Sei sulla pagina 1di 13

Polymer International Polym Int 56:145–157 (2007)

Review
Biodegradable polymers applied in tissue
engineering research: a review
Monique Martina1 and Dietmar W Hutmacher2∗
1 Department of Orthopedic Surgery, Yong Loo Lin School of Medicine, National University of Singapore, 10 Kent Ridge Crescent,
Singapore 119260
2 Division of Bioengineering, Faculty of Engineering, Department of Orthopaedic Surgery, Yong Loo Lin School of Medicine,

National University of Singapore, 10 Kent Ridge Crescent, Singapore 119260

Abstract: Typical applications and research areas of polymeric biomaterials include tissue replacement, tissue
augmentation, tissue support, and drug delivery. In many cases the body needs only the temporary presence of
a device/biomaterial, in which instance biodegradable and certain partially biodegradable polymeric materials
are better alternatives than biostable ones. Recent treatment concepts based on scaffold-based tissue engineering
principles differ from standard tissue replacement and drug therapies as the engineered tissue aims not only to
repair but also regenerate the target tissue. Cells have been cultured outside the body for many years; however, it
has only recently become possible for scientists and engineers to grow complex three-dimensional tissue grafts to
meet clinical needs. New generations of scaffolds based on synthetic and natural polymers are being developed and
evaluated at a rapid pace, aimed at mimicking the structural characteristics of natural extracellular matrix. This
review focuses on scaffolds made of more recently developed synthetic polymers for tissue engineering applications.
Currently, the design and fabrication of biodegradable synthetic scaffolds is driven by four material categories:
(i) common clinically established polymers, including polyglycolide, polylactides, polycaprolactone; (ii) novel
di- and tri-block polymers; (iii) newly synthesized or studied polymeric biomaterials, such as polyorthoester,
polyanhydrides, polyhydroxyalkanoate, polypyrroles, poly(ether ester amide)s, elastic shape-memory polymers;
and (iv) biomimetic materials, supramolecular polymers formed by self-assembly, and matrices presenting
distinctive or a variety of biochemical cues. This paper aims to review the latest developments from a scaffold
material perspective, mainly pertaining to categories (ii) and (iii) listed above.
 2006 Society of Chemical Industry

Keywords: scaffolds; biodegradable polymers; tissue engineering; matrices

INTRODUCTION degradation and mechanical property customization;


A great number of current tissue engineering strate- (iii) polymers that are regulatory approved for spe-
gies are based on the development of a cell–scaffold cific applications and/or are in clinical trials, e.g.
construct whose role is to repair and regenerate tissue polyorthoester (POE), polyanhydrides, polyhydrox-
defects (Fig. 1). During the first phase of tissue engi- yalkanoate (PHA), and newly synthesized polymeric
neering in the 1990s research utilized either US Food biomaterials, such as polypyrroles (PPy), poly(ether
and Drug Administration (FDA)/CE mark approved ester amide)s (PEEA), and elastic shape-memory poly-
devices or used so-called conventional scaffold fabrica- mers; and (iv) biomimetic materials, supramolecular
tion technologies in combination with FDA/CE mark polymers formed by self-assembly, and matrices pre-
approved biomaterials of synthetic and natural origin. senting distinctive or a variety of biochemical cues.
This work has been reviewed in detail elsewhere.1 – 3 The aim of this review is to capture the latest develop-
Currently, the design and fabrication of syn- ments from a scaffold material point of view covering
thetic scaffolds is driven by four material categories: the aforementioned categories (ii) and (iii).
(i) biodegradable and bioresorbable polymers, which
have been effectively used for clinically established
products, including polyglycolide (PGA), polylactides DI- AND TRI-BLOCK POLYMERS BASED ON
(PLA), poly-L-lactic acid (PLLA), poly-D,L-lactic ALIPHATIC POLYESTERS
acid (PDLA), polycaprolactone (PCL); (ii) novel di- In attempts to tune aliphatic polymer properties
and tri-block polymers which predominantly incor- to wider applications for scaffold-based tissue engi-
porate PGA, PLA, and other resorbable polymers neering, di- and tri-block polymers are seen as
in different chain arrangements which confer both promising alternatives. By modifying the backbone


Correspondence to: Dietmar W Hutmacher, Division of Bioengineering, Faculty of Engineering, Department of Orthopaedic Surgery, Yong Loo Lin School of
Medicine, National University of Singapore, 10 Kent Ridge Crescent, Singapore 119260
E-mail: biedwh@nus.edu.sg
(Received 3 October 2005; revised version received 28 February 2006; accepted 21 March 2006)
Published online 13 October 2006; DOI: 10.1002/pi.2108

 2006 Society of Chemical Industry. Polym Int 0959–8103/2006/$30.00


M Martina, DW Hutmacher

Figure 1. Schematic of scaffold-based tissue engineering.

of the polymers, some of the characteristics, such to its elasticity (elongation up to 250% and recovery
as degradation properties, mechanical properties, and up to 98% after applied strain of 120%) Lee et al.5
even biocompatibility, of the polymers can be changed predicted the potential of this di-block in muscle tis-
to a significant extent.4 – 6 sue engineering. Preliminary in vitro studies using rat
Li et al.4 developed a series of di- and tri-block smooth muscle cells (SMC) displayed growth and tis-
polymers based on PCL, PLA, and poly(ethylene gly- sue formation on the PCL/PGA scaffold. A summary
col) (PEG)/poly(ethylene oxide) (PEO). They used a of the properties of this system is given in Table 1.
tri-block made of PLA/PEO/PLA, with PEO as the PCL-based copolymers for soft tissue engineering
hydrophilic and swollen component; and PLA chains have been investigated by Cohn’s group,7 who
as the nanometric nodes in the gel network. These synthesized tri-blocks based on PCL/PEO/PCL (soft
physically crosslinked hydrogels possessed interesting segment with hexamethylene diisocyanate chain
degradation properties. Incorporation of PLA blocks extension as the hard segment). Incorporation of PEO
at the end of the PEO segments decreased the degra- enhanced water permeability. These tri-blocks have
dation rate when compared to pure PEO. Weight a water uptake of up to 94%, which also influences
losses of 49% after 10 days and 77% after 30 days the mechanical properties and biodegradability. The
were observed. A fast initial loss was observed due to length of the blocks affect polymer properties: for
the release of PEO-rich segments. Enzymatic degra- example, longer PEO segments lead to a decrease
dation promoted a much faster degradation rate which in the degree of crystallinity, an increase in water
reached 59% weight loss after 80 h of immersion, com- uptake, and an increase in degradation rate. Possible
pared to 5% under hydrolytic degradation conditions. applications include soft tissue replacement and drug
The author of this review and co-workers prepared a delivery systems. The cell biocompatibility of this
series of scaffolds from these materials using a rapid system was studied by the authors of this review
prototyping system (Fig. 2) and the scaffolds were (Fig. 3(a)–(c)).
studied in vitro and in vivo.6 One of the current challenges facing polymer
PCL/PGA di-block systems5 have certain elastic usage in cell therapy is the ability to add bioac-
properties. Scaffolds with pore sizes of 250 ± 50 µm tive molecules to enhance cytocompatibility. Con-
and a porosity of 93% were produced using a sol- ventional biodegradable polymers lack this charac-
vent casting and particulate leaching method. Due teristic. A strategy to overcome this limitation is

146 Polym Int 56:145–157 (2007)


DOI: 10.1002/pi
DOI: 10.1002/pi
Table 1. Summary of polymer properties

Mechanical Degradation Biocompatibility Biocompatibility


Polymer Abbreviation properties rate Application in vivo in vitro Reference

Polym Int 56:145–157 (2007)


Poly(glycolide-co- PGCL >250% elongation at 50% mass loss in Smooth muscle tissue Subcutaneous Rat SMC 5
caprolactone) break 6 weeks (PBS, engineering implantation using
37 ◦ C) mouse model,
appropriate SM
tissue formation
after 3 weeks
Poly(ester urethane)urea PEUU 0.97–1.64 MPa tensile 13.37–20.7% mass Soft tissue engineering Rat SMC 23
strength (3D porous loss in 8 weeks (in
scaffold) PBS, 37 ◦ C)
Poly(ether ester PEEUU 0.59–1.68 MPa tensile 25.4–47.3% mass Soft tissue engineering Rat SMC 23
urethane)urea strength (3D porous loss in 8 weeks (in
scaffold) PBS, 37 ◦ C)
Poly(ester urethane)urea PEUU 13 ± 4 MPa tensile Soft tissue engineering Rat SMC 24
strength (nanofibre)
Lysine diisocyanate-glucose LDI-glucose 67.7% mass loss in Bone tissue Subcutaneous Rabbit BMSC 22
60 days (PBS, engineering implantation using
37 ◦ C). Approx. rat model, no tissue
90–100% mass loss necrosis,
in 60 days (in vivo) vascularized, no
giant cells
Poly(propylene fumarate) PPF 354.1–1024.2 MPa 13.6–27.9 ± 1.8% Bone tissue Subcutaneous and Rat osteoblast 26–28
compression modulus mass loss in engineering cranial implantation
12 weeks (in PBS, using rabbit model,
37 ◦ C) 5–15 layers of
fibrous capsule, and
inflammatory cells
are seen

147
Biodegradable polymers in tissue engineering
148
M Martina, DW Hutmacher

Table 1. Continued

Mechanical Degradation Biocompatibility Biocompatibility


Polymer Abbreviation properties rate Application in vivo in vitro Reference

Poly(ethyl PPHOS/HA 200 MPa compression Less than 5% mass Bone tissue Rabbit tibial model, Osteoblast cell lines 12
glycinate/p-methyl modulus, and remains loss in 12 weeks (in engineering minimal
phenoxy) phosphazene/ after 2 weeks PBS, 37 ◦ C) inflammatory
hydroxyapatite response
Poly(trimethylene Poly(TMC/CL) 1–10 MPa Young’s Approx. 6% after Cardiovascular Subcutaneous 16,17,19
carbonate/ε- modulus 52 weeks’ implantation using
caprolactone) implantation rat model, sterile
inflammation with
normal foreign body
reaction
Poly(glycerol sebacate) PGS 0.282 ± 0.025 MPa 17 ± 6% mass loss in Microvasculature, soft Subcutaneous Fibroblast cell lines 33,34
tensile Young’s 60 days (in vitro). tissue implantation using
modulus Totally absorbed rat model, similar
(in vivo) inflammatory
response to PLGA
Poly(desamino Poly(DTE carbonate) 57 ± 8 MPa ultimate No mass loss after Anterior cruciate Subcutaneous Fibroblast human cell 10
tyrosyl-tyrosine ethyl ester tensile strength 30 weeks (in PBS, ligament implantation using lines and primary
carbonate) 37 ◦ C) rat model, tissue rabbit fibroblast
ingrowth similar to
collagen. Rabbit
model

DOI: 10.1002/pi
Polym Int 56:145–157 (2007)
Biodegradable polymers in tissue engineering

Figure 2. Polyester–polyether block copolymers composed of PCL or PLA and poly(ethylene glycol) (PEG) have attracted much attention as they
offer the possibility of varying the ratio of hydrophobic/hydrophilic constituents to modulate the degradability and hydrophilicity of the polymer
matrix and surface. Scaffolds were fabricated using a rapid prototyping (RP) machine built in-house at the National University of Singapore. System
details are reported elsewhere.7 Briefly, in contrast to traditional RP systems such as fused deposition modelling, three-dimensional printing,
stereolithography, and selective laser sintering, which mainly focus on a single mode of material processing, this system can accommodate a much
larger variety of synthetic and/or natural biomaterials. A lay-down pattern of 0/90 was used to form the honeycomb patterns of square with a single
fill gap (FG) of 1 mm. Porous sheets measuring 40 × 40 × 3 mm were fabricated using the system, with a 0.5 mm diameter nozzle. PCL–PEG
diblock copolymer was synthesised by Li et al.4 The extrusion temperature was set at 110 ◦ C for PCL–PEG. PCL–PEG was dispensed with an air
pressure of 3.5 bar and a speed of 30 mm min−1 in the x/y axis. Scanning electron micrographs (right) of melt extruded PCL–PEG scaffolds display
typical honeycomb morphology (A). Alternate layers of filaments were positioned at right angles to one another creating pore sizes of 600 µm.
Confocal laser microscopy of scaffold/cell constructs was used to study cell attachment and proliferation.

to use polymers containing functional side groups. of the tyrosine-derived polycarbonates that was not
Guan and co-workers8 attempted to synthesize an only resorbable, but also possessed high strength. The
ABA tri-block polymer that had this particular structure of this material is shown in Fig. 4(a). How-
property. Fabrication from ABA-type tri-block poly- ever, in that study, the degradation rate was shown to
mer PLGBG–PEG–PLGBG consisting of PEG and be very slow, with no mass loss observed after 30 weeks
poly[(lactic acid)-co-(glycolic acid)-alt-(γ -benzyl-L- of incubation in PBS, at 37 ◦ C.
glutamic acid)] that had undergone catalytic hydro- In the first phase, Bourke et al. fabricated this mate-
genation resulted in PLGG–PEG–PLGG, which rial by melt extrusion at 60–90 ◦ C, which was above
contained carboxyl pendant groups. These carboxyl the glass transition temperature. In the second phase,
pendant groups may be useful to facilitate the attach- they tried the melt-spinning technique at 181–183 ◦ C.
ment of oligosaccharides, drug molecules, or short The fabricated fibres were aligned to mimic the
peptides. Due to its amphiphilic nature, the polymer structure of ACL (anterior cruciate ligament). The
formed micelles and may find useful applications in ultimate tensile strength of this material was found
drug delivery systems. to be comparable with natural ACL (57 MPa). Sub-
Another new tri-block system are thermogelling cutaneous implantation using a rat model revealed
copolymers9 made of PEG/PCL/PEG. At a copolymer tissue ingrowth 4 weeks post-operative with mainly
concentration of less than 1% and at a temperature fibroblast-like structures observed surrounding the
of 20 ◦ C, micelle aggregates of 10 and 23 nm coex- fibres. Interestingly, although the structure was intact
isted, but at 25–45 ◦ C micelle aggregates of 23 nm 8 weeks post-implantation, strength retention dropped
were dominant. Increasing the copolymer concentra- to 40%; this was in contrast to in vitro studies which
tion also increased the micelle aggregation size. Upon showed almost 90% strength retention, which implied
varying the PCL block’s length, sol–gel temperature that the in vivo environment induced molecular relax-
decreased due to the hydrophobic interaction that ation, causing reduction in strength.
drove gel formation. This system is appropriate for Further studies based on polycarbonate’s good
in situ gel forming whereby entrapment and depot biocompatibility and ease of biochemical modification
formation can be envisaged with minimally invasive towards cell adhesivity11 have been undertaken.
therapy. Using the polycondensation technique, copolymers of
poly ethylglycol/poly(desamino tyrosyl-tyrosine ethyl
ester carbonate) (PEG/poly(DTE carbonate)) were
obtained to study keratinocyte migration. It was
NEWLY SYNTHESIZED OR STUDIED concluded that PEG provided a better attachment and
POLYMERIC BIOMATERIALS migration substrate for fibroblasts and keratinocytes,
Polycarbonate whereas poly(DTE carbonate) acted to maintain the
Polycarbonate in its pure form is an amorphous structural integrity of the graft.
polymer that possesses low moisture absorption,
and is not susceptible to microbial attack, which Polyphosphazene
implies non-biodegradability of the polymer. How- The search for polymeric materials that are versatile
ever, the mechanical properties of the polymer attract for various hard and soft tissue engineering appli-
researchers keen to develop polycarbonate-based cations has generated interest in another group of
material that can withstand mechanical loading while polyphosphazene-based materials.12 – 14 The unique
tissue is being regenerated. Since resorbability is the properties arise from the unusual flexible backbone
main problem, Bourke et al.10 investigated a member allowing torsional and angular freedom within the

Polym Int 56:145–157 (2007) 149


DOI: 10.1002/pi
M Martina, DW Hutmacher

(a)

(b)

(c)

Figure 3. (a) Soft-PCL scaffold fabricated via a rapid prototyping technique (unpublished data). (i) High flexibility of scaffold is demonstrated; (ii, iii)
swelling ability is demonstrated (soft segments in polymer chain allow scaffold to swell up to 1.5 times). (b) Light micrographs of soft-PCL
(P) alginate/thrombin (M) construct seeded with human adipose derived precursor cells (hADAS). Formulations of cell aggregates and extracellular
material (white arrows) illustrate cell proliferation upon culturing: (i) day 1, (ii) day 7, (iii) day 22, (iv) day 42. Bar represents 100 µm. (c) Environmental
scanning electron micrograph (left) of scaffold/cell construct after 3 weeks of culturing exhibited intact cell/alginate/thrombin structure (M) inside the
soft-PCL pore morphology. Confocal laser micrograph (right) of soft-PCL (P) alginate/thrombin construct after 3 weeks of culturing. Cell viability is
shown by life/death assay (FDA; green fluorescence in confocal laser micrograph) inside the alginate/thrombin matrix. (Cells are indicated by white
arrows).

P–N skeletal system (Fig. 5(a)). The most com- technology enables fabrication of this polymer derived
mon fabrication technique is thermal ring-opening from polydichlorophosphazene using two different
polymerization of dichlorophosphazene. The current substituents. These substituents induce highly tuned

150 Polym Int 56:145–157 (2007)


DOI: 10.1002/pi
Biodegradable polymers in tissue engineering

(a) (b)

(c)

(d)

Figure 4. Structures of (a) poly(desamino tyrosyl-tyrosine ethyl ester carbonate) (poly(DTE carbonate)), modified from of Bourke et al.;10
(b) PPF/PPF-DA crosslinked, modified from of Horch et al.;25 (c) poly DTH-adipate for R = hexyl and y = 4, modified from of Schachter and Kohn;42
(d) soft segment and hard segment of poly(ether ester amide), modified from Deschamps et al.44

(a)

(b)

(c)

Figure 5. Preparation schemes of (a) polydichlorophosphazene, modified from Ambrosio et al.;12 (b) PEUU, modified from Stankus et al.;24
(c) poly(glycerol sebacate), modified from Wang et al.33

Polym Int 56:145–157 (2007) 151


DOI: 10.1002/pi
M Martina, DW Hutmacher

properties such as crystallinity, degradability, and opportunity for this material to be utilized in appli-
hydrophilicity/hydrophobicity which provide the ver- cations such as coatings to enhance tissue integration
satility of the polymers. and wound dressings.
Ambrosio et al.12 explored amino acid ester
polyphosphazene in bone tissue engineering. They Trimethylene carbonate-based materials
chose poly[(ethylglycinate phosphazene)-co-(p-me- The elastomeric properties of poly(trimethylene car-
thylphenoxy phosphazene)] (PPHOS) in combination bonate) (poly(TMC)) make it a potential can-
with hydroxyapatite (HA) to develop a composite. didate for scaffold-based soft tissue engineering
Degradation was achieved by incorporating side chains applications.16 – 19 Interestingly, this polymer exhibits
that sensitized the polymer backbone to hydrolysis; slow degradation in vitro, but rapid degradation
in this case, the amino acid side chains. Phenoxy in vivo.16 Poly(TMC) showed negligible mass loss
side chains inhibited degradation, and in this study, after 2 years in physiological solution, in which degra-
PPHOS/HA remained stable mechanically and main- dation mainly occurred due to hydrolysis of ester
tained a compressive modulus of around 200 MPa, bonds. In contrast, in vivo implantation of polymer
after 12 weeks’ degradation in vitro. PLAGA/HA con- discs into Wistar rats resulted in a high degradation
structs that were used as a control showed almost rate based on a cell-triggered surface erosion mecha-
insignificant compressive modulus. Degradation prod- nism. After 52 weeks, implanted poly(TMC) was 1%
ucts could easily be detoxified by the body as the of its initial mass.
degradation by-products are amino acids, phosphates, Incorporation of other monomers such as D,L-
and ammonia. PPHOS possesses a combination of lactide acid (DLLA) or ε-caprolactone (CL) mod-
surface- and bulk-eroding properties. ulated the degradation properties in vitro as well as
Other types of polyphosphazene, those with in vivo. In vivo degradation characteristics of these
polyethyloxybenzoate and polypropyloxybenzoate, copolymers coincided well with in vitro degradation
were combined with HA resulting in composites behaviour, in which hydrolysis was mainly responsi-
suitable for bone grafts.15 Syntheses of these polyphos- ble for both conditions. TMC/DLLA decreased to
phazene/HA composites were achieved through 1% of the initial mass after 52 weeks, with rapid loss
acid–base reactions to form HA precursors, which observed after week 12, while TMC/CL displayed a
increased the pH of the solution and in the presence linear and continuous reduction throughout a year of
of polyphosphazenes resulted in carboxyl formation in investigation, with mass loss less than 7% of the initial
the surface layer. These carboxyl groups then reacted mass. The difference in degradation behaviour of the
with calcium ions to form calcium crosslinks on the two copolymers was caused by autocatalytic activity of
surface, which played a role in nucleation and deposi- DLLA, which did not occur in TMC/CL.
tion of HA. Furthermore, the in vivo tissue response of these
In another study, a PPHOS/PLAGA blend exhibited copolymers is similar to sterile inflammatory reactions
a higher pH in the degradation solution compared to followed by normal foreign body reactions, which are
PLAGA.13 The buffering phenomenon was attributed commonly seen after implantation of biodegradable
to phosphates which were shown to be present in the polymers.
solution using 31 P NMR. Cell adhesion experiments Due to its flexibility, poly(TMC/DLLA) was tai-
using MC3T3-E1 cells revealed PPHOS/HA to lored towards heart tissue engineering applications.17
possess a similar cell attachment and proliferation Poly(TMC/DLLA), although glassy at room temper-
when compared to tissue culture plastic (TCPS). ature, was rubbery at body temperature which made it
Nair and colleagues14 investigated the effect of sol- difficult to achieve a porous structure. A combination
vent, needle diameter, solution concentration, and of coprecipitation, compression moulding, and salt
applied voltage on the fabrication of polyphosphazene leaching techniques were used to produce a porous
nanofibre scaffolds. Using chloroform as the solvent stable scaffold with a pore size of around 100 µm, a
produced the most uniform fibres. Decreasing the nee- porosity of 85–95%, and a compressive modulus up
dle diameter resulted in decreased fibre diameter, but to 430 kPa. In vitro degradation studies in physiolog-
too fine a needle produced fibres with beads. Increas- ical solution showed that porous scaffolds shrunk by
ing the solution concentration generally resulted in up to 65% after the third week. Preliminary investi-
larger fibre diameters. The applied voltage affected the gation of cell compatibility using rat cardiomyocytes
shape of the nanofibres generated: 27–30 kV resulted showed average cell attachment and proliferation on
in a curly or distorted shape, 33 kV resulted in cylindri- the copolymer surface compared to normal TCPS
cal and rod-like structures, while 36 kV gave even finer plates.
and more distinct rod-like structures. Investigation Poly(TMC/CL) was explored for nerve grafting.16
of the polyphosphazene nanofibres extended to cell Human dermal fibroblasts (HDFs) were subjected
compatibility studies. In vitro cell adhesion of bovine to extracts of the polymer. Results showed that HDFs
coronary artery endothelial cells showed adhesion 24 h subjected to the extraction vehicle maintained high cell
post-seeding. The materila also exhibited cytocom- viability and cell metabolism activity. Schwann cells
patibility to osteoblast cell lines, MC3T3-E1, which were plated onto copolymer discs and demonstrated
adhered and proliferated for 7 days. This presents an good proliferation, with a higher proliferation rate after

152 Polym Int 56:145–157 (2007)


DOI: 10.1002/pi
Biodegradable polymers in tissue engineering

6 days compared with those cells plated onto PLGA. v/v) are harmless to the cell.21 Moreover, in contrast
Implanted tubular poly(TMC/CL) scaffolds were still to PLA and PLGA degradation mechanisms, the
smooth and flexible at 60 days post-implantation, and study showed no significant increase in pH of the
adhered well to the surrounding tissue. In another solution. PEUU degradation products were also
in vivo study, a lesion in the spinal cord of female shown to be non-toxic to endothelial cells. The
Wistar rats exhibited extensive growth of Schwann polymer showed a linear degradation with no signs
cells and extracellular matrix after implantation of of autocatalytic effects when compared to PLA
poly(TMC/CL) to the lesion.20 or PLGA degradation behaviour.23 Scaffolds were
Grijpma et al.19 also explored the possibility of a prepared by thermally induced phase separation and a
photocrosslinkable polymer via UV functionalization subsequent solvent extraction technique. The results
of poly(TMC/CL) or poly(TMC/DLLA) with fumaric were porous structures (porosity >80%) with pore
acid monoethyl ester (FAME). diameters ranging from 12 to 232 µm. The pore shape
was dependent upon the quenching temperature and
Polyurethanes polymer concentration. The tensile strength of these
Polyurethanes are another area of investigation, polymers during degradation decreased by about 12%
especially for soft tissue engineering applications, after 1 week and 31% after 2 weeks from the initial
in contrast to aliphatic linear polyesters that are value of 0.57–1.69 MPa.
better suited to hard tissue engineering due to their Stankus et al.24 were able to machine these
high glass transition temperature and high modulus. polyurethanes via electrospinning to produce scaffolds.
Polyurethanes exhibit a wide range of properties The tensile strength of the nanofibres produced ranged
through variability of the hard segment (diisocyanate), from 2 to 13 MPa and breaking strains from 160 to
the soft segment (polyethers or polyesters), the chain 280%. Incorporation of type I bovine collagen reduced
extenders, and the ratios in which they are reacted. the tensile strength, but also reduced water contact
Previously, polyurethanes had a limited usage due angles. This represents an increase in hydrophilicity
to the toxicity of their degradation product (2,4- and, in turn, enhanced attachment of rat smooth
diaminotoluene). Hence, the challenge presented was muscle cells.
to develop polyurethanes with non-toxic degradation In vitro and in vivo observations demonstrated that
products. these polymers had no harmful effects on cell viability,
Non-toxic polyurethanes with diisocyanate replace- growth, and proliferation. Subcutaneous implantation
ments which could give rise to non-toxic degrada- using rat models revealed that LDI-glucose polymer
tion products have been developed by at least two did not enhance capsule formation, accumulation of
groups.21 – 24 Guan et al.23 synthesized the polymers foreign body giant cells, or tissue necrosis.22
made from PCL and 1,4-diisocyanatobutane (BDI) The versatility of these polyurethanes was demon-
with putrescine as chain extender (poly(etherurethane strated through attempts to enhance cell and tissue
urea), PEUU) (Fig. 5(b)). BDI was used because, compatibility via: addition of soft segments with PEG
upon degradation, it would release putrescine, a to the backbone in order to enhance hydrophilicity;
polyamine that is essential for cell growth and pro- mixing with collagen type I for better cell attachment;
liferation. Zhang et al.22 synthesized polyurethane and addition of ascorbic acid to the polymer mixture
from highly pure lysine diisocyanate (hard segment) to enhance osteoblast lineage progression. All of these
and polymerized it with glucose, which resulted in proved to be feasible. Polyurethane-based materials
major degradation products lysine and glucose (LDI- are consequently considered promising candidates for
glucose). biomaterial-based applications.
Scaffold fabrication routes for these polyurethanes
include thermally induced phase separation, electro- Polyfumarate
spinning, and water foaming. These create differ- Polyfumarate-based materials have been developed
ent porosities, surface-to-volume ratios, and three- mainly for bone tissue engineering.25 – 28 The main
dimensional structures, with concomitant changes in advantages of these materials are their injectable
mechanical properties which are wide-ranging and can and in situ crosslinkable properties.29,30 With
be varied to suit potential applications in the biomed- N-vinylpyrrolidone (N-VP) as crosslinker, Payne
ical field, such as engineering blood vessels and bone et al.28 showed during an in vitro study that fully
(Table 1). crosslinked poly(propylene fumarate) (PPF) had
The degradation mechanisms of the polymers potential as a substrate for supporting rat osteoblasts.
are important and need to be investigated further. Cell proliferation, ALP activity, and osteocalcin and
Non-toxic degradation products are necessary and, calcium production of cells on fully crosslinked PPF
moreover, mechanical properties are also influenced did not exhibit significant differences with those
by degradation mechanisms. LDI-glucose polymer, cells grown on TCPS. Poly(caprolactone fumarate)
for example, is degraded by hydrolysis of urethane (PCLF) and poly(ethylene glycol fumarate) (PEGF)
bonds to liberate lysine, glucose, ethanol, and CO2 . were also investigated as injectable, self-crosslinkable
Ethanol could inhibit cell–cell adhesion, but a study polymers which circumvented the requirement for
reported that concentrations less than 30 mM (0.5% crosslinking agents that may be toxic. The polymers

Polym Int 56:145–157 (2007) 153


DOI: 10.1002/pi
M Martina, DW Hutmacher

were shown to harden and self-crosslink when under properties of the composite. Without proper disper-
physiological conditions, and tissue compatibility sion, the inorganic particles tended to aggregate and
studies using rat models demonstrated no inflamma- create crack propagation sites, which, in turn, made
tory reactions.29,30 compressive fracture and flexural fracture strength
Bone and soft tissue responses have been inves- worse than that of PPF/PPF-DA.
tigated in New Zealand white rabbit models.27 A Flexural testing was done on samples fabricated
PEG block was added in order to make a hydrogel by injecting a nanocomposite mixture into a mould,
(oligo(PEG fumarate), OPF). Tissue response and and crosslinking the mixture using UV radiation.
in vitro degradation behaviour of four groups were The samples that showed enhanced dispersion and
investigated. The groups included longer PEG blocks, covalent bonding with PPF/PPF-DA matrix exhibited
higher crosslinking density, and the addition of cell- the best flexural modulus of more than threefold
adhesive peptides to the hydrogels. Hydrogels with higher (5.4 GPa) compared with blank PPF/PPF-DA,
higher crosslinking were observed to have the least which has a flexural strength of 1.5 GPa. Compressive
mass loss after 12 weeks in vitro. As the degradation strength, however, was not significantly affected by
product was fumaric acid, a drop in pH was expected. these modifications.
A pH drop of not more than 0.5 was reported (which Incorporation of β-tricalcium phosphate (β-TCP)
was considered to be harmless to the cells). In contrast, was also attempted to enhance the mechanical
in vivo results suggested that hydrogels with longer properties of polyfumarate-based polymers.26 The
PEG block resulted in more tissue infiltration, and polymer was produced by radical polymerization using
benzoyl peroxide and dimethyltoluidine as initiator
hence more active degradation. Fibrous capsule for-
and accelerator. N-VP was used as crosslinking
mation with 5–15 cell layers including inflammatory
reagent. Scaffolds were then fabricated by mixing
cells was observed for all groups.
PPF with β-TCP and leachable porogen NaCl.
Degradation rate and degree of swelling were also
Groups with β-TCP concentrations of 0.5 g g−1 of
dependent on the degree of crosslinking per macromer
PPF exhibited bending strengths of up to 16 MPa,
chain. Less crosslinking gave rise to an increase in
compressive strengths of up to 79 MPa, moduli
water uptake and increased swelling. As degradation
in bending of up to 1270 MPa, and moduli in
occurred mainly due to the cleavage of ester bonds compression of up to 1020 MPa. An approximately
within the crosslinked network by hydrolysis, a greater twofold increase in bending and compressive strength
number of chains were accessible to water and so and bending and compression modulus for a twofold
the polymer degraded faster. The higher the water increase in β-TCP concentration was observed.
uptake, the faster the degradation. One type of OPF Mechanical testing showed compression and bending
investigated even appeared soft and jelly-like after modulus of elasticity to be of the same order of
12 weeks under physiological conditions. magnitude as that of trabecular bone.
As the applications of polyfumarate-based materials
were first investigated for bone substitutes, mechanical Polyorthoester
properties are an ultimate concern. Cortical bone has Polyorthoester is a hydrophobic polymer fabricated by
a compressive modulus in the range 17–20 GPa, com- polycondensation of diketene acetals and diols. This
pressive strength of 106–144 MPa, flexural modulus fabrication creates ortho-ester bonds that are stable at
of 15.5 GPa, and flexural strength of about 180 MPa. neutral pH, but hydrolyse rapidly at phagosomal pH,
This presents a major challenge for the tissue engineer i.e. pH 5.5. Today, it is mainly used in drug delivery
using polyfumarate-based materials. This unique char- systems.31,32
acteristic of bone is a result of its composite make-up
comprising the interaction of inorganic material, i.e. Poly(glycerol sebacate)
HA, with organic material such as collagen fibres. The search for soft and mechanically stable elastomeric
Researchers have considered ways to optimize the materials to be implanted in dynamic environments
mechanical properties of poyfumarate-based polymers led to the investigation of polymers that are anal-
using the principle of bone architecture.25,26 ogous with vulcanized rubber, having a crosslinked
Horch et al.25 developed PPF/poly(propylene three-dimensional network in combination with ran-
fumarate diacrylate) (PPF/PPF-DA) (Fig. 4(b)) with dom coil characteristics. Polycondensation of glycerol
a surface modification using carboxylate alumoxane and sebacic acid renders a polymer having hydro-
nanoparticles. The polymer composites were gener- gen bonding interactions through hydroxyl proline
ated by mixing in a chloroform solvent, and the solvent hydroxyl groups.33,34 With building blocks made of
was then removed by rotary evaporation and high- glycerol (a basic building block of lipids) and sebacic
vacuum drying. Crosslinking was achieved by UV acid (a natural metabolic intermediate in ω-oxidation
radiation. Interactions of the inorganic and organic of medium- to long-chain fatty acids), such materi-
matrix were achieved by covalent bonding, and the als are expected to be biocompatible and non-toxic
presence of organophilic chains in the organic matrix (Fig. 5(c)).
enhanced dispersion. Covalent bonding alone was not The hydrophilic characteristics of the material are
enough; dispersion played a key role in the mechanical a result of the hydroxyl groups attached to its

154 Polym Int 56:145–157 (2007)


DOI: 10.1002/pi
Biodegradable polymers in tissue engineering

backbone. The material is insoluble but swells in Another development by Lendlein and co-workers38
water by approximately 2%. It is totally amorphous as regards elastic shape-memory polymers was light-
at 37 ◦ C like vulcanized rubber, a thermoset polymer. induced shape-memory polymers. With a similar
However, the uncrosslinked polymer can be melted to mechanism to that of a heat-sensitive shape-memory
a liquid form which is soluble in common organic polymer, this too consisted of two segments: molecular
solvents. It is tougher than hydrogels with tensile switches that were photoresponsive, and netpoints for
strength of less than 0.5 MPa and tensile strain covalent crosslinking that determined the permanent
more than 300%.33 In vivo and in vitro investigations shape, using cinnamic acid as a molecular switch.
showed an acceptable biocompatibility. In vitro studies When the polymer was stretched, the coiled segments
using NIH 3T3 fibroblast cell lines grown in PGS- of amorphous polymer chain were elongated. Upon
coated Petri dishes exhibited higher cell growth exposure to UV radiation of wavelength >260 nm,
compared with PLGA-coated Petri dishes. An in vivo the elongated segments were partially fixed due to the
study with Sprague-Dawley rats showed less fibrous formation of new photoresponsive crosslinks. If the
capsule formation compared to PLGA control after loading was removed and the polymer exposed to UV
subcutaneous implanation. radiation of wavelength <260 nm, the crosslinks were
While simple foams were fabricated using a salt cleaved and the polymer returned to its original shape.
leaching technique, lithography was used to fabricate The discovery of this material eliminated temperature
capillary networks.34 The surface of the wafer-like constraints associated with thermally induced shape-
scaffold was coated by a pentapeptide derived from memory polymers for medical applications.
fibronectin to improve cell attachment. Adherence
of the HUVEC to the PGS was observed and
proliferation occurred for at least 10 days. After
Polypyrrole
14 days, the surface of the network was nearly
Electronic interaction is one of the factors that may
confluent and cultures were kept for up to 4 weeks.
influence neuronal tissue regeneration and growth.
PGS has another potential application as nerve
guides.35 In vitro studies demonstrated PGS as A class of polymers that have also been explored in
non-toxic to Schwann cells and supported better the field of nerve tissue engineering are conducting
proliferation compared to PLLA. In vivo studies polymers. Polypyrrole (PPy) is an electrodeposited
revealed less inflammatory response possibly due to polymer that can be doped to modify its physical,
its degradation profile via surface erosion. chemical, and electrical properties, and has conse-
quently emerged as a potential candidate for scaffolds
in neuronal tissue regeneration.39 Using PPy doped
Elastic shape-memory polymers with polystyrene-sulfonate (PSS) and sodium dode-
Lendlein and Langer36,37 investigated polymers that cylbenzene sulfonate (NaDBS) at different deposition
could change shape by increases in temperature. This temperatures and solvents, George et al.39 investi-
thermally induced shape-memory effect required two gated the influence of these parameters on neural
components: a switching segment, having transition tissue growth in vivo. Immunofluorescence analysis of
temperature to fix temporary shape, and a segment for the neural tissue displayed a more complete bridging
crosslinking to determine the permanent shape. The when compared to a Teflon implant, which is cur-
switching segment used was oligo(ε-caprolactone)diol rently a gold standard for neural implants. The data
or dimethacrylates, while crosslinking segments were presented implied that PPy implants generally exhib-
either n-butylacrylate or oligo(p-dioxanone)diol. ited greater tissue integration and less inflammation.
The mechanism of changing the shape from The feasibility of incorporation of neuronal growth
permanent to temporary relies on the transition factors (NGFs) was also shown, with even more neu-
temperature of the material. A polymer in a permanent ral tissue formation observed for NGF-incorporated
shape is heated above Ttrans while applying an external implants.
stress. The temporary shape is obtained by reducing Moreover, the conductivity of PPy can be exploited
the temperature to below Ttrans . Releasing the external for the fabrication of bioelectrical circuits that integrate
stress and heating up the material will reform the electrical and neural signals. Cui et al. investigated
temporary shape back to the permanent shape. the use of PPy as a neural probe.40,41 These neural
Using this mechanism, prior to surgery, implants probes facilitated the functional stimulation and
can be compressed to a smaller and more compact recording from the peripheral or central nervous
temporary shape, inserted by minimally invasive system. The conductivity of PPy might induce selective
surgery, and then using body heat, it will expand back neurons to attach to the electrode and achieve
to the permanent shape. Another possible application neuronal tissue regeneration. However, problems were
is as sutures in endoscopic surgery, whereby the suture encountered, including loss of the ability to record
knot can be applied loosely in its temporary shape, neural activity with time. Modifications by patterning
followed by an increase in the temperature, which peptide or peptide/protein polymer blends on the
would tighten the knot as it goes back to its permanent surface were explored to enhance interaction and
shape. anchorage between electrode and neurons.

Polym Int 56:145–157 (2007) 155


DOI: 10.1002/pi
M Martina, DW Hutmacher

Polyarylates A porous scaffold was also fabricated from this mate-


Schachter and Kohn introduced an interesting rial using compression moulding followed by a salt
concept of using a group of polyarylates for drug leaching technique.
delivery vehicles as an alternative to conventional Another amphiphilic drug delivery system uses a
PLA, PGA, or PLGA systems.42 The polyarylates PCL-based polymer construct containing hydrophilic
(Fig. 4(c)) used were tyrosine-derived polymers which PEEA.45 The possibility of using PCL macromers
possessed sites for interaction with peptides. Some of different molecular weights changed the PEEA
possible interactions included hydrogen bonding and properties including crystallinity and the hydrophilic-
hydrophobic interactions. Addition of PLGA system ity/hydrophobicity, and this would affect the drug
as a ‘delayed excipient’ induced a decrease of release rate. In vitro drug release studies using three
internal pH during its degradation, and weakened the different drugs of different natures revealed that PEEA
sensitive interaction between peptide and polymer. successfully enhanced the completion of drug release.
Using Integrillin as the drug model, the glass An acidic drug was released rapidly after 2 h incuba-
transition temperature of the system varied in the tion, while a drug of a basic nature exhibited longer
range 36–40 ◦ C as the amount of loaded drug varied. controlled release.
The delayed time of the Integrillin release increased
as the initial molecular weight of the PLGA increased. Poly(amido amine)
While degradation of PLGA affected the release of the The use of hydrogels as matrices for tissue engineering
peptide, degradation of the polyarylates did not have applications and the search for versatile, easily fab-
such an effect. Preliminary in vitro studies showed the ricated, biodegradable, and biocompatible matrices
polymer to be versatile towards drug release kinetics continue. One novel hydrogel is made of poly(amido
by tuning in the pendant chain and the backbone unit. amine), a polycationic polymer in nature. It con-
tains ter-amino and amido groups regularly arranged
along the polymer chain, obtained by Michael-type
polyaddition of primary or secondary amines to bis-
Poly(ether ester amide) acrylamides. The method of fabrication made it
Another candidate for drug delivery systems arises possible to have side substituents with biomimick-
from the family of poly(ether ester amide)s (PEEAs). ing properties, such as carboxyl, ter-amino, hydroxyl,
They are mainly fabricated by polycondensation of allyl, or bioactive molecules like proteins and peptides.
PEG and diester-diamide to create an amphiphilic In vitro assays performed by Ferruti et al.46 inves-
system.43,44 Diester-diamide acts as a hydrophobic tigated cytotoxicity, compatibility, and proliferation
block for creating reversible physical crosslinks that of the materials and the cells in contact with them.
account for stronger mechanical properties in the Biological evaluation results exhibited non-toxicity,
swollen state. In this case, PEG is the ‘soft’ segment comparable cell proliferation (more than 70%) to nor-
and diester-diamide acts as the ‘hard’ segment mal TCPS plates, and normal cell morphology.
(Fig. 4(d)). Furthermore, degradation products were non-toxic
The structure gives versatility allowing the tailoring and the rate of degradation could be fine tuned
of hydrogels to suit different drug release profiles. depending on the structure and degree of crosslinking,
Bezemer et al.43 fabricated the polymer by a two-step which, in turn, affected the degradation via a hydrolysis
mechanism. The first step involved transesterification mechanism. For example, hybrids of PAA and BSA
of diamide-dimethyl ester monomers with PEG, and (bovine serum albumin) degraded fully up to 8 months
the second step was polycondensation at 220 ◦ C. The in simulated body fluids.46 It was suspected that this
resulting polymers had an intrinsic viscosity ranging was because of the protective characteristics of BSA.
from 0.58 to 0.78 dL g−1 , and they were not completely Apart from providing a protective layer, BSA made the
amorphous; both depended on the length of the hybrid substrate more supportive to cell adhesion and
polymer blocks. It was shown that the soft segment proliferation. An attempt to make it more bioactive was
length in a microsphere influenced the degree of achieved by incorporating agmatine, a decarboxylated
swelling, in vitro degradation, and release rate.43 product of arginine, derived from RGD sequences.47
Cytocompatibility was investigated using Cell adhesion and proliferation assays exhibited up to
HUVEC.44 These surfaces did not perform as pos- 80% capacity compared to normal polystyrene culture
itively as TCPS, possibly due to the PEO con- plates.
tent. Deschamps et al.44 implanted a low-content-
PEG PEEA subcutaneously in rats. Mass loss was
7–12 wt% at 14 weeks post-implantation and a slow CONCLUSIONS
decrease in intrinsic viscosity led to bulk hydrolysis Biodegradable synthetic polymers, from a material
degradation. Tissue responses were found to be sim- standpoint, are a key area of interest for the devel-
ilar to tissue reactions observed upon implantation of opment of new scaffold-based tissue engineering
other biodegradable polymers. Fibrous capsules were strategies. A critical issue in scaffold-based tissue
observed, with macrophages and blood vessel infiltra- engineering is the assembly of cells and extracel-
tion accompanied by cracks on the polymer surface. lular material into a three-dimensional architecture

156 Polym Int 56:145–157 (2007)


DOI: 10.1002/pi
Biodegradable polymers in tissue engineering

that allows for both structure and functionality that 10 Bourke SL, Kohn J and Dunn MG, Tissue Eng 10:43
mimics the native tissue that is being replaced (2003).
11 Sharma RI, Kohn J and Moghe PV, J Biomed Mater Res A
and/or repaired. As the scaffolds for tissue engi- 69:114 (2004).
neering will be implanted in the human body, the 12 Ambrosio AM, Sahota JS, Runge C, Kurtz SM, Lakshmi S,
scaffold materials should be non-antigenic, non- Allcock HR, et al, IEEE Eng Med Biol Sept/Oct: 18 (2003).
carcinogenic, non-toxic, non-teratogenic, and possess 13 Ambrosio AM, Allcock HR, Katti DS and Laurencin CT,
Biomaterials 23:1667 (2002).
high cell/tissue biocompatibility so that they will
14 Nair SL, Bhattacharyya S, Bender JD, Greish YE, Brown W,
not trigger pathological reactions after implantation. Allcock HR, et al, Biomacromolecules 5:2215 (2004).
In addition to materials issues, the macro- and 15 Greish YE, Bender JD, Lakshmi S, Brown PW, Allcock HR and
microstructural properties of the scaffold are also very Laurencin CT, Biomaterials 26:1 (2005).
important. In general, the scaffolds require individual 16 Pego AP, Van Luyn MJ, Brouwer LA, van Wachem PB,
Poot AA, Grijpme DW, et al, J Biomed Mater Res A 67:1044
external shape and well-defined internal structure with
(2003).
interconnected porosity to host most cell types. From 17 Pego AP, Siebum B, Van Luyn MJ, Gallego y Van Seijen XJ,
a biological point of view the designed matrix should Poot AA, Grijpma DW, et al, Tissue Eng 9:981 (2003).
serve various functions, including (1) as an immobi- 18 Fabre T, Schappacher M, Bareille R, Dupuy B, Soum A,
lization site for transplanted cells, (2) formation of a Bertrand-Barat J, et al, Biomaterials 22:2951 (2001).
19 Grijpma DW, Hou Q and Feijen J, Biomaterials 26:2795 (2005).
protective space to prevent unwanted tissue growth 20 Lietz M, Ullrich A, Schulte-Eversum C, Oberhoffner S, Fricke
into the wound bed and allow healing with dif- C, Muller HW, et al, Biotechnol Bioeng 93:99 (2006).
ferentiated tissue, (3) directing migration or growth 21 Zhang JY, Doll BA, Beckman EJ and Hollinger JO, Tissue Eng
of cells via surface properties of the scaffold, and 9:1143 (2003).
(4) directing migration or growth of cells via release of 22 Zhang JY, Beckman EJ, Hu J, Yang GG, Agawal S and
Hollinger JO, Tissue Eng 8:771 (2002).
soluble molecules such as growth factors, hormones, 23 Guan J, Fujimoto KL, Sacks MS and Wagner WR, Biomaterials
and/or cytokines. Some technology platforms of FDA- 26:3961 (2005).
approved polymeric degradable scaffold systems have 24 Stankus JJ, Guan J and Wagner WR, J Biomed Mater Res A
already found promising clinical applications. New 70:603 (2004).
25 Horch RA, Shahid N, Mistry AS, Timmer MD, Mikos AG and
polymers are under development and their applica-
Barron AR, Biomacromolecules 5:1990 (2004).
tions need to be a part of the systemic approach 26 Porter BD, Oldham JB, He SL, Zobits ME, Payne RG, An KN,
to tissue engineering; namely the need to coordi- et al, J Biomech Eng 122:286 (2000).
nate interaction between the parameters of scaffolds, 27 Shin H, Quinten Ruhe P, Mikos AG and Jansen JA, Biomaterials
cells, bioreactors, and biomolecular factors as well as 24:3201 (2003).
28 Payne RG, McGonigle JS, Yaszemski MJ, Yasko AW and
the controlling features of the host response to re-
Mikos AG, Biomaterials 23:4381 (2002).
implanted constructs, including phenomena of angio- 29 Jabbari E, Wang S, Lu L, Gruetzmacher JA, Ameenudin S,
genesis, inflammation, and immune response. Based Hefferan TE, et al, Biomacromolecules 6:2503 (2005).
on these results, novel therapeutic options in the area 30 Wang S, Lu L, Gruetzmacher JA, Currier BL and Yazsem-
of polymeric scaffold-based tissue replacement can be ski MJ, Biomaterials 27:832 (2006).
31 Deng JS, Li L, Tian Y, Ginsburg E, Widman M and Myers A,
expected in the 21st century. Pharm Dev Technol 8:31 (2003).
32 Wang C, Ge Q, Ting D, Nguyen D, Shen HR, Chen J, et al,
Nature Mater 3:190 (2004).
ACKNOWLEDGEMENTS 33 Wang Y, Ameer GA, Sheppard BJ and Langer R, Nature Biotech
20:602 (2002).
The work reported in this review was supported in part 34 Fidkowski C, Kaazempur-Mofrad MR, Borenstein J, Vacanti
by the Biomedical Research Council (BMRC) grants JP, Langer R and Wang Y, Tissue Eng 11:302 (2005).
R-397-000-005-305 (to DWH). The authors thank 35 Sundback CA, Shyu JY, Wang Y, Faquin WC, Langer RS,
Maria Woodruff for editing the manuscript. Vacanti JP, et al, Biomaterials 26:5454 (2005).
36 Lendlein A and Langer R, PNAS 98:842 (2000).
37 Lendlein A and Langer R, Science 296:1673 (2002).
38 Lendlein A, Jiang H, Junger O and Langer R, Nature 434:879
REFERENCES (2005).
1 Hutmacher DW, Goh JC and Teoh SH, Ann Acad Med 39 George PM, Lyckman AW, LaVan DA, Hegde A, Leung Y,
Singapore 30:183 (2001). Avasare R, et al, Biomaterials 26:3511 (2005).
2 Liu X and Ma PX, Ann Biomed Eng 32:477 (2004). 40 Cui X, Wiler J, Dzaman M, Altshuler RA and Martin DC,
3 Di Martino A, Sittinger M and Risbud MV, Biomaterials Biomaterials 24:777 (2003).
26:5983 (2005). 41 Cui X, Lee VA, Raphael Y, Wiler JA, Hetke JF, Anderson DJ,
4 Li S, Molina I, Martinez MB and Vert M, J Mater Sci: Mater et al, J Biomed Mater Res 56:261 (2001).
Med 13:81 (2002). 42 Schachter DM and Kohn J, J Control Release 78:143 (2002).
5 Lee SH, Kim BS, Kim SH, Choi SW, Jeong SI, Kwon IK, et al, 43 Bezemer JM, Oude Weme P, Grijpma DW, Dijkstra PJ, van
J Biomed Mater Res A 66:29 (2003). Blitterswijk CA and Feijen J, J Biomed Mater Res 52:8 (2000).
6 Cohn D, Stern T, Gonzales MF and Epstein J, J Biomed Mater 44 Deschamps AA, van Apeldoom AA, de Bruijn JD, Grijpma DW
Res 59:273 (2002). and Feijen J, Biomaterials 24:2643 (2003).
7 Huang MH, Li S, Hutmacher DW, Schantz JT, Vacanti CA, 45 Barbato F, La Rotona MI, Maglio G, Palumbo R and Quaglia
Braud C, et al, J Biomed Mater Res A 69:417 (2004). F, Biomaterials 22:1371 (2001).
8 Guan H, Xie Z, Zhang P, Deng C, Chen X and Jing X, 46 Ferruti P, Bianchi S, Ranucci E, Chiellini F and Caruso V,
Biomacromolecules 6:1954 (2005). Macromol Biosci 5:613 (2005).
9 Hwang JM, Suh JM, Bae YH, Kim SW and Jeong B, Biomacro- 47 Ferruti P, Bianchi S and Ranucci E, Biomacromolecules 6:2229
molecules 6:885 (2005). (2005).

Polym Int 56:145–157 (2007) 157


DOI: 10.1002/pi

Potrebbero piacerti anche