Open access

Cartilage Tissue Engineering: the Application of Nanomaterials and Stem Cell Technology

Written By

Adelola O. Oseni, Claire Crowley, Maria Z. Boland, Peter E. Butler and Alexander M. Seifalian

Submitted: 19 November 2010 Published: 17 August 2011

DOI: 10.5772/22453

From the Edited Volume

Tissue Engineering for Tissue and Organ Regeneration

Edited by Daniel Eberli

Chapter metrics overview

6,365 Chapter Downloads

View Full Metrics

1. Introduction

Replacement and reconstruction of pathological or absent cartilage within the human body has been a clinical challenge for many years. The avascular nature of cartilage tissue in all areas of the human body means it has little capacity for regeneration or repair beyond the production of functionally inferior fibrocartilage. Cartilage is injured in a number of ways; in the joint region, repetitive stress can cause irreparable damage, eventually resulting in Osteoarthritis, a debilitating disorder managed only with pain medication or joint replacement. A rise in the incidence of cancer has increased the prevalence of tracheal and nasal cancers, both frequently requiring radical resection as part of aggressive treatment regimes. Congenital disorders, such as Treacher Collins syndrome and Aperts syndrome can cause severe malformation of the ear and nose. It is evident that each of these clinical scenarios involves extensive damage to crucial skeletal cartilage and it is for these reasons that a drive for advancements in cartilage tissue engineering exists.

Tissue engineering uses principles of cell biology, engineering and medicine to generate constructs that can successfully recapitulate the function of native tissues in terms of histology, mechanics and morphology. There is a need for a suitable scaffold that can provide a 3D environment for cells to proliferate and adhere. Debate still continues over the key characteristics needed for the ideal scaffold, but they are likely to differ according to the type and location of cartilage to be engineered. Should it be biodegradable/non biodegradable, natural/synthetic, and what impact do these features have on the flexibility and strength of neocartilage constructs produced? There are many scaffolds that have been extensively investigated in cartilage tissue engineering research from natural collagen and alginate, to the synthetic Polyhydroxyacids and PEG hydrogels. Nonetheless, despite advancements in scaffold design, neocartilage constructs are still mechanically inferior to their natural counterparts, and in vivo problems of poor biointegration, and deterioration in tissue quality over time limit there translation into clinical use.

Nanomaterial science has introduced new methods for improving scaffold quality. Scaffolds can now be engineered on the nanoscale, using techniques such as electrospinning and 3D fibre deposition. Likewise, the incorporation of nanoparticles into polymeric material has allowed the addition of nanoscale features into the matrix structure. Both of these methods produce scaffolds that more closely replicate the extra cellular matrix environment found in native cartilage. It is hoped that this will increase cellular interaction with the scaffold and improve the quality of constructs produced.

With regards to the cell population to be used for engineering these constructs, there is a continued excitement over the possible application of stem cell technology. Stem cells are highly replicative and have multi lineage differentiation capacity. The traditional source of mesenchymal stem cells (precursor of chondrocytes) is bone marrow and various adjuncts to their propagation and differentiation have been explored in detail, such as growth hormones, biomaterials and environmental factors such as shear stress and oxygen tension that are important for culture techniques and bioreactor design. The discovery of new sources of mesenchymal stem cells, such as blood, adipose tissue or the synovium opens up a plethora of possibilities for clinical application, where methods of isolation and differentiation are being optimized.

In light of the numerous advancements that have been made in the last decade, this chapter aims to give a detailed account of cartilage tissue engineering strategy, looking with particular focus at the effect of scaffolds on cell growth, the evolution of stem cell technology and the expansion of bioreactor design and application. We will also explore how an integration of this revolutionary and innovative bench work can be translated into much needed clinical application in the not too distant future.

Advertisement

2. Cartilage in the human body

2.1. Cartilage tissue biology

Cartilage is a flexible connective tissue found in many areas of the human body, including the joints, ribs, nose, ear, trachea and intervertebral discs (Fig 1). In these regions cartilage can act as structural support, maintain shape or absorb shock during physical exercise. Unlike most other connective tissues, cartilage is largely avascular leading to hypoxic environments that limit the rate of cellular growth and tissue regeneration (115; 116). This in turn limits the capacity of cartilage to repair itself in the event of damage. The main cellular component of cartilage are chondrocytes, highly specialized cells that lie within spaces called lacunae and secrete the extracellular matrix (ECM) of cartilage. As with most connective tissues, the ECM of cartilage consists of a meshwork of macromolecules including collagens, elastin, glycoproteins and proteoglycans, each of which is present in varying amounts, depending on the type and function of cartilage. There are several cell surface receptors that allow chondrocytes to bind these proteins including the integrins, CD44, and the proteoglycan family of receptors e.g. syndecan (144).

The three types of cartilage are hyaline, elastic and fibrocartilage. Hyaline is the most abundant type, white-blue in colour and macroscopically smooth on its surface. It is present on the articular surfaces of joints and in the nasal septum. Hyaline cartilage is covered externally by a fibrous membrane known as the perichondrium, and in the joint especially, it is diffusion from the synovial fluid that provides this tissue with nutrition. It is rich in collagen type II, which forms a meshwork that encases giant aggregates of proteoglycans (Proteins with glycosaminoglycan (GAG) side chains e.g. aggrecan, biglycan, decorin in the extracellular matrix; syndecan, CD44 and fibroglycan as cell surface receptors; serglycan in intracellular tissues) (20; 21). These GAG side chains, keratan and chondroitin sulphate are able to retain water. Cyclical pressures from joint loading are crucial for normal hyaline cartilage function, and encourage the passage of water and nutrients between cartilage and synovial fluid. Elastic cartilage however, is more flexible, due to its rich elastin fibre content that is woven into a collagen mesh (20; 21). Elastin is an insoluble protein polymer that when cross linked with desmosine and isodesmosine make up the elastin fibres themselves. This type of cartilage is also surrounded by perichondrium and is more commonly found in the pinna, Eustachian tube, larynx and epiglottis, providing crucial structural support and flexibility. The third type of cartilage is fibrocartilage, which contains abundant thick collagen type I in addition to type II, that are interlaced into a mesh work of longitudinal and circumferential fibres (20; 21). These collagen bundles impart a great ability for this type of cartilage to withstand high tensile stresses. Fibrocartilage is usually found with the intervertebral discs, sacroiliac joints, pubic symphysis and costochondral joints.

Figure 1.

A diagrammatic representation of cartilaginous regions in the human body

2.2. Development of cartilage

Central to effective tissue engineering practice is the understanding of tissue origin and development. This is based on the widely accepted hypothesis that natural tissue regeneration recapitulates developmental processes (118); hence embryological study can give an insight into the regulatory processes and patterns that govern tissue type and function, in addition to forming a foundation for understanding the degeneration and damage seen in tissues of the human body. We can only give a brief outline of the development of cartilaginous tissue specifically, however interested parties are advised to consult specific reviews (17; 70; 121) and books that devote entire chapters to this topic.

One of the epicentres of skeletal cartilage development is the growth plate which produces long bones via the cartilage template in a process known as endochondral ossification. It is important to note that this process is specific to the articular regions of bones and is followed by the eventual replacement by bony tissue. A milieu of hormones and paracrine factors govern a complex interplay of chondrocyte proliferation and differentiation, and the process can be divided into five stages, with the first three mainly being crucial for cartilage formation (144). Mesenchymal stem cells (MSCs) are first committed to becoming chondrocytes by paracrine factors that induce the expression of key transcription factors Pax 1 + scleraxis, which in turn activate cartilage specific genes. During the second stage, the committed MSCs condense into compact nodules and differentiate into chondrocytes. Chondrocytes then proliferate rapidly during the third stage, increase their cytoplasmic contents and secrete large amounts cartilage specific ECM. Their volume increases 5-10 fold, proliferation slows and they are known as hypertrophic. After this stage, the expression profile of the cells change and collagen type X and fibronectin are secreted, enabling mineralization by calcium carbonate and osteoblast infiltration to make bone. Vascular infiltration leads the way to terminal differentiation and bone development, resulting in chondrocyte apoptosis and osteoblastic differentiation. Facial cartilage development is very different, as it is embryologically derived from the cranial neural crest cells that originate from the anterior hindbrain. These cells migrate to specific locations and differentiate under the instruction of an array of Hox genes, the complexities of which are outside the scope of this chapter.

2.3. Clinical need for cartilage

Due to the limited self healing capacity of human cartilage, the repair of defects caused by degenerative joint diseases, cancer or trauma gives rise to a challenging clinical problem. In the joint region in particular, lesions of the articular cartilage are frequently associated with debilitating pain and reduced functionality. If not successfully treated long term disability can only be averted by total replacement of the joint. Damage to facial cartilaginous structures such as the nose or ear are only resolved with a prosthesis or autologous transplantation surgery that results in the formation of a donor site and frequently requires a number of revision surgeries. Large scale damage to the trachea has even less options for reconstruction with stents and tracheotomy tubes being the mainstay of treatment.

Cartilage regeneration has always been a key therapeutic target for treating articular cartilage damage in particular. Popular marrow stimulating techniques using micro-fracture or subchondral drilling of the bone have been developed to encourage the invasion of mesenchymal progenitor cells (MPC) into the affected site for spontaneous cartilage repair (94; 109). Unfortunately the outcome of such techniques varies greatly due to the lack of biological instructions for the MPCs to follow. This results in the formation of fibrocartilage which compared with hyaline tissue, has reduced durability and functionality (87; 140). The later invention of cell based therapies such as Autologous Chondrocyte Implantation (ACI) provided an important breakthrough treatment of articular cartilage damage and paved the way perhaps for more complex tissue engineering approaches with matrix assisted ACI introduced later on (16). ACI involves harvesting and propagating a population of autologous mature articular chondrocytes in vitro and re-introducing them into the defect site in cell suspension or in a supported matrix. They are then expected to lay down ECM to repair the site of injury (12; 102). Clinically the results of such procedures are good as they appear to provide symptomatic relief for patients. However, histologically the tissue produced is far inferior to native hyaline, being fibrotic in nature, again with limited functionality and durability (19; 66; 120; 141). Further evaluations of such techniques has evidenced a strong correlation between the quality of tissue produced and the symptomatic relief of the patient from swelling and pain, once again highlighting the importance of tissue quality in cartilage regeneration. It can be postulated that these clinical breakthroughs buttressed the drive for advancements in cartilage tissue engineering technique.

It is also essential to note that the desired characteristics of engineered cartilage depend heavily on the site to be reconstructed. For instance, in tracheal constructs mechanical integrity, strength, flexibility and durability are all crucial for function, whereas in the facial region the aesthetic properties of the constructs may be equally as important. Taking the specific requirements of the tissue to be regenerated into consideration informs the tissue engineering strategy and the expected out comes of such undertakings.

2.4. Tissue engineering cartilage

Cartilage tissue engineering paradigm is based on the isolation of chondrocytes/ chondrocyte precursors from a tissue biopsy, expanding the cell number in culture, seeding them onto 3D scaffold, incubating for a period of time before placing the construct inside a patient. Many studies over the last decades have demonstrated that animal cells, when utilised in this way can produce tissues approaching the biomechanical and histological properties of native cartilage, even after implantation in vivo (3; 8; 24; 48; 58; 64; 77; 99; 110; 113; 152). However challenges do arise regarding the translation of such academic success into the clinical scenario. Challenges include isolating and propagating primary human cells, gaining relevant and reproducible construct morphology and size and ensuring good durability of the construct in vivo. Cell phenotype regulation, in vitro expansion of cell numbers, scaffold design and suitability, bioreactor design are all crucial components of the tissue engineering process that need to be optimized to advance cartilage tissue engineering from a mere academic prologue, into a clinical reality and success. These challenges will be discussed at length in the rest of this chapter.

Figure 2.

A diagrammatic outline of tissue engineering strategy in cartilage tissue engineering for articular joint repair (an example of clinical usage)

Advertisement

3. Stem cell technology

Stem cells are unspecialised cells with a very high replication capacity. For cartilage regeneration, mesenchymal stem cells are the cell type of choice as they are multipotent stem cells, capable of differentiating into a number of lineages of the musculoskeletal system, including osteoblasts (bone cells), chondrocytes (cartilage cells) and adipocytes (fat cells). Although not immortal, these cells are capable of expanding through many passages in culture while retaining their growth and multi-lineage potential.

MSCs can originate from various tissues including the bone marrow (11), skeletal muscle, adipose (106; 159) synovium (134), the embryo and periosteum. The optimal cell source for cartilage tissue engineering is still being identified. When selecting an ideal cell source, it is important to achieve a number of criteria, including: (i) easy access to/harvesting of the source of MSCs, (ii) extensive self-renewal or expansion capability of the cells (to generate sufficient quantities of cells for large scale tissue engineering, (iii) the ability to readily differentiate into the chondrocytic lineage when induced, and (iv) a lack of or minimal immunogenicity or ‘tumourigenic’ tendencies. The two most commonly used MSC sources are adipose tissue and bone marrow. Unlike other sources such as embryonic tissue, there are few ethical issues associated with harvesting and using these tissues in research and development. Additionally, bone marrow MSCs (BMSCs) and adipose derived (ADMSCs) are relatively easy to source compared with synovium-derived- or periostium-derived MSCs.

Interestingly, bone marrow is the only organ in which at least two types of stem cells exist; hematopoietic and mesenchymal stem cells (137). The MSCs are found arrayed around the central sinus in the bone marrow. The cells can be isolated from the marrow using standardised techniques and expanded in culture through many generations, while retaining their capacity to differentiate along these pathways when exposed to appropriate culture conditions. Adipose tissue is an abundant, readily available source of MSCs. The cells can be isolated from fat that has been excised or ‘liposuctioned’ (lipoaspirate). There are advantages and disadvantages to both techniques. Anecdotally, it is thought that excised fat provides a higher yield of MSCs compared with lipoaspirate. This is due minimal mechancal impact upon cell membranes, which would ordinarily cause cell rupture, during the isolation process. Conversely, lipoaspirate is accessible without creating a large donor site defect, a major reason for pursuing tissue engineering methods at the outset. Some studies have compared adipose-derived MSCs and bone marrow-derived MSC (107; 122) and found that both BMSCs and ADASCs are capable of chondrogenic differentiation. There is some debate over which is the superior cell source, with numerous papers highlighting each source at optimum (107).

Mesenchymal stem cells can be identified using a number of methods. These include i) examination of cell morphology, ii) FACS (fluorescence activated cell sorting) analysis to detect the expression of MSC specific markers and iii) proving their differentiation capacity by differentiating the cells into a number of lineages, namely; osteoblastic, adipocytic, and chondrocytic. For FACS analysis, the presence of MSC-specific cell surface proteins such as the following are sought; CD 105 (SH2), CD 90 (THY1) and CD 73 (SH3/4). Similarly, negative markers are used to mark and remove cells expressing cell surface proteins not typically seen on MSCs, such as CD 45, CD 34, and CD 14 (9).

3.1. Stem cell differentiation to chondrocytes

Chondrogenesis is the term used to describe the process by which a stem cell is differentiated into a mature chondrocyte and is one of the earliest morphogenetic events of embryonic development (112). The stages where introduced earlier in the section on cartilage tissue biology. They include: MSC condensation, the rise of chondroprogenitors, chondrogenesis, terminal differentiation of progenitor cells and in skeletal development ossification (29).

Figure 3.

Schematic diagram of the stages of chondrogenesis, the main growth factors involved in each stage and the accompanying alterations in ECM. (Adapted from Vinatier C. 2009 (144))

Condensation is directed by cell-cell and cell-matrix interactions as well as secreted factors interacting with their related receptors. Prior to condensation, the prechondrocytic MSCs secrete an extracellular matrix (ECM) which is rich in hyaluronan and collagen type I that prevents intimate cell-cell interaction (60). When condensation is initiated, there is an increase in hyaluronidase activity, thus causing a decrease in hyaluronan in the ECM. It is thought that hyaluronan facilitates cell movement, therefore, the increase in hylauronidase and subsequent decrease in hyaluronan allows for close cell-cell interactions. The establishment of the cell-cell interactions is thought to be involved in triggering one or more of the signal transduction pathways that initiate chondrogenic differentiation. Cell adhesion molecules: N-cadherin (neural cadherin) and N-CAM (neural cell adhesion molecule) are also expressed in the condensing mesenchyme, but disappear in the differentiated cartilage, and later can be found only in the perichondrium. Perturbing the function of N-cadherin or N-CAM causes a reduction or alterations in chondrogenesis both in vitro and in vivo, which further evidences a role for these cell adhesion molecules in mediating the mesenchymal condensation step (46; 60).

As mentioned earlier, cell-matrix interactions also play an important role in mesenchymal cell condensation. Fibronectin is a component of the ECM which has the ability to regulate N-CAM. TGF-β1, one of the earliest signals in chondrogenic condensation, stimulates the synthesis of fibronectin. The expression of fibronectin is increased in areas of condensation and decreased as cytodifferentiation proceeds. Syndecan binds to fibronectin and down-regulates N-CAM thereby setting the condensation boundaries. One study showed that fibronectin mRNA undergoes alternative splicing during chondrogenesis. The isoform containing exon EIIIA is present during condensation but disappears once differentiation begins. This suggests that this isoform switching is important for cytodifferentiation to occur. A later study by the same group determined that the function of the fibronectin EIIIA exon seems to regulate mesenchymal cell spreading, therefore permitting and/or promoting adequate cell-cell interaction to take place during the condensation phase of chondrogenesis (46).

The differentiation of chondroprogenitors is characterized by the deposition of cartilage matrix containing collagens II, IX, XI and aggrecan. SOX9 is one of the earliest markers expressed in cells undergoing condensations. It is required for the expression of the type II collagen gene (Col2a1) and other cartilage-specific matrix proteins, including Col11a2 and CD-RAP prior to other matrix deposition in the cartilage anlagen. Two other Sox family members L-Sox5 and Sox6, are not expressed in early mesenchymal condensations, but are co-expressed with Sox 9 during chondrocyte differentiation. Figure 2 outlines the stages of chondrogenesis and the accompanying alterations to the ECM (46; 60).

3.2. Inducing chondrogenesis

3.2.1. Biochemical stimuli for cartilage tissue engineering

Specific biomolecules are essential for cartilage tissue engineering. The role of these biomolecules is primarily to induce chondrogenesis and to maintain the chondrocyte phenotype. There are five main families of growth factors involved in chondrogenic differentiation. These are: the transforming growth factor-β super-family (TGFβ), the fibroblast growth factor family (FGF), the insulin-like growth factor family (IGF), the wingless family (Wnt) and the hedgehog family(HH) (144). Below is a brief introduction to each growth family. However, for a more detailed description, the following references are recommended (46; 144). Figure 3 outlines the sequence in which the transcription factors are involved in each stage of chondrogenesis.

The transforming growth factor beta super-family is a family of proteins which have been shown to play a huge role in cartilage formation (11). Members include TGF-β, bone morphogenetic proteins (BMPs), inhibins and activins. All members have been shown to regulate cell growth, differentiation and apoptosis of a large number of different cell types including osteoblasts, chondrocytes, neural and epithelial cells. TGF-β is a secreted protein and exists in five isoforms TGF- β1-5. TFG- β 1,2 and 3 are thought to stimulate proteoglycans and type II collagen synthesis in chondrocytes as well as to induce chondrogenic differentiation of MSCs (144). Studies have also shown that TGF- β isoforms differ in their effects on various cell types. For example, TGF- β1 has been shown to be responsible for the initial cell-cell interactions between condensing progenitor cells and TGF-β2 mediates hypertrophic differentiation (32).

BMPs are also members of the TGF-β super-family and comprise of a group of 20 proteins each one playing an important role in chondrogenesis and osteogenesis during skeletal development. BMP -2,-4,-6,-7 and -13 have all proven their ability to stimulate chondrogenesis in MSCs (144). BMP2 in particular, has been found to be expressed in the condensing mesenchyme of the developing limb (101). It regulates chondrogenic development of mesenchymal progenitors (25) as well as stimulates the synthesis of chondrocyte matrix components by articular cartilage in vitro. Even combinations of many growth factors have been to enhance chondrogenesis, for example, BMP-2 with TGF-β3 and BMP-6 with TGF-β3 have been proven to stimulate chondrogenic differentiation and result in chondrogenic lineage development (73; 127).

The FGF family is a group of growth factors consisting of 22 members. Most FGFs are secreted, except for FGF1, 2 and FGF 11 and 14 (144). Signalling by FGF18 and FGF receptor 3 have demonstrated regulation, proliferation, differentiation and matrix production of articular cartilage and growth plate chondrocytes in vivo and in vitro (45).

The IGF family is a group of proteins which have a high similarity to insulin. IGF-1 has the ability to mediate chondrogenesis by increasing proteoglycan and collagen type II production (144). Combining TGF-β3 and IGF-1, has been shown to enhance chondrogenic induction (72). One study examined the effect that IGF-1 has on the chondrogenesis of bone marrow MSCs in the presence and absence of TGF-β signalling. It showed that IGF-1 could modulate MSC chondrogenesis by stimulating proliferation, regulating cell apoptosis and inducing expression of chondrocyte markers. In addition, it proved that the chondroinductive actions of IGF-1 were equally potent to TGF-β1 and independent from the TGF-beta signalling (98). Another similar study investigated the effects of IGF-1 on TGF-β1 induced chondrogenesis. It was found that the combination of IGF-1 and TGF-β1 produced higher amounts of glycosaminoglycan than TGF-beta1 alone at 8 weeks (124).

3.2.2. Mechanotransduction in cartilage tissue engineering

In vivo, articular cartilage experiences a variety of stresses and strains on a daily bases. Thus, many groups have extensively researched into the various mechanical stimulation methods for enhancing chondrogenic differentiation and cartilage tissue engineering (67; 68; 135).Examples of the types of mechanical stimuli examined include: hydrostatic pressure (4), cyclic mechanical compression (4), shear stress (139; 143), pulsed ultrasound (130) and dynamic compressive strain (23).

It has been found that exposing differentiating MSCs to various mechanical stimuli results in a shift in the types of protein expressed during chondrogenesis. For example, the application of cyclic, mechanical compression has been shown to result in an increase in proteoglycan and collagen contents as well as a higher amount of proteoglycan-rich, extracellular matrix production. Similarly, the application of shear stress by perfusion to differentiating BM-MSCs results in an enhanced ECM deposition and an increased collagen type II production (143). Low intensity pulsed ultrasound treated cell scaffold constructs show a significant increase of chondrogenic marker gene expression and extracellular matrix deposition in differentiating human BM-MSCs (130).

Collectively, the research shows that MSCs are mechanically sensitive and the chondrogenic differentiation can be modulated and enhanced by mechanical stimulation.

Advertisement

4. Biomaterials

It is well established that cells reside, proliferate and differentiate inside a complex 3-dimensional (3D) ECM environment. In cartilage, chondrocytes are surrounded by a highly hydrated matrix of proteins which informs many of their phenotypic states. For example, research has shown that isolated chondrocytes will loose their differentiated phenotypes if cultured in 2-dimension (42). These chondrocytes display a shift towards a fibroblastic phenotype, evident on protein assays and histological evaluations. Type I collagen expression is increased and the typical rounded morphology of the chondrocyte becomes spindle in shape (128). This process has been shown to be reversible upon relocation to 3D matrix environments such as pellet and micro-mass culture systems, which mimic the high cell density phenomenon seen during MSC condensation, a crucial stage of cartilage development (47; 74; 92).

It is in light of this that biomaterials have been proposed as engineered 3D environments in which chondrocytes can reside. For years, material scientists along with cell biologists have worked to optimize the tissue engineering characteristics of various biomaterials. There are a number of characteristics that are thought to be necessary for general tissue engineering attempts. These include a need for biomaterials to allow adequate cell adhesion and migration, with subsequent proliferation and differentiation. The overall architecture of the scaffold should guide and frame tissue formation, whilst providing mechanical support akin to that of native tissue. The scaffold should be porous, as porosity is thought to be crucial in maintaining the phenotype of the differentiated chondrocytes, considering their preference for 3D environments. It would also allow for mass transfer of nutrients and waste products. The scaffold should also be biocompatible with the ability to integrate into surrounding native tissue.

Biomaterial scaffolds can be broadly divided into natural and synthetic scaffolds. In this section we will give a brief overview of existing natural and synthetic scaffolds used for cartilage tissue engineering research, focusing on the regulatory influence these scaffolds have on cell behaviour and the potential application of nanomaterial science to this research.

4.1. Natural

Natural materials used as bioactive scaffolds include agarose, alginate, collagen, Hyaluronic acid and acellular cartilage matrix (Table.1). The potential for clinical use of these scaffold matrices is hampered however by poor mechanical strength and flexibility, in addition to a potential for disease transfer and immune system reactivity if allogenically sourced. Their biochemical make-up leaves them prone also to host-related degradations.

Agarose: a linear polysaccharide consisting of repeating units of agarobiose, derived from Asian seaweeds and capable of supporting the chondrogenic phenotype. Its ability to form hydrogels allows it to encapsulate chondrocytes providing a 3D matrix for their growth and development. A group in Germany performed allograft transplants of chondrocytes in agarose gel into osteochondral defects in the knee of rabbits. There we no graft versus host rejections, and after 18 months, 47% of grafts had morphologically stable hyaline –like cartilage (117).

Ref.FabricationMethodCells SourceOutcome
Natural
ALGINATE
(49)Chondrocytes in suspension with 2% sodium-alginateIn vivo; 500µl of suspension injected subcutaneously into dorsa of nude mice. Calcium chloride then injected into this area to stimulate cross linking of the scaffold. Cartilage harvested from 14 to 38 weeksHuman nasal septal chondrocytesGross analysis showed that 14/15 constructs resembled native human cartilage. 6 of the explants had histologically homogenous resemblance to native cartilage. The neo-constructs stained positively for Col II.
(97)3D alginate scaffold prepared by freeze dryingIn vitro; Cells were cultured in the alginate for 1-4 weeks in a bioreactorPorcine articular chondrocytesRT-PCR analysis showed the cells maintained their differentiated phenotype for up to 4 weeks. The cell also proliferated increasing from 5 x 105 cells to 3 x 107.
(151)3D alginate gelsIn vitro; cell/gel constructs were cultured for 0, 6, 12, 18 and 24 daysHuman MSCsResults of qRT-PCR analysis provided a temporal analysis for marker expression during chondrogenesis. Stage I (days 0–6): Col I and VI, Sox 4, and BMP-2. Stage II (days 6–12): Cartilage oligomeric matrix protein, HAPLN1, Col XI, and Sox 9. Stage III (days 12 -18): Matrilin 3, Ihh, Hbx 7, chondroadherin, and WNT 11. Stage IV (days 18–24): aggrecan, collagen IX, II, and X, osteocalcin, fibromodulin, PTHrP and alkaline phosphatise.
(91)Alginate gel layerIn vitro; To evaluate the effect of low-intensity ultrasound (LIUS) on cell viability during chondrogenic differentiationHuman MSCsWhen the cell/alginate construct was cultured with TGF-β1, cell viability decreased. However, addition of LIUS enhanced viability and inhibited apoptosis under the same conditions. Demonstrated by the expression profiles of apoptosis genes, p53, bax and bcl-2.
(30)HydrogelIn vitro; Chondrocytes were seeded onto alginate after 1, 2 and 3 passages in a monolayer. Human nasal septal chondrocytesAlginate stimulated GAG and Col I deposition supporting the chondrocytic phenotype. Results did not also support other research showing that culture with alginate beads can redifferentiate cells.
CHITOSAN
(114)Fibrous scaffold vs. spongeIn vitro; constructs analysed 3 days, 10 days and 21 days after cell seedingMouse BMSC lineAt 10 and 21 days the cells were embedded but did not aggregate, with fibrous scaffolds containing more ECM. The cells had a round morphology. Histology revealed cell and ECM distribution was not homogenous. mRNA expression for Col II was 3 times greater for the fibrous scaffold compared with the sponge at 21 days
(22)Chitosan scaffold and Chitosan microspheresIn vitro; Scaffold and microspheres used asTGF-β1 carrier to see the effect of this growth factor on chondrogenic potentialRabbit articular chondrocytesEncapsulation efficiency of TGF-β1 was 90.1%. TGF-β1 was released from chitosan in a multiphase fashion. TGF-β1 loaded microspheres significantly improved cell proliferation rate and Col II production, compared with controls with no microspheres or controlled TGF-β1 release.
(61)Chitosan scaffold synthesized via freeze drying In vitro; cells seeded on to chitosan of varying porosity; <10µm, 10-50µm and 70-120µm. Cultured for 28days in a rotating bioreactorPorcine articular chondrocytesChitosan scaffolds remained intact compar4ed with the positive control PGA. However cartilage specific DNA levels and GAG were lower in the Chitosan groups compared with PGA. Chitosan also had the largest pores, with more Chondrocytes, but on histological analysis, the composition of cartilage produced on PGA resembled the structure of native cartilage more than chitosan constructs.
COLLAGEN
(38)PLGA mesh and Collagen spongeIn vitro; hybrid disks of PLGA/Collagen scaffold with different structures. In vivo; week old cultured constructs implanted into dorsa of athymic nude mice and harvested after 2, 4 and 8 weeksBovine articular chondrocytesHomogenous cell distribution with natural chondrocyte morphology. Abundant ECM production. Levels of GAG and Collagen II DNA, and aggrecan mRNA increased on the scaffolds with more collagen (semi, collagen on one side of PLGA and sandwich, collagen on both sides). Semi and sandwich compared with natural articular cartilage, had a Young’s modulus of 54.8% and 49.3% respectively. 76.8% and 62.7% in stiffness.
(156)HydrogelIn vivo: comparison of collagen hydrogel and collagen-alginate hydrogel. Gel injected subcutaneously into rabbit backs.BM-MSCHomogenous distribution of cells with chondrocyte characteristics demonstrated the chondrogenic differentiation of BM-MSCs. Both collagen hydrogel and collagen alginate hydrogel may induce chondrogenesis. Expression profile of cartilage specific genes differed between collagen hydrogel and collagen alginate, indicating that induction of chondrogenesis is materials dependent.
(153)3D collagen spongedIn vitro; Cells seeded onto collagen sponges and cultured in either standard or serum free culture conditions for 1, 2 and 4 weeksBovine articular chondrocytesOverall chondrogenesis in serum free culture (Nutridoma replacement) was equivalent or better than control cultures in serum. Insulin-transferrinselenium (ITS+3) serum replacement cultures were poor due to decreased cell viability. The porous 3D collagen sponges were able to maintain chondrocyte viability, shape, and synthetic activity with evidence from quantitative assays for cartilage-specific gene expression and biochemical measures of chondrogenesis.
FIBRIN
(83)Fibrin gelIn vivo: ACI on 30 patients using minimally invasive injection techniques. Mix of fibrin gel and chondrocytes. Autologous adult chondrocytesPatients evaluated 24 months post operatively using the Cincinnati knee ligament rating scores, for which 10 patients had excellent result, 17 with good results, two fair and one poor result. Further arthroscopy in 10 patients demonstrated good fill and integration in grafted areas.
(131)PLGA/Fibrin hybrid scaffold In vitro; PLGA scaffold soaked in chondrocyte-fibrin suspension (polymerized by thrombin CaCl2 solution), Constructs were cultured for a maximum of 21 days.Rabbit articular chondrocytesCell proliferation increased steadily until day 14, but declined by day 21. Cartilage formation evident at day 14, confirmed by the presence of cartilaginous cells embedded in basophilic ECM filled lacunae. Proteoglycan and GAG presence was confirmed. Suppression of the cart dedifferentiation marker Col 1 observed after 2 and 3 weeks in culture. sGAG production greater in fibrin/PLGA compared with PLGA control.
(28)PLGA/Fibrin hybrid scaffoldIn vivo; PLGA scaffold soaked in chondrocyte-fibrin suspension (polymerized by thrombin CaCl2 solution) and constructs implanted subcutaneously into dorsum of nude mice for 4 weeks after culture for 3 weeks. Analysis performed at 1, 2 and 4 weeks. Rabbit articular chondrocytesConstructs maintained their shape and there was no significant difference between fibrin/PLGA and control PLGA. All exhibited smooth cartilage like properties 1, 2 and 4 weeks after implantation. Presence of proteoglycans and GAG was confirmed. The constructs were also strongly positive for Col II. Notably, sGAG production was greater on fibrin/PLGA scaffold than the control. Overall, both fibrin/PLGA and PLA showed comparable potential in sustaining the chondrogenic phenotype.
HYALURONIC ACID(HA)
(33)***Hyaff®-11, biodegradable polymer, nonwoven meshIn vitro; Chondrocytes were harvested from OA patients and seeded onto Hyaff®. Constructs remained in culture for 28 days, analysed on day 0, 7, 14, 21 and 28. Human Autologous chondrocytesViability and proliferation of OA chondrocytes similar to cells from normal subjects. Immunohistochemistry showed no signs of ageing or degeneration in cartilage produced by OA cells. The experimental groups and controls both had significantly raised Col II, Sox 9 and aggrecan. Suggests OA cells benefit from the HA rich environment.
(154)Hydrogel (in vitro), beads(in vivo)In vitro and In vivo; implanted into nude mice. Constructs were cultured in vitro for 2 weeks prior to implantation. Constructs remained implanted for 2 weeks.Human MSCBoth in vitro and in vivo cultures of MSC-laden HA hydrogels enabled chondrogenesis. This was measured by the early gene expression and production of cartilage specific matrix proteins (aggrecan, Col II). HA hydrogels were compared to relatively inert poly(ethylene glycol) (PEG) hydrogels, and showed enhanced expression of cartilage specific markers
(107)HA immobilized on surface of PLGA scaffoldIn vitro; biodegradable macroporous PLGA scaffolds chemically conjugated to the surface exposed amine groups of the PLGA. Incubation times varied for each assay.Bovine articular chondrocytesEnhanced cellular attachment was observed compared with PLGA controls. GAG and total Col synthesis was significantly increased for HA/PLGA compared to the control. The HA/PLGA constructs exhibited morphological characteristics of cartilage and had cartilage specific Col II expression.
Synthetic
PLGA- Poly(lactic-co-glycolic) acid
(107)PLGA scaffoldsIn vivo: PLGA scaffolds were seeded with AD-MSC, cultured in TGFβ1 containing medium for 3 weeks, prior to implantation in the subcutaneous pockets of nude mice for 8 weeks. Human AD-MSCRT-PCR demonstrated the increased expression profiles of chondrospecific marker mRNA, compared with control samples after 3 weeks in vitro and 8 weeks in vivo.
(150)HA modified porous PLGA scaffoldIn vitro; cells seeded onto HA/PLGA scaffolds and cultured for a total of 5 days.Human AD-MSCThe AD-MSC cultured in HA coated wells showed enhanced expression of cartilage specific mRNA. HA-modified PLGA did not affect cell adherence and viability, but did enhance gene expression after 1, 3 and 5 days in culture. GAG and Col I production enhanced after 4 weeks in culture compared with PLGA control.
(10)PLGA scaffoldsIn vivo; cells were pre-cultured on poly-HEMA coated dish, then seeded onto PLGA. The construct was implanted into the subcutaneous pockets of nude mice for 16 weeks.ChondrocytesMacroscopic signs of neo cartilage formation appeared at 8 weeks, and completed by 16 weeks. All constructs showed viable chondrocytes with normal lacunae and ECM. They stained positively for Col II. Control was a cell-free scaffold implanted into the other side of the dorsum on the same mouse.
(75)PLGA microspheresIn vivo; PLGA microsphere seeded with rabbit Chondrocytes injected subcutaneously into dorsa of athymic female miceAutologous rabbit ChondrocytesThe PLGA microsphere permitted cell adhesion. 4 and 9 weeks post-implantation there was macroscopic and histological evidence of cartilage formation on the seeded PLGA microsphere compared with nothing on the PLGA and chondrocyte controls.
(142)PLGA porous scaffold disksIn vivo; MSC seeded PLGA scaffold disks implanted into 36 week old Japanese white rabbits. Constructs were harvested after 4 and 12 weeks.Rabbit BM-MSC
Engineered cartilage from autologous BM-MSC and PLGA scaffold filled the defects in the rabbit knees. The constructs were macroscopically and histologically similar to hyaline cartilage at 12 weeks post transplantation.
PCL- Poly(carprolactone)
(82)3 porous PCL scaffold types investigated (1)PCL/Pluronic F127, (2)PCL collagen and (3)PCL/Pluronic F127/collagen, in addition to (4) PCL onlyIn vitro; 3 porous PCL scaffold modifications investigated (1) PCL/Pluronic F127, (2) PCL collagen and (3) PCL/Pluronic F127/collagen, in addition to (4) PCL only. Cultured for 3 weeks.Human BM-MSCThe 3 surface treated scaffolds had higher chondrospecific DNA content than the PCL only. GAG concentrations were also higher than in the PCL only, and RT-PCR showed that Sox 9 and Col IIA1 were remarkably elevated in the modified PCLs. Notably, Col IA1 and ColI0A1 mRNA levels were lower in the three modified scaffolds than in the PCL, suggestion prevention of the dedifferentiated phenotype.
(95)Electrospun 3D nanofibrous scaffoldIn vitro; MSC seeded onto pre-fabricated nanofibrous scaffold for 21 daysHuman BM-MSCHistological analysis was congruent with cartilage formation when cells were grown in medium containing TGFβ1. The cartilage specific gene profile (Aggrecan, Col II and Col X) was low, but improved significantly in chondrogenic medium with TGFβ1. Col X levels were paradoxically down regulated. There was positive immunohistochemistry for cartilage specific ECM molecules.
PGA-Polyglycolic acid
(158)Porous PGA and high density polyethylene composite scaffoldIn vivo; High-density polyethylene carved into cylindrical rods (internal support), with non-woven PGA sheets wrapped around the rods to form the scaffold. Implanted subcutaneously into nude mice.Porcine BMSC8 weeks post-implantation the constructs had formed mature cartilage with an abundant deposition of ECM on SEM. The experimental groups showed a positive histological likeness to cartilage with large number of lacunae and good expression of Col II.
(157)PGA-HA composite scaffoldIn vivo; MSC were seeded onto the PGA-HA and co-cultured for 72hours. There were then implanted into full thickness cartilage defects in the intercondylar fossa of rabbit femurs. Constructs were then harvested after 16 or 32 weeks of surgery.Rabbit MSCGrossly, the constructs demonstrated hyaline cartilage formation and at 16 weeks, there appeared to be integration with surrounding normal cartilage and subchondral bone. At 32 weeks there was no sign of degradation of the neoconstruct.
(160)PGA vs. PLA bio-resorbable nonwoven scaffoldsIn vivo; Cells seeded onto scaffolds and cultured for 7 days in serum free media, before implantation into subcutaneous nude mice for 6 and 12 weeks Human articular chondrocytesAggrecan synthesis always higher in the PGA groups. mRNA gene expression for Col II significantly higher in the PGA groups after 6 and 12 weeks. Expression of Col X and cartilage oligomeric matrix protein increased on both scaffolds.
PEG- Poly (ethylene glycol)
(125)PEG-peptide copolymer gelsIn vitro; RGD and KLER sequences chosen as motifs to modify PEG gels. (KLER is a binding site from decorin protein, known to bind strongly to Col II, RGD promotes survival of encapsulated cells). Cells were encapsulated in the PEG peptide gel and cultured for 6weeksHuman MSCsAfter 14 days, cells in RGD and KLER functionalized gels produced 2.5 times as much GAG as those only containing RGD. hMSCs also produced 27x as much hydroxyproline (a major component of collagen) than scrambled sequence gel controls. Col II was more prominent in KLER gels on immunostaining and RT-PCR analysis demonstrated higher levels of Col II and aggrecan synthesis.
(111)HydrogelIn vitro; cells were encapsulated in the PEG hydrogel and allowed to free swell for 24hrs. Bovine temporomand-ibular chondrocytesCondylar chondrocyte viability was maintained within the constructs during cell culture. RTPCR analysis showed the expression of cartilage specific markers, namely Col II, aggrecan and Col I was maintained

Table 1.

Summary of in vitro and in vivo studies that have used various scaffolds to engineer cartilage (2005-2010). Abbreviations: AD-MSC, adipose derived mesenchymal stem cells. BMSC, bone marrow stromal cells. BM-MSC, bone marrow derived mesenchymal stem cells. Col, collagen. ECM, extra cellular matrix. GAG, glycosaminoglycan. Hbx, homeobox. Ihh, Indian hedgehog. OA, osteoarthritis. PTHrp, parathyroid hormone replacement hormone. RT-PCR, real-time polymerase chain reaction. SEM, scanning electron microscopy. sGAG, sulphated glycosaminoglycan

Alginate: derived from brown marine algae and is consists of 1, 4-linked β-D-mannuronnic and α-L-guluronic residues, which are soluble in aqueous solutions. Cross-linking with bivalent cations such as Ba2+ or Ca2+ allows it to form stable gels.

Chitin: a polysaccharide based analogue of GAG found in the exoskeleton of arthropods. Relatively unexplored bioactive scaffold for tissue engineering, perhaps because it is degraded in vivo by lysozyme; an enzyme found in many human bodily fluids.

Collagen 1 and II: As the principle ECM components of cartilage, seeded chondrocytes can bind using inherent cell-surface receptors and use standard signalling pathways to regulate proliferation and growth. Can be fabricated as a sponge, foam, or gel, but like chitin, is subject to enzymatic breakdown.

Fibrin: Can be derived from autologous blood samples, and has a comprehensive history of biocompatibility in its clinical use as a wound adhesive. Chondrocytes have integrins that can bind directly to fibrin, much like with collagen.

Gelatin: A porous substance derived from hydrolysis of collagen. Its application as a scaffold for cartilage tissue engineering is relatively unchartered.

Hyaluronic Acid: a non-sulphated GAG, found in abundantly within the cartilaginous ECM. It is crucial for maintaining the biophysical properties of the cartilage ECM for optimum chondrocyte growth and proliferation.

4.2. Synthetic

The main aim of biomimetic materials (synthetic biomaterials) is to generate 3D scaffolds that support essential cell functions in addition to mimicking the biomechanical properties of host tissues, whilst avoiding host immune responses (Table.1). These are two characteristics more difficult to find in natural scaffold alternatives. When considering clinical applications, susceptibility to vascular invasion is a key consideration and there is continued debate between groups about the need for biodegradation. Persistence and stability have been the focal aims for tissue engineering cartilage with the mechanical and biochemical properties of synthetic materials being more amenable to modification than natural scaffolds.

Polyhydroxyacids: polyhydroxyacids such as PLLA [poly (L-lactic acid)], PCL [poly (L-lactide-ε-caprolactone)] and PGA [poly (glycolic acid)] have been well studied as potential cartilage scaffold matrices, where they are easily extruded into fibrous or open lattice sponges. PGA is reportedly highly biodegradable (5 weeks); PLLA can stay in vivo up to 3 years. PCL and PGA used to fabricate ear templates for tissue engineering auricular cartilage (133).

Elastomeric polyurethanes: Well documented history of use in a variety of biomedical instruments, ranging from urinary and vascular catheters to intra-aortic balloons and mammary implants. Can be fabricated in a biodegradable form, and have been shown to support chondrocyte attachment and growth.

PEG [poly (ethylene glycol)]: FDA (Food and Drugs Administration) approved, and extensive research into its ability to promote chondrogenesis

4.3. Regulatory influence of scaffolds on cell behaviour

It is widely appreciated that soluble biochemicals such as cytokines, growth factors and chemokines affect the growth and development of all tissues including cartilage. Transforming growth factor beta (TGFβ) and bone morphogenic proteins (BMPs) have been evidenced as highly potent stimulators of cartilage tissue generation (78; 96; 118). In addition to such signalling mechanisms, ECM proteins such as collagens, glycosaminoglycans and proteoglycans exert an array of instructions on cells via transmembrane receptors that affect expression and therefore, cell behaviour. Much of this instruction will crosstalk with growth factor signalling (37; 44). Additional studies have also shown chondrocytes to be particularly receptive to mechanical loading, with this parameter having been evidenced as a crucial factor in the chondrogenic differentiation of MSCs during critical cartilage development. Repetition of these loads and varying the duration and force of the load has positive effects on the structural organization of cartilage ECM (7; 62). The effects of mechanobiology on chondrogenesis have been discussed in detail in the section on stem cells.

In recent times, tissue engineering research had broadened its horizons to understand the effect of scaffold physical properties on cell behaviour. Properties considered include; roughness (88; 89; 147), micro and nanotopography (reviewed in (132)), porosity (155) and surface energy (80; 147). The stiffness of the substrate (scaffold matrix) has been demonstrated as a crucial regulator of stem cell behaviour (15; 48; 52; 119). It is thought that the stiffness or elasticity of a matrix can act as a ‘passive’ cue for cell processes via a phenomenon known as mechano-transduction. This is a method by which cells convert mechanical stimuli into a chemical response, thus affecting their own behaviour. For detailed reviews see (2; 54). Cells bind to the matrix using integrins. The intra cellular domain connects to the actin and myosin (contractile) cytoskeleton of the cell, and the extracellular domain to the biomaterial. When cells are bound, they apply mechanical forces onto the matrix using their contractile cytoskeleton. Integrins cluster which in turn recruits structural and signalling proteins at the site of contact between cell and matrix, known as a focal adhesion. If a matrix is relatively hard, there is more resistance to the forces applied by the cellular cytoskeleton. This results in a more organized cytoskeleton, more integrin clustering and thus focal adhesions that are greater in maturity. Comparatively, if cells are seeded onto a soft matrix, there is little resistance to counterbalance the cell forces, therefore reduced development of the actin-myosin cytoskeleton. This phenomenon is fundamental considering that changes in cytoskeletal organization affect signalling, thereby translating mechanical processes into chemical responses.

So how can this trend be used in cartilage tissue engineering technology? Let us consider the application of stem cells in tissue engineering cartilage. Stem cells extracted from human or animal sources are frequently expanded in culture. Culturing stem cells on traditional tissue culture plastic could result in preconditioning of the cells in accordance with the stiffness of the plate (51; 52). Depending on the experimental aims, it may be wiser to culture and expand on softer substrates with stiffness comparable to that of native tissue. However conflicting data has shown that stiffer substrates increase the rate of proliferation, whereas soft substrates promote the dedifferentiation of cells (15).This suggests the stiffness of the material used for cartilage tissue engineering is an important parameter not just in terms of mechanical support but also in terms of propagating chondrocyte growth and matrix deposition.

4.4. Nanomaterials

Cell coverage over a matrix layer is directly correlated to the spread of microscale ECM proteins over its surface, irrespective of the geometric patterning of such proteins (93). This theory applies at the microscale level of tissue engineering, but at the nanoscale, there is increasing evidence to indicate that cells are able to alter their behaviour differentially in response to changes in nanotopographical surfaces. These changes can be cytoskeletal or a change in morphology, focal adhesions, motility, gene expression and differentiation. Much like the mechano-transduction discussed earlier, there is support for some sort of topography-dependent transduction that communicates independent of chemical signalling from ECM molecules (35). Studies have since demonstrated that this cellular response is heavily related to the pattern and spacing of adhesive ligands (36;41;79;148).

In light of the revelation that nanotopography plays a major role in the governance of cell-matrix interactions, many physical and chemical methods have been developed to engineer geometrically defined nanopatterns on biocompatible scaffolds. Crude methods of acid treatment (85), bonding with calcium cations (63), and coating with nanoparticles (reviewed in (126)) allowed scientists to introduce nanofeatures into the surface topography of scaffolds. Surface modifications with Lanthanum phosphate (LaPO4) nanoparticles increased osteoblast adhesion to traditional bioceramics; Hydroxyapatite and Tricalcium phosphate (53). Likewise with chondrocytes, the levels of adhesion increased on 70%/30% (wt) PLGA/titanium composite scaffolds manufactured to have a nanosurface (76). However the advancement of nanoscience allows more precise pattering of various nanofeatures to further affect cell behaviour. Nanofeatures now come in many forms ranging from nanopits and grooves, to nanopillars, nanodots and traditional nanoparticles. The pattern in which they are arranged is also on the nanoscale. The latest techniques used for nanosurface pattering (reviewed in (132)) include photolithography, electron beam lithography (40), Dip-pen lithography (71) and imprint lithography.

Nanopatterning to mimic the surface density and arrangement of integrin-binding epitopes as seen in the ECM has been a challenge not yet beaten. Studies have shown that integrin mediated signalling operates with a minimum surface density however; the exact spatial organization of these ligands in vivo has not been elucidated. The nearest estimate has come from a group that developed a block-copolymer micelle nanolithography technique to label surfaces with hexagonal arrays of gold nanodots coated with one RGD peptide (found in adhesive glycoproteins such as fibronectin and vitronectin) (26;27). Upon cell seeding they found that only 28nm and 58nm spacing between the nanodots would allow adequate clustering of integrins, which are approximately 8-12nm in size (138). Additional studies on RGD-coated gold nanoparticles have shown that the velocity of migrating cells decreases with an increased particle density, with a peak velocity at circa 120nm, suggesting the boost in particle density increased levels of adhesion (6;65). Interestingly enough, research has also shown the MSC osteoblastic differentiation can be hampered by regularly arranged hexagonal nanopits arrays compared with arrays with a slight irregularity (39). Similar results were found by Biggs et al 2007, where highly ordered nanopits resulted in decreased formation and length of focal adhesions, compared with controlled disorder increasing focal adhesion formation and size (13).

With more research being conducted into the in vitro effects of surface nanopatterning on cell behaviour, there are implications for cartilage tissue engineering research. Data shows that nanostructured PLGA can accelerate chondrocyte attachment, growth and proliferation in addition to improving ECM production (76). In our lab, chondrocytes seeded onto nanocomposite polymer POSS-PCU (UCL nanoBio™) have a faster rate of proliferation compared with controls lacking the nano modification (unpublished data). And though the current research into nanomaterials and cartilage tissue engineering is just evolving, there are many lessons to be learnt from bone (80;81), skin (31) and vascular (100;108) tissue engineering research.

Advertisement

5. Bioreactors

Bioreactors are devices in which biological and/or biochemical processes develop under controlled and monitored environmental and operating conditions (104). It is the exceptional control over environmental conditions that makes bioreactor use particularly pertinent in tissue engineering research where specific factors need to be controlled in order to optimise tissue growth. Bioreactors can maintain physiological boundaries at desired levels, enhance nutrient and waste transport rates, and provide specific stimuli to promote optimum growth.

The use of bioreactors has provided a promising method for tackling some causes for poor research outcomes in tissue engineering practice. Restricted, unspecific, or impermanent cell differentiation and poor tissue formation/ remodelling in cartilage tissue engineering largely results from a lack of correct physical stimulation in vitro (86). For example, mechanotransduction, the transduction of mechanical stresses into biochemical signals, affects chondrocyte function. Modifying the mechanical stressors applied to cells in vitro may therefore improve the quality of tissue constructs produced. In early parts of this chapter, the effect of cyclical loading, especially within the articular region has been shown to improve the ECM content of constructs, and therefore the overall construct viability. Mimicking some of these forces in bioreactor systems could also dramatically improve tissue growth. Studies which evidenced the effect of adaptive physical stimulation on mechanotransduction, led to the development of bioreactor devices that transmit forces including shear stress, hydrostatic pressure and compression to articular cartilage in vitro (129).

5.1. Mechanical forces

Key to tissue engineering in the joint region specifically is the use of exogenous mechanical forces to simulate loading forces (exerted during daily movement and exercise), which in turn increases the metabolic activity of and ECM production by chondrocytes. Shear stress, compressive forces, tensile forces and hydrostatic pressures are all parameters that can be modulated to influence the quality of cartilaginous constructs engineered. The effect of these mechanical forces on chondrogenesis, have been described earlier in the chapter. We will examine briefly the bioreactors that have been used to study the effects of shear stress however, the different types of bioreactors available for exerting other forces are expertly reviewed in Schulz RM 2007.

Figure 4.

Experimental appraoches to bioreactor tissue engineering. Representation of static, stirred and rotating vessels. Fc, Fd, Fg refer to centrifugal, drag and net forces respectively. (Adapted from Vunjak-Novakovic G 1999 (145))

The easiest method for examining the effects of shear stress in bioreactor systems is by placing constructs in culture either on a petri dish or in a dynamic or orbital shaker (56;57). Other methods developed have included using spinner flasks or vessels with magnetic stirrers (14;18;136). Extensive study in the nineties looked at the application of shear stress by comparing static and orbital shakers, stirred vessels with rotating vessels (Fig.4) (145). Rotating vessels are more advanced systems where constructs float freely within culture medium, whilst the whole vessel rotates around a central axis at a constants speed. Chondrocytes were seeded onto 97% porous scaffold discs and cultured in the aforementioned vessels for 8 weeks. Results showed that freely cultured constructs were larger than those cultured in static or stirred vessels. They formed cartilaginous ECM with the greatest concentration of GAGs and collagen. Their mechanical properties where also shown to be superior.

5.2. Oxygen tension

Optimizing O2 tension within culture systems is an area of great importance in bioreactor design. O2 is the partial pressure of oxygen dissolved in a liquid such as blood. Cells in culture require nutrients and oxygen to proliferate and this is usually achieved through mass transport (net movement of mass from one location to another). When oxygen and nutrients are limiting factors, larger grafts tend to contain a hypoxic, necrotic centre, surrounded by a rim of viable cells (Martin I 2004). In a tissue graft, the density of cells may be higher than the distance oxygen can freely diffuse across by mass transport to provide sufficient oxygen for the inner cells; therefore they are starved of oxygen. Limited O2 diffusion can also affect the spatial distribution of cells and as the O2 concentration gradient decreases from the surface of the tissue compartment to its centre (34). In humans, this problem is solved by the circulatory system and thus nutrients are provided to all cells via a complex network of vessels, slowly decreasing in size the deeper into tissues they enter. It is the proximity of capillaries to somatic (body) cells that allows their mass transfer requirements to be met (105). So how is this problem solved in tissue engineering practice? The introduction of simple stirred flask bioreactors enables the mixing of oxygen and nutrients throughout the medium. So not only does it provide a shear stress which is known to be beneficial for chondrocyte growth and proliferation, but it also reduces the concentration boundary layer of oxygen at the construct surface (14;18;104). In a static culture environment, oxygen would diffuse into cells and carbon dioxide out. The medium in closest proximity to the cells would have a steadily decreasing O2 tension with a conversely increasing CO2 tension. This in turn limits the overall rate diffusion as O2 moves from areas of high tension to areas of low tension. Thus if the culture medium is not circulated or replenished the rate of diffusion will decrease and eventually cease at the point where there is no longer a concentration gradient, leading to cell death.

Studies have also used bioreactors to investigate the effect of different partial pressures of O2 and pH levels on gene and protein expression, as well as the metabolic activity of chondrocytes. Results showed chondrocyte sensitivity to acidic conditions where reduced expression of Coll Type 1, SOX9 and VEGF (vascular endothelial growth factor) were observed. Conversely in hypoxic conditions, VEGF levels were found to be higher, with a pH dependent reduction in Coll Type 1 (43). Culture in bioreactors at low oxygen tension increases the production and retention of glycosaminoglycan (GAG) within the cartilage matrix without affecting chondrocyte proliferation or collagen deposition which typically would requires higher partial pressures of O2 (123). These studies highlight the twofold applications of bioreactors, in maximizing cell growth and tissue generation for clinical use and in research and development to investigate the effect of different biological factors on cell growth.

5.3. Growth factors

It has also bee suggested that bioreactors provide suitable environments to add growth stimulating factors to constructs to improve chondrogenesis. For example, transducing human MSCs with an adenoviral vector containing SOX9 and subjecting the construct to mechanical stimulation could increase GAG synthesis (90). Growth factor application of BMP-2, IGF-1 and TGF-β1 in a bioreactor system can increase the compressive and tensile biomechanical properties of engineered tissue (50). The efficiency of chondrocyte proliferation from low initial seeding densities can also be enhanced by adding various growth factor combinations in to automated bioreactors systems (55). Chitosan scaffolds were used to engineer articular cartilage with the aid of a chondrogenic differentiation factor, BMP-6. Results showed that proliferated cells contained a higher value of GAG, Coll type II and DNA indicating improved chondrogenesis (1). Alternatively, inhibiting the expression of some factors, namely interleukin 6, has also been investigated with the aim of improving tissue growth in bioreactors. In 2010, Wang P et al demonstrated how high levels of interleukin – 6 have been found in osteoarthritic cartilage and suggested that inhibiting this expression may improve cartilage construct culturing in bioreactors (146).

Advertisement

6. Challenges for the clinical application of regenerated cartilage

Over the past two decades the amount of data on cartilage tissue engineering strategies has risen exponentially. There is now a plethora of exciting in vitro data evaluating chondrocyte/MSC seeded biomaterial constructs. Perhaps one of the most iconic studies in cartilage tissue engineering research was produced by Cao and Vacanti’s group in 1996, when they implanted an auricular shaped cartilaginous construct onto the back of a mouse (24). Even with all the advancements in stem cell and biomaterial technology, the invention of various bioreactor systems, little has progressed beyond this scientifically historic event. Most constructs fail to develop beyond immature, inflexible neocartilage that lacks the durability essential to most clinical applications.

There are a number of reasons for the stagnation in translation to clinical practice. Many of which have been discussed throughout the course of this chapter. On a cellular level, reasons for poor research outcomes could also include; (i)Regenerative cells being lost through leakage of the cell suspension (149), (ii) inflammatory cytokine, matrix metalloproteinase, nitric oxide mediated apoptosis and necrosis at the site of injury. These biochemical factors are released as part of the normal inflammatory and wound healing process, especially at the interface between host and repair tissue, which can also adversely affect biointegration of the neo tissue. The use of anti-apoptotic factors would be crucial in maintaining cell numbers but also in creating a favourable environment for biointegration (5). The poor migration capacity of chondrocytes could also be responsible for hampered infiltration of repair tissue into the host environment. The naturally slow rates of chondrocyte ECM production could slow down integration as well disparities in the organization of neocartilage matrix compared with the zonal arrangement of native cartilage tissue (69;84). Dedifferentiation of chondrogenic cells is another problem, and is likely responsible to the highly fibrotic nature of neocartilage produced, suggesting that over time, cells may have dedifferentiated into fibroblasts or incompletely differentiated into chondrocytes. Solutions would include seeding with cells that have been fully differentiated in vitro, but again there would be difficulties with motility, proliferation and shelf life.

In addition to the cell based scientific problems associated with cartilage engineering tissue research, ambiguous regulatory guidelines currently hamper the flow of development from laboratories to clinics and operating theatres. The EU regulation on Advanced Therapy Medicinal Products (ATMP), which includes tissue engineered constructs, is still in its infancy having only been formally established in December 2008. ATMP regulation aims to proved a coherent and tailored framework for tissue engineered products, however the nascent and fast growing nature of the tissue engineering field means that there is a constant threat of irrelevance over the guidelines developed under this regulation. Tissue engineering technology needs to reach a level of quality controlled and quality assured reproducibility to allow for not just clinical efficiency, but also commercial viability. Methods of stem cell differentiation, cell seeding, scaffold fabrication and bioreactor development/implementation all need to be governed by Good Manufacturing Practice (GMP). Additionally, methods of commercialization ought to be better established, to avoid uncertainty in the markets, improve regulatory approval and clinical uptake/use (103).

References

  1. 1. AkmanA. C.SedaT. R.GumusdereliogluM.NohutcuR. M.Bone morphogenetic protein-6-loaded chitosan scaffolds enhance the osteoblastic characteristics of MC3T3-E1 cells. Artif. Organs 3420106574
  2. 2. AlenghatF. J.IngberD. E.Mechanotransductionall.signalspoint.tocytoskeleton.matrixintegrinsSci. STKE. 2002: e6, 2002
  3. 3. AlhadlaqA.MaoJ. J.Tissue-engineered neogenesis of human-shaped mandibular condyle from rat mesenchymal stem cells. J Dent. Res 822003951956
  4. 4. AngeleP.SchumannD.AngeleM.et al.Cyclicmechanical.compressionenhances.chondrogenesisof.mesenchymalprogenitor.cellsin.tissueengineering.scaffoldsBiorheology 412004335346
  5. 5. ArcherC. W.RedmanS.KhanI.BishopJ.RichardsonK.Enhancing tissue integration in cartilage repair procedures. J Anat. 2092006481493
  6. 6. ArnoldM.Hirschfeld-WarnekenV. C.LohmullerT.et al.Induction of cell polarization and migration by a gradient of nanoscale variations in adhesive ligand spacing. Nano. Lett. 8200820632069
  7. 7. ArokoskiJ. P.JurvelinJ. S.VaatainenU.HelminenH. J.Normal and pathological adaptations of articular cartilage to joint loading. Scand. J Med. Sci. Sports 102000186198
  8. 8. AtalaA.Engineeringtissues.organscellsJ.TissueEng.MedRegen.83962007
  9. 9. AugelloA.KurthT. B.De BariC.Mesenchymalstem.cellsa.perspectivefrom.invitro.culturesto.invivo.migrationnichesEur. Cell Mater. 202010121133
  10. 10. BaekC. H.KoY. J.Characteristics of tissue-engineered cartilage on macroporous biodegradable PLGA scaffold. Laryngoscope 116200618291834
  11. 11. BarryF.BoyntonR. E.LiuB.MurphyJ. M.Chondrogenic differentiation of mesenchymal stem cells from bone marrow: differentiation-dependent gene expression of matrix components. Exp. Cell Res 2682001189200
  12. 12. BentleyG.BiantL. C.CarringtonR. W.etal. A.prospectiverandomised.comparisonof.autologouschondrocyte.implantationversus.mosaicplastyfor.osteochondraldefects.inthe.kneeJ Bone Joint Surg Br 852003223230
  13. 13. BiggsM. J.RichardsR. G.GadegaardN.WilkinsonC. D.DalbyM. J.Regulation of implant surface cell adhesion: characterization and quantification of S-phase primary osteoblast adhesions on biomimetic nanoscale substrates. J Orthop Res 252007273282
  14. 14. BouchetB. Y.ColonM.PolotskyA.ShikaniA. H.HungerfordD. S.FrondozaC. G.Betaintegrinexpression.byhuman.nasalchondrocytes.inmicrocarrier.spinnerculture.J Biomed Mater Res 522000716724
  15. 15. BreulsR. G.JiyaT. U.SmitT. H.Scaffold stiffness influences cell behavior: opportunities for skeletal tissue engineering. Open. Orthop J 22008103109
  16. 16. BrittbergM.LindahlA.NilssonA.OhlssonC.IsakssonO.PetersonL.Treatment of deep cartilage defects in the knee with autologous chondrocyte transplantation. N. Engl. J Med. 3311994889895
  17. 17. BrochhausenC.LehmannM.ZehbeR.etal. [.Tissue engineering of cartilage and bone : growth factors and signaling molecules]. Orthopade 38200910531062
  18. 18. BrownA. N.KimB. S.AlsbergE.MooneyD. J.Combining chondrocytes and smooth muscle cells to engineer hybrid soft tissue constructs. Tissue Eng 62000297305
  19. 19. BrunP.DickinsonS. C.ZavanB.CortivoR.HollanderA. P.AbatangeloG.Characteristics of repair tissue in second-look and third-look biopsies from patients treated with engineered cartilage: relationship to symptomatology and time after implantation. Arthritis Res Ther. 10: R132, 2008
  20. 20. BuckwalterJ. A.MankinH. J.Articularcartilage.tissuedesign.chondrocyte-matrixinteractions.Instr. Course Lect. 471998477486
  21. 21. BuckwalterJ. A.MankinH. J.Articularcartilage.degenerationosteoarthritisrepair.regenerationtransplantationInstr.CourseLect. 4.4875041998
  22. 22. CaiD. Z.ZengC.QuanD. P.et al.Biodegradable chitosan scaffolds containing microspheres as carriers for controlled transforming growth factor-beta1 delivery for cartilage tissue engineering. Chin Med. J (Engl. ) 1202007197203
  23. 23. CampbellJ. J.LeeD. A.BaderD. L.Dynamic compressive strain influences chondrogenic gene expression in human mesenchymal stem cells. Biorheology 432006455470
  24. 24. CaoY.VacantiJ. P.PaigeK. T.UptonJ.VacantiC. A.Transplantation of chondrocytes utilizing a polymer-cell construct to produce tissue-engineered cartilage in the shape of a human ear. Plast. Reconstr. Surg 1001997297302
  25. 25. CarlbergA. L.PucciB.RallapalliR.TuanR. S.HallD. J.Efficient chondrogenic differentiation of mesenchymal cells in micromass culture by retroviral gene transfer of BMP-2. Differentiation 672001128138
  26. 26. Cavalcanti-AdamE. A.MicouletA.BlummelJ.AuernheimerJ.KesslerH.SpatzJ. P.Lateral spacing of integrin ligands influences cell spreading and focal adhesion assembly. Eur. J Cell Biol. 852006219224
  27. 27. Cavalcanti-AdamE. A.TomakidiP.BezlerM.SpatzJ. P.Geometric organization of the extracellular matrix in the control of integrin-mediated adhesion and cell function in osteoblasts. Prog. Orthod. 62005232237
  28. 28. CavalloC.DesandoG.FacchiniA.GrigoloB.Chondrocytes from patients with osteoarthritis express typical extracellular matrix molecules once grown onto a three-dimensional hyaluronan-based scaffold. J Biomed Mater Res A 9320108695
  29. 29. ChenW. H.LaiM. T.WuA. T.et al.In vitro stage-specific chondrogenesis of mesenchymal stem cells committed to chondrocytes. Arthritis Rheum. 602009450459
  30. 30. ChiaS. H.HomiczM. R.SchumacherB. L.et al.Characterization of human nasal septal chondrocytes cultured in alginate. J Am. Coll. Surg 2002005691704
  31. 31. ChongE. J.PhanT. T.LimI. J.et al.Evaluation of electrospun PCL/gelatin nanofibrous scaffold for wound healing and layered dermal reconstitution. Acta Biomater. 32007321330
  32. 32. ChungC.BurdickJ. A.Engineeringcartilage.tissueAdv.DrugDeliv.Rev6.2432622008
  33. 33. ChungC.BurdickJ. A.Influence of three-dimensional hyaluronic acid microenvironments on mesenchymal stem cell chondrogenesis. Tissue Eng Part A 152009243254
  34. 34. CurcioE.MacchiariniP.De B. L.Oxygen mass transfer in a human tissue-engineered trachea. Biomaterials 31201051315136
  35. 35. CurtisA. S.DalbyM.GadegaardN.Cell signaling arising from nanotopography: implications for nanomedical devices. Nanomedicine. (Lond) 120066772
  36. 36. CurtisA. S.GadegaardN.DalbyM. J.RiehleM. O.WilkinsonC. D.AitchisonG.Cells react to nanoscale order and symmetry in their surroundings. IEEE Trans. Nanobioscience. 320046165
  37. 37. CzyzJ.WobusA.Embryonic stem cell differentiation: the role of extracellular factors. Differentiation 682001167174
  38. 38. DaiW.KawazoeN.LinX.DongJ.ChenG.The influence of structural design of PLGA/collagen hybrid scaffolds in cartilage tissue engineering. Biomaterials 31201021412152
  39. 39. DalbyM. J.GadegaardN.TareR.et al.The control of human mesenchymal cell differentiation using nanoscale symmetry and disorder. Nat. Mater 620079971003
  40. 40. DalbyM. J.GiannarasD.RiehleM. O.GadegaardN.AffrossmanS.CurtisA. S.Rapid fibroblast adhesion to 27nm high polymer demixed nano-topography. Biomaterials 2520047783
  41. 41. DalbyM. J.RiehleM. O.SutherlandD. S.AgheliH.CurtisA. S.Use of nanotopography to study mechanotransduction in fibroblasts--methods and perspectives. Eur. J Cell Biol. 832004159169
  42. 42. DarlingE. M.AthanasiouK. A.Rapid phenotypic changes in passaged articular chondrocyte subpopulations. J Orthop Res 232005425432
  43. 43. DasR. H.van OschG. J.KreuknietM.OostraJ.WeinansH.JahrH.Effects of individual control of pH and hypoxia in chondrocyte culture. J Orthop Res 282010537545
  44. 44. DattaN.HoltorfH. L.SikavitsasV. I.JansenJ. A.MikosA. G.Effect of bone extracellular matrix synthesized in vitro on the osteoblastic differentiation of marrow stromal cells. Biomaterials 262005971977
  45. 45. DavidsonD.BlancA.FilionD.et al.Fibroblast growth factor (FGF) 18 signals through FGF receptor 3 to promote chondrogenesis. J Biol. Chem. 28020052050920515
  46. 46. De LiseA. M.FischerL.TuanR. S.Cellular interactions and signaling in cartilage development. Osteoarthritis. Cartilage. 82000309334
  47. 47. DenkerA. E.NicollS. B.TuanR. S.Formation of cartilage-like spheroids by micromass cultures of murine C3H10T1/2 cells upon treatment with transforming growth factor-beta 1. Differentiation 5919952534
  48. 48. DischerD. E.JanmeyP.WangY. L.Tissue cells feel and respond to the stiffness of their substrate. Science 310200511391143
  49. 49. DobratzE. J.KimS. W.VoglewedeA.ParkS. S.Injectablecartilage.usingalginate.humanchondrocytes.Arch. Facial. Plast. Surg 1120094047
  50. 50. ElderB. D.AthanasiouK. A.Systematic assessment of growth factor treatment on biochemical and biomechanical properties of engineered articular cartilage constructs. Osteoarthritis. Cartilage. 172009114123
  51. 51. EnglerA. J.RehfeldtF.SenS.DischerD. E.Microtissueelasticity.measurementsby.atomicforce.microscopyitsinfluence.oncell.differentiationMethods Cell Biol. 832007521545
  52. 52. EnglerA. J.SenS.SweeneyH. L.DischerD. E.Matrix elasticity directs stem cell lineage specification. Cell 1262006677689
  53. 53. ErgunC.LiuH.WebsterT. J.Osteoblast adhesion on novel machinable calcium phosphate/lanthanum phosphate composites for orthopedic applications. J Biomed Mater Res A 892009727733
  54. 54. ForgacsG.YookS. H.JanmeyP. A.JeongH.BurdC. G.Role of the cytoskeleton in signaling networks. J Cell Sci. 117200427692775
  55. 55. FrancioliS. E.MartinI.SieC. P.et al.Growth factors for clinical-scale expansion of human articular chondrocytes: relevance for automated bioreactor systems. Tissue Eng 13200712271234
  56. 56. FreedL. E.MartinI.Vunjak-NovakovicG.Frontiers in tissue engineering. In vitro modulation of chondrogenesis. Clin. Orthop Relat Res S46S58, 1999
  57. 57. FreedL. E.Vunjak-NovakovicG.LangerR.Cultivation of cell-polymer cartilage implants in bioreactors. J Cell Biochem. 511993257264
  58. 58. FuchsJ. R.HannoucheD.TeradaS.VacantiJ. P.FauzaD. O.Fetal tracheal augmentation with cartilage engineered from bone marrow-derived mesenchymal progenitor cells. J Pediatr. Surg 382003984987
  59. 59. GabayO.SanchezC.TaboasJ. M.Update in cartilage bio-engineering. Joint Bone Spine 772010283286
  60. 60. GoldringM. B.TsuchimochiK.IjiriK.The control of chondrogenesis. J Cell Biochem. 9720063344
  61. 61. GriffonD. J.SedighiM. R.SchaefferD. V.EurellJ. A.JohnsonA. L.Chitosanscaffolds.interconnectivepore.sizecartilageengineering.Acta Biomater. 22006313320
  62. 62. GrodzinskyA. J.LevenstonM. E.JinM.FrankE. H.Cartilage tissue remodeling in response to mechanical forces. Annu. Rev. Biomed Eng 22000691713
  63. 63. HanawaT.KonM.UkaiH.MurakamiK.MiyamotoY.AsaokaK.Surface modifications of titanium in calcium-ion-containing solutions. J Biomed Mater Res 341997273278
  64. 64. HildnerF.AlbrechtC.GabrielC.RedlH.van G. M.State of the art and future perspectives of articular cartilage regeneration: a focus on adipose-derived stem cells and platelet-derived products. J Tissue Eng Regen. Med. 2011
  65. 65. Hirschfeld-WarnekenV. C.ArnoldM.Cavalcanti-AdamA.Lopez-GarciaM.KesslerH.SpatzJ. P.Cell adhesion and polarisation on molecularly defined spacing gradient surfaces of cyclic RGDfK peptide patches. Eur. J Cell Biol. 872008743750
  66. 66. HorasU.PelinkovicD.HerrG.AignerT.SchnettlerR.Autologous chondrocyte implantation and osteochondral cylinder transplantation in cartilage repair of the knee joint. A prospective, comparative trial. J Bone Joint Surg Am. 85A: 185-192, 2003
  67. 67. HuangA. H.FarrellM. J.MauckR. L.Mechanics and mechanobiology of mesenchymal stem cell-based engineered cartilage. J Biomech. 432010128136
  68. 68. HuangA. H.SteinA.TuanR. S.MauckR. L.Transient exposure to transforming growth factor beta 3 improves the mechanical properties of mesenchymal stem cell-laden cartilage constructs in a density-dependent manner. Tissue Eng Part A 15200934613472
  69. 69. HunterC. J.LevenstonM. E.Maturation and integration of tissue-engineered cartilages within an in vitro defect repair model. Tissue Eng 102004736746
  70. 70. HunzikerE. B.Mechanism of longitudinal bone growth and its regulation by growth plate chondrocytes. Microsc. Res Tech. 281994505519
  71. 71. HuoF.ZhengZ.ZhengG.GiamL. R.ZhangH.MirkinC. A.Polymerpen.lithographyScience.321658 2008
  72. 72. IndrawattanaN.ChenG.TadokoroM.et al.Growth factor combination for chondrogenic induction from human mesenchymal stem cell. Biochem. Biophys. Res. Commun. 3202004914919
  73. 73. IndrawattanaN.ChenG.TadokoroM.et al.Growth factor combination for chondrogenic induction from human mesenchymal stem cell. Biochem. Biophys. Res. Commun. 3202004914919
  74. 74. JohnstoneB.HeringT. M.CaplanA. I.GoldbergV. M.YooJ. U.In vitro chondrogenesis of bone marrow-derived mesenchymal progenitor cells. Exp. Cell Res 2381998265272
  75. 75. KangS. W.JeonO.KimB. S.Poly(lactic-co-glycolicacid.asmicrospheresan.injectablescaffold.forcartilage.tissueengineering.Tissue Eng 112005438447
  76. 76. KayS.ThapaA.HaberstrohK. M.WebsterT. J.Nanostructured polymer/nanophase ceramic composites enhance osteoblast and chondrocyte adhesion. Tissue Eng 82002753761
  77. 77. KayakabeM.TsutsumiS.WatanabeH.KatoY.TakagishiK.Transplantation of autologous rabbit BM-derived mesenchymal stromal cells embedded in hyaluronic acid gel sponge into osteochondral defects of the knee. Cytotherapy. 82006343353
  78. 78. KesslerM. W.GrandeD. A.Tissue engineering and cartilage. Organogenesis. 420082832
  79. 79. KhangD.KimS. Y.Liu-SnyderP.PalmoreG. T.DurbinS. M.WebsterT. J.Enhanced fibronectin adsorption on carbon nanotube/poly(carbonate) urethane: independent role of surface nano-roughness and associated surface energy. Biomaterials 28200747564768
  80. 80. KhangD.LuJ.YaoC.HaberstrohK. M.WebsterT. J.The role of nanometer and sub-micron surface features on vascular and bone cell adhesion on titanium. Biomaterials 292008970983
  81. 81. KhangD.SatoM.PriceR. L.RibbeA. E.WebsterT. J.Selective adhesion and mineral deposition by osteoblasts on carbon nanofiber patterns. Int. J Nanomedicine. 120066572
  82. 82. KimH. J.LeeJ. H.ImG. I.Chondrogenesis using mesenchymal stem cells and PCL scaffolds. J Biomed Mater Res A 922010659666
  83. 83. KimM. K.ChoiS. W.KimS. R.OhI. S.WonM. H.Autologous chondrocyte implantation in the knee using fibrin. Knee Surg Sports Traumatol Arthrosc. 182010528534
  84. 84. KleinT. J.SchumacherB. L.SchmidtT. A.et al.Tissue engineering of stratified articular cartilage from chondrocyte subpopulations. Osteoarthritis. Cartilage. 112003595602
  85. 85. KnabeC.BergerG.GildenhaarR.KlarF.ZreiqatH.The modulation of osteogenesis in vitro by calcium titanium phosphate coatings. Biomaterials 25200449114919
  86. 86. KorossisS. A.WilcoxH. E.WattersonK. G.KearneyJ. N.InghamE.FisherJ.In-vitro assessment of the functional performance of the decellularized intact porcine aortic root. J Heart Valve Dis. 142005408421
  87. 87. KreuzP. C.SteinwachsM. R.ErggeletC.et al.Results after microfracture of full-thickness chondral defects in different compartments in the knee. Osteoarthritis. Cartilage. 14200611191125
  88. 88. KunzlerT. P.DrobekT.SchulerM.SpencerN. D.Systematic study of osteoblast and fibroblast response to roughness by means of surface-morphology gradients. Biomaterials 28200721752182
  89. 89. KunzlerT. P.HuwilerC.DrobekT.VorosJ.SpencerN. D.Systematic study of osteoblast response to nanotopography by means of nanoparticle-density gradients. Biomaterials 28200750005006
  90. 90. KupcsikL.StoddartM. J.LiZ.BennekerL. M.AliniM.Improvingchondrogenesis.potentiallimitationsof. S. O. X.genetransfer.mechanicalstimulation.forcartilage.tissueengineering.Tissue Eng Part A 16201018451855
  91. 91. LeeH. J.ChoiB. H.MinB. H.ParkS. R.Low-intensity ultrasound inhibits apoptosis and enhances viability of human mesenchymal stem cells in three-dimensional alginate culture during chondrogenic differentiation. Tissue Eng 13200710491057
  92. 92. LefebvreV.SmitsP.Transcriptional control of chondrocyte fate and differentiation. Birth Defects Res C. Embryo. Today 752005200212
  93. 93. LehnertD.Wehrle-HallerB.DavidC.et al.Cell behaviour on micropatterned substrata: limits of extracellular matrix geometry for spreading and adhesion. J Cell Sci. 11720044152
  94. 94. LevyA. S.LohnesJ.SculleyS.Le CroyM.GarrettW.Chondral delamination of the knee in soccer players. Am. J Sports Med. 241996634639
  95. 95. LiW. J.TuliR.HuangX.LaquerriereP.TuanR. S.Multilineage differentiation of human mesenchymal stem cells in a three-dimensional nanofibrous scaffold. Biomaterials 26200551585166
  96. 96. LiebermanJ. R.DaluiskiA.EinhornT. A.The role of growth factors in the repair of bone. Biology and clinical applications. J Bone Joint Surg Am. 84A: 1032-1044, 2002
  97. 97. LinY. J.YenC. N.HuY. C.WuY. C.LiaoC. J.ChuI. M.Chondrocytes culture in three-dimensional porous alginate scaffolds enhanced cell proliferation, matrix synthesis and gene expression. J Biomed Mater Res A 8820092333
  98. 98. LongobardiL.O’RearL.AakulaS.et al.Effect-Iof. I. G. F.inthe.chondrogenesisof.bonemarrow.mesenchymalstem.cellsin.thepresence.orabsence.ofT. G.F-betasignaling.J. Bone Miner. Res. 212006626636
  99. 99. LuH. H.SubramonyS. D.BoushellM. K.ZhangX.Tissue engineering strategies for the regeneration of orthopedic interfaces. Ann. Biomed Eng 38201021422154
  100. 100. LuJ.RaoM. P.MacDonald. N. C.KhangD.WebsterT. J.Improved endothelial cell adhesion and proliferation on patterned titanium surfaces with rationally designed, micrometer to nanometer features. Acta Biomater. 42008192201
  101. 101. LyonsK. M.PeltonR. W.HoganB. L.Organogenesis and pattern formation in the mouse: RNA distribution patterns suggest a role for bone morphogenetic protein-2A (BMP-2A). Development 1091990833844
  102. 102. MarlovitsS.ZellerP.SingerP.ResingerC.VecseiV.Cartilagerepair.generationsof.autologouschondrocyte.transplantationEur. J Radiol. 5720062431
  103. 103. MartinI.SmithT.WendtD.Bioreactor-based roadmap for the translation of tissue engineering strategies into clinical products. Trends Biotechnol. 272009495502
  104. 104. MartinI.WendtD.HebererM.The role of bioreactors in tissue engineering. Trends Biotechnol. 2220048086
  105. 105. MartinY.VermetteP.Bioreactors for tissue mass culture: design, characterization, and recent advances. Biomaterials 26200574817503
  106. 106. MehlhornA. T.NiemeyerP.KaiserS.et al.Differential expression pattern of extracellular matrix molecules during chondrogenesis of mesenchymal stem cells from bone marrow and adipose tissue. Tissue Eng 12200628532862
  107. 107. MehlhornA. T.SchmalH.KaiserS.et al.Mesenchymal stem cells maintain TGF-beta-mediated chondrogenic phenotype in alginate bead culture. Tissue Eng 12200613931403
  108. 108. MillerD. C.ThapaA.HaberstrohK. M.WebsterT. J.Endothelial and vascular smooth muscle cell function on poly(lactic-co-glycolic acid) with nano-structured surface features. Biomaterials 2520045361
  109. 109. MitchellN.ShepardN.The resurfacing of adult rabbit articular cartilage by multiple perforations through the subchondral bone. J Bone Joint Surg Am. 581976230233
  110. 110. MizutaH.KudoS.NakamuraE.OtsukaY.TakagiK.HirakiY.Active proliferation of mesenchymal cells prior to the chondrogenic repair response in rabbit full-thickness defects of articular cartilage. Osteoarthritis. Cartilage. 122004586596
  111. 111. NicodemusG. D.VillanuevaI.BryantS. J.Mechanical stimulation of TMJ condylar chondrocytes encapsulated in PEG hydrogels. J Biomed Mater Res A 832007323331
  112. 112. PelttariK.SteckE.RichterW.The use of mesenchymal stem cells for chondrogenesis. Injury 39 Suppl 1: S58S65, 2008
  113. 113. PerettiG. M.PozziA.BallisR.DepontiD.PellacciF.Current Surgical Options for Articular Cartilage Repair. Acta Neurochir. Suppl 1082011213219
  114. 114. RagetlyG. R.SlavikG. J.CunninghamB. T.SchaefferD. J.GriffonD. J.Cartilage tissue engineering on fibrous chitosan scaffolds produced by a replica molding technique. J Biomed Mater Res A 9320104655
  115. 115. RaghunathJ.RolloJ.SalesK. M.ButlerP. E.SeifalianA. M.Biomaterials and scaffold design: key to tissue-engineering cartilage. Biotechnol. Appl. Biochem. 4620077384
  116. 116. RaghunathJ.SalacinskiH. J.SalesK. M.ButlerP. E.SeifalianA. M.Advancing cartilage tissue engineering: the application of stem cell technology. Curr. Opin. Biotechnol. 162005503509
  117. 117. RahfothB.WeisserJ.SternkopfF.AignerT.von derM. K.BrauerR.Transplantation of allograft chondrocytes embedded in agarose gel into cartilage defects of rabbits. Osteoarthritis. Cartilage. 619985065
  118. 118. ReddiA. H.Cartilagemorphogenetic.proteinsrole.injoint.developmenthomoeostasis.regenerationAnn. Rheum. Dis. 62 Suppl 2: ii73ii78, 2003
  119. 119. RehfeldtF.EnglerA. J.EckhardtA.AhmedF.DischerD. E.Cell responses to the mechanochemical microenvironment--implications for regenerative medicine and drug delivery. Adv. Drug Deliv. Rev. 59200713291339
  120. 120. RobertsS.Mc CallI. W.DarbyA. J.et al.Autologous chondrocyte implantation for cartilage repair: monitoring its success by magnetic resonance imaging and histology. Arthritis Res Ther. 5: R60R73, 2003
  121. 121. RobertsonW. W.Jr Newest knowledge of the growth plate. Clin. Orthop Relat Res 2702781990
  122. 122. RonziereM. C.PerrierE.Mallein-GerinF.FreyriaA. M.Chondrogenic potential of bone marrow- and adipose tissue-derived adult human mesenchymal stem cells. Biomed. Mater. Eng 202010145158
  123. 123. SainiS.WickT. M.Effect of low oxygen tension on tissue-engineered cartilage construct development in the concentric cylinder bioreactor. Tissue Eng 102004825832
  124. 124. SakimuraK.MatsumotoT.MiyamotoC.OsakiM.ShindoH.Effects of insulin-like growth factor I on transforming growth factor beta1 induced chondrogenesis of synovium-derived mesenchymal stem cells cultured in a polyglycolic acid scaffold. Cells Tissues. Organs 18320065561
  125. 125. SalinasC. N.AnsethK. S.The enhancement of chondrogenic differentiation of human mesenchymal stem cells by enzymatically regulated RGD functionalities. Biomaterials 29200823702377
  126. 126. SchmidtR. C.HealyK. E.Controlling biological interfaces on the nanometer length scale. J Biomed Mater Res A 90200912521261
  127. 127. SchmittB.RingeJ.HauplT.etal. B. M. P.initiateschondrogenic.lineagedevelopment.ofadult.humanmesenchymal.stemcells.inhigh-density.cultureDifferentiation 712003567577
  128. 128. SchnabelM.MarlovitsS.EckhoffG.et al.Dedifferentiation-associated changes in morphology and gene expression in primary human articular chondrocytes in cell culture. Osteoarthritis. Cartilage. 1020026270
  129. 129. SchulzR. M.BaderA.Cartilage tissue engineering and bioreactor systems for the cultivation and stimulation of chondrocytes. Eur. Biophys. J 362007539568
  130. 130. SchumannD.KujatR.NerlichM.AngeleP.Mechanobiological conditioning of stem cells for cartilage tissue engineering. Biomed Mater Eng 16: S37S52, 2006
  131. 131. Sha’banM.KimS. H.IdrusR. B.KhangG.Fibrin and poly(lactic-co-glycolic acid) hybrid scaffold promotes early chondrogenesis of articular chondrocytes: an in vitro study. J Orthop Surg Res 3: 17, 2008
  132. 132. ShekaranA.GarciaA. J.Extracellular matrix-mimetic adhesive biomaterials for bone repair. J Biomed Mater Res A 2010
  133. 133. ShiehS. J.TeradaS.VacantiJ. P.Tissue engineering auricular reconstruction: in vitro and in vivo studies. Biomaterials 25200415451557
  134. 134. ShirasawaS.SekiyaI.SakaguchiY.YagishitaK.IchinoseS.MunetaT.In vitro chondrogenesis of human synovium-derived mesenchymal stem cells: optimal condition and comparison with bone marrow-derived cells. J. Cell Biochem. 9720068497
  135. 135. SmithR. L.CarterD. R.SchurmanD. J.Pressure and shear differentially alter human articular chondrocyte metabolism: a review. Clin. Orthop Relat Res S89S95, 2004
  136. 136. StadingM.LangerR.Mechanical shear properties of cell-polymer cartilage constructs. Tissue Eng 51999241250
  137. 137. SterodimasA.De FariaJ.CorreaW. E.PitanguyI.Tissue engineering in plastic surgery: an up-to-date review of the current literature. Ann. Plast. Surg. 62200997103
  138. 138. TakagiJ.SpringerT. A.Integrin activation and structural rearrangement. Immunol. Rev. 1862002141163
  139. 139. TarngY. W.CasperM. E.FitzsimmonsJ. S.et al.Directional fluid flow enhances in vitro periosteal tissue growth and chondrogenesis on poly-epsilon-caprolactone scaffolds. J Biomed Mater Res A 952010156163
  140. 140. TemenoffJ. S.MikosA. G.Reviewtissue.engineeringfor.regenerationof.articularcartilage.Biomaterials 212000431440
  141. 141. TinsB. J.Mc CallI. W.TakahashiT.et al.Autologous chondrocyte implantation in knee joint: MR imaging and histologic features at 1-year follow-up. Radiology 2342005501508
  142. 142. UematsuK.HattoriK.IshimotoY.et al.Cartilage regeneration using mesenchymal stem cells and a three-dimensional poly-lactic-glycolic acid (PLGA) scaffold. Biomaterials 26200542734279
  143. 143. vesda.SilvaM. L.MartinsA.Costa-PintoA. R.et al.Chondrogenic differentiation of human bone marrow mesenchymal stem cells in chitosan-based scaffolds using a flow-perfusion bioreactor. J Tissue Eng Regen. Med. 2010
  144. 144. VinatierC.MrugalaD.JorgensenC.GuicheuxJ.NoelD.Cartilageengineering. a.crucialcombination.ofcells.biomaterialsbiofactorsTrends Biotechnol. 272009307314
  145. 145. Vunjak-NovakovicG.MartinI.ObradovicB.et al.Bioreactor cultivation conditions modulate the composition and mechanical properties of tissue-engineered cartilage. J Orthop Res 171999130138
  146. 146. WangP.ZhuF.LeeN. H.KonstantopoulosK.Shear-inducedinterleukin.synthesisin.chondrocytesroles.ofE.prostanoid. E. P. .inE. P.P/proteinc. A. M.kinaseA.P. I.Akt-dependentK.F-kappaN.activationB.J Biol. Chem. 28520102479324804
  147. 147. WashburnN. R.YamadaK. M.SimonC. G.Jr KennedyS. B.AmisE. J.High-throughput investigation of osteoblast response to polymer crystallinity: influence of nanometer-scale roughness on proliferation. Biomaterials 25200412151224
  148. 148. WebsterT. J.AhnE. S.Nanostructured biomaterials for tissue engineering bone. Adv. Biochem. Eng Biotechnol. 1032007275308
  149. 149. WoodJ. J.MalekM. A.FrassicaF. J.et al.Autologouscultured.chondrocytesadverse.eventsreported.tothe.UnitedStates.FoodDrugAdministration.J. J., Malek, M. A., Frassica, F. J. et al. Autologous cultured chondrocytes: adverse events reported to the United States Food and Drug Administration. J Bone Joint Surg Am. 882006503507
  150. 150. WuS. C.ChangJ. K.WangC. K.WangG. J.HoM. L.Enhancement of chondrogenesis of human adipose derived stem cells in a hyaluronan-enriched microenvironment. Biomaterials 312010631640
  151. 151. XuJ.WangW.LudemanM.et al.Chondrogenic differentiation of human mesenchymal stem cells in three-dimensional alginate gels. Tissue Eng Part A 142008667680
  152. 152. YanagaH.ImaiK.YanagaK.Generative Surgery of Cultured Autologous Auricular Chondrocytes for Nasal Augmentation. Aesthetic Plast. Surg 2009
  153. 153. YatesK. E.AllemannF.GlowackiJ.Phenotypic analysis of bovine chondrocytes cultured in 3D collagen sponges: effect of serum substitutes. Cell Tissue Bank. 620054554
  154. 154. YooH. S.LeeE. A.YoonJ. J.ParkT. G.Hyaluronic acid modified biodegradable scaffolds for cartilage tissue engineering. Biomaterials 26200519251933
  155. 155. ZeltingerJ.SherwoodJ. K.GrahamD. A.MuellerR.GriffithL. G.Effect of pore size and void fraction on cellular adhesion, proliferation, and matrix deposition. Tissue Eng 72001557572
  156. 156. ZhengL.FanH. S.SunJ.et al.Chondrogenic differentiation of mesenchymal stem cells induced by collagen-based hydrogel: an in vivo study. J Biomed Mater Res A 932010783792
  157. 157. ZhouX. Z.LeungV. Y.DongQ. R.CheungK. M.ChanD.LuW. W.Mesenchymal stem cell-based repair of articular cartilage with polyglycolic acid-hydroxyapatite biphasic scaffold. Int. J Artif. Organs 312008480489
  158. 158. ZhuL.WuY.JiangH.LiuW.CaoY.ZhouG.Engineered cartilage with internal porous high-density polyethylene support from bone marrow stromal cells: A preliminary study in nude mice. Br J Oral Maxillofac Surg 2009
  159. 159. ZukP. A.ZhuM.MizunoH.et al.Multilineage cells from human adipose tissue: implications for cell-based therapies. Tissue Eng 72001211228
  160. 160. ZwingmannJ.MehlhornA. T.SudkampN.StarkB.DaunerM.SchmalH.Chondrogenic differentiation of human articular chondrocytes differs in biodegradable PGA/PLA scaffolds. Tissue Eng 13200723352343

Written By

Adelola O. Oseni, Claire Crowley, Maria Z. Boland, Peter E. Butler and Alexander M. Seifalian

Submitted: 19 November 2010 Published: 17 August 2011