Schmitz JP. Macrophysiologic roles of a delivery


system for vulnerary factors needed for bone regeneration. Ann
New York Acad Sci 1997;831:427}37.
[47] Hollinger JO, Battistone GC. Biodegradable bone repair materials. Clin Orthop Rel Res 1986;207:290}305.
[48] Grizzi I, Garreau H, Li S, Vert M. Hydrolytic degradation of
devices based on poly(DL-lactic acid) size dependence. Biomaterials 1995;16:305}11.
[49] Vert M, Li S, Garreau H. New insights on the degradation of
bioresorbable polymeric devices based on lactic and glycolic
acids. Clin Mater 1992;10:3}8.
[50] Burg KJL, Shalaby SW. Physicochemical changes in degrading polylactide "lms. J Biomater Sci - Polym Ed
1997;9(1):15}29.
[51] Mainil-Varlet P, Rahn B, Gogolewski S. Long-term in vivo
degradation and bone reaction to various polylactides. 1. Oneyear results. Biomaterials 1997;18:257}66.
[52] Boyan BD, Hummert TW, Dean DD, Schwartz Z. Role of
material surfaces in regulating bone and cartilage cell response.
Biomaterials 1996;17(2):137}46.
[53] Mikos AG, Thorsen AJ, Czerwonka LA, Bao Y, Langer R.
Preparation and characterization of poly(L-lactic acid) foams.
Polymer 1994;35:1068}77.
[54] Burg KJL, Mikos AG, Beiler RJ, Culberson CR, Greene KG,
Loebsack AB, Roland WD, Wyatt S, Halberstadt CR, Holder Jr.
WD, Burg TC. Particulate selection and importance to cell
adhesion in solvent-cast, particulate-leached polymeric constructs. Transactions of the 25th Annual Meeting of the Society
of Biomaterials, April 1999, Providence, RI.
[55] Holy CE, Dang SM, Davies JE, Shoichet MS. In vitro degradation of a novel poly(lactide-co-glycolide) 75/25 foam. Biomaterials 1999;20:1177}85.
[56] Reddi AH. Role of morphogenetic proteins in skeletal tissue
engineering and regeneration. Nat Biotechnol 1998;16:247}52.
[57] Whang K, Tsai DC, Nam EK, Aitken M, Sprague SM, Patel PK,
Healy KE. Ectopic bone formation via rhBMP-2 delivery from
porous bioabsorbable polymer sca!olds. J Biomed Mater Res
1998;42:491}9.
[58] Holmes RE. Bone regeneration within a coralline hydroxyapatite implant. Plast Reconstr Surg 1979;63:626}33.
[59] Friedman CD, Costantino PD, Takagi S, Chow LC. BoneSource
hydroxyapatite cement: a novel biomaterial for craniofacial skeletal tissue engineering and reconstruction. J Biomed Mater Res
(Appl Biomater) 1998;43:428}32.
[60] Tancred DC, McCormack BAO, Carr AJ. A synthetic bone
implant macroscopically identical to cancellous bone. Biomater
1998;19:2303}11.
[61] Brekke JH, Toth JM. Principles of tissue engineering applied
to programmable osteogenesis. J Biomed Mater Res
1998;43:380}98.
[62] Peter SJ, Miller MJ, Yasko AW, Yaszemski MJ, Mikos AG.
Polymer concepts in tissue engineering. J Biomed Mater Res
1998;43:422}7.
[63] Van de Putte KA, Urist MR. Experimental mineralization of
collagen sponge and decalci"ed bone. Clin Orthop Rel Res
1965;40:48}56.
[64] Van de Putte KA, Urist MR. Osteogenesis in the interior of
intramuscular implants of decalci"ed bone matrix. Clin Orthop
Rel Res 1966;43:257}70.
[65] Urist MR, Wallace TH, Adams T. The function of "brocartilaginous fracture callus. JBJS 1965;47B:304}18.
[66] Urist MR, Silverman BF, Buring K, Dubuc FL, Rosenberg JM.
The bone induction principle. Clin Orthop Rel Res
1967;53:243}83.
[67] Urist MR, Mikulski A, Lietze A. Solubilized and insolubilized
bone morphogenetic protein. Proc Nat Acad Sci USA
1979;76:1828}32.

[68] Urist MR, Dawson E. Intertransverse process fusion with the aid
of chemosterilized autolyzed antigen-extracted allogeneic (AAA)
bone. Clin Orthop Rel Res 1980;154:97}113.
[69] Russell JL, Block JE. Clinical utility of demineralized bone
matrix for osseous defects, arthrodesis, and reconstruction: impact of processing techniques and study methodology. Orthopaedics 1999;22:524}31.
[70] Gogolewski S, Mainil-Varlet P. E!ect of thermal treatment on
sterility, molecular and mechanical properties of various polylactides. 2. Poly(L/D-lactide) and poly(L/DL-lactide). Biomater
1997;18:251}5.
[71] Elissee! J, Anseth K, Sims D, McIntosh W, Randolph M, Langer R. Transdermal photopolymerization for minimally invasive
implantation. Proc Nat Acad Sci USA 1999;96:3104}7.
[72] Meinig RP, Buesing CM, Helm J, Gogolewski S. Regeneration
of diaphyseal bone defects using resorbable poly(L/DL-lactide)
and poly(D-lactide) membranes in the Yucatan pig model. J Orthop Trauma 1997;11:551}8.
[73] Dunnen WFA, Schakenraad JM, Zondervan GJ, Pennings AJ,
van der Lei B, Robinson PH. A new PLLA/PCL copolymer for
nerve regeneration. J Mater Sci Mater Med 1993;4:521}5.
[74] Nakamura T, Shimizu Y, Matsui T, Okumura N, Hyon SH,
Nishiya K. A novel bioabsorbable mono"lament surgical suture
made from (e-caprolactone, L-lactide) copolymer. In: Planck H,
Dauner M, Renardy M, editors. Degradation phenomena on
polymeric biomaterials. Springer: Berlin, 1992. p. 153}62.
[75] Ekholm M, Hietanen J, Lindqvist C, Rautavuori J, Santavirta S,
Suuronen R. Histological study of tissue reactions to e-caprolactone-lactide copolymer in paste form. Biomater 1999;20:1257}62.
[76] Muggli DC, Lee HR, Keyser SA, Anseth KS. Photocrosslinkable
polyanhydride networks for use in orthopedic applications. In:
Peppas NA, Mooney DJ, Mikow AG, Brannon-Peppas L, editors. Biomaterials, carriers for drug delivery, and sca!olds for
tissue engineering. New York: AIChE, 1997. p. 275}7.
[77] Peter SJ, Miller MJ, Yasko AW, Yaszemski MJ, Mikos AG.
Polymer concepts in tissue engineering. J Biomed Mater Res
1998;43:422}7.
[78] Crane GM, Ishaug SL, Mikos AG. Bone tissue engineering. Nat
Med 1995;1:1322}4.
[79] Peter SJ, Suggs LJ, Yaszemski MJ, Engel PS, Mikos AG. Synthesis of poly(propylene fumarate) by acylation of propylene
glycol in the presence of a proton scavenger. J Biomater Sci
Polym Ed 1999;10:363}73.
[80] Lewandrowski K-U, Cattaneo MV, Gresser JD, Wise DL,
White RL, Bonassar L, Trantolo DJ. E!ect of a poly(propylene
fumarate) foaming cement on the healing of bone defects. Tissue
Eng 1999;5:305}16.
[81] Peter SJ, Kim P, Yasko AW, Yaszemski MJ, Mikos AG. Crosslinking characteristics of an injectable poly(propylene fumarate)/b-tricalcium phosphate paste and mechanical properties of
the crosslinked composite for use as a biodegradable bone cement. J Biomed Mater Res 1999;44:314}21.
[82] Bennett S, Connolly K, Lee DR, Jiang Y, Buck D, Hollinger JO,
Gruskin EA. Initial biocompatibility studies of a novel degradable polymeric bone substitute that hardens in situ. Bone
1996;19:101S}7S.
[83] Zhang R, Ma PX. Porous poly(L-lactic acid)/apatite composites
created by biomimetic process. J Biomed Mater Res
1999;45:285}93.
[84] Zhang R, Ma PX. Poly(a-hydroxyl acids)/hydroxyapatite porous composites for bone-tissue engineering. I. Preparation and
morphology. J Biomed Mater Res 1999;44:446}55.
[85] Roweton SL. A new approach to the formation of tailored
microcellular foams and microtextured surfaces of absorbable
and non-absorbable thermoplastic biomaterials. Master of
Science Thesis, Department of Bioengineering, Clemson University, 1994.

K.J.L. Burg et al. / Biomaterials 21 (2000) 2347}2359


[86] Hammerle CH, Olah AJ, Schmid J, Fluckiger L, Gogolewski S,
Winkler JR. The biological e!ect of natural bone mineral on
bone neoformation on the rabbit skull. Clin Oral Implants Res
1997;8:198}207.
[87] Garg R, Prud'homme RK, Aksay IA. 3-D stereolithography of
ceramic materials as orthopaedic implants. In: Peppas NA,
Mooney DJ, Mikow AG, Brannon-Peppas L, editors. Biomaterials, carriers for drug delivery, and sca!olds for tissue
engineering. New York: AIChE, 1997. p. 281}3.
[88] Urist MR, Mikulski AJ. A soluble bone morphogenetic protein
extracted from bone matrix with a mixed aqueous and
nonaqueous solvent. Proc Soc Exp Biol Med 1979;162:48}53.
[89] Urist MR, Strates BS. Bone morphogenetic protein. J Dent Res
1971;50:1392}406.
[90] Sampath TK, Maliakal JC, Hauschka PV, Jones WK, Sasak H,
Tucker RF. Recombinant human osteogenic protein-1 (hOP-1)
induces new bone formation in vivo with a speci"c activity
comparable with natural bovine osteogenic protein and stimulates osteoblast proliferation and di!erentiation in vitro. J Biol
Chem 1992;267:20352}62.
[91] Takagi K, Urist MR. The role of bone marrow in bone morphogenetic protein-induced repair of femoral massive diaphyseal
defects. Clin Orthop Rel Res 1982:224}31.
[92] Wang EA, Rosen V, D'Alessandro JS, Bauduy M, Cordes P,
Harada T. Recombinant human bone morphogenetic protein
induces bone formation. Proc Nat Acad Sci 1990;87:2220}4.
[93] Wozney JM, Rosen V. Bone morphogenetic protein and bone
morphogenetic protein gene family in bone formation and repair. Clin Orthop Rel Res 1998;346:26}37.
[94] Miyamoto S, Takaoka K, Okada T, Yoshikawa H, Hashimoto J,
Suzuki S. Polylactic acid-polyethylene glycol block copolymer.
A new biodegradable synthetic carrier for bone morphogenetic
protein. Clin Orthop Relat Res 1993:333}43.
[95] Maurer RC, Dillin L. Multistaged surgical management of posttraumatic segmental tibial bone loss. Clin Orthop Rel Res
1987;216:162}70.
[96] Johnson EE, Urist MR. One-stage lengthening of femoral
nonunion augmented with human bone morphogenetic protein.
Clin Orthop Rel Res 1998:105}16.
[97] Johnson EE, Urist MR, Finerman GA. Distal metaphyseal tibial
nonunion. Deformity and bone loss treated by open reduction,
internal "xation, and human bone morphogenetic protein
(hBMP). Clin Orthop Rel Res 1990:234}40.
[98] Stevenson S, Cunningham N, Toth J, Davy D, Reddi AH. The
e!ect of osteogenin (a bone morphogenetic protein) on the
formation of bone in orthotopic segmental defects in rats. J Bone
Jt Surg Am 1994;76:1676}87.

2359

[99] Johnson EE, Urist MR, Finerman GA. Resistant nonunions and
partial or complete segmental defects of long bones. Treatment
with implants of a composite of human bone morphogenetic
protein (BMP) and autolyzed, antigen-extracted, allogeneic
(AAA) bone. Clin Orthop Rel Res 1992:229}37.
[100] Miyamoto S, Takaoka K, Okada T, Yoshikawa H, Hashimoto J,
Suzuki S. Evaluation of polylactic acid homopolymers as carriers for bone morphogenetic protein. Clin Orthop Rel Res
1992:274}85.
[101] Takaoka K, Nakahara H, Yoshikawa H, Masuhara K, Tsuda T,
Ono K. Ectopic bone induction on and in porous hydroxyapatite combined with collagen and bone morphogenetic protein.
Clin Orthop Rel Res 1988:250}4.
[102] Nakahara H, Takaoka K, Koezuka M, Sugamoto K, Tsuda T,
Ono K. Periosteal bone formation elicited by partially puri"ed bone morphogenetic protein. Clin Orthop Rel Res
1989:299}305.
[103] Kirker-Head CA, Gerhart TN, Armstrong R, Schelling SH,
Carmel LA. Healing bone using recombinant human bone morphogenetic protein 2 and copolymer. Clin Orthop Rel Res
1998;349:205}17.
[104] Yaylaoglu MB, Yildiz C, Korkusuz F, Hasirci V. A novel osteochondral implant. Biomaterials 1999;20:1513}20.
[105] Du C, Cui FZ, Zhu XD, de Groot K. Three-dimensional nanoHAp/collagen matrix loading with osteogenic cells in organ
culture. J Biomed Mater Res 1999;44:407}15.
[106] Puelacher WC, Vacanti JP, Ferraro NF, Schloo B, Vacanti CA.
Femoral shaft reconstruction using tissue-engineered growth of
bone. Int J Oral Maxillofac Surg 1996;25:223}8.
[107] Holder Jr. WD, Gruber HE, Roland WD, Moore AL, Culberson
CR, Loebsack AB, Burg KJL, Mooney DJ. Increased vascularization and heterogeneity of vascular structures occurring in polyglycolide matrices containing aortic endothelial cells implanted
in the rat. Tissue Eng 1997;3:149}60.
[108] Mooney DJ, Mazzoni CL, Breuer C, McNamara K, Hern D,
Vacanti JP, Langer R. Stabilized polyglycolic acid "bre-based
tubes for tissue engineering. Biomater 1996;17:115}24.
[109] Isogai N, Landis W, Kim TH, Gerstenfeld LC, Upton J, Vacanti
JP. Formation of phalanges and small joints by tissue-engineering. J Bone Jt Surg 1999;81-A:306}16.
[110] Attawia MA, Herbert KM, Uhrich KE, Langer R, Laurencin
CT. Proliferation, morphology, and protein expression by osteoblasts cultured on poly(anhydride-co-imides). J Biomed Mater Res (Appl Biomater) 1999;48:322}7.
[111] Solchaga LA, Dennis JE, Goldberg VM, Caplan AI. Hyaluronic
acid-based polymers as cell carriers for tissue-engineered repair
of bone and cartilage. J Orthop Res 1999;17:205}13.

You might also like