Elsevier

Acta Biomaterialia

Volume 5, Issue 4, May 2009, Pages 1006-1018
Acta Biomaterialia

Density–property relationships in mineralized collagen–glycosaminoglycan scaffolds

https://doi.org/10.1016/j.actbio.2008.11.029Get rights and content

Abstract

Mineralized collagen–glycosminoglycan scaffolds have previously been fabricated by freeze-drying a slurry containing a co-precipitate of calcium phosphate, collagen and glycosaminoglycan. The mechanical properties of the scaffold are low (e.g. the dry Young’s modulus for a 50 wt.% mineralized scaffold is roughly 780 kPa). Our previous attempt to increase the mechanical properties of the scaffold by increasing the mineralization (from 50 to 75 wt.%) was unsuccessful due to defects in the more mineralized scaffold. In this paper, we describe a new technique to improve the mechanical properties by increasing the relative density of the scaffolds. The volume fraction of solids in the slurry was increased by vacuum-filtration. The slurry was then freeze-dried in the conventional manner to produce scaffolds with relative densities between 0.045 and 0.187 and pore sizes of about 100–350 μm, values appropriate for bone growth. The uniaxial compressive stress–strain curves of the scaffolds indicated that the Young’s modulus in the dry state increased from 780 to 6500 kPa and that the crushing strength increased from 39 to 275 kPa with increasing relative density. In the hydrated state, the Young’s modulus increased from 6.44 to 34.8 kPa and the crushing strength increased from 0.55 to 2.12 kPa; the properties were further increased by cross-linking. The modulus and strength were well described by models for cellular solids.

Introduction

Scaffolds for tissue regeneration are defined as: “three-dimensional open-cell porous structures synthesized from either natural or synthetic polymers which have the potential to support attachment, migration and multiplication of living cells” [1]. Although unproven, a widely believed design paradigm for scaffolds is that mimicking the composition of the natural tissue as closely as possible improves the capacity for regeneration [2]. The ability of a scaffold to regenerate tissue depends on its pore size, pore shape, porosity, biodegradability and mechanical properties. The average pore diameter must be large enough for cells to migrate through the pores yet small enough to retain an appropriate specific surface area for sufficient cell binding. For example, pore sizes in excess of 100 μm are optimal for bone growth [2], [3], [4]. Equiaxed pore shape and homogeneity are optimum for uniform cell adhesion and distribution of extracellular matrix proteins. Scaffolds must have large enough porosity (generally greater than 90%) and interconnectivity for effective transfer of cells and metabolites [5]. The degradation rate of the scaffold has to be roughly equal to the regeneration rate of the tissue. Furthermore, cells have been observed to be sensitive to the mechanical properties of the scaffold, which in turn affects the overall construct bioactivity [6].

Ideally, scaffolds should be similar to their natural counterparts in terms of chemical composition and physical structure. For this reason, natural polymers such as collagen are of major interest. To this end, collagen–glycosaminoglycan (CG) scaffolds have been developed and used clinically for skin regeneration and experimentally for nerve regeneration over the past three decades [7], [8], [9], [10], [11], [12], [13], [14], [15], [16], [17], [18], [19]. Composite scaffolds of collagen or gelatin with ceramics (e.g. hydroxyapatite and tricalcium phosphate), i.e. mineralized CG (MCG) scaffolds, have been developed to regenerate hard tissues such as bone [20], [21], [22], [23]. The most recent fabrication technique improves upon this mineralization process by forming a triple co-precipitate of mineral, collagen and glycosaminoglycan, without using a titrant, by controlling the molarity of the reactant acid and molar ratios of the different calcium sources [24], [25], [26], [27], [28], [29], [30]. Due to the in situ co-precipitation of the mineral phase, calcium phosphate crystals form within the collagen fibers, resulting in a more uniformly mineralized scaffold. Freeze-drying is then used to fabricate porous scaffolds from the triple co-precipitated slurry. These MCG scaffolds have regenerated subchondral bone at 16 weeks in a 4 mm diameter and 6 mm deep defect site at the knee joint in a goat model [31].

Extensive microstructural and mechanical characterization of CG and of MCG scaffolds of varying mineral content has been reported by Harley et al. [32] and by Kanungo et al. [30], respectively. Critical mechanical properties of scaffolds include elastic modulus, E, compressive crushing strength, σ, and compressive crushing strain, ε. The mechanical properties of different MCG scaffolds (along with the triple co-precipitated scaffolds) have been compared in the literature [30]. The triple co-precipitated mineralized scaffolds have relative densities (the density of the cellular solid, ρ, divided by that of the solid from which it is made, ρs) of roughly 0.03–0.04; that of trabecular bone varies from 0.05 to 0.60 [33], [34]. The mechanical properties of human compact and trabecular bone, along with 50 wt.% MCG scaffold (with a relative density, ρ/ρs, of 0.04), are listed in Table 1.

It is critical that the scaffold should have sufficient stiffness and strength to maintain its shape and size during surgical procedures such as implantation and to enhance bone in-growth while preventing encroachment of non-osseous tissue and competing cell types after implantation [35]. The optimal requirements for the above properties vary depending on the defect site and there are no established optimal magnitudes of the mechanical properties for bone scaffolds [36]. The current triple co-precipitated MCG scaffold (with a ρ/ρs of 0.04) can be crushed by hard thumb pressure. Hence, it is critical to improve the mechanical properties of MCG scaffolds such that they can be functionally suitable for bone regeneration. The mechanical properties (E and σ) of the scaffold depend on those of the solid (Es and σfs) they are made from as well as the relative density of the scaffold, (ρs) [5], [30], [32], [37]. The overall properties of the scaffolds can be improved by either improving the properties of the solid it is made from or by increasing the relative density of the scaffold. Previous attempts to increase the mechanical properties of the scaffold by increasing the mineral content led to scaffolds with poorer mechanical properties due to the introduction of defects [30]. Our previous attempts to improve the mechanical properties by increasing the volume fraction of the components of the slurry have not been successful due to the difficulty in mixing the viscous slurry at higher volume fractions of the mineral, collagen and GAG [32]. In this paper we describe a new technique to improve the mechanical properties by increasing the relative density of the scaffold by a vacuum filtration technique.

Section snippets

Fabrication of mineralized collagen–glycosaminoglycan suspension

A mineralized CG suspension (50 wt.% mineral) was fabricated using microfibrillar, type I collagen isolated from bovine achilles tendon (Sigma–Aldrich Chemical Co, St. Louis, MO), chondroitin-6-sulfate (GAG) isolated from shark cartilage (Sigma–Aldrich), phosphoric acid (H3PO4) (EMD Chemicals Inc., Gibbstown, NJ), calcium nitrate (Ca(NO3)2·4H2O) and calcium hydroxide (Ca(OH)2) (Sigma–Aldrich). The suspension was prepared by combining collagen (0.019 wt.%), GAG (0.002 wt.%), calcium nitrate

Microstructural characterization

The measured overall scaffold dry densities (ρ) for the MCG scaffolds are listed in Table 2. The starting densities of the co-precipitate in the slurry were less than the dry scaffold densities (e.g. for the 1× scaffold, 0.042 g ml−1 and 0.076 g cm−3 are the slurry and dry densities, respectively) partly because some of the solvent was soaked into the collagen reducing the solidified solvent volume (and hence increasing the dry density) [49], [50], [51]. The relative densities can be calculated

Discussion

We were able to increase substantially the Young’s modulus and the crushing strength of the mineralized scaffold in both the dry and hydrated state by increasing the relative density by a factor of 3. The denser scaffolds, 2× and 3×, in the dry state sustained hard thumb pressure; this is a critical criterion for scaffold implantation at a defect site. Cross-linking the scaffolds with EDAC further increased the properties of the hydrated, but not the dry scaffolds. We were able to achieve an

Conclusions

50 wt.% MCG scaffolds with four different relative densities (0.045, 0.098, 0.137 and 0.187) were fabricated via a three step process: (i) titrant-free triple co-precipitation of the slurry; (ii) vacuum filtering the slurry to the desired density; and (iii) freeze-drying the slurry to obtain the dry scaffold. The MCG scaffolds had an open-cell pore structure with both walls and struts, and interconnected pores. While we have not definitively demonstrated by cell adhesion, migration,

Conflicts of interest statement

Lorna J. Gibson has a financial interest in Orthomimetics, a start-up firm that resulted from a previous collaboration on a similar mineralized collagen scaffold. However, the authors did not receive any financial support from Orthomimetics for this project.

Acknowledgments

Funding for this project was provided by the National Science Foundation, Grant No. CMS-0408259. The authors are grateful to Prof. Ioannis Yannas and Prof. Simona Socrate in the MIT Department of Mechanical Engineering, Prof. Elazer Edelman in the Harvard-MIT Division of Health Sciences & Technology, Mr. Philip Seifert in the CBSET Department of Pathology, Mr. Alan Schwartzman in the MIT Department of Materials Science and Engineering, Mr. Kaustuv DeBiswas in the MIT Department of Architecture,

References (65)

  • F.J. O’Brien et al.

    The effect of pore size on cell adhesion in collagen–GAG scaffolds

    Biomaterials

    (2005)
  • F.J. O’Brien et al.

    Influence of freezing rate on pore structure in freeze-dried collagen–GAG scaffolds

    Biomaterials

    (2004)
  • P.E. Danielsson

    Euclidean distance mapping

    Comput Graph Image Process

    (1980)
  • L.J. Gibson

    Biomechanics of cellular solids

    Journal of Biomechanics

    (2005)
  • M. Kikuchi et al.

    Self-organization mechanism in a bone-like hydroxyapatite/collagen nanocomposite synthesized in vitro and its biological reaction in vivo

    Biomaterials

    (2001)
  • C.H. Turner et al.

    The elastic properties of trabecular and cortical bone tissues are similar: results from two microscopic measurement techniques

    J. Biomech.

    (1999)
  • J.Y. Rho et al.

    Elastic properties of human cortical and trabecular lamellar bone measured by nanoindentation

    Biomaterials

    (1997)
  • Y. Chevalier et al.

    Validation of a voxel-based FE method for prediction of the uniaxial apparent modulus of human trabecular bone using macroscopic mechanical tests and nanoindentation

    J Biomech

    (2007)
  • M.J. Silva et al.

    The effects of non-periodic microstructure and defects on the compressive strength of two-dimensional cellular solids

    Int J Mech Sci

    (1997)
  • M.J. Silva et al.

    The effects of nonperiodic microstructure on the elastic properties of 2-dimensional cellular solids

    Int J Mech Sci

    (1995)
  • X.E. Guo et al.

    Behavior of intact and damaged honeycombs: a finite element study

    Int J Mech Sci

    (1999)
  • V. Karageorgiou et al.

    Porosity of 3D biomaterial scaffolds and osteogenesis

    Biomaterials

    (2005)
  • Yannas IV, Spector M. Design of medical devices and implants. MIT Course No.: 3.961 2006, Massachusetts Institute of...
  • J.O. Hollinger et al.

    Bone tissue engineering

    (2005)
  • Harley BA. Cell–matrix interactions: collagen-GAG scaffold fabrication, characterization, and measurement of cell...
  • I.V. Yannas et al.

    Synthesis and characterization of a model extracellular-matrix that induces partial regeneration of adult mammalian skin

    Proc Nat Acad Sci USA

    (1989)
  • I.V. Yannas

    Prompt, long-term functional replacement of skin

    Trans Am Soc Artif Int Organs

    (1981)
  • I.V. Yannas et al.

    Design of an artificial skin. 2. Control of chemical-composition

    J Biomed Mater Res

    (1980)
  • I.V. Yannas et al.

    Design of an artificial skin. 1. Basic design principles

    J Biomed Mater Res

    (1980)
  • Yannas IV. Facts and theories of induced organ regeneration. In: Regenerative medicine I: theories, models and methods....
  • I.V. Yannas

    Similarities and differences between induced organ regeneration in adults and early foetal regeneration

    J R Soc Interf

    (2005)
  • I.V. Yannas

    Biologically-active analogs of the extracellular-matrix—artificial skin and nerves

    Angewandte Chemie-Int Ed English

    (1990)
  • Cited by (0)

    View full text