28.07.2018 Views

G. Wei 2004)

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

ARTICLE IN PRESS<br />

Biomaterials 25 (<strong>2004</strong>) 4749–4757<br />

Structure and properties of nano-hydroxyapatite/polymer composite<br />

scaffolds for bone tissue engineering<br />

Guobao <strong>Wei</strong> a , Peter X. Ma a,b,c, *<br />

a Department of Biomedical Engineering, University of Michigan, Ann Arbor, MI 48109-2009, USA<br />

b Department of Biologic and Materials Sciences, University of Michigan, Ann Arbor, MI 48109-1078, USA<br />

c Macromolecular Science and Engineering Center, University of Michigan, Ann Arbor, MI 48109-1055, USA<br />

Received 9 June 2003; accepted 11 November 2003<br />

Abstract<br />

To better mimic the mineral component and the microstructure of natural bone, novel nano-hydroxyapatite (NHAP)/polymer<br />

composite scaffolds with high porosity and well-controlled pore architectures were prepared using thermally induced phase<br />

separation (TIPS) techniques. The morphologies, mechanical properties and protein adsorption capacities of the composite scaffolds<br />

were investigated. The high porosity (90% and above) was easily achieved and the pore size was adjusted by varying phase<br />

separation parameters. The NHAP particles were dispersed in the pore walls of the scaffolds and bound to the polymer very well.<br />

NHAP/polymer scaffolds prepared using pure solvent system had a regular anisotropic but open 3D pore structure similar to plain<br />

polymer scaffolds while micro-hydroxyapatite (MHAP)/polymer scaffolds had a random irregular pore structure. The introduction<br />

of HAP greatly increased the mechanical properties and improved the protein adsorption capacity. In a dioxane/water mixture<br />

solvent system, NHAP-incorporated poly(l-lactic acid) (PLLA) scaffolds developed a fibrous morphology which in turn increased<br />

the protein adsorption three fold over non fibrous scaffolds. The results suggest that the newly developed NHAP/polymer composite<br />

scaffolds may serve as an excellent 3D substrate for cell attachment and migration in bone tissue engineering.<br />

r 2003 Elsevier Ltd. All rights reserved.<br />

Keywords: Nano; Hydroxyapatite; Phase separation; Scaffold; Protein adsorption<br />

1. Introduction<br />

There is a growing need for bone regeneration due to<br />

various clinical bone diseases such as bone infections,<br />

bone tumors and bone loss by trauma [1]. Current<br />

therapies for bone defects include autografts, allografts,<br />

xenografts and other artificial substitutes such as metals,<br />

synthetic cements and bioceramics [2,3]. However, these<br />

substitutes are far from ideal and each has its specific<br />

problems and limitations. For example, autografts are<br />

associated with donor shortage and donor site morbidity<br />

whereas allografts and xenografts have the risk of<br />

disease transmission and immune response [4], and<br />

synthetic materials wear and do not behave like true<br />

bone. As a result, recent research has been devoted to<br />

bone tissue engineering in which a 3D porous scaffold is<br />

loaded with specific living cells and/or tissue-inducing<br />

*Tel.: +1-734-764-2209; fax: +1-734-647-2110.<br />

E-mail address: mapx@umich.edu (P.X. Ma).<br />

factors to launch a tissue regeneration or replacement in<br />

a natural way [5,6].<br />

In the tissue engineering approach, the temporary 3D<br />

scaffold serves an important role in the manipulation of<br />

the functions of osteoblasts and a central role in the<br />

guidance of new bone formation into desired shapes. In<br />

principle, a biodegradable matrix with sufficient mechanical<br />

strength, optimized architecture and suitable<br />

degradation rate, which could finally be replaced by<br />

newly formed bone, is most desirable. The scaffolding<br />

materials for bone tissue engineering should also be<br />

osteoconductive so that osteoprogenitor cells can adhere<br />

and migrate on the scaffolds, differentiate, and finally<br />

form new bone [7,8]. Biodegradable polymers, mainly<br />

polyesters such as poly(lactic acid) (PLA), poly(glycolic<br />

acid) (PGA) and their copolymers (PLGA), have been<br />

widely used to develop porous 3-D scaffolds using<br />

various fabrication techniques [8–14]. These synthetic<br />

materials have been demonstrated to be biocompatible<br />

and degrade into non-toxic components with a<br />

0142-9612/$ - see front matter r 2003 Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.biomaterials.2003.12.005


4750<br />

ARTICLE IN PRESS<br />

G. <strong>Wei</strong>, P.X. Ma / Biomaterials 25 (<strong>2004</strong>) 4749–4757<br />

controllable degradation rate in vivo [15]. Another<br />

major class of biomaterials for bone repair is ceramics<br />

such as hydroxyapatite (HAP) and tricalcium phosphate<br />

(TCP) [16,17]. Being similar to the mineral component<br />

of natural bone, they showed good osteoconductivity<br />

and bone bonding ability [18]. However, the main<br />

limitation for the use of HAP ceramics was their<br />

inherent brittleness and difficulty for processing [19].<br />

To combine the osteoconductivity of calcium phosphates<br />

and good biodegradability of polyesters, polymer/ceramic<br />

composite scaffolds have been developed<br />

for bone tissue engineering either by direct mixing or by<br />

a biomimetic approach [7,8,20]. Compared to plain<br />

polymer scaffolds in which neo tissue matrix was formed<br />

only in the surface layer (o240 mm) [21], the composite<br />

scaffolds supported cells growth and neo tissue formation<br />

throughout the scaffold including in the very center<br />

of the scaffold [6].<br />

Polymer/ceramic composite scaffolds mimic the natural<br />

bone to some extent. Natural bone is composed of<br />

inorganic compound (mainly partially carbonated<br />

HAP on the nanometer scale) and organic compounds<br />

(mainly collagen). The nanometer size of the inorganic<br />

component (mainly bone-like apatite) in natural<br />

bone is considered to be important for the mechanical<br />

properties of the bone [22]. Recent research in this<br />

field also suggested that better osteoconductivity<br />

would be achieved if synthetic HAP could resemble<br />

bone minerals more in composition, size and<br />

morphology [23,24]. In addition, nano-sized HAP may<br />

have other special properties due to its small size and<br />

huge specific surface area. Webster et al. [25,26]<br />

have shown significant increase in protein adsorption<br />

and osteoblast adhesion on the nano-sized ceramic<br />

materials compared to traditional micron-sized ceramic<br />

materials.<br />

In previous works, the Ma group has developed a<br />

variety of scaffolds using thermally induced phase<br />

separation (TIPS) [8–10]. Both pore structure and pore<br />

wall morphology can be controlled by phase separation<br />

parameters. They have also demonstrated that the<br />

addition of micron-sized hydroxyapatite (MHAP) increase<br />

the adsorption of proteins and extracellular<br />

matrix (ECM) components [27]. The cell seeding<br />

uniformity into the scaffold was improved substantially<br />

due to the introduction of osteoconductive hydroxyapatite<br />

[6]. The MHAP/PLLA composite scaffolds were<br />

shown to suppress apoptosis of osteoblasts by a possible<br />

mechanism of enhanced adsorption of serum proteins<br />

such as vitronectin and fibronectin onto the scaffolds<br />

[27, 28]. In this paper, we further mimicked the size scale<br />

of hydroxyapatite in natural bone and aim to fabricate<br />

novel and improved composite scaffolds. The pore<br />

structure, pore wall morphology, mechanical properties<br />

and protein adsorption capacity were systematically<br />

investigated.<br />

2. Materials and methods<br />

2.1. Materials<br />

Poly(l-lactic acid) (PLLA) with an inherent viscosity<br />

of 1.4–1.8 dl/g was purchased from Boehringer Ingelheim<br />

(Ingelheim, Germany) and was used as received.<br />

PLGA85 (Medisorb s , LA:GA=85:15) with an inherent<br />

viscosity of 0.6 was obtained from Alkermes Inc.<br />

(Wilmington, OH). NHAP were purchased from Berkeley<br />

Advanced Biomaterials Inc. (San Leandro, CA).<br />

Conventional HAP particles (MHAP, particle size on<br />

the micrometer scale), tetrahydrofuran (THF), dioxane<br />

and benzene were obtained from aldrich chemical<br />

company (Milwaukee, WI).<br />

2.2. Fabrication method<br />

NHAP/PLLA composite scaffolds were fabricated<br />

using a phase separation technique similar to that<br />

reported previously [8]. Briefly, the HAP nanocrystal<br />

powder was dispersed in a solvent by sonication for 30 s<br />

at 15 W (Virsonic 100, Cardiner, NY). After that, PLLA<br />

was dissolved in the NHAP suspended solvent at about<br />

60 C to make homogeneous solutions with desired<br />

concentrations. 2.5 ml polymer/NHAP mixture was<br />

added into a Teflon vial, sonicated again and then<br />

transferred into a freezer at a preset temperature to<br />

induce solid–liquid or liquid–liquid phase separation.<br />

The solidified mixture was maintained at that temperature<br />

overnight and then transferred into a freeze-drying<br />

container which was maintained at a temperature<br />

between 5 C and 10 C by salt–ice bath for freezedrying.<br />

The samples were freeze-dried at 0.5 mmHg for 7<br />

days to remove solvent. The final composition of the<br />

composite foam was determined by the concentration of<br />

the polymer solution and NHAP content in the mixture.<br />

2.3. SEM characterization<br />

The porous morphologies of the scaffold were<br />

examined with scanning electron microscopy (SEM)<br />

(S-3200N, Hitachi, Japan) at 20 or 15 kV. The samples<br />

were cut by a razor blade or fractured after being frozen<br />

in liquid nitrogen for several minutes and then were<br />

coated with gold for 200 s using a sputter coater (Desk-<br />

II, Denton Vacuum Inc.).<br />

2.4. Porosity<br />

The melting behavior of the matrices was characterized<br />

with a differential scanning calorimeter (DSC-7, Perkin-<br />

Elmer, Norwalk, CT) as detailed previously [9]. The<br />

degree of crystallinity, X c , of a sample was calculated as<br />

X c ¼ DH m =DH 0 m ;


ARTICLE IN PRESS<br />

G. <strong>Wei</strong>, P.X. Ma / Biomaterials 25 (<strong>2004</strong>) 4749–4757 4751<br />

where DH m was the measured enthalpy of melting and<br />

DHm 0 was the enthalpy of melting for 100% crystalline<br />

polymer. For PLLA, DHm 0 ¼ 203:4 J=g: The estimated<br />

density and porosity of the matrix were obtained as<br />

follows. The diameter and height of the matrix were<br />

measured to calculate the volume. The mass of the matrix<br />

was measured with an analytical balance. The density was<br />

calculated from the volume and mass. The porosity, e;<br />

was calculated from the measured overall densities D m of<br />

the matrix and the skeletal density D s . For a composite<br />

scaffold, the skeletal density was determined using the<br />

densities of the polymer and HAP powder. The porosity<br />

was given by<br />

e ¼ D s D m<br />

;<br />

D s<br />

where D s was calculated from the following formula:<br />

1<br />

D s ¼<br />

;<br />

ð1 X h Þ=D p þ X h =D h<br />

where D h is the density of the HAP powder with a value<br />

of 3.16 g/ml and X h is the percentage of HAP in the<br />

composite while D p is the density of the polymer. D p was<br />

calculated from<br />

1<br />

D p ¼<br />

;<br />

ð1 X c Þ=D a þ X c =D c<br />

where X c was the degree of crystallinity of the polymer.<br />

The density of amorphous PLLA (D a ) is 1.248 g/ml and<br />

the density of 100% crystalline PLLA (D c ) is 1.290 g/ml.<br />

2.5. Mechanical properties<br />

The compressive mechanical properties of the scaffolds<br />

were tested using a MTS Synergie 200 mechanical<br />

tester (MTS systems Co. Eden Prairie, Minnesota). The<br />

scaffolds have dimensions of about 16 mm in diameter<br />

and 3 mm in thickness. A crosshead speed of 0.5 mm/<br />

min was used. The compressive modulus was defined as<br />

the initial linear modulus. At least five specimens were<br />

tested for each sample. The averages and standard<br />

deviations were graphed. A two-tail Student’s t-test<br />

(assuming equal variances) was performed to determine<br />

the statistical significance (po0.05) of the differences in<br />

mechanical properties.<br />

2.6. Protein adsorption<br />

Protein adsorption was performed by incubating the<br />

scaffolds in phosphate buffered saline (PBS, 0.1 m,<br />

pH=7.4) containing 2.5% or 5% fetal bovine serum<br />

(FBS). The disk-like specimens with dimensions of 15.5–<br />

16.5 mm in diameter and 370.1 mm in thickness were<br />

used. Before incubation in the medium containing<br />

proteins, the specimens were pretreated by ethanol for<br />

30 min and then washed by PBS three times overnight<br />

under gentle shaking. Specimens were then put into 24-<br />

well culture plates (one specimen for each well) and<br />

1.5 ml FBS/PBS solution was added into each well. The<br />

scaffolds were incubated at 37 C for a chosen incubation<br />

time. The concentration of the protein in the FBS/<br />

PBS solution was then quantified with a commercial<br />

protein assay kit, BCA (Pierce, Rockford, IL), using<br />

bovine serum albumin (BSA) standards. The amount of<br />

absorbed proteins was determined by subtracting the<br />

amount of proteins left in the FBS/PBS solution after<br />

adsorption from the amount of proteins in control FBS/<br />

PBS solution (without specimen) under the same<br />

incubation conditions.<br />

3. Results<br />

3.1. Porosity<br />

NHAP/polymer composite scaffolds with high porosity<br />

have been fabricated using thermally induced phase<br />

separation techniques (Tables 1 and 2). Two solvent<br />

systems were used to obtain composite scaffolds with<br />

different pore structures. One is a mono-solvent system<br />

of pure dioxane or benzene and the other, a solvent/nonsolvent<br />

mixture system of dioxane and water in different<br />

proportions. The densities and porosities of scaffolds are<br />

listed in Tables 1 and 2. Incorporation of nano-sized<br />

HAP only decreased the porosity slightly. This was<br />

especially true when NHAP content was low. Regardless<br />

of detailed morphology and composition, all scaffolds<br />

have a high porosity of at least 89%, which was<br />

considered to be beneficial for cell ingrowth and<br />

survival. The solvent system did not have much<br />

effect on the density and porosity. The quenching<br />

temperature did not affect the porosity significantly<br />

either (p=0.12).<br />

3.2. Morphology<br />

The morphology and microstructure of the scaffolds<br />

were examined using SEM (Fig. 1): micrographs of<br />

PLLA scaffolds (a,b), MHAP/PLLA (c,d) and NHAP/<br />

PLLA (e–h) composite scaffolds fabricated with pure<br />

Table 1<br />

Densities and porosities of NHAP/PLLA composite scaffolds prepared<br />

from dioxane solutions, phase separated at 18 C<br />

Scaffolds<br />

PLLA:HAP<br />

(w/w)<br />

D p<br />

(g/cm 3 )<br />

D s<br />

(g/cm 3 )<br />

Porosity<br />

(%)<br />

PLLA 100:0 1.2577 0.08812 93.0<br />

PLLA–NHAP 90:10 1.3383 0.09708 92.8<br />

PLLA–NHAP 70:30 1.5394 0.1176 92.3<br />

PLLA–NHAP 50:50 1.7993 0.1475 91.8<br />

PLLA–NHAP 30:70 2.1737 0.2304 89.4<br />

PLLA–MHAP 50:50 1.7993 0.1505 91.6


4752<br />

ARTICLE IN PRESS<br />

G. <strong>Wei</strong>, P.X. Ma / Biomaterials 25 (<strong>2004</strong>) 4749–4757<br />

Table 2<br />

Porosities of NHAP/PLLA scaffolds fabricated using dioxane/water mixture solvent system and phase separated at various temperatures<br />

Scaffolds PLLA:HAP (w/w) Dioxane:H 2 O (v/v) Quenching temp. ( C) Porosity (%)<br />

PLLA–NHAP 70:30 100:0 18 92.3<br />

PLLA–NHAP 70:30 95:5 18 91.6<br />

PLLA–NHAP 70:30 90:10 18 93.1<br />

PLLA–NHAP 70:30 87:13 4 94.5<br />

PLLA–NHAP 70:30 87:13 18 93.1<br />

PLLA–NHAP 70:30 87:13 70 92.8<br />

PLLA–NHAP 70:30 87:13 196 (LN 2 ) 92.5<br />

dioxane or benzene. NHAP/PLLA composite scaffolds<br />

maintained a regular internal ladder-like pore structure<br />

similar to plain PLLA scaffolds, a typical morphology<br />

formed by solid–liquid phase separation [8,10]. By<br />

controlling the heat transfer direction, NHAP/PLLA<br />

scaffolds with tubular pore architecture were also<br />

obtained using benzene as a solvent. Nano-sized HAP<br />

particles were distributed within the pore walls of the<br />

scaffolds and no large aggregates appeared in pores. In<br />

MHAP/PLLA scaffolds, the HAP platelets with size<br />

ranging from 10 to 50 mm were found randomly<br />

distributed in the PLLA matrix. Some were embedded<br />

in the pore wall and some others were piled together<br />

between pores or in the pores. The pore size of all these<br />

scaffolds ranged from 50 to 100 mm. NHAP/PLGA85<br />

composite scaffolds prepared with pure dioxane had<br />

leaf-like morphology and well-interconnected pore<br />

structures (Fig. 2).<br />

Very different morphologies appeared in composite<br />

scaffolds fabricated with dioxane and water mixture<br />

solvent system (Fig. 3). The proportion of water in the<br />

solvent systems had a significant effect on the pore size<br />

and overall morphology of the scaffold. Even a small<br />

amount of water (less than 5%) in the solvent system<br />

significantly reduced the pore size from 100 mm to about<br />

10 mm and random pore structure replaced the regular<br />

ladder-like pore structure (Figs. 3a, band c). Another<br />

obvious change was the pore wall surface texture.<br />

Unlike relative solid and smooth pore wall in NHAP/<br />

PLLA scaffold prepared from mono-solvent system,<br />

fibrous and loose network became the characteristics of<br />

the pore structure of the scaffolds prepared from the<br />

mixed solvent system. When the water volume was 13%<br />

in the mixture solvent, the scaffolds consisted of three<br />

ranges of pore sizes. Large macropores were developed<br />

in a range of 200–500 mm while much smaller micropores<br />

(tens of microns) between the pore walls of the<br />

large pores. The pore wall surface was not smooth, and<br />

appeared as loose fibrous structure with several microns<br />

in between. NHAP particles were entrapped in the<br />

fibrous network of the porous structures and had little<br />

effect on the changes in morphology with different water<br />

proportions since plain PLLA scaffolds had the same<br />

morphological changes (data not shown).<br />

3.3. Mechanical properties<br />

The compressive modulus of NHAP/PLLA scaffolds<br />

increased with NHAP content (Fig. 4). The plain PLLA<br />

scaffolds prepared using solid–liquid phase separation<br />

techniques had a compressive modulus of 4.3 MPa. The<br />

addition of NHAP into the scaffold improved the<br />

mechanical properties. The modulus increased significantly<br />

when the NHAP proportion reached 30% of the<br />

composite and reached 8.3 MPa when the ratio of<br />

NHAP to PLLA was 50:50. The solvent system had a<br />

large effect on the mechanical properties of the porous<br />

scaffolds (Fig. 5). The introduction of a non-solvent<br />

component (in this case, water) in the polymer/HAP/<br />

solvent mixture likely induced liquid–liquid phase<br />

separation instead of solid–liquid phase separation.<br />

Since regular and orientated pore architectures were<br />

replaced by random pore structures in scaffolds<br />

fabricated using mixed solvent system, the compressive<br />

modulus decreased significantly. However, the modulus<br />

recovered substantially when dioxane/water (87:13, v/v)<br />

system was used. This might have resulted from the<br />

uniform structure of the scaffolds (Figs. 3 d–f).<br />

3.4. Protein adsorption<br />

NHAP/PLLA composite scaffolds with different<br />

amounts of NHAP were incubated in FBS/PBS solution<br />

to investigate protein adsorption. For all NHAP/PLLA<br />

composite scaffolds, the protein adsorption reached<br />

equilibrium in 25 h and there was no significant increase<br />

of protein adsorption during further incubation<br />

(Fig. 6a). The addition of NHAP increased the protein<br />

adsorption. NHAP/PLLA 50:50 and NHAP/PLLA<br />

70:30 composite scaffolds absorbed 2.4- and 3.2-fold,<br />

respectively, more serum proteins than plain PLLA<br />

scaffold (Fig. 6b). The size of HAP particles also<br />

affected protein adsorption onto the scaffolds containing<br />

a high percentage of HAP. Serum protein adsorption<br />

was significantly greater on NHAP/PLLA scaffolds<br />

than on MHAP/PLLA scaffolds when the HAP/PLLA<br />

was 50:50 or higher (po0.05, Fig. 6c). However, when<br />

the ratio of HAP in the composite was lower than 50%,<br />

the differences in the protein adsorption between NHAP


ARTICLE IN PRESS<br />

G. <strong>Wei</strong>, P.X. Ma / Biomaterials 25 (<strong>2004</strong>) 4749–4757 4753<br />

Fig. 1. SEM micrographs of plain PLLA, NHAP/PLLA and MHAP/PLLA scaffolds fabricated from 5% dioxane (a–e) and benzene (f,g) solution.<br />

(a,b) Pure PLLA scaffold, 50, 400; (c,d) MHAP/PLLA 50:50 scaffold, 100, 500; (e,f) NHAP/PLLA 50:50 scaffold, 100, 1000;<br />

(g) Tubular NHAP/PLLA scaffold, cross-section, 200; (h) Tubular NHAP/PLLA scaffold, longitudinal section, 100.


4754<br />

ARTICLE IN PRESS<br />

G. <strong>Wei</strong>, P.X. Ma / Biomaterials 25 (<strong>2004</strong>) 4749–4757<br />

Fig. 2. SEM micrographs of NHAP/PLGA85 (30:70) scaffolds fabricated from 10% dioxane solution quenched at 70 C. (a) 150 and (b) 500.<br />

and MHAP incorporated scaffolds were not statistically<br />

significant. With the dramatic changes in morphology of<br />

scaffolds prepared using different solvent systems, the<br />

protein adsorption also changed greatly. Scaffolds with<br />

fibrous texture absorbed about four times greater<br />

amount of protein than those with smooth pore wall<br />

surface. These results suggested that the protein<br />

adsorption on the scaffolds could be regulated by the<br />

content of HAP, its size scale, and the pore wall<br />

morphology of the scaffolds.<br />

4. Discussion<br />

The present studies described a phase separation<br />

technique to fabricate polymer/NHAP-composite scaffolds<br />

with high porosity and controlled pore architecture.<br />

Different solvent systems were used to obtain<br />

scaffolds with different microarchitectures and properties.<br />

When dioxane was used alone, the porous structure<br />

resulted from a solid–liquid phase separation of the<br />

polymer solution. During quenching, the solvent crystallized<br />

and the polymer was expelled from the solvent<br />

crystallization front. Solvent crystals became pores after<br />

subsequent sublimation. The characteristics of the pores<br />

were determined by the morphologies of solvent crystals<br />

under such quenching conditions. The temperature<br />

gradient along the solvent crystallization direction led<br />

to anisotropic structure. The introduction of HAP<br />

particles into the polymer solution perturbed the solvent<br />

crystallization to some extent and thereby made the pore<br />

structure more irregular and isotropic. As expected,<br />

MHAP/PLLA composite scaffold exhibited an isotropic<br />

and irregular pore structure. The perturbation by<br />

NHAP particles, however, was small even in high<br />

proportion up to 50% due to their nanometer size scale<br />

and uniform distribution. As a result, NHAP/PLLA<br />

composite scaffolds maintained the main characteristic<br />

pore architecture of solid–liquid phase separation which<br />

was anisotropic and regular. In contrast to NHAP/<br />

PLLA, the regular anisotropic pore structure was<br />

obtained only when the HAP content was very low in<br />

MHAP/PLLA scaffolds, e.g., MHAP/PLLAo10/90<br />

(images not shown). In this case, low content of HAP<br />

did not affect the solvent crystallization significantly<br />

enough to alter the pore structure.<br />

When mixed solvent (dioxane/water) was used to<br />

prepare 5% PLLA solutions, a liquid–liquid phase<br />

separation was induced and this process dominated the<br />

final pore structure of the system [29], resulting in<br />

random and interconnected pore architecture. The pore<br />

size was also much smaller. With the increase of water<br />

content in the system, larger quench depth allowed the<br />

system to continue on a coarsening process to further<br />

reduce the interfacial free energy between two phases<br />

separated in earlier stage [30]. The coarsening effect<br />

might have resulted in the formation of macropores in<br />

addition to the micropores (Figs. 3d–f). It is therefore<br />

possible to optimize various phase separation parameters<br />

to achieve a macroporous structure in composite<br />

scaffold. The mechanism of fibrous texture formation is<br />

not fully understood. Since PLLA is a semicrystalline<br />

polymer, the crystallization in the polymer-rich phase<br />

may present a possible explanation for the formation of<br />

the fibrous structure [9].<br />

Most mammalian cells are anchorage-dependent cells<br />

and they need a biocompatible substrate for attachment,<br />

migration and differentiation to form new tissues.<br />

Recent research demonstrated that cell adhesion and<br />

survival could be modulated by protein pre-adsorption<br />

on the substrate [25–28]. Therefore, protein adsorption<br />

is of importance in evaluating a scaffold for tissue<br />

engineering. In this paper, we have shown that the<br />

addition of hydroxyapatite particles into the composite<br />

scaffolds greatly increased protein adsorption. In addition<br />

to the high affinity of hydroxyapatite itself for


ARTICLE IN PRESS<br />

G. <strong>Wei</strong>, P.X. Ma / Biomaterials 25 (<strong>2004</strong>) 4749–4757 4755<br />

Fig. 3. SEM micrographs of NHAP/PLLA (30:70) scaffolds fabricated using dioxane/water mixture solvents. (a) Dioxane:water=95:5, 500. (b,c)<br />

Dioxane:water=90:10, 500, 8000. (d,e,f) Dioxane:water=87:13, 45, 500, 10000.<br />

proteins, incorporation of HAP also altered the pore<br />

surface morphology of the composite scaffolds and may<br />

have made them more suitable for protein adsorption.<br />

Furthermore, NHAP/PLLA scaffold absorbed significantly<br />

more proteins than MHAP/PLLA at high HAP<br />

content but not at low HAP content. It was likely that<br />

most particles were embedded in the polymer pore walls<br />

and the effect of HAP size was not evident when the<br />

HAP content was low in the composite. With the<br />

increase of HAP content, more and more particles were<br />

exposed on the surfaces of the pore walls. This<br />

phenomenon was most evident when the scaffold<br />

exhibited fibrous morphology. Fibrous scaffolds had<br />

much greater surface to volume ratio than scaffolds with<br />

solid pore walls, which might have further increased the<br />

protein adsorption capacity.<br />

5. Conclusion<br />

The present study investigated new NHAP/polymer<br />

composite scaffolds developed using thermally induced<br />

phase separation techniques. Nano-sized HAP particles<br />

were successfully incorporated into the PLLA porous


4756<br />

ARTICLE IN PRESS<br />

G. <strong>Wei</strong>, P.X. Ma / Biomaterials 25 (<strong>2004</strong>) 4749–4757<br />

scaffolds. The incorporation of NHAP improved the<br />

mechanical properties and protein adsorption of the<br />

composite scaffolds while maintaining high porosity and<br />

suitable microarchitecture. These results suggest that the<br />

newly developed NHAP/polymer composite scaffolds<br />

may be superior for bone tissue engineering.<br />

Fig. 4. Effects of NHAP content on the compressive modulus of<br />

NHAP/PLLA scaffolds fabricated from 5% dioxane solutions<br />

% po0.05, w p=0.29, n=5.<br />

Fig. 5. Effects of mixture solvent composition and quenching<br />

temperature on the compressive moduli of NHAP/PLLA scaffolds<br />

fabricated from 5% dioxane/water solvent systems.<br />

Fig. 6. Protein adsorption on porous composite scaffolds incubated in<br />

FBS/PBS solution. (a) as a function of time, 2.5% FBS/PBS, labels<br />

indicate the w/w ratios of NHAP to PLLA; (b) as a function of NHAP<br />

content, 2.5% FBS; (c) Comparison of MHAP and NHAP, 5.0% %<br />

FBS/PBS. po0.05, w p=0.71, n=3. (d) Comparison of NHAP/PLLA<br />

(30:70) scaffolds prepared using different solvent systems, 2.5% %<br />

FBS/PBS. po0.05, n=3.


ARTICLE IN PRESS<br />

G. <strong>Wei</strong>, P.X. Ma / Biomaterials 25 (<strong>2004</strong>) 4749–4757 4757<br />

References<br />

[1] Braddock M, Houston P, Campbell C, Ashcroft P. Born again<br />

bone: tissue engineering for bone repair. News Physiol Sci<br />

2001;16:208–13.<br />

[2] Goldberg VM, Stevenson S. Natural history of autografts and<br />

allografts. Clin Orthop 1987;225:7–16.<br />

[3] Costantino PD, Friedman CD. Synthetic bone graft substitutes.<br />

Otolaryngol Clin North Am 1994;27:1037–74.<br />

[4] Bonfiglio M, Jeter WS. Immunological responses to bone. Clin<br />

Orthop 1972;87:19–27.<br />

[5] Langer R, Vacanti JP. Tissue engineering. Science 1993;260:920–6.<br />

[6] Ma PX, Zhang R, Xiao G, Franceschi R. Engineering new bone<br />

tissue in vitro on highly porous poly(alpha-hydroxyl acids)/<br />

hydroxyapatite composite scaffolds. J Biomed Mater Res<br />

2001;54:284–93.<br />

[7] Zhang R, Ma PX. Porous poly(l-lactic acid)/apatite composites<br />

created by biomimetic process. J Biomed Mater Res 1999;45:285–93.<br />

[8] Zhang R, Ma PX. Poly(alpha-hydroxyl acids)/hydroxyapatite<br />

porous composites for bone tissue engineering. I. Preparation and<br />

morphology. J Biomed Mater Res 1999;44:446–55.<br />

[9] Ma PX, Zhang R. Synthetic nano-scale fibrous extracellular<br />

matrix. J Biomed Mater Res 1999;46:60–72.<br />

[10] Ma PX, Zhang R. Micro-tubular architecture of biodegradable<br />

polymer scaffolds. J Biomed Mater Res 2001;56:469–77.<br />

[11] Mikos AG, Thorsen AJ, Czerwonka LA, Bao Y, Langer R,<br />

Winslow DN, Vacanti JP. Preparation and characterization of<br />

poly(l-lactic acid) foams. Polymer 1994;35:1068–77.<br />

[12] Whang K, Thomas CH, Healy KE, Nuber G. A novel method to<br />

fabricate bioabsorbable scaffolds. Polymer 1995;36:837–42.<br />

[13] Mooney DJ, Baldwin DF, Suh NP, Vacanti JP, Langer R. Novel<br />

approach to fabricate porous sponges of poly(d,l-lactic-coglycolic<br />

acid) without the use of organic solvents. Biomaterials<br />

1996;17:1417–22.<br />

[14] Borden M, Attawia M, Laurencin CT. The sintered microsphere<br />

matrix for bone tissue engineering: in vitro osteoconductivity<br />

studies. J Biomed Mater Res 2002;61:421–9.<br />

[15] Visscher GE, Robison RL, Maulding HV, Fong JW, Pearson JE,<br />

Argentieri GJ. Biodegradation of and tissue reaction to 50:50<br />

poly(DL-lactide-co-glycolide) microcapsules. J Biomed Mater Res<br />

1985;19:349–65.<br />

[16] Li SH, De Wijn JR, Layrolle P, de Groot K. Synthesis of<br />

macroporous hydroxyapatite scaffolds for bone tissue engineering.<br />

J Biomed Mater Res 2002;61:109–20.<br />

[17] Flautre B, Descamps M, Delecourt C, Blary MC, Hardouin P.<br />

Porous HA ceramic for bone replacement: role of the pores and<br />

interconnections—experimental study in the rabbit. J Mater Sci-<br />

Mater Med 2001;12:679–82.<br />

[18] LeGeros RZ. Properties of osteoconductive biomaterials: calcium<br />

phosphates. Clin Orthop 2002;395:81–98.<br />

[19] Cooke FW. Ceramics in orthopedic surgery. Clin Orthop<br />

1992;276:135–46.<br />

[20] Thomson RC, Yaszemski MJ, Powers JM, Mikos AG. Hydroxyapatite<br />

fiber reinforced poly(alpha-hydroxy ester) foams for<br />

bone regeneration. Biomaterials 1998;19:1935–43.<br />

[21] Ishaug SL, Crane GM, Miller MJ, Yasko AW, Yaszemski MJ,<br />

Mikos AG. Bone formation by three-dimensional stromal<br />

osteoblast culture in biodegradable polymer scaffolds. J Biomed<br />

Mater Res 1997;36:17–28.<br />

[22] Rho JY, Kuhn-Spearing L, Zioupos P. Mechanical properties<br />

and the hierarchical structure of bone. Med Eng Phys 1998;20:<br />

92–102.<br />

[23] Du C, Cui FZ, Zhu XD, de Groot K. Three-dimensional nano-<br />

HAp/collagen matrix loading with osteogenic cells in organ<br />

culture. J Biomed Mater Res 1999;44:407–15.<br />

[24] Du C, Cui FZ, Feng QL, Zhu XD, de Groot K. Tissue response<br />

to nano-hydroxyapatite/collagen composite implants in marrow<br />

cavity. J Biomed Mater Res 1998;42:540–8.<br />

[25] Webster TJ, Ergun C, Doremus RH, Siegel RW, Bizios R.<br />

Specific proteins mediate enhanced osteoblast adhesion on<br />

nanophase ceramics. J Biomed Mater Res 2000;51:475–83.<br />

[26] Webster TJ, Siegel RW, Bizios R. Osteoblast adhesion on<br />

nanophase ceramics. Biomaterials 1999;20:1221–7.<br />

[27] Woo KM, Zhang R, Deng H, Ma PX. Protein-mediated<br />

osteoblast survival and migration on biodegradable<br />

polymer/hydroxyapatite composite scaffolds. Transaction of the<br />

28th Annual Meeting of the Society for Biomaterial, vol. 605,<br />

2002.<br />

[28] Woo KM, <strong>Wei</strong> G, Ma PX. Enhancement of fibronectin- and<br />

vitronectin-adsorption to polymer/hydroxyapatite scaffolds suppresses<br />

the apoptosis of osteoblasts. J Bone Miner Res<br />

2002;17(Suppl 1):M49.<br />

[29] Zhang R, Ma PX. Processing of polymer scaffolds: phase<br />

separation. In: Atala A, Lanza R, editors. Methods of tissue<br />

engineering. San Diego, CA: Academic Press; 2001. p. 715–24.<br />

[30] Nam YS, Park TG. Porous biodegradable polymeric scaffolds<br />

prepared by thermally induced phase separation. J Biomed Mater<br />

Res 1999;47:8–17.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!